BIOENERGY RESEARCH
Copyright
Preface
Foreword
List of Contributors
Acknowledgments
3. Use of Agroindustrial Residues for Bioethanol Production
Introduction
Raw Material
Sugar-Containing Residues
Starch-Containing Residues
Cellulose-Containing Residues
Sugar Production and Fermentation
Separate Hydrolysis and Fermentation
Simultaneous Saccharification and Fermentation
Concluding Remarks
Acknowledgments
References
Acknowledgments
Acknowledgments
Acknowledgments
Acknowledgments
Acknowledgments
Acknowledgments
26. Biochar Processing for Sustainable Development in Current and Future Bioenergy Research
Introduction
Theoretical Income Streams
Renewable Energy and Fuel Generation
Carbon Sequestration of Biochars and Carbon Markets
Agricultural Benefits
Economic Analysis
Can Biochar Be a Cost-effective Fertilizer Substitute?
Can Biochar Be a Cost-Effective Approach to Increase Grain Crop Primary Productivity?
Can Biochars Increase Livestock Growth Rates, or Provide a New Market for Semiarid Forestry?
A Comparison of Biochar Carbon Value for Different Potential Income Streams
Conclusion
Disclaimers
Acknowledgments
References
Index
A
B
C
D
E
F
G
H
I
J
K
L
M
N
O
P
R
S
T
U
V
W
X
Z

Author: Gupta VC.   Tuohy M.G.   Kubiсek K.P.   Saddler D.   Xu F.  

Tags: bioenergy  

ISBN: 978-0-444-59561-4

Year: 2014

Text
                    BIOENERGY
RESEARCH:
ADVANCES AND
APPLICATIONS
Edited by

VIJAI K. GUPTA, MARIA G. TUOHY, CHRISTIAN P. KUBICEK,
JACK SADDLER, FENG XU

AMSTERDAM • BOSTON • HEIDELBERG • LONDON • NEW YORK • OXFORD
PARIS • SAN DIEGO • SAN FRANCISCO • SYDNEY • TOKYO


Elsevier 225, Wyman Street, Waltham, MA 02451, USA The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK Radarweg 29, PO Box 211, 1000 AE Amsterdam, The Netherlands Copyright Ó 2014 Elsevier B.V. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher Permissions may be sought directly from Elsevier’s Science & Technology Rights Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting Obtaining permission to use Elsevier material Notice No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress ISBN: 978-0-444-59561-4 For information on all Elsevier publications visit our web site at store.elsevier.com Printed and bound in Great Britain 14 15 16 17 10 9 8 7 6 5 4 3 2 1
Preface The finite nature of fossil fuels and the emission of greenhouse gases as result of the consumption, these resources provide the impetus to seek alternative sources of clean energy, which can be produced in a sustainable manner. This important quest underpins the essential requirement for research and development on various types of bioenergy. Bioethanol production has been the focus of considerable research in the context of liquid fuels for transportation. The use of starchbased (first-generation) agricultural products as substrates as bioethanol feedstocks is possible but raises some concerns because of potential competition with food production. Although numerous investigations on bioenergy have been performed over the past decades to clarify the potential of, and to develop processes for the use of agricultural crops and biomass as feedstock for fuel and energy, the recent period has seen a renewed intensity of research on biomass to bioenergy conversion technologies and processes, with the aim of developing economical and sustainable solutions at commercial scale. To support economic sustainability, biorefinery systems have been implemented to convert renewable materials, such as wood or agricultural crops, into additional valuable products such as platform and feedstock chemicals, and pharma compounds. It is envisaged that the biorefinery concept should enable a transition from the traditional fossil fuel-based platforms for production of commodity products to more environmentally favorable and sustainable bio-based processes. For researchers and industrialists alike, the biorefinery approach brings both significant scientific and technical challenges and much opportunity for technological innovation. Second-generation bioenergy uses the lignocellulose present in woody biomass, forestry residue, agricultural residues, food wastes, agricultural wastes and animal wastes. Agricultural residues include the straw from wheat and rice, sugar cane bagasse, stem and roots from food crops, the top ends of trees like eucalyptus not used in paper manufacture, and fast developing tall grasses (e.g. Miscanthus spp., coastal grasses, etc.). A detailed understanding of the composition of the lignocellulosic waste is essential to develop and optimize mechanistic models for its conversion. Inclusion of pretreatment processes to aid the integration of waste streams into the raw materials for ethanol plants in such models is essential to increase both fuel (ethanol)/bioenergy yields, recover valuable coproducts and biorefinery feedstocks, as well as to reduce process costs. Hydrolysis of lignocellulosic materials is the first step for either digestion to biogas (methane) or fermentation to ethanol. Hydrolysis using enzymes (generally derived from microbial sources) is the preferred option as enzymes can be used to selectively convert carbohydrate-rich biopolymers in biomass to fermentable sugars, without formation of by-products that inhibit downstream bioenergy and biorefinery conversion processes. However, pretreatment of the lignocellulose to reduce its recalcitrance to enzymatic and microbial conversion is essential. Pretreatment by physical, chemical or biological means is an essential process for ethanol production from lignocellulosic materials. Pretreatment also enhances the biodegradability of the wastes for ethanol and biogas production and increases accessibility of the enzymes to the biopolymers present in the biomass/waste feedstocks. Research is necessary to improve process efficiencies in the areas of pretreatment and bioconversion, and to explore new technologies for conversion of lignocellulose to bioenergy. Similarly, the major challenge for microalgal biodiesel production is the high cost of producing microalgal biomass, and the current significant environmental, safety and sustainability concerns surrounding the recovery and extraction of lipid fractions used for biodiesel production. In this sector, the key issues to be solved are the costs for harvesting the algae, protection of the high-oil microalgae from the contamination by other algae, and the development of environmentally and operationally more benign extraction processes. Another important issue for both lignocellulosic ethanol and microalgal biodiesel processes involves the development of technologies for the utilization of coproducts and residues formed through primary bioconversion processes which should increase overall process economics. Utilization of each fraction in biomass agricultural wastes provides an effective way to minimize environmental pollution, address food security problems and improve agricultural waste management approaches. ix
x PREFACE This book focuses on current innovative methods and technological developments which are aimed at overcoming the bottlenecks in biofuel and bioenergy processes. Reviews of the potential of lignocellulosics for the production of (bio)chemicals are also included. Chapters on biorefining routes resulting in a product with higher market value than ethanol have been included. It is envisaged that once such approaches have reached viable commercial scale, global dependence on petroleum for a host of products used in day-to-day applications will be reduced, and a more sustainable global bioeconomy will result. Editors
Foreword Our present industrial civilization relies on the consumption of enormous amounts of energy and much of today’s economic wealth is based on a petroleum-based economy. Petroleum not only is used as energy in transport but also is the starting material of many other products of our daily life including such diverse products as plastics, pharmaceuticals, solvents, fertilizers, pesticides and clothing up to the tarmac, which we use for the transport of these products. However, our continued reliance on fossil fuels will make it impossible to reduce greenhouse gas emissions to stop environmental problems such as global warming. Without decisive actions, the global usage of energy and energy-related emissions of carbon dioxide is predicted to double by 2050. Although there is an active debate when the demand for oil will exceed its supply (Peak Oil), it is clear that our present economic system will need a major shift to develop effective alternatives including a more sustainable economy. This sustainable development will be based on renewable energy and biomass sources as well as more efficient ways to use these. Traditionally, biomass has been used to produce food, feed and wood fiber. But biomass can also provide energy in the form of (bio)fuels and it can be used as a source of feedstock chemicals replacing the petroleum-based products. The development of such a biobased economy is occurring already at a relatively rapid pace and some of its products are already on the market including first-generation biofuels. The commercial viability of this approach will depend largely on the availability of cost-competitive technologies capable of converting (waste) biomass within a holistic concept of a biorefinery to biofuels and other bio-based products. Biorefiningdthe sustainable processing of biomass into food/feed ingredients, chemicals, materials and bioenergydaims to use the available biomass resources as efficient as possible. At the moment, a wide range of biomass conversion technologies are under development to improve efficiencies, lower costs along the whole supply chain and improve the environmental performance. But there is also a need for further technological innovation leading to more efficient and cleaner conversion of a more diverse range of feedstocks. These include not only existing lignocellulosic waste residues from forestry, agriculture and urban communities but also the generation of new feedstocks from energy crops or microalgae. A first wave of cellulosic biofuels demonstration plants is now reaching completion producing transportation fuels from agro-, forestry and process residues. To make the overall process more market competitive, these plants co-produce added-value biobased products thereby supplying processes that are less energy or chemically intensive compared to their petroleum-based counterparts. Increasing deployment of biomass will include also other challenges for our society including an increasing competition for land, questions of biodiversity and soil quality or the availability of water resources. But biomass will be an important part of the future energy mix thereby contributing to a low CO2 future. Excluding biomass from the energy mix would significantly increase the cost of decarbonizing our energy system. This book has been initiated to describe the current stage of knowledge on bioenergy research from various perspectives, thereby outlining also those areas where further progress is needed. Dr. Bernhard Seiboth Professor, Head of Molecular Biotechnology, Vienna University of Technology, Vienna, Austria xi
List of Contributors Bruno C. Aita Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Y. Allahverdiyeva Department of Biochemistry, University of Turku, Turku, Finland Samuel Amartey Division of Biology, Imperial College of Science, Technology and Medicine, South Kensington, London, UK M. Anusree Biotechnology Division, National Institute for Interdisciplinary Science and Technology (NIIST), CSIR, Trivandrum, Kerala, India E.M. Aro Department of Biochemistry, University of Turku, Turku, Finland Rama Raju Baadhe Department of Biotechnology, National Institute of Technology, Warangal, Andhra Pradesh, India Mikhail Balakshin Prussia, PA, USA Renmatix, R&D Department, King of Ciarán John Forde AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland Michael P. Garver Department of Paper and Bioprocess Engineering, College of Environmental Science and Forestry, State University of New York, Syracuse, NY, USA Juliana M. Gasparotto Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Maria Gavrilescu Department of Environmental Engineering and Management, Gheorghe Asachi Technical University of Iasi, Iasi, Romania; Academy of Romanian Scientists, Bucharest, Romania Nishant Gopalan Biotechnology Division, National Institute for Interdisciplinary Science and Technology (NIIST), CSIR, Trivandrum, Kerala, India Alex Berlin Novozymes, Protein Chemistry Department, Davis, CA, USA Vipin Gopinath Biotechnology Division, National Institute for Interdisciplinary Science and Technology (NIIST), CSIR, Trivandrum, Kerala, India Susan Boland AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland Richard J.A. Gosselink Food and Biobased Research, Wageningen UR, Wageningen, The Netherlands John Bosco Carrigan AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland Maria Aparecida F. Cesar-Oliveira Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, Curitiba, Paraná, Brazil Daniel P. Chielle Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Rhykka Connelly UT Algae Science and Technology Facility, University of Texas at Austin, Austin, TX, USA Claudiney S. Cordeiro Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, Curitiba, Paraná, Brazil Ed de Jong Netherlands Avantium Chemicals, Amsterdam, The Kiran S. Dhar Biotechnology Division, National Institute for Interdisciplinary Science and Technology (NIIST), CSIR, Trivandrum, Kerala, India Hanshu Ding Department of Protein Chemistry, Novozymes Inc., Davis, California, USA Thaddeus Chukwuemeka Ezeji The Ohio State University, Department of Animal Sciences and Ohio State Agricultural Research and Development Center (OARDC), Wooster, OH, USA Tingyue Gu Department of Chemical and Biomolecular Engineering, Ohio University, Athens, OH, USA Vijai K. Gupta Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland Patrick C. Hallenbeck Département de Microbiologie et Immunologie, Université de Montréal, Montréal, Québec, Canada Daniel J. Hassett Department of Molecular Genetics, Biochemistry and Microbiology, University of Cincinnati, College of Medicine, Cincinnati, OH, USA Alan Hernon AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland Charles Hyland Department of Civil & Environmental Engineering, The University of Auckland, Auckland, New Zealand Tao Jin Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), College of Environmental Science and Engineering, Nankai University, Tianjin, China Vasiliki Kachrimanidou Department of Food Science and Human Nutrition, Agricultural University of Athens, Athens, Greece Rodrigo Klaic Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil xiii
xiv LIST OF CONTRIBUTORS Nikolaos Kopsahelis Department of Food Science and Human Nutrition, Agricultural University of Athens, Athens, Greece Shirley Nakagaki Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, Curitiba, Paraná, Brazil S.N. Kosourov Department of Biochemistry, University of Turku, Turku, Finland K. Apostolis A. Koutinas Department of Food Science and Human Nutrition, Agricultural University of Athens, Athens, Greece Christian P. Kubicek Research Area Biotechnology and Microbiology, Institute of Chemical Engineering, TU Wien, Gumpendorferstrasse Wien, Austria Jyothi Kumaran Human Health Therapeutics, National Research Council Canada, Ottawa, ON, Canada; School of Environmental Sciences, University of Guelph, Guelph, ON, Canada Gustavo B. Leite Département de Microbiologie et Immunologie, Université de Montréal, Montréal, Québec, Canada Madhavan Nampoothiri Biotechnology Division, National Institute for Interdisciplinary Science and Technology (NIIST), CSIR, Trivandrum, Kerala, India W.J. Oosterkamp Netherlands Oosterkamp Oosterbeek Octooien, The Anthonia O’Donovan Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland Irmene Ortı́z Departamento de Procesos y Tecnologı́a, Universidad Autónoma Metropolitana - Cuajimalpa, México D.F., México Ravichandra Potumarthi Department of Chemical Engineering, Monash University, Clayton, Victoria, Australia Wensheng Qin Department of Biology, Lakehead University, ON, Canada Xiangling Li Aquatic and Crop Resource Development, National Research Council Canada, Ottawa, ON, Canada; College of Chinese Medicine, Guangzhou University of Chinese Medicine, Guangzhou, China Rodolfo Quintero Departamento de Procesos y Tecnologı́a, Universidad Autónoma Metropolitana - Cuajimalpa, México D.F., México Shijie Liu Department of Paper and Bioprocess Engineering, College of Environmental Science and Forestry, State University of New York, Syracuse, NY, USA Nasib Qureshi United States Department of Agriculture, National Center for Agricultural Utilization Research, ARS, Bioenergy Research, Peoria, IL, USA Fan Lu College of Bioengineering, Hubei University of Technology, Wuhan, Hubei Province, China Luiz P. Ramos Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, Curitiba, Paraná, Brazil Miranda Maki Department of Biology, Lakehead University, ON, Canada Nirupama Mallick Agricultural and Food Engineering Department, Indian Institute of Technology, Kharagpur, West Bengal, India Shovon Mandal Section of Ecology, Behavior and Evolution, University of California, San Diego, CA, USA Marcio A. Mazutti Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Mark P. McHenry School of Engineering and Information Technology, Murdoch University, Perth, Western Australia, Australia Marie Meaney AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland Naveen Kumar Mekala Department of Biotechnology, National Institute of Technology, Warangal, Andhra Pradesh, India Clive Mills AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland Jéssica M. Moscon Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Adrian Muller Research Institute of Organic Farming FiBL, Zurich, Switzerland; Institute for Environmental Decisions, Swiss Federal Institutes of Technology (ETH), Zurich, Switzerland Gabrielly V. Ribeiro Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Paulo R.S. Salbego Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Ajit K. Sarmah Department of Civil & Environmental Engineering, The University of Auckland, Auckland, New Zealand Gauri Dutt Sharma garh, India Bilaspur University, Bilaspur, Chattis- Dong Shen Tong Research Group for Advanced Materials & Sustainable Catalysis (AMSC), Breeding Base of State Key Laboratory of Green Chemistry Synthesis Technology, College of Chemical Engineering and Materials Science, Zhejiang University of Technology, Hangzhou, Zhejiang, China Fabiane M. Stringhini Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil Maria G. Tuohy Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland Victor Ujor The Ohio State University, Department of Animal Sciences and Ohio State Agricultural Research and Development Center (OARDC), Wooster, OH, USA Luiz J. Visioli Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil
LIST OF CONTRIBUTORS Hongyu Wang Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), College of Environmental Science and Engineering, Nankai University, Tianjin, China Colin Webb School of Chemical Engineering and Analytical Science, University of Manchester, Manchester, England, United Kingdom Lin Mei Wu Research Group for Advanced Materials & Sustainable Catalysis (AMSC), Breeding Base of State Key Laboratory of Green Chemistry Synthesis Technology, College of Chemical Engineering and Materials Science, Zhejiang University of Technology, Hangzhou, Zhejiang, China Fernando Wypych Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, Curitiba, Paraná, Brazil Feng Xu Department of Protein Chemistry, Novozymes Inc., Davis, California, USA Trent Chunzhong Yang Aquatic and Crop Resource Development, National Research Council Canada, Ottawa, ON, Canada Jie Yang Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), College of xv Environmental Science and Engineering, Nankai University, Tianjin, China Yanbin Yin Department of Biological Sciences, Northern Illinois University, DeKalb, IL, USA Wei Hua Yu Research Group for Advanced Materials & Sustainable Catalysis (AMSC), Breeding Base of State Key Laboratory of Green Chemistry Synthesis Technology, College of Chemical Engineering and Materials Science, Zhejiang University of Technology, Hangzhou, Zhejiang, China Chun Hui Zhou Research Group for Advanced Materials & Sustainable Catalysis (AMSC), Breeding Base of State Key Laboratory of Green Chemistry Synthesis Technology, College of Chemical Engineering and Materials Science, Zhejiang University of Technology, Hangzhou, Zhejiang, China; The Institute for Agriculture and the Environment, University of Southern Queensland, Queensland, Australia Minghua Zhou Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), College of Environmental Science and Engineering, Nankai University, Tianjin, China
C H A P T E R 1 Current Bioenergy Researches: Strengths and Future Challenges Naveen Kumar Mekala 1, Ravichandra Potumarthi 2,*, Rama Raju Baadhe 1, Vijai K. Gupta 3 1 Department of Biotechnology, National Institute of Technology, Warangal, Andhra Pradesh, India, Department of Chemical Engineering, Monash University, Clayton, Victoria, Australia, 3Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland *Corresponding author email: ravichandra.potumarthi@monash.edu; pravichandra@gmail.com 2 O U T L I N E Introduction Different Forms of Bioenergy 1 3 Biopellets 3 Bioethanol Feedstock for Bioethanol Pretreatment of Lignocelluloses Biological Pretreatment Physical Pretreatment Chemical Pretreatment Bioethanol Fermentation Molecular Biology Trends in Bioethanol Production Development Bioreactors in Ethanol Production Immobilization of Cells for Ethanol Production 3 3 4 5 6 6 7 8 8 9 10 10 11 12 12 13 Biogas Biogas Feedstock Household Digesters for Biogas Fixed Dome Digesters Floating Drum Digesters Social and Environmental Aspects of Biogas Digesters 14 15 15 15 16 17 Conclusion 17 References 18 9 INTRODUCTION rate of consumption, crude oil reserves, natural gas and liquid fuels were expected to last for around 60 and 120 years, respectively (British Petroleum Statistical Review, 2011). An additional challenge with fossil fuel consumption is emission of greenhouse gases (GHGs). In 2010, an average of 450 g of CO2 was emitted by production of 1 kWh of electricity from the coal (Figure 1.1) (International Energy Agency Statistics, 2012). It is also clear that coal’s share of the global Modern world is facing two vital challenges, energy crisis and environmental pollution. Energy is a key component for all sectors of modern economy and plays an elementary role in improving the quality of life (US DOE, 2010). In current situations, approximately 80% of world energy supplies rely on rapidly exhausting nonrenewable fossil fuels. At the current Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00001-2 Biodiesel Feedstocks for Biodiesel Biodiesel from Pure Vegetable Oil Biodiesel from Animal Fat Wastes Other Waste Cooking Oils Algae as a Biodiesel Source Bioreactors for Biodiesel Production 1 Copyright Ó 2014 Elsevier B.V. All rights reserved.
2 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES Energy production Level : World Legend Mtoe : [ 2 347.03 ; 2 347.03 ] [ 1 993.36 ; 2 347.03 ] [ 1 605.20 ; 1 993.36 ] [ 1 253.92 ; 1 605.20 ] [ 1 160.87 ; 1 253.92 ] [ 1 066.08 ; 1 160.87 ] [ 727.64 ; 1 066.08 ] [ 63.86 ; 727.64 ] FIGURE 1.1 Global energy production chart signifies the growing demand for energy. Source: IEA, 2012. (For color version of this figure, the reader is referred to the online version of this book.) energy continues to rise, and by 2017 coal will come close to surpassing oil as the world’s top energy source. China and India lead the growth in coal consumption over the next 5 years. Research says China will surpass the rest of the world in coal demand during the outlook period, while India will become the largest seaborne coal importer and second largest consumer, surpassing the United States (IEA, 2012). Growing global energy needs, release of environmental pollutants from fossil fuels and national security have finally tuned the attention in clean liquid fuel as a suitable alternative source of energy. The alternative bioenergy sources, not only cut the dependence on oil trade and reduce the doubts caused by the fluctuations in oil price, but also secure reductions in environmental pollution due to their high oxygen content (Huang et al., 2008; Boer et al., 2000). In this context, the availability of bioenergy in its two main appearances, wood and agro energy can offer cleaner energy services to meet basic energy requirements. This century could see a remarkable switchover from fossil fuel-based energy to bioenergy-based economy, with agriculture and forestry as the main sources of feedstock for biofuels such as wood pellets, fuelwood, charcoal, bioethanol, and biodiesel (Agarwal, 2007). Moreover, energy crops can be part of highly specialized and various agricultural production chains and biorefineries, where a variety of bioproducts could be obtained besides bioenergy, which are important for their economic competitiveness (United Nations Environment Program, 2006). The exploitation of currently unused by-products and growing energy crops can address other existing environmental concerns. Perennial energy crops and plantations are generally characterized by higher biodiversity compared with conventional annual crops. By providing more continuous soil cover, they reduce the impact of rainfall and sediment transport, thereby preventing soil erosion. The introduction of annual energy crops into crop systems allows for diversification and expansion of crop rotations, replacing less favorable monocropping systems (Kheshgi et al., 1996). Deforested, degraded and marginal land can be rehabilitated with bioenergy plantations, thus helping to combat desertification and hopefully reducing market and geosocial pressures on high-quality arable land. Biofuels can be obtained in bulk when they are derived from agricultural crops, crop residues and processing wastes from agroindustries, forests, etc. Despite this immense potential, existing biofuel policies have been very costly; they produce slight reductions in fossil fuel use and increase, rather than decrease, in GHG emissions (Wuebbles and Jain, 2001). However, recent volatility and rise in international fossil fuel prices, make biomass increasingly competitive as energy feedstock. Current bioenergy research around the globe should direct us toward reduced production cost, higher energy conversion efficiency and greater costeffectiveness of biofuels. After all we are aware of a fact “use of biomass as a potentially large source of energy in the 21st century will have a significant impact in rural, agricultural and forestry development” (UNEP, 2006).
3 BIOETHANOL Different Forms of Bioenergy Organic matter holding bioenergy sources in side is known as biomass. We can utilize this biomass in many different ways, through something as simple as burning wood for heat, or as complex as growing genetically modified microbes to produce cellulosic ethanol (Adler et al., 2009). Since nearly entire bioenergy can be traced back to energy from sunlight, bioenergy has the key advantage of being a renewable energy source. Here, in this chapter we will discuss various forms of bioenergy and their application in detail. BIOPELLETS Today, wood pellets are an imperative and wellaccepted fuel in lots of different countries and the according markets are likely to rise even further in future. For these reasons, it is feared that the inadequate availability of cheap wood as a feedstock for pellets will limit this market increase (Marina et al., 2011; Larsson et al., 2008). As alternative, autumn leaves from urban areas, as a seasonal available waste material, are the possible substitutes for or additives to wood. In lot of Western countries, wood pellets become a more and more significant fuel for the use in small furnaces for household buildings or in industries as a climate-neutral alternative to crude oil or natural gas (Verma et al., 2012; Nielsen et al., 2009). This pelletized biomass has a number of advantages like tolerance against microbial degradation, high transport and storage density of bioenergy, and the process of pelletization is quite simpler (Figure 1.2). BIOETHANOL Bioethanol is the most common biofuel worldwide. It is produced by simple fermentation of sugars derived from wheat, corn, sugar beets, sugarcane, molasses and any sugar or starch sources that alcoholic beverages can be made from (Cara et al., 2008). Bioethanol can be used in petrol engines as a substitute for gasoline. Bioconversion of lignocellulosics into fermentable sugars is a biorefining area in which enormous research labors have been invested, as it is a prerequisite for the subsequent bioethanol production (Broder et al., 1992). Although extensive research has been carried out to meet the potential challenges of bioenergy generation, there is no self-sufficient process or technology available today to convert the lignocellulosic biomass to bioethanol (Tu et al., 2007). Use of bioethanol-blended fossil fuel for automobiles can significantly cut the petroleum use and exhaust GHG emission. Bioethanol can be produced from different kinds of raw materials and these raw materials are classified into three categories of agricultural raw materials: simple sugars, starch and lignocelluloses (Mustafa and Havva, 2009). Bioethanol from sugarcane, under proper conditions, is essentially a clean fuel and has several advantages over petroleum-derived gasoline in reducing GHG emissions and improving air quality in metropolitan cities. Conversion technologies for producing bioethanol from cellulosic biomass resources such as forest materials, agricultural residues and urban wastes are under development and have not yet been established commercially (Demirbas, 2008). Feedstock for Bioethanol Across the globe, there is a rising need to find out new and cheap carbohydrate sources for bioethanol production (Mohanty et al., 2009). Presently, a serious focus is on biofuels made from renewable energy crops such as sugarcane, corn, wheat, soybeans, etc. In a given production line, the comparison of the biomass includes several issues: (1) cultivation practices, (2) chemical composition of the biomass, (3) use of resources, (4) emission of GHGs, (5) availability of land and land use practices, (6) soil erosion, (7) energy balance, (8) price of the (b) (a) Species 1 Species 2 ... Drying Milling Conditioning Species n Pelletizing Leaf mixture Analyzing Leaf pellets FIGURE 1.2 (a) Experimental flow sheet for pelletization of leaves; (b) leaf pellets. (For color version of this figure, the reader is referred to the online version of this book.)
4 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES FIGURE 1.3 Major lignocellulosic feedstock explored for bioethanol production. Source: Taherzadeh and Karimi, 2007. biomass, (9) contribution to biodiversity and landscape value losses, (10) direct economic value of the feedstock, (11) water requirements and water availability, (12) creation or maintain of employment, and (13) logistic cost (transport and storage of the biomass) (Gnansounou et al., 2005). Bioethanol feedstock can be divided into three major groups: (1) sugar-based feedstock (e.g. sugarcane, beet sugar, sorghum and fruits), (2) starchy feedstock (e.g. corn, sweet potato, rice, potatoes, cassava, wheat and barley), and (3) lignocellulosic feedstock (e.g. wood, straw, grasses, and corncob). In short term, production of bioethanol as a fuel is almost entirely dependent on starch and sugars from existing food crops (Smith, 2008; Potumarthi et al., 2012). The negative part in producing bioethanol from starch and sugar is that the feedstock tends to be costly and demanded by other applications as well (Enguidanos et al., 2002). Lignocellulosic biomass is envisaged to provide a major portion of the raw materials for bioethanol production in the long term due to its low cost and high availability (Gnansounou et al., 2005). Up to 2003, about 60% of global bioethanol was obtained from sugarcane and 40% from all other crops (Dufey, 2006). Brazil utilizes sugarcane for bioethanol production, while the United States and other western countries mainly use starch from corn, wheat and barley (Linde et al., 2008). Brazil is the largest producer of sugarcane with about 672,157,000 tons of global production followed by India, second largest producer with 285,029,000 ton production (Food & Agricultural Organization of United Nations, 2013). Bioethanol production in Brazil is less expensive than in the United States from corn or in Europe from sugar beet, because of lower labor costs, shorter processing time, lower transport costs, and other input costs. After sugarcane, starch is the high-yield feedstock for bioethanol production, but pretreatment is necessary to produce bioethanol by fermentation (Pongsawatmanit et al., 2007). Starch is a homopolymer consisting monomers of D-glucose and for bioethanol production it is necessary to break down this carbohydrate for obtaining glucose syrup, which can be further transformed into bioethanol by yeasts. Starch-based feedstock are the most utilized for bioethanol production in America and Europe. Biomass from agricultural waste (wheat straw, sugarcane bagasse, etc.), wood, and energy crops are attractive materials for bioethanol production since it is the most abundant reproducible assets on earth (Figure 1.3). The existing biomass from crops could produce up to 442 American billion liters per year of bioethanol (Bohlmann, 2006). Thus, the total possible bioethanol production from crop residues and wasted crops is 491 American billion liters per year, about 16 times higher than the existing world bioethanol production. Advantages of biofuels are as follows: (1) biofuels are easily available from common biomass sources, (2) biofuels have a considerable environmentally friendly potential, and (3) they are biodegradable and contribute to sustainability (Balat, 2007; Mekala et al., 2008). Although lignocellulosic biomass is the best alternative source, initial pretreatment is a must to attain simple sugars for simultaneous ethanol fermentation. Pretreatment of Lignocelluloses Woody materials including bark, wood, and mixture of other residues from the forest contain cellulose, hemicelluloses, lignin and small amount of other biomass
5 BIOETHANOL of these lignocelluloses separates the sugars and lignin and disrupts the crystalline portion of the biopolymers (Hu et al., 2008). Different pretreatment methods have been explored for achieving the optimistic situations with different biomass. In general, pretreatment methods can be divided into biological pretreatment, physical pretreatment, and chemical pretreatment according to the pretreatment procedure. Some pretreatments combine two or more of broadly explored methodologies. Table 1.1 recaps some of the broadly explored pretreatment methods as per the classification (Sun and Cheng, 2002). FIGURE 1.4 Chemical composition of lignocellulosic biomass (SW, soft wood; HW, hard wood). contents (Figure 1.4). Cellulose is the major component in plant biomass and it is made of anhydroglucopyranose or glucose residues, which can be converted to glucose and act as major source of hexoses in woody feedstock (Alvira et al., 2010). Due to the hydrogen bonds in it, cellulose is a highly crystalline structure and it is difficult to hydrolyze. Unlike cellulose, hemicelluloses are heteropolymers composed of both fivecarbon sugars and six-carbon sugars, including glucose, mannose, arabinose, xylose and others (Bochek, 2003). Due to its amorphous structure, hemicellulose is easily breakable by dilute acid or base. Lignin is the third major part in wood and comprises the glue that guards woody biomass from pathogens. Lignin mainly consists of phenolic units and with available technology we cannot use lignin as a source of bioethanol. Pretreatment Biological Pretreatment Most pretreatment technologies require selected and expensive amenities or equipment that have high power requirements, depending on the process. In particular, physical and chemical processes require rich energy for biomass conversion, whereas, biological treatment via microorganisms is a safe and environmentally friendly method and is increasingly being promoted as a process that does not require high energy, even for lignin removal from a lignocellulosic biomass (Okano et al., 2005; Potumarthi et al., 2013; Ravichandra et al., 2013). Phanerochaete chrysosporium is one among the best microbial models to study the lignin degradation by white rot fungi. Fungi breaks down lignin anaerobically through a family of extracellular enzymes collectively termed as lignases (Howard et al., 2003). Two families of lignolytic enzymes are generally considered to play vital role in the enzymatic degradation: peroxidases (lignin peroxidase) and phenol oxidase TABLE 1.1 Pretreatment Methods of Lignocellulosic Feedstock Energy Pretreatment Source Means Effect Biological pretreatment Microorganisms Actinomysis, fungi Removes lignin and reduces the degree of polymerization (DP) of celluloses Physical pretreatment Comminution Ball milling, compression milling, colloidal milling Decreases the particle size, crystallinity and DP of cellulose Steam explosion High-pressure steam Partially hydrolyze cellulose and hemicelluloses Ultrasonic radiation Electron beam, gamma rays, microwave Increases the surface area and softens the lignin Acid Hydrochloric acid, hydrofluoric acid, nitric acid, sulfuric acid, peracetic acid, etc. Alkaline Sodium hydroxide, sodium carbonate, ammonia, ammonium sulfate, lime, etc. Decreases in crystallinity and DP of cellulose; partial or complete degradation of hemicellulose; delignification Gases Chlorine dioxide, nitrogen dioxide Cellulose solvents DMSO, cadoxen, CMCS Chemical pretreatment Source: Moiser et al., 2005.
6 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES (Malherbe and Cloete, 2003). Other enzymes are not fully explored including glucose oxidase, methanol oxidase, glyoxal oxidase, and oxidoreductase (Eriksson, 2000). Another best example was Trichoderma reesei, which is a mesophilic cellulolytic fungus isolated in the mid-1950s. By the mid-1970s, an impressive collection of more than 14,000 cellulolytic fungi were isolated against cellulose and other insoluble fibers (Coyne et al., 2013). Trichoderma reesei, although a good producer of hemi and cellulolytic enzymes, is unable to degrade lignin (Mekala et al., 2008; Gupta et al., 2013). Actinomycetes are also best tested for their task in lignin biodegradation. Lignolytic enzymes like peroxidases, ligninase and manganese peroxidase were discovered in P. chrysosporium (Saritha et al., 2012). Based on this, P. chrysosporium was taken up for biological delignification of wood and paddy straw in ethanol production. But, the extent of delignification was inadequate to expose a significant portion of cellulose for enzymatic hydrolysis. Thus, from the reports available, it is evident that white rot fungi and actinomycete can be used jointly to remove lignin from lignocellulosic substrates, and further studies are required to shorten the incubation time and to optimize the delignification process. Physical Pretreatment MECHANICAL COMMINUTION The objective is to cut the particle size and crystallinity of lignocellulosic biomass in order to increase the surface area and reduce the degree of polymerization. Methods like chipping, grinding and milling can be used to improve the further enzymatic hydrolysis. However, this process is not economically feasible due to the high energy requirement (Tassinari et al., 1980). During comminution, vibratory ball milling is found to be more efficient in breaking down the cellulose molecules of spruce and aspen chips and improving the digestibility of the biomass than ordinary ball milling (Sun and Cheng, 2002). The power requirement of mechanical comminution of agricultural materials depends on the final particle size and the waste biomass characteristics. STEAM EXPLOSION It is a hydrothermal pretreatment in which the lignocellulose is subjected to pressurized water vapors for few seconds to several minutes, and then suddenly depressurized. In this process, combination with the partial hydrolysis of hemicelluloses and solubilization, the lignin is redistributed and removed up to certain level from the material (Pan et al., 2005). Although this technique is cost-effective, it generates toxic by-products and the hemicelluloses degradation is partial (Saritha et al., 2012). ULTRASONIC PRETREATMENT This technique is extensively used for the treatment of sludge from wastewater treatment plants. An experiment on carboxyl methyl cellulose with irradiation increased the rate of enzymatic hydrolysis up to 200% approximately (Imai et al., 2004). The mechanism of action in ultrasonic treatment remains unknown. One guess is that, the hydrogen bonds in the crystalline cellulose structures were broken due to irradiation energy, whose energy is higher than the hydrogen bond energy (Bochek, 2003). EXTRUSION This process disrupts the crystal structure of lignocellulose and increases the accessibility of carbohydrates to enzyme. In this case, materials are subjected to heating, mixing and shearing resulting in physical and chemical modifications in biomass structure (Karunanithy et al., 2008). However, the process is novel and not widely applied. Chemical Pretreatment ACID HYDROLYSIS During acid hydrolysis, concentrated acids like HCl and H2SO4 have been used to pretreat lignocellulosic biomass. Although acids are influential agents for cellulose hydrolysis, intense acids are poisonous, corrosive, and require chemical reactors that are resistant to corrosion. In addition, concentrated acid must be removed after hydrolysis of celluloses into simple sugars, which simultaneously enter into alcoholic fermentation (Potumarthi et al., 2013; Ravichandra et al., 2013). Hydrolysis using dilute acid has been industriously developed for pretreatment of lignocellulosic biomass (O’Donovan et al., 2013). The dilute sulfuric acid pretreatment can attain high reaction rates and drastically improve cellulose hydrolysis. Dilute acids at lower temperatures, saccharification suffered from low yields because of sugar decomposition (Chen et al., 2009). High temperatures, with dilute acids are favorable for cellulose hydrolysis. In recent times, dilute acid hydrolysis processes use less harsh environment and achieve high xylan to xylose conversion rates. Achieving high xylan to xylose conversion yields is required to achieve favorable economics, because xylan is the third most promising carbohydrate in many lignocellulosic feedstocks (Sun and Cheng, 2002). Primarily two types of dilute acid pretreatment processes are well studied: high-temperature (T > 160  C), continuous flow process for low solids loading (5e15% (weight of biomass/weight of reaction mixture)) (Converse et al., 1989), and low-temperature (T < 160  C), batch process for high solids loading (10e40%) (Esteghlalian et al., 1997). Although dilute acid hydrolysis can significantly improve the cellulose
7 BIOETHANOL breakdown, overall cost is typically higher when compared with few other physicochemical pretreatment processes such as steam explosion. ALKALINE HYDROLYSIS Usually alkaline hydrolysis was carried out at low temperature and pressure and it may be completed even at ambient conditions. The only drawback of this process is time; it might be hours or even days to complete the hydrolysis. During lime pretreatment, some calcium is tainted into nonrecoverable salts or included in the biomass (Chang et al., 2001). Other alkaline pretreatment methods include calcium, potassium, sodium and ammonium hydroxides as reactants based on biomass category. Among these reactants, sodium hydroxide receives the most attention followed by lime, due to its advantage of being low cost and secure to use, as well as it is easily recoverable from water as insoluble CaCO3 by reaction with CO2. Further delignification of feedstocks can be enhanced by supplying surplus air/ oxygen (Hu et al., 2008). We can compare alkaline pretreatment of feedstocks to Kraft pulping, where lignin was removed efficiently, thus improving the reactivity of polysaccharides. Alkaline hydrolysis also effectively removes acetyl groups and uronic substitutions from hemicellulose; thus, the surface of hemicellulose becomes more accessible to the hydrolytic enzymes. AMMONIA HYDROLYSIS Ammonia has abundant desirable characteristics as a pretreatment reagent. It is a valuable swelling reagent for lignocellulosic biomass, and ammonia has high selectivity for reactions with lignin over those with carbohydrates (Kim et al., 2003). It is one of the most extensively used commodity chemicals with about one-fourth the cost of sulfuric acid on molar basis. Its high volatility makes it easy to recover and recycle. It is a nonpolluting and noncorrosive chemical. One of the known reactions of aqueous ammonia with lignin is cleavage of ether (CeOeC) bonds in lignin as well as ester bonds in the ligninecarbohydrate complex (Lewin and Roldan, 1971). This above reaction indicates that ammonia pretreatment selectively cuts the lignin content in biomass. Lignin is believed to be a major hindrance in enzymatic hydrolysis and there are several advantages by removing lignin early in the conversion process before it faces the biological treatment. OZONOLYSIS Ozone is a leading oxidant that demonstrates high delignification efficiency. This ozonolysis is done at room temperature and at normal pressure. In this case we do not locate any inhibitory by-products, which affect the simultaneous fermentation steps (Saritha et al., 2012). An important drawback is ozone requirement in large quantities, which can make the process economically unapproachable (Sun and Cheng, 2002). Bioethanol Fermentation Once the lignocelluloses were hydrolyzed into simple sugars, they have to be fermented to ethanol. The hydrolyzate now contains various hexoses and pentoses, mainly glucose and xylose, depending on the substrate and the pretreatment method applied. Currently, fermentation of simple sugars is mostly done using yeast cultures (Saccharomyces cerevisiae), because of its well-known characteristics, toughness and high ethanol yield. However, S. cerevisiae can only ferment hexoses and not the pentoses. The pentose sugars can be fermented in an additional step by another microorganism or by S. cerevisiae itself through genetic engineering approaches, so that it is able to ferment pentoses as well (Van Zyl et al., 2007). List of most popular yeast strains used for ethanol fermentation are mentioned in Table 1.2. Besides a high yield, an important aspect with fermentation is alcohol tolerance in the fermenting organisms. A strategy to defeat this crisis is to have a system where the ethanol is recovered at regular intervals to keep the alcohol concentrations under control. Another problem is inhibitory compounds that TABLE 1.2 Yeast Species That Produce Ethanol as the Main Fermentation Product Strain/Species Temperature ( C) pH Carbon Source/Concentration (g/l) Incubation Time (h) Ethanol Concentration Produced (g/l) 27817- S. cerevisiae 30 5.5 Glucose/(50e200) 18e94 91.8 L-041- S. cerevisiae 30e35 e Sucrose/(100) 24 50 ATCC 24860-S. cerevisiae 30 4.5 Molasses/(1.6e5.0) 24 18.5 Bakers’ yeastdS. cerevisiae 28 5.0 Sucrose/(220) 96 96.71 CMI237- S. cerevisiae 30 4.5 Sugar/(160) 30 70 27774- Kluyveromyces fragilis 30 5.5 Glucose/(20e120) 18e94 48.6 Source: Lin and Tanaka, 2006.
8 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES TABLE 1.3 Comparison between Biodiesel and Petroleum Diesel Advantages (1) (2) (3) (4) (5) (6) (7) Domestically produced from nonpetroleum, renewable resources Can be used in most diesel engines, especially in recent ones Less air pollutants (other than nitrogen oxides) Less greenhouse gas emissions (e.g. B20 reduces CO2 by 15%) Biodegradable Nontoxic Safer to handle are produced during the pretreatment. As mentioned above they can be reduced by an additional detoxification step, but this is an expensive operation (Van Maris et al., 2006). Molecular Biology Trends in Bioethanol Production Development In the last few years technologies breakthrough has compelled us for an alternative feedstock due to considerable shortage in agricultural land. In this sense, advances in metabolic pathway engineering/genetic engineering have led to the development of microbes skilled enough to convert biomass into ethanol (Das Neves et al., 2007). Generally, such development depends on expansion of the substrate range and inclusion of other biomass sources like arabinose or xylose in strains that cannot ferment sugars other than glucose. Examples of such microorganisms include genetically modified Escherichia coli, Saccharomyces sp., and Zymomonas mobilis, etc. (Davis et al., 2006). In cellulosic ethanol industry, aside from Pichia stipitis, natural xylose fermenting yeast, more efforts are being taken in obtaining recombinant bacterial and yeast strains that are able to ferment pentose sugars, such as arabinose and xylose. Figure 1.5 is one among the best examples depicting recombination process in microbes, where the tail end in E. coli and Klebsiella oxytoca or the front end of S. cerevisiae and Z. mobilis can be recombined for improved production of ethanol (Hagerdal et al., 2006). Disadvantages (1) (2) (3) (4) (5) (6) Use of blends above B5 not yet approved by many auto makers Lower fuel economy and power (10% lower for B100, 2% for B20) Currently more expensive B100 generally not suitable for use in low temperatures Concerns about B100’s impact on engine durability Slight increase in nitrogen oxide emissions possible in some circumstances Moreover, genetic engineering of plants is another promising area, which most likely plays a key role in biofuel industry. The latest hybrid varieties have helped us considerably in improving starch yield from energy crops. For example, 25 kg of corn contains about 15 kg of starch. In the near future, that same 25 kg may contain as much as 17 kg of starch through hybrid corn. This would result in a gain of nearly $2 million in annual income by processing the same amount of corn in a 120 million liter per year ethanol production (DOE, 2007). Bioreactors in Ethanol Production A major commitment in cost-effective lignocellulosic bioethanol production is to employ reactor systems yielding the maximal cellulose conversion with the minimal enzyme. As a result, one of the most vital parameters for the fabrication and operation of bioreactors for lignocellulosic conversion is the efficient use of the enzymes to gain high specific rates of cellulose conversion (yield of glucose attained/amount of enzymes). The maximization of the product concentration, i.e. the amount of glucose obtained per liquid volume, is also a significant parameter as well as the optimization of the volumetric productivity. When hydrolysis is carried out with biomass comprised of high cellulose levels, the product concentration will drive up. For this reason, few researchers are attempting the enzymatic biomass conversion with high biomass loads (Jorgensen et al., 2007). The most imperative FIGURE 1.5 Strains that can be metabolically engineered for ethanol production. Source: Hagerdal et al., 2006. (For color version of this figure, the reader is referred to the online version of this book.)
9 BIODIESEL difficulty in high biomass loads is related to the viscosity of reaction mixture, which also influences the rheology of the mixture. In particular, mixing and mass transfer limitations and presumably increased inhibition by intermediates come into play. A variety of fed-batch strategies have been adopted with the scope of supplying the substrate without reaching excessive viscosities and unproductive enzyme binding to the substrate (Rudolf et al., 2005). General criteria in bioreactor design and in the choice of the operating conditions could be use of bioreactors or reaction regimes that allow a rapid decrease in the glucose concentration; running of the reactions at low to medium substrate concentrations in order to maintain higher conversion rates and thus obtain higher volumetric output of the reactor (Andric et al., 2010). The combination of the bioreactor with a separation unit has obtained prospective results with product inhibited or equilibrium limited enzyme-mediated conversions, because it potentially removes the products as they are accumulated (Gan et al., 2002). In this regard, membrane bioreactors could be a feasible process configuration. Unlike the Solid State Fermentation (SSF) approach in which the glucose consumption is carried out by the microbes simultaneously accessible in the hydrolyzate, the use of membrane bioreactors would finish the same function without any compromise in the reaction parameters. A membrane bioreactor (Figure 1.6) is a multitasking reactor that combines the reaction with a separation, namely, in this case the product was taken away by membrane separation, as one integrated unit (in situ removal) or alternatively in two or more separate units. The membrane bioreactors used for this separation processes are mainly ultra- and nanofiltration types (Pinelo et al., 2009). However, the use of this technology is restricted by the accumulation of unreacted lignocellulosics in large level and/or continuous processing (Andric et al., 2010). Already in the past, few scientists enhanced the efficiency of the continuous stirred tank bioreactor by incorporating membrane separation technologies during the reactor design. Recently, an advanced reactor system was intended that removes the reducing sugars during the enzymatic hydrolysis of cellulose through a system consisting in a tubular reactor, in which the substrate was retained with a porous filter at the bottom and buffer entered at the top through a distributor (Yang et al., 2006). This hollow fiber ultrafiltration module with polysulfone membrane enabled the permeation and the separation of the sugars. To keep the volume constant in the tubular reactor, the entire buffer was recycled back from the ultrafiltration membrane and the makeup buffer was continuously supplied from the reservoir. In some applications an additional microfiltration unit has exceptionally been used to retain the unconverted lignin-rich solid fraction due to the presence of firmly bound enzymes or has been employed to remove the unconverted substrate from the reactor. These setups result in slightly complex process layouts for the hydrolysis (Knutsen and Davis, 2004). It is obvious that the optimization of the reactor designs will allow overcoming both the rheological and inhibition limit of the bioconversion and maximizing the enzymatic conversion. Therefore, the reactor design becomes more relevant for large-scale processing of cellulosic biomass. Immobilization of Cells for Ethanol Production For bioreactor application, immobilization of cells is a technique that has proved augmented ethanol productivity, operation stability and easier downstream processing, compared to processes using suspended cells (Das Neves et al., 2007). However, the specific advantages of immobilized cells depend on the nature of cells, reactor design and nature of the process. Entrapment of cells in natural polymers by ionic gelation (alginate) or by thermal precipitation (carrageenan and agar) is a method commonly used for cell immobilization (Ogbonna et al., 1991). Immobilization by passive adhesion to surfaces has great potential for industrial application since the immobilization method is relatively simple. The use of cheap carriers ensures that this method can be exploited with minimal increase in the overall production cost. Thus, one limiting factor of this technology is that it can only be adapted for practical industrial production if the expected increase in bioethanol productivity can overcome the increase in the production costs (cost of the carrier and immobilization) (Ogbonna et al., 1996). BIODIESEL FIGURE 1.6 Schematic of membrane bioreactor integrated with membrane distillation (MD) process for alcohol distillation. Source: Gryta, 2012. (For color version of this figure, the reader is referred to the online version of this book.) Biodiesel is a form of diesel fuel manufactured from vegetable oils, animal fats, or recycled restaurant greases. It is safe, biodegradable, and produces less air
10 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES pollutants than petroleum-based diesel. Biodiesel can be used in its pure form (B100) or blended with petroleum diesel. Common blends include B2 (2% biodiesel), B5, and B20. Biodiesel is an ideal biofuel contender that eventually could replace petroleum based diesel. Currently, biodiesel production is still too costly to be commercialized. Due to the static cost associated with oil extraction and biodiesel processing and the variability in biomass production, future cost-saving efforts for biodiesel production should focus on the production of oil-rich feedstocks like microalgae, nonedible oils, etc. As discussed above, biodiesel is costlier than conventional diesel fuel, although it is rarely quoted as being competitive, as it will be if existing fluctuations in feedstocks/product prices are favorable. Using the distribution of these prices over the last 20 years, less than 5% of costebenefit analyses based on fixed prices over the project life will show a positive result in producing biodiesel. If the feedstocks/product prices are varied each year, as will be the case in reality, biodiesel production will always be more expensive than conventional diesel (Duncan, 2003). Feedstocks for Biodiesel Biodiesel can be made from any oil/lipid source; the major components of these sources are triacylglycerol molecules. In general, biodiesel feedstocks can be categorized into three groups: pure vegetable oils, animal fats, and waste cooking oils. TABLE 1.4 Biodiesel from Pure Vegetable Oil The first group is pure oils derived from various crops and plants such as soybean, canola (rapeseed), corn, cottonseed, flax, sunflower, peanut, and palm. These are the most widely used feedstocks by commercial biodiesel producers. The oil composition from vegetable crops is pure; this cuts down on preprocessing steps and makes for a more consistent quality of biodiesel product. However, there is an obvious disadvantage for vegetable oils as biodiesel feedstocks: wide scale production of crops for biodiesel feedstocks can cause an increase in worldwide food and commodity prices. Such a “food vs fuels” debate has reached national attention when using vegetable oils for biodiesel production. Alternative feedstocks usually arise out of necessity from regions of the world where the above materials are not locally available or as part of a concerted attempt to reduce reliance on imported petroleum. JATROPHA CURCAS (JATROPHA) The nonedible oil from Jatropha curcas (Jatropha) has recently attracted extensive attention as a feedstock for biodiesel production in India and other climatically parallel regions of the world (Kumartiwari et al., 2007; Kalbande et al., 2008). The Jatropha tree is a perennial shrub belonging to the Euphorbiaceae family whose seeds contain up to 30 wt% oil. This plant can be found in tropical and subtropical regions such as Africa, Indian subcontinent, Central America, and other countries of Asia. Since Jatropha oil contains a relatively elevated percentage of saturated fatty acids (Table 1.4), the corresponding methyl esters display relatively poor low Biodiesel Production from Feedstocks High in Free Fatty Acids Feedstock FFA (wt%) Pretreatment Catalyst for Transesterification Yield (wt%) References Pongamia pinnata Up to 20 H2SO4 KOH 97 Naik et al. (2008) Jatropha curcas 14/<1 H2SO4 KOH 99þ Kumartiwari et al. (2007) Madhuca indica 20 None Pseudomonas cepacia 96þ** Kumari et al. (2007) Nicotiana tabacum 35/<2 H2SO4 KOH 91 Veljkovic et al. (2006) Calophyllum inophyllum 22/<2 H2SO4 KOH 85 Sahoo et al. (2007) Zanthoxylum bungeanum 45.5/1.16* None H2SO4 98 Zhang and Jiang (2008) Brown grease 40/<1 Diarylammonium catalysts NaOCH3 98þ** Ngo et al. (2008) Waste cooking oil 7.25/<1* H2SO4 NaOH 90** Meng et al. (2008) Waste fryer grease 5.6 H2SO4 KOH 90þ Issariyakul et al. (2007) Sorghum bug oil 10.5 None H2SO4 77e94 Mariod et al. (2006) * Acid value (mg KOH/g) was given instead of FFA. ** Conversion to esters (wt%) is provided instead of yield.
BIODIESEL temperature operability, as evidenced by pour point (PP) value of 2  C (Kumartiwari et al., 2007). PONGAMIA PINNATA (KARANJA) Another nonedible biomass originated in India is Pongamia pinnata (Karanja), which is a medium-sized deciduous plant that grows fast in damp and subtropical environments and matures in 5e7 years to tender fruit that contains two kidney-shaped kernels (Mohibbeazam et al., 2005). The oil content of Karanja kernels ranges between 25 wt% and 40 wt% (Karmee et al., 2005; Mohibbeazam et al., 2005). The primary fatty acid found in Karanja oil is oleic acid (45e70 wt%), followed by palmitic, linoleic, and stearic acids (Karmee et al., 2005; Naik et al., 2008). The low-temperature operability of the parallel methyl esters from karanja is superior to that of jatropha oil methyl esters as a result of the fairly high percentage of oleic acid in karanja oil, as evidenced by cloud point (CP) and PP values of 2  C and 6  C, respectively (Srivastava and Verma, 2008). MADHUCA INDICA (MAHUA) Madhuca indica, commonly known as “Mahua”, is a tropical plant found frequently in the central and northern plains and forests of India. It belongs to the family Sapotaceae and grows rapidly up to 20 m in height, possesses evergreen or semievergreen foliage, and is well adapted to dry environments (Ghadge and Raheman, 2006; Kumari et al., 2007). The fruit is nonedible, obtained from the tree in 4e7 years and contains one to two kidney-shaped kernels (Mohibbeazam et al., 2005). The oil content of dried Mahua seeds is about 50 wt%. Mahua oil is characterized by free fatty acid (FFA) content of around 20 wt% and a comparatively high percentage of saturated fatty acids such as stearic (14.0 wt%) and palmitic (17.8 wt%) acids (Ghadge and Raheman, 11 2006). The remaining fatty acids are mostly spread among unsaturated components such as linoleic (17.9 wt%) and oleic (46.3 wt%) acids (Singh and Singh, 1991). The relatively high percentage of saturated fatty acids (35.8 wt%) found in Mahua oil results in relatively poor low-temperature properties of the parallel methyl esters, as evidenced by PP value of 6  C (Ghadge and Raheman, 2006). NICOTIANA TABACUM (TOBACCO) Nicotiana tabacum, commonly referred as tobacco, is a commercial shrub with pink flowers and green capsules containing abundant small seeds grown in a large number of countries around the world. The foliage of the plant is the commercial product and used in the preparation of cigarettes and other tobacco-containing products. The oil content of the seeds, a by-product from tobacco, ranges from 36 wt% to 41 wt% (Usta, 2005). This tobacco seed oil contains more than 17 wt% FFAs (Veljkovic et al., 2006) and is high in linoleic acid (69.5 wt%), along with oleic (14.5 wt%) and palmitic (11.0 wt%) acid in significant amounts. Due to high linoleic acid content of tobacco seed oil, the corresponding methyl esters display relatively low kinematic viscosity (3.5 mm2/s) in comparison to most other biodiesel fuels (Usta, 2005). Biodiesel from Animal Fat Wastes The feedstock issues are very critical, which affect the economic potential of biodiesel production, since feedstock accounts around 75% of the biodiesel total cost (Figure 1.7). Recently, alternative lipid residues such as waste frying oil and nonedible animal fats have also received substantial attention from the biofuel sector. To take benefit of these low-cost and low-quality resources, a suitable act would be to reuse residues in FIGURE 1.7 Biodiesel production cost summary sheet. Source: Pruszko, 2007. (For color version of this figure, the reader is referred to the online version of this book.)
12 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES order to integrate sustainable energy supply and waste management in food processing facilities. Animal fats are typically considered as waste by-products and less expensive than commodity vegetable oils, which make them attractive as feedstock for biodiesel production. These animal wastes are collected from chicken, cow, pork lard, and other animals such as fish and insects. BEEF TALLOW AND CHICKEN FAT Animal fats like beef tallow and chicken fat are byproducts from the meat industry and stand for cheap feedstock for biodiesel production. The key fatty acids found in beef tallow were oleic (47.2 wt%), palmitic (23.8 wt%), and stearic (12.7 wt%) acids. The prime fatty acids contained in chicken fat include oleic (40.9 wt%), palmitic (20.9 wt%), and linoleic (20.5 wt%) acids (Wyatt et al., 2005). Due to very low concentration of polyunsaturated fatty acid in beef tallow, the corresponding methyl esters illustrate excellent oxidative stability, as evidenced by an oil stability index (OSI) value of 69 h at 110  C. In addition, other physical properties of beef tallow methyl esters include kinematic viscosity (40  C) of 5.0 mm2/s, a flash point (FP) of 150  C, and CP, PP and cold filter plugging point (CFPP) values of 11, 13, and 8  C respectively (Moser, 2009). In chicken fat, due to high polyunsaturated fatty acid content, the corresponding methyl esters display poor oxidative stability, as evidenced by an OSI value of 3.5 h at 110  C. Burning the B20 blends of beef tallow and chicken fat methyl esters results in NOx exhaust emissions of only 2.4% versus 6.2% of B20 blend of soybean methyl esters (SME) (Wyatt et al., 2005). PORK LARD Pork lard is a by-product of the food industry and symbolizes a low-cost feedstock for biodiesel production. The main fatty acids in pork lard includes stearic (121 wt%), linoleic (127 wt%), oleic (44.7 wt%), and palmitic (26.4 wt%) acids (Jeong et al., 2009). Due to high saturated fatty acid content in pork lard, the corresponding methyl esters exhibit quite high CFPP value of 8  C and a relatively low iodine value (IV) of 72, along with a typical kinematic viscosity (40  C) of 4.2 mm2/s. Another study determined that pork lard methyl esters have a kinematic viscosity (40  C) of 4.8 mm2/s, FP of 160  C, OSI value of 18.4 h at 110  C, and CP, PP, and CFPP values of 11, 13, and 8  C, respectively (Wyatt et al., 2005). Furthermore, combustion of B20 blends of pork lard methyl esters results in NOx exhaust emissions of only 3.0% versus 6.2% for a B20 blend of SME. Other Waste Cooking Oils Waste oils may include a variety of low-worth materials such as used cooking or frying oils, acid oils, tall oil, vegetable oil soapstocks, and other waste materials. Waste oils are usually characterized by relatively high FFA and water contents and potentially contain a variety of solid materials that must be removed by filtration prior to conversion to biodiesel (Moser, 2009). In the case of used cooking or frying oils, hydrogenation to increase the useful cooking lifetime of the oil may result in the introduction of relatively high-melting trans constituents, which influence the physical properties of the resulting biodiesel. Used frying or cooking oil is mainly acquired from restaurants and may cost between free to 50% less expensive than commodity vegetable oils, depending on the source and the availability (Predojevic, 2008). The physical properties of methyl esters prepared from used cooking or frying oils include kinematic viscosities (40  C) of 4.23 (Meng et al., 2008), 4.79, and 4.89 mm2/s; FP of 171  C; cetane number of 55, IV of 125, CFPP values of 1 and 6  C (Cetinkaya and Karaosmanoglu, 2004), CP values of 9 and 3  C, and PP values of 3 and 0  C (Phan and Phan, 2008). The disparities in the physical property data among the various studies may be a result of feedstock origin or due to differences in product purity. Algae as a Biodiesel Source Algae can also be used to produce energy in a number of ways. One of the most competent ways is through exploitation of the algal oils to produce biodiesel. Algal biomass contains three major components: carbohydrates, proteins, and lipids/natural oils (Dunahay et al., 1996). Because the natural oil made by microalgae is in the form of triacylglycerol molecule, which is the right kind of oil for producing biodiesel, microalgae are the exclusive focus in the algae to biofuel arena. Actual biodiesel yield per hectare is about 80% of the yield of the parent crop oil given in Table 1.5. In view of Table 1.5, microalgae emerged to be the only source of biodiesel that has the potential to completely replace petroleum diesel. Unlike other oil crops, microalgae grow extremely rapidly and many are exceedingly rich in oil. Microalgae commonly double their biomass within 24 h. Biomass doubling times during exponential growth are commonly as short as 3e4 h. Oil content in microalgae can exceed 70% by weight of dry biomass (Metting, 1996; Spolaore et al., 2006). Oil levels up to 50% are quite common. Oil productivity, the mass of oil produced per unit volume of the microalgal broth per day, depends on the algal growth rate and the oil content of the biomass. Microalgae with high oil productivities are desired for producing biodiesel. CHEMICAL TRANSESTERIFICATION PROCESS FOR BIODIESEL PRODUCTION The source oil used in making biodiesel consists of triglycerides (Figure 1.7), in which three fatty acid
BIODIESEL TABLE 1.5 Comparison between Few Biodiesel Sources Crop Oil Yield (l/hectare)*** Land Required (M he) Corn 172 1540 Soybean 446 594 Jatropha 1892 140 Coconut 2698 99 Oil palm 5950 45 Microalgae* 136,900 2 Microalgae** 58,700 4.5 * About 70% oil (by wt) in biomass. ** About 30% oil (by wt) in biomass. *** For meeting 50% of all transport fuel needs of the United States. molecules are esterified with a molecule of glycerol. In biodiesel production, triglycerides are reacted with methanol in a reaction known as transesterification or alcoholysis. Transesterification produces methyl esters of fatty acids that are biodiesel and glycerol (Figure 1.7). The reaction occurs stepwise: triglycerides are first converted to diglycerides, then to monoglycerides and finally to glycerol. At equilibrium, transesterification needs 3 mol of alcohol for every mole of triglyceride to produce 1 mol of glycerol and 3 mol of methyl esters (Figure 1.8). Industrial processes use 6 mol of methanol for each mole of triglyceride (Fukuda et al., 2001). This large excess of methanol ensures that the reaction is driven in the direction of methyl esters, i.e. toward biodiesel. Yield of methyl esters exceeds 98% on a weight basis. Transesterification is catalyzed by acids and alkalis (Fukuda et al., 2001). Alkali-catalyzed transesterification is about 4000 times quicker than the acid-catalyzed reaction. Thus, alkalis such as sodium and potassium hydroxide are frequently used as commercial catalysts at a concentration of about 1% by weight of oil. Alkoxides such as sodium methoxide (CH3ONa) act like better catalysts than sodium hydroxide and are being increasingly used. Use of lipases offers significant advantages, but it is currently not feasible because of the relatively high cost of the catalyst (Chisti, 2007). Alkali-catalyzed transesterification is carried out at about 60  C under one atmospheric pressure, as methanol boils off at 65  C at atmospheric pressure. Under these conditions, reaction takes about 90 min to complete (Meher et al., FIGURE 1.8 Transesterification of oil to biodiesel. 13 2006). A higher temperature can be used in combination with higher pressure, but the process becomes expensive. During reaction, methanol and oil do not mix; hence, the reaction mixture shows two liquid phases. Other alcohols can be used, but methanol is the least expensive. To stop yield loss due to saponification reactions (soap formation), the oil and alcohol must be dry and the oil should have a least of FFAs. Biodiesel is recovered by repeated washing with water to remove glycerol and methanol. ENZYMATIC TRANSESTERIFICATION PROCESS FOR BIODIESEL PRODUCTION Lipases (triacylglycerol hydrolase, EC 3.1.1.3.) are enzymes that catalyze the breakdown of carboxylic ester link in the triacylglycerol molecule to form FFAs, diand monoglycerides and glycerol. Although their purpose is to catalyze hydrolysis of ester links, they can also catalyze the esterification, the conception of this link between alcohol hydroxyl groups and carboxyl groups of carboxylic acids. Therefore, they can catalyze hydrolysis, alcoholysis, esterification and transesterification and they have a wide spectrum of biotechnological applications (Kirk et al., 2002). Lipases are also highly specific as regio, chemo and enantioselective catalysts. Thanks to protein engineering, it is possible to enhance catalytic potential of lipases and “tailor” them to exact application and process situation, enabling further expansion of their industrial applications (B van Beilen and Li, 2002). Among lipases from animal, plant and microbial origins, the most commonly used are microbial lipases. They have abundant advantages over lipases from animal and plant sources. Using microbes it is possible to achieve a higher yield of enzymes, and to genetically control the strain in obtaining a lowcost lipase with preferred properties for the conversion of fats and oils into biodiesel. In addition, the enzymatic yield is independent of potential seasonal variations and it is possible to achieve rapid growth of microbes in lowcost media (Gupta et al., 2004). Bioreactors for Biodiesel Production Microalgae are unicellular microscopic organisms, like simple plants with no leaves and roots that grow through photosynthesis process. They capture carbon dioxide during photosynthesis and convert it into feedstock that can be used as food, fertilizer, a source of medicine and biodiesel (Chojnacka and Marquez-Rochaet, 2004). Growing algae in open pond system raise several concerns such as impossibility to control growth settings and contamination threats. Algal cells in open ponds are exposed to the environment, light deficiency, subject to risk of contamination, and heterogenous medium
14 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES Exhaust gas Culture medium Harvest Airlift system Air CO2 Samples Solar receiver FIGURE 1.9 Schematic of tubular photobioreactor with airlift system. Source: Molina et al., 2001. depending upon the mixing mechanism, the shape of the ponds and the depth of the pond (Chojnacka and Marquez-Rochaet, 2004). On the other hand, closed ponds (photobioreactors) mitigate fluid culture contamination, and enhance full control over algal growth parameters such as homogenous culture, pH, light penetration, and carbon dioxide input. They would use less space with high algal biomass yield. However, they are costly to build and maintain (Mulumba and Farag, 2012). Design of tubular photobioreactor (TPBR) for algal cell growth was depicted in Figure 1.9. It has a main tank connected to two spiral tubes set in sequence. Both spiral parts were clear polyvinyl chloride (PVC) tubes of 100 external diameter and 3/400 internal diameter. The capacity of both spiral parts was 3.4 gallons. The main tank served as a feeding point of medium to the PVC tubes with a maximum capacity of 5 gallons (Chisti, 2007). Culture medium was pumped into the tubings at a fixed flow rate. These tubes provided an area of 20 ft2 exposed to the fluorescence light. Air compressor supplies air to the TABLE 1.6 system for aeration and to serve as a source of carbon dioxide (CO2). The air flow rate was set in the ranges of 190e210 gallons/h. In TPBR, selected algal strain was cultured using fresh medium with no modification. Algal growth and pH were measured over a period of time varying between 12 days and 14 days. A sample was taken every 2 days to quantify the turbidity using a spectrophotometer at 682 nm and cell counts were performed using a microscope. The pH of culture was measured using pH test strips. The selected algal strain shows the typical growth curve of other microbes, which include lag, exponential or log, stationary and lytic phases. The length of each phase depends on light penetration, nutrients concentration, mixing mechanism, and the solubility of oxygen in medium. After reaching a stationary or lysis phase, algal culture was harvested by centrifugation followed by lyophilization to produce dry algal feedstock. Crude lipid from dried algal biomass was extracted using either modified Folch method (Cooksey et al., 1987) or Soxhlet extractor (Mulumba, 2010; Chojnacka and Marquez-Rochaet, 2004). In both methods, polar and nonpolar solvents such as methanol and chloroform/ hexane were used (Table 1.6). The combination of polar and nonpolar solvents enhances the extraction of both polar and nonpolar lipid. BIOGAS Biogas is obtained by anaerobic digestion (AD) of organic materials, which occurs inside the anaerobic biodigester. Chemical composition of this biogas depends on several parameters, such as type of digester employed, the kind of organic material and the constancy of the feeding process of the biodigester. The most significant biogas components are methane (CH4), carbon dioxide (CO2) and sulfuric components (H2S). The composition of biogas is a crucial parameter, Biodiesel Production with Various Lipases Lipase Source Acyl Acceptor Solvent Yield (%) References Candida antarctica B Waste cooking palm oil Methanol tert-butanol 79.1 Halim et al. (2009) Thermomyces lanuginosus Soybean oil Ethanol n-hexane/solvent free 70e100 Rodrigues et al. (2010) Pseudomonas fluorescens, Candida rugosa Jatropha seed oil Ethanol Solvent free 98 Shah and Gupta (2007) Rhizomucor miehei, Penicillium cyclopium Soybean oil Methanol Solvent free 68e95 Guan et al. 2010 Candida antarctica Sunflower oil Methyl acetate Solvent free >95 Ognjanovic et al. (2009) Thermomyces lanuginosus Rapeseed oil Methanol Solvent free 95 Li et al. (2006) Candida antarctica Jatropha seed oil, karanja oil Ethyl acetate Solvent free >90 Modi et al. (2007)
15 BIOGAS because it allows identifying the suitable purification system, which aims to remove sulfuric gases and reduce the water volume, contributing to recover the combustion fuel conditions (Boe et al., 2007). Other important data collected from biogas analysis is referent to the low heat value, that combined to the efficiency and biogas consumption is important to estimate the electric generation potential. However, biogas production is much variable because it depends on several parameters, such as the kind of organic material (Liu et al., 2004). Biogas production involves three steps: fermentation, which includes hydrolysis and acid genesis, acetone genesis and methane genesis. In the fermentation process, during the hydrolysis the organic material is converted into smaller molecules and this material is transformed in soluble acids by acidogenese. Next step is acetanogenese process, transforming the products obtained in the first step into acetic acid, hydrogen and carbon dioxide. The last step is referent to metanogenese process, producing methane gas through anaerobic bacteria (Figure 1.10) (Seadi et al., 2008; Boe et al., 2007). Biogas Feedstock Awide range of biomass types can be used as substrates for the production of biogas by AD. The most common biomass categories used in biogas production are listed below and in Table 1.7. Animal manure and slurry, agricultural residues and by-products, digestible organic wastes from food and agro industries, organic fraction of municipal waste and from catering, and sewage sludge, etc. are best study sources for biogas production. Recently, various novel feed stocks has been tested and introduced for biogas synthesis in many countries, the dedicated energy crops (DECs), crops grown specifically for energy and biogas production. DECs can be herbaceous (grass, maize, and raps) and also woody crops (willow, poplar, and oak), although the woody crops need particular delignification treatment before AD. In AD, substrates can be classified according to the following criteria: methane yield, origin, dry matter (DM) content, etc. Table 1.7 gives a summary on the Carbohydrate Sugar Household Digesters for Biogas It is difficult to accept one particular type of digester for household biogas production. The design of the digesters is diversified based on the availability of substrate, geographical location, and climatic conditions. For example, a digester designed in mountainous regions has less gas volume in order to avoid gas loss. For tropical countries, it is recommended to have digesters underground due to the geothermal energy (Bin, 1989). Of all the different digesters developed, the fixed dome model developed in China and the floating drum model developed in India sustained to perform well until today (Rajendran et al., 2012). Recently, plug flow digesters are gaining attention due to its portability and easy operation. Fixed Dome Digesters The fixed dome digesters (Figure 1.11) is also called “hydraulic” or “Chinese” digesters and it is the most frequent model developed and used in China for biogas production (Rajendran et al., 2012). In this case, digester is filled through the inlet pipe until the level reaches the base level of the expansion chamber. Biogas that is produced is accumulated at the upper part of the digester called storage part. The difference in the levels between the slurry inside the digester and the expansion chamber develops pressure inside due to accumulation of biogas. This accumulated biogas requires space and presses the substrate apart and enters into the expansion chamber. The slurry flows back into the digester straight away after Carbon acids, alcohols Fats Fatty acids Proteins Amino acids Hydrolysis characteristics of some digestible feedstocks. Substrates with DM content less than 20% are used for what is called wet digestion (wet fermentation), which includes animal slurries and manure as well as various wet organic wastes from food industries. When the DM content is as high as 35%, it is called dry digestion (dry fermentation), and it is typical for energy crops and silages. The choice of types and amounts of feedstock for the AD substrate mixture depends on their DM content as well as the content of sugars, lipids and proteins. Acidogenesis Hydrogen carbon dioxide ammonia Acetic acid hydrogen carbon dioxide Acetogenesis Methane carbon dioxide Methanogenesis FIGURE 1.10 Biochemical process in anaerobic digester. (For color version of this figure, the reader is referred to the online version of this book.)
16 TABLE 1.7 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES Characteristics of Some Digestible Feedstocks Type of Feedstock Organic Content C:N Ratio DM (%) VS% of DM Biogas Yield (m3/kg VS*) Unwanted Physical Impurities Cattle slurry Carbohydrate, protein, lipid 6e20 5e12 80 0.20e0.30 Bristles, soil, water, straw, wood Poultry slurry Carbohydrate, protein, lipid 3e10 10e30 80 0.35e0.60 Grit, sand, feathers Stomach/intestine content Carbohydrate, protein, lipid 3e5 15 80 0.40e0.68 Animal tissues Concentrated whey 75e80% lactose 20e25% protein e 20e25 90 0.80e0.95 Transportation impurities Flotation sludge 65e70% proteins 30e35% lipids e e e e Animal tissues Straw Carbohydrates, lipids 80e100 70e90 80e90 0.15e0.35 Sand, grit Food remains Carbohydrate, protein, lipid e 10 80 0.50e0.60 Bones, plastic * VS - volatile solids Source: Seadi et al., 2008. Floating Drum Digesters gas is released (Adeoti et al., 2000). Fixed dome digesters are usually built underground and the size of the digester depends on the place, number of households, and the amount of substrate available every day. Generally size of these digesters normally varies between 5 m3 and 150 m3 in various parts of Asia (Tomar, 1994). Instead of having a digester for each home, a large-volume digester is used to produce biogas for 10 to 20 homes, and is called community-type biogas digesters. In countries where houses are clustered as in Africa, these types of biogas digesters are more viable (Adeoti et al., 2000). The floating drum digester was first time constructed by Khadi and Village Industries Commission and this model was developed in 1962 (Figure 1.12). Although the model is old, it is one of the most extensively used designs for household purposes in India. This design includes a movable inverted drum placed on a wellshaped digester. The inverted steel drum acts as a storage tank, which can move up and down depending on the quantity of accumulated biogas at the top of the digester. The weight of this inverted drum applies the Outlet for bio-gas Slurry of cattle dung and water S Mixing tank Slab cover Dome M V Gas valve Slab cover D Ground level Overflow tank F Bio-gas I Inlet chamber Spent slurry Dung and water mixture T O Outlet chamber Underground digester tank FIGURE 1.11 Schematic sketch of fixed dome digester. Source: GMI, India, 2013. (For color version of this figure, the reader is referred to the online version of this book.)
17 CONCLUSION Gas control valve Gas stove FIGURE 1.12 Floating drum digester. Source: Working of biogas plants Working of biogas plants, 2013, www.tutorvista.com. (For color version of this figure, the reader is referred to the online version of this book.) Gas holder Mixing tank Over flow tank Inlet pipe Outlet pipe Digester tank Inlet tank Partition wall pressure needed for biogas flow through the pipeline (Singh and Sooch, 2004). Floating drum digesters manufacture biogas at a stable pressure with variable volume. In floating drum reactor, by position of the drum, the amount of biogas accumulated under the drum is easily noticeable. However, the floating drum needs to be coated with paint at regular intervals to avoid rusting. Additionally, fibrous materials in biomass will block the movement of the digester. Hence, their accumulation must be avoided if possible (Adeoti et al., 2000). In Thailand, the floating dome has been customized with two cement jars on each side of the floating drum. The average size of these digesters is around 1.2 m3 (Gosling, 1982). For small and medium-size farms the size varies from around 5 to 15 m3. Singh and Singh (1991) compared 14 different biogas plants with a floating drum model and optimized the various parameters for maximum biogas production. Social and Environmental Aspects of Biogas Digesters Change in the global climate is a major threat that the world is facing today. The nonrenewable energy consumption in the past has led to global warming that needs to be addressed (Bilen et al., 2008). The household digesters could reduce the pressure on the environment by dropping deforestation and GHG emissions followed by loss of cultivable land, and soil erosion (Gautam et al., 2009). Biogas production in rural areas can partly reduce global warming (Pei-dong et al., 2007). By using biogas in rural households economical, environmental, and social benefits were achieved (Yang et al., 2011). Even though both carbon dioxide and methane are major contributors to the greenhouse effect, the global warming effect of methane is 21 times greater than that of carbon dioxide (Dhingra et al., 2011). However, houses equipped with biogas systems exhibit leakage of gases in the biogas systems. Fortunately, the households with biogas plants have 48% less emissions compared to households without biogas systems (Pathak et al., 2009). It is worth talking about 10% of households, which had methane leakage (Yang et al., 2011). Research has already shown that by replacing firewood and coal with biogas, the emission of CO2 and SO2 would be reduced by 4193 thousand tons, and 62.0 thousand tons, respectively (Pei-dong et al., 2007). CONCLUSION We conclude that by accelerating research in areas of bioenergy, we can make significant contributions to sustainable development and use of feedstock. We must realize that by maximizing biomass conversion efficiency, we can minimize raw material requirements, while at the same time the financial position of various market sectors (e.g. energy, agriculture, and forestry) are strengthened. There is an international agreement on the fact that the feedstock accessibility is inadequate so that the raw materials should be used as competently as possible, i.e. expansion of multipurpose industries (biorefineries) that can utilize variable biomass sources as raw materials for bioenergy production. The main constraint in making this biorefinery a successful path is bringing the stakeholders together, who normally operate in different market sectors (e.g. energy, agriculture and forestry, fuel transportation, etc.). Above all, the government should make policies to help overcome the threshold by dropping production costs in the form of feedstock in tariffs, feedstock
18 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES in premiums, tax exemptions, etc. These encouragements can be targeted at different parts of the supply chain like feedstock producers, energy producers, and distributors. References Adeoti, O., Ilori, M.O., Oyebisi, T.O., Adekoya, L.O., 2000. Engineering design and economic evaluation of a family-sized biogas project in Nigeria. Technovation 20, 103e108. Adler, P.R., Sanderson, M.A., Weimer, P.J., Vogel, K.P., 2009. Plant species composition and biofuel yields of conservation grasslands. Ecol. Appl. 19, 2202e2209. Agarwal, A.K., 2007. Biofuels (alcohols and biodiesel) applications as fuels for internal combustion engines. Prog. Energy Combust. Sci. 33, 233e271. Alvira, P., Tomas, P.E., Ballesteros, M., Negro, M.J., 2010. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour. Technol. 101, 4851e4861. Andric, P., Meyer, A.S., Jensen, P.A., Johansen, K.D., 2010. Reactor design for minimizing product inhibition during enzymatic lignocellulose hydrolysis II. Quantification of inhibition and suitability of membrane reactors. Biotechnol. Adv. 28 (3), 407e425. Balat, M., 2007. An overview of biofuels and policies in the European Union countries. Energy Sources Part B 2, 167e181. Bilen, K., Ozyurt, O., Bakırc, K., Karsl, S., Erdogan, S., Yılmaz, M., Comakl, O., 2008. Energy production, consumption, and environmental pollution for sustainable development: a case study in Turkey. Renewable Sustainable Energy Rev. 12, 1529e1561. Bin, C., 1989. The current status of agricultural geothermal utilization in China. Biomass 20, 69e76. Bochek, A.M., 2003. Effect of hydrogen bonding on cellulose solubility in aqueous and nonaqueous solvents. Russ. J. Appl. Chem. 76 (11), 1711e1719. Boe, K., Steyer, J.P., Angelidaki, I. 2007. Monitoring and Control of the Biogas Process Based on Propionate Concentration Using Online VFA Measurement. Presented at 11th IWA World Congress on Anaerobic Digestion, 23e27 September 2007, Brisbane, Australia. Boer, G.J., Flato, G., Reader, M.C., Ramsden, D., 2000. A transient climate change simulation with greenhouse gas and aerosol forcing: experimental design and comparison with the instrumental record for the 20th century. Clim. Dyn. 16, 405e425. Bohlmann, G.M., 2006. Process economic considerations for production of ethanol from biomass feedstocks. Ind. Biotechnol. 2, 14e20. British Petroleum Statistical Review, 2011. BP statistical review of world energy. Available at. http://www.bp.com/en/global/ corporate/about-bp/statistical-review-of-world-energy-2013/ statistical-review-1951-2011.html (Accessed on 10.03.2013). Broder, J.D., Barrier, J.W., Lightsey, G.R., 1992. Conversion of cotton trash and other residues to liquid fuel from renewable resources. In: Cundiff, J.S. (Ed.), Proceedings of an Alternative Energy Conference. American Society of Agricultural Engineers, St. Joseph, MI, pp. 189e200. Cara, C., Ruiz, E., Mercedes, B., Paloma, M., Ma, J.N., Eulogio, C., 2008. Production of fuel ethanol from steam-explosion pretreated olive tree pruning. Fuel 87 (6), 692e700. Cetinkaya, M., Karaosmanoglu, F., 2004. Optimization of basecatalyzed transesterification reaction of used cooking oil. Energy Fuel 18, 1888e1895. Chang, V.S., Kaar, W.E., Burr, B., Holtzapple, M.T., 2001. Simultaneous saccharification and fermentation of lime treated biomass. Biotechnol. Lett. 23 (6), 1327e1333. Chen, M., Zhao, J., Xia, L., 2009. Comparison of four different chemical pretreatments of corn stover for enhancing enzymatic digestibility. Biomass Bioenergy 33, 1381e1385. Chisti, Y., 2007. Biodiesel from microalgae. Biotechnol. Adv. 25, 294e306. Chojnacka, K., Marquez-Rocha, F.J., 2004. Kinetic and stoichiometric relationships of the energy and carbon metabolism in the culture of microalgae. Biotechnology 3, 21e34. Converse, A.O., Kwarteng, I.K., Grethlein, H.E., Ooshima, H., 1989. Kinetics of thermochemical pretreatment of lignocellulosic materials. Appl. Biochem. Biotechnol. 20 (21), 63e78. Cooksey, K.E., Guckert, J.B., Williams, S.A., Callis, P.R., 1987. Fluorometric determination of the neutral lipid content of microalgal cells using nile red. J. Microbiol. Methods 6, 333e345. Coyne, J., Gupta, V.K., O’Donovan, A., Tuohy, M.G., 2013. The role of fungal enzymes in global biofuel production technologies. In: Gupta, V.K., Tuohy, M.G. (Eds.), Biofuel Technologies. Springer Science Publishers, USA, pp. 121e143. Das Neves, M.A., Kimura, T., Shimizu, N., Nakajima, M., 2007. State of the art and future trends of bioethanol production. In: da Silva, J.A.T. (Ed.), Dynamic Biochemical Process Biotech Molecular Biology. Global Science Books Ltd, Middlesex, UK, pp. 1e14. Davis, L., Rogers, P., Pearce, J., Peiris, P., 2006. Evaluation of Zymomonas-based ethanol production from a hydrolysed waste starch stream. Biomass Bioenergy 30, 809e814. Demirbas, A., 2008. The importance of bioethanol and biodiesel from biomass. Energy Sources, Part B 3, 177e185. Department of Energy, 2007. Office of Energy Efficiency and Renewable Energy. Washington DC, US. Available at: http://www.doe. gov. (accessed 14.03.2013). Dhingra, R., Christensen, E.R., Liu, Y., Zhong, B., Wu, C.F., Yost, M.G., Remais, J.V., 2011. Greenhouse gas emission reductions from domestic anaerobic digesters linked with sustainable sanitation in rural China. Environ. Sci. Technol. 45, 2345e2352. Dufey, A., 2006. Biofuels Production, Trade and Sustainable Development: Emerging Issues. Sustainable Markets Discussion, Paper No. 2. International Institute for Environment and Development, London. Dunahay, T.G., Jarvis, E.E., Dais, S.S., Roessler, P.G., 1996. Manipulation of microalgal lipid production using genetic engineering. Appl. Biochem. Biotechnol. 57, 223e231. Duncan, J. 2003. Costs of biodiesel production. Prepared for: Energy Efficiency and Conservation Authority: 1e26. Enguidanos, M., Soria, A., Kavalov, B., Jensen, P., 2002. Techno-Economic Analysis of Bioalcohol Production in the EU: A Short Summary for Decision-Makers. Report EUR 20280 EN. IPTS/JRC, Sevilla. Eriksson, K.E.L., 2000. Lignocellulose, lignin, ligninases. In: Schaechter, M. (Ed.), Encyclopedia of Microbiology. Academic Press, San Diego, pp. 39e48. Esteghlalian, A., Hashimoto, A.G., Fenske, J.J., Penner, M.H., 1997. Modeling and optimization of the dilute-sulfuric-acid pretreatment of corn stover, poplar and switchgrass. Bioresour. Technol. 59, 129e136. FAO, United Nations, 2013. Food and Agricultural Organization of United Nations: Economic and Social Department: The Statistical Division. Available at: http://faostat.fao.org/. (accessed 15.03.2013). Fukuda, H., Kondo, A., Noda, H., 2001. Biodiesel fuel production by transesterification of oils. J. Biosci. Bioeng. 92, 405e416. Gan, Q., Allen, S.J., Taylor, G., 2002. Design and operation of an integrated membrane reactor for enzymatic cellulose hydrolysis. Biochem. Eng. J. 12 (3), 223e229.
REFERENCES Gautam, R., Baral, S., Herat, S., 2009. Biogas as a sustainable energy source in Nepal: present status and future challenges. Renewable Sustainable Energy Rev. 13, 248e252. Ghadge, S.V., Raheman, H., 2006. Process optimization for biodiesel production from mahua (Madhuca indica L.) oil using response surface methodology. Bioresour. Technol. 97, 379e384. GMI, India, 2013. Global Methane Initiative, India. Available at: http://www.methanetomarketsindia.com/1/agri.htm. (accessed 20.03.2013). Gnansounou, E., Bedniaguine, D., Dauriat, A. 2005. Promoting bioethanol production through clean development mechanism: findings and lessons learnt from ASIATIC project. In Proceedings of the 7th IAEE European Energy Conference, 28e30 August 2005. Bergen, Norway. Gosling, D., 1982. Biogas for Thailand’s rural development: transferring the technology. Biomass 2, 309e316. Gryta, M., 2012. Desalination of industrial effluents using integrated membrane processes. In: Ning, R.Y. (Ed.), Advancing Desalination. InTech, Rijeka, Croatia, pp. 37e56. Guan, F., Peng, P., Wang, G., Yin, T., Peng, Q., Huang, J., Guan, G., Li, Y., 2010. Combination of two lipases more efficiently catalyzes methanolysis of soybean oil for biodiesel production in aqueous medium. Process Biochem. 45 (10), 1667e1682. Gupta, R., Gupta, N., Rathi, P., 2004. Bacterial lipases: an overview of production, purification and biochemical properties. Appl. Microbiol. Biotechnol. 64 (6), 763e781. Gupta, V.K., Tuohy, M.G., Sharma, G.D., 2013. Biotechnology of Trichoderma: an overview. In: Gupta, V.K., Tuohy, M.G., Sharma, G.D., Gaur, S. (Eds.), Applications of Microbial Genes in Enzyme Technology. Nova Science Publishers, USA, pp. 375e393. Hagerdal, B.H., Grauslund, G.M., Liden, G., Zacchi, G., 2006. Bioethanoldthe fuel of tomorrow from the residues of today. Trends Biotechnol. 24, 549e556. Halim, S.F.A., Kamaruddin, A.H., Fernando, W.J.N., 2009. Continuous biosynthesis of biodiesel from waste cooking palm oil in a packed bed reactor: optimization using response surface methodology (RSM) and mass transfer studies. Bioresour. Technol. 100 (2), 710e716. Howard, R.L., Abotsi, E., Jansen van Rensberg, E.L., Howard, S., 2003. Lignocellulose biotechnology: issues of bioconversion and enzyme production. Afr. J. Biotechnol. 2, 602e619. Hu, G., Heitmann, J.A., Rojas, O.J., 2008. Feedstock pretreatment strategies. Bioresources 3 (1), 270e294. Huang, H.J., Ramaswamy, S., Tschirner, U.W., Ramarao, B.V., 2008. A review of separation technologies in current and future biorefineries. Sep. Purif. Technol. 62, 1e21. Imai, M., Ikari, K., Suzuki, I., 2004. High-performance hydrolysis of cellulose using mixed cellulose species and ultrasonication pretreatment. Biochem. Eng. J. 17 (2), 79e83. International Energy Agency Statistics, 2012. CO2 Emissions from the Consumption of Coal. Available at. http://www.eia.gov/cfapps/ ipdbproject/IEDIndex3.cfm?tid¼1&pid¼1&aid¼8 (Accessed on 10.03.2013). Issariyakul, T., Kulkarmi, M.G., Dalai, A.K., Bakhshi, N.N., 2007. Production of biodiesel from waste fryer grease using mixed methanol/ethanol system. Fuel Process. Technol. 88, 429e436. Jeong, G.W., Yang, H.S., Park, D.H., 2009. Optimization of transesterification of animal fat ester using response surface methodology. Bioresour. Technol. 100, 25e30. Jorgensen, H., Vibe-Pedersen, J., Larsen, J., Felby, C., 2007. Liquefaction of lignocellulose at high solids concentrations. Biotechnol. Bioeng. 96 (5), 862e870. Kalbande, S.R., More, G.R., Nadre, R.G., 2008. Biodiesel production from non-edible oils of jatropha and karanj for utilization in electrical generator. Bioenergy Res. 2, 170e178. 19 Karmee, S.K., Chadha, A., 2005. Preparation of biodiesel from crude oil of Pongamia pinnata. Bioresour. Technol. 96 (13), 1425e1429. Karunanithy, C., Muthukumarappan, K., Julson, J.L. 2008. Influence of High Shear Bioreactor Parameters on Carbohydrate Release from Different Biomasses. American Society of Agricultural and Biological Engineers Annual International Meeting, 29 Junee2 July, 2008, Providence, Rhode Island. Kheshgi, H.S., Jain, A.K., Wuebbles, D.J., 1996. Accounting for the missing carbon sink with the CO2 fertilization effect. Clim. Change 33, 31e62. Kim, T.H., Kim, J.S., Sunwoo, C., Lee, Y.Y., 2003. Pretreatment of corn stover by aqueous ammonia. Bioresour. Technol. 90, 39e47. Kirk, O., Borchert, T.V., Fuglsang, C.C., 2002. Industrial enzyme application. Curr. Opin. Biotechnol. 13 (4), 345e351. Knutsen, J.S., Davis, R.H., 2004. Cellulase retention and sugar removal by membrane ultrafiltration during lignocellulosic biomass hydrolysis. Appl. Biochem. Biotechnol. 113e116, 585e599. Kumari, V., Shah, S., Gupta, M.N., 2007. Preparation of biodiesel by lipase-catalyzed transesterification of high free fatty acid containing oil from Madhuca indica. Energy Fuel 21, 368e372. Kumartiwari, A.K., Kumar, A., Raheman, H., 2007. Biodiesel production from jatropha oil (Jatropha curcas) with high free fatty acids: an optimized process. Biomass Bioenergy 31, 569e575. Larsson, S.H., Thyrel, M., Geladi, P., Lestander, T.A., 2008. High quality biofuel pellet production from pre-compacted low density raw materials. Bioresour. Technol. 99, 7176e7182. Lewin, M., Roldan, L.G., 1971. The effect of liquid anhydrous ammonia in the structure and morphology of cotton cellulose. J. Polym. Sci. Part C 36, 213e229. Li, L., Du, W., Liu, D., Wang, L., Li, Z., 2006. Lipase-catalyzed transesterification of rapeseed oils for biodiesel production with a novel organic solvent as the reaction medium. J. Mol. Catal. B: Enzym. 43 (1e4), 58e62. Lin, Y., Tanaka, S., 2006. Ethanol fermentation from biomass resources: current state and prospects. Appl. Microbiol. Biotechnol. 69, 627e642. Linde, M., Galbe, M., Zacchi, G., 2008. Bioethanol production from non-starch carbohydrate residues in process streams from a drymill ethanol plant. Bioresour. Technol. 99, 6505e6511. Liu, J., Olsson, G., Mattiasson, B., 2004. Control of an anaerobic reactor towards maximum biogas production. Water Sci. Technol. 50 (11), 189e198. Malherbe, S., Cloete, T.E., 2003. Lignocellulosic biodegradation: fundamentals and applications: a review. Environ. Sci. Biotechnol. 1, 105e114. Marina, S., Verena, E.M.S., Martin, K., 2011. Pelletizing of autumn leavesdpossibilities and limits. Biomass Convers. Biorefin. 1, 173e187. Mariod, A., Klupsch, S., Hussein, H., Ondruschka, B., 2006. Synthesis of alkyl esters from three unconventional Sudanese oils for their use as biodiesel. Energy Fuel 20, 2249e2252. Meher, L.C., Vidya-Sagar, D., Naik, S.N., 2006. Technical aspects of biodiesel production by transesterificationda review. Renewable Sustainable Energy Rev. 10, 248e268. Mekala, N.K., Singhania, R.R., Sukumaran, R.K., Pandey, A., 2008. Cellulase production under solid-state fermentation by Trichoderma reesei RUT C30: statistical optimization of process parameters. Appl. Biochem. Biotechnol. 151, 122e131. Meng, X., Chen, G., Wang, Y., 2008. Biodiesel production from waste cooking oil via alkali catalyst and its engine test. Fuel Process. Technol. 89, 851e857. Metting, F.B., 1996. Biodiversity and application of microalgae. J. Ind. Microbiol. 17, 477e489. Modi, M.K., Reddy, J.R.C., Rao, B.V.S.K., Prasad, R.B.N., 2007. Lipasemediated conversion of vegetable oils into biodiesel using ethyl acetate as acyl acceptor. Bioresour. Technol. 98 (6), 1260e1264.
20 1. CURRENT BIOENERGY RESEARCHES: STRENGTHS AND FUTURE CHALLENGES Mohanty, S.M., Behera, S., Swain, M.R., Ray, R.C., 2009. Bioethanol production from mahula (Madhuca latifolia L.) flowers by solidstate fermentation. Appl. Energy 86, 640e644. Mohibbeazam, M.M., Waris, A., Nahar, N.M., 2005. Prospects and potential of fatty acid methyl esters of some non-traditional seed oils for use as biodiesel in India. Biomass Bioenergy 29, 293e302. Molina, E., Fernandez, J., Acien, F.G., Chisti, Y., 2001. Tubular photobioreactor design for algal cultures. J. Biotechnol. 92, 113e131. Moser, B.R., 2009. Biodiesel production, properties, and feedstocks. In Vitro Cell. Dev. Biol.: Plant 45, 229e266. Mosier, N., Wyman, C., Dale, B., Elander, R., Lee, Y.Y., Holtzapple, M., 2005. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96 (6), 673e686. Mulumba, N. 2010. Production of Biodiesel from Microalgae, M.S. Thesis, Chemical Engineering Dept., University of New Hampshire, Durham, NH, USA. Mulumba, N., Farag, I.H., 2012. Tubular photobioreactor for microalgae biodiesel production. Int. J. Eng. Sci. Technol. 4 (2), 703e709. Mustafa, B., Havva, B., 2009. Recent trends in global production and utilization of bio-ethanol fuel. Appl. Energy 86, 2273e2282. Naik, M., Meher, L.C., Naik, S.N., Das, L.M., 2008. Production of biodiesel from high free fatty acid karanja (Pongamia pinnata) oil. Biomass Bioenergy 32, 354e357. Ngo, H.L., Zafiropoulos, N.A., Foglia, T.A., Samulski, E.T., Lin, W., 2008. Efficient two-step synthesis of biodiesel from greases. Energy Fuel 22, 626e634. Nielsen, N.P.K., Gardner, D.J., Poulsen, T., Felby, C., 2009. Importance of temperature, moisture content, and species for the conversion process of wood into fuel pellets. Wood Fiber Sci. 41, 414e425. Ogbonna, J.C., Matsumura, M., Kataoka, H., 1991. Effective oxygenation of immobilized cells through reduction in bead diameters: a review. Process Biochem. 26, 109e121. Ogbonna, J.C., Tomiyama, S., Tanaka, H., 1996. Development of a method for immobilization of non-flocculating cells in loofa (Luffa cylindrica) sponge. Process Biochem. 31, 737e744. Ognjanovic, N., Bezbradica, D., Knezevic-Jugovic, Z., 2009. Enzymatic conversion of sunflower oil to biodiesel in a solvent-free system: process optimization and immobilized system stability. Bioresour. Technol. 100 (21), 5146e5154. Okano, K., Kitagaw, M., Sasaki, Y., Watanabe, T., 2005. Conversion of Japanese red cedar (Cryptomeria japonica) into a feed for ruminants by white-rot basidiomycetes. Anim. Feed Sci. Technol. 120 (3), 235e243. O’Donovan, A., Gupta, V.K., Tuohy, M.G., 2013. Recent updates in acid pretreatments and SEM analysis of acid pretreated grass biomass. In: Gupta, V.K., Tuohy, M.G. (Eds.), Biofuel Technologies; Recent Developments. Springer Science Publishers, USA, pp. 97e118. Pan, X., Xie, D., Gilkes, N., Gregg, D.J., Saddler, J.N., 2005. Strategies to enhance the enzymatic hydrolysis of pretreated softwood with high residual lignin content. Appl. Biochem. Biotechnol. A 124, 1069e1079. Pathak, H., Jain, N., Bhatia, A., Mohanty, S., Gupta, N., 2009. Global warming mitigation potential of biogas plants in India. Environ. Monit. Assess. 157, 407e418. Pei-dong, Z., Guomei, J., Gang, W., 2007. Contribution to emission reduction of CO2 and SO2 by household biogas construction in rural China. Renewable Sustainable Energy Rev. 11, 1903e1912. Phan, A.N., Phan, T.M., 2008. Biodiesel production from waste cooking oils. Fuel 87, 3490e3496. Pinelo, M., Jonsson, G., Meyer, A.S., 2009. Review: membrane technology for purification of enzymatically produced oligosaccharides: molecular features affecting performance. Sep. Purif. Technol. 70 (1), 1e11. Pongsawatmanit, R., Temsiripong, T., Suwonsichon, T., 2007. Thermal and rheological properties of tapioca starch and xyloglucan mixtures in the presence of sucrose. Food Res. Int. 40, 239e248. Potumarthi, R., Baadhe, R.R., Jetty, A., 2012. Mixing of acid and base pretreated corncobs for improved production of reducing sugars and reduction in water use during neutralization. Bioresour. Technol. 119, 99e104. Potumarthi, R., Baadhe, R.R., Jetty, A., 2013. Simultaneous pretreatment and saccharification of rice husk by Phanerochaete chrysosporium for improved production of reducing sugars. Bioresour. Technol. 128, 113e117. Ravichandra, R., Baadhe, R.R., Bhattacharya, S., 2013. Fermentable sugars from lignocellulosic biomass: technical challenges. In: Gupta, V.K., Tuohy, M. (Eds.), Biofuel Technologies: Recent Developments. Springer, Germany, pp. 3e27. Predojevic, Z.J., 2008. The production of biodiesel from waste frying oils: a comparison of different purification steps. Fuel 87, 3522e3528. Pruszko, R. 2007. Alternative feedstocks and biodiesel production. Presented at Practical Biodiesel Blueprint Conference (23e24 Jan, 2007), Kuala Lumpur, Malaysia. Rajendran, K., Aslanzadeh, S., Taherzadeh, M.J., 2012. Household biogas digestersda review. Energies 5, 2911e2942. Rodrigues, R.C., Pessela, B.C.C., Volpato, G., Fernandez-Lafuente, R., Guisan, J.M., Ayub, M.A.Z., 2010. Two step ethanolysis: a simple and efficient way to improve the enzymatic biodiesel synthesis catalyzed by an immobilized-stabilized lipase from Thermomyces lanuginosus. Process Biochem. 45 (8), 1268e1273. Rudolf, A., Alkasrawi, M., Zacchi, G., Liden, G., 2005. A comparison between batch and fedbatch simultaneous saccharification and fermentation of steam pretreated spruce. Enzyme Microb. Technol. 37 (2), 195e204. Sahoo, P.K., Das, L.M., Babu, M.K.G., Naik, S.N., 2007. Biodiesel development from high acid value polanga seed oil and performance evaluation in a CI engine. Fuel 86, 448e454. Saritha, M., Arora, A., Lata, 2012. Biological pretreatment of lignocellulosic substrates for enhanced delignification and enzymatic digestibility. Indian J. Microbiol. 52 (2), 122e130. Seadi, T.A., Rutz, D., Prassl, H., Kottner, M., 2008. Biogas Handbook. University of Southern Denmark, Esbjerg, Denmark. Shah, S., Gupta, M.N., 2007. Lipase catalyzed preparation of biodiesel from Jatropha oil in a solvent free system. Process Biochem. 42 (2), 409e414. Singh, A., Singh, I.S., 1991. Chemical evaluation of mahua (Madhuca indica (M. longifolia)) seeds. Food Chem. 40, 221e228. Singh, K.J., Sooch, S.S., 2004. Comparative study of economics of different models of family size biogas plants for state of Punjab, India. Energy Convers. Manage. 45, 1329e1341. Smith, A.M., 2008. Prospects for increasing starch and sucrose yields for bioethanol production. Plant J. 54, 546e558. Spolaore, P., Joannis-Cassan, C., Duran, E., Isambert, A., 2006. Commercial applications of microalgae. J. Biosci. Bioeng. 101, 87e96. Srivastava, P.K., Verma, M., 2008. Methyl ester of karanja oil as an alternative renewable source energy. Fuel 87, 1673e1677. Sun, Y., Cheng, J., 2002. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresour. Technol. 83 (1), 1e11. Taherzadeh, M.J., Karimi, K., 2007. Acid-base hydrolysis process for ethanol from lignocellulosic material: a review. Bioresour. Technol. 2 (3), 472e499. Tassinari, T., Macy, C., Spano, L., Ryu, D.D.Y., 1980. Energy requirements and process design considerations in compression milling pretreatment of cellulosic wastes for enzymatic hydrolysis. Biotechnol. Bioeng. 22, 1689e1705.
REFERENCES Tomar, S.S., 1994. Status of biogas plant in India. Renewable Energy 5, 829e831. Tu, M.B., Chandra, R.P., Saddler, J., 2007. Evaluating the distribution of cellulases and the recycling of free cellulases during the hydrolysis of lignocellulosic substrates. Biotechnol. Prog. 23, 398e406. United Nations Environment Program, 2006. Introducing the International Bioenergy Platform (IBEP). Available at: http://www. unep.org/training/programmes/Instructor%20Version/Part_2/ Activities/Innovations_and_Technology/Energy/Supplemental/ Bioenergy.pdf. (accessed 11.03.2013). US DOE, 2010. U.S. Department of Energy’s Bioenergy Research Centers: An Overview of the Science, DOE/SC-0127. Office of Biological and Environmental Research within the DOE Office of Science. Available at: genomicscience.energy.gov/centers/ brcbrochure.pdf. (accessed 15.03.2013). Usta, N., 2005. Use of tobacco seed oil methyl ester in a turbocharged indirect injection diesel engine. Biomass Bioenergy 28, 77e86. van Beilen, J.B., Li, Z., 2002. Enzyme technology: an overview. Curr. Opin. Biotechnol. 13 (4), 338e344. Van Maris, A.J.A., Abbott, D.A., Bellissimi, E., Brink, J.V.D., 2006. Alcoholic fermentation of carbon sources in biomass hydrolysates by Saccharomyces cerevisiae: current status. Anton. Leeuw. Int. J. G. 90 (4), 391e418. Van-Zyl, W.H., Lynd, L.R., Haan, R.D., McBride, J.E., 2007. Consolidated bioprocessing for bioethanol production using Saccharomyces cerevisiae. Biofuels 108, 205e235. 21 Veljkovic, V.B., Lakicevic, S.H., Stamenkovic, Z.B., Todorovic, O.S., Lazic, M.L., 2006. Biodiesel production from tobacco (Nicotiana tabacum) seed oil with a high content of free fatty acids. Fuel 85, 2671e2675. Verma, V.K., Bram, S., Delattin, F., Laha, P., Vandendael, O., Hubin, A., 2012. Agropellets for domestic heating boilers: standard laboratory and real life performance. Appl. Energy 90 (1), 17e23. Working of biogas plants. 2013. Tutorvista.com. Available at: http:// www.tutorvista.com/content/physics/physics-ii/fission-andfusion/biogas-plants.php#construction-of-the-floating-gas-holdertype-plant. (accessed 21.03.2013). Wuebbles, D.J., Jain, A.K., 2001. Concerns about climate change and the role of fossil fuel use. Fuel Process. Technol. 71, 99e119. Wyatt, V.T., Hess, M.A., Dunn, R.O., Foglia, T.A., Haas, M.J., Marmer, W.M., 2005. Fuel properties and nitrogen oxide emission levels of biodiesel produced from animal fats. J. Am. Oil Chem. Soc. 82, 585e591. Yang, S., Ding, W., Chen, H., 2006. Enzymatic hydrolysis of rice straw in a tubular reactor coupled with UF membrane. Process. Biochem. 41 (3), 721e725. Yang, J., Chen, W., Chen, B., 2011. Impacts of biogas projects on agroecosystem in rural areasda case study of Gongcheng. Front. Earth Sci. 5, 1e6. Zhang, J., Jiang, L. 2008. Acid-catalyzed esterification of Zanthoxylum bungeanum seed oil with high free fatty acids for biodiesel production. Bioresource. Tech.
C H A P T E R 2 Bioenergy Research: An Overview on Technological Developments and Bioresources Vijai K. Gupta 1,*, Ravichandra Potumarthi 2, Anthonia O’Donovan 1, Christian P. Kubicek 3, Gauri Dutt Sharma 4, Maria G. Tuohy 1,* 1 Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland, 2Department of Chemical Engineering, Monash University, Clayton, Victoria, Australia, 3Research Area Biotechnology and Microbiology, Institute of Chemical Engineering, TU Wien, Gumpendorferstrasse Wien, Austria, 4Bilaspur University, Bilaspur, Chattisgarh, India *Corresponding author email: vijai.gupta@nuigalway.ie, maria.tuohy@nuigalway.ie O U T L I N E Introduction 23 Current Bioenergy Practices 25 Main Biofuel Technologies and Current Processes 26 Technological Routes for Bioenergy Production Biomass Pretreatment Hydrolysis Fermentation Combined Pretreatment, Hydrolysis and Fermentation Strategies Advanced Biomass-to-Biofuels Development Platform 28 28 29 29 29 Bioenergy Resources and Biofuels Development Program Sustainability Bioenergy Feedstocks and Dedicated Biofuel Crops Lignocellulosic Feedstocks Dedicated Bioenergy Crops Feedstocks for Biodiesel 36 Conclusions 41 References 41 37 37 38 39 30 INTRODUCTION nontoxic; biofuel spillages present far less risk than fossil fuel spillages (Hahn-Hagerdal et al., 2006). Bioenergy is a term broadly used to describe gaseous, liquid or solid energy products that, for the most part, are derived from biological raw materials (biomass). In the 1990s bioethanol was a promising technological option to reduce transportation sector greenhouse gas (GHG) emissions (Lynd, 1996); most of this ethanol was derived from so-called first-generation, starchand sucrose-rich feedstocks. Bioethanol is readily made from starchy seeds, tubers, or roots of plants such as maize (Zea mays), barley (Hordeum vulgare), Fossil fuels such as petrol, diesel or crude oil, are nonrenewable sources of fuel and are not natural resources in many countries making these nations dependent on fossil fuel-rich countries at enormous expense. The rising cost and simultaneous depletion of fossil fuels, in addition to political instability in key countries, means the competitiveness of biomass-derived energy has increased considerably. Additionally bioenergy sources, including biofuels, pose a reduced threat to the environment because they are biodegradable and Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00002-4 33 23 Copyright Ó 2014 Elsevier B.V. All rights reserved.
24 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES wheat (Triticum aestivum), rice (Oryza sativa), potato (Solanum tuberosum), sweet potato (Ipomea batatas), cassava (Manihot esculenta), Jerusalem artichoke (Helianthus tuberosus), etc. and from the sugar-rich stems and roots of sugarcane (Saccharum officinarum), sweet sorghum (Sorghum vulgare), and sugar beet (Beta vulgaris). Indeed, the basic technology for making ethanol from such crops is centuries old (Lemus and Parrish, 2009). Currently bioethanol is produced commercially by fermentation of sugars derived from corn, sugar cane and sugar beet. It is expected that, in light of the increase in global population and the “food versus fuel” debate, there will be limits to the supply of these feedstocks for biofuel production in the near future ultimately making first-generation biofuels an unsustainable approach to meet future energy needs (O’Donovan et al., 2013; Groom et al., 2008; Simpson et al., 2008). Perennial herbaceous energy crops make good feedstocks because they do not require annual reseeding once established, need fewer energy inputs (such as fertilizers and pesticides) than annual cropland, and can be grown on marginal lands (Dien et al., 2005). They also have environmental benefits that include reduced soil erosion, enhanced carbon sequestration, and conservation of wildlife habitats (Lemus and Lai, 2005). The major herbaceous energy crops that have been selected for bioethanol production in the United States are switch grass (Panicum virgatum), Miscanthus (Miscanthus spp. Anderss.), canary grass (Phalaris arundinacea), giant reed (Arundo donax L.), and alfalfa (Medicago sativa L.). They are considered to have energetic, economic, and environmental advantages over food crops for ethanol production. While these dedicated energy crops contain substantial amounts of holocellulose (cellulose and hemicelluloses) in their cell walls, their feedstock quality for livestock makes them less attractive options for fuel ethanol and bioenergy generation (Hill et al., 2006). It is essential to approach renewable energy (REN) production through the application of complementary technologies. The recent “food versus fuel” debate has motivated the development of technologies to utilize nonfood crops as well as food wastes and agriprocessing wastes, in biomass to bioenergy strategies; therefore, lignocellulosic residues and other nonfood plant biomass types are considered attractive alternative feedstocks. Large efforts are being made worldwide in order to develop technologies that generate clean, sustainable energy sources from nonfood biomass feedstocks that could substitute fossil fuels (Ragauskas et al., 2006; Levin et al., 2006). At present, in the United States, biomass provides about 40 times as much energy as photovoltaics (Banerjee, 2011) and represents 78% of the total REN generated worldwide (International Energy Agency (IEA), 2010). Biofuels are the only viable energy source for the foreseeable future and can provide sustainable development in a manner that will address socioeconomic and environmental concerns (Demirbas, 2005). Bionergy derived from second-generation feedstocks, i.e. lignocellulosic materials is now the prime target for commercial biofuel production (O’Donovan et al., 2013; Demain, 2009). Lignocellulosic biomass is an abundant, domestic, renewable feedstock source rich in complex carbohydrates, which can be converted to liquid transportation fuel and other chemicals by strategies involving carbohydrate degradation and subsequent fermentation. More than 70% of lignocellulosic biomass is made up of the complex biopolymers, cellulose, hemicellulose, and lignin. The organization of these structural polymers in the plant cell wall makes such feedstocks highly recalcitrant to bioconversion and difficult to use as a raw material in ethanol production compared with starch (O’Donovan et al., 2013; Abramson et al., 2010; Somerville et al., 2010). However, lignocellulosic biomass in the form of wood and agricultural residues is virtually inexhaustible (Sarkar et al., 2012; Zhang et al., 2007; Zhang and Lynd, 2006; Lynd et al., 2002; Kuhad et al., 1997). Agricultural residuals or by-products are annually renewable, abundantly available and account for more than 180 million tons of biomass per year (Kapdan and Kargi, 2006). The most abundant lignocellulose agricultural residues are corncobs, corn stover, wheat, rice, barley straw, sorghum stalks, coconut husks, sugarcane bagasse, switchgrass, pineapple and banana leaves (Demain et al., 2005; Kim and Dale, 2004). Cereal crops, pulse crops and harvestable palm oil biomass are also being produced in large amounts worldwide annually (Rajaram and Verma, 1990). In addition, wood and paper industries generate huge amounts of residual lignocellulosic biomass. Along with agricultural and forestry wastes and residues, locally available nonfood plant biomass and municipal solid wastes are potential candidates to meet demands for biofuel and bioenergy production, since additional costs for cultivation and harvesting are not involved. It is likely that the diversity of raw materials will support the decentralization of fuel production with geopolitical, economic, and social benefits (Wyman, 2007), thus bringing further socioeconomic benefits. The concept of replacing fossil fuels with alternative biobased energy sources and fuels has been markedly enhanced by the realization that plant biomass also has the potential to provide a wide range of feedstock (bio)chemicals that can yield high-value commodity products and offset bioenergy production costs in lignocellulose-based biorefinery approaches (Zhu and Zhuang, 2012; Cherubini, 2010; FitzPatrick et al., 2010; Percival Zhang, 2008; Taylor, 2008; Kamm and Kamm, 2004). Bioenergy production processes (e.g. anaerobic
CURRENT BIOENERGY PRACTICES digestion and thermochemical treatments) can also generate organic wastes that still have significant market potential, for example, as organic fertilizers and biochars, which are most important for soil enrichment. The developments in biorefining have underpinned several recent and promising advances in bioenergy (Aden et al., 2002). Bioethanol, biobutanol and biomethane are promising future bioenergy and biofuel sources. Biomethane is produced most frequently through anaerobic digestion, in which biomass is converted by consortia of bacteria via hydrolysis, fermentation, acetogenesis and methanogenesis reaction steps to methane and smaller amount of other gases (Keating et al., 2012; Liew et al., 2012; McHugh et al., 2003; Mata-Alvarez et al., 2000). Liquid fuels that are being produced from biomass are typically of higher quality and burn more cleanly than petroleum-based diesel and jet fuels. Biofuels also reduce the release of volatile organic compounds, as the addition of ethanol to gasoline oxygenates the fuel mixture causing it to burn more completely. Biodiesel is another important biofuel usually produced from oleaginous crops, such as rapeseed, soybean, sunflower, palm and from microalgae through a chemical transesterification process of their oils with shortchain alcohols, mainly methanol (Antolı́n et al., 2002). Thus, a shift to biofuels for current fuel needs would reduce energy dependency on oil imports and could boost rural development, providing farmers and crop producers with an additional source of income. CURRENT BIOENERGY PRACTICES Biofuels account for the major proportion of bioenergy production worldwide, with most of the fuels being derived through biochemical processes. For this reason, this review will focus in the main on current practices used in the production of the main biofuels. The major producers of bioethanol are Brazil and the United States, both of which account for about 89% of world production (World Development Report, 2008; Lichts, 2010), while the European Union is the world’s largest producer of biodiesel (OECD-FAO, 2009). The United States has been the world’s largest producer of ethanol fuel since 2005 and the world’s largest exporter since 2010. In 2011, the United States produced 52.6 billion liters (13.9 billion US liquid gallons) of ethanol, while Brazil produced 21.1 billion liters (5.57 billion US liquid gallons), representing 24.9% of the world’s total ethanol used as fuel (Renewable Fuels Association, 2012). Fuel ethanol production is considerably more modest in the European Union, where France, Germany and Spain are the largest producers of bioethanol producing 950, 581 and 346 million liters, respectively, in 2008 25 (European Bioethanol Fuel Association, 2009). Countries such as Poland, Hungary and Slovakia have also increased their bioethanol output producing 200, 150 and 94 million liters of bioethanol, respectively (European Bioethanol Fuel Association, 2009). Sweden is the leading country in Europe in terms of the use of ethanol as fuel, the impetus for which is driven by government policy. Although most of the ethanol is imported, Swedish gas stations are required by an act of parliament to offer at least one alternative fuel. Furthermore, reductions in biofuel prices to the consumer have also encouraged biofuel consumption. Government incentives for biofuel replacement of gasoline are now being implemented in other countries worldwide, motivated by ever-increasing oil costs, depleting fossil fuel resources, GHG emission targets and the need for greater diversification to support agricultural and rural development (Mussato et al., 2010). The major feedstock for bioethanol in Brazil is sugarcane including bagasse, while corn grain/maize is the main feedstock used for bioethanol production in the United States. As mentioned earlier, bioethanol can be produced from any sugar or starch crop in firstgeneration processes, but other potential resources for bioethanol include sugar beet, cassava, maize, oil palm, rapeseed, soybean, corn stover, grass, leaves, agricrop residues and various locally available nonfood plant biomass like Jatropha, Miscanthus, willow, hemp and switchgrass. Table 2.1 summarizes the major crops/biomass currently (ranked in order of importance) in use for biofuel and bioenergy production in different countries. Shapouri (1995), Shapouri et al. (2002) concluded that the energy content of bioethanol was higher than the energy required to produce it, although other researchers would argue as to the economic viability of bioethanol in the absence of an accompanying high-value biorefinery process. Production of ethanol from lignocellulosic biomass is a complex process where the biomass often requires pretreatment to render the holocellulose more accessible to a mixture of enzymes, which are utilized to saccharify or hyrolyze the complex polysaccharides to fermentable sugars. Pretreatment processes can be expensive, toxic and corrosive and may require a subsequent costly detoxification step (Agbor et al., 2011; Zhang and Lynd, 2004; Sun and Cheng, 2002). In addition, preparation of fermentable sugars and the inhibitory effect of lignin and carbohydrate-derived compounds, formed during pretreatment of the lignocelluloses, are the major bottlenecks in bioconversion processes (Viikari et al., 2007). However, since biomass energy is derived from renewable resources, its production can still be advantageous if proper management technologies are utilized in biomass harvesting, pretreatment and processing, and if biomass feedstocks are produced sustainably. Plant
26 TABLE 2.1 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Major Crops Used for the Production of Biofuels Crop Countries Cassava Nigeria/Brazil/Thailand/Indonesia Corn Stover United States/Latin America Hemp Ireland/United Kingdom/United States/other European countries/China Jatropha India/China/other Asian countries/Africa Maize United States/China/Brazil/Mexico Miscanthus Ireland/United Kingdom/other European countries Oil Palm Malaysia/Indonesia/Nigeria/Thailand Rapeseed China/Canada/India/Germany Soybean United States /Brazil/Argentina/China Sugar Beet France/United States/Germany/Russia Sugarcane Brazil/India/China/Thailand Switch Grass Ireland/United Kingdom/United States/other European countries/China Willow Ireland/United Kingdom/other European countries Source: Müller et al., 2008; De Fraiture and Berndes 2009. biomass to energy or chemicals can be economical only if all of the components in the biomass are converted into fuel, chemicals or other value-added components in a true biorefinery approach (FitzPatrick et al., 2010; Cherubini, 2010; Percival Zhang, 2008; Kamm and Kamm, 2004). MAIN BIOFUEL TECHNOLOGIES AND CURRENT PROCESSES Biofuels, biogas and syngas are energy carriers that store the energy derived from biomass. A wide range of biomass sources can be used to produce bioenergy in a variety of forms (see Figure 2.1). As mentioned, food, fiber and wood process residues from the industrial sector; energy crops, short rotation crops and agricultural wastes from the agriculture sector; and residues from the forestry sector are being used to produce biofuels, generate electricity, heat, combined heat and power, and other forms of bioenergy (Department of Energy (DOE), 2011; Chu and Majumdar, 2012). Figure 2.1 provides an overview of the main biological and chemical technology platforms that are currently in use for the production of different types of bioenergy, including biofuels. Liquid biofuels produced from agricultural crops, e.g. cereals, maize, sugarcane, sugar beet and sweet sorghum and from vegetable oil, as well as biogas, are referred to as first-generation biofuels, while those produced from lignocellulosic biomass and nonfood crops are referred to as second-generation biofuels (Yuan et al., 2008). Biofuels produced from algae are termed third-generation biofuels (Brennan and Owende, 2010; see Table 2.2). The many sources of biomass used for energy purposes are often scattered across large and diverse geographical areas. Even today, most energy derived from biomass originates from by-products of food, fodder and fiber production. For instance, the main byproducts of forest processing industries are used to produce fuels wood and charcoal, and black liquor (a by-product of pulp mills) is a major fuel source for bioelectricity generation in countries such as Brazil, Canada, Finland, Sweden and the United States. A considerable amount of heat and power is derived from recovered and/or recycled woody biomass and increasing amounts of energy are being recovered from biomass derived from cropland (straw and cotton stalks) and forest land (wood chips and pellets). In sugar- and FIGURE 2.1 Main technology platforms for bioenergy production (including biofuels) and by-/coproducts from various feedstocks. Source: Department of Energy (DOE), 2011; Chu and Majumdar, 2012. (For color version of this figure, the reader is referred to the online version of this book.)
27 MAIN BIOFUEL TECHNOLOGIES AND CURRENT PROCESSES TABLE 2.2 Bioenergy Outputs, Feedstocks Utilized and By-/Coproducts Bioenergy Category Feedstock(s) Used Bioenergy Outputs Solid Biofuels Woody material, dried manure Used as dried biomass for energy 1. Rapeseed oil, sunflower oil and other 1. As transport fuel 2. For generation of electricity By-/Coproducts FIRST GENERATION Plant Oils (Vegetable Oil) plant oil, waste vegetable oil 2. Rapeseed oil, palm oil, Jatropha and other plant oil Biodiesel Bioethanol Biogas Oil cake as animal feed and heat in decentralized combined heat and power (CHP) stations 1. 2. 3. 4. Europe: rapeseed, sunflower, soya United States: soya, sunflower, Canada: soya, rapeseed, canola South and Central America: soya, palm, Jatropha, castor 5. Africa: palm, soya, sunflower, Jatropha 6. Asia: palm, soya, rapeseed, sunflower, Jatropha Transesterification of oils and fats to provide fatty acid methyl ester and to use transport fuel 1. 2. 3. 4. 5. Europe: cereals, sugar beets United States: corn Canada: maize, cereals Brazil: sugarcane South and Central America: sugarcane, cassava 6. Africa: sugarcane, maize 7. Asia: sugarcane, cassava Fermentation (sugar), hydrolysis and fermentation (starch) used in transport fuel Energy crops (viz. maize, Miscanthus, wood from short rotation and multiple cropping system), agriculture residues (waste leaves, twigs, plant material), biodegradable waste material including sewage from any source Fermentation of biomass used in decentralized system for energy requirement or through supply into gas pipeline (as purified biomethane) 1. For generation of electricity and heat in 1. Oil cakes as animal feed 2. Glycerin 3. Oil cakes as residue for energy recovery 1. Maize and cereals yield animal feedstocks 2. Sugarcane bagasse used for energy recovery Residues used as fertilizer (soil conditioners and nutrient recycling) CHP-based power stations 2. For transport fuel: either 100% biogas fuel or blending with natural gas used as fuel Solid Biofuels Wood, grass cuttings, switchgrass, perennial rye grass, grass press cake, grains, straw, charcoal, domestic refuse and dried manure 1. Densification of biomass by carbonization (charcoal) 2. Residuals and waste for generation of electricity and heat (e.g. industrial waste in CHP) Residues used as fertilizer (soil conditioners and nutrient recycling) SECOND GENERATION Bioethanol Lignocellulosic biomass like straw, stalks of wheat, corn stover and wood, bioenergy crops, e.g. Miscanthus, willow, witchgrass (Panicum virgatum L.), reed canary grass (Phalaris arundinacea L.), and alfalfa (Medicago sativa L.), agave, Jatropha Biodiesel and Range of Lignocellulosic biomass like straw, Biofuels such as wood and secondary raw materials Biohydrogen, Biomethane; 2,5-Dimethylfuran, Dimethyl ether, Mixed Alcohols Breakdown of lignocellulosic biomass in many steps via hydrolysis to fermentation to produce bioethanol (Both biochemical and thermal platforms are being used as technological innovation) 1. Residues used as fertilizer (soil conditioners and nutrient recycling) 2. Range of biochemicals Gasification of low-moisture biomass provides syngas (mix of CO, CO2, H2, CH4, and hydrocarbons) from which liquid fuels and other chemicals are being derived Various feedstocks for chemical industry to produce range of biochemical and plastics Bioreactors for ethanol, transesterification and pyrolysis for biodiesel Liquid biofuels processing Biopolymers, high-protein animal feedstocks, agricultural fertilizers THIRD-GENERATION BIOFUELS Biodiesel, Aviation Fuels, Macroalgae, microalgae Bioethanol, Biobutanol, Biohydrogen, Bio-Oils from Algae Source: Farine et al., 2011; Stefan et al., 2009.
28 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES coffee-producing countries, bagasse and coffee husks are used for direct combustion to produce heat energy and steam (http://www.thebioenergysite.com/articles/172/ biofuels-and-agriculture). In terms of bioenergy, however, the big growth area in recent years has been in the production of liquid biofuels for transport using agricultural crops as feedstocks. The bulk of the biofuel output is in the form of bioethanol (sugar crops or starchy crops), or biodiesel (oil crops, e.g. oilseed rape and Jatropha). Table 2.2 provides an overview of some of the main feedstocks used for target bioenergy outputs under first-, second- and third-generation biofuel production strategies. TECHNOLOGICAL ROUTES FOR BIOENERGY PRODUCTION Suitable conversion technologies are needed in order to effectively breakdown or deconstruct biomass into simple sugars, carbohydrate derivatives or bio-oils that are more easily converted into fuels in combination with downstream conversion technologies to subsequently upgrade these intermediates into bioenergy, biofuels and value-added bioproducts. At present, the bioconversion of lignocellulose is carried out in four major steps viz. pretreatment, hydrolysis, fermentation and separation/purification to recover bioenergy/biofuels and residues (more recently, the recovery of coproducts has becoming increasingly important). The pretreatment of lignocellulose materials is considered a key step in bioenergy production and indeed in biorefining as it accelerates the hydrolysis procedure, by enhancing cellulose accessibility and increasing pore size, which, in theory, leads to higher sugar yields for fermentation. An ideal pretreatment should remove lignin and thus reduce the crystallinity of cellulose (Lynd et al., 2002), increase porosity and accessibility of the cellulose (and hemicellulose) to enzymatic hydrolysis, release/generate low levels of inhibitory compounds, be low cost and have low energy requirements. The overall result should be a reduction in the recalcitrance of lignocellulose and an increase in accessibility to enzymes. In general, accessibility of cellulose is achieved through the removal of lignin and hemicellulose polymers through various pretreatment methods, which can be defined as chemical, physical or biological (O’Donovan et al., 2013; Dashtban et al., 2009; Taherzadeh and Karimi, 2008; Ong, 2004; Howard et al., 2003). Biomass Pretreatment Pretreatments vary from hot-water extraction and steam pretreatments (often with an oxidant or other chemical) to weak and strong acid and alkali pretreatments (Sun and Cheng, 2002). Physical pretreatments include mechanical communition, milling and ultrasound methods (Agbor et al., 2011; Balat, 2011), as well as irradiation. Chemical pretreatment methods include ammonia fiber explosion (AFEX), organosolv treatment and the addition of either acid or alkali (Isroi et al., 2011; Ong, 2004; Dashtban et al., 2009). The use of acid as a catalyst, normally H2SO4, targets the hemicellulose to dissolve with lignin and cellulose remaining as solids, whereas the addition of alkali, normally NaOH, mainly targets lignin, leaving mainly cellulose as a solid with hemicelluloses (Dashtban et al., 2009; Ong, 2004). Physicochemical pretreatments combine a mild chemical treatment with high pressure and temperature and include methods ranging from uncatalyzed solvolysis (hydrothermolysis) to steam explosion with chemical additives such as carbon dioxide or sulfur dioxide, AFEX and “popping” techniques (Mosier et al., 2005; Pan et al., 2005; Wi et al., 2011). Recent developments include pretreatments based on alkali soaking (NaOH) coupled with extrusion (Karunanithy and Muthukumarappan, 2011). Steam explosion consists of steaming the lignocellulose at high pressure followed by either a rapid or slow reduction in pressure to dissolve the hemicelluloses into solution and allow the cellulose and lignin to remain as solids (Ong, 2004; Dashtban et al., 2009). SO2 or CO2 can be used as catalysts, although SO2 can be highly toxic to downstream fermentation microorganisms (Ong, 2004). Although physical and chemical pretreatments can effectively reduce the recalcitrance of lignocellulosic compounds within a shorter time frame, they result in many environmental and cost concerns for industries. They require high-energy input alongside highpressure reactors and can produce toxic compounds and wastewaters (Isroi et al., 2011). Biological pretreatment methods include the use of microorganisms in order to delignify the lignocellulose material (Ravichandra et al. 2013; Dashtban et al., 2009). The enzymes produced by the microorganisms selectively disrupt the fibril and lignin structures of the plant cell wall and provide the advantages of lower energy demands, minimal waste production and reduced effects on the environment (Isroi et al., 2011; Dashtban et al., 2009). Microbial delignification is a gentle and effective approach to remove up to 31.59% lignin from biomass such as corn stover (Wan and Li, 2010) but results in a low rate of downstream hydrolysis. Pretreatment times required for direct microbiological methods are lengthy, being typically from 18 to 35 d. Nonetheless, enzymatic delignification is an alternative option and different “-omics” technologies are likely to yield new enzymatic delignification systems from
TECHNOLOGICAL ROUTES FOR BIOENERGY PRODUCTION different white rot and brown rot fungi (Martinez et al., 2009). The method chosen for pretreatment is dependent upon the lignocellulosic material and the hydrolysis to be carried out afterward. If the hydrolysis step is accompanied by microbial enzymes, which are optimized at a lower pH (4e6), an acidic pretreatment is preferred as the first step in the bioconversion process (Dashtban et al., 2009). Hydrolysis Hydrolysis is the process by which the lignocellulose polymers are reduced (saccharified) to yield fermentable sugars (hexoses and pentoses) (Harris and DeBolt, 2010). There are two methods of hydrolysis used within the bioenergy and biorefining processes, namely, acid hydrolysis and enzymatic hydrolysis (Potumarthi et al., 2013, 2012; Dashtban et al., 2009; Ong, 2004). Acid hydrolysis is the older method of the two and has been implemented on an industrial scale since World War I. In this particular process, dilute or concentrated acid, normally H2SO4 as it is cheapest, is used to hydrolyze the cellulose with the reaction temperatures dependent upon the molarity; dilute acids require higher temperatures (above 200  C) while concentrated acids require lower temperatures. The acid hydrolysis approaches are less attractive due to low yields with dilute acid and the recovery and environmental factors involved with use of concentrated acids (Ong, 2004). In enzymatic hydrolysis, the lignocellulose is broken down into the corresponding monomeric sugars by specific enzymes produced from bacteria or fungi ( Coyne et al., 2013; Gupta et al., 2013; Ong, 2004; Dashtban et al., 2009). This approach is more complex, expensive and time consuming, in comparison to the acid hydrolysis approach, but has the advantage of little or no byproducts to dispose of at the end of the biorefining process (Ong, 2004) and it can be used for more selective fractionation in a biorefinery context (Menon and Rao, 2012). Pretreatment of lignocelluloses with acid or alkali partially removes the lignin and hemicellulose but also substantially disrupts the fibrillar structure of biomass. Therefore, acid or alkali pretreated lignocellulosic biomass can be saccharified enzymatically to produce fermentable sugars. This results in faster hydrolysis rates and higher glucan enzymatic digestibility. A common belief is that lignin removal in particular promotes faster and more efficient enzymatic cellulose hydrolysis (Zhu et al., 2008). Fermentation Fermentation, the third step of bioconversion, converts the hydrolysates, mainly glucose, xylose, arabinose 29 and mannose to bioethanol using microorganisms. In addition to bioethanol, fermentation can be used to generate other useful end products, e.g. biobutanol, fatty acids, lactic acid, bioplastics or other biochemicals. The hydrolysates are often detoxified before fermentation due to the production of inhibitory compounds, such as phenolics and furan derivatives, in the pretreatment and hydrolysis steps (Dashtban et al., 2009; Ong, 2004). Saccharomyces cerevisiae is the most commonly used microorganism as it has a high fermentation rate and the application of recombinant DNA techniques has enabled the bioengineering of strains, capable of converting arabinose and xylose, as well as glucose, to bioethanol (Dashtban et al., 2009). This allows utilization of a larger amount of the hydrolysates, thus giving a higher percentage yield of bioethanol. Combined Pretreatment, Hydrolysis and Fermentation Strategies Different combinations of the first three bioconversion steps have been investigated in order to reduce production costs, increase end-product yield and reduce time required for bioconversion. Sequential hydrolysis and fermentation provides the opportunity of optimizing each process separately, although it can result in the use of large amounts of enzymes such as b-glucosidase to overcome end-product inhibition during the hydrolysis making this a costly process (Blanch, 2012; Dashtban et al., 2009). Simultaneous saccharification and fermentation (SSF) combines both steps into one reaction, which in theory allows direct fermentation of hydrolysates into bioethanol with a reduction in enzyme costs. However, involved both reactions and end-product yields can be compromised in SSF (Dashtban et al., 2009; Ong, 2004). Another method termed consolidated bioprocessing can be used to combine all three steps into one with the use of one or many microorganisms (Hasunuma et al., 2013; Matano et al., 2013; Amore and Faraco, 2012; Blanch, 2012; Hasunuma and Kondo, 2012; Girio et al., 2010; Dashtban et al., 2009). This particular process possesses the potential to reduce bioethanol production costs to competitive fuel levels. Although significant advances have been made with regard to CBP (Hansunuma et al., 2013; Hyeon et al., 2013; Matano et al., 2013; Olson et al., 2012), more research into the microbial cell factories, enzymes and physicochemical and catalytic conditions (pH, temperature, and synergies) is required (Olson et al., 2012; Menon and Rao, 2012; Van Dyck and Pletschke, 2012). However, key technologies are available to convert a variety of biomass into electricity, gas, or different liquid fuels (Table 2.3). These technologies use various types of feedstocks, and are produced in different ways (Farine et al., 2011).
30 TABLE 2.3 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Biomass to Bioenergy Routes for Important Feedstocks Conversion Route Energy Product Type Starch (Wheat, Sorghum, Barley, Oat and Triticale Grain) Fermentation Ethanol Sucrose (C-Molasses and Sugarcane Sugar) Fermentation Ethanol Oil (Canola, Animal Tallow, Waste Oil Mixture, Algae, Pongamia Seed) Transesterification Biodiesel Lignocellulose (Stubble from Annual Crops, Bagasse, Sugarcane (Whole Plant), Products and Residues from Native Forest, and Hardwood and Softwood Plantations, Wood Waste Mixture and Coppice Eucalyptus) Enzymatic fermentation Ethanol Lignocellulose (Stubble from Annual Crops, Bagasse, Sugar Cane (Whole Plant), Products and Residues from Native Forest, and Hardwood and Softwood Plantations, Wood Waste Mixture and Coppice Eucalyptus) Combustion Electricity Feedstock/Biomass Source: Farine et al., 2011. Advanced Biomass-to-Biofuels Development Platform The lignocellulosic substrates include woody substrates such as hardwood (birch, aspen, etc.) and softwood (spruce, pine, etc.), agri residues (wheat straw, sugarcane bagasse, corn stover, etc.), dedicated energy crops (switchgrass, willow, hemp, Miscanthus, etc.), weedy materials (Eichhornia crassipes, Lantana camara, etc.), and municipal solid waste (food and kitchen waste, etc.). The diversity of raw materials will allow the decentralization of fuel production with geopolitical, economic, and social benefits (Van Dyck and Pletschke, 2012; Wyman, 2007). Despite the success achieved in the laboratory, there are limitations to success with lignocellulosic substrates on a commercial scale (Chandel and Singh, 2011) as each source of biomass brings a unique technological challenge. The advanced biomass-to-biofuels development platform has multiple goals, including the use of new enzymes to take full advantage of available carbohydrates, the development of new lines of bioenergy crops with increased fermentation productivity (Carpita, 2012; Abramson et al., 2010), the development of new uses for coproducts, and the reduction of processing and energy costs. Lignocelluloses have three main components: cellulose, hemicelluloses, and lignin. Cellulose is the most abundant organic polymer on the earth. It is a homopolymer of sugars containing six carbon atoms linked together in a chain that constitutes the largest proportion of the plant cell wall. Hemicelluloses are heteropolysaccharides consisting of short branched chains of hexoses, e.g. mannose units in mannans and pentoses such as xylose units in xylans (Chandel et al.,. 2010; Girio et al., 2010; Kuhad et al., 1997). Table 2.4 summarizes the basic cell wall composition of some important lignocellulosic biomass used in bioenergy generation. In general, hardwoods contain 18e25% lignin, 45e55% cellulose, and 24e40% hemicelluloses, while softwoods contain 25e35% lignin, 45e50% cellulose, and 5e35% hemicelluloses. Grasses normally contain 10e30% lignin, 25e40% cellulose, and 25e50% hemicelluloses (Balat, 2011; Sanchez, 2009; Howard et al., 2003; Malherbe and Cloete, 2003; Betts et al., 1991). Agri-biomass commonly comprises about 40% cellulose, 25% hemicellulose and 18% lignin. The structure and components of the cell walls of weeds are significantly different from those of most plant species, which may influence digestibility during the bioconversion process to bioethanol (Van Dyck and Pletschke, 2012; Chandel and Singh, 2011; Sarkar et al., 2009). The hydrolytic breakdown of cellulose in nature involves the use of enzymes including cellobiohydrolases, endoglucanases and b-glucosidases produced by microbes or other biological agents, alone or in combination (Turner et al., 2010; Kuhad et al., 1997). More recent studies have shown that additional oxidoreductase enzymes (glycosyl hydrolase family 61 polysaccharide monooxygenases and cellobiose dehydrogenase) are essential components in a complete cellulose-degrading enzyme system (Horn et al., 2012; Kittl et al., 2012; Langston et al., 2011). The sugar chains of cellulose can be hydrolyzed to produce glucose and cellooligosaccharides, most of which can be fermented using ordinary baker’s yeast. To attain economic feasibility a high ethanol yield is a necessity. Producing monomer sugars from cellulose and hemicellulose at high yields is far more difficult than deriving sugars from sugar- or starchcontaining crops, e.g. sugarcane or maize (Van Dyck and Pletschke, 2012; Tuohy et al., 1994). Therefore, although the cost of lignocellulosic biomass is far lower than that of sugar and starch crops, the cost of obtaining sugars from such materials for fermentation into bioethanol has historically been far too high to attract industrial interest. For this reason, it is crucial to solve the problems involved in the conversion of lignocellulosic biomass to sugar and further to ethanol (Agbor et al., 2011; Galbe and Zacchi, 2002). The heterogeneity in feedstock and the influence of different process conditions on microorganisms and enzymes makes the biomass-to-ethanol process
31 TECHNOLOGICAL ROUTES FOR BIOENERGY PRODUCTION TABLE 2.4 Cell Wall Compositions (%) of Different Lignocellulosic Sources Biomass Type Cellulose Hemicellulose Lignin References Birch 40.0 23.0 21.0 Olsson and Hahn-Hägerdal, 1996 Willow 37.0 23.0 21.0 Olsson and Hahn-Hägerdal, 1996 Aspen 51.0 29 16 Olsson and Hahn-Hägerdal, 1996 Spruce 43 26 29 Olsson and Hahn-Hägerdal, 1996 Pine 44e46.4 8.8e26 29.4 Wayman and Parekh, 1990; Olsson and Hahn-Hägerdal, 1996 Hemlocks 47.5 22.0 28.5 Wayman and Parekh, 1990 HARD WOOD SOFT WOOD AGRICULTURAL FEEDSTOCKS/RESIDUE Sugarcane Bagasse 33 30 29 Neureiter et al., 2002 Sorghum Bagasse 44.4 35.5 3.9 Dogaris et al., 2009 Wheat Straw 37e38.2 21.2e29 15e23.4 Wiselogel et al., 1996; Lee et al., 2007a Corn Stover 37.5e26 22.4e29 17.6e19 Zhu et al., 2008; Lee et al., 2007a Rice Straw 33.0 26.0 7.0 Severe and ZoBell, 2012 Barley Straw 43.3 29.6 7.7 Severe and ZoBell, 2012 Oat Straw 41.0 16.0 11.0 Mussatto and Teixeira, 2010; Severe and ZoBell, 2012 Sunflower 34.06e42.1 5.18e29.7 7.72e13.4 Mussatto and Teixeira, 2010; Tutt and Olt, 2011 Silage 39.27 25.96 9.02 Tutt and Olt, 2011 Jerusalem Artichoke 20.95e25.99 4.50e5.48 5.05e5.70 Tutt and Olt, 2011 Reed 49.40 31.50 8.74 Tutt and Olt, 2011 Coffee Grounds 8.6 37.6 NA Mussatto et al., 2011 Rye Straw 37.6 30.5 19.0 Mussatto and Teixeira, 2010 Soya Stalks 34.5 24.8 19.8 Mussatto and Teixeira, 2010 Leaves (Mixed Biomass) 15e20 80e85 0 Sun and Cheng, 2002; Harmsen et al., 2010 Nut Shells 25e30 25e30 30e40 Sun and Cheng, 2002; Harmsen et al., 2010 Orchard Grass 52.3 42.9 6.6 Jung and Vogel, 1986 Smooth Bromegrass 49.8 41.9 7.6 Jung and Vogel, 1986 Indiangrass 49.8 43.1 6.7 Jung and Vogel, 1986 Big Bluestem 47.6 47.4 4.5 Jung and Vogel, 1986 Ensiled Grass 37.85 27.33 9.65 Tutt and Olt, 2011 Coastal Bermuda grass 25 35.7 6.4 Sun and Cheng, 2002; Harmsen et al., 2010 Grasses (Mixed Biomass) 25e40 35e50 10e30 Sun and Cheng, 2002; Harmsen et al., 2010 Switchgrass (Perennial Grass) 31.0e37 20.4e29 17.6e19 Wiselogel et al., 1996; Lee et al., 2007b; Tutt and Olt, 2011 Miscanthus 40e42 18e30.15 7e25 Sørensen et al., 2008; Tutt and Olt, 2011 Alfalfa 33 18 8 Sreenath et al., 2001 ENERGY CROPS (Continued )
32 TABLE 2.4 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Cell Wall Compositions (%) of Different Lignocellulosic Sourcesdcont’d Biomass Type Cellulose Hemicellulose Lignin References Hemp 53.86 10.60 8.76 Tutt and Olt, 2011 Jatropha 34 10 12 Singh et al., 2008, Abreu, 2009; Jingura et al., 2010; Yamamura et al., 2012 Algae 7.1 16.3 1.52 Ververis et al., 2007 Saccharum spontaneum 45.10 22.75 24.38 Chandel et al., 2009 Lantana camara 45.1 17.0 27.25 Pasha et al., 2007 Prosopis juliflora 45.5 20.38 24.65 Gupta et al., 2009 Eichhornia crassipes 18.2 48.7 3.50 Kumar et al., 2009 Crofton Weed Stem 37.6 22.4 16.4 Zhao et al., 2007 C. odorata (Siam Weed) 41.0 17.3 20.7 Zhao et al., 2010 Processed Paper/Black Paper 47 25 12 Ackerson et al., 1991 Waste Papers from Chemical Pulps 60e70 10e20 5e10 Sun and Cheng, 2002; Harmsen et al., 2010 Newspaper 40e61 25e40 18e30 Ackerson et al., 1991; Sun and Cheng, 2002 Brown Bin Waste/Food Waste 42.51e49.53 0.73e7.41 10.9e14.33 Komilis and Ham, 2003; Lamborn, 2009 Sorted Refuse 60 20 20 Sun and Cheng, 2002; Harmsen et al., 2010 Primary Wastewater Solids 8e15 NA 24e29 Sun and Cheng, 2002; Harmsen et al., 2010 Solid Cattle Manure 1.6e4.7 1.4e3.3 2.7e5.7 Sun and Cheng, 2002; Harmsen et al., 2010 Poultry Waste 11 16 4 FAO, 1980 Spent Mushroom Compost 38 19 25 Jordan et al., 2008 Swine Waste 6.0 28 NA Sun and Cheng, 2002; Harmsen et al., 2010 Dried Distilled Grains with Solubles (DDGS) 16e22 8.2e15 0e3.1 Blaschek and Ezej, 2007; Kim et al., 2008; Pasangulapati et al., 2012, Eastern Red Cedar 40.3 8.5 35.9 Pasangulapati et al., 2012 Poplar 39.8 14.8 29.1 Blaschek and Ezej, 2007 WEEDS SOLID WASTE FOREST RESIDUE NA, data not available. complex. Ethanol can be produced from lignocellulosic materials in various ways. The main difference between the process alternatives is the hydrolysis steps, which as mentioned previously, can be performed by dilute acid, concentrated acid or enzymatically. Some of the process steps are more or less the same, independent of the hydrolysis method used. For example, enzyme production will be omitted in an acid hydrolysis process; likewise, the recovery of acid is not necessary in an enzyme hydrolysis process (Galbe and Zacchi, 2002). To achieve lower production costs, the sustainable supply of cheap raw materials is a necessity. It is also essential to ensure that all components of the biomass are utilized and resulting by-products and wastes are used in a biorefinery system. When lignocellulosic raw materials are used, the main by-product is lignin, which can be used as an ash-free solid fuel for production of heat and/or electricity, for which there are no foreseeable market limits. However, in addition, lignin can be used for a range of additional high-value products that have the potential to enhance overall process economics significantly (Azadi et al., 2013; Lange et al., 2013; Doherty et al., 2011; Collinson and Thielemans, 2010). Accordingly, it will only be possible to produce large amounts of low-cost ethanol if lignocellulosic feedstocks such as fast-growing trees, grass, aquatic plants, waste products (including agricultural and forestry residues) and municipal and industrial waste are used
BIOENERGY RESOURCES AND BIOFUELS DEVELOPMENT PROGRAM (Van Dyck and Pletschke, 2012; Wheals et al., 1999). The potential of using lignocellulosic biomass for energy production is even more apparent when one realizes that it is the most abundant renewable organic component in the biosphere (Claassen et al. 1999). Currently enzyme hydrolysis has high yields (70e85%) of bioconversion, and improvements are still possible (85e95%) (Van Dyck and Pletschke, 2012; Sills and Gossett, 2011; Redding et al., 2010; Hu and Wen, 2008). BIOENERGY RESOURCES AND BIOFUELS DEVELOPMENT PROGRAM Current bioenergy resources consist of residues from forestry and agriculture, various organic waste streams and dedicated biomass production from pasture land, wood plantations and sugar cane (Figure 2.2). At present, the main biomass feedstocks for electricity and heat generation are forestry and agricultural residues and municipal waste in cogeneration and cofiring power plants. In the longer term, lignocellulosic crops could provide bioenergy resources for second-generation biofuels, which are considered more sustainable, provide land use opportunities and will reduce the competition with food crops (http://www.ga.gov.au/image_cache/ GA16706.pdf). Major feedstock sources for future biofuel production are likely to be high biomass producing plant species such as poplar, pine, switchgrass, sorghum maize, Miscanthus, hemp, Jatropha, willow and cassava. With 33 growing interest in the utilization of plant biomass for the production of ethanol and other biofuels, the use of plant species as biofuel feedstocks has become a focal point in research. Due to concerns about diverting grain and seed from human food and livestock feed to biofuel feedstock production, emphasis has shifted to the use of lignocellulose-derived biofuel production, and research is now directed at improving not only lignocellulosic yield but also quality traits in these species (Banerjee, 2011; Mueller et al., 2011; Tyner, 2010). A long-term opportunity exists to produce fuels from nonedible lignocellulosic biomass from plants (Heather and Somerville, 2012). Sugarcane, energy cane, elephant grass, switchgrass, and Miscanthus have intrinsically higher light, water and nitrogen use efficiency and are fast-growing biomass/crops for bioenergy work program. Work on perennial grasses such as switchgrass (Panicum spp.), prairie cordgrass (Spartina spp.), big bluestem (Andropogon spp.), little bluestem (Schizachyrium spp.) and others could produce significant biomass in a variety of biomass throughout the northern plains and southeastern grasslands in the United States (Gonzalez-Hernandez et al., 2009). Woody biomass can be harvested sustainably for lumber and paper and may, therefore, provide biofuel feedstock for some regions (Malmsheimer et al., 2011). Table 2.5 summarizes the countrywise contribution of current biofuel yield from different feedstocks. As mentioned previously, biomass energy can come from numerous sources and produce several types of fuels. Ethanol is typically produced from biomass FIGURE 2.2 Share of biomass sources in the world. Source: IEA, 2009. (For color version of this figure, the reader is referred to the online version of this book.)
34 TABLE 2.5 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Countrywise Contribution of Current Biofuel Yield from Their Available Feedstocks Crop Yield Conversion Efficiency Total Biofuel Yield Crop Global/Country Wise Estimates Biofuel Types (Tons/ha) (l/ton) (l/ha) Sugar Beet Global Bioethanol 46.0 110 5060 SugarCane Global Bioethanol 65.0 70 4550 Cassava Global Bioethanol 12.0 180 2070 Maize Global Bioethanol 4.9 400 1960 Rice Global Bioethanol 4.2 430 1806 Wheat Global Bioethanol 2.8 340 952 Sorghum Global Bioethanol 1.3 380 494 Sugar Cane Brazil Bioethanol 73.5 74.5 5476 Sugar Cane India Bioethanol 60.7 74.5 4522 Oil Palm Malaysia Biodiesel 20.6 230 4736 Oil Palm Indonesia Biodiesel 17.8 230 4092 Maize United States Bioethanol 9.4 399 3751 Maize China Bioethanol 5.0 399 1995 Cassava Brazil Bioethanol 13.6 137 1863 Cassava Nigeria Bioethanol 10.8 137 1480 Soybean United States Biodiesel 2.7 205 552 Soybean Brazil Biodiesel 2.4 205 491 Jatropha India Biodiesel 2.0 340 680 Sources: Rajagoapl and Zilberman, 2007, Naylor et al., 2007, FAO, 2008. high in carbohydrates (sugar, starch and cellulose) during a fermentation process. Recent developments in fermentation processes now allow almost any plant type to be used to produce ethanol. The most promising natural oils, such as rapeseed oil, have been used to produce biodiesel, which performs much like petroleum-derived diesel fuel. Apart from agricultural, forestry and other by-products, the main source of lignocellulosic biomass for second-generation biofuels is likely to be from “dedicated biomass feedstocks”, such as certain perennial grass and forest tree species. Genomics, genetic modifications and other biotechnologies are all being investigated as tools to produce plants with desirable characteristics for secondgeneration biofuel production, for example, plants that produce less lignin (a compound that cannot be fermented into liquid biofuel), plants that produce enzymes themselves for cellulose and/or lignin degradation, or plants that produce increased cellulose or overall biomass yields. Grass, leaves, agri crops, agricrop residues and currently available nonfood plant biomass are the dominant source of lignocellulosic materials (Carpita, 2012; Ambavaram et al., 2011; Abramson et al., 2010; Davison et al., 2006; Nguyen et al., 1999, 2000). Bioenergy resources used in current biofuels development programs, potential future resources and the related bioenergy outputs are summarized in Table 2.6. Bioenergy resources are difficult to estimate due to their multiple and competing uses. Production statistics exist for current commodities such as grain, sugar, pulp wood and saw logs; however, these commodities are currently largely committed to food, animal feed and materials markets. Potential feedstocks for the future include modified strains of existing crops, new tree crops and algae. There are many factors to be taken into account for each bioenergy resource, such as moisture content, resource location and distribution, and type of conversion process that is most suitable. Different sources of biomass require very different production systems and therefore a variety of sustainability issues can arise. These range from very positive benefits (e.g. use of waste material, or growing woody biomass on degraded agricultural land) through to large-scale diversion of high-input agricultural food crops for biofuels (O’Connell et al., 2009).
TABLE 2.6 Potential Resources and the Bioenergy Outputs Biomass Groups Current Resources Agriculture- Related Wastes and By-Products Livestock wastes: • Manure • Abattoir wastes solids By-products: • Wheat starch • Used cooking oil Electricity and heat generation Transport biofuel production Crop and food residues from harvesting and processing: • Large scale: rice husks, cotton ginning, and cereal straw • Small scale: maize cobs, coconut husks and nut shells • Crop stubble: The residue remaining after the harvest of grain crops such as wheat, barley and lupins • Grasses (various varieties including wild sorghum, kangaroo grass, tall fescue, perennial ryegrass) Electricity and heat generation Sugar Cane Bagasse (the stem residue remaining after the crushing to remove sugarrich juice from sugar cane), fibrous residues of sugar cane milling process sugar and C-molasses Electricity and heat generation Transport biofuel production Trash, leaves and tops from harvesting Electricity and heat generation Energy Crops High yield, short rotation crops grown specifically: • Sugar and starch crops • Oil-bearing cropsdsunflower, canola, juncea and soya beans • Palm oil • Jatropha (plant that produces seeds containing inedible oil content of 30e40% seed weight) Transport biofuel production Woody crops, genetically modified (GM) crops, tree crops, coppice (short rotation tree species, e.g. eucalyptus, poplar), woody weeds (e.g. camphor, laurel), new oilseed (Pongamia, camelina, and cotton seed), sugar (agave) crops, algae (micro and macro), and Halophytes (salt water and coastal/desert plant varieties, e.g. salicornia, marsh grasses, mangroves) Electricity and heat generation Transport biofuel production Forest and Forest Residues Wood from plantation forests Electricity and heat generation Wood from plantation forests, native forestry operations, bark, sawdust, pulpwood (wood used for processing into paper and related products) and harvest residues Electricity and heat generation Transport biofuel production Wood-Related Waste Saw mill residues: • Wood chips and saw dust Pulp mill residues: • Black liquor and wet wastes Electricity and heat generation Commercial and industrial waste, food-related wastes, garden organics, palettes, furniture, paper and cardboard material and urban timber Electricity and heat generation Meat and livestock by-product Electricity and heat generation Electricity and heat generation Landfill Gas Methane emitted from landfills mainly municipal solid wastes and industrial wastes Electricity and heat generation Sewage Gas Methane emitted from the solid organic components of sewage Electricity and heat generation Tallow Bioenergy Type Transport biofuel production 35 Source: Sustainable Aviation Fuel Road Map 2011; Batten and O’Connell 2007; IEA, 2006. Future Resources BIOENERGY RESOURCES AND BIOFUELS DEVELOPMENT PROGRAM Urban Solid Waste Bioenergy Type
36 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES SUSTAINABILITY Consideration of the sustainability of biomass to bioenergy programs based on utilizing lignocellulosic feedstocks is both timely and important in terms of the current plans for commercial valorization of this sector (Third International Conference on Lignocellulosic Ethanol; http://www.biofuelstp.eu/events/3rd-icleapril-2013.pdf). Sustainability of second-generation bioenergy is also been driven and supported by European and International directives and certification programs, including the Renewable Energy Directive 2009/28/EC (EU-RED), International Sustainability and Carbon Certification programs and standards, the Roundtable on Sustainable Biofuels and the Global Bioenergy Partnership (Scarlat and Dallemand, 2011). The sustainability of biomass to bioenergy programs has been a subject of great interest in Sweden, Canada and the western United States as well as in some Asian countries (Nguyen et al., 1999, 2000; Wu et al., 1999). The ecological and sustainable potential of biomass sources for fuel production is estimated to reach 130 TWh/year in Sweden by around 2020 (Parrika, 1997). Issues such as land use, environmental impact, logistics and resource management must be considered in terms of feedstock production. In addition, the sustainability of the bioconversion process(es) and downstream outputs, and the ability to meet REN and GHG emission targets must be carefully evaluated. High on the priority list of most national governments is the need to support rural development and sustain the local and national economies. Consequently, biomass to bioenergy programs need to be subjected to detailed life cycle analysis (LCA), where all of the aforementioned considerations are evaluated. LCA can also help derisk biomass to bioenergy processes (Buonocore et al., 2012). The use of conventional crops for energy use can also be expanded, with careful consideration of land availability and food demand. For sustainable bioenergy development lignocellulosic crops (both herbaceous and woody) could be produced on marginal, degraded and surplus agricultural lands and, in theory, could provide the bulk of the biomass resource in the medium term along with aquatic biomass (algae) as a significant contribution in the longer term (Richardson, 2008). However, significant progress needs to be made to scale-up algal production and processing in an economic manner to make algal biomass to bioenergy a commercially viable option. First-generation biofuels face both social and environmental challenges, largely because they use food crops that could lead to food price increases and possibly indirect land use change (ILUC). Nonfood biomass, e.g. lignocellulosic feedstocks such as organic wastes, forestry residues, high-yielding woody or grass energy crops and algae have the potential to provide possible solution to this problem, if developed and managed in a sustainable manner. The use of these feedstocks for second-generation biofuel production would significantly decrease the potential pressure on land use, improve GHG emission reductions when compared to some first-generation biofuels, and result in lower environmental and social risks (Bauen et al., 2009 IEA Report). The environmental impacts of conventional crop production have been researched in far greater detail than those of lignocellulosic crop production. Technically, the potential supply of energy from lignocellulosic biomass depends largely on the amount of land that is available for growing energy crops. In parallel, the need to meet the growing worldwide demand for food, protect biodiversity, manage soil and water reserves sustainably and fulfill additional socioeconomic objectives must be addressed. Bioenergy crop production can have positive impacts, for example, it can help to improve the soil structure and fertility of degraded lands. However, conversion of areas with sparse vegetation to high-yielding lignocellulosic plantations or ILUC may lead to substantial reductions in ground water recharge and water supply, which may lead to deteriorating conditions in water-scarce areas (Upham et al., 2011; Cabral et al., 2010; Smeets and Faaij, 2010). The cultivation of short rotation biomass crops may lead to nutrient removal or depletion (van den Broek et al., 2000), and important habitats may be lost through both land conversion and intensification (Pedroli et al., 2012). Aesthetic considerations also need to be considered in terms of the impact of cultivating and harvesting short rotation bioenergy crops (Hardcastle, 2006). Sound agricultural methods exist that can achieve major increases in feedstock productivity in neutral or positive environmental conditions in order to provide a continuous supply of energy crops/biomass waste, which can support the important role of bioenergy chains in socioeconomic development (Figure 2.3; Dornburg et al., 2008). The issue of biomass logistics is also a factor that needs careful consideration in terms of feedstock supply, processing technology selection, sitting of commercial production facilities and overall sustainability (Stephen et al., 2010). Recent studies have shown the potential of recycled wastewater for biomass production in an integrated natural water treatment approach (Fedler and Duan, 2011), which suggests that through innovative and careful consideration of environmental impacts solutions can be found that have multiple potential benefits. It has been suggested that the application of strict sustainability criteria, standards and a requirement for certification (Scarlat and Dallemand, 2011; Schubert and Blasch, 2010; van Dam et al., 2010) of feedstocks, land use and
SUSTAINABILITY 37 FIGURE 2.3 Sustainability of bioenergy crop supply chains and environmental effects. Source: Dornburg et al., 2008. (For color version of this figure, the reader is referred to the online version of this book.) bioenergy programs globally could both alleviate concerns and provide a more harmonized framework globally for sustainable development of secondgeneration bioenergy (Cornelissen et al., 2012; Van Stappen et al., 2011). Bioenergy Feedstocks and Dedicated Biofuel Crops There are two principal sources of biomass-based REN for second-generation bioenergy and biofuels: (1) wastes and residues from agriculture and forestry and (2) dedicated bioenergy crops. Wastes such as wood and agricultural residues, municipal wastes, and poultry litter are typically less expensive to supply to end point users, and are likely to play an important role in early development of commercial-scale REN supplies. However, analyses of future demand for REN indicates that these wastes may be capable of supplying only 14e30% of the total potential production of cellulosic ethanol and only approximately 18e60% of the production potential that could be derived from producing energy crops on currently idle or potentially available agricultural lands (Robert and Abbott, 2012; Brown, 2009; Lynd et al., 1991). Thus dedicated energy crops will be required to meet the demands of a growing REN market. Such crops, grown in the vicinity of the end point industrial user and specifically for the conversion process being used, offer important advantages of more systematic control of fuel quality, supply, and price stability than wastes derived from dispersed sources, which will be subject to alternative competitive end point uses and associated price fluctuations. The potential feedstocks for second-generation biofuel production considered in this study are biomass from crop residues, other nonfood energy crops, wood/forestry residues, Miscanthus, willow, hemp, Jatropha, switchgrasses and algae (Bauen et al., 2009). Lignocellulosic Feedstocks The major components of lignocellulosic feedstocks are cellulose and hemicellulose that can be converted to sugars through a series of thermochemical and biological processes and eventually fermented to bioethanol, other solvent biofuel or biogas. Therefore, lignocellulosic feedstocks are mainly categorized as agricultural residues (e.g. crop residues and sugarcane bagasse), forest residues, herbaceous and woody energy
38 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES crops. There are three principal technological end points for bioenergy crops: (1) conversion to liquid fuels, (2) combustion (alone or in combination with fossil fuels) to produce heat, steam, or electricity, and (3) gasification to simpler gaseous products that can have various uses. Agricultural residues can differ significantly in their chemical composition, which can lead to different bioenergy and biofuel yields per unit feedstock (Carriquiry et al., 2010). Commonly available agricultural residues for bioenergy production are barley straw, corn stover, rice straw, sorghum straw, wheat straw and sugarcane bagasse and having available carbohydrate contents of 70%, 58.3%, 49.3%, 61%, 54% and 67.2%, respectively. If the available carbohydrate of these feedstocks is fully converted to bioenergy, they can potentially yield approximately 367, 503, 392, 199, 1413 and 3133 liters biofuel/hectare, respectively (Carriquiry et al., 2010; US-DOE, 2008a, b; Kim and Dale, 2004). Forest residues include logging residues produced from harvest operations, fuel wood extracted from forestlands, and primary and secondary wood-processing mill residues (Perlack et al., 2005). Some important forest residues include hardwoods (black locust and hybrid poplar), softwoods (eucalyptus and pine) and switchgrass, which comprise approximately 57.15e66.45% carbohydrate, depending on the residue (Menon and Rao, 2012; van Dyck and Pletschke, 2012; Alves et al., 2010; Carriquiry et al., 2010; US-DOE, 2008a, b; Merino and Cherry, 2007; Hamelinck et al., 2005 Howard et al., 2003). However, several factors restrict the potential use of forest residues for biofuel production (Perlack et al., 2005). The first factor is the economic cost of transportation as limited accessibility largely increases the operational costs of logging/collection activities. Another factor is a potential reduction in recoverability of harvest areas due to environmental considerations (Richardson, 2008). Therefore a shift in research efforts to dedicated biofuel crops is taking place in order to have a continuous supply of feedstock for secondgeneration bioenergy requirements. Dedicated Bioenergy Crops From the industrial perspective, both fuel cost and quality relative to corresponding fossil fuels are essential considerations (Menon and Rao, 2012; Graham et al., 1995). Dedicated bioenergy crops can be broadly categorized into grassy (herbaceous or forage) and woody (tree) crops. Perennial forage crop species are a promising feedstock source for second-generation biofuels. Switchgrass (P. virgatum L.) is frequently mentioned because of its relatively low water and nutrition input and costs, positive environmental impact and adaptability to low-quality land (Keshwani and Cheng, 2009). Switchgrass has a wide and natural distribution from Central America to Southern Canada. Switchgrass as an energy crop is classified as a lignocellulosic crop because the cell walls primarily are digested to form sugars, which can subsequently be fermented to produce liquid fuels. Switchgrass research has been conducted cooperatively by the US Department of Agriculture and the University of Nebraska since the mid-1930s, with a primary focus on bioenergy since 1990 at several institutions. Progress has been made in switchgrass breeding and genetics, molecular genetics, fertility management, production economics and energetics, harvest and storage management, ecosystem services and ethanol yield. Other perennial forage crops such as alfalfa (M. sativa L.), reed canary grass (P. arundinacea L.), napier grass (Pennisetum purpureum Schumach.), and Bermuda grass (Cynodon spp.) also have potential as dedicated bioenergy crops (Carriquiry et al., 2010). The rationale for developing purpose-grown lignocellulosic crops for energy is that less intensive production techniques and poorer quality land can be used for these crops, thereby avoiding competition with food production on better quality land. A potential limitation of some forms of bioenergy is that biochemical composition, energy content, and contamination with alkali metals can limit their usefulness for certain industrial applications (Miles et al., 1993). Analysis of the energy content, the levels of alkali and ash and combustion properties of switchgrass indicates that this biomass is a versatile feedstock that is well suited for use in combustion, gasification, and liquid fuel production systems (Alonso et al., 2010; McKendry, 2002; McLaughlin et al., 1996). A complete field-validated biomass production system has been developed for the Midwest and Central Plains of the United States. Even with favorable economic and sustainability results from field trials, switchgrass for bioenergy has not been adopted on a large or commercial scale as yet. This is likely to be due to a number of factors, including the need for efficient conversion technologies, farmers’ reluctance to plant switchgrass without a viable bioenergy market and reluctance to build commercial biorefineries without a viable long-term feedstock supply already in place. Production, economic, net energy and sustainability research completed to date fully supports the use of switchgrass as a biomass energy crop (Mitchell et al., 2012). Another potential biofuel crop is Miscanthus sp., which has been used for forage and thatching in Japan for thousands of years, and managed through burning and grazing in vast prairies similar to those managed by Native American tribes in the central United States (Stewart et al., 2009). Miscanthus is a grass native to Asia and is a compelling herbaceous biomass feedstock for Europe (Lewandowski et al., 2003), in part because of its cold tolerance and low nitrogen requirements.
SUSTAINABILITY A drawback to the use of this species is that it takes 2e3 years to start full production as it must be established and propagated by rhizome cuttings. Other major limitations are (1) limited availability of genotype, (2) losses over winter, and (3) high costs in establishing the crop (Carriquiry et al., 2010; Lewandowski et al., 2003). Giant Miscanthus has been studied in the European Union and is used commercially in some member states for bedding, heat and electricity generation (Jones and Walsh, 2001). Most production currently occurs in England with some production also in Spain, Italy, Hungary, France, and Germany. Recently, a renewed interest in this native species has occurred in China and Japan and multiple research and commercialization projects have commenced. In the United States, research on the use of Miscanthus began at the University of Illinois at Urbana-Champaign in 2001 (Pyter et al., 2007) and has expanded rapidly to other US universities. Giant Miscanthus has been proposed for use in the United States in combined heat and power generation, as a supplement or on its own (Khanna et al., 2008; Heaton et al., 2008). It is also a leading candidate feedstock for cellulosic ethanol (Department of Energy (DOE), 2006). Although it is widely touted for cellulosic ethanol, giant Miscanthus has traits that are likely to make it better suited for thermochemical conversion processes over biological fermentation, at least using existing technology (Williams and Douglas, 2011). Reed canary grass is commonly used for hay and forage. It is well adapted to temperate agroeconomic regions and to weathered soils (Carlson et al., 1996). Reed canary grass can be slow to establish and can become an invasive species in native wetland (Merigliano and Lesica, 1998). Alfalfa is a forage crop that can be used to both supply biomass feedstock and high-quality animal feed (Delong et al., 1995). Several other subtropical and tropical grasses have been explored as potential biomass feedstocks in the United States, including Bermuda grass (Boateng et al., 2007), napier grass (Schank et al., 1993), eastern gamagrass and prairie cordgrass (Carriquiry et al., 2010; Boe and Lee, 2007; Springer and Dewald, 2004). Dedicated fast-growing woody energy crops with potential include fast-growing tree species. Important attributes include the relatively high yield potential, wide geographical distribution and relatively low levels of nutrient and manpower input needed when compared to annual crops, as well as their versatility as a source of solid and liquid energy has also been highlighted (Smeets et al., 2007). Poplar (Populus spp.), willow (Salix spp.), and eucalyptus are among the species most frequently mentioned in this latter category. Biofuels produced from short rotation coppice species like willow could help reduce dependence on fossil fuels. To maximize yields per hectare, light interception and utilization of the plant canopy need to be optimized 39 (Cunniff and Cerasuolo, 2011; Carriquiry et al., 2010). Poplar and willow have been grown successfully using municipal waste-derived fertilizers and irrigated with municipal or industrial wastewater, thereby decreasing two waste streams yet achieving nutrient and water inputs needed for high yields (Powlson et al., 2005). Pressure to increase the use of woody biomass for bioenergy and biofuel production could lead to conversion of forests to plantations with short rotation tree species, e.g. poplar (Populus spp.) and willow (Salix spp.) (Karp et al.,2011; Carriquiry et al., 2010; Zalesny et al., 2009). Cellulosic ethanol is derived from grasses, agricrop and wood residues and fast-growing trees (such as poplar or willow) and typically yields >10 times more energy than is needed to produce the fuel (Carriquiry et al., 2010); Powlson et al., 2005. However, with the case of crop and forest residues, the logistics of feedstocks obtained from dedicated energy crops is still a challenging issue to be resolved as these feedstocks are bulky and difficult to transport. Feedstocks for Biodiesel Jatropha (Jatropha curcas) is an oilseed species that has generated the most excitement in recent years in terms of its potential as a feedstock for biodiesel production. It is a multipurpose bush or low-growing tree, native to tropical America that can be used as a hedge, to reclaim land and as a commercial crop (Carriquiry et al., 2010; Azam et al., 2005; Openshaw, 2000). Jatropha is now grown in many tropical and subtropical regions within Asia and Africa. The oil derived from Jatropha has been shown to yield a biodiesel that meets European and US quality standards (Pandey et al., 2012; Akbar et al., 2009; Azam et al., 2005). Jatropha is known as a diesel fuel plant; the seed can yield a substantial quantity of oil that can be converted to biodiesel without prior refining (Carriquiry et al., 2010; Becker and Makkar, 2009). This plant is currently underutilized but could help in meeting the challenges of global biofuel demand (37 billion gallon) by 2016. Jatropha can be grown in semiarid conditions and/or marginal soils without large investment inputs (Jongschaap et al., 2007). While nonedible and toxic to humans and some animals (toxic substances include toxalbumin curcin, phorbol, saponins, trypsin inhibitor and a toxic lectin; Rakshit et al., 2013; Pimentel et al., 2012; Carels, 2009), its oil can be burnt directly or processed into biodiesel, which makes it an especially attractive biofuel crop in remote rural areas (Akbar et al., 2009; Jongschaap et al., 2007). The interest in Jatropha has been fueled by very optimistic claims of a concurrent capability to producing high oil yields and recovering wasteland (Achten et al., 2008). However, to date, critical questions remain regarding its ability to be economically viable
40 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES when grown in poor environmental conditions. Attainment of consistently high yields has only been achieved with relatively high levels of nutrient inputs and on good soils (International Energy Agency (IEA) Bioenergy, 2008). Nonetheless, the possibility of cultivating energy crops such as J. curcas L. has the potential to enable some smallholder farmers, producers and processors to improve their economic and social conditions, and support rural development. In addition to growing on degraded and marginal lands, this crop has special appeal, in that it grows under drought conditions and animals do not graze on it (Pandey et al., 2012; Carriquiry et al., 2010). Another important biodiesel feedstock are microalgae, which comprise a diverse group of aquatic photosynthetic microorganisms that grow rapidly and have the capability to yield large quantities of lipids adequate for biodiesel production (Ahmad et al., 2011; Amaro et al., 2011; Singh et al., 2011; Carriquiry et al., 2010; Mata et al., 2010; Li et al., 2008; Chisti, 2007; World Watch Institute (WWI), 2007). Algae were initially investigated as a potential source of fuel during the gas scare of the 1970s (Li et al., 2008). The National Renewable Energy Laboratory (NREL) started its algae feedstock studies in the late 1970s, but its research program was discontinued in 1996. Recent renewed interest has led the NREL to restart its research into the bioenergy/biodiesel potential of algae (Donovan and Stowe, 2009). The potential for algae to provide biomass for biodiesel production is now widely accepted. Furthermore, algae are recognized among the most efficient raw material for this purpose, and some studies (Carriquiry et al., 2010; Chisti, 2007) assert microalgae represent the “only source of biodiesel that has the potential to completely displace fossil diesel”. One of the main advantages is the ability of microalgae to produce large amounts of biomass per unit of land. In addition, microalgae can be grown in saline water, coastal seawater, freshwater and on nonarable land, hence reducing the competition for land with conventional agriculture (Khan et al., 2009), and creating economic opportunities in arid or salinity affected regions (Carriquiry et al., 2010; Schenk et al., 2008) Cultivation of microalgae, which is considered one of the major bottlenecks to commercial development, is being done mainly on open ponds, on closed bioreactors, and in hybrid systems (Brennan and Owende, 2010; Mata et al., 2010; Ugwu et al., 2008). While conventional open ponds are old systems for biomass production and account for the majority of microalgae cultivated today, closed bioreactors that achieve higher biomass productivity are being developed (Khan et al., 2009; Schenk et al., 2008). Open ponds are often perceived to be less expensive than bioreactors, as they require less capital and are cheaper to operate (Carriquiry et al., 2010; Khan et al., 2009). However, open ponds are more susceptible to contamination from unwanted species (Schenk et al., 2008), suffer from high water losses due to evaporation and reduced process control and reproducibility. Algal biomass production systems can be adapted to various levels of operational and technological skills; some microalgae yield chemically useful fatty acid profiles and an unsaponifiable fraction, which supports biodiesel production with high oxidation stability (Natrah et al., 2007; Minowa et al., 1995; Dote et al., 1994; Milne et al., 1990). In a biorefinery context, the lipid profiles of microalgae can also provide a valuable source of omega-3 fatty acids, such as docosahexaenoic acid and eicosapentaenoic acid (Yen et al., 2013; Doughman et al., 2007). Some important microalgal species are listed in Table 2.7 with their corresponding oil content. The physical and fuel properties of biodiesel from algal oil are comparable, in general, to those of fuel diesel (Amin, 2009; Rana and Spada, 2007; Miao and Wu, 2006). TABLE 2.7 Oil Content of Some Algae Species Oils (% Dry Matter of Lipid) FRESHWATER MICROALGAE Scenedesmus obliquus 11e55 Scenedesmus dimorphus 6e40 Chlorella vulgaris 14e56 C. emersonii 25e63 C. protothecoides 23/55 C. sorokiana 22 C. minutissima 57 Spirulina maxima 4e9 MARINE MICROALGAE Crypthecodinium cohnii 20e51.1 Dunaliella bioculata 8 D. salina 14e20 D. tertiolecta 16.7e71 Dunaniella sp. 17.5e67 Nannochloris sp. 20e56 Nannochloropsis sp. 12e53 Neochloris oleoabundans 29e65 Phaeodactylum tricornutum 18e57 Pyremnesium parvum 22e38 Skeletonema costatum 13.5e51.3 Tetraselmis suecica 8.5e23 Sources: Mata et al., 2010; Bruton et al., 2009; Gouveia and Oliveira 2009.
REFERENCES The use of microalgae could be a suitable alternative in the future, if improved high-rate production systems are available at scale, because these algae are one of the most efficient biological producers of oils on the planet and are a versatile biomass source (Demirbas, 2011; Mata et al., 2010; Macedo, 2007; Campbell, 1997). In fact, microalgae with a lower oil content (w30% of the dry biomass) could yield 58,700 L oil/hectare per year or 51,927 kg biodiesel/hectare per year. In comparison, Jatropha (J. curcas L.), with an oil content of 28% (dry weight), can yield 741 L oil/hectare or a biodiesel productivity of 656 kg biodiesel/hectare per year (Mata et al., 2010). On average, the biodiesel production yield from microalgae can be 10e20 times higher than the yield obtained from oleaginous seeds and/or vegetable oils (Mata et al., 2010; Gouveia and Oliveira, 2009; Chisti, 2007; Tickell, 2000). Therefore, in the future microalgae may become one of the Earth’s most important renewable fuel feedstocks for an number of reasons: their higher photosynthetic efficiency, biomass productivities, faster growth rates (in comparison with terrestrial plants), higher CO2 fixation and O2 production rates, and ability to grow in liquid medium, in variable climates and in ponds on nonarable land including marginal areas unsuitable for agricultural purposes (e.g. desert and seashore lands). Microalgae can also grow in nonpotable water or even in systems to combine waste treatment and biomass production (Zeng et al., 2012). They also use far less water than traditional crops and do not displace food crops; their production is not seasonal and biomass can be harvested daily (Chisti, 2007, 2008; Spolaore et al., 2006; Campbell, 1997). CONCLUSIONS In summary, achieving the feedstock yields to meet bioenergy requirements will generally require lignocellulosic crops rather than food crops. Pretreatment is likely to be required, and could be conducted close to the site of harvesting, as the pretreated biomass would be reduced in bulk, and thus cheaper to transport. The ideal pretreatment should be low cost, yield minimum levels of inhibitory compounds, result in a minimum loss of the main polysaccharides and enable maximum recovery of different fractions from the biomass. Pretreated biomass is also more amenable to downstream enzymatic bioconversion. There are major challenges ahead to reduce bioenergy production costs, many of which can provide significant opportunities for fundamental research and innovation in science and engineering. Bioenergy production, especially from second- and third- generation feedstocks, can yield many socioeconomic benefits. Selection of 41 the appropriate feedstocks in combination with positive sustainable agronomic and resource management approaches will reduce global dependency on fossil fuels. However, well-integrated and well-conceived strategies are required so that bioenergy can maintain the environment, support biodiversity, conserve water resources, lead to a reduction in emissions and enable rural development. Lignocellulosic biomass has several advantages over conventional sugar- and starch-based raw materials and has been projected to be one of the main sources of bioenergy and biofuels in the near future. With the application of existing technologies and future advances, biomass to bioenergy can provide a significant positive alternative in the energy and biofuel sector. Acknowledgments The authors are grateful for research funding from Enterprise Ireland and the Industrial Development Authority, through the Technology Centre for Biorefining and Bioenergy (TCBB), as part of the Competence Centre program under the National Development Plan 2007e2013. The support of Mr B. Bonsall, Technology Leader (TCBB), and Prof. V. O’Flaherty, Chair of Microbiology, School of Natural Sciences, & Deputy Director of the Ryan Institute for Environmental, Marine and Energy Research at NUI Galway, Ireland, is gratefully acknowledged. References Abramson, M., Shoseyov, O., Shani, Z., 2010. Plant cell wall reconstruction toward improved lignocellulosic production and processability. Plant. Sci. 178, 61e72. Abreu, F., 2009. Alternative By-Products from Jatropha. http://www. ifad.org/events/Jatropha/harvest/F_Abreu.ppt. Achten, W.M.J., Verchot, L., Franken, Y.J., Mathijs, E., Singh, V.P., Aerts, R., Muys, B., 2008. Jatropha bio-diesel production and use. Biomass Bioenergy 32, 1063e1084. Ackerson, M.D., Clausen, E.C., Gaddy, J.L., 1991. Production of ethanol from MSW via concentrated acid hydrolysis of the lignocellulosic fraction. Energy Biomass Wastes 15, 725e743. Aden, M., Ruth, K., Ibsen, J., et al., 2002. Lignocellulosic Biomass to Ethanol Process Design and Economics Utilizing Co-Current Dilute Acid Pre-hydrolysis and Enzymatic Hydrolysis for Corn Stover (Report No. NREL/TP-510e32438). National Renewable Energy Laboratory, USA. Agbor, V.B., Cicek, N., Sparling, R., Berlin, A., Levin, D.B., 2011. Biomass pretreatment: fundamentals towards application. Biotechnol. Adv. 29, 675e685. Ahmad, A.L., Yasin, N.H.M., Derek, C.J.C., Lim, J.K., 2011. Microalgae as a sustainable energy source for biodiesel production: review. Renewable Sustainable Energy Rev. 15, 584e593. Akbar, E., Yaakub, Z., Kamarudin, S.K., Ismail, M., Salimon, J., 2009. Characteristics and composition of Jatropha curcas oil seed from Malaysia and its potential as biodiesel feedstock. Eur. J. Sci. Res. 29, 396e403. Alonso, D.M., Bond, J.Q., Dumesic, J.A., 2010. Catalytic conversion of biomass to biofuels. Green Chem. 12, 1493e1513. Alves, F.F., Bose, S.K., Francis, R.C., Colodette, J.L., Iakovlev, M., Heiningen, A.V., 2010. Carbohydrate composition of eucalyptus, bagasse and bamboo by a combination of methods. Carbohydr. Polym. 82, 1097e1101.
42 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Amaro, H.M., Guedes, A.C., Malcata, F.X., 2011. Advances and perspectives in using microalgae to produce biodiesel. Appl. Energy 88, 3402e3410. Ambravaram, M.M.R., Krishnan, A., Trijatmiko, K.R., Pereira, A., 2011. Coordinated activation of cellulose and repression of lignin biosynthetic pathways in rice. Plant. Physiol. 155, 916e931. Amin, S., 2009. Review on biofuel oil and gas production processes from microalgae. Energy Conversion and Management 50, 1834e1840. Amore, A., Faraco, V., 2012. Potential of fungi as category I consolidated BioProcessing organisms for cellulosic ethanol production. Renewable Sustainable Energy Rev. 16, 3286e3301. Antolı́n, G., Tinaut, F.V., Briceño, Y., Castaño, V., Pérez, C., Ramı́rez, A.I., 2002. Optimisation of biodiesel production by sunflower oil transesterification. Bioresour. Technol. 83, 111e114. Azadi, P., Inderwildi, O.R., Farnood, R., King, D.A., 2013. Liquid fuels, hydrogen and chemicals from lignin: a critical review. Renewable Sustainable Energy Rev. 21, 506e523. Azam, M.M., Waris, A., Nahar, N.M., 2005. Prospects and potential of fatty acid methyl ester of some nontraditional seed oils for use as biodiesel. Ind. Biomass Bioenergy 29, 293e302. Balat, M., 2011. Production of bioethanol from lignocellulosic materials via the biochemical pathway: a review. Energy Convers. Manage. 52, 858e875. Banerjee, A., 2011. Food, feed, fuel: transforming the competition for grains. Dev. Change 42, 529e557. Batten, D., O’Connell, D., 2007. Biofuels in Australia: Some Economic and Policy Considerations. RIRDC Publication No07/177. Rural Industries Research and Development Corporation and CSIRO, Canberra. Bauen, A., Gerndes, G., Junginger, M., Londo, M., Vuille, F., Ball, R., Bole, T., Chudziak, C., Faaij, A., Mozafferian, H., 2009. Bioenergy e A Sustainable and Reliable Energy Source: A Review of Status and Prospects. IEA BIOENERGY. ExCo: 2009:06. Becker, K., Makkar, H.P.S., 2009. Jatropha curcas: a potential source for tomorrow’s oil and biodiesel. Lipid Toxicol. 20, 104e107. Betts, W.B., Dart, R.K., Ball, A.S., Pedlar, S.L., 1991. Biosynthesis and structure of lignocellulose. In: Betts, W.B. (Ed.), Biodegradation: Natural and Synthetic Materials. Springer, Berlin, Germany, pp. 139e155. Blanch, H., 2012. Bioprocessing for biofuels. Curr. Opin. Biotechnol. 23, 390e395. Blaschek, H.P., Ezej, T.C., 2007. Science of alternative feedstocks. In: Corn-Based Ethanol in Illinois and the US: A Report from the Department of Agricultural and Consumer Economics. University of Illinois, pp. 112e128. Boateng, A., Anderson, W., Phillips, J., 2007. Bermuda grass for biofuels: effect of two genotypes on pyrolysis product yield. Energy Fuels 21, 1183e1187. Boe, A., Lee, D., 2007. Genetic variation for biomass production in prairie cordgrass and switchgrass. Crop. Sci. 47, 929e934. Brennan, L., Owende, P., 2010. Biofuels from microalgae e a review of technologies for production, processing and extraction of biofuels and co-products. Renewable Sustainable Energy Rev. 14, 557e577. Brown, J.N., 2009. Development of a Lab-Scale Auger Reactor for Biomass Fast Pyrolysis and Process Optimization Using Response Surface Methodology. Graduate Theses and Dissertations. Paper 10996, Iowa State University, USA. Bruton, T., Lyons, H., Lerat, Y., Stanley, M., BoRasmussen, M., 2009. A Review of the Potential of Marine Algae as a Source of Biofuel in Ireland. Sustainable Energy, Ireland. Buonocore, E., Franzese, P.P., Ulgiati, S., 2012. Assessing the environmental performance and sustainability of bioenergy production in Sweden: a life cycle assessment perspective. Energy 37, 69e78. Cabral, O.M.R., Rocha, H.R., Gash, J.H.C., LIgo, M.A.V., Freitas, H.C., Tatsch, J.D., 2010. The energy and water balance of a Eucalyptus plantation in southeast Brazil. J. Hydrol. 388, 208e216. Campbell, C.J., 1997. The Coming Oil Crisis. Multi-science Publishing Company and petroconsultants S.A, Essex, England. Carels, N., 2009. Jatropha curcas: a review. Adv. Bot. Res. 50, 39e86. Carlson, I.T., Oram, R.N., Surprenant, J., 1996. Reed canary grass and other Phalaris species. In: Moser, L.E., Buxton, D.R., Casler, M.D. (Eds.), Cool-Season Forage Grasses. Am. Soc. Agron, Madison, WI, pp. 569e604. Carpita, N., 2012. Progress in the biological synthesis of the plant cell wall: new ideas for improving biomass for bioenergy. Curr. Opin. Biotechnol. 23, 330e337. Carriquiry, Miguel A., Xiaodong, Du, Timilsina, Govinda R., 2010. Second-Generation Biofuels: Economics and Policies. Policy Research Working Paper 5406. http://econ.worldbank.org. Chandel, Anuj K., Singh, Om V., 2011. Weedy lignocellulosic feedstock and microbial metabolic engineering: advancing the generation of ‘Biofuel’. Appl. Microbiol. Biotechnol. 89, 1289e1303. Chandel, A.K., Narasu, M.L., Chandrasekhar, G., Manikeyam, A., Rao, L.V., 2009. Use of Saccharum spontaneum (wild sugarcane) as biomaterial for cell immobilization and modulated ethanol production by thermotolerant Saccharomyces cerevisiae VS3. Bioresour. Technol. 100, 2404e2410. Chandel, A.K., Singh, O.V., Rao, L.V., 2010. Biotechnological applications of hemicellulosic derived sugars: state-of-the-art. In: Singh, O.V., Harvey, S.P. (Eds.), Sustainable Biotechnology: Renewable Resources and New Perspectives. Springer, Netherland, pp. 63e81. Cherubini, F., 2010. The biorefinery concept: using biomass instead of oil for producing energy and chemicals. Energy Convers. Manage. 51, 1412e1421. Chisti, Y., 2007. Biodiesel from microalgae. Biotechnol. Adv. 25, 294e306. Chisti, Y., 2008. Biodiesel from microalgae beats bioethanol. Trends. Biotechnol. 26, 126e131. Chu, S., Majumdar, A., 2012. Opportunities and challenges for a sustainable energy future. Nature 488, 294e303. Claassen, P.A.M., Sijtsma, L., Stams, A.J.M., De Vries, S.S., Weusthuis, R.A., 1999. Utilisation of biomass for the supply of energy carriers. Appl. Microbiol. Biotechnol 52, 741e755. Collinson, S.R., Theilemans, W., 2010. The catalytic oxidation of biomass to new materials focusing on starch, cellulose and lignin. Coord. Chem. Rev. 254, 1854e1870. Cornelissen, S., Koper, M., Deng, Y.Y., 2012. The role of bioenergy in a fully sustainable global energy system. Biomass Bioenergy 41, 21e33. Coyne, J., Gupta, V.K., O’Donovan, A., Tuohy, M.G., 2013. The role of fungal enzymes in global biofuel production technologies. In: Gupta, V.K., Tuohy, M.G. (Eds.), Biofuel Technologies. Springer Science Publishers, USA, pp. 121e143. Cunniff, J., Cerasuolo, M., 2011. Lighting the way to willow biomass production. J. Sci. Food Agric. 91, 1733e1736. Dashtban, M., Schraft, H., Qin, W., 2009. Fungal bioconversion of lignocellulosic residues; opportunities & perspectives. Int. J. Biol. Sci. 5, 578e595. Davison, B., Drescher, S., Tuskan, G., Davis, M., Ngheim, N., 2006. Variation of S/G ratio and lignin content in a Populus family influences the release of xylose by dilute acid hydrolysis. Appl. Biochem. Biotechnol 130, 427e435. De Fraiture, C., Berndes, G., 2009. Biofuels and water. In: Howarth, R.W., Bringezu, S. (Eds.), Biofuels: Environmental Consequences and Interactions with Changing Land Use. Proceedings of the Scientific Committee on Problems of the Environment (SCOPE) International Biofuels Project Rapid Assessment, 22e25 September 2008, Gummersbach Germany. Cornell University, Ithaca NY, USA, pp. 139e153. http://cip.cornell.edu/ biofuels/.
REFERENCES Delong, M.M., Swanberg, D.R., Oelke, E.A., Hanson, C., Onischak, M., Schmid, M.R., Wiant, B.C., August, 1995. Sustainable biomass energy production and rural economic development using alfalfa as a feedstock. In: Klass, D.L. (Ed.), August, 1995. Second Biomass Conf. of the Americas: Energy, Environment, Agriculture, and Industry, vol. 21e24, pp. 1582e1591. Portland, OR, USA. Demain, A.L., Newcomb, M., Wu, D.J.H., 2005. Cellulase, clostridia, and ethanol microbiology. Mol. Biol. Rev. 69, 124e154. Demain, A.L., 2009. Biosolutions to the energy problem. J.Ind.Microbiol. Biotechnol 36, 319e332. Demirbas, M.F., 2011. Biofuels from algae for sustainable development. Appl. Energy 88, 3473e3480. Demirbas, A., 2005. Bioethanol from cellulosic materials: a renewable motor fuel from biomass. Energy Sources 27, 327e337. Department of Energy (DOE), 2011. In: Quadrennial Technology Review Report. Technology Assessments, vol. 2. US Department of Energy. http://energy.gov/articles/department-energy-releasesinaugural-quadrennial-technology-review-report. Department of Energy (DOE), 2006. Breaking the Biological Barriers to Cellulosic Ethanol. A Joint Research Agenda (DOE/SC-0095). U.S. Department of Energy, Rockville, Maryland. Dien, B.S., Iten, L.B., Skory, C.D., 2005. Converting herbaceous energy crops to bioethanol; a review with emphasis on pretreatment processes. In: Hou, C.T. (Ed.), Handbook of Industrial Biocatalysis,. Taylor and Francis, Boca Raton, FL, pp. 1e11. (Chapter 23). Dogaris, I., Vakontios, G., Kalogeris, E., Mamma, D., Kekos, D., 2009. Induction of cellulases and hemicellulases from Neurospora crassa under solid-state cultivation for bioconversion of sorghum bagasse into ethanol. Ind. Crops. Prod. 29, 404e411. Doherty, W.O.S., Mousavioun, P., Fellows, C.M., 2011. Value-adding to cellulosic ethanol: lignin polymers. Ind. Crops Prod. 33, 259e276. Donovan, J., Stowe, N., 2009. Is the Future of Biofuels in Algae? RenewableEnergyWorld.com 06/12/2009. http://www. renewableenergyworld.com/rea/news/article/2009/06/is-thefuture-ofbiofuels-in-algae. Dornburg, V., Faaij, A., Langeveld, H., van de Ven, G., Wester, F., van Keulen, H., van Diepen, K., Ros, J., van Vuuren, D., van den Born, G.J., van Oorschot, M., Smout, F., Aiking, H., Londo, M., Mozaffarian, H., Smekens, K., Meeusen, M., Banse, M., Lysen, E., van Egmond, S., 2008. Biomass Assessment: Assessment of Global Biomass Potentials and Their Links to Food, Water, Biodiversity, Energy Demand and Economy. Report 500102 012. Dote, Y., Sawayama, S., Inoue, S., Minowa, T., Yokoyama, S., 1994. Recovery of liquid fuel from hydrocarbon-rich microalgae by thermo-chemical liquefaction. Fuel 73, 1855e1857. Doughman, S.D., Krupanidhi, S., Sanjeevi, C.B., 2007. Omega-3 fatty acids for nutrition and medicine: considering microalgae oil as a vegetarian source of EPA and DHA. Curr. Diabetes Rev. 3, 198e203. European Bioethanol Fuel Association, November, 2009. Production Data. FAO, 1980. Feed from Animal Wastes: State of Knowledge. http:// www.fao.org/docrep/004/X6518E/X6518E00.htm#TOC. FAO, 2008. The State of Food and Agriculture 2008. Biofuels: Prospects, Risks and Opportunities, Rome. Farine, D., O’Connell, D., Raison, J., May, B., O’Connor, M., Crawford, D., Alexander Herr - Herry, Herr, Taylor, J., Jovanovic, T., Campbell, P., Dunlop, M., Rodriguez, L., Poole, M., Braid, A., Kriticos, D., 2011. An assessment of biomass for bioelectricity and biofuel, and for greenhouse gas emission reduction in Australia. GCB Bioenergy 4, 148e175. Fedler, C.B., Duan, R., 2011. Biomass production for bioenergy using recycled wastewater in a natural waste treatment system. Resour., Conserv. Recycl. 55, 793e800. FitzPatrick, M., Champagne, P., Cunningham, M.F., Whitney, R.A., 2010. A biorefinery processing perspective: treatment of 43 lignocellulosic materials for the production of value-added products. Bioresour. Technol. 101, 8915e8922. ftp://ftp.fao.org/ docrep/fao/011/i0100e/i0100e02.pdf. Galbe, M., Zacchi, G., 2002. A review of the production of ethanol from softwood. Appl. Microbiol. Biotechnol. 59, 618e628. Girio, F.M., Fonseca, C., Caravalheiro, F., Duarte, L.C., Marques, A., Bogel-Lukasik, R., 2010. Hemicelluloses for fuel ethanol: a review. Bioresour. Technol. 101, 4775e4800. Gonzalez-Hernandez, J.L., Sarath, G., Stein, J.M., Owens, V., Gedye, K., Boe, A., 2009. A multiple species approach to biomass production from native herbaceous perennial feedstocks. In Vitro Cell. Dev. Biol.: Plant 45, 267e281. Gouveia, L., Oliveira, A.C., 2009. Microalgae as a raw material for biofuels production. J. Ind. Microbiol. Biotechnol. 36, 269e274. Graham, R.L., Lichtenberg, E., Roningen, V.O., Shapouri, H., Walsh, M., 1995. The Economics of Biomass Production in the United States. Proc. Second Biomass of the Americas Conference, Portland, OR. Groom, M.J., Gray, E.M., Townsend, P.A., 2008. Biofuels and biodiversity: principles for creating better policies for biofuel production. Conserv. Biol. 22, 602e609. Gupta, R., Sharma, K.K., Kuhad, R.C., 2009. Separate hydrolysis and fermentation (SHF) of Prosopis juliflora, a woody substrate, form the production of cellulosic ethanol by Saccharomyces cerevisiae and Pichia stipitis NCIM 3498. Bioresour. Technol. 100, 1214e1220. Gupta, V.K., Tuohy, M.G., Sharma, G.D., 2013. Biotechnology of Trichoderma: an overview. In: Gupta, V.K., Tuohy, M.G., Sharma, G.D., Gaur, S. (Eds.), Applications of Microbial Genes in Enzyme Technology. Nova Science Publishers, USA, pp. 375e393. Hahn-Hagerdal, B., Galbe, M., Gorwa-Grauslund, M., Liden, G., Zacchi, G., 2006. Bio-ethanol the fuel of tomorrow from the residues of today. Trends. Biotechnol. 24, 549e556. Hamelinck, C., Hooijdonk, G., Faaij, A., 2005. Ethanol from lignocellulosic biomass: techno-economic performance in short-, middleand long-term. Biomass Bioenergy 28, 384e410. Hardcastle, P.D., 2006. A Review of the Potential Impacts of Short Rotation Forestry. Forestry Commission, Farnham, UK. Harmsen, P.F.H., Huijgen, W.J.J., Bermúdez López, L.M., Bakker, R.R.C., 2010. Literature Review of Physical and Chemical Pretreatment Processes for Lignocellulosic Biomass. BioSynergy project. Food & Biobased Research Centre, Wageningen University, The Netherlands. Harris, D., DeBolt, S., 2010. Synthesis, regulation and utilization of lignocellulosic biomass. Plant Biotechnol. J. 8, 244e262. Hasunuma, T., Kondo, A., 2012. Development of yeast cell factories for consolidated bioprocessing of lignocellulose to bioethanol through cell surface engineering. Biotechnol. Adv. 30, 1207e1218. Hasunuma, T., Okazaki, F., Okai, N., Hora, K.Y., Ishii, K., Kondo, A., 2013. A review of enzymes and microbes for lignocellulosic biorefinery and the possibility of their application to consolidated bioprocessing. Bioresour. Technol. 135, 513e522. Heather, Y., Somerville, C., 2012. Development of feedstocks for cellulosic biofuels. F1000 Biology Reports. 4 (10). Heaton, E.A., Flavell, R.B., Mascia, P.N., Thomas, S.R., Dohleman, F.G., Long, S.P., 2008. Herbaceous energy crop development: recent progress and future prospects. Curr. Opin. Biotechnol. 19, 202e209. Hill, J., Nelson, E., Tilman, D., Polasky, S., Tiffany, D., 2006. Environmental, economic, and energetic costs and benefits of biodiesel and ethanol biofuels. Proc. Nat. Acad. Sci. USA 103, 11206e11210. Horn, S., Vaaje-Kolstad, G., Westereng, B., Eijsink, V., 2012. Novel enzymes for the degradation of cellulose. Biotechnol. Biofuels 5, 45. Howard, R., Abotsi, E., Jansen van Rensburg, E., Howard, S., 2003. Lignocellulose biotechnology: issues of bioconversion and enzyme production. Afr. J. Biotechnol. 2, 602e619. Http://www.ga.gov.au/image_cache/GA16706.pdf
44 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Http://www.thebioenergysite.com/articles/172/biofuels-andagriculture Hu, Z., Wen, Z., 2008. Enhancing enzymatic digestibility of switchgrass by microwave-assisted alkali pretreatment. Biochem. Eng. J. 38, 369e378. Hyeon, J.E., Jeon, S.D., Han, S.O., in press. Cellulosome-based, Clostridium-derived multi-functional enzyme complexes for advanced biotechnology tool development: Advances and applications. Biotechnol. Adv. (Available online: 3.04.13.). IEA, 2006. Energy Technology Perspectives e Scenario’s and Strategies to 2050. OECD/IEA, Paris, France. IEA, 2009. World Energy Balances 2009. OECD/IEA, Paris. International Energy Agency (IEA) Bioenergy, November, 2008. From First- to Second-Generation Biofuel Technologies: An Overview of Current Industry and RD&D Activities. IEA. International Energy Agency (IEA), 2010. Energy Technology Perspectives: Scenarios and Strategies to 2050. OECD/IEA, Paris, 706. Isroi, Millati, R., Syamsiah, S., Niklasson, C., Nur Cahyanto, M., Lundquist, K., Taherzadeh, M., 2011. Biological pretreatments of lignocelluloses with white-rot fungi and its applications: a review. BioResources 6, 5224e5259. Jingura, R.M., Musademba, D., Matengaifa, R., 2010. An evaluation of utility of Jatropha curcas L. as a source of multiple energy carriers. Int. J. Eng. Sci. Technol. 2, 115e122. Jones, M.B., Walsh, M., 2001. Miscanthus for Energy and Fibre. James and James Ltd., London. Jongschaap, R.E.E., Corre, W.J., Bindraban, P.S., Brandenburg, W.A., 2007. Claims and facts on Jatropha curcas L. Report 158. Plant Res. Int. (Wageningen). Jordan, S.N., Mullen, G.J., Murphy, M.C., 2008. Composition variability of spent mushroom compost in Ireland. Bioresour. Technol. 99, 411e418. Jung, H.G., Vogel, K.P., 1986. Influence of lignin on digestibility of forage cell wall material. J. Anim. Sci. 62, 1703e1712. Kamm, B., Kamm, M., 2004. Principles of biorefineries. Appl. Microbiol. Biotechnol. 64, 137e145. Kapdan, L.K., Kargi, F., 2006. Biohydrogen production from waste materials. Enzyme Microb. Technol. 38, 569e582. Karp, A., Hanley, S.J., Trybush, S.O., Macalpine, W., Pei, M., Shield, I., 2011. Genetic improvement of willow for bioenergy and biofuels. J. Integr. Plant Biol. 53, 151e165. Karunanithy, C., Muthukumarappan, K., 2011. Optimization of alkali soaking and extrusion of prairie cord grass for maximum sugar recovery by enzymatic hydrolysis. Biochem. Eng. J. 564, 71e82. Keating, C., Cysneiros, D., Mahony, T., O’Flaherty, V., 2012. The hydrolysis and biogas production of complex cellulosic substrates using three anaerobic biomass sources. Water. Sci. Technol. 67, 293e298. Keshwani, D., Cheng, J., 2009. Switchgrass for bioethanol and other value-added applications: a review. Bioresour. Technol. 100, 1515e1523. Khan, S.A., Rashmi, Hussain, M.Z., Prasad, S., Banerjee, U.C., 2009. Prospects of biodiesel production from microalagae in India. Renewable Sustainable Energy Rev. 13, 2361e2372. Khanna, M., Dhungana, B., Clifton-Brown, J., 2008. Costs of producing miscanthus and switchgrass for bioenergy in Illinois. Biomass Bioenergy 32, 482e493. Kim, E., Martinez Amezcua, J.C., Utterback, P.L., Parsons, C.M., 2008. Phosphorus bioavailability, true metabolizable energy, and amino acid digestibility of high protein corn distillers dried grains and dehydrated corn germ. Poult. Sci. 87, 700e705. Kim, S., Dale, E.B., 2004. Global potential bioethanol production from wasted crops and crop residues. Biomass Bioenergy 26, 361e375. Kittl, R., Kracher, D., Burgstaller, D., Haltrich, D., Ludwig, R., 2012. Production of four Neurospora crassa lytic polysaccharide monooxygenases in Pichia pastoris monitored by a fluorimetric assay. Biotechnol. Biofuels 5, 79. Komilis, D.P., Ham, R.K., 2003. The effect of lignin and sugars to the aerobic composting of solid waste. Waste Manage. 23, 419e423. Kuhad, R.C., Singh, A., Eriksson, K.E.L., 1997. Microorganisms and enzymes involved in the degradation of plant fiber cell walls. In: Eriksson, K.E.L. (Ed.), 1997. Advances in Biochemical Engineering Biotechnology, vol. 57, pp. 46e125. Kumar, A., Singh, L.K., Ghosh, S., 2009. Bioconversion of lignocellulosic fraction of water-hyacinth (Eichhornia crassipes) hemicellulose acid hydrolysate to ethanol by Pichia stipitis. Bioresour. Technol. 100, 3293e3297. Lamborn, J., 2009. Characterisation of Municipal Solid Waste Composition into Model Inputs. Third Workshop on the Hydroe PhysicoeMechanical Properties of Wastes (HPM3), Braunschweig, Germany. Lange, H., Decina, S., Crestini, C., in press. Oxidative upgrade of lignin e recent routes reviewed. Eur. Polym. J. Langston, J.A., Shaghasi, T., Abbate, E., Xu, F., Vlasenko, E., Sweeney, M.D., 2011. Oxidoreductive cellulose depolymerization by the enzymes cellobiose dehydrogenase and glycoside hydrolase 61. Appl. Environ. Microbiol. 77, 7007e7015. Lee, D.K., Owens, V.N., Boe, A., Jeranyama, P., 2007a. Composition of Herbaceous Biomass Feedstocks. A report by Plant Science Department, South Dakota State University. http://agbiopubs. sdstate.edu/articles/SGINC1-07.pdf. Lee, D.K., Owens, V.N., Doolittle, J.J., 2007b. Switchgrass and soil carbon sequestration response to ammonium nitrate, manure, and harvest frequency on conservation reserve program land. Agron. J. 99, 462e468. Lemus, R., Parrish, D.J., 2009. Herbaceous crops with potential for biofuel production in the USA. CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources 4 (no. 057). Lemus, R., Lai, R., 2005. Bioenergy crops and carbon sequestration. Crit. Rev. Plant Sci. 24, 1e21. Levin, D.B., Islam, R., Cicek, N., Sparling, R., 2006. Hydrogen production by Clostridium thermocellum 27405 from cellulosic biomass substrates. Int. J. Hydrogen Energy 31, 1496e1503. Lewandowski, I., Scurlock, J., Lindvall, E., Christou, M., 2003. The development and current status of perennial rhizomatous grasses as energy crops in the US and Europe. Biomass Bioenergy 25, 335e361. Li, Y., Horsman, M., Wu, N., Lan, C.Q., Dubois-Calero, N., 2008. Biofuels from microalgae. Biotechnol. Prog. 24, 815e820. Lichts, F.O., 2010. Industry Statistics: 2010 World Fuel Ethanol Production. Renewable Fuels Association. Liew, L.N., Shi, J., Li, Y., 2012. Methane production from solid-state anaerobic digestin of lignocellulosic biomass. Biomass Bioenergy 46, 125e132. Lynd, L.L., Cushman, J.H., Nichols, R.J., Wyman, C.F., 1991. Fuel ethanol from cellulosic biomass. Science 231, 1318e1323. Lynd, L.R., 1996. Overview and evaluation of fuel ethanol from cellulosic biomass: technology, economics, the environment, and policy. Annu. Rev. Energy. Environ. 21, 403e465. Lynd, L.R., Weimer, P.J., van Zyl, W.H., Pretorius, I.S., 2002. Microbial cellulose utilization: fundamentals and biotechnology. Microbiol. Mol. Biol. Rev. 66, 506e577. Macedo, I.C., 2007. Etanol de Cana de Acucar no Brasil. Presentation at the Seminar on Technologies for Future Ethanol Production in Brazil. Instituto Tecnologia Promon, Sao Paulo, Brazil. Malherbe, S., Cloete, T.E., 2003. Lignocellulose biodegradation: fundamentals and applications. Environ. Sci. Bioechnol. 1, 105e114. Malmsheimer, R.W., Bowyer, J.L., Fried, J.S., Gee, E., Izlar, R.L., Miner, R.A., Munn, I.A., Oneil, E., Stewart, W.C., 2011. Stewart: managing forests because carbon matters: integrating energy, products, and land management policy. J. For. 109, S7eS48.
REFERENCES Martinez, A.T., Ruiz-Duenas, F.J., Martinez, M.J., del Rio, J.C., Gutierrez, A., 2009. Enzymatic delignification of plant cell walls: from nature to mill. Curr. Opin. Biotechnol. 20, 348e357. Mata, T.M., Martins, A.A., Caetano, N.S., 2010. Microalgae for biodiesel production and other applications: a review. Renewable Sustainable Energy Rev. 14, 217e232. Mata-Alvarez, J., Mace, S., Llabres, P., 2000. Anaerobic digestion of organic solid wastes. An overview of research achievements and perspectives. Bioresour. Technol. 74, 3e16. Matano, Y., Hasunuma, T., Kondo, A., 2013. Cell recycle batch fermentation of high-solid lignocellulose using a recombinant cellulose-displaying yeast strain for high yield ethanol production in consolidated bioprocessing. Bioresour. Technol. 135, 403e409. McHugh, S., Carton, M., Mahony, T., O’Flaherty, V., 2003. Methanogenic population structure in a variety of anaerobic bioreactors. FEMS. Microbiol. Lett. 219, 297e304. McKendry, Peter, 2002. Energy production from biomass (part 3): gasification technologies. Bioresour. Technol. 83, 55e63. McLaughlin, S.B., Samson, R., Bransby, D., Weislogel, A., Sept. 1996. Evaluating Physical, Chemical, and Energetic Properties of Perennial Grasses as Biofuels. Proc. Bioenergy 96, Nashville, TN, pp. 1e8. Menon, V., Rao, M., 2012. Trends in bioconversion of lignocellulose: biofuels, platform chemicals & biorefinery concept. Prog. Energy Combust. Sci. 38, 522e550. Merigliano, M.F., Lesica, P., 1998. The native status of reed canary grass (Phalaris arundinacea L.) in the inland Northwest USA. Nat. Areas J. 18, 223e230. Merino, S.T., Cherry, J., 2007. Progress and challenges in enzyme development for biomass utilization. Adv. Biochem. Eng. Biotechnol. 108, 95e120. Miao, X., Wu, Q., 2006. Biodiesel production from heterotrophic microalgal oil. Bioresour. Technol. 97, 841e846. Miles, T.R., et al., 1993. Alkali Slagging Problems in Biomass Fuels. Proc. First Biomass Conference of the Americas, Burlington, VT, pp. 406e421. Milne, T.A., Evans, R.J., Nagle, N., 1990. Catalytic conversion of microalgae and vegetable oils to premium gasoline, with shape selective zeolites. Biomass 21, 219e232. Minowa, T., Yokoyama, S.Y., Kishimoto, M., Okakurat, T., 1995. Oil production from algal cells of Dunaliella tertiolecta by direct thermochemical liquefaction. Fuel 74, 1735e1738. Mitchell, Rob, Vogel Kenneth, P., Uden, Daniel R., 2012. The feasibility of switchgrass for biofuel production. Biofuels 3, 47e59. Mosier, N., Wyman, C.E., Dale, B.D., Elander, R.T., Lee, Y.Y., Holtzapple, M., Ladisch, M.R., 2005. Features of promising technologies for pretreatment of lignocelluolsic biomass. Bioresour. Technol. 96, 673e686. Mueller, S.A., Anderson, J.E., Wallington, T.J., 2011. Impact of biofuel production and other supply and demand factors on food price increases in 2008. Biomass Bioenergy 35, 1623e1632. Müller, A., Schmidhuber, J., Hoogeveen, J., Steduto, P., 2008. Some insights in the effect of growing bio energy demand on global food security and natural resources. Water Policy 10, 83e94. Mussatto, S.I., Teixeira, J.A., 2010. Lignocellulose as raw material in fermentation processes. In: Méndez-Vilas, A. (Ed.), Current Research, Technology and Education Topics in Applied Microbiology and Microbial Biotechnology, Microbiology Series, vol. 2. FORMATEX. Mussato, S.I., Dragone, G., Guimaraes, P.M.R., Silva, J.P.A., Carneiro, L.M., Roberto, I.C., Vicente, A., Domingues, L., Teixeira, J.A., 2010. Technological trends, global market, and challenges of bio-ethanol production. Biotechnology Advances 28, 817e830. Mussatto, S.I., Carneiro, L.M., Silva, J.P.A., Roberto, I.C., Teixeira, J.A., 2011. A study on chemical constituents and sugars extraction from spent coffee grounds. Carbohydr. Polym. 83, 368e374. 45 Natrah, F., Yosoff, F.M., Shariff, M., Abas, F., Mariana, N.S., 2007. Screening of Malaysian indigenous microalgae for antioxidant properties and nutritional value. J. Appl. Phycol. 19, 711e718. Naylor, R., Liska, A.J., Burke, M.B., Falcon, W.P., Gaskell, J.C., Rozelle, S.D., Cassman, K.G., 2007. The ripple effect: biofuels, food security, and the environment. Environment 49, 31e43. Neureiter, M., Danner, H., Thomasser, C., Saidi, B., Braun, R., 2002. Dilute acid hydrolysis of sugarcane bagasse at varying conditions. Appl. Biochem. Biotechnol. 98, 49e58. Nguyen, Q.A., Tucker, M.P., Keller, F.A., Eddy, F.P., 2000. Two-stage dilute acid pretreatment of softwoods. Appl. Biochem. Biotechnol. 84e86, 561e576. Nguyen, Q.A., Tucker, M.P., Keller, F.A., Beaty, D.A., Connors, K.M., Eddy, F.P., 1999. Dilute acid hydrolysis of softwoods. Appl. Biochem. Biotechnol. 77e79, 133e142. O’Connell, D., Braid, A., Raison, J., Handberg, K., Cowie, A., Rodriguez, L., George, B., November 2009. Sustainable Production of Bioenergy; a Review of Global Bioenergy Sustainability Frameworks and Assessment Systems. RIRDC Publication No.09/167. Rural Industries Research and Development Corporation and CSIRO, Canberra. O’Donovan, A., Gupta, V.K., Tuohy, M.G., 2013. Recent updates in acid pretreatments and SEM analysis of acid pretreated grass biomass. In: Gupta, V.K., Tuohy, M.G. (Eds.), Biofuel Technologies; Recent Developments. Springer Science Publishers, USA, pp. 97e118. OECD/FAO, 2009. Agricultural Outlook, 2009e2018. Olson, D.G., McBride, J.E., Shaw, J., Lynd, L.R., 2012. Recent progress in consolidated bioprocessing. Curr. Opin. Biotechnol. 23, 396e405. Olsson, L., Hahn-Hägerdal, B., 1996. Fermentation of lignocellulosic hydrolysates for ethanol production. Enzyme Microb. Technol. 18, 312e331. Ong, L.K., 2004. Conversion of lignocellulosic biomass to fuel ethanol a brief review. Planter 80, 517e524. Openshaw, K., 2000. A review of Jatropha curcas: an oil plant of unfulfilled promise. Biomass Bioenergy 19, 1e15. Pandey, V.M., Singh, K., Singh, J.S., Kumar, A., Singh, B., Singh, R.P., 2012. Jatropha curcas: a potential biofuel plant for sustainable environmental development. Renewable Sustainable Energy Rev. 16, 2870e2883. Pan, X., Xie, D., Gilkes, N., Gregg, D.J., Saddler, J.N., 2005. Strategies to enhance the enzymatic hydrolysis of pretreated softwood with high residual lignin content. Appl. Biochem. Biotechnol. (part A) 124, 1069e1079. Parrika, M., 1997. Biosimse a Method for the Estimation of Woody Biomass for Fuel in Sweden (Ph.D. thesis). Department of ForestryIndustry-Market Studies, Swedish University of Agricultural Sciences, Uppsala, Sweden. Pasangulapati, V., Ramachandriya, K.D., Kumar, A., Wilkins, M.R., Jones, C.L., Huhnke, R.L., 2012. Effects of cellulose, hemicellulose and lignin on thermochemical conversion characteristics of the selected biomass. Bioresour. Technol. 114, 663e669. Pasha, C., Valli, N., Rao, L.V., 2007. Lantana camara for fuel ethanol production using thermotolerant yeast. Lett. Appl. Microbiol. 44, 666e672. Pedroli, B., Elbersen, B., Frederiksen, P., Grandin, U., Heikkilä, R., Krogh, P.H., Izakraovicová, Z., Johansen, A., Meiresonne, L., Spijker, J., 2012. Is energy cropping in Europe compatible with biodiversity? Opportunities and threats to biodiversity from land-based production of biomass for bioenergy purposes. Biomass Bioenergy. Percival Zhang, Y.-H., 2008. Reviving the carbohydrate economy via multi-product lignocellulose biorefineries. J. Ind. Microbiol. Biotechnol. 35, 367e375. Perlack, R., Wright, L., Turhollow, A., Graham, R., Stokes, B., Erbach, D., April, 2005. Biomass as Feedstock for a Bioenergy and
46 2. BIOENERGY RESEARCH: AN OVERVIEW ON TECHNOLOGICAL DEVELOPMENTS AND BIORESOURCES Bioproducts Industry: The Technical Feasibility of a Billion-Ton Annual Supply. Department of Energy/GO-102005e2135. Pimentel, L.A., Riet-Correa, B., Dantas, A.F., Medeiros, R.M.T., RietCorrea, F., 2012. Poisoning by Jatropha ribifolia in goats. Toxicon 59, 587e591. Potumarthi, R., Baadhe, R.R., Jetty, A., 2012. Mixing of acid- and basepretreated corn cobs for improved production of reducing sugars and reduction in water use during neutralization. Bioresour. Technol. 119, 99e104. Potumarthi, R., Baadhe, R.R., Nayak, P., Jetty, A., 2013. One step biological pretreatment of rice husk by Phanerochaete chrysosporium for the production of reducing sugars. Bioresour. Technol. 128, 113e117. Powlson, D.S., Richie, A.B., Shield, I., 2005. Biofuels and other approaches for decreasing fossil fuel emissions from agriculture. Ann. Appl. Biol. 146, 193e201. Pyter, R., Voigt, T.B., Heaton, E.A., Dohleman, F.G., Long, S.P., 2007. Giant miscanthus: biomass crop for Illinois. In: Janick, J., Whipkey, A. (Eds.), Issues in New Crops and New Uses. ASHS Press, Alexandria, VA, pp. 39e42. Ragauskas, A.J., Williams, C.K., Davison, B.H., Britovsek, G., Cairney, J., Eckert, C.A., Frederick, W.J.J.R., Hallett, J.P., Leak, D.J., Liotta, C.L., 2006. The path forward for biofuels and biomaterials. Science 311, 484e489. Rajagopal, D., Zilberman, D., 2007. Review of Environmental, Economic and Policy Aspects of Biofuels. Development Research Group, The World Bank. Policy Research Working Paper WPS4341. Rajaram, S., Verma, A., 1990. Production and characterization of xylanase from Bacillus thermoalkalophilus growth on agricultural wastes. Appl. Microbiol. Biotechnol. 34, 141e144. Rakshit K., Devappa, J.S., Roach, H.P.S., Makkar, K.B., in press. Occular and dermal toxicity of Jatropha curcas phorbal esters. Ecotoxicol. Environ. Saf. Rana, R., Spada, V., 2007. Biodiesel production from ocean biomass. In: Proceedings of the Fifteenth European conference and exhibition. Berlin. Ravichandra, P., Rama, R.B., Sankar, B., 2013. Fermentable sugars from lignocellulosic biomass: technical challenges. In: Gupta, V.K., Tuohy, M.G. (Eds.), Biofuel Technologies: Recent Developments. Springer Verlag, Germany, pp. 3e27. Redding, A.P., Wang, Z., Keshwani, D.R., Cheng, J.J., 2011. High temperature dilute acid pretreatment of coastal Bermuda grass for enzymatic hydrolysis. Bioresour. Technol. 102, 1415e1424. Renewable Fuels Association, 2012. Accelerating Industry Innovationd2012, Ethanol Industry Outlook. Renewable Fuels Association, pp. 3e23. Richardson, J., 2008. Production of biomass for energy from sustainable forestry systems: Canada and Europe. Short Rotation Crops International Conference. Robert, F., Abbott, Z., 2012. Energy Crop Opportunities in the Western Upper Peninsula of Michigan. Michigan Technological University, USA. Sanchez, C., 2009. Lignocellulosic residues: biodegradation and bioconversion by fungi. Biotechnol. Adv. 27, 185e194. Sarkar, P., Bosneaga, E., Auer, M., 2009. Plant cell walls throughout evolution: towards a molecular understanding of their design principles. J. Exp. Bot. 60, 3615e3635. Sarkar, N., Ghosh, S.K., Bannerjee, S., Aikat, K., 2012. Bioethanol production from agricultural wastes. An overview. Renewable Energy 37, 19e27. Scarlat, N., Dallemand, J.-F., 2011. Recent developments of biofuels/ bioenergy sustainability certification: a global overview. Energy Policy 39, 1630e1646. Schank, S., Chenowyth, D., Turick, C., Mendoza, P., 1993. Napier grass genotypes and plant parts for biomass energy. Biomass Bioenergy 4, 1e7. Schenk, P., Thomas-Hall, S.R., Stephens, E., Marx, U.C., Mussgnug, J.H., Posten, O.Kruse, C., Hankamer, B., 2008. Second generation biofuels: high-efficiency microalgae for biodiesel production. Bioenergy Res. 1, 20e43. Schubert, R., Blasch, J., 2010. Sustainability standards for bioenergy e a means to reduce climate change risks? Energy Policy 38, 2797e2805. Severe, J., ZoBell, D.R., 2012. Technical Aspects for the Utilization of Small Grain Straws as Feed Energy Sources for Ruminants: Emphasis on Beef Cattle. AG/BeefCattle/2012e03. Department of Extension and Agriculture, Utah State University, USA. http://extension.usu.edu/files/publications/publication/AG_ BeefCattle_2012-03.pdf. Shapouri, H., Duield, J.A., Grabosk, M.S., 1995. Estimating the Net Energy Balance of Corn Ethanol. Agricultural Economic Report 721. US Department of Agriculture, USA, Washington, DC. Shapouri, H., Duield, J.A., Wang, M., 2002. The Energy Balance of Corn Ethanol: An Update. Agricultural Economic Report 813. US Department of Agriculture, USA, Washington, DC. Sills, D.L., Gossett, J.M., 2011. Assessment of commercial hemicellulases for saccharification of alkaline pretreated biomass. Bioresour. Technol. 102, 1389e1398. Simpson, T.W., Sharpley, A.N., Howarth, R.W., Paerl, H.W., Mankin, K.R., 2008. The new gold rush: fueling ethanol production while protecting water quality. Environ. Qual. 37, 318e324. Singh, A., Nigam, P.S., Murphy, J.D., 2011. Mechanism and challenges in commercialization of algal biofuels. Bioresour. Technol. 102, 26e34. Singh, R.N., et al., 2008. SPRERI experience on holistic approach to utilize all parts of Jatropha curcas fruit for energy. Renewable Energy 33, 1868e1873. Smeets, E.M.W., Faaij, A.P.C., 2010. The impact of sustainability criteria on the costs and potentials of bioenergy production e applied for case studies in Brazil and Ukraine. Biomass Bioenergy 34, 319e333. Smeets, E.M.W., Faaij, A., Lewandowski, I.M., Turkenburg, W.C., 2007. A bottom up assessment and review of global bio-energy potentials for 2050. Prog. Energy Combust. Sci. 33, 56e106. Somerville, C., Youngs, H., Taylor, C., Davis, S.C., Long, S.P., 2010. Feedstocks for lignocellulosic biofuels. Science 329, 790e792. Sørensen, A., Teller, P.J., Hilstrøm, T., Ahring, B.K., 2008. Hydrolysis of Miscanthus for bioethanol production using dilute acid presoaking combined with wet explosion pre-treatment and enzymatic treatment. Bioresour. Technol. 99, 6602e6607. Spolaore, P., Joannis-Cassan, C., Duran, E., Isambert, A., 2006. Commercial applications of microalgaedreview. J. Biosci. Bioeng. 101, 87e96. Springer, T., Dewald, C., 2004. Eastern gamagrass and other Tripsacum species. In: Moser, L., Burson, B., Sollenberger, L. (Eds.), Warmseason (C4) Grasses. American Society of Agronomy, Madison, WI. Sreenath, H.K., Koegel, R.G., Moldes, A.B., Jeffries, T.W., Straub, R.J., 2001. Ethanol production from alfalfa fiber fractions by saccharification and fermentation. Process Biochem. 36, 1199e1204. Stefan, B., Helmut, S., O’Brien, M., Kauppi, L., Howarth, R.W., McNeely, J., 2009. Towards Sustainable Production and Use of Resources: Accessing Biofuels. International Panel for Sustainable Resource Management. United Nations Environment Programme (UNEP). www.unep.org. Stephen, J.D., Mabee, W.E., Saddler, J.N., 2010. Biomass logistics as a determinant of second-generation biofuel facility scale, location and technology selection. Biofuels, Bioproducts Biorefin. 4, 503e518. Stewart, R.J., Toma, Y., Fernandez, F.G., Nishiwaki, A., Yamada, T., Bollero, G.A., 2009. The ecology and agronomy of Miscanthus sinensis, a species important to bioenergy crop development, in its native range in Japan: a review. GCB Bioenergy 1, 126e153.
REFERENCES Sun, Y., Cheng, J., 2002. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresour. Technol. 83, 1e11. Sustainable Aviation Fuel Road Map, 2011. Towards Establishing a Sustainable Aviation Fuels Industry in Australia and New Zealand. CSIRO Energy Transformed Flagship, PO Box 330, Newcastle, NSW 2300 Australia. Taherzadeh, M.J., Karimi, K., 2008. Pretreatment of lignocellulosic wastes to improve ethanol and biogas production: a review. Int. J. Mol. Sci. 9, 1621e1651. Taylor, G., 2008. Biofuels and the biorefinery concept. Energy Policy 36, 4406e4409. Tickell, J., 2000. From the Fryer to the Fuel Tank. The complete guide to using vegetable oil as an alternative fuel. Tallahasseee, USA. Third International Conference on Lignocellulosic Ethanol. 2013. http://www.biofuelstp.eu/events/3rd-icleapril-2013.pdf Tuohy, M.G., Laffey, C.D., Coughlan, M.P., 1994. Characterisation of the individual components of the xylanolytic enzyme system of Talaromyces emersonii. Bioresour. Technol. 50, 37e42. Turner, K.M., Ayyachamy, M., Brebion, J., Gupta, V.K., Tuohy, M.G., 2010. Fungal thermozymes: a new route for the synthesis of functional prebiotics. In: Gupta, V.K., Tuohy, M.G., Gaur, R.K. (Eds.), Fungal Biochemistry and Biotechnology. Lambert Academic Publishing, Germany. Tutt, M., Olt, J., 2011. Suitability of various plant species for bioethanol production. Agron. Res. Biosyst. Eng. 1, 261e267. Tyner, W.E., 2010. The integration of energy and agricultural markets. Agric. Econ. 41, 193e201. Ugwu, C.U., Aoyagi, H., Uchiyama, H., 2008. Photobioreactors for mass cultivation of algae. Bioresour. Technol. 99, 4021e4028. Upham, P., Riesch, H., Tomei, J., Thornley, P., 2011. The sustainability of forestry biomass supply for EU bioenergy: a post-normal approach to environmental risk and uncertainty. Environ. Sci. Policy 14, 510e518. US-DOE, 2008a. Biomass Program. Department of Energy, The Office of Energy Efficiency and Renewable Energy, Washington, DC. http://www1.eere.energy.gov/biomass/biofuels_initiative.html. US-DOE, 2008b. Theoretical Ethanol Yield Calculator. http://www1. eere.energy. gov/biomass/ethanol_yield_calculator.html. Van Dam, J., Junginger, M., Faaij, A.P.C., 2010. From the global efforts on certification of bioenergy towards an integrated approach based on sustainable land use planning. Renewable Sustainable Energy Rev. 14, 2445e2472. Van den Broek, R., van den Burg, T., van Wijk, A., Turkenburg, W., 2000. Electricity generation from eucalyptus and bagasse by sugar mills in Nicaragua: a comparison with fuel oil electricity generation on the basis of costs, macro-economic impacts and environmental emissions. Biomass Bioenergy 19, 311e335. Van Dyck, J.S., Pletschke, B.I., 2012. A review of lignocellulose bioconversion using enzymatic hydrolysis and synergistic cooperation between enzymes e factors affecting enzymes, conversion and synergy. Biotechnol. Adv. 30, 1458e1480. Van Stappen, F., Brose, I., Schenkel, Y., 2011. Direct and indirect land use change issues in European sustainability initiatives: state-ofthe-art, open issues and future developments. Biomass Bioenergy 35, 4824e4834. Ververis, C., Georghiou, K., Danielidis, D., Hatzinikolaou, D.G., Santas, P., Santas, R., Corleti, V., 2007. Cellulose, hemicelluloses, lignin and ash content of some organic materials and their suitability for use as paper pulp supplements. Bioresour. Technol. 98, 296e301. Viikari, L., Alapuranen, M., Puranen, T., Vehmaanperä, J., SiikaAho, M., 2007. Thermostable enzymes in lignocellulose hydrolysis. Adv. Biochem. Eng. Biotechnol. 108, 121e145. Wan, C., Li, Y., 2010. Microbial pretreatment of corn stover with Ceriporiopsis subvermispora for enzymatic hydrolysis and ethanol production. Bioresour. Technol. 101, 6398e6403. 47 Wayman, M., Parekh, S.R., 1990. Biotechnology of Biomass Conversion; Fuel and Chemicals from Renewable Resources. Open University Press, Milton Keynes, 181e232. Wheals, A.E., Basso, L.C., Alves, D.M.G., Amorim, H.V., 1999. Fuel ethanol after 25 years. Trends Biotechnol. 17, 482e487. Wi, S.G., Chung, B.Y., Lee, Y.G., Yang, D.J., Bae, H.-J., 2011. Enhanced enzymatic hydrolysis of rapeseed straw by popping pretreatment for bioethanol production. Bioresour. Technol. 102, 5788e5793. Williams, M.J., Douglas, J., 2011. Planting and managing giant miscanthus as a biomass energy crop. USDA plant materials program, Technical Note No. 4. Wiselogel, A., Tyson, S., Johnson, D., 1996. Biomass feedstock resources and composition. In: Wyman, C.E. (Ed.), Handbook on Bioethanol: Production and Utilization. Taylor & Francis, Washington, DC, pp. 105e118. World Development Report, 2008. Biofuels: The Promise and the Risks. The World Bank, 70e71. World Watch Institute (WWI), 2007. In: Hunt, S., Sterling, V.A. (Eds.), Biofuels for Transport: Global Potential and Implications for Sustainable Energy and Agriculture. Wu, M.M., Chang, K., Gregg, D.J., Boussaid, A., Beatson, R.P., Saddler, J.N., 1999. Optimization of steam explosion to enhance hemicellulose recovery and enzymatic hydrolysis of cellulose in softwoods. Appl. Biochem. Biotechnol. 77 (79), 47e54. www. ieabioenergy.com/DownLoad.aspx?DocId¼6494. Wyman, C.E., 2007. What is (and is not) vital to advancing cellulosic ethanol. Trends Biotechnol. 25, 153e157. Yamamura, M., Akashi, K., Yokota, A., Hattori, T., Suzuki, S., Shibata, D., Umezawa, T., 2012. Characterization of Jatropha curcas lignins. Plant Biotechnol. 29, 179e183. Yen, H.-W., Hu, I.-C., Chen, C.-Y., Ho, S.-H., Lee, D.-J., Chang J-, S., 2013. Microalgae-based biorefinery e from biofuels to natural products. Bioresour. Technol. 135, 166e174. Yuan, J.S., Tiller, K.H., Al-Ahmad, H., Stewart, N.R., Stewart, C.N., 2008. Plants to power: bioenergy to fuel the future. Trends Plant Sci. 13, 421e429. Zalesny Jr., R.S., Wiese, A.H., Bauer, E.O., Riemenschneider, D.E., 2009. Ex situ growth and biomass of Populus bioenergy crops irrigated and fertilized with landfill leachate. Biomass Bioenergy 33, 62e69. Zeng, X., Danquah, M., Zheng, C., Potumarthi, R., Xiao, D.C., Lu, Y., 2012. NaCS-PDMDAAC immobilized autotrophic cultivation of Chlorella sp. for wastewater nitrogen and phosphate removal. Chem. Eng. J. 187, 185e192. Zhang, Y.H.P., Ding, S.-Y., Mielenz, J.R., Cui, J.-B., Elander, R.T., Laser, M., Himmel, M.E., McMillan, J.R., Lynd, L.R., 2007. Fractionating recalcitrant lignocellulose at modest reaction conditions. Biotechnol. Bioeng. 97, 214e223. Zhang, Y.H.P., Lynd, L.R., 2004. Toward an aggregated understanding of enzymatic hydrolysis of cellulose: non-complexed cellulase systems. Biotechnol. Bioeng. 88, 797e824. Zhang, Y.H.P., Lynd, L.R., 2006. A functionally based model for hydrolysis of cellulose by fungal cellulose. Biotechnol. Bioeng 94, 888e898. Zhao, X., Zhang, L., Liu, D., 2010. Pretreatment of Siam weed stem by several chemical methods for increasing the enzymatic digestibility. Biotechnol. J. 5, 493e504. Zhao, X., Zhang, L., Liu, D., 2007. Comparative study on chemical pretreatment methods for improving enzymatic digestibility of crofton weed stem. Bioresour. Technol. 99, 3729e3736. Zhu, J.Y., Zhuang, X.S., 2012. Conceptual net energy output for biofuel production from lignocellulosic biomass through biorefining. Prog. Energy Combust. Sci. 38, 583e598. Zhu, L., O’Dwyer, J.P., Chang, V.S., Granda, C.B., Holtzapple, M.T., 2008. Structural features affecting biomass enzymatic digestibility. Bioresour. Technol. 99, 3817e3828.
C H A P T E R 3 Use of Agroindustrial Residues for Bioethanol Production Luiz J. Visioli, Fabiane M. Stringhini, Paulo R.S. Salbego, Daniel P. Chielle, Gabrielly V. Ribeiro, Juliana M. Gasparotto, Bruno C. Aita, Rodrigo Klaic, Jéssica M. Moscon, Marcio A. Mazutti* Department of Chemical Engineering, Federal University of Santa Maria, Santa Maria, Brazil *Corresponding author email: mazutti@ufsm.br O U T L I N E Introduction 49 Raw Material Sugar-Containing Residues Starch-Containing Residues Cellulose-Containing Residues 50 51 51 52 Sugar Production and Fermentation 52 Separate Hydrolysis and Fermentation Simultaneous Saccharification and Fermentation Concluding Remarks 55 References 55 INTRODUCTION toward the use of fuels derived from plant feedstock (Ferreira-Leitão et al., 2010). Agroindustrial and forestry residues, which are byproducts of key industrial and economical activities, stand out as potential raw materials for the production of renewable fuels, chemicals and energy (FerreiraLeitão et al., 2010). Biofuels can also be derived from fishery products or municipal wastes, also including by-products and wastes originated from agroindustry, food industry and food services (Nigam and Singh, 2011). The key advantage of the utilization of renewable sources for the production of biofuels is the utilization of natural bioresources (that are geographically more evenly distributed than fossil fuels) and the produced bioenergy provides independence and security of energy supply (Nigam and Singh, 2011). The use of agricultural residue and waste substrates as raw materials is advantageous as their availability is not hindered by a requirement for arable land for the production of food The last years have verified a pronounced demand for fossil fuels worldwide due to increase in industrialization and motorization (Agrawal et al., 2007). Nowadays, fossil fuels represent around 80% of all primary energy consumed in the world, where 58% is employed in the transport sector (Escobar et al., 2009). The estimates show that the global energy demand is projected to grow by more than 50% by 2025, with much of this increase in demand emerging from several rapidly developing nations. Clearly, increasing demand for finite petroleum resources cannot be a satisfactory policy for the long term (Ragauskas et al., 2006). Biofuels are a renewable energy source produced from natural (plant) materials, which can be used as a substitute for petroleum fuels (Demirbas, 2011). The global demand for liquid biofuels more than tripled in last decade, indisputably showing the increasing trend Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00003-6 52 55 49 Copyright Ó 2014 Elsevier B.V. All rights reserved.
50 3. USE OF AGROINDUSTRIAL RESIDUES FOR BIOETHANOL PRODUCTION and feed. Reusing agricultural waste products is one goal of environmental sustainability and has become an option to add value to producers (Manique et al., 2012). In addition, waste utilization prevents its accumulation, which is of great environmental concern due to its potential for contamination of rivers and underground water (Ferreira-Leitão et al., 2010). The most well-known first-generation biofuel is ethanol (Nigam and Singh, 2011), which is currently being produced from sugarcane or corn and will often be referred to as bioethanol (Demirbas, 2011). Ethanol has long been considered as a suitable alternative to fossil fuels either as a sole fuel in cars with dedicated engines or as an additive in fuel blends with no engine modification requirement when mixed up to 30%. Today, bioethanol is the most dominant biofuel and its global production showed an upward trend over the last 25 years. Worldwide production capacity in 2006 was about 49  109 liters per year, and total output in 2015 is forecast to reach over 115  109 liters (Talebnia et al., 2010). Feedstock containing significant amounts of sugar, or materials that can be converted into sugars, such as starch or cellulose, can also be used in the production of ethanol (Nigam and Singh, 2011). The production of ethanol from cellulosic feedstock has a growing interest worldwide. Cellulosic biomass is an abundant renewable resource on earth and includes various agricultural residues. Some of these agricultural residues such as straw, corn husk, and sugarcane residue represent an abundant, inexpensive, and readily available source of renewable lignocellulosic biomass. At the present time, this readily available biomass is considered as a waste and is disposed of through agricultural burning after harvest (Dawson and Boopathy, 2007). Agricultural residues are produced in large quantities throughout the world. Approximately, 1 kg of residue is produced for each kilogram of grain harvested. These residues are renewable resources that could be used to produce ethanol and many other value-added products (Dawson and Boopathy, 2007). Among these residues, ethanol can be produced from biomass feedstocks such as sucrose-containing feedstocks (e.g. sugar beet, sweet sorghum, and sugarcane), starchy materials (e.g. wheat, corn, barley, cassava, and rice), and lignocellulosic biomass (e.g. wood, straw, grasses, and various crop residues). These biomass feedstocks can reduce about 50% of the price of the ethanol produced, depending on the type of the biomass used (Hong and Yoon, 2011). Lignocellulosic waste materials obtained from energy crops, wood and agricultural residues represent the most abundant global source of renewable biomass. Among the agricultural residues, wheat straw is the largest biomass feedstock in Europe and the second largest in the world after rice straw. About 21% of the world’s food depends on the wheat crop and its global production needs to be increased to satisfy the growing demand of human consumption; therefore, wheat straw would serve as a great potential feedstock for production of ethanol in the twenty-first century (Talebnia et al., 2010). The use of lignocellulosic energy crops, and particularly low-cost biomass residues, offers excellent perspectives for large-scale application of ethanol in transportation fuels. These materials will increase the ethanol production capacity and reduce production cost to a competitive level. Bioethanol from these materials provides a highly costeffective option for CO2 emission reduction in the transportation sector (Patle and Lal, 2007). The utilization of lignocellulosic biomass has been closely associated with a new technological concept, so-called biorefinery, which emerges as key to the significant expansion of the desired production of ethanol. Fermentative processes stand out, where microbial metabolism is used for the transformation of simple raw materials in products with high aggregate value. Experts believe that the biorefineries are likely to be a key industry of the twenty-first century, even responsible for a new industrial revolution, because of the importance of the technologies they employ and their effects on the actual industrial model (Santos et al., 2010). Regarding crop residues that have proper application in energy supply, the energetic generation cost for useful energy is a matter for consideration. Studies done so far suggest that, when transport distances are similar, the most efficient energetic use of lignocellulosic materials such as agricultural residues is the application for the generation of electricity. Applied in this way, crop residues are most efficient in replacing fossil fuels, much more so than when crop residues are converted to ethanol for use in cars. However, when road transport distances to power-generating plants are very large, it may be that energetic uses that require a much lower input of transport fuel become energetically more attractive (Reijnders, 2008). Based on these aspects, the main objective of this work is to present an overview about bioethanol production from agroindustrial residues, which were based on low-priced feedstocks such as crop residues, municipal/industrial solid waste, and food residues. For this purpose, papers and overviews since 2006 were reported. RAW MATERIAL Bioethanol is usually produced out of organic-based matter with high contents of sugar fermentation by enzymes produced from yeast. The yeast converts
51 RAW MATERIAL six-carbon sugars (mainly glucose) to ethanol, because starch is much easier than cellulose to convert to glucose (Nigam and Singh, 2011). Bioethanol is produced similarly to other alcohols such as spirits using natural products like wheat, maize and sugar beet. Hence, the suitable raw materials required for bioethanol production could be any of those materials that contain considerable amounts of carbohydrates to provide fermentable sugars for bioconversion into bioethanol. Then an optimized microbial fermentation process can be used for the bioconversion of sugars released from carbohydrates into ethanol (Nigam and Singh, 2011). Agricultural waste materials are inexpensively found outside the human food chain in large amounts and can be obtained throughout the year. These agricultural biomasses are the potential feedstocks for bioethanol production, including the cellulosic biomass, as well as starchy waste agricultural materials, and they provide low-cost and uniquely sustainable resources, improvement on energy security, development of the economy, as well as cleaning the environment and atmosphere by the disposing of problematic solid wastes and getting wealth out of wastes. Synthetically, 7% ethanol can be made from petroleum resources and 93% ethanol through fermentation process using microorganisms to convert biomass materials into ethanol (Kahn et al., 2011). Sugar-Containing Residues Raw materials (containing saccharides) include sweet sorghum, sugar beet, banana, mango, watermelon, sugarcane and other fruits; these are examples of sugar feedstock. The wastes of these sugar-containing sources can be fermented by using different microorganisms of interest. However, the use of these materials for bioethanol production is highly expensive and humans use them as food (Kahn et al., 2011). The first research about bioethanol showed in Table 3.1 reports the production of biofuel using residues of fruits and vegetables. These are an important source of sugar for ethylic fermentation because the processing of fruits have a great potential to generate residues that can be used. The authors report the use of enzymes in the process; this is necessary because these residues have a lot of fiber that can be hydrolyzed (Patle and Lal, 2007). The use of wastes is the main aspect of two other studies (Gouvea et al., 2009; Ge et al., 2011). Gouvea et al. (2009) reported the use of a residue very abundant in Brazil, that is, the coffee husk. As well as in the case of fruit and vegetable residues the floating seaweed wastes have other kinds of carbohydrates that were hydrolyzed by enzymes (Wu et al., 2011). Ge et al. (2011) reported the use of a principal sugar source for ethylic fermentation, the stalk juice of sugarcane. This juice has great quantity of glucose in its composition and has a low cost. More studies about sugarcane juice were omitted since the main subject of this chapter is agroindustrial residues. Starch-Containing Residues Starch materials are also potential resources for the bioethanol production. Starch molecules are polysaccharides made up of long chains of glucose units covalently linked. Before the fermentation process, the starchy materials are broken into simple glucose molecules after which the simple sugar units are easily fermented by the microbes. Examples of starchy materials used for bioethanol production include cereal grains, potato, sweet potato, beans, cassava, maize, wheat and cereal grains. As these materials are also too expensive and included in the human food TABLE 3.1 Sugar Raw Material Raw Material Sugar Production Operation Ethanol Yield References Fruit and vegetable residues (apple, carrot, mango, orange, pineapple, sapota and tomato) Acid and enzymatic hydrolysis was used previously to the fermentation to release the sugar from residues Separated hydrolysis and fermentation (SHF) Apple: yield of 79.8% Carrot: yield of 70% Mango: yield of 82.8% Orange: yield of 91.3% Pineapple: yield of 75.4% Sapota: yield of 97.7% Tomato: yield of 72.4% Patle and Lal (2007) Coffee husks Not applicable Not applicable Ethanol production was 84.9  2.9 g/kg dry basis Gouvea et al. (2009) Floating seaweed wastes Acid pretreatment (H2SO4), enzymatic hydrolysis fermentation Separated hydrolysis and fermentation (SHF) The maximum yield of glucose reached 277.5 mg/g (FSW) that was 80.8% fermented to ethanol Ge et al. (2011) Sugarcane stalk juice Not applicable Not applicable The theoretical yield was 88% Wu et al. (2011)
52 TABLE 3.2 3. USE OF AGROINDUSTRIAL RESIDUES FOR BIOETHANOL PRODUCTION Starch Raw Material Raw Material Sugar Production Operation Ethanol Yield References Food residues The hydrolysis was made by enzymes Simultaneous saccharification and fermentation (SSF) The ethanol concentrations were 19, 35, and 39 g/dm3 when the concentrations of food residue were 50, 100, and 120 g/dm3 Hong and Yoon (2011) Raw corn starch Enzymatic saccharification Simultaneous saccharification and fermentation (SSF) The largest production of ethanol was 85 g/kg (raw material) Moukamnerd et al. (2010) Potato starch residue Hydrolysis acid (HCl and H2SO4) Separated hydrolysis and fermentation (SHF) The maximum yield of ethanol (5.52 g/l) Hashem and Darwish (2010) chain, wastes are collected from places where they are crushed into flour or from industries where they are used for various products (Kahn et al., 2011). Table 3.2 reports the use of starchy residues in ethanol production. As starch is an important component of the food chain, it is expected that the wastes of the processing be more used for ethylic fermentation (Moukamnerd et al., 2010). Another important point that can be observed is the predominant use of simultaneous saccharification and fermentation (SSF) in the use of starch for energy generation and in both cases hydrolysis was achieved by enzymes (Hong and Yoon, 2011; Moukamnerd et al., 2010). Hashem and Darwish (2010) reported the use of potato starch residue stream produced during chips manufacture and the authors have used separate hydrolysis and fermentation (SHF) and a very low acid hydrolysis of the starch to reduce the cost associated with this necessary treatment. Cellulose-Containing Residues Another way to produce bioethanol is using cellulosic materials. Examples of cellulosic materials are bagasse, straw, paper, cardboard, wood and materials of plant cellulosic fibers such hemp, giant reed, eucalyptus tree and Miscanthus. Cellulosic resources are immensely widespread and found abundantly everywhere. These cellulosic materials have the potential to be used for the production of bioethanol since they are not commonly used in the human food chain and exist in large amounts. Moreover, these materials are inexpensive as compared to the sugar and starchy feedstocks and preferably used for bioethanol production. Cellulosic materials are called lignocelluloses because they are composed of lignin, cellulose and hemicelluloses (Kahn et al., 2011). The cellulosic residue more reported in Table 3.3 for ethanol production is sugarcane bagasse (Dawson and Boopathy, 2007; Santos et al., 2012; Wu et al., 2011; Buaban et al., 2010). This can be explained because of the great use of its juice for sugar ethylic fermentation and the residue is generated just in the alcohol manufacture (Santos et al., 2012). Furthermore, the quantity of this waste available is very great and its direct combustion is not the best economical way to use this resource (Wu et al., 2011). The hydrolysis process more employed for saccharification of the bagasse is enzymatic. This is because of the inefficiency of acid hydrolysis in a very complex matrix. Moreover many studies are trying SSF, which can reduce one step of the process. The agricultural wastes (apart from sugarcane) are studied much for the ethanol production. This is because they are present around the world (comes from diverse agricultural crops). Thus these residues can mean the energetic independence of many countries. Rice plant (Kitamoto et al., 2011) and Lycoris radiata Herbert (Liu et al., 2012) are examples of how much singular are the wastes that the researches are using to produce biofuels. Similarly to sugarcane, Talebnia et al. (2010) have tested SSF and SHF of wheat straw, and obtained good results in the two cases. In Table 3.3 it is possible to see that pretreatment (physical or chemical) is usually necessary, and this is a difficulty in this production. SUGAR PRODUCTION AND FERMENTATION Separate Hydrolysis and Fermentation SHF uses different stages for enzyme production, cellulose hydrolysis and fermentation of glucose. In this process, the hydrolysis of cellulose occurs before the fermentation of glucose, which is carried out in different reactors. In this case, the temperature of hydrolysis and fermentation can be optimized individually (Hamelinck et al., 2005). For this reason, this process is preferably used for ethanol production from lignocellulosic material, as can be seen in Table 3.3, since the optimal temperature for acid or enzymatic hydrolysis is different from that of fermentation.
TABLE 3.3 Cellulosic Raw Material Sugar Production Operation Ethanol Yield References Postharvest sugarcane residue Alkaline pretreatment (H2O2) to remove lignin and acid hydrolysis (H2SO4) Separated hydrolysis and fermentation (SHF) Ethanol production of 335.67 mg/l after 12 days of fermentation Dawson and Boopathy (2007) Agricultural residues and hay (wheat, barley, and triticale straw and barley, triticale, pearl millet, and sweet sorghum hay) Chemical pretreatment (NaOH or H2SO4) and enzymatic hydrolysis Separated hydrolysis and fermentation (SHF) Production between 52.00% and 65.82% of the theoretical ethanol yield Chen et al. (2007) Cotton stalk, triticale hay, barley, triticale, and wheat straw Enzymatic hydrolysis of lignocelluloses Separated hydrolysis and fermentation (SHF) Ethanol yields range between 0.21 and 0.28 (g/g reducing sugars) Chen et al. (2007) Wheat straw The several kinds of pretreatment and hydrolysis are found in the reference Simultaneous saccharification and fermentation (SSF) The largest ethanol yield reported by the authors was 81% Talebnia et al. (2010) Wheat straw The several kinds of pretreatment and hydrolysis are found in the reference Separated hydrolysis and fermentation (SHF) Ethanol yield ranging from 65% to 99% of theoretical value Talebnia et al. (2010) Production canola residue Acid (H2SO4) and alkali (NaOH) pretreatment, followed by enzymatic hydrolysis Separated hydrolysis and fermentation (SHF) The best yield obtained in the fermentation was 45% for ethanol production or around 95 l per dry ton of raw material George et al. (2010) Corncob residues Chemical pretreatment (acid or alkali) and enzymatic hydrolysis Simultaneous saccharification and fermentation (SSF) The yield range was between 25.2% and 27.1% of theoretical value to ethanol production Liu et al. (2010) Floating seaweed wastes (FSWs) Acid pretreatment (H2SO4), enzymatic hydrolysis fermentation Separated hydrolysis and fermentation (SHF) The maximum yield of glucose reached 277.5 mg/g FSW that was 80.8% fermented to ethanol Ge et al. (2011) Sugarcane bagasse Pretreated by steam explosion at 200  C and delignification with NaOH, and enzymatic hydrolysis Simultaneous saccharification and fermentation (SSF) Ethanol concentration was higher than 25 g/l Santos et al. (2012) SUGAR PRODUCTION AND FERMENTATION Raw Material (Continued) 53
54 TABLE 3.3 Cellulosic Raw Materialdcont’d Sugar Production Operation Ethanol Yield References Rice plants Physical and chemical pretreatment and enzymatic saccharification Simultaneous saccharification and fermentation (SSF) Ethanol in fresh matter (169 g/kg dry matter) was produced Kitamoto et al. (2011) Lycoris radiata Herbert (Amarylllidaceae) residues The residue was hydrolyzed by enzymes before the fermentation Separated hydrolysis and fermentation (SHF) Not determined Liu et al. (2012) Sugarcane bagasse Enzymatic hydrolysis Separated hydrolysis and fermentation (SHF) The theoretical yield was 88% Wu et al. (2011) Soybeans hulls Enzymatic hydrolysis without pretreatment Simultaneous saccharification and fermentation (SSF) Ethanol concentrations of 25e30 g/l were obtained Mielenz et al. (2009) Sugarcane bagasse Mechanical pretreatment by ball milling, with enzymatic hydrolysis Separated hydrolysis and fermentation (SHF) Yield of 56.9% of theoretical production Buaban et al. (2010) Sugarcane bagasse Mechanical pretreatment by ball milling, with enzymatic hydrolysis Simultaneous saccharification and fermentation (SSF) Yield of 52.9% of the theoretical production Buaban et al. (2010) Corn stover Enzymatic hydrolysis Simultaneous saccharification and fermentation (SSF) The yield was between 69% and 98% of the theoretical ethanol production Yoo et al. (2012) SHF, separated hydrolysis and fermentation; SSF, simultaneous saccharification and fermentation. 3. USE OF AGROINDUSTRIAL RESIDUES FOR BIOETHANOL PRODUCTION Raw Material
55 REFERENCES The main characteristic of SHF is that the technique allows a high number of steps. Thereby the hydrolysis can be carried out with better efficiency than in SSF. Another important aspect to the use of this one is the unnecessity of development or trials of new microorganism or enzymes that are able to produce ethanol or sugar at different medium of that when usually fermentation happens (George et al., 2010). Tables 3.2 and 3.3 show that the SHF is the predominant technique used for saccharification of any source of carbohydrates. Simultaneous Saccharification and Fermentation Research in ethanol has been targeted for the development of second-generation technology, including the strategy of SSF process, which combines in a single unit the cellulose enzymatic hydrolysis and the ethanol fermentation (Santos et al., 2010). In the SSF process, glucose released by cellulase action is directly converted to ethanol by the fermenting microorganisms, which alleviates problems caused by the end product. The consumption of glucose and the presence of ethanol in the culture medium would reduce the risk of undesired contamination by glucose-dependent organisms. Recently, consolidated bioprocessing, which combines enzyme production, saccharification and fermentation in a single step, has gained recognition as a potential bioethanol production system, because the costs of capital investment, substance and other raw materials, and utilities associated with enzyme production can be avoided using microorganisms with the capability for efficient cellulose hydrolysis and ethanol production (Hasunuma and Kondo, 2012). Recently, there are many reports that SSF is superior to the traditional saccharification and subsequent fermentation in the ethanol production because the SSF process can improve ethanol yields by removing end-product inhibition of saccharification process and decrease the enzyme loading. Moreover, SSF requires a single fermenter for the entire process and eliminates the need for separating reactors for saccharification and fermentation leading to reduce the investment cost (Boonsawang et al., 2012). Difference between SHF and SSF is in an incipient step of their development. It is possible to note that a significant number of studies reported in Tables 3.2 and 3.3 are making a comparison between the two techniques, which shows that the SSF researches are trying to develop an efficient process to substitute the SHF method. On the other hand the starchy raw materials have a great use in SSF fermentation; this can be explained by the simplicity of this substrate compared to cellulosic (the efficiency of enzymatic hydrolysis is better) and the conditions of operation can be milder, facilitating the adaptation of a fermentation microorganism. CONCLUDING REMARKS As can be seen from the tables above, there is a growing interest in ethanol production from agroindustrial residues of a variety of sources including grains, straws, stalks and husks such as cotton, barley, triticale, wheat, coffee, rice, canola, sugarcane and other fruits and vegetables. In terms of volume, lignocellulosic material is the predominant raw material for secondgeneration ethanol. However, the production costs associated with the use of lignocellulosic ethanol is high, making it necessary to develop an efficient process for hydrolysis and fermentation, where the use of simultaneous saccharification and hydrolysis is seen a promising technology, but there is also the necessity to genetically modify a microorganism to grow at high temperatures or obtain an enzyme to carry out the hydrolysis at normal fermentation temperature. Low-cost biomass residues offer excellent perspective for largescale application of ethanol. Acknowledgments The authors thank CAPES for the scholarships and SCIT-RS and CNPq for the financial support of this work. References Agrawal, R., Singh, N.R., Ribeiro, F.H., Delgass, W.N., 2007. Sustainable fuel for the transportation sector. Proc. Natl. Acad. Sci. 104, 4828e4833. Boonsawang, P., Subkaree, Y., Srinorakutara, T., 2012. Ethanol production from palm pressed fiber by prehydrolysis prior to simultaneous saccharification and fermentation (SSF). Biomass Bioenergy 40, 127e132. Buaban, B., Inoue, H., Yano, S., Tanapongpipat, S., Ruanglek, V., Champreda, V., Pichyangkura, R., Rengpipat, S., Eurwilaichitr, L., 2010. Bioethanol production from ball milled bagasse using an on-site produced fungal enzyme cocktail and xylose-fermenting Pichia stipitis. J. Biosci. Bioeng. 110, 18e25. Chen, Y., Sharma-Shivappa, R.R., Chen, C., 2007. Ensiling agricultural residues for bioethanol production. Appl. Biochem. Biotechnol. 143, 80e92. Dawson, L., Boopathy, R., 2007. Use of post-harvest sugarcane residue for ethanol production. Bioresour. Technol. 98, 1695e1699. Demirbas, A., 2011. Competitive liquid biofuels from biomass. Appl. Energy 88, 17e28. Escobar, J.C., Lora, E.S., Venturini, O.J., Yanez, E.r E., Castillo, E.F., Almazan, O., 2009. Biofuels: environment, technology and food security. Renewable Sustainable Energy Rev. 13, 1275e1287. Ferreira-Leitão, V., Gottschalk, L.M.F., Ferrara, M.A., Nepomuceno, A.L., Molinari, H.B.C., Bon, E.P.S., 2010. Biomass residues in Brazil: availability and potential uses. Waste Biomass Valorization 1, 65e76.
56 3. USE OF AGROINDUSTRIAL RESIDUES FOR BIOETHANOL PRODUCTION Ge, L., Wang, P., Mou, H., 2011. Study on saccharification techniques of seaweed wastes for the transformation of ethanol. Renewable Energy 36 (84), 89. George, N., Yang, Y., Wang, Z., Sharma-Shivappa, R., Tungate, K., 2010. Suitability of canola residue for cellulosic ethanol production. Energy Fuels 24, 4454e4458. Gouvea, B.M., Torres, C., Franca, A.S., Oliveira, L.S., Oliveira, E.S., 2009. Feasibility of ethanol production from coffee husks. Biotechnol. Lett. 31, 1315e1319. Hamelinck, C.N., van Hooijdonk, G., Faaij, A.P., 2005. Ethanol from lignocellulosic biomass: techno-economic performance in short, middle, and long term. Biomass Bioenergy 28, 384e410. Hashem, M., Darwish, S.M.I., 2010. Production of bioethanol and associated by-products from potato starch residue stream by Saccharomyces cerevisiae. Biomass Bioenergy 34, 953e959. Hasunuma, T., Kondo, A., 2012. Consolidated bioprocessing and simultaneous saccharification and fermentation of lignocellulose to ethanol with thermotolerant yeast strains. Process Biochem. 47, 1287e1294. Hong, Y.S., Yoon, H.H., 2011. Ethanol production from food residues. Biomass Bioenergy 35, 3271e3275. Kahn, R.A., Kahn, A.N., Ahmed, M., Kahn, M.R., Shah, M.S., Azam, N., Sadullah, F., Dian, F., Ullah, S., Kahn, N., 2011. Bioethanol sources in Pakistan: a renewable energy resource. Afr. J. Biotechnol. 10, 19850e19854. Kitamoto, H.K., Horita, M., Cai, Y., Shinozaki, Y., Sakaki, K., 2011. Silage produces biofuel for local consumption. Biotechnol. Biofuels 4, 46. Liu, K., Lin, X., Yue, J., Li, X., Fang, X., Zhu, M., Lin, J., Qu, Y., Xiao, L., 2010. High concentration ethanol production from corncob residues by fed-batch strategy. Bioresour. Technol. 101, 4952e4958. Liu, S., Ding, Z.G., Zhang, L., Gu, Z., Wang, X., Sun, X., Shi, G., 2012. Ethanol production from Lycoris radiata Herbert (Amarylllidaceae) residues as a new resource. Biomass Bioenergy 37, 237e242. Manique, M.C., Faccini, C.S., Onorevoli, B., Benvenutti, E.V., 2012. Rice husk ash as an adsorbent for purifying biodiesel from waste frying oil. Fuel 92, 56e61. Mielenz, J.R., Bardsley, J.S., Wyman, C.E., 2009. Fermentation of soybean hulls to ethanol while preserving protein value. Bioresour. Technol. 100, 3532e3539. Moukamnerd, C., Kino-oka, M., Sugiyama, M., Kaneko, Y., Boonchird, C., Harashima, S., Noda, H., Ninomiya, K., Shioya, S., Katakura, Y., 2010. Ethanol production from biomass by repetitive solid-state fed-batch fermentation with continuous recovery of ethanol. Appl. Microbiol. Biotechnol. 88, 87e94. Nigam, P.S., Singh, A., 2011. Production of liquid biofuels from renewable resources. Prog. Energy Combust. Sci. 37, 52e68. Patle, S., Lal, B., 2007. Ethanol production from hydrolysed agricultural wastes using mixed culture of Zymomonas mobilis and Candida tropicalis. Biotechnol. Lett. 29, 1839e1843. Ragauskas, A.J., Williams, C.K., Davison, B.H., Britovsek, G., Cairney, J., Eckert, C.A., Frederick Jr, W.J., Hallett, J.P., Leak, D.J., Liotta, C.L., Mielenz, J.R., Murphy, R., Templer, R., Tschaplinski, T., 2006. The path forward for biofuels and biomaterials. Science 311, 484. Reijnders, L., 2008. Ethanol production from crop residues and soil organic carbon. Resour., Conserv. Recycl. 52, 653e658. Santos, D.S., Camelo, A.C., Rodrigues, K.C.P., Carlos, L.C., Pereira, N., 2010. Ethanol production from sugarcane bagasse by Zymomonas mobilis using simultaneous saccharification and fermentation (SSF) process. Appl. Biochem. Biotechnol. 161, 93e105. Santos, J.R.A., Lucena, M.S., Gusmão, N.B., Gouveia, E.R., 2012. Optimization of ethanol production by Saccharomyces cerevisiae UFPEDA 1238 in simultaneous saccharification and fermentation of delignified sugarcane bagasse. Ind. Crops Prod. 36, 584e588. Talebnia, F., Karakashev, D., Angelidaki, I., 2010. Production of bioethanol from wheat straw: an overview on pretreatment, hydrolysis and fermentation. Bioresour. Technol. 101, 4744e4753. Wu, L., Li, Y., Arakane, M., Ike, M., Wada, M., Terajima, Y., Ishikawa, S., Tokuyasu, K., 2011. Efficient conversion of sugarcane stalks into ethanol employing low temperature alkali pretreatment method. Bioresour. Technol. 102, 11183e11188. Yoo, C.G., Kuo, M., Kim, T.H., 2012. Ethanol and furfural production from corn stover using a hybrid fractionation process with zinc chloride and simultaneous saccharification and fermentation (SSF). Process Biochem. 47, 319e326.
C H A P T E R 4 Recent Advancements in Pretreatment Technologies of Biomass to Produce Bioenergy Irmene Ortı´z*, Rodolfo Quintero Departamento de Procesos y Tecnologı́a, Universidad Autónoma Metropolitana - Cuajimalpa, México D.F., México *Corresponding author email: irmene@correo.cua.uam.mx O U T L I N E Lignocelullosic Biomass 57 Pretreatment of Lignocelullosic Biomass for Biofuels Production Trends in Pretreatments Other Pretreatments 62 62 58 Pretreatment Modeling 65 58 58 59 60 Environmental and Economical Aspects 65 Concluding Remarks 66 References 66 Types of Pretreatments Biological Pretreatments Physical Pretreatments Chemical Pretreatments Physicochemical Pretreatments 61 LIGNOCELULLOSIC BIOMASS resulting in the aggregation of chains into elementary crystalline fibrils of 36 cellulose chains, while hemicelluloses are complex branched heterogeneous polysaccharides composed of monomeric residues: D-glucose, D-galactose, D-mannose, D-xylose, L-arabinose, D-glucuronic acid and 4-O-methyl-D-glucuronic acid; and lignin is a complex amorphous network formed by polymerization of phenyl propane units and constitutes the most abundant nonpolysaccharide fraction in lignocellulose (Jørgensen et al., 2007; Lewis et al., 2005). Biofuels produced from native lignocellulose are known as second-generation biofuels. In this process the cellulose is converted into glucose, which is easily fermented to ethanol, while the hemicellulosic fraction is converted into monomeric sugars (mainly pentoses), a fermentation that is considerably harder to accomplish (Dias et al., 2011). The physicochemical and structural compositions of native lignocellulose are, however, recalcitrant to direct enzymatic hydrolysis of cellulose (Mosier et al., 2005). Therefore, a pretreatment step is Lignocellulosic biomass is composed primarily of cellulose, hemicelluloses (mainly xylan), lignin and smaller amounts of other compounds. Typically, the composition of lignocellulosic biomass by weight is 40e50% cellulose, 20e40% hemicellulose, 10e30% lignin and other components such as minerals, oils, soluble sugars, pectins, proteins, and ashes (Jørgensen et al., 2007; Lewis et al., 2005; Wyman et al., 2005). Cellulose, hemicelluloses and lignin are present in varying amounts in the different parts of the plant and they are intimately associated to form the structural framework of the plant cell wall; also, the content of the different sugars of the hemicelluloses varies significantly between different plants depending on plant species, age and growth conditions (Jørgensen et al., 2007). Cellulose is the most abundant constituent of the plant cell wall; its linear structure enables the formation of both intra- and intermolecular hydrogen bonds Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00004-8 57 Copyright Ó 2014 Elsevier B.V. All rights reserved.
58 4. RECENT ADVANCEMENTS IN PRETREATMENT TECHNOLOGIES OF BIOMASS TO PRODUCE BIOENERGY invariably required to render the cellulose amenable to enzymatic hydrolysis (Zheng et al., 2009). The total estimated availability of usable biomass in the world is about 2 billion dry tons per year (Lewis et al., 2005). Therefore, the enormous potential of secondgeneration fuels and the increasing interest toward developing effective, low-cost and environmentally friendly pretreatments for breaking down the close association of the structures of the biomass. by-products, e.g. lignin; (9) to be cost-effective by operating in reactors of moderate size, minimizing the heat and power requirements, chemicals and capital equipment; and (10) be scalable to industrial size (Alvira et al., 2010; Brodeur et al., 2011; Jørgensen et al., 2007; Yang and Wyman, 2008). This chapter reviews the advances in the most studied pretreatments and those recently proposed in scheme of the biochemical biorefinery, including kinetics, mechanistic and economical models proposed for describing some of these pretreatment processes. PRETREATMENT OF LIGNOCELULLOSIC BIOMASS FOR BIOFUELS PRODUCTION TYPES OF PRETREATMENTS One of the most promising emerging biorefinery platforms is the biochemical path that focuses on fermentation of sugars extracted from lignocellulosic feedstocks (Carvalheiro et al., 2008). This technology involves three basic steps: (1) conversion of biomass to sugar or other fermentation-feedstock, (2) bioconversion of these biomass intermediates using biocatalysts, and (3) process products to yield added value chemicals, fuelgrade ethanol and other fuels, heat and/or electricity (Carvalheiro et al., 2008). The first step involves a pretreatment process the goal of which is to alter or remove structural and compositional impediments to hydrolysis in order to improve the rate of enzyme hydrolysis and increase yields of fermentable sugars from cellulose or hemicellulose (Mosier et al., 2005). The effectiveness of enzymatic hydrolysis of pretreated lignocellulosic biomass can be significantly enhanced if lignin and its derivatives are removed or effectively modified before adding enzymes because lignin and its derivatives interfere with the path for cellulases action and they are also toxic to microorganisms, slowing down enzymatic hydrolysis (Qing et al., 2010; Yang and Wyman, 2008). The ideal pretreatment process produces a disrupted, hydrated substrate that is easily hydrolyzed but avoids the formation of sugar degradation products and fermentation inhibitors (Agbor et al., 2011). Furthermore, there is an overall consensus on the several technical, operational and economical characteristics that a successful pretreatment should accomplish, including (1) maximize the production of highly digestible solids that enhances sugar yields during enzyme hydrolysis; (2) avoid the degradation of sugars including those derived from hemicellulose; (3) not require the addition of toxic compounds or minimize their use; (4) fermentation compatibility, minimizing the formation of inhibitors for the enzymes or microorganisms in the subsequent steps; (5) effectiveness at low moisture content; (6) broad applicability for multiple crops, sites ages and harvesting times; (7) not required size reduction of biomass; (8) maximize the production of other valuable The pretreatment methods cause physical and/or chemical changes in the lignocellulosic biomass; thus, pretreatment technologies are usually classified into physical, chemical, physicochemical, and biological. For the purposes of classification, steam and water are excluded from being considered chemical agents for pretreatment since extraneous chemicals are not added to the biomass (Mosier et al., 2005). This chapter focuses on chemical, physical and physicochemical pretreatments but a brief description of biological treatments are include in order to contrast and compare them. Biological Pretreatments Biological pretreatments employ microorganisms mainly brown, white and soft-rot fungi, which degrade lignin, hemicellulose and cellulose in small proportion (Alvira et al., 2010). Recently, this approach has received renewed attention as biological pretreatments have several advantages over conventional physical/chemical pretreatment methods, such as they are considered as environmentally friendly, low capital cost, low energy, no chemicals requirement, and mild environmental conditions (Saritha and Lata, 2011). However, the main drawbacks to develop biological methods are the low hydrolysis rate obtained in most biological materials and the relatively long time of the pretreatment compared to physical/chemical methods. Consequently, more space and longer processes are required, which increase the operating costs (Alvira et al., 2010; Saritha and Lata, 2011). The white-rot fungi are able to decompose all wood fractions, including lignin because they produce various enzymes involved in lignin degradation such as lignin peroxidase, laccase, manganese peroxidase, versatile peroxidase, and H2O2-forming enzymes such as glyoxal oxidase and aryl alcohol oxidase. White-rot fungi also produce cellulases, xylanases and other hemicellulases that are required in the hydrolysis. Almost all
TYPES OF PRETREATMENTS white-rot fungi produce manganese peroxidase and laccase, but only some of them produce lignin peroxidase (Isroi et al., 2011). Several white-rot fungi such as Phanerochaete chrysosporium, Ceriporia lacerata, Cyathus stercolerus, Ceriporiopsis subvermispora, Pycnoporus cinnarbarinus, Pleurotus ostreaus, Dichomitus squalens, Coriolus versicolor, Trichoderma reesei, Aspergillus terreus, Aspergillus awamori, Bjerkandera adusta, Phlebia tremellosus, Fusarium proliferatum, and Pleurotus florida have been examined on different lignocellulosic biomass (Alvira et al., 2010; Cui et al., 2012; Isroi et al., 2011; Kuhar et al., 2008; Pinto et al., 2012; Saritha and Lata, 2011; Wan and Li, 2011). Recently, some bacterial laccases have also been characterized from Azospirillum lipoferum and Bacillus subtilis (Saritha and Lata, 2011). However, they face three major challenges associated with lignin structure: (1) the lignin polymer is large; therefore, ligninolytic systems must be extracellular, (2) lignin structure comprises interunit carbonecarbon and ether bonds; therefore, the degradation mechanism must be oxidative rather than hydrolytic, and (3) lignin polymer is stereoirregular, therefore the ligninolytic agents must be much less specific than degradative enzymes (Isroi et al., 2011). As mentioned before, one of the most main drawbacks of biological pretreatments is the time of the pretreatment; reported time of treatment is between 7 and 60 days (Giles et al., 2011; Mahalaxmi et al., 2010; Wan and Li, 2011). After 18 days of pretreatment, C. subvermispora effectively delignified corn stover, switchgrass, and hardwood with glucose yields during enzymatic hydrolysis that reached 56.50%, 37.15%, and 24.21%, respectively (Wan and Li, 2011). Also, glucose yield after 21 days of pretreatment with Poria subvermispora and Irpex lacteus reached 69% and 66% of cellulose available in the wheat straw, respectively, with an ethanol yield of 62% in both cases (Salvachua et al., 2011). Biological pretreatment has also been used before pyrolysis of biomass to produce fuel. The biological pretreatment of corn stover can optimize the thermal decomposition, decrease the reaction temperature and reduce the gas contamination (SOx), making the biomass pyrolysis more efficient and environmentally friendly. Biological pretreatment can decrease the activation energy and reacting temperature of the hemicellulose and cellulose pyrolysis (up to 36  C), shorten the temperature range of the active pyrolysis (up to 14  C), and increase the thermal decomposition rate (Isroi et al., 2011; Yang et al., 2011). A cost-competitive biological pretreatment of lignocellulose requires continuous studying and testing more microorganisms for their ability to delignify the plant material quickly and efficiently (Saritha and Lata, 2011). Also, integrated methods, such as, cotreatment with 59 organic solvents, diluted acids, supercritical CO2 and ionic liquids (ILs); mutation breeding and crossbreeding of fungal mycelia to obtain engineering strains; and integration of fungal pretreatment with simultaneous saccharification and fermentation to produce biofuels and value-added products should be studied (Tian et al., 2012). Physical Pretreatments Physical methods involve breakdown of biomass size by coarse size reduction, chipping, shredding, grinding, and milling in order to increase the available specific surface area and reduce the degree of polymerization, enhancing the digestibility of lignocellulosic biomass (Agbor et al., 2011; Brodeur et al., 2011). However, it has been shown that further reduction of biomass particle size below 0.4 mm has little effect on the rates and yields of biomass hydrolysis (Agbor et al., 2011). Chipping reduces heat and mass transfer limitations; grinding and milling are more effective at reducing the particle size and cellulose crystallinity than chipping, probably as result of the shear forces generated during milling. Vibratory ball milling has been used with more effective results in reducing cellulose crystallinity than ordinary ball milling. Also, disk milling, which produces fibers, has been reported as more efficient in enhancing cellulose hydrolysis than hammer milling, which produces finer bundles (Agbor et al., 2011). Stirring ball milling also could significantly damage the structure of biomass, resulting in the variation of surface morphology, the increase in amorphous region ratio and hydrogen bond energy, and the decrease in crystallinity and crystalline size (Liao et al., 2011). The energy requirements for physical pretreatments are dependent on the biomass characteristics, final particle size and reduction in crystallinity, for example, hardwoods require more energy input than agricultural residues (Agbor et al., 2011). Taking into account the high-energy requirement of milling on an industrial scale and the rise in energy demands, this method is not economically feasible and likely will not be used in a full-scale process (Agbor et al., 2011; Saritha and Lata, 2011). In most cases where the only option available for pretreatment is physical, the required energy is higher than the theoretical energy content available in the biomass (Brodeur et al., 2011). The pretreated biomasses by physical methods are subjected to heating, mixing and shearing resulting in physical and chemical modifications (Karunanithy and Muthukumarappan, 2011; Lamsal et al., 2010; Saritha and Lata, 2011). Also, Agbor et al., in 2011, suggested that the materials can be milled after chemical pretreatment with significantly reduction of (1) milling energy consumption, (2) reduce cost of solid liquid separation
60 4. RECENT ADVANCEMENTS IN PRETREATMENT TECHNOLOGIES OF BIOMASS TO PRODUCE BIOENERGY because the pretreated chips can be easily separated, (3) eliminate energy-intensive mixing of pretreatment slurries, (4) liquid to solid ratio and (5) did not result in the production of fermentation inhibitors. Another physical method is extrusion that disrupts the lignocellulose structure and increases the accessibility of carbohydrates to enzyme attack. This method has reported the improvement on sugar recovery up to 63.5% (Karunanithy and Muthukumarappan, 2011). Other physical pretreatments involve the use of gamma rays that cleave the b-1,4 glycosidic bonds, thus giving a larger surface area and lower crystallinity. This method will undoubtedly be very expensive on a large scale with huge environmental and safety concerns (Agbor et al., 2011). Proton beam irradiation has also been tested reporting a glucose conversion of 68% of the theoretical maximum at 72 h (Kim et al., 2011b). The effect of microwave and microwave-chemical pretreatments on densification characteristics and physical quality of pellets has also been investigated showing that microwave pretreatment was significantly able to disintegrate the lignocellulosic structure of wheat and barley straw grinds (Kashaninejad and Tabil, 2011). These pretreatments also have been tested on barley husks, sweet sorghum bagasse, bamboo, coconut husk and garden biomass (Choudhary et al., 2012; Ding et al., 2012; Gabhane et al., 2011; Jackowiak et al., 2011; JankerObermeier et al., 2012; Roos et al., 2009; Wu et al., 2012). However, some restrictions on energy consumption must be accomplished to obtain a positive energy balance with these pretreatments. Ultrasounds and ultrasoundassisted alkaline pretreatment also has been reported (Sun et al., 2002; Velmurugan and Muthukumar, 2012b). On the other hand, some continuous system includes a wet disk milling of rice straw (Hideno et al., 2012) and pulsed electric field of wood chip and switchgrass (Kumar et al., 2011). Chemical Pretreatments In general, chemical pretreatments show a high degree of selectivity for the biomass component they degrade; they also involve relatively harsh reaction conditions, which may not be ideal in a biorefinery scheme due to the possible production of toxic substances and their possible effects on downstream biological processing (FitzPatrick et al., 2010). Degradation of lignin has been observed in most chemical pretreatments, and particularly in dilute-acid and lime pretreatments (Samuel et al., 2011). Acid treatments solubilize the hemicellulose, and by this, make the cellulose more accessible. The main reaction that occurs during acid pretreatment is the hydrolysis of hemicellulose, especially xylan as glucomannan is relatively acid stable. The condensation and precipitation of solubilized lignin components is an unwanted reaction, as it decreases digestibility (Hendriks and Zeeman, 2009). Dilute-acid pretreatment is considered as one of the promising pretreatment methods despite its highenergy (steam or electricity) requirements and/or corrosion-resistant high-pressure reactors, and extensive washing, which increases the cost (Isroi et al., 2011). On the other hand, pretreatments with strong acids for the ethanol production is not an attractive option, because there is a risk of formation of inhibiting compounds (Hendriks and Zeeman, 2009). Other weak organic acids such as lactic acid and phosphoric acid have been also investigated (Monauari et al., 2011). During alkaline pretreatment the first reactions taking place are solvation and saponification. This causes a swollen state of the biomass and makes it more accessible for enzymes and bacteria. At “strong” alkali concentrations, dissolution, “peeling” of end groups, alkaline hydrolysis, degradation and decomposition of dissolved polysaccharides can take place. Alkali extraction can also cause solubilization, redistribution and condensation of lignin and modifications in the crystalline state of the cellulose (Hendriks and Zeeman, 2009). ILs are generally defined as salts that melt at or below 100  C, providing liquids exclusively composed of ions. Simple inorganic salts (e.g. NaCl) melt at very high temperatures (803  C), rendering unfeasible their routine use as solvents for organic chemical processing (Tadesse and Luque, 2011). ILs have been termed green solvents due to their negligible vapor pressure (Patel and Lee, 2012). The application of ILs to biomass valorization and pretreatment recently started to attract a great deal of attention because they are capable of disrupting the hydrogen bonds between different polysaccharide chains, thus decreasing the compactness of cellulose and making the carbohydrate fraction more susceptible to hydrolysis (Tadesse and Luque, 2011). Additionally, the recovering and the recycling of ILs has been proposed for decreasing the cost of the pretreatment process (Tadesse and Luque, 2011). However, cost and energyintensive recycling of the solvents are major constraints preventing ILs from commercial viability (Fu and Mazza, 2011). Another drawback of ILs is the fact that cellulases are inactivated even at low concentrations of ILs (Wang et al., 2011a). The ILs pretreatment has been tested using coadjuvant metal or acid catalysts to obtain higher conversion and/or yields of intermediates (Tadesse and Luque, 2011). Organosolv pretreatment is the process to extract lignin from lignocellulosic feedstocks with organic solvents or their aqueous solutions. This process is similar to that used in industrial paper-making processes but the degree of delignification for pretreatment is not demanded to be as high as that of pulping. Generally, organosolv processes are conducted at high temperatures
TYPES OF PRETREATMENTS (100e250  C) using low boiling point solvents (methanol and ethanol), high boiling point alcohols (ethylene glycol, glycerol, tetrahydrofurfuryl alcohol) and other classes of organic compounds including ethers, ketones, phenols, organic acids, and dimethyl sulfoxide (Agbor et al., 2011). This pretreatment removes extensive lignin and nearly complete hemicellulose, enhancing the enzymatic digestibility as a consequence of the increase in accessible surface area and pore volume (Agbor et al., 2011; Zhao et al., 2009). The organosolv pretreatment is more expensive at present than the leading pretreatment processes; however, organosolv can provide some valuable byproducts that might lead it to be a promising pretreatment for biorefining lignocellulosic feedstock in the future (Zhao et al., 2009). The advantages of organosolv pretreatment includes, organic solvents are always easy to recover by distillation and recycled for pretreatment; the chemical recovery in organosolv pulping processes can isolate lignin as a solid material and carbohydrates as syrup, both of which show promise as chemical feedstocks. However, there are inherent drawbacks to the organosolv pretreatment, such as air and water pollution, the pretreated solids always need to be washed with organic solvent before water washing in order to avoid the reprecipitation of dissolved lignin, which leads to cumbersome washing arrangements (Zhao et al., 2009). Physicochemical Pretreatments The objective of steam pretreatment, steam explosion or liquid hot water, is to solubilize the hemicellulose to make the cellulose better accessible for enzymatic hydrolysis and to avoid the formation of inhibitors (Hendriks and Zeeman, 2009). During steam pretreatment parts of the hemicellulose hydrolyze and form acids, which could catalyze the further hydrolysis of the hemicellulose. To avoid the formation of inhibitors, the pH should be kept between 4 and 7 during the pretreatment (Hendriks and Zeeman, 2009). The aqueous fractionation of native lignocellulosic materials with hot, compressed water (also known as hydrothermal processing or autohydrolysis) has been proposed as a fractionation method for biorefineries, as it enables the simultaneous removal of watersoluble extractives and the solubilization of hemicelluloses, yielding a solid phase enriched in lignin and cellulose (Gullón et al., 2012). Liquid hot water has the major advantage that solubilized hemicellulose and lignin products are present in lower concentrations, when compared to steam pretreatment, due to higher water input. These lower concentrations reduce the risk on degradation products like furfural and the condensation and precipitation of lignin compounds (Hendriks and Zeeman, 2009). 61 Wet oxidation is another oxidative pretreatment method, which uses oxygen as oxidation agent. The soluble sugars produced during wet-oxidation pretreatment are mainly polymers opposite to the monomers produced during steaming or acid hydrolysis as pretreatment. Phenolic monomers are no end products during wet oxidation but are further degraded to carboxylic acids (Hendriks and Zeeman, 2009; Martin and Thomsen, 2007). Carbon dioxide pretreatment is conducted with highpressure carbon dioxide at high temperatures of up to 200  C with duration of several minutes. Explosive steam pretreatment with high-pressure carbon dioxide causes the liquid to be acidic and this acid hydrolyses especially the hemicellulose. Carbon dioxide is also applied as supercritical carbon dioxide (35  C, 73 bars) for depolymerization of the sugars present in biomass, increasing the glucose yield probably caused by increase in pore size (Hendriks and Zeeman, 2009). This method is considered as a “green” pretreatment because it does not require neutralization or pH adjustment prior to enzymatic hydrolysis (King et al., 2012). Ammonia fiber explosion (AFEX), ammonia recycled percolation (ARP) and soaking aqueous ammonia (SAA) are alkaline pretreatment methods that use liquid ammonia to pretreat biomass. The difference between AFEX and ARP processes is that the first is carried out in liquid ammonia and the second one in an aqueous ammonia solution. AFEX is a physicochemical pretreatment process in which lignocellulosic biomass is exposed to liquid ammonia at high temperature and pressure for a period of time, and then the pressure is suddenly reduced (Kumar et al., 2009). The AFEX pretreatment simultaneously reduces lignin content and removes some hemicellulose while decrystallizing cellulose and it has the advantage of ammonia being recyclable due to its high volatility (Yang and Wyman, 2008). AFEX has been shown to complete conversion of cellulose to fermentable sugars but removes or loses little lignin or hemicellulose. In a typical AFEX process, the dosage of liquid ammonia is 1e2 kg of ammonia/kg of dry biomass, the temperature is 90  C, and the residence time is 30 min (Kumar et al., 2009). However, AFEX pretreatment at 40  C and longer residence times, up to 8 h, has also been proposed with comparable yields of sugar and ethanol (Bals et al., 2012). AFEX treatment is a batch process while continuous processing in an extruder is an approach called FIBEX (fiber extrusion) that significantly reduces both the time required for treatment and the ammonia levels required with similar hydrolysis results to those for AFEX (Yang and Wyman, 2008). ARP is another process based on ammonia, which recycles aqueous ammonia solution (5e15 wt%) through
62 4. RECENT ADVANCEMENTS IN PRETREATMENT TECHNOLOGIES OF BIOMASS TO PRODUCE BIOENERGY a reactor packed with biomass at elevated temperatures (80e180  C). Ammonia in aqueous solution and at high temperature breaks down lignin via the ammoniolysis reaction but has virtually no effect on carbohydrates (Geddes et al., 2011). A major challenge for ARP is to reduce liquid loadings to keep energy costs low (Yang and Wyman, 2008). SAA is a modified version of AFEX but it uses moderate temperatures (25e60  C) to reduce the liquid amount during pretreatment. At ambient temperatures the duration could be up to 10e60 days while at higher temperatures (150e190  C) the duration of pretreatment is reduced to minutes (Agbor et al., 2011). The cost of ammonia, and especially of ammonia recovery, drives the cost of the ammoniarelated pretreatments (Kumar et al., 2009). TRENDS IN PRETREATMENTS A search made in the database Current Contents ConnectÒ for key words “lignocellulosic biomass” and “pretreatment” resulted in a total of 1217 articles published from 2000 to 2012. From the total, 91.2% corresponded to research papers and 8.8% to review papers (Figure 4.1). The number of published papers has increased exponentially from 2007 to 2012 as shown in Figure 4.1, indicating the relevance that the topic has gained in the recent years. Among the reviews documents, 37.2% corresponded to reviews directly related to pretreatments, pointing out the importance of this step in the concept of biorefinery. The remaining documents corresponded to reviews on the general topic of biofuels from lignocellulosic biomass. FIGURE 4.2 Classification of published papers from 2000 to 2012 in the topic of pretreatments of lignocellulosic biomass. Source: With data of Current Contents ConnectÒ. For the research papers, the search was refined to select only papers that focus on the pretreatment processes obtaining a total of 692 papers and around 54% of these papers were published in 2012. The research papers related to pretreatment were classified into nine categories as shown in Figure 4.2. From this analysis we can conclude that alkaline, acid, thermal and IL pretreatments are the most reported, the comparison and combination of them is also widely reported. However, this variety of studied pretreatments indicates that there is not a prevalent pretreatment suggesting that further investigation on the topic is required. Other Pretreatments FIGURE 4.1 Published papers from 2000 to 2012 in the topic ) Original papers and ( ) review lignocellulosic biomass. ( papers. Source: With data of Current Contents ConnectÒ. The pretreatments described above such as steam explosion, liquid hot water, dilute acid, lime, and ammonia pretreatments are the most studied methods because they have potential as cost-effective pretreatments (Kazi et al., 2010; Mosier et al., 2005; Piccolo and Bezzo, 2009; Tao et al., 2011; Wyman et al., 2005). Other alternatives such as biological, ultrasonication, microwave, organosolvs, ILs, and combinatorial methods are also essayed; however, they are either low effective, long-time treatment or too expensive, and further investigation and improvements have to be reached before they can be competitive. In this section multiple or combinatorial pretreatments and other alternative pretreatments will be discussed. Biological pretreatments must decrease the time of the process in order to be competitive in an industrial concept of biorefinery; to reach this objective its combination with chemicals and/or physical methods
TRENDS IN PRETREATMENTS has been proposed by several studies. For example, the combination of a biological pretreatment by I. lacteus or P. subvermispora with a mild alkali pretreatment improved significantly ethanol production without the production of inhibitor compounds for downstream processes (Salvachua et al., 2011; Zhong et al., 2011). Other two-step pretreatment proposed consisted in a mild physical or chemical step (ultrasonic and H2O2) and a subsequent biological treatment by P. ostreatus, increasing significantly the lignin degradation compared to those of one-step pretreatments (Yu et al., 2009); also, pretreatment by white-rot fungi has been combined with organosolv pretreatment in an ethanol production process from beech wood chips (Salvachua et al., 2011); the combination of biological and mild acid pretreatment was reported as a promising method to improve enzymatic hydrolysis and ethanol production from water hyacinth with low lignin content (Ma et al., 2010). Another combination of biological pretreatment with thermal processing for wheat straw consisted in a first phase of biodegradation by P. chrysosporium (10 days) and a thermal decomposition using pyrolysis (Zeng et al., 2011). Also, a combination of fungal treatment with liquid hot water treatment was conducted to enhance the enzymatic hydrolysis of Populus tomentosa (Wang et al., 2012). Sugarcane bagasse is one of the most promising biomass considered in biorefineries; thus, several studies have proposed combined pretreatments. The ultrahigh-pressure explosion combined with alkaline treatment (0.5% NaOH) at 125  C for 120 min significantly decreased the particle size and disrupted the microstructure, with a significant delignification and increased enzymatic digestibility of sugarcane bagasse (Chen et al., 2010). A combined treatment with dilute sulfuric acid and microwave heating up to 190  C for 5 min has also been studied. This treatment resulted in an increment of the specific surface area of bagasse, almost complete removal of hemicellulose and significant reduction of the crystalline structure of cellulose (Chen et al., 2011), while microwaveealkali treatment at 450 W for 5 min resulted in almost 90% of lignin removal from the bagasse (Binod et al., 2012). Also, bagasse has been subjected to sono-assisted alkaline pretreatment (Velmurugan and Muthukumar, 2012a). Acid, alkaline or sequential acid/alkaline solutions have been tested to conversion into bio-oil in a pyrolysis process at low-temperature conversion under He or O2/He atmospheres at 350e450  C (Cunha et al., 2011). A twostage process for delignification of sugarcane bagasse uses alkali and peracetic acid combination (Teixeira et al., 2000; Zhao et al., 2011b). Same strategies (acidic/alkaline) have been proposed for corn stover. For example, a two-stage process consists of use of 0.07 wt% sulfuric acid at 170  C, 2.5 ml/min for 30 min and ARP (15 wt% ammonia) 63 at the same temperature, 5.0 ml/min for 60 min. In the first stage hemicellulose was recovered while in the following stage lignin was recovered. This treatment brought about enzymatic digestibility of 90% using 60 filter paper units/g glucan cellulase enzyme loadings (Kim, 2011). Another combined treatment proposed for corn stover is SAA (15 wt% ammonia) with solution containing also 20 wt% ethanol at 60  C for 24 h preserving the hemicellulose in solid form (Kim et al., 2009). Also, the use of NaOH (0.3 N) and a step of particle size homogenization has reported a significant enhancement of enzymatic hydrolysis (Li et al., 2004). The synergistic effect of preimpregnation by sulfuric acid (3 wt%) and steam explosion (190  C) has been investigated; after 48 h of digestion the yield of glucose was 93% of the theoretical (Zimbardi et al., 2007). Sequential stages of autohydrolysis and ethanole water mixtures were used to pretreat olive tree trimmings recovering up to 42% of the polysaccharides contained in the raw material (Requejo et al., 2011). Also, this process has been tested with uncatalyzed ethanolewater solutions of Eucalyptus globulus wood (Romani et al., 2011). Mixtures of ethanol/water/acetic acid in an autoclave have been also used (Teramoto et al., 2008). This combined process causes the solubilization of hemicelluloses and lignin, leaving solids enriched in cellulose. A treatment of ethanosolv catalyzed with FeCl3 (0.1 M) at 170  C for 72 h has been proposed for barley straw allowing enzymatic digestibility of 89%. This treatment had a particularly strong effect on enzymatic digestibility and cellulose recovery (Kim et al., 2010). Another pretreatment at pH 1 (hydrochloric acid) and subsequently at pH 13 (sodium hydroxide) released 69% and 95% of the theoretical maximal amounts of glucose and xylose, respectively, from the straw and removal of 68% of the lignin (Pedersen et al., 2010). The opposite sequence alkaline stage (ammonia) followed by acidic stage (dilute sulfuric acid by percolation) has also been used to treated rice straw (Kim et al., 2011a). Microwave-based heating (190  C) was used to pretreat switchgrass presoaked in alkali solutions (0.1 g/g) resulting in release of 90% of maximum potential sugars. This value was significantly higher than the one obtained with conventional heat and it was attributed to the disruption of recalcitrant structures under microwave heating (Hu and Wen, 2008). Significant disintegration of lignocellulosic structure of wheat, barley straw grinds, switchgrass and coastal bermuda grass has been reported with the microwaveechemical (NaOH or Ca(OH)2) pretreatments (Kashaninejad and Tabil, 2011; Keshwani and Cheng, 2010). Also, microwave-assisted pretreatment of woody biomass with ammonium molybdate activated by H2O2 has also been proposed resulting in a selective delignifying system (Verma et al., 2011).
64 4. RECENT ADVANCEMENTS IN PRETREATMENT TECHNOLOGIES OF BIOMASS TO PRODUCE BIOENERGY For hydrogen production from Miscanthus by Thermotoga elfii, high delignification values were obtained by the combination of mechanical (one-step extrusion) and chemical pretreatments (NaOH at 70  C) resulting in a 33% conversion into monosaccharides of the initial biomass after enzymatic hydrolysis (de Vrije et al., 2002). A two-stage pretreatment method was proposed and tested for deconstruction of Miscanthus; first, biomass is pretreated at 50  C, 1.0e4.0% alkaline peroxide solutions to remove up to 64% of hemicellulose and 64% of lignin. The remaining solids were subjected to a second pretreatment at 121  C with electrolyzed water (Wang et al., 2010). On the other hand, application of a dehydration process to the mechanochemical pretreatment process of the bioethanol production system has been proposed for energy saving and cost reduction. However, the dehydration process has problems with the loss of sugars eluted in the liquid phase during the hydrothermal process (Yanagida et al., 2011). Combination of hot compressed water (hydrothermal treatment) and mechanochemical milling, including a dewatering step for Eucalyptus and rice straw, has been proposed for ethanol production (Fujimoto et al., 2008; Hideno et al., 2012). Torrefaction is a mild thermal pretreatment (T < 300 C) that improves biomass milling and storage properties (Chen et al., 2012; Fisher et al., 2012). This treatment has gained attention in recent years and some biomasses that have been treated include oil palm fiber and eucalyptus, Norwegian birch, spruce, Miscanthus and white oak sawdust; residues from coffee grain, sugarcane, sawdust and rice husk bagasse (Chen et al., 2012; Lu et al., 2012; Medic et al., 2012; Protasio et al., 2012; Srinivasan et al., 2012; Tapasvi et al., 2012; Tumuluru et al., 2012). Wet torrefaction (hot compressed water 200e260  C) and dry (nitrogen, 250e300  C) has been tested with Loblolly pine with mass yield of solid product ranging between 57% and 89%, and energy densification to 108e136% of the original feedstock (Yan et al., 2009). Extrusion has also been used in combination with alkali (1.70%, w/v NaOH) soaking for pretreatment of prairie cord grass at a barrel temperature of 114  C, 122 rpm screw speed resulted in an 82% of sugar recovery after enzymatic hydrolysis (Karunanithy and Muthukumarappan, 2011). An alkali-combined extrusion pretreatment of corn stover obtained glucose and xylose sugar yields of 86.8% and 50.5%, respectively. The conditions used were alkali loading of 0.04 g/g dry biomass, a screw speed of 80 rpm, residence time for extrusion is 27 min, temperature of 140  C and washed with water (Zhang et al., 2012b). Also, glucose conversion of 95% was reported from soybean hulls using a thermomechanical extrusion pretreatment (screw speed 350 rpm, 80  C and in-barrel moisture content 40% wt) (Yoo et al., 2011). A study of high-temperature (110e130  C), concentratedacid (5e30 wt.%) hydrolysis kinetics was undertaken for pretreated pine in a corotating twin-screw extruder reactor, obtaining more than 50% of the theoretical glucose in roughly 25 min (Miller and Hester, 2007). A successive pretreatment of ball-milled bamboo consisted in ultrasound treatment in ethanol solution at 20  C from 0 up to 50 min. After that the samples were dissolved with 7% NaOH/12% urea solutions at 12  C, followed by successive extractions with dioxane, ethanol, and dimethyl sulfoxide (Li et al., 2010). Other treatments, such as SAA and proton beam irradiation, have been tested with rice straw and approximately 90% of the theoretical glucose conversion was obtained at 12 h (Kim et al., 2011b). Microwave pretreatment also has been combined with alkali to pretreat cashew apple bagasse founding that alkali exerted influence on glucose formation (Rodrigues et al., 2011). A pretreatment method using ammonia and ILs reported a synergy effect for rice straw, achieving 82% of the cellulose recovery and 97% of the enzymatic glucose conversion with recycling of the ILs (Nguyen et al., 2010). Pretreatment of wheat straw with combined sulfuric acid (0e3%, w/v) and Tween-20 (concentration, 0e1%) was evaluated with modification of lignin surface (Qi et al., 2010). Other surfactants, such as, Tween-80, dodecylbenzene sulfonic acid, and polyethylene glycol 4000, have also been used combined with diluted acid to treat corn stover and bagasse (Qing et al., 2010; Sindhu et al., 2012). Other pretreatments include technology used in kraft pulp mills for the efficient conversion of lignocellulosic biomass into ethanol (Gonzalez et al., 2011). Sulfite pretreatment to overcome recalcitrance of lignocellulose consists of sulfite treatment of wood chips under acidic conditions followed by mechanical size reduction using disk refining (Li et al., 2012; Zhang et al., 2012a). Pretreatment of corn stalk with sulfite (7%) at a temperature of 180  C for 30 min was successfully performed (Liu et al., 2011; Zhu et al., 2009). Silage preparation is a well-known procedure for preserving plant material; the effects of Fe(NO3)3 pretreatment conditions on sugar yields were investigated for corn stover silage. Ensiling techniques, with and without supplemental enzymes, also have been reported as a cost-effective pretreatment (Chen et al., 2007; Sun et al., 2011; Thomsen et al., 2008). Also, FeSO4 (0.1 mol/L at 180  C for 20 min) was investigated as a catalyst for the pretreatment of corn stover, observing significantly increased hemicellulose degradation in aqueous solutions with high xylose recovery and low cellulose removal (Zhao et al., 2011a). Lignocellulose pretreatment featuring modest reaction conditions (50  C and atmospheric pressure) was demonstrated to fractionate lignocellulose to amorphous
ENVIRONMENTAL AND ECONOMICAL ASPECTS cellulose, hemicellulose, lignin, and acetic acid by using a nonvolatile cellulose solvent (concentrated phosphoric acid), a highly volatile organic solvent (acetone) and water (Zhang et al., 2007). PRETREATMENT MODELING For a rational design of pretreatment processes is required experimental investigation of physical changes and chemical reactions that occur during pretreatment; however, due to the wide range of pretreatments and biomass available for biorefinery, the development of effective and mechanistic models can provide a large amount of information to optimize operational conditions. Furthermore, several key criteria regarding technical, economical, and environmental considerations should be critically analyzed when adapting these technologies for the nascent biorefinery industry (Sousa et al., 2009). In this section some models that particularly focus on pretreatments in the scheme of a biorefinery plant are discussed. The most commonly developed models for the pretreatment are kinetic models with assumptions of a first-order dependence of reaction rate on biomass components and an Arrhenius-type correlation between rate constant and temperature (Wang et al., 2011b). In view of the heterogeneous nature of the reactions involved in the pretreatment, the uses of severity factor, artificial neural network, and fuzzy inference systems, represent alternative approaches for predicting the behavior of the systems (Wang et al., 2011b). A multiscale model of hydrothermal pretreatment methods, including microscale, mesoscale and macroscale, was used to elucidate the mechanisms involved in the breakage of hemicellulose of wood (Hosseini and Shah, 2009). A model that simulates a biorefinery plant integrating first- and second-generation ethanol production process from sugarcane, surplus bagasse and trash included a selected pretreatment method followed, or not, by a delignification step. The simulation indicated that the best results were obtained for steam explosion pretreatment at high solids loading and hydrolysis time between 24 and 48 h (Dias et al., 2011). Also, a mathematical model for a countercurrent shrinking-bed reactor for pretreatment/hydrolysis of hardwood cellulose predicts that dilute sulfuric acid (0.08 wt%) and with optimal adjustment of other operating parameters resulted in 80e90% yield with 2e4 wt% product concentration. This model also indicates that acid concentration and temperatures acutely affect the reactor performance in cellulose hydrolysis. In contrast, hemicellulose hydrolysis is less sensitive to acid concentration and temperature allowing broader latitude in operating 65 conditions (Lee et al., 2000). Also, methods of optimization have been used to acid-catalyzed pretreatment process showing that the sulfuric acid concentration plays the major role during the pretreatment of areca nut husk (57% contribution) followed by the duration of operation (24.98% contribution) and solid loading (14.3% contribution) (Sasmal et al., 2011). Other authors have reported that for alkali pretreatment of cereal crop residues, the temperature had the greatest impact on sugar release, followed by alkali concentration and treatment time (Vancov and McIntosh, 2011). A statistical optimization method proposed variables such as temperature, sulfuric acid concentration and reaction time to release xylose from sugarcane bagasse as a useful means of trading off the combined effects of these three variables on total xylose recovery yields (Um and Bae, 2011). Other works considered also the reduction of the acid concentrations and reaction times to optimize the pretreatment process. This kinetic model predicted optimum conditions to pretreatment of corn stover of 150  C, 0.6% HNO3 and 1 min of reaction time for maximal xylose, glucose and arabinose yields and minimal yield of acetic acid and furfural (Zhang et al., 2011). To deeply understand the factors that affect the conversion of lignocellulosic biomass to fermentable sugars, experimental results should be bridged with process simulations (Wang et al., 2011b). ENVIRONMENTAL AND ECONOMICAL ASPECTS A renewable biofuel economy is projected as a pathway to decrease dependence on fossil fuels as well as to reduce greenhouse gas (GHG) emissions. Ethanol produced on large scale from lignocellulosic raw materials is considered the most potential next-generation automotive fuel. However, the environmental impact of biorefinery processes must be assessed. The GHG emissions have been evaluated in some processes of ethanol production for many authors, reporting that ethanol emissions could be 20e90% lower than those from fossil fuels, depending on the scheme of production and the biomass used. Some works have proposed the life cycle assessment model to evaluate the environmental implications of the production of ethanol founding that compared to conventional gasoline, life cycle GHG emissions are lower for ethanol blends, specifically up to 145% lower for E85 mixture (85% ethanol and 15% gasoline v/v) derived from Ethiopian mustard, associated with the low intensive energy and high biomass yield of this crop (Gonzalez-Garcia et al., 2010). Production of fuel ethanol from lignocellulosic feedstock has been modeled for design optimization through mass and energy balances in terms of ethanol yield and
66 4. RECENT ADVANCEMENTS IN PRETREATMENT TECHNOLOGIES OF BIOMASS TO PRODUCE BIOENERGY power generation as well as from a financial point of view in order to identify critical parameters of the processes productivity and profitability (Piccolo and Bezzo, 2009). Technoeconomic analysis will play a key role in process development and targeting of technical and economic barriers for these new fuels and feedstocks (Aden and Foust, 2009). Some of these economic analyses indicate that about 18e20% of the total projected cost for biological production of cellulosic ethanol can be attributed to pretreatment; then, reducing ethanol cost requires optimizing pretreatment strategies and conditions to the most economical possible and to accelerate commercial applications (Banerjee et al., 2010; Yang and Wyman, 2008). Most of the current process design and economic results are described for dilute-acid pretreatment followed by enzymatic hydrolysis and fermentation. The projection made in 2007 of some models for ethanol costs in 2012 at commercial scale of corn stover conversion process was $0.35 per liter (Aden and Foust, 2009). However, commercial plants from lignocellulosic materials are still under development. The variation in estimated ethanol production cost is considerable, ranging from about 0.13 to 0.81 US$ per liter ethanol. This can be explained to a large extent by actual process differences and variations in the assumptions underlying the technoeconomic evaluations (Galbe et al., 2007). Other studies indicate that dilute-acid pretreatment process has the lowest product value compared to hot water and AFEX pretreatments for three downstream process variations (pervaporation, separate five-carbon and six-carbon sugars fermentation, and onsite enzyme production) (Kazi et al., 2010). The technical and economic challenges for softwood to ethanol processes with SO2-catalyzed steam explosion and ethanol organosolv pretreatments have been analyzed concluding that organosolv pretreatment has the advantage of high-value coproduct from lignin (Mabee et al., 2006). CONCLUDING REMARKS The technical feasibility of extracting sugars from lignocellulosic biomass applying a pretreatment process and the subsequent steps of enzymatic hydrolysis and fermentation to obtain bioenergy and other products has been established for the past decades. However, the economical feasibility of this process is still under development. The biochemical and chemical paths of biorefineries are not economically competitive with first-generation biofuels. The steam explosion, liquid hot water, dilute acid, lime, and ammonia pretreatments are the most studied methods because they have potential as cost-effective pretreatments. Furthermore, other alternatives such as biological, ultrasonication, microwave, organosolvs, and IL pretreatments require further investigation and improvements have to be reached before they can be competitive. The several options proposed in pretreatments indicate that there is not a prevalent technology and it is probable that each biorefinery plant should have to select the best option according to their feedstock and target products. Further research must be performed in order to evaluate if one pretreatment or combined stages of pretreatment are economical and environmentally feasible prior to enzymatic hydrolysis step. Also, consolidated process of pretreatment, hydrolysis or even including fermentation process, should be evaluated from both economical and environmental points of view. References Aden, A., Foust, T., 2009. Technoeconomic analysis of the dilute sulfuric acid and enzymatic hydrolysis process for the conversion of corn stover to ethanol. Cellulose 16, 535e545. Agbor, V.B., Cicek, N., Sparling, R., Berlin, A., Levin, D.B., 2011. Biomass pretreatment: fundamentals toward application. Biotechnol. Adv. 29, 675e685. Alvira, P., Tomas-Pejo, E., Ballesteros, M., Negro, M.J., 2010. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour. Technol. 101, 4851e4861. Bals, B.D., Teymouri, F., Campbell, T., Jin, M., Dale, B.E., 2012. Low temperature and long residence time AFEX pretreatment of corn stover. BioEnergy Res. 5, 372e379. Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., Pandey, R.A., 2010. Commercializing lignocellulosic bioethanol: technology bottlenecks and possible remedies. Biofuels, Bioprod. Biorefin. 4, 77e93. Binod, P., Satyanagalakshmi, K., Sindhu, R., Janu, K.U., Sukumaran, R.K., Pandey, A., 2012. Short duration microwave assisted pretreatment enhances the enzymatic saccharification and fermentable sugar yield from sugarcane bagasse. Renewable Energy 37, 109e116. Brodeur, G., Yau, E., Badal, K., Collier, J., Ramachandran, K.B., Ramakrishnan, S., 2011. Chemical and physicochemical pretreatment of lignocellulosic biomass: a review. Enzyme Res. 2011, 1e17. Carvalheiro, F., Duarte, L.C., Gı́rio, F.M., 2008. Hemicellulose biorefineries: a review on biomass pretreatments. J. Sci. Ind. Res. 67, 849e864. Chen, D., Guo, Y., Huang, R.B., Lu, Q., Huang, J., 2010. Pretreatment by ultra-high pressure explosion with homogenizer facilitates cellulase digestion of sugarcane bagasses. Bioresour. Technol. 101, 5592e5600. Chen, W.-H., Lu, K.-M., Tsai, C.-M., 2012. An experimental analysis on property and structure variations of agricultural wastes undergoing torrefaction. Appl. Energy 100, 318e325. Chen, W.H., Tu, Y.J., Sheen, H.K., 2011. Disruption of sugarcane bagasse lignocellulosic structure by means of dilute sulfuric acid pretreatment with microwave-assisted heating. Appl. Energy 88, 2726e2734. Chen, Y., Sharma-Shivappa, R.R., Chen, C., 2007. Ensiling agricultural residues for bioethanol production. Appl. Biochem. Biotechnol. 143, 80e92.
REFERENCES Choudhary, R., Umagiliyage, A.L., Liang, Y., Siddaramu, T., Haddock, J., Markevicius, G., 2012. Microwave pretreatment for enzymatic saccharification of sweet sorghum bagasse. Biomass Bioenergy 39, 218e226. Cui, Z., Wan, C., Shi, J., Sykes, R.W., Li, Y., 2012. Enzymatic digestibility of corn stover fractions in response to fungal pretreatment. Ind. Eng. Chem. Res. 51, 7153e7159. Cunha, J.A., Pereira, M.M., Valente, L.M.M., de la Piscina, P.R., Homs, N., Santos, M.R.L., 2011. Waste biomass to liquids: low temperature conversion of sugarcane bagasse to bio-oil. The effect of combined hydrolysis treatments. Biomass Bioenergy 35, 2106e2116. de Vrije, T., de Haas, G.G., Tan, G.B., Keijsers, E.R.P., Claassen, P.A.M., 2002. Pretreatment of Miscanthus for hydrogen production by Thermotoga elfii. Int. J. Hydrogen Energy 27, 1381e1390. Dias, M.O.S., Cunha, M.P., Maciel, R., Bonomi, A., Jesus, C.D.F., Rossell, C.E.V., 2011. Simulation of integrated first and second generation bioethanol production from sugarcane: comparison between different biomass pretreatment methods. J. Ind. Microbiol. Biotechnol. 38, 955e966. Ding, T.Y., Hii, S.L., Ong, L.G.A., 2012. Comparison of pretreatment strategies for conversion of coconut husk fiber to fermentable sugars. Bioresources 7, 1540e1547. Fisher, E.M., Dupont, C., Darvell, L.I., Commandre, J.M., Saddawi, A., Jones, J.M., Grateau, M., Nocquet, T., Salvador, S., 2012. Combustion and gasification characteristics of chars from raw and torrefied biomass. Bioresour. Technol. 119, 157e165. FitzPatrick, M., Champagne, P., Cunningham, M.F., Whitney, R.A., 2010. A biorefinery processing perspective: treatment of lignocellulosic materials for the production of value-added products. Bioresour. Technol. 101, 8915e8922. Fu, D.B., Mazza, G., 2011. Aqueous ionic liquid pretreatment of straw. Bioresour. Technol. 102, 7008e7011. Fujimoto, S., Inoue, H., Yano, S., Sakaki, T., Minowa, T., Endo, T., Sawayama, S., Sakanishi, K., 2008. Bioethanol production from lignocellulosic biomass requiring no sulfuric acid: mechanochemical pretreatment and enzymic saccharification. J. Jpn. Pet. Inst. 51, 264e273. Gabhane, J., William, S., Vaidya, A.N., Mahapatra, K., Chakrabarti, T., 2011. Influence of heating source on the efficacy of lignocellulosic pretreatment - a cellulosic ethanol perspective. Biomass Bioenergy 35, 96e102. Galbe, M., Sassner, P., Wingren, A., Zacchi, G., 2007. Process engineering economics of bioethanol production. Biofuels 108, 303e327. Geddes, C.C., Nieves, I.U., Ingram, L.O., 2011. Advances in ethanol production. Curr. Opin. Biotechnol. 22, 312e319. Giles, R.L., Galloway, E.R., Elliott, G.D., Parrow, M.W., 2011. Two-stage fungal biopulping for improved enzymatic hydrolysis of wood. Bioresour. Technol. 102, 8011e8016. Gonzalez, R., Treasure, T., Phillips, R., Jameel, H., Saloni, D., 2011. Economics of cellulosic ethanol production: green liquor pretreatment for softwood and hardwood, greenfield and repurpose scenarios. Bioresources 6, 2551e2567. Gonzalez-Garcia, S., Moreira, M.T., Feijoo, G., 2010. Comparative environmental performance of lignocellulosic ethanol from different feedstocks. Renewable Sustainable Energy Rev. 14, 2077e2085. Gullón, P., Romani, A., Vila, C., Garrote, G., Parajo, J.C., 2012. Potential of hydrothermal treatments in lignocellulose biorefineries. Biofuels, Bioprod. Biorefin. 6, 219e232. Hendriks, A., Zeeman, G., 2009. Pretreatments to enhance the digestibility of lignocellulosic biomass. Bioresour. Technol. 100, 10e18. Hideno, A., Inoue, H., Yanagida, T., Tsukahara, K., Endo, T., Sawayama, S., 2012. Combination of hot compressed water 67 treatment and wet disk milling for high sugar recovery yield in enzymatic hydrolysis of rice straw. Bioresour. Technol. 104, 743e748. Hosseini, S.A., Shah, N., 2009. Multiscale modelling of biomass pretreatment for biofuels production. Chem. Eng. Res. Des. 87, 1251e1260. Hu, Z.H., Wen, Z.Y., 2008. Enhancing enzymatic digestibility of switchgrass by microwave-assisted alkali pretreatment. Biochem. Eng. J. 38, 369e378. Isroi Millati, R., Syamsiah, S., Niklasson, C., Cahyanto, M.N., Lundquist, K., Taherzadeh, M.J., 2011. Biological pretreatment of lignocelluloses with white-rot fungi and its applications: a Review. Bioresources 6, 5224e5259. Jackowiak, D., Bassard, D., Pauss, A., Ribeiro, T., 2011. Optimisation of a microwave pretreatment of wheat straw for methane production. Bioresour. Technol. 102, 6750e6756. Janker-Obermeier, I., Sieber, V., Faulstich, M., Schieder, D., 2012. Solubilization of hemicellulose and lignin from wheat straw through microwave-assisted alkali treatment. Ind. Crops Prod. 39, 198e203. Jørgensen, H., Kristensen, J.B., Felby, C., 2007. Enzymatic conversion of lignocellulose into fermentable sugars: challenges and opportunities. Biofuels, Bioprod. Biorefin. 1, 119e134. Karunanithy, C., Muthukumarappan, K., 2011. Optimization of alkali soaking and extrusion pretreatment of prairie cord grass for maximum sugar recovery by enzymatic hydrolysis. Biochem. Eng. J. 54, 71e82. Kashaninejad, M., Tabil, L.G., 2011. Effect of microwave-chemical pretreatment on compression characteristics of biomass grinds. Biosyst. Eng. 108, 36e45. Kazi, F.K., Fortman, J.A., Anex, R.P., Hsu, D.D., Aden, A., Dutta, A., Kothandaraman, G., 2010. Techno-economic comparison of process technologies for biochemical ethanol production from corn stover. Fuel 89, S20eS28. Keshwani, D.R., Cheng, J.J., 2010. Microwave-based alkali pretreatment of switchgrass and coastal Bermuda grass for bioethanol production. Biotechnol. Prog. 26, 644e652. Kim, J.W., Kim, K.S., Lee, J.S., Park, S.M., Cho, H.Y., Park, J.C., Kim, J.S., 2011a. Two-stage pretreatment of rice straw using aqueous ammonia and dilute acid. Bioresour. Technol. 102, 8992e8999. Kim, S.B., Kim, J.S., Lee, J.H., Kang, S.W., Park, C., Kim, S.W., 2011b. Pretreatment of rice straw by proton beam irradiation for efficient enzyme digestibility. Appl. Biochem. Biotechnol. 164, 1183e1191. Kim, T.H., 2011. Sequential hydrolysis of hemicellulose and lignin in lignocellulosic biomass by two-stage percolation process using dilute sulfuric acid and ammonium hydroxide. Korean J. Chem. Eng. 28, 2156e2162. Kim, T.H., Nghiem, N.P., Hicks, K.B., 2009. Pretreatment and fractionation of corn stover by soaking in ethanol and aqueous ammonia. Appl. Biochem. Biotechnol. 153, 171e179. Kim, Y., Yu, A., Han, M., Choi, G.W., Chung, B., 2010. Ethanosolv pretreatment of barley straw with iron(III) chloride for enzymatic saccharification. J. Chem. Technol. Biotechnol. 85, 1494e1498. King, J.W., Srinivas, K., Guevara, O., Lu, Y.-W., Zhang, D., Wang, Y.-J., 2012. Reactive high pressure carbonated water pretreatment prior to enzymatic saccharification of biomass substrates. J. Supercrit. Fluids 66, 221e231. Kuhar, S., Nair, L.M., Kuhad, R.C., 2008. Pretreatment of lignocellulosic material with fungi capable of higher lignin degradation and lower carbohydrate degradation improves substrate acid hydrolysis and the eventual conversion to ethanol. Can. J. Microbiol. 54, 305e313. Kumar, P., Barrett, D.M., Delwiche, M.J., Stroeve, P., 2009. Methods for pretreatment of lignocellulosic biomass for efficient hydrolysis and biofuel production. Ind. Eng. Chem. Res. 48, 3713e3729.
68 4. RECENT ADVANCEMENTS IN PRETREATMENT TECHNOLOGIES OF BIOMASS TO PRODUCE BIOENERGY Kumar, P., Barrett, D.M., Delwiche, M.J., Stroeve, P., 2011. Pulsed electric field pretreatment of switchgrass and wood chip species for biofuel production. Ind. Eng. Chem. Res. 50, 10996e11001. Lamsal, B., Yoo, J., Brijwani, K., Alavi, S., 2010. Extrusion as a thermomechanical pre-treatment for lignocellulosic ethanol. Biomass Bioenergy 34, 1703e1710. Lee, Y.Y., Wu, Z.W., Torget, R.W., 2000. Modeling of countercurrent shrinking-bed reactor in dilute-acid total-hydrolysis of lignocellulosic biomass. Bioresour. Technol. 71, 29e39. Lewis, R.S., Datar, R.P., Huhnke, R.L., 2005. Biomass to ethanol. In: Encyclopedia of Chemical Processing, Vol. 1, pp. 142e151. Li, M.F., Fan, Y.M., Xu, F., Sun, R.C., Zhang, X.L., 2010. Cold sodium hydroxide/urea based pretreatment of bamboo for bioethanol production: characterization of the cellulose rich fraction. Ind. Crops Prod. 32, 551e559. Li, X., Luo, X., Li, K., Zhu, J.Y., Fougere, J.D., Clarke, K., 2012. Effects of SPORL and dilute acid pretreatment on substrate morphology, cell physical and chemical wall structures, and subsequent enzymatic hydrolysis of lodgepole pine. Appl. Biochem. Biotechnol. 168, 1556e1567. Li, Y., Ruan, R., Chen, P.L., Liu, Z., Pan, X., Lin, X., Liu, Y., Mok, C.K., Yang, T., 2004. Enzymatic hydrolysis of corn stover pretreated by combined dilute alkaline treatment and homogenization. Trans. ASAE 47, 821e825. Liao, Z.D., Huang, Z.Q., Hu, H.Y., Zhang, Y.J., Tan, Y.F., 2011. Microscopic structure and properties changes of cassava stillage residue pretreated by mechanical activation. Bioresour. Technol. 102, 7953e7958. Liu, Y., Wang, G.S., Xu, J.L., Zhang, Y., Liu, C.F., Yuan, Z.H., 2011. Effect of sulfite pretreatment to overcome the recalcitrance of lignin (SPORL) on enzymatic saccharification of corn stalk. Bioresources 6, 5001e5011. Lu, K.-M., Lee, W.-J., Chen, W.-H., Liu, S.-H., Lin, T.-C., 2012. Torrefaction and low temperature carbonization of oil palm fiber and eucalyptus in nitrogen and air atmospheres. Bioresour. Technol. 123, 98e105. Ma, F.Y., Yang, N., Xu, C.Y., Yu, H.B., Wu, J.G., Zhang, X.Y., 2010. Combination of biological pretreatment with mild acid pretreatment for enzymatic hydrolysis and ethanol production from water hyacinth. Bioresour. Technol. 101, 9600e9604. Mabee, W.E., Gregg, D.J., Arato, C., Berlin, A., Bura, R., Gilkes, N., Mirochnik, O., Pan, X.J., Pye, E.K., Saddler, J.N., 2006. Updates on softwood-to-ethanol process development. Appl. Biochem. Biotechnol. 129-132, 55e70. Mahalaxmi, S., Jackson, C.R., Williford, C., Burandt, C.L., 2010. Estimation of treatment time for microbial preprocessing of biomass. Appl. Biochem. Biotechnol. 162, 1414e1422. Martin, C., Thomsen, A.B., 2007. Wet oxidation pretreatment of lignocellulosic residues of sugarcane, rice, cassava and peanuts for ethanol production. J. Chem. Technol. Biotechnol. 82, 174e181. Medic, D., Darr, M., Shah, A., Potter, B., Zimmerman, J., 2012. Effects of torrefaction process parameters on biomass feedstock upgrading. Fuel 91, 147e154. Miller, S., Hester, R., 2007. Concentrated acid conversion of pine sawdust to sugars. Part II: high-temperature batch reactor kinetics of pretreated pine sawdust. Chem. Eng. Commun. 194, 103e116. Monauari, S., Galbe, M., Zacchi, G., 2011. Influence of impregnation with lactic acid on sugar yields from steam pretreatment of sugarcane bagasse and spruce, for bioethanol production. Biomass Bioenergy 35, 3115e3122. Mosier, N., Wyman, C., Dale, B., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., 2005. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96, 673e686. Nguyen, T.A.D., Kim, K.R., Han, S.J., Cho, H.Y., Kim, J.W., Park, S.M., Park, J.C., Sim, S.J., 2010. Pretreatment of rice straw with ammonia and ionic liquid for lignocellulose conversion to fermentable sugars. Bioresour. Technol. 101, 7432e7438. Patel, D.D., Lee, J.-M., 2012. Applications of ionic liquids. Chem. Rec. 12, 329e355. Pedersen, M., Vikso-Nielsen, A., Meyer, A.S., 2010. Monosaccharide yields and lignin removal from wheat straw in response to catalyst type and pH during mild thermal pretreatment. Process Biochem. 2010, 1181e1186. Piccolo, C., Bezzo, F., 2009. A techno-economic comparison between two technologies for bioethanol production from lignocellulose. Biomass Bioenergy 33, 478e491. Pinto, P.A., Dias, A.A., Fraga, I., Marques, G., Rodrigues, M.A.M., Colaco, J., Sampaio, A., Bezerra, R.M.F., 2012. Influence of ligninolytic enzymes on straw saccharification during fungal pretreatment. Bioresour. Technol. 111, 261e267. Protasio, T.d.P., Bufalino, L., Mendes, R.F., Ribeiro, M.X., Trugilho, P.F., Leite, E.R.d.S., 2012. Torrefaction and carbonization of briquettes made with residues from coffee grain. Rev. Bras. Eng. Agric. Ambient 16, 1252e1258. Qi, B.K., Chen, X.R., Wan, Y.H., 2010. Pretreatment of wheat straw by nonionic surfactant-assisted dilute acid for enhancing enzymatic hydrolysis and ethanol production. Bioresour. Technol. 101, 4875e4883. Qing, Q., Yang, B., Wyman, C.E., 2010. Impact of surfactants on pretreatment of corn stover. Bioresour. Technol. 101, 5941e5951. Requejo, A., Peleteiro, S., Rodriguez, A., Garrote, G., Parajo, J.C., 2011. Second-generation bioethanol from residual woody biomass. Energy Fuels 25, 4803e4810. Rodrigues, T.H.S., Rocha, M.V.P., de Macedo, G.R., Goncalves, L.R.B., 2011. Ethanol production from cashew apple bagasse: improvement of enzymatic hydrolysis by microwave-assisted alkali pretreatment. Appl. Biochem. Biotechnol. 164, 929e943. Romani, A., Garrote, G., Lopez, F., Parajo, J.C., 2011. Eucalyptus globulus wood fractionation by autohydrolysis and organosolv delignification. Bioresour. Technol. 102, 5896e5904. Roos, A.A., Persson, T., Krawczyk, H., Zacchi, G., Stalbrand, H., 2009. Extraction of water-soluble hemicelluloses from barley husks. Bioresour. Technol. 100, 763e769. Salvachua, D., Prieto, A., Lopez-Abelairas, M., Lu-Chau, T., Martinez, A.T., Martinez, M.J., 2011. Fungal pretreatment: an alternative in second-generation ethanol from wheat straw. Bioresour. Technol. 102, 7500e7506. Samuel, R., Foston, M., Jaing, N., Cao, S.L., Allison, L., Studer, M., Wyman, C., Ragauskas, A.J., 2011. HSQC (heteronuclear single quantum coherence) (13)C-(1)H correlation spectra of whole biomass in perdeuterated pyridinium chloride-DMSO system: an effective tool for evaluating pretreatment. Fuel 90, 2836e2842. Saritha, M., Lata, A.A., 2011. Biological pretreatment of lignocellulosic substrates for enhanced delignification and enzymatic digestibility. Indian J. Microbiol. Sasmal, S., Goud, V.V., Mohanty, K., 2011. Optimisation of the acid catalysed pretreatment of areca nut husk fibre using the Taguchi design method. Biosyst. Eng. 110, 465e472. Sindhu, R., Kuttiraja, M., Binod, P., Preeti, V.E., Sandhya, S.V., Vani, S., Sukumaran, R.K., Pandey, A., 2012. Surfactant-assisted acid pretreatment of sugarcane tops for bioethanol production. Appl. Biochem. Biotechnol. 167, 1513e1526. Sousa, L.D., Chundawat, S.P.S., Balan, V., Dale, B.E., 2009. ’Cradleto-grave’ assessment of existing lignocellulose pretreatment technologies. Curr. Opin. Biotechnol. 20, 339e347. Srinivasan, V., Adhikari, S., Chattanathan, S.A., Park, S., 2012. Catalytic pyrolysis of torrefied biomass for hydrocarbons production. Energy Fuels 26, 7347e7353. Sun, R.C., Sun, X.F., Ma, X.H., 2002. Effect of ultrasound on the structural and physiochemical properties of organosolv soluble hemicelluloses from wheat straw. Ultrason. Sonochem. 9, 95e101.
REFERENCES Sun, Y.S., Lu, X.B., Zhang, R., Wang, X.Y., Zhang, S.T., 2011. Pretreatment of corn stover silage with Fe(NO(3))(3) for fermentable sugar production. Appl. Biochem. Biotechnol. 164, 918e928. Tadesse, H., Luque, R., 2011. Advances on biomass pretreatment using ionic liquids: an overview. Energy Environ. Sci. 4, 3913e3929. Tao, L., Aden, A., Elander, R.T., Pallapolu, V.R., Lee, Y.Y., Garlock, R.J., Balan, V., Dale, B.E., Kim, Y., Mosier, N.S., Ladisch, M.R., Falls, M., Holtzapple, M.T., Sierra, R., Shi, J., Ebrik, M.A., Redmond, T., Yang, B., Wyman, C.E., Hames, B., Thomas, S., Warner, R.E., 2011. Process and technoeconomic analysis of leading pretreatment technologies for lignocellulosic ethanol production using switchgrass. Bioresour. Technol. 102, 11105e11114. Tapasvi, D., Khalil, R., Skreiberg, O., Khanh-Quang, T., Gronli, M., 2012. Torrefaction of Norwegian birch and spruce: an experimental study using macro-TGA. Energy Fuels 26, 5232e5240. Teixeira, L.C., Linden, J.C., Schroeder, H.A., 2000. Simultaneous saccharification and cofermentation of peracetic acid-pretreated biomass. Appl. Biochem. Biotechnol. 84-6, 111e127. Teramoto, Y., Tanaka, N., Lee, S.H., Endo, T., 2008. Pretreatment of eucalyptus wood chips for enzymatic saccharification using combined sulfuric acid-free ethanol cooking and ball milling. Biotechnol. Bioeng. 99, 75e85. Thomsen, M.H., Holm-Nielsen, J.B., Oleskowicz-Popiel, P., Thomsen, A.B., 2008. Pretreatment of whole-crop harvested, ensiled maize for ethanol production. Appl. Biochem. Biotechnol. 148, 23e33. Tian, X.-f., Fang, Z., Guo, F., 2012. Impact and prospective of fungal pre-treatment of lignocellulosic biomass for enzymatic hydrolysis. Biofuels, Bioprod. Biorefin. 6, 335e350. Tumuluru, J.S., Boardman, R.D., Wright, C.T., Hess, J.R., 2012. Some chemical compositional changes in Miscanthus and white oak sawdust samples during torrefaction. Energies 5, 3928e3947. Um, B.H., Bae, S.H., 2011. Statistical methodology for optimizing the dilute acid hydrolysis of sugarcane bagasse. Korean J. Chem. Eng. 28, 1172e1176. Vancov, T., McIntosh, S., 2011. Effects of dilute acid pretreatment on enzyme saccharification of wheat stubble. J. Chem. Technol. Biotechnol. 86, 818e825. Velmurugan, R., Muthukumar, K., 2012a. Sono-assisted enzymatic saccharification of sugarcane bagasse for bioethanol production. Biochem. Eng. J. 63, 1e9. Velmurugan, R., Muthukumar, K., 2012b. Ultrasound-assisted alkaline pretreatment of sugarcane bagasse for fermentable sugar production: optimization through response surface methodology. Bioresour. Technol. 112, 293e299. Verma, P., Watanabe, T., Honda, Y., 2011. Microwave-assisted pretreatment of woody biomass with ammonium molybdate activated by H2O2. Bioresour. Technol. 102, 3941e3945. Wan, C., Li, Y., 2011. Effectiveness of microbial pretreatment by Ceriporiopsis subvermispora on different biomass feedstocks. Bioresour. Technol. 102, 7507e7512. Wang, B., Wang, X.J., Feng, H., 2010. Deconstructing recalcitrant Miscanthus with alkaline peroxide and electrolyzed water. Bioresour. Technol. 101, 752e760. Wang, W., Yuan, T., Wang, K., Cui, B., Dai, Y., 2012. Combination of biological pretreatment with liquid hot water pretreatment to enhance enzymatic hydrolysis of Populus tomentosa. Bioresour. Technol. 107, 282e286. Wang, Y., Radosevich, M., Hayes, D., Labbe, N., 2011a. Compatible ionic liquid-cellulases system for hydrolysis of lignocellulosic biomass. Biotechnol. Bioeng. 108, 1042e1048. Wang, Z.Y., Xu, J.L., Cheng, J.J., 2011b. Modeling biochemical conversion of lignocellulosic materials for sugar production: a review. Bioresources 6, 5282e5306. 69 Wu, Y., Zhang, C., Liu, Y., Fu, Z., Dai, B., Yin, D., 2012. Biomass char sulfonic acids (BC-SO3H)-catalyzed hydrolysis of bamboo under microwave irradiation. Bioresources 7, 5950e5959. Wyman, C.E., Dale, B., Richard, T.E., Holtzapple, M., Ladisch, M., Lee, Y.Y., 2005. Coordinated development of leading biomass pretreatment technologies. Bioresour. Technol. 96, 1959e1966. Yan, W., Acharjee, T.C., Coronella, C.J., Vasquez, V.R., 2009. Thermal pretreatment of lignocellulosic biomass. Environ. Prog. Sustainable Energy 28, 435e440. Yanagida, T., Fujimoto, S., Bespyatko, L., Tsukahara, K., Hideno, A., Sawayama, S., Minowa, T., 2011. Introduction of dehydration process into mechanochemical pretreatment for bioethanol production. J. Jpn. Pet. Inst. 54, 215e221. Yang, B., Wyman, C.E., 2008. Pretreatment: the key to unlocking lowcost cellulosic ethanol. Biofuels, Bioprod. Biorefin. 2, 26e40. Yang, X.W., Ma, F.Y., Yu, H.B., Zhang, X.Y., Chen, S.L., 2011. Effects of biopretreatment of corn stover with white-rot fungus on lowtemperature pyrolysis products. Bioresour. Technol. 102, 3498e3503. Yoo, J., Alavi, S., Vadlani, P., Amanor-Boadu, V., 2011. Thermo-mechanical extrusion pretreatment for conversion of soybean hulls to fermentable sugars. Bioresour. Technol. 102, 7583e7590. Yu, J., Zhang, J.B., He, J., Liu, Z.D., Yu, Z.N., 2009. Combinations of mild physical or chemical pretreatment with biological pretreatment for enzymatic hydrolysis of rice hull. Bioresour. Technol. 100, 903e908. Zeng, J.J., Singh, D., Chen, S.L., 2011. Thermal decomposition kinetics of wheat straw treated by Phanerochaete chrysosporium. Int. Biodeterior. Biodegrad. 65, 410e414. Zhang, C., Zhu, J.Y., Gleisner, R., Sessions, J., 2012a. Fractionation of forest residues of Douglas-fir for fermentable sugar production by SPORL pretreatment. Bioenergy Res. 5, 978e988. Zhang, R., Lu, X.B., Sun, Y.S., Wang, X.Y., Zhang, S.T., 2011. Modeling and optimization of dilute nitric acid hydrolysis on corn stover. J. Chem. Technol. Biotechnol. 86, 306e314. Zhang, S., Keshwani, D.R., Xu, Y., Hanna, M.A., 2012b. Alkali combined extrusion pretreatment of corn stover to enhance enzyme saccharification. Ind. Crops Prod. 37, 352e357. Zhang, Y.H.P., Ding, S.Y., Mielenz, J.R., Cui, J.B., Elander, R.T., Laser, M., Himmel, M.E., McMillan, J.R., Lynd, L.R., 2007. Fractionating recalcitrant lignocellulose at modest reaction conditions. Biotechnol. Bioeng. 97, 214e223. Zhao, J., Zhang, H.M., Zheng, R.P., Lin, Z.X., Huang, H., 2011a. The enhancement of pretreatment and enzymatic hydrolysis of corn stover by FeSO(4) pretreatment. Biochem. Eng. J. 56, 158e164. Zhao, X.B., Cheng, K.K., Liu, D.H., 2009. Organosolv pretreatment of lignocellulosic biomass for enzymatic hydrolysis. Appl. Microbiol. Biotechnol. 82, 815e827. Zhao, X.B., Wu, R.C., Liu, D.H., 2011b. Production of pulp, ethanol and lignin from sugarcane bagasse by alkali-peracetic acid delignification. Biomass Bioenergy 35, 2874e2882. Zheng, Y., Pan, Z., Zhang, R., 2009. Overview of biomass pretreatment for cellulosic ethanol production. Int. J. Agric. Biol. Eng. 2, 51e58. Zhong, W.X., Yu, H.B., Song, L.L., Zhang, X.Y., 2011. Combined pretreatment with white-rot fungus and alkali at near roomtemperature for improving saccharification of corn stalks. Bioresources 6, 3440e3451. Zhu, J.Y., Pan, X.J., Wang, G.S., Gleisner, R., 2009. Sulfite pretreatment (SPORL) for robust enzymatic saccharification of spruce and red pine. Bioresour. Technol. 100, 2411e2418. Zimbardi, F., Viola, E., Nanna, F., Larocca, E., Cardinale, M., Barisano, D., 2007. Acid impregnation and steam explosion of corn stover in batch processes. Ind. Crops Prod. 26, 195e206.
C H A P T E R 5 Biofuels and Bioproducts Produced through Microbial Conversion of Biomass Trent Chunzhong Yang 1, Jyothi Kumaran 2,3, Samuel Amartey 4, Miranda Maki 5, Xiangling Li 1,6, Fan Lu 7, Wensheng Qin 5,* 1 Aquatic and Crop Resource Development, National Research Council Canada, Ottawa, ON, Canada, 2Human Health Therapeutics, National Research Council Canada, Ottawa, ON, Canada, 3School of Environmental Sciences, University of Guelph, Guelph, ON, Canada, 4Division of Biology, Imperial College of Science, Technology and Medicine, South Kensington, London, UK, 5Department of Biology, Lakehead University, ON, Canada, 6 College of Chinese Medicine, Guangzhou University of Chinese Medicine, Guangzhou, China, 7 College of Bioengineering, Hubei University of Technology, Wuhan, Hubei Province, China *Corresponding author email: wqin@Lakeheadu.ca O U T L I N E Lignocellulosic Biomass and its Pretreatment Nonbiological Pretreatment Physical Pretreatments Chemical Pretreatments Physicochemical Pretreatments Biological Pretreatment with Microorganisms Potential Advantages over Nonbiological Pretreatment Biological Degradation of Lignin Commonly used Microorganisms for Biological Pretreatment Natural Microorganisms and Practical Applications in Bioconversion Application of White-Rot Fungus in Treatment of Different Biomasses White-Rot Fungus Pretreatment of Biomass for Animal Feed White-Rot Fungus Pretreatment in Biological Pulping White-Rot Fungus Pretreatment of Biomass for Biofiber Brown-Rot Fungi Soft-Rot Fungi Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00005-X 72 72 72 72 72 73 73 73 73 74 74 75 75 75 75 76 Bacteria Genetically Modified Microorganisms for Biomass Conversion Rational Engineering Metabolic Engineering of Microbial Pathways for Enhanced Bioproduct Production Strategies of Using Microbial Pretreatment to Enhance Sugar Release for Biofuel and Bioproduct Production Application of Microbial Pretreatment for Biogas Production Application of Microbial Pretreatment for Biomass Conversion Strategies for Microorganism Application in Biomass Commonly Used Microorganisms in Biomass Conversion and Some Application Examples Other Bioproducts Produced by Microbial Conversion of Biomass: Introduction References 77 77 77 78 79 80 81 81 82 84 87 71 Copyright Ó 2014 Elsevier B.V. All rights reserved.
72 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS LIGNOCELLULOSIC BIOMASS AND ITS PRETREATMENT Lignocellulose is the primary building block of plant cell walls and is composed mainly of cellulose, hemicelluloses, lignin and small quantities of pectin, proteins, extractives and ash. The cellulose, hemicelluloses and lignin are present in varying amounts in the different parts of the plant and are intimately associated to form the complex structural framework of the plant cell wall where cellulose and hemicellulose are bound together with lignin and other components to form a tight matrix. The composition of lignocellulose depends on plant species as well as growth conditions and age. Lignocellulose biomass is a renewable, sustainable, abundant and cheap resource for producing renewable biofuels and bioproducts. However, their conversion into fermentable sugar before fermentation is a major hurdle due to its complex structure and recalcitrant nature. While hydrolysis of cellulose and hemicellulose yields fermentable sugars, they are not easily accessible due to the crystalline structure of cellulose and interference by the phenyl-propanoid polymer, lignin. Bioconversion of carbohydrates from lignocellulosic feedstocks into fermentable sugars is a key challenge in the biorefinery process. Efficient, cost-effective and environmentally benign pretreatment and hydrolysis methods are required. The primary purpose of pretreatment is to change the architecture of the cell wall by delignification and disrupting the cellulose structure and making the lignocellulosic biomass accessible and reactive to allow high rates and yields on enzymatic hydrolysis. Pretreatment has been considered as one of the most expensive processing steps in biomass to fermentable sugar conversion (Mosier et al., 2005). This article focuses mainly on biological conversion of biomass with microorganisms. However, nonbiological pretreatments, as well as the most frequently studied and applied procedures, will also be discussed. Nonbiological Pretreatment A variety of nonbiological pretreatment methods have been extensively reviewed. These include physical, chemical, physicochemical and other combinations of procedures (Alvira et al., 2010; Chandra et al., 2007; da Costa Sousa et al., 2009; Hendriks and Zeeman, 2009; Sun and Cheng, 2002; Taherzadeh and Karimi, 2008). Based on their effects on biomass structure, pretreatments can be divided into different categories: those that increase enzyme accessibility to crystalline cellulose by decreasing the fiber’s degree of polymerization or by facilitating hemicellulose and/or lignin removal to create pores in the cellulose fibrils. Since hemicellulose and lignin are the two main protective coats surrounding cellulose, they have to be removed or altered in order to achieve fast enzymatic hydrolysis of the biomass. However, to obtain high sugar yield for both hexoses and pentoses, an ideal pretreatment procedure should efficiently remove or modify lignin and also hydrolyze hemicellulose, but not degrade these hemicellulose sugars (Ohgren et al., 2007). Some of the most widely investigated procedures are briefly described. Physical Pretreatments These include mechanical methods to chip, grind and mill the biomass to reduce particle size and, potentially, the crystallinity and degree of polymerization of lignocellulose in order to maximize the downstream enzyme hydrolysis process (Tassinari et al., 1980). Recently, a novel extrusion method was developed where the biomass materials are subjected to heating, mixing and shearing to cause both physical and chemical modifications to the material in order to increase cellulose accessibility (Karunanithy and Muthukumarappan, 2010a,b; Karunanithy et al., 2012). Chemical Pretreatments These are mainly alkali and acid pretreatments. Alkali pretreatments increase cellulose digestibility by enhancing lignin solubilization and decreasing cellulose crystallinity. This method is more effective on agricultural biomass than on wood material (Kumar et al., 2009; Playne, 1984). Acid pretreatment, mostly diluted acid pretreatments, increase cellulose accessibility mainly by solubilizing hemicellulose. It can be used as either a pretreatment or a direct hydrolysis process but leads to toxic degradation products that inhibit downstream fermentation (Alvira et al., 2010). On the contrary, ozonolysis uses the powerful oxidant ozone to delignify lignocellulosic materials at room temperature and does not form inhibitory compounds, yet it is economically unviable due to large amounts of ozone consumed (Sun and Cheng, 2002). On the other hand, organosolv process can efficiently remove lignin and result in minimal cellulose loss. This is a promising process if economic solvents are available at commercial scales (Wood and Saddler, 1988; Zhao et al., 2009). Physicochemical Pretreatments Steam explosion is the most studied and commonly used physicochemical method and extensively reviewed (Hsu, 1996; McMillan, 1994; Saddler et al., 1993). During this hydrothermal procedure, biomass is subjected to pressurized steam for a short time and then suddenly depressurized. The process leads to hemicellulose degradation and lignin transformation and as a result, increases pore volumes in the pretreated biomass, leading to enhanced enzymatic accessibility (Grous et al., 1986). It is recognized as one of the most cost-effective
COMMONLY USED MICROORGANISMS FOR BIOLOGICAL PRETREATMENT processes for hardwoods and agricultural residues, but less effective for softwoods (Sun and Cheng, 2002). Another disadvantage is the production of inhibitory compounds. Addition of diluted acids can decrease pretreatment time and temperature thus reducing the production of inhibitory compounds and also enhancing softwood pretreatment efficiency (Ballesteros et al., 2006; Duff and Murray, 1996; Jørgensen, 2007; Kumar et al., 2009; Stenberg et al., 1998). As a relatively energy and environmentally friendly procedure, steam explosion had been scaled up and used in pilot-scale production at Iogen (Canada) and is to be used in many of the planned commercial size facilities worldwide. Other physicochemical methods explored include ammonia fiber explosion (AFEX) (Alizadeh et al., 2005; Teymouri et al., 2004, 2005), carbon dioxide explosion (Zheng et al., 1995, 1998), liquid hot water (LHW) pretreatment (Kim et al., 2009; Mosier et al., 2005), ultrasound pretreatment (Gonzalez-Fernandez et al., 2012; Sasmal et al., 2012), and microwave pretreatment (Azuma et al., 1984; Ma et al., 2009; Ooshima et al., 1984). For practical application, different pretreatment methods have to be tested for each specific biomass to determine the best procedure that is compatible with the downstream hydrolytic enzyme cocktail. For example, in a recent report describing switchgrass hydrolysis, different pretreatment methods were tested including ammonia fiber expansion (AFEX), dilute acid (DA), LHW, lime, lime þ ball milling, soaking in aqueous ammonia, and sulfur dioxide (SO2). It was demonstrated that lime þ ball milling lead to the highest overall sugar yield (98.3%) from pretreated biomass with xylanase addition (Falls et al., 2011). Biological Pretreatment with Microorganisms Potential Advantages over Nonbiological Pretreatment Microbial pretreatment by solid state cultivation (SSC) has the potential to be a low-cost, environmentally friendly alternative to chemical approaches. Existing nonbiological pretreatment methods as described above have largely been developed on the basis of physicochemical technologies such as steam explosion, microwave radiation, ionizing radiation, dilute acid, alkali, and oxidation or various combinations of these methodologies (Mosier et al., 2005). Most of these methods require expensive, complicated, high-pressure and corrosion-resistant equipment and may consume large amounts of energy and water. Furthermore, chemical pretreatments can be detrimental to subsequent enzymatic hydrolysis and microbial fermentation in addition to producing acidic or alkaline waste water, which requires predisposal treatment to ensure environmental safety (Keller et al., 2003). Due to its low energy and 73 material costs, mild reaction conditions with simple equipment, and environmental benefits, microbial/biological pretreatment has received increased attention as an alternative to physicochemical or thermochemical pretreatments (Kumar and Wyman, 2009; Rabinovich et al., 2004; Sanchez, 2009; Saritha et al., 2012a; Shi et al., 2008; Sun and Cheng, 2002; Zeng et al., 2011). Biological Degradation of Lignin Lignin is a complex, heterogeneous phenylpropanoid polymer that is linked to both hemicelluloses and cellulose to form an impenetrable physical and chemical barrier for biodegradative systems (Sanchez, 2009; Blanchette, 1991). Unless lignin is modified or removed, hydrolytic enzymes cannot penetrate and effectively degrade woody substrates. In addition to producing the extracellular polysaccharide degradative enzymes, such as cellulases, xylanases, and mannanases, saprophytic fungi have a unique oxidative and extracellular lignolytic system called Fenton’s reagents to degrade lignin and open phenyl rings (Green and Highley, 1997; Jensen et al., 2001; Arantes et al., 2012; Contreras et al., 2007; Irbe et al., 2011; Kramer et al., 2004; Ray et al., 2010; Suzuki et al., 2006; Yanase et al., 2010b). In addition to cellulase and hemicellulases, lignolytic enzymes have also been detected in some strains. Particularly, species among the Basidiomycotina fungi that cause white rots of wood may simultaneously degrade lignin and cell wall carbohydrates (Sanchez, 2009). Furthermore, a small number of the white-rot fungi preferentially degrade lignin leading to little to no loss of cellulose (Blanchette, 1991). For practical applications, these species that can selectively remove lignin without extensive cellulose degradation are of special interest. The most widely studied white-rot fungus, Phanerochaete chrysosporium, can significantly degrade lignin and simultaneously degrade a small fraction of cellulose and hemicellulose, whereas others such as Ceriporiopsis subvermispora tend to remove lignin in advance of cellulose and hemicellulose (Blanchette et al., 1992; Hatakka, 1994; Sanchez, 2009). COMMONLY USED MICROORGANISMS FOR BIOLOGICAL PRETREATMENT Microbial pretreatment makes use of microorganisms and their enzyme systems to breakdown lignin and/or hemicellulose present in lignocellulosic biomass. So far, the isolated and identified lignocellulolytic microorganisms mainly include fungi and a few bacterial strains. Fungi including brown-, white-, and soft-rot fungi are the predominant organisms responsible for lignocellulose degradation, and among the fungi, the Basidiomycetes that cause both white and brown rots
74 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS are the most rapid degraders (Bennet et al., 2002; Loguercio-Leite et al., 2008; Rabinovich et al., 2004; Sanchez, 2009; ten Have and Teunissen, 2001). Several Basidiomycetes such as P. chrysosporium, C. subvermispora, Phlebia subserialis, Pleurotus ostreatus, and Irpex lacteus have been shown to efficiently degrade lignin in different lignocellulosic materials (Hatakka and Usi-Rauva, 1983; Keller et al., 2003; Sawada et al., 1995; Taniguchi et al., 2005; Zeng et al., 2011). Natural Microorganisms and Practical Applications in Bioconversion Application of White-Rot Fungus in Treatment of Different Biomasses CORN STOVER When corn stover is pretreated with C. subvermispora for downstream bioethanol production, lignin is selectively degraded up to 31.59% with a limited cellulose loss of less than 6% during an 18-day pretreatment. Longer pretreatment time was found to increase lignin removal, resulting in correspondingly higher glucose yields from enzymatic hydrolysis. The highest overall ethanol yield of 57.80% was obtained with 35-daypretreated corn stover (Wan and Li, 2010). In a later study, the effectiveness of C. subvermispora pretreatment on different types of feedstocks, including corn stover, wheat straw, soybean straw, switchgrass, and hardwood was tested. After an 18-day pretreatment, corn stover, switchgrass, and hardwood were effectively delignified, leading to a two- to threefold increase in glucose yield over those of the untreated raw materials. In contrast, wheat straw and soybean straw did not show glucose yield increase after undergoing the same pretreatment, suggesting the importance of using a specific strain for pretreatment of specific biomass (Wan and Li, 2011). Pretreatments of corn stover with the white-rot fungus I. lacteus CD2 also resulted in significant lignin degradation with limited cellulose loss (Zeng et al., 2011). Pretreatment of corn stover with Cyathus stercoreus led to a three- to fivefold improvement in enzymatic cellulose digestibility (Keller et al., 2003). Pretreatment of corn stover with a newly isolated white-rot fungus, Trametes hirsuta yj9, led to selective lignin degradation up to 71.49% and a significant increase in enzymatic digestibility of 73.99% after a 42-day pretreatment (Sun et al., 2011). Pretreatment of corn stover fractions (leaves, cobs, and stalks) with the white-rot fungus C. subvermispora showed that the leaves were the least recalcitrant to fungal pretreatment with a 45% lignin degradation as well as higher carbohydrate degradation after 30 days of pretreatment. However, corn cobs produced the highest sugar yield after fungal pretreatment (Cui et al., 2012). SOFTWOOD The effect of pretreatment on the softwood Pinus densiflora by three white-rot fungi, Ceriporia lacerata, Stereum hirsutum, and Polyporus brumalis, has been investigated. Among the three white-rot fungi tested, S. hirsutum selectively degraded the lignin rather than the holocellulose component. Consistently, extracellular enzymes from S. hirsutum showed higher activity of ligninase and lower activity of cellulase than those from the other white-rot fungi. In addition, the available pore size and surface area in the pretreated wood were increased, possibly due to degradation of lignin and a small portion of hemicellulose by the secreted enzymes. Sugar yield of the S. hirsutum pretreated wood also greatly increased compared to a nonpretreated sample, indicating S. hirsutum might be a potentially effective fungus for use in biological pretreatment of woody biomass (Lee et al., 2007). COTTON STALKS Conditions for pretreatment of cotton stalks using P. chrysosporium by SSC have also been explored. While substrate moisture content significantly affects lignin degradation, supplementation with modified salts did not affect the reaction process. Over a period of 14 days, SSCat 75% moisture content without salts resulted in 27.6% lignin degradation, 71.1% solids recovery and 41.6% availability of carbohydrates, suggesting that microbial pretreatment by SSC has the potential to be a low-cost, environmentally friendly alternative to chemical approaches (Shi et al., 2008). RICE STRAW Fungal pretreatment of rice straw for improved enzymatic saccharification has been reported. Yamagishi et al. (2011) tested 17 C. stercoreus isolates for their ability to treat rice straw for improved enzymatic hydrolysis. A negative correlation was found between cellulase and xylanase activity in these isolates and enzymatic saccharification yields in the pretreated straw. A 25-day pretreatment with the strain C. stercoreus TY-2 led to a more than fivefold increase in enzymatic saccharification yield compared to untreated control samples, suggesting this isolate has the potential for biological pretreatment of rice straw under conditions of low energy input. A 15-day pretreatment of rice straw with P. chrysosporium in an optimized media resulted in a treated biomass with an enzymatic digestibility of 64.9% of the theoretical maximum glucose yield. When the fungal-pretreated rice straw was used as a substrate in simultaneous saccharification and fermentation (SSF), a 9.49 g/l ethanol concentration, 58.2% of the theoretical maximum production yield, and 0.40 g/l/h productivity were achieved after 24 h and a 62.7% of the theoretical maximum ethanol yield was expected after 96 h (Bak et al., 2009).
COMMONLY USED MICROORGANISMS FOR BIOLOGICAL PRETREATMENT When rice straw was pretreated with the wood-rot fungus, Dichomitus squalens, for 15 days, an enzymatic digestibility of 58.1% of theoretical glucose yield was reached for the treated biomass. When the pretreated rice straw was used as a substrate for ethanol production in SSF, the ethanol production yield and productivity were 54.2% of the theoretical maximum and 0.39 g/l/h, respectively, after 24 h (Bak et al., 2009). Taniguchia et al. (Taniguchi et al., 2005) reported the effect on rice straw composition and susceptibility to enzymatic hydrolysis after pretreatment with four white-rot fungi (P. chrysosporium, Trametes versicolor, C. subvermispora, and P. ostreatus). Among the four strains, P. ostreatus selectively degraded the lignin fraction of rice straw rather than the cellulose component. A 60-day pretreatment of rice straw with P. ostreatus led to a total weight loss of 25% and 41% lignin degradation, but only a 17% loss of cellulose and a 48% loss of hemicellulose. A 48-h enzymatic hydrolysis lead to 52% holocellulose and 44% cellulose solubilization in the pretreated rice straw corresponding to a net sugar yield of 33% from holocellulose and 32% from cellulose. PADDY STRAW A recent report of a study on the pretreatment of paddy straw with the white-rot fungus T. hirsuta (Microbial Type Culture Collection) MTCC 136 showed high ligninase and low cellulase activities. It showed that within 10 days of solid state fermentation, the carbohydrate content was enhanced by 11.1% and a much higher yield of sugars was obtained after enzymatic hydrolysis. Saccharification efficiency of the biologically pretreated paddy straw with the commercial enzyme AcceleraseÒ 1500 reached 52.69% within 72 h suggesting the delignification potential of T. hirsuta for pretreatment of lignocellulosic substrate and facilitating efficient enzymatic digestibility of cellulose (Saritha et al., 2012b). White-Rot Fungus Pretreatment of Biomass for Animal Feed Pretreatment of lignocellulosic biomass with the white-rot fungi increases biodegradability and leads to high-quality ruminant feed. For example, white-rot fungi-treated cedar wood shows significant improvement for rumen digestibility (Okano et al., 2005). When high-lignin forages such as grass, oat straw and alfalfa stems were treated with various white-rot fungi, substantial improvements in digestibilities have also been obtained (Akin et al., 1995, 1993; Jung et al., 1992). White-Rot Fungus Pretreatment in Biological Pulping White-rot fungi have also been used in biological pulping (biopulping) to reduce the utilization of chemicals in the pulping industry and decrease the environmental hazard caused by the traditional pulping process (Singh 75 et al., 2010). Biopulping process removes not only lignin and hemicellulose but also some of the wood extractives. It can also improve paper quality and significantly reduce the electrical energy and cooking time required for pulping wood chips (Ali and Sreekrishnan, 2001; Hunt et al., 2004; Singh et al., 2010). When C. subvermispora was used for biopulping of agricultural residues including rice, wheat and barley straw samples, the tensile strength and burst factor of hand sheets produced from the biopulping process improved significantly compared to the chemical process (Yaghoubi et al., 2008). Blanchette et al. (Blanchette et al., 1992) evaluated the potential application in biopulping of 19 strains of P. chrysosporium and 9 strains of C. subvermispora. For the P. chrysosporium isolates, only a few strains preferentially removed large amounts of lignin from wood while the majority of the isolates removed all cell wall components nonselectively. In contrast, all nine isolates of C. subvermispora led to moderate weight losses and preferential degradation of lignin in aspen, birch and loblolly pine wood. White-Rot Fungus Pretreatment of Biomass for Biofiber Microbial pretreatment can also improve the feature of the fiber in biomass for biocomposite production. For example, corn stalk pretreated with the white-rot fungus Trametes hirsuta has been used to produce fiberboard by hot pressing without adhesive. The corn stalk-based fiberboard made of the pretreated biomass has an increase of 3.40- and 8.87-fold in moduli of rupture and elasticity, respectively, over the fiberboard made from untreated corn stalk. Further analyses showed that the increase in the mechanical properties of the fiberboard resulted from the pretreated biomass possessing more than twice the number of hydroxyl groups, an 18% higher crystallinity, and twice the polysaccharide content of untreated corn stalk (Wu et al., 2011). Brown-Rot Fungi Brown-rot fungi are Basidiomycete fungi that, unlike white-rot fungi, selectively modify and then completely hydrolyze lignocellulose polysaccharides, typically without secreting an exoacting glucanase and without removing lignin (Schilling et al., 2009; Tewalt and Schilling, 2010). The wood decay resulting from the action of brown-rot fungi leads to an increased volume of pores in the wood cell wall and decreased degree of polymerization of holocellulose along with a dramatic weight loss (Flournoy et al., 1991). Depolymerization of holocellulose occurs rapidly during the early decay process leading to an extensive degradation of holocellulose in wood (Blanchette, 1995; Irbe et al., 2011; Kumar et al., 2009) and as high as 75% wood strength loss even when only 1% weight loss has occurred (Green and Highley, 1997; Richards, 1954; Wilcox, 1978).
76 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS The exact mechanism for brown-rot decay is still unclear. For the selective removal of polysaccharides, a two-step procedure has been proposed: a nonenzymatic radical-based modification of the wood cell wall through small molecules, followed by secretion of enzymes to catalyze the breakdown of polysaccharides into their sugar monomers (Green and Highley, 1997; Tewalt and Schilling, 2010). However, cellulose and hemicellulose removal by brown-rot fungi does not open up cell walls to facilitate enzyme penetration (Flournoy et al., 1991). Primarily because enzymes are too large to penetrate the decayed wood, attack by cellulolytic enzymes may only be limited to a localized, superficial area (Baldrian and Valaskova, 2008; Flournoy et al., 1991). It has been proposed that Fenton’s reagents and not enzymes are responsible for rapid wood decomposition early in brown-rot decay (Green and Highley, 1997; Jensen et al., 2001; Ray et al., 2010). Other study results also support that hydroxyl radicals (HO_) generated through Fenton chemistry (H2O2eFe(II)) initiate lignocellulose breakdown (Arantes et al., 2012; Contreras et al., 2007; Hammel et al., 2002; Kaneko et al., 2005; Kramer et al., 2004; Suzuki et al., 2006). Consequently, this suggests that reactive oxygen species play an important role in the early stages of wood degradation by brown-rot fungi (Irbe et al., 2011). In brown-rot wood decay, hemicellulose is removed considerably faster than cellulose (Curling et al., 2002; Highley, 1987; Monrroy et al., 2011). Consistently, the total secretome hemicellulase expression and activity for brown-rot fungi peak prior to cellulase activity (Lyr, 1960; Martinez et al., 2009). Hemicellulose is embedded in cellulose microfibrils and its prior removal may facilitate cellulose degradation and removal (Green and Highley, 1997). Continual degradation of holocellulose by brown-rot fungi leads to gradually increased weight loss but the percent crystallinity in decayed wood increases apparently at an early stage, peaks between 2 and 4 weeks and then decreases implying structural changes of cellulose chains during fungal attack (Howell et al., 2009). Towards the end of brown-rot decay, nearly 100% of carbohydrates can be removed; however, most of the lignin remains (Eriksson et al., 1990). Only a small fraction of the lignin is oxidized, demethylated and depolymerized, often leading to lignin-derived volatile components (Ewen et al., 2004; Irbe et al., 2011; Schilling et al., 2012). Recently, the potential application of brown-rot fungi for the pretreatment of biomass to increase downstream enzymatic hydrolysis has been explored. When spruce and pine woods were treated with one of two brownrot fungi, Gloeophyllum trabeum or Fomitopsis pinicola, saccharification efficiency was increased significantly even though total sugar yield was low, probably due to low enzyme loading (Schilling et al., 2009). In another effort, G. trabeum-treated pine wood block only led to a maximum 22% glucose release even though 60 FPU Celluclast was loaded, suggesting brown-rot fungus G. trabeum modification of pine wood may not be sufficient to increase cellulose accessibility (Tewalt and Schilling, 2010). Similarly, when the brown-rot fungi G. trabeum and Laetoporeus sulphureus were used for the pretreatment of the wood Pinus radiate and Eucalyptus globules, the highest glucose yield was 14% after 8 weeks of biodegradation (Monrroy et al., 2011). On the other hand, when G. trabeum was used to pretreat different biomass including aspen, spruce, or corn stover, sugar yield was significantly increased up to threefold. In the best case, a 2-week pretreatment of aspen by G. trabeum led to a 72% cellulose-to-glucose yield corresponding to 51% yield relative to original glucan. For corn stover, a weak colonization with minor degradation by another tested brown-rot fungus, Postia placenta, resulted in more than a twofold increase in sugar yield (Schilling et al., 2012). Similar to wood biomass, when corn stover is pretreated with the brown-rot fungus Fomitopsis sp. IMER2, the amorphous regions of the cellulose are preferentially degraded in contrast to the significant lignin degradation by the white-rot fungus I. lacteus CD2 (Zeng et al., 2011). In another successful case, simple pretreatment of Scots pine (Pinus sylvestris) with the brown rot fungus Coniophora puteana for 15 days permitted recovery of greater than 70% of the glucose present in the biomass, with a total wood mass loss of 9%, suggesting great potential for use of this specific group of fungi in lignocellulosic biomass pretreatment (Ray et al., 2010). Brown-rot fungi therefore hold significant potential for practical application in biological pretreatment. Soft-Rot Fungi Even though the process of wood decay by many common white- and rot fungi has been well characterized, other types of decay caused by soft-rot fungi or bacteria are still not well understood (Blanchette et al., 2002, 2004). Soft rot is caused by fungi taxonomically classified in the phylum Ascomycota, including related asexual taxa. The term soft rot is used because it was first identified from soft, decayed wood surfaces in contact with excessive moisture (Findlay, 1984). Soft rot can also occur in dry environments (Blanchette, 2000) and seems to predominate in extreme environments such as excessively wet or dry sites, where white- and brown-rot fungi growth is inhibited, and in substrates that do not favor the growth and development of other types of fungi (Blanchette, 1995; Blanchette et al., 2004). Soft-rot fungi attack the lignocellulose matrix in wood by formation of cavities (type I) or cell wall erosion (type II). Cellulases and hemicellulases, but not ligninases, are involved in soft-rot attack leading to extensive loss of the carbohydrate polymers; high amounts of lignin remain even in advanced stages of
COMMONLY USED MICROORGANISMS FOR BIOLOGICAL PRETREATMENT soft rot (Blanchette, 1995; Eriksson et al., 1990; Nilsson et al, 1989). The most studied and applied soft-rot fungus, Trichoderma reesei, and its mutants, are mainly used for large-scale commercial production of cellulases and hemicellulases (Durand et al., 1988; Esterbauer et al., 1991; Tomme et al., 1988). Bacteria Bacteria degrade plant cell walls through three main morphological forms: tunneling, erosion, and cavitation (Blanchette, 1995; Daniel et al., 1987; Singh and Butcher, 1991, 1985; Singh et al., 1990). An early study has confirmed that the Gram-positive filamentous bacterium Streptomyces viridosporus degrades softwood lignin into low molecular weight fragments (Crawford et al., 1982). Furthermore, enzymes similar to the fungal system such as peroxidases, ligninases and manganese peroxidases have been implicated in bacterial biomass delignification (Glenn and Gold, 1983; Kirk et al., 1986). Interestingly, some bacteria can attack high lignin-containing hard wood that is considered durable and resistant to fungal decay (Nilsson et al., 1992; Singh and Butcher, 1991). However, compared to fungi, bacteria are not as efficient for lignocellulosic biomass pretreatment, as shown by a recent work comparing eight microorganisms including fungi and bacteria, for pretreatment of sugarcane waste (Singh et al., 2008). Genetically Modified Microorganisms for Biomass Conversion Since the 1990s, bacteria, fungi and yeasts have been genetically engineered for the industrial production of biofuels and bioproducts. More conventionally, the improvement of microorganisms for biomass conversion has been done using classical chemical mutagenesis, a random approach followed by the screening and selection of a desired trait. Nevertheless, with advancements in molecular biology and biotechnology approaches, the improvement of microorganisms via rational engineering of proteins and metabolic engineering of pathways has become more prevalent (Strohl, 2001). This is due to the economic needs of the industry, which demands the development of strains that produce greater yields and a different variety of products. Specifically, in the bioconversion of biomass, researchers face challenges related to the substrate such as appropriate enzymes for conversion and microorganisms that produce them, fermentation of nonglucose sugars (i.e. xylose), and “consolidated bioprocessing”, where the production of enzymes for biomass conversion (i.e. cellulose production), hydrolysis or modification of the biomass (i.e. cellulose hydrolysis), and fermentation of solubilized carbohydrates occur in a single step (Lynd et al., 1999). Therefore, prior to engineering microorganisms for 77 biomass conversion it is important to select host organisms with desired characteristics; with emphasis on strains that can utilize low-cost substrates, have high product yield, competitive fitness, and are more robust to environmental stresses (Lynd et al., 1999). Once a good host has been selected based on targeted physiological characteristics and functionalities, one can identify the additionally desirable characteristic that will then be engineered into the host, whether targeting proteins such as enzymes through rational engineering or changing the metabolism and/or metabolic flux through metabolic engineering (Zhang et al., 2009). Rational Engineering Generally speaking, rational engineering refers to planned biochemical changes to a protein through the use of protein sequence and structure information, which in theory corresponds to a physiological or functional change in the proteins behavior. The engineered changes are usually predicted using computational biology and protein sequence data. However, there is limited structural information available for enzymes, for example, in structureefunction relationshipdso predictions on behavioral changes after rational engineering still remain in a trial-like state (Maki et al., 2009). Nonetheless, with increasing knowledge of biomass substrates and a rigorous test of our knowledge about enzyme interactions with plant-based biomass, rational engineering can be a valuable tool in the economical production of biofuels and value-added by-products. Briefly, rational design of proteins can be summed up in three simple steps: (1) a suitable enzyme is chosen based on desired characteristics, (2) using computational biology or a high resolution crystallographic structure, the amino acid sites to be changed are identified, and (3) mutants produced from rationally engineered proteins are characterized (Percival Zhang et al., 2006). Moreover, rational modifications to enzymes often include amino acids substitutions using site-directed mutagenesis, which can be used to increase the stability of enzymes (i.e. thermostability), substrate specificity, cofactor specificity, and the elucidation of enzymatic mechanisms (Bornscheuer and Pohl, 2001). In the field of biomass conversion to biofuels and bioproducts, the use of rational design has pioneering examples as outlined here. For the most part, there are numerous reviews that summarize studies that revealed the mechanism of cellulase and other biomass-converting genes through the use of site-directed mutagenesis (Schulein, 2000; Wilson, 2004; Wither, 2001). On the contrary, very few researchers have reported increasing cellulase and other biomass-converting activities or enhancing properties through site-directed mutagenesis. However, Baker et al. were able to improve the activity of endoglucanase
78 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS Cel5A of Acidothermus celluloyticus toward microcrystalline cellulose by 20% (Baker et al., 2005). This was accomplished utilizing a high-resolution crystallographic structure (Sakon et al., 1996) to determine a series of mutations designed to alter the active cleft through a change in chemistry of the product-leaving side. As a result, structural information allowed endproduct inhibition to be alleviated by a substitution of a nonaromatic residue at site 245; a Y245G mutant increased the KI of cellobiose by 15-fold. In a similar study, site-directed mutagenesis was used to improve the catalytic activity of endo/exocellulase Cel9A in Thermobifida fusca by 40% with soluble and amorphous cellulose, such as carboxymethyl cellulose (CMC) and swollen cellulose. Through the use of computer modeling, the conserved phenylalanine residue F476 (one of three residues) was found at the end of the carbohydrate binding module and appeared to play an important role in the initial binding of the cellulase to substrate. Also, computer modeling was used to predict that a new hydrogen bond could be created as a result of mutating the conserved phenylalanine residue F476 to a tyrosine. Therefore, the observed increase in catalytic activity of mutant F476Y is thought to be attributed to better binding properties, which are key for placing the soluble and amorphous cellulose chains in the carbohydrate binding domain (Escovar-Kousen et al., 2004). Rational engineering of enzymes can also be used to improve characteristics such as thermostability and alkalinity in addition to specific activity. The roles of highly conserved residues (Asp 60, Tyr 35 and Glu 141), near the catalytic site, were investigated in the pHdependent activity of xylanase XYL1p from Scytalidium acidophilum using site-directed mutagenesis. In doing so, three single mutants, D60N, Y35W and E141A, were created and the activities of three combined xylanase mutants DN/YW, DN/EA and YW/EA were evaluated at different pHs and temperatures. An increased pH optimum of 0.5e1.5 pH units and lower specific activities were observed in all the mutants except one. Mutant E141A exhibited a 50% increase in specific activity at pH 4.0 and had an overall higher catalytic efficiency than wild-type enzyme (Al Balaa et al., 2009). This work presents some important knowledge in acidophilic adaptation and, at the same time, is a prime example of how rational engineering can lead to the development of enzymes more suitable for the bioconversion industry environment, with competitive catalytic efficiency maintained. Finally, the possibility of using rational engineering to improve the pH optimum and catalytic efficiency of laccase enzymes, involved in the oxidation of lignin, has been increasing as several researchers explore important residues conserved in laccases from fungi (Rogers et al., 2009). In one compelling example, researchers replaced an Asp residue in position 206 with an Asn residue in a laccase from T. versicolor, using site-directed mutagenesis. Upon expression of mutants in the yeast Yarrowia lipolytica, it was noted that catalytic activity was significantly affected as the pH optimum was raised by 1.4 pH units (Madzak et al., 2006), highlighting the interaction between the reducing substrate and the binding pocket of laccase. This study, like those discussed previously, pave the way for future development of efficient biomass-converting enzymes. Metabolic Engineering of Microbial Pathways for Enhanced Bioproduct Production Contrary to rational engineering, partial and/or additional metabolic pathways of microorganisms can be engineered to enhance bioproduct production. The term “metabolic engineering” was first coined by Bailey and was described as a vast variety of manipulations and experimental procedures to improve the productivity of a desired metabolite by an organism (Bailey, 1991). More specifically, examples of metabolic engineering can include increased productivity and/or yield, improvement of substrate uptake, widening the scope of substrate range for an organism, modification of metabolic flux, and elimination of unnecessary or competing metabolic pathways (Stephanopoulos, 1999). Metabolic engineering, similar to rational engineering, requires the selection of a good host/microorganism as a candidate for the production of biofuels and/or bioproducts from biomass. This could include engineering desired pathways into well-studied host microorganisms such as Escherichia coli and Saccharomyces cerevisiae; these microorganisms have been used for industrial-scale production for several years. However, some experts suggest that engineering desired pathways into microorganisms that already possess industrial properties may be more successful. This is due to the potential for metabolic burden to the cell; new metabolic pathways require amino acids, redox cofactors, and energy for synthesis and function of its enzymes (Lee et al., 2008a). Furthermore, metabolic engineering poses several general challenges for researchers including the development of recombinant DNA technologies for selected host microorganisms, development of quantitative tools, methods to understand flux modification in complex biological systems, and the development of quantitative techniques to determine changes in fluxes or metabolite concentrations (Cameron and Tong, 1993). A few successful examples of metabolic engineering to improve general host and select host microorganisms metabolism for the digestion and conversion of biomass are outlined below. Recently, the development of genome-scale modeling permits the prediction of how new metabolic pathways
STRATEGIES OF USING MICROBIAL PRETREATMENT TO ENHANCE SUGAR RELEASE FOR BIOFUEL AND BIOPRODUCT PRODUCTION may impact growth and product production using metabolic models. These models result in a more rational approach to metabolic engineering (Patil et al., 2004). Moreover, stoichiometric models can be defined by established equations through the use of metabolic flux analysis (MFA); this is established by measuring exchange fluxes experimentally (Lee et al., 2008b). For example, the native metabolism of E. coli under different growth conditions (Kayser et al., 2005) and during recombinant protein production (Ozkan et al., 2005) has been determined using MFA. For efficient application in biofuel and bioproduct production, genome-scale models should be developed with constraints to optimize flux in desired pathways, while balancing important cofactors and energy metabolites (Lee et al., 2008b). Host microorganisms such as E. coli and S. cerevisae have been improved time and again for the fermentation of sugars to ethanol. In particular, due to the broad range of carbohydrates metabolized by E. coli, it has been a potential candidate for the expression of ethanologenic pathways in some studies. For example, a portable cassette called the production of ethanol operon (PET operon) was used to genetically engineer the homoethanologenic pathway from Zymomonas mobilis into E. coli, which included the pyruvate decarboxylase and alcohol dehydrogenase B genes. Using the PET system, these genes were integrated into the chromosome of E.coli at the pfl locus. Meanwhile the fumarate reductase (frd) gene was deleted to eliminate succinate production, therefore preventing carbon loss. These metabolic changes resulted in the recombinant strain KO11, which produced ethanol yields as high as 95% in complex medium (Jarboe et al., 2007; Ohta et al., 1991). However, host strains such as E.coli may encounter metabolic burdens and are often not naturally adapted to the toxicity of end products like ethanol. Thus, there have also been some attempts to metabolically engineer known biomass-converting bacteria or fungal strains. Typically, bacteria produce more desirable end products through facultative and anaerobic digestion, as is the case for bacteria belonging to the class Clostridia. Much of the metabolic engineering in these species focuses on product formation, which may include the elimination of undesirable products such as in the case of an engineering project conducted on Clostridium acetobutylicumda well-known ethanogenic strain studied often for the production of butanol. In brief, the acetoacetate decarboxylase gene (adc) was disrupted in the hyperbutanol-producing strain C. acetobutylicum EA 2018 using TargeTron technology (Sigma Aldrich) (Jiang et al., 2009). TargeTron is a group II intron developed for rapid and site-specific gene disruption in prokaryotes. The disruption of adc led to an increase in butanol ratio from 70% to 80.05%, with a simultaneous reduction in acetone of 0.21 g/l (Jiang et al., 2009). 79 In contrast, one can implement metabolic engineering to improve native metabolism in microorganisms by engineering entirely novel pathways for desired product formation, which is more practically done in hosts able to hydrolyze biomass, such as the example with Clostridium cellulolyticum. Recently, Higashide et al. demonstrated the production of isobutanol from crystalline cellulose in C. cellulolyticum (Higashide et al., 2011). In this study, the development of valine biosynthesis pathway required the expression of five genes, alsS, ilvC, ilvD, kivD, and ahdA, to convert pyruvate into isobutanol. Consequently, only the expression and function of kivD (2-keto-acid decarboxylase) and alsS (alphaacetolactate synthase) were confirmed; nonetheless modified C. cellulolyticum produced up to 660 mg/l of isobutanol over a 7- to 9-day growth period (Higashide et al., 2011). These examples of engineering and modeling to improve the metabolic capabilities of strains helped lay the foundation for future development of biomassconverting microorganisms. Combined with the ability to rationally design enzymes with greater stability and/or increased specific activity the modification of microorganisms in industrial production of biofuels and bioproducts looks promising. STRATEGIES OF USING MICROBIAL PRETREATMENT TO ENHANCE SUGAR RELEASE FOR BIOFUEL AND BIOPRODUCT PRODUCTION The advantages of biological pretreatment include minimum facility cost, low energy requirement and mild environmental conditions. However, for practical application, there are two major disadvantages associated with this process. First, fungi growth consumes holocellulose as an energy source leading to significant carbohydrate loss; second, most biological pretreatments are long processes due to slow microbial growth and delignification reaction rates. Since lignin breakdown in the biomass would lead to enzyme access to cellulose and hemicellulose, selective lignin degradation by white-rot fungi hold some promise for real application in biomass pretreatment if the procedure can be cut shorter and sugar consumption can be controlled to an insignificantly low level. However, not even white-rot fungi can use lignin as a sole carbon and energy source; fungi growth inevitably results in carbohydrate loss (Fan et al., 2012; Sanchez, 2009). Strategies taken to shorten biological pretreatment time and decrease carbohydrate consumption include (1) selection for naturally occurring white-rot fungi that preferentially attack lignin (Ander Eriksson, 1977; Kirk and Moore, 1972; Lee et al., 2007; Muller and Trosch, 1986; Salvachua
80 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS et al., 2011), (2) selection of cellulase-deficient mutants (Akin et al., 1993; Eriksson et al., 1980; Ruel et al., 1981), or (3) repression of cellulase and hemicellulase expression (Yang et al., 1980). As an example of strain selection, among 22 screened Basidiomycetes, mostly the white-rot fungi Pleurotus sp. “florida” preferentially attacks lignin in wheat straw to increase cellulose accessibility. After 90 days pretreatment with Pleurotus sp. “florida”, the resulting biomass can release the same amount of glucose as Avicel, the lignin-free cellulose (Muller and Trosch, 1986). However, pretreatment using this strain is still time consuming. Furthermore, there are many limitations to the strategies for strain improvement. First, carbohydrate consumption is needed for microbial growth; therefore, strains can only be selected for increased delignification and decreased sugar loss and not for minimal sugar loss. In addition, decreasing the secretion of carbohydrate hydrolysis enzymes would lower the reaction rate and lead to even longer pretreatment time. Genetic modification of white-rot fungi to improve the required features may help resolve some of the drawbacks, but the technical process is quite challenging (Fan et al., 2012). Another way to improve the biological pretreatment process is through optimization of nutrients, temperature, and preprocessing time to reach a balance between maximum sugar release and minimum sugar loss within the shortest possible time. Based on the enzymatic activity profile obtained in a 28-day pretreatment analysis, switchgrass is pretreated with P. chrysosporium for 7 days. The pretreatment of switchgrass led to higher glucan, xylan, and total sugar yields than the unpretreated sample, suggesting enzyme profile assays may be utilized for initial estimation of pretreatment time in order to enhance sugar yields and reduce sugar loss (Mahalaxmi et al., 2010). By monitoring compositional changes during biological pretreatment, a 15-day pretreatment time was selected for the pretreatment of the woody biomasses Prosopis juliflora and Lantana camara with the white-rot fungus Pycnoporus cinnabarimus (Gupta et al., 2011). This 15-day pretreatment resulted in a relatively small weight loss in the pretreated feedstocks with decreased lignin and increased holocellulose contents. Enzymatic hydrolysis of the pretreated biomass led to sugar releases of 389 and 402 mg per gram of dried solid. Alternatively, as a compromise, preliminary microbial pretreatment of biomass can be used in combination with downstream thermochemical, chemical or other pretreatment. This procedure would reduce, for example, the amount of acid needed combined with lower temperature and shorter time, thus reducing energy and chemical costs. In addition, there would be less biomass degradation and inhibitor production compared to conventional thermochemical pretreatment. Preliminary tests showed that after corn stover pretreatment with P. chrysosporium, the shear forces needed to obtain the same shear rates of 3.2e7 rev/s were reduced 10- to 100-fold, respectively. The digestibility of C. stercoreuspretreated corn stover showed a three- to fivefold improvement in enzymatic cellulose digestibility (Keller et al., 2003). Sawada et al. reported that combination of fungal pretreatment with less severe steam explosion maximizes enzymatic saccharification of beech wood meal (Sawada et al., 1995). Compared to steam explosion alone, combined pretreatments improve saccharification by 20e100% of the polysaccharide in the wood. However, 17% of the holocellulose was degraded during fungal pretreatment, and there was an unspecified holocellulose loss during steam explosion at optimum 215  C for 6.5 min (Sawada et al., 1995). Pretreatment of wheat straw with P. juliflora followed by acid hydrolysis led to a reduction in acid load and an increase in sugar release as well as ethanol yield (Kuhar et al., 2008). Interestingly, a recent study showed that by simply changing the pretreatment sequence, i.e. when the wood Pimus radiata biomass was treated first with steam explosion followed by fungi pretreatment, a 10-fold increase in glucose yield was achieved after enzymatic hydrolysis (Vaidya and Singh, 2012). A combination of selected fungal pretreatment with a mild alkali treatment of wheat straw led to a maximum of 69% glucose yield and an ethanol yield of 62% with no inhibitor formation during the pretreatment (Salvachua et al., 2011). Also, a combination of the white-rot fungus Lenzites betulina C5617 pretreatment with LHW treatment enhanced the enzymatic hydrolysis of the poplar wood Populus tomentosa led to the highest hemicellulose removal of 92.33%, which was almost two times higher than that of LHW treatment alone and a 2.66-fold increase in glucose yield (Wang et al., 2012). Application of Microbial Pretreatment for Biogas Production A promising application for microbial pretreatment of lignocellulosic materials is for increasing biogas yield in the anaerobic fermentation process. Anaerobic digestion of organic waste and residues not only provides a good solution for the sustainable processing and treatment of large amounts of biomaterials, but also leads to value-added renewable energy production. Natural lignocellulosic materials can only be converted to biogas at a very low efficiency due to their resistance to anaerobic digestion. The low biogas conversion rate results from the resistance to enzymatic attack by the biomass due to the tight association of lignin, cellulose, and hemicellulose. Under anaerobic conditions, cellulose and hemicellulose can be degraded during biogas production but not lignin (Fernandes et al., 2009). Pretreatment procedures to increase the accessibility of holocellulose are necessary
STRATEGIES OF USING MICROBIAL PRETREATMENT TO ENHANCE SUGAR RELEASE FOR BIOFUEL AND BIOPRODUCT PRODUCTION to increase biogas production. Different pretreatment methods, including physical and chemical pretreatments, effectively enhance anaerobic digestion, but these procedures have disadvantages as described beforehand. A microbial pretreatment followed by another step of biological process seems very promising and close to practical application as shown by some following examples. Pretreatment of wheat straw with Pleurotus sp. "florida" doubles both cellulase digestibility of the treated biomass and the resulting biogas yield, compared with untreated wheat straw (Muller and Trosch, 1986). Pretreatment of softwood in the presence of wheat bran with the white-rot fungus C. subvermispora, which can effectively degrade the lignin component, enhanced methane fermentation of softwood to 35% of the theoretical yield, based on holocellulose content of the biomass. In contrast, pretreatment with Pleurocybella porrigens, which has a lower ability to decompose lignin, led to no significant changes (Amirta et al., 2006). Application of a lignocellulose degrading composite microbial system with high xylanase activity (XDC-2), instead of a pure culture of microorganisms for biomass pretreatment has also been tested. XDC-2 is composed of 26 different clones from three phyla: Clostridiales, Proteobacteria, and Bacteriodetes. However, these degrade mainly carbohydrate but not lignin. After a 5-day pretreatment with XDC-2, corn stalk was efficiently degraded by nearly 45%, and the cellulose and hemicellulose contents were decreased by 22.7% and 74.1%, respectively. Biodegradability of the pretreated biomass is improved resulting from changes in chemical structure due to decreased holocellulose content. Compared with untreated corn stalks, total biogas production and methane yield were increased by 68.3% and 87.9%, respectively, and the technical digestion time (T80) was shortened by 35.7% (Yuan et al., 2011). Effectiveness of biological pretreatments in enhancing corn straw biogas production has also been reported with complex microbial agents including yeast (S. cerevisiae, Coccidioides immitis, and Hansenula anomala), cellulolytic bacteria (Bacillus licheniformis, Pseudomonas sp., Bacillus subtilis, and Pleurotus florida), and the lactic acid bacteria Lactobacillus deiliehii. A 15-day pretreatment of corn straw at ambient temperature led to reduced contents of total lignin, cellulose, and hemicellulose, and increased content of hot-water extractives. Anaerobic digestion of the pretreated material resulted in 33.07% more biogas yield, 75.57% more methane yield, and 34.6% shorter technical digestion time compared with the untreated sample (Zhong et al., 2011). In conclusion, under proper conditions, microbial/ biological pretreatment can be an effective method for improving biodegradability and enhancing downstream biological conversion efficiency of biomass into bioenergy and other value-added bioproducts. 81 Application of Microbial Pretreatment for Biomass Conversion Strategies for Microorganism Application in Biomass Most naturally occurring microorganisms cannot utilize untreated lignocellulose efficiently for the production of biofuel or bioproducts due to the inaccessibility of the carbohydrate polymers, even though many of them secrete a variety of hydrolytic enzymes. For efficient utilization, biomass must first be pretreated to open up the cell wall and then hydrolyzed by acidic or enzymatic processes to fermentable sugar monomers. In addition to monomeric sugars, the pretreatment and acidic hydrolysis processes may also produce low molecular weight organic acids like acetic acid, furfural, hydroxymethylfurfural and various lignin-degradation products that are potent inhibitors of microbial metabolism (Larsson et al., 1999; Palmqvist and HahnHägerdal, 2000). For an economically viable manufacturing process from lignocellulosic biomass, both hexose and pentose sugars produced during hydrolysis of both cellulose and hemicelluloses need to be utilized efficiently. In the course of cellulosic biomass conversion into biofuels and bioproducts, four biologically mediated processes are involved: (1) saccharolytic enzyme production, (2) enzymatic hydrolysis of biomass, (3) fermentation of hexose sugars, and (4) fermentation of pentose sugars (Lynd et al., 2005, 2002). For an industrially viable process, each of the four steps must be rapid and efficient. As suggested by a recent calculation, an economically competitive fermentation process for industrial application needs to approach an anaerobic yield of w95% of the theoretical yield, produce around 100 g/l of end product with a productivity of more than 2 g/l/h (Sheridan, 2009). DIFFERENT PROCESSES OF MICROORGANISMMEDIATED BIOMASS CONVERSION For enzymatic hydrolysis and fermentation, different strategies have been explored including separate hydrolysis and fermentation (SHF), SSF nonisothermal simultaneous saccharification and fermentation (NSSF), simultaneous saccharification and cofermentation (SSCF), or consolidated bioprocessing (CBP) (Lynd et al., 2002; Taherzadeh and Karimi, 2007). Each process has advantages and disadvantages. For SHF, the main advantage is the possibility to separately optimize hydrolysis and fermentation steps and the main drawback is the inhibition of cellulase activity by the released sugars, mainly cellobiose and glucose (Taherzadeh and Karimi, 2007). SSF, different from SHF, combines the enzymatic hydrolysis and fermentation in one step, thus minimizing the product inhibition of cellulase enzymes as the released sugars are immediately
82 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS consumed by the microorganism. In addition, cellulase production and fermentation of hemicellulose hydrolysis products occur in two additional, discrete process steps. This process has many advantages over SHF such as increased ethanol yield, decreased enzyme loading, decreased contamination, and lower capital cost. The disadvantages are differences between optimum temperatures for enzyme hydrolysis and fermentation and inhibition of cellulase by the produced ethanol (Lynd et al., 2002; Olofsson et al., 2008). To solve the issue of temperature difference, the NSSF process was proposed (Wu and Lee, 1998) in which saccharification and fermentation occur simultaneously but in two separate reactors, each operated at its own optimum temperature. Compared to SSF, NSSF increased ethanol yield and productivity with a reduced overall enzyme loading of 30e40%. The disadvantage is increased capital cost for extra equipment. In SSCF, enzymatic biomass hydrolysis and fermentation of both cellulose and hemicellulose hydrolysis products all occur in a single bioreactor with a single microorganism (Teixeira et al., 2000). It is considered an improved process compared to SSF, which requires two bioreactors with two different microorganisms and two different biomass production setups (Hamelinck et al., 2005; McMillan, 1997; McMillan et al., 1999). However, SSCF usually requires a metabolically engineered microorganism that can robustly coferment both glucose and xylose (Teixeira et al., 2000) without synthesis of side products. For example, when a naturally occurring strain, Lactobacillus pentosus (American Type Culture Collection, ATCC 8041), was used in an SSCF process using pretreated corn stover as substrate and the commercial cellulase Spezyme-CP for hydrolysis, the maximum yield of lactic acid was >90% of the theoretical maximum on the basis of all available fermentable sugars. However, acetic acid was also produced through a different metabolic pathway that assimilates pentoses (xylose and arabinose). Another drawback of the process is the difficulty in improving lactic acid concentration due to end-product inhibition of the nonengineered strain (Zhu et al., 2007). All the above-mentioned processes require a separate enzyme production step or an external supply of enzymes for biomass hydrolysis. In CBP, enzyme production, biomass hydrolysis, and fermentation of pentoses and hexoses are accomplished in a single reactor by mono- or cocultures of microorganisms (Lynd et al., 2002). The obvious advantages of CBP are decreased capital costs and no extra cost for enzyme production or purchasing (Hamelinck et al., 2005; Lynd et al., 2005). However, since naturally occurring microorganisms cannot simultaneously synthesize enough of the necessary saccharolytic enzymes and convert released sugars into the desired end products, the CBP configuration requires the development of engineered microorganisms (Hasunuma and Kondo, 2012a; Xu et al., 2009). Such “superbugs” need to not only secrete high titer, robust enzymes, but also efficiently produce ethanol and other bioproducts at high yields under harsh environments containing toxic compounds. CBP is gaining increasing recognition as a potential breakthrough for low-cost biomass processing (Hasunuma and Kondo, 2012a; van Zyl et al., 2007). The company Mascoma Corporation claims to have successfully engineered microorganisms for industrial CBP application (http://www.mascoma. com/). Commonly Used Microorganisms in Biomass Conversion and Some Application Examples A large number of microorganisms are capable of degrading plant cell walls including bacteria and fungi. With few exceptions, two distinct cellulolytic strategies have been adapted by the aerobic and anaerobic groups. While aerobic bacteria and fungi produce numerous individual, extracellular enzymes with many of them acting in synergy for effective hydrolysis, anaerobic bacteria and fungi possess a unique extracellular multienzyme complex, termed the cellulosome, that can efficiently hydrolyze crystalline cellulose (Bayer et al., 2004, 1998; Doi and Kosugi, 2004; Fontes and Gilbert, 2010; Lamed et al., 1983; Lynd et al., 2002; Schwarz, 2001; Shoham et al., 1999; Steenbakkers et al., 2003). Metabolic utilization of the monomeric sugars from hydrolyzed biomass leads to the natural production of biofuels and bioproducts, mostly as side products by different microorganisms. For ethanol fermentation of lignocellulosic biomass, most frequently considered microorganisms include the bacteria E. coli, Z. mobilis and Clostridium phytofermentans; themophilic bacteria such as Clostridium thermocellum; yeasts such as S. cerevisiaeand Pichia stipitis; and filamentous fungi (Amore and Faraco, 2012; Hahn-Hagerdal et al., 2007; Weber et al., 2010; Xu et al., 2009). Like ethanol, the majority of other potential biofuels and bioproducts are naturally produced by various microorganisms as side products. The viability of a fermentation process for industrial application depends on its cost-competitiveness. As listed in Table 5.1, most microorganisms cannot use polymeric carbohydrates directly as fermentation substrates; therefore, biomass has to be broken down into monomeric sugars to be used as fermentation substrates. For an economically viable manufacturing process of biofuels from lignocellulosic biomass, pentose utilization is essential. Therefore, an optimal microorganism should be able to simultaneously ferment both hexose and pentose sugars and give rise to high productivities and yields. In addition, it should have high tolerance to fermentation inhibitors and end products and resist microbial
STRATEGIES OF USING MICROBIAL PRETREATMENT TO ENHANCE SUGAR RELEASE FOR BIOFUEL AND BIOPRODUCT PRODUCTION 83 TABLE 5.1 Typical Features of Representative Microorganisms for Biofuel Production Strain Pros Cons References E. coli Pentose utilization Not resistant to environmental stress, low ethanol and butanol tolerance (Jeffries, 1983; Knoshaug and Zhang, 2009; Shin et al., 2010; Trinh and Srienc, 2009; Yomano et al., 1998, 2008) Z. mobilis High ethanol yield and productivity; high ethanol tolerance Cannot metabolize pentose sugars (Rogers et al., 1982; Weber et al., 2010) Clostridium phytofermentans (ethanol), Clostridium acetobutylicum (butanol) Saccarify cellulose and hemicellulose, ferment hexose and pentose sugars Slow growth rate, low productivity, sensitive to bacteriophage infection (Jones et al., 2000; Lee et al., 2008a,b; Maki et al., 2009; Warnick et al., 2002) S. cerevisiae High robustness, highly resistant to toxic inhibitors and end products Cannot naturally ferment pentose sugars (Olofsson et al., 2008; Yanase et al., 2010a,b) P. stipitis Naturally ferment xylose Lower sugar consumption rate than S. cerevisiae; sequential fermentation of glucose and xylose (Agbogbo and Coward-Kelly, 2008; Jeffries, 1983; Jeffries et al., 2007; Parekh and Wayman, 1986) Kluyveromyces marxianus Thermotolerance allowing higher fermentation temperature, optimum SSF process at lower enzyme loading, lower operation cost, potential application in CBP Poor xylose fermentation, undesirable side product (Babiker et al., 2010; Banat et al., 1992; Hasunuma and Kondo, 2012a,b; Yanase et al., 2010a,b) Clostridium thermocellum Thermophilic anaerobe that grows fast on crystalline cellulose, both cellulolytic and ethanologenic, hydrolyze homocellulose and directly ferment hexose sugars to ethanol and organic acids, no need for external enzyme addition No pentose fermentation, branched fermentation pathways lead to acetate and lactate byproducts, low ethanol production efficiency, low ethanol tolerance (Demain et al., 2005; Lynd et al., 2005; Ng et al., 1981; Raman et al., 2011; Roberts et al., 2010; Zhang and Lynd, 2005) T. reesei Hyper producer of cellulolytic enzymes, extensive knowledge and tools for genetic manipulation and practical application Extensive efforts needed for strain development, low ethanol yield and productivity, low ethanol tolerance (Amore and Faraco, 2012; Xu et al., 2009) contamination, e.g. bacteriophage infections (Weber et al., 2010). No naturally occurring microorganism has all the required features. Promising means to develop a microorganism for sustainable bioethanol/bioproduct production include breeding technologies, genetic engineering and the search for undiscovered species (Weber et al., 2010). For production of a particular product from a specific biomass, native organisms can be selected from a group of different species of microbes based on their fermentation performance, such as substrate utilization efficiency, inhibitor resistance, and productivity (Rumbold et al., 2010, 2009). The yeast S. cerevisiae is by far the most widely used organism in the existing fermentation industry. To improve its application in bioethanol fermentation from biomass, targeted evolution strategy has been applied to obtain inhibitor-tolerant S. cerevisiae that can resist individual or multiple inhibitors (Ding et al., 2012; Heer and Sauer, 2008; Liu, 2006). When adaptation and selection processes were applied to the parental fungus Rhizopus oryzae, a new strain was obtained that exhibited significantly improved efficiency of substrate utilization and enhanced production of L-(þ)lactic acid from corncob hydrolysate. The final product concentration, yield, and volumetric productivity more than doubled compared with its parental strain (Bai et al., 2008). Applications of thermotolerant mesophilic microorganisms in the fermentation process have considerable potential for cost-effective ethanol and other bioproduct production. The thermotolerant yeast Kluyveromyces marxianus grows well at temperatures as high as 45e52  C and can efficiently ferment ethanol at temperatures of between 38 and 45  C. A 5  C increase in the fermentation temperature can greatly decrease fuel
84 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS ethanol production costs (Babiker et al., 2010). Results from solid state fermentation of sweet sorghum stalk to ethanol with the thermotolerant yeast strain Issatchenkia orientalis IPE 100A showed great potential for its practical application in large-scale, deep-bed solid state fermentation (Kwon et al., 2011). The thermotolerant Bacillus coagulans strain 36D1 can ferment both hexoses and pentoses from enzymatically hydrolyzed biomass at 50e55  C and pH 5.0 producing L (þ)-lactic acid as the primary fermentation product. Since such conditions are closer to the optimum fungal enzyme functioning requirements, the amount of enzyme required for cellulose conversion is significantly reduced in comparison with yeast or lactic acid bacteria currently used by the industry as microbial biocatalysts. In addition, both biomass conversion efficiency and product yield are greatly increased with a dramatically decreased fermentation time, thus reducing the cost of both the process and final product (Ou et al., 2009). The anaerobic mesophilic bacterium C. phytofermentans (ATCC 700,394) is a promising native microorganism for biomass conversion since its genome encodes the highest number of enzymes for degradation of lignocellulosic material among sequenced Clostridial genomes (Warnick et al., 2002; Weber et al., 2010). It secretes noncomplex, individual enzymes to hydrolyze both cellulose and hemicelluloses to both hexose and pentose sugars, which are mostly directly consumed, producing ethanol and acetate as the major products (Warnick et al., 2002; Weber et al., 2010). When used in the CBP process with pretreated corn stover as substrate, at optimal conditions with low solid loading (0.5% w/w), C. phytofermentans hydrolyzed 76% of glucan and 88.6% of xylan in 10 days. These values reach 87% and 102% of those obtained by SSCF process using commercial enzymes and S. cerevisiae 424A with an ethanol titer of 2.8 g/l corresponding to 71.8% of that yielded by SSCF (3.9 g/l) (Jin et al., 2011a). However, using a similar process with high solid loading (4% w/w), the side product acetate became a major product (Jin et al., 2012). Even though C. thermocellum seems a good candidate for ethanol fermentation from cellulosic biomass, there are a few disadvantages as listed in Table 5.1. Despite its ability to degrade lignocellulosic waste to both hexose and pentose sugars, it can only utilize hexose sugars from cellulose and not the pentose sugars derived from hemicellulose (Lynd et al., 2002; Taylor et al., 2009). This drawback could be solved by the use of mixed cultures for the degradation and fermentation of all sugars derived from lignocellulosic materials. For example, the anaerobic thermophile Thermoanaerobacterium saccharolyticum, which can ferment xylan and almost all soluble biomass sugars, would be a good candidate for coculture with C. thermocellum. A twofold reduction of the bioethanol production cost from lignocellulose could be achieved when using thermophilic anaerobic mixed cultures (Demain et al., 2005; Lynd et al., 2002). Since there is currently no perfect CBP microbe that can degrade lignocellulosic biomass efficiently and at the same time utilize all the sugars released from biomass to produce mostly ethanol, coculture or community/ mixed fermentation may be a suitable option (Barnard et al., 2010; Demain, 2009; Jin et al., 2011a). Chen reviewed 35 coculture systems for ethanol production by cofermentation of glucose and xylose and concluded that even though still in its infancy, this strategy is promising as it can increase ethanol yield and productivity, shorten fermentation time, and reduce process costs (Chen, 2011). FUTURE PERSPECTIVES For a particular product made from lignocellulosic biomass fermentation, it will be difficult to predict which particular microorganism should be finally used in commercial production. For different processes, it is possible that different species may be required. For bioethanol production, S. cerevisiae has some advantages since it is already widely used in large-scale, first-generation bioethanol production with wellestablished processes and technology. An ideal biomass sugar fermentation process needs to reach high product yield by fermenting all biomass sugars including glucose, xylose, arabinose, mannose, and galactose with an optimal microorganism that is resistant to toxic materials/chemicals in biomass hydrolysates such as acids, phenolics, salts, and sugar oligomers. In addition, the microorganism should be robust, resistant to contamination and environmental stresses, with minimal metabolic by-product production. To achieve these goals, metabolic engineering, or extensive physiological reprogramming of the producing organisms may provide solutions. Other Bioproducts Produced by Microbial Conversion of Biomass: Introduction The use of microorganisms in conversion processes to produce usable material from biomass sources has been ongoing for several decades. Most of the reports in the literature discuss the development of bioprocesses that are involved in the production of simple sugars, which are then used to produce bioethanol or related compounds for use as biofuels. However, there are new trends emerging for the use of biomass conversion by microbes, as shown in Table 5.2. Biomass conversion processes may eventually be implemented to produce a much greater array of useful bioproducts, in addition to biofuels.
STRATEGIES OF USING MICROBIAL PRETREATMENT TO ENHANCE SUGAR RELEASE FOR BIOFUEL AND BIOPRODUCT PRODUCTION 85 TABLE 5.2 List of Bioproducts Produced by Different Microorganisms Bioproduct Organism Conversion References Biofuel Clostridium thermosaccharolyticum Xylose to ethanol (Mistry and Cooney, 1989) Engineered Escherichia coli Cell wall sugars to biofuel (Doran-Peterson et al., 2008) Lactobacillus buchneri NRRL B-30929 Xylose and glucose to ethanol and chemicals (Liu et al., 2009) Saccharomyces cerevisiae Heptanal to heptanol (Verma et al., 2010) Saccharomyces cerevisiae AM12 Spent shiitake mushroom medium (using Meicelase) into ethanol (Asada et al., 2011) Pichia stiptis NCIM3498 and Saccharomyces cerevisiae-VS3 Hemicellulosic hydrolysate to ethanol (Chandel et al., 2011) Methanosarcinales and Methanomicrobiales Coal to methane (Wawrik et al., 2012) Saccharomyces cerevisiae daughter strains Pretreated pine to ethanol (Hawkins and Doran-Peterson, 2011) Trichoderma reesei xylanase Wheat biomass to bioethanol (Juodeikiene et al., 2012) Saccharomyces cerevisiae Lignocellulose-derived sugars to ethanol (Madhavan et al., 2012) Clostridium saccharoperbutylacetonicum n-butyrate to n-butanol (Richter et al., 2012) Burkholderia sp. C20 Microalgal oil to biodiesel (Tran et al., 2012) Cyathus stercoreus and Ceriporiopsis subversmispora Grass stem pretreatment (Akin et al., 1995) Ceriporia lacerata, Stereum hirsutum, and Polyporus brumalis Softwood pretreatment (Lee et al., 2007) Ceriporiopsis subvermispora Corn stover pretreatment for enzymatic hydrolysis and ethanol production (Wan and Li, 2010) Trametes versicolor Canola straw pretreatment for biofuel production (Canam et al., 2011) Pleurotus ostreatus Wood degradation (Piskur et al., 2011) Irpex lacteus Straw saccharification (Pinto et al., 2012) Tramete hirsuta Paddy straw pretreatment for improved enzymatic saccharification (Saritha et al., 2012b) Phanerochaete chrysosporium Pretreatment of cornstalk to enhance enzymatic saccharification and hydrogen production (Zhao et al., 2012) Aureobasidium pullulans (yeastlike mold strain) Glucose to gluconic acid (Anastassiadis et al., 2003) Enterobacter aerogenes 230S L-Psicose (Rao et al., 2008) Debaryomyces hansenii D-xylose and sugarcane bagasse hemicellulose to xylitol (Prakash et al., 2011) Agromyces sp. C42 and Stenotrophomonas sp. A10b (from yellow mealworm gut) Lignocellulose to reducing sugars (Qi et al., 2011) Ustilago maydis Fungal lignocellulosic biomass to glucose and other sugars (Couturier et al., 2012) Pretreated/delignified biomass Simple sugars to L-tagatose (Continued)
86 TABLE 5.2 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS List of Bioproducts Produced by Different Microorganismsdcont’d Bioproduct Lipids Organic chemicals Other Organism Conversion References Debaryomyces hansenii NRRL Y-7426 Distilled grape marc hemicellulosic hydrolysates to xylitol (Salgado et al., 2012) Candida athensensis SB18 D-xylose and horticultural waste hemicellulosic hydrolysate to xylitol (Zhang et al., 2012a) Acidotermus celluloyticus endoglucanase Cellulose to glucose (Zhang et al., 2012b) Cellulolytic fungus of Aspergillus oryzae A-4 Wheat straw to lipid (Lin et al., 2010) Engineered Escherichia coli Simple sugars to fatty esters, fatty alcohols and waxes (Steen et al., 2010) Ustilago maydis Crude glycerol to glycolipids (Liu et al., 2011) Cryptococcus curvatus Crude glycerol to oleic acid, palmitic acid, stearic acid and linoleic acid (Thiru et al., 2011) Trichosporon coremiiforme Organic acids and residual sugars (following butanol fermentation) to oil (Chen et al., 2012a) Trichosporon cutaneum Corncob acid hydrolysate to oil (Chen et al., 2012b) Lipomyces starkeyi Cellobiose and xylose into intracellular lipids (Gong et al., 2012) Rhodococcus opacus DSM 1069 and PD630 Lignin model compounds to triglycerides (Kosa and Ragauskas, 2012) Clostridium lentocellum SG6 Cellulose to acetic acid (Tammali et al., 2003) Saccharomyces uvarum SW-58 Ethyl 4,4,4-trifluoroacetoacetate to ethyl (R)-4,4,4-trifluoro3-hydroxybutanoate [(R)-2] (He et al., 2007) Engineered E. coli Glucose to glucuronic and glucaric acid (Moon et al., 2009) Phanerochaete chrysosporium Rice straw biodelignification in the presence of dirhamnolipid biosurfactant (Liang et al., 2010) Schizophyllum commune Cinnamic acid derivatives to phenols (Nimura et al., 2010) Aspergillus parasiticus speare BGB Glycyrrhizinic acid in liquorice to 18-beta glycyrrhetinic acid (Wang et al., 2010) Gliocladium spp. and E. coli Cellulosic biomass to hydrocarbons (Ahamed and Ahring, 2011) Actinobacillus succinogenes Sugarcane bagasse hemicellulose hydrolysate to succinic acid (Borges and Pereira, 2011) Engineered Thermobifida fusca Untreated lignocellulosic biomass to 1-propanol (Deng and Fong, 2011) Plasticicumulans acidivorans/Thauera selenatis mixed culture Lactate, lactate/acetate mix to poly3-hydroxy butyrate (Jiang et al., 2011) Klebsiella pneumoniae Glycerol and xylose cofermentation to 1,3-propanediol (Jin et al., 2011b) Clostridium ragsdalei Acetone to isopropanol (Ramachandriya et al., 2011) Pseudonocardia carboxydivorans Compactin to pravastatin (Lin et al., 2011) Ganoderma sp. rckk02 Wheat straw to nutritive ruminant feed (Shrivastava et al., 2012) Brevundimonas sp. SGJ L-Tyrosine (Surwase et al., 2012) to L-dihydroxyphenylalanine Lactobacillus brevis TCCCC13007 Monosodium glutamate to gammaaminobutyric acid (Zhang et al., 2012c)
REFERENCES References Agbogbo, F.K., Coward-Kelly, G., 2008. Cellulosic ethanol production using the naturally occurring xylose-fermenting yeast, Pichia stipitis. Biotechnol. Lett. 30, 1515e1524. Ahamed, A., Ahring, B.K., 2011. Production of hydrocarbon compounds by endophytic fungi Gliocladium species grown on cellulose. Bioresour. Technol. 102, 9718e9722. Akin, D.E., Rigsby, L.L., Sethuraman, A., Morrison, W.H., Gamble, G.R., Eriksson, E.K., 1995. Alterations in structure, chemistry, and biodegradability of grass lignocellulose treated with the white rot fungi Ceriporiopsis subvermispora and Cyathus stercoreus. Appl. Environ. Microbiol. 61, 1591e1598. Akin, D.E., Sethuraman, A., Morrison, W.H., Martin, S.A., Eriksson, E.K., 1993. Microbial delignification with white rot fungi improves forage digestibility. Appl. Environ. Microbiol. 59, 4274e4282. Al Balaa, B., Brijs, K., Gebruers, K., Vandenhaute, J., Wouters, J., Housen, I., 2009. Xylanase XYL1p from Scytalidium acidophilum: site-directed mutagenesis and acidophilic adaptation. Bioresour. Technol. 100, 6465e6471. Ali, M., Sreekrishnan, T.R., 2001. Aquatic toxicity from pulp and paper mill effluents: a review. Adv. Environ. Res. 5, 175e196. Alizadeh, H., Teymouri, F., Gilbert, T.I., Dale, B.E., 2005. Pretreatment of switchgrass by ammonia fiber explosion (AFEX). Appl. Biochem. Biotechnol. 121-124, 1133e1141. Alvira, P., Tomas-Pejo, E., Ballesteros, M., Negro, M.J., 2010. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour. Technol. 101, 4851e4861. Amirta, R., Tanabe, T., Watanabe, T., Honda, Y., Kuwahara, M., 2006. Methane fermentation of Japanese cedar wood pretreated with a white rot fungus, Ceriporiopsis subvermispora. J. Biotechnol. 123, 71e77. Amore, A., Faraco, V., 2012. Potential of fungi as category I consolidated bioprocessing organisms for cellulosic ethanol production. Renewable Sustainable Energy Rev. 1, 3286e3301. Anastassiadis, S., Aivasidis, A., Wandrey, C., 2003. Continuous gluconic acid production by isolated yeast-like mould strains of Aureobasidium pullulans. Appl. Microbiol. Biotechnol. 61, 110e117. Ander, P., Eriksson, K.E., 1977. Selective degradation of wood components by white-rot fungi. Physiol. Plant 41, 239e248. Arantes, V., Jellison, J., Goodell, B., 2012. Peculiarities of brown-rot fungi and biochemical Fenton reaction with regard to their potential as a model for bioprocessing biomass. Appl. Microbiol. Biotechnol. 94, 323e338. Asada, C., Asakawa, A., Sasaki, C., Nakamura, Y., 2011. Characterization of the steam-exploded spent shiitake mushroom medium and its efficient conversion to ethanol. Bioresour. Technol. 102, 10052e10056. Azuma, J.I., Tanaka, F., Koshijima, T., 1984. Enhancement of enzymatic susceptibility of lignocellulosic wastes by microwave irradiation. J. Ferment. Technol. 62, 377e384. Babiker, M.A., Banat, A., Hoshida, H., Ano, A., Nonklang, S., Akada, R., 2010. High-temperature fermentation: how can processes for ethanol production at high temperatures become superior to the traditional process using mesophilic yeast? Appl. Microbiol. Biotechnol. 85, 861e867. Bai, D.M., Li, S.Z., Liu, Z.L., Cui, Z.F., 2008. Enhanced L-(þ)-lactic acid production by an adapted strain of Rhizopus oryzae using corncob hydrolysate. Appl. Biochem. Biotechnol. 144, 79e85. Bailey, J.E., 1991. Toward a science of metabolic engineering. Science 252, 1668e1675. Bak, J.S., Ko, J.K., Choi, I.G., Park, Y.C., Seo, J.H., Kim, K.H., 2009. Fungal pretreatment of lignocellulose by Phanerochaete chrysosporium to produce ethanol from rice straw. Biotechnol. Bioeng. 104, 471e482. 87 Baker, J.O., McCarley, J.R., Lovett, R., Yu, C.H., Adney, W.S., Rignall, T.R., Vinzant, T.B., Decker, S.R., Sakon, J., Himmel, M.E., 2005. Catalytically enhanced endocellulase Cel5A from Acidothermus cellulolyticus. Appl. Biochem. Biotechnol. 121-124, 129e148. Baldrian, P., Valaskova, V., 2008. Degradation of cellulose by basidiomycetous fungi. FEMS Microbiol. Rev. 32, 501e521. Ballesteros, I., Negro, M.J., Oliva, J.M., Cabanas, A., Manzanares, P., Ballesteros, M., 2006. Ethanol production from steam-explosion pretreated wheat straw. Appl. Biochem. Biotechnol. 129-132, 496e508. Banat, I.M., Nigam, P., Marchant, R., 1992. Isolation of thermotolerant, fermentative yeasts growing at 52  C and producing ethanol at 45  C and 50  C. World J. Microbiol. Biotechnol. 8, 259e263. Barnard, D., Casanueva, A., Tuffin, M., Cowan, D., 2010. Extremophiles in biofuel synthesis. Environ. Technol. 31, 871e888. Bayer, E.A., Belaich, J.P., Shoham, Y., Lamed, R., 2004. The cellulosomes: multienzyme machines for degradation of plant cell wall polysaccharides. Annu. Rev. Microbiol. 58, 521e554. Bayer, E.A., Shimon, L.J., Shoham, Y., Lamed, R., 1998. Cellulosomes structure and ultrastructure. J. Struct. Biol. 124, 221e234. Bennet, J.W., Wunch, K.G., Faison, B.D., 2002. Use of fungi in biodegradation. In: Hurst, C.J. (Ed.), Manual of Environmental Microbiology. ASM Press, Washington, D.C. Blanchette, R.A., 1991. Delignification by wood-decay fungi. Annu. Rev. Phytopathol. 29, 381e398. Blanchette, R.A., 1995. Degradation of the lignocellulose complex in wood. Can. J. Bot. 73 (suppl. 1), 999e1010. Blanchette, R.A., 2000. A review of microbial deterioration found in archaeological wood from different environments. Int. Biodeterior. Biodegrad. 46, 189e204. Blanchette, R.A., Burnes, T.A., Erdmans, M.M., Akhtar, M., 1992. Evaluating isolates of Phanerochaete chrysosporium and Ceriporiopsis subvermispora for use in biological pulping processes. Holzforschung 46, 109e115. Blanchette, R.A., Held, B.W., Farrell, R.L., 2002. Defibration of wood in the expedition huts of Antarctica: an unusual deterioration process occurring in the polar environment. Polar Rec. 38, 313e322. Blanchette, R.A., Held, B.W., Jurgens, J.A., McNew, D.L., Harrington, T.C., Duncan, S.M., Farrell, R.L., 2004. Wooddestroying soft rot fungi in the historic expedition huts of Antarctica. Appl. Environ. Microbiol. 70, 1328e1335. Borges, E.R., Pereira Jr., N., 2011. Succinic acid production from sugarcane bagasse hemicellulose hydrolysate by Actinobacillus succinogenes. J. Ind. Microbiol. Biotechnol. 38, 1001e1011. Bornscheuer, U.T., Pohl, M., 2001. Improved biocatalysts by directed evolution and rational protein design. Curr. Opin. Chem. Biol. 5, 137e143. Cameron, D.C., Tong, I.T., 1993. Cellular and metabolic engineering. An overview. Appl. Biochem. Biotechnol. 38, 105e140. Canam, T., Town, J.R., Tsang, A., McAllister, T.A., Dumonceaux, T.J., 2011. Biological pretreatment with a cellobiose dehydrogenasedeficient strain of Trametes versicolor enhances the biofuel potential of canola straw. Bioresour. Technol. 102, 10020e10027. Chandel, A.K., Singh, O.V., Narasu, M.L., Rao, L.V., 2011. Bioconversion of Saccharum spontaneum (wild sugarcane) hemicellulosic hydrolysate into ethanol by mono and co-cultures of Pichia stipitis NCIM3498 and thermotolerant Saccharomyces cerevisiae-VS(3). New Biotechnol. 28, 593e599. Chandra, R.P., Bura, R., Mabee, W.E., Berlin, A., Pan, X., Saddler, J.N., 2007. Substrate pretreatment: the key to effective enzymatic hydrolysis of lignocellulosics? Adv. Biochem. Eng. Biotechnol. 108, 67e93. Chen, X.F., Huang, C., Xiong, L., Chen, X.D., Chen, Y., Ma, L.L., 2012a. Oil production on wastewaters after butanol fermentation by oleaginous yeast Trichosporon coremiiforme. Bioresour. Technol. 118, 594e597.
88 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS Chen, X.F., Huang, C., Xiong, L., Chen, X.D., Ma, L.L., 2012b. Microbial oil production from corncob acid hydrolysate by Trichosporon cutaneum. Biotechnol. Lett. 34, 1025e1028. Chen, Y., 2011. Development and application of co-culture for ethanol production by co-fermentation of glucose and xylose: a systematic review. J. Ind. Microbiol. Biotechnol. 38, 581e597. Contreras, D., Rodriguez, J., Freer, J., Schwederski, B., Kaim, W., 2007. Enhanced hydroxyl radical production by dihydroxybenzenedriven Fenton reactions: implications for wood biodegradation. J. Biol. Inorg. Chem. 12, 1055e1061. Couturier, M., Navarro, D., Olive, C., Chevret, D., Haon, M., Favel, A., Lesage-Meessen, L., Henrissat, B., Coutinho, P.M., Berrin, J.G., 2012. Post-genomic analyses of fungal lignocellulosic biomass degradation reveal the unexpected potential of the plant pathogen Ustilago maydis. BMC Genomics 13, 57. Crawford, D.L., Barder, M.J., Pometto III, A.L., Crawford, R.L., 1982. Chemistry of softwood lignin degradation by Streptomyces viridosporus. Arch. Microbiol. 131, 140e145. Cui, Z.F., Wan, C.X., Sykes, R., Shi, J., Li, Y.B., 2012. Enzymatic digestibility of corn stover fractions in response to fungal pretreatment. Ind. Eng. Chem. 51, 7153e7159. Curling, S.F., Clausen, C.A., Winandy, J.E., 2002. Relationships between mechanical properties, weight loss, and chemical composition of wood during incipient brown-rot decay. For. Prod. J. 52, 34e39. da Costa Sousa, L., Chundawat, S.P., Balan, V., Dale, B.E., 2009. ’Cradle-to-grave’ assessment of existing lignocellulose pretreatment technologies. Curr. Opin. Biotechnol. 20, 339e347. Daniel, G., Nilsson, T., Singh, A.P., 1987. Degradation of lignocellulosics by unique tunnel-forming bacteria. Can. J. Microbiol. 33, 943e948. Demain, A.L., 2009. Biosolutions to the energy problem. J. Ind. Microbiol. Biotechnol. 36, 319e332. Demain, A.L., Newcomb, M., Wu, J.H., 2005. Cellulase, clostridia, and ethanol. Microbiol. Mol. Biol. Rev. 69, 124e154. Deng, Y., Fong, S.S., 2011. Metabolic engineering of Thermobifida fusca for direct aerobic bioconversion of untreated lignocellulosic biomass to 1-propanol. Metab. Eng. 13, 570e577. Ding, M.-Z., Wang, X., Yang, Y., Yuan, Y.-J., 2012. Comparative metabolic profiling of parental and inhibitors-tolerant yeasts during lignocellulosic ethanol fermentation. Metabolomics 8, 232e243. Doi, R.H., Kosugi, A., 2004. Cellulosomes: plant-cell-wall-degrading enzyme complexes. Nat. Rev. Microbiol. 2, 541e551. Doran-Peterson, J., Cook, D.M., Brandon, S.K., 2008. Microbial conversion of sugars from plant biomass to lactic acid or ethanol. Plant J. 54, 582e592. Duff, S.J.B., Murray, W.D., 1996. Bioconversion of forest products industry waste cellulosics to fuel ethanol: a review. Bioresour. Technol. 55, 1e33. Durand, H., Tiraby, G., 1988. Genetic improvement of Trichoderma reesei for large scale cellulase production. Enzyme Microb. Technol. 10, 341e346. Eriksson, K.E., Grunewald, A., Nilsson, T., Vallander, L., 1980. A scanning electron microscopy study of the growth and attack on wood by three white-rot fungi and their cellulaseless mutants. Holzforschung 34, 207e213. Eriksson, K.E.L., Blanchette, R.A., Ander, P., 1990. Microbial and Enzymatic Degradation of Wood and Wood Components. Springer-Verlag, Berlin, Germany. Escovar-Kousen, J.M., Wilson, D., Irwin, D., 2004. Integration of computer modeling and initial studies of site-directed mutagenesis to improve cellulase activity on Cel9A from Thermobifida fusca. Appl. Biochem. Biotechnol. 113-116, 287e297. Esterbauer, H., Steiner, W., Labudova, I., Hermann, A., Hayn, M., 1991. Production of Trichoderma cellulase in laboratory and pilot scale. Bioresour. Technol. 36, 51e65. Ewen, R.J., Jones, P.R., Ratcliffe, N.M., Spencer-Phillips, P.T., 2004. Identification by gas chromatography-mass spectrometry of the volatile organic compounds emitted from the wood-rotting fungi Serpula lacrymans and Coniophora puteana, and from Pinus sylvestris timber. Mycol. Res. 108, 806e814. Falls, M., Shi, J., Ebrik, M.A., Redmond, T., Yang, B., Wyman, C.E., Garlock, R., Balan, V., Dale, B.E., Pallapolu, V.R., et al., 2011. Investigation of enzyme formulation on pretreated switchgrass. Bioresour. Technol. 102, 11072e11079. Fan, Z., Wu, W., Hildebrand, A., Kasuga, T., Zhang, R., Xiong, X., 2012. A novel biochemical route for fuels and chemicals production from cellulosic biomass. PLoS One 7, e31693. Fernandes, T.V., Bos, G.J., Zeeman, G., Sanders, J.P., van Lier, J.B., 2009. Effects of thermo-chemical pre-treatment on anaerobic biodegradability and hydrolysis of lignocellulosic biomass. Bioresour. Technol. 100, 2575e2579. Findlay, W.P.K., 1984. Soft rot of timberda review. J. Indian Acad. Wood Sci. 15, 1e11. Flournoy, D.S., Kirk, T.K., Highley, T.L., 1991. Wood decay by brownrot fungi: changes in pore structure and cell wall volume. Holzforschung 45, 383e388. Fontes, C.M., Gilbert, H.J., 2010. Cellulosomes: highly efficient nanomachines designed to deconstruct plant cell wall complex carbohydrates. Annu. Rev. Biochem. 79, 655e681. Glenn, J.K., Gold, M.H., 1983. Decolorization of several polymeric dyes by the lignin-degrading basidiomycete Phanerochaete chrysosporium. Appl. Environ. Microbiol. 45, 1741e1747. Gong, Z., Wang, Q., Shen, H., Hu, C., Jin, G., Zhao, Z.K., 2012. Co-fermentation of cellobiose and xylose by Lipomyces starkeyi for lipid production. Bioresour. Technol. 117, 20e24. Gonzalez-Fernandez, C., Sialve, B., Bernet, N., Steyer, J.P., 2012. Comparison of ultrasound and thermal pretreatment of Scenedesmus biomass on methane production. Bioresour. Technol. 110, 610e616. Green, F., Highley, T.L., 1997. Mechanism of brown-rot decay: paradigm or paradox. Int. Biodeterior. Biodegrad. 39, 113e124. Grous, W.R., Convers, A.O., Grethlein, H.E., 1986. Effect of steam explosion pretreatment on pore size and enzymatic hydrolysis of poplar. Enzyme Microb. Technol. 8, 274e280. Gupta, R., Mehta, G., Khasa, Y.P., Kuhad, R.C., 2011. Fungal delignification of lignocellulosic biomass improves the saccharification of cellulosics. Biodegradation 22, 797e804. Hahn-Hagerdal, B., Karhumaa, K., Fonseca, C., Spencer-Martins, I., Gorwa-Grauslund, M.F., 2007. Towards industrial pentosefermenting yeast strains. Appl. Microbiol. Biotechnol. 74, 937e953. Hamelinck, C.N., Van Hooijdonk, G., Faaij, A.P.C., 2005. Ethanol from lignocellulosic biomass: techno-economic performance in short-, middle- and long-term. Biomass Bioenergy 28, 384e410. Hammel, K.E., Kapich, A.N., Jensen, K.A., Ryan, Z.C., 2002. Reactive oxygen species as agents of wood decay by fungi. Enzyme Microb. Tech. 30, 445e453. Hasunuma, T., Kondo, A., 2012a. Consolidated bioprocessing and simultaneous saccharification and fermentation of lignocellulose to ethanol with thermotolerant yeast strains. Process Biochem. 47, 1287e1294. Hasunuma, T., Kondo, A., 2012b. Development of yeast cell factories for consolidated bioprocessing of lignocellulose to bioethanol through cell surface engineering. Biotechnol. Adv. 30, 1207e1218. Hatakka, A., 1994. Lignin-modifying enzymes from selected white-rot fungi - production and role in lignin degradation. FEMS Microbiol. Rev. 13, 125e135. Hatakka, A.I., Usi-Rauva, A.K., 1983. Degradation of 14C-labelled poplar wood lignin by selected white-rot fungi. Eur. J. Appl. Microbiol. Biotechnol. 17, 235e242. Hawkins, G.M., Doran-Peterson, J., 2011. A strain of Saccharomyces cerevisiae evolved for fermentation of lignocellulosic biomass
REFERENCES displays improved growth and fermentative ability in high solids concentrations and in the presence of inhibitory compounds. Biotechnol. Biofuels 4, 49. He, J., Mao, X., Sun, Z., Zheng, P., Ni, Y., Xu, Y., 2007. Microbial synthesis of ethyl (R)-4,4,4-trifluoro-3-hydroxybutanoate by asymmetric reduction of ethyl 4,4,4-trifluoroacetoacetate in an aqueous-organic solvent biphasic system. Biotechnol. J. 2, 260e265. Heer, D., Sauer, U., 2008. Identification of furfural as a key toxin in lignocellulosic hydrolysates and evolution of a tolerant yeast strain. Microb. Biotechnol. 1, 497e506. Hendriks, A.T., Zeeman, G., 2009. Pretreatments to enhance the digestibility of lignocellulosic biomass. Bioresour. Technol. 100, 10e18. Higashide, W., Li, Y., Yang, Y., Liao, J.C., 2011. Metabolic engineering of Clostridium cellulolyticum for production of isobutanol from cellulose. Appl. Environ. Microbiol. 77, 2727e2733. Highley, T., 1987. Changes in chemical components of hardwood and softwood by brown-rot fungi. Mater. Org. 21, 39e45. Howell, C., Steenkjær Hastrup, A.C., Goodell, B., Jellison, J., 2009. Temporal changes in wood crystalline cellulose during degradation by brown rot fungi. Int. Biodeterior. Biodegrad. 63, 414e419. Hsu, T.-A., 1996. Pretreatment of biomass. In: Wyman, C.E. (Ed.), Handbook on Bioethanol, Production and Utilization. Taylor and Francis, USA. Hunt, C., Kenealy, W., Horn, E., Houtman, C., 2004. A biopulping mechanism: creation of acid groups on fiber. Holzforschung 58, 434e439. Irbe, I., Andersone, I., Andersons, B., Noldt, G., Dizhbite, T., Kurnosova, N., Nuopponen, M., Stewart, D., 2011. Characterisation of the initial degradation stage of Scots pine (Pinus sylvestris L.) sapwood after attack by brown-rot fungus Coniophora puteana. Biodegradation 22, 719e728. Jarboe, L.R., Grabar, T.B., Yomano, L.P., Shanmugan, K.T., Ingram, L.O., 2007. Development of ethanologenic bacteria. Adv. Biochem. Eng. Biotechnol. 108, 237e261. Jeffries, T.W., 1983. Utilization of xylose by bacteria, yeasts, and fungi. Adv. Biochem. Eng. Biotechnol. 27, 1e32. Jeffries, T.W., Grigoriev, I.V., Grimwood, J., Laplaza, J.M., Aerts, A., Salamov, A., Schmutz, J., Lindquist, E., Dehal, P., Shapiro, H., et al., 2007. Genome sequence of the lignocellulose-bioconverting and xylose-fermenting yeast Pichia stipitis. Nat. Biotechnol. 25, 319e326. Jensen Jr., K.A., Houtman, C.J., Ryan, Z.C., Hammel, K.E., 2001. Pathways for extracellular Fenton chemistry in the brown rot basidiomycete Gloeophyllum trabeum. Appl. Environ. Microbiol. 67, 2705e2711. Jiang, Y., Marang, L., Kleerebezem, R., Muyzer, G., van Loosdrecht, M.C., 2011. Polyhydroxybutyrate production from lactate using a mixed microbial culture. Biotechnol. Bioeng. 108, 2022e2035. Jiang, Y., Xu, C., Dong, F., Yang, Y., Jiang, W., Yang, S., 2009. Disruption of the acetoacetate decarboxylase gene in solvent-producing Clostridium acetobutylicum increases the butanol ratio. Metab. Eng. 11, 284e291. Jin, M., Balan, V., Gunawan, C., Dale, B.E., 2011a. Consolidated bioprocessing (CBP) performance of Clostridium phytofermentans on AFEX-treated corn stover for ethanol production. Biotechnol. Bioeng. 108, 1290e1297. Jin, P., Li, S., Lu, S.G., Zhu, J.G., Huang, H., 2011b. Improved 1,3-propanediol production with hemicellulosic hydrolysates (corn straw) as cosubstrate: impact of degradation products on Klebsiella pneumoniae growth and 1,3-propanediol fermentation. Bioresour. Technol. 102, 1815e1821. Jin, M., Gunawan, C., Balan, V., Dale, B.E., 2012. Consolidated bioprocessing (CBP) of AFEX-pretreated corn stover for ethanol 89 production using Clostridium phytofermentans at a high solids loading. Biotechnol. Bioeng. 109, 1929e1936. Jones, D.T., Shirley, M., Wu, X., Keis, S., 2000. Bacteriophage infections in the industrial acetone butanol (AB) fermentation process. J. Mol. Microbiol. Biotechnol. 2, 21e26. Jørgensen, H., Kristensen, J.B., Felby, C., 2007. Enzymatic conversion of lignocellulose into fermentable sugars: challenges and opportunities. Biofuels Bioprod. Bioref 1, 119e134. Jung, H.G., Valdez, F.R., Abad, A.R., Blanchette, R.A., Hatfield, R.D., 1992. Effect of white rot basidiomycetes on chemical composition and in vitro digestibility of oat straw and alfalfa stems. J. Anim. Sci. 70, 1928e1935. Juodeikiene, G., Basinskiene, L., Vidmantiene, D., Makaravicius, T., Bartkiene, E., 2012. Benefits of beta-xylanase for wheat biomass conversion to bioethanol. J. Sci. Food Agric. 92, 84e91. Kaneko, S., Yoshitake, K., Itakura, S., Tanaka, H., Enoki, A., 2005. Relationship between production of hydroxyl radicals and degradation of wood, crystalline cellulose, and a lignin-related compound or accumulation of oxalic acid in cultures of brown-rot fungi. J. Wood Sci. 51, 262e269. Karunanithy, C., Muthukumarappan, K., 2010a. Effect of extruder parameters and moisture content of switchgrass, prairie cord grass on sugar recovery from enzymatic hydrolysis. Appl. Biochem. Biotechnol. 162, 1785e1803. Karunanithy, C., Muthukumarappan, K., 2010b. Influence of extruder temperature and screw speed on pretreatment of corn stover while varying enzymes and their ratios. Appl. Biochem. Biotechnol. 162, 264e279. Karunanithy, C., Muthukumarappan, K., Gibbons, W.R., 2012. Extrusion pretreatment of pine wood chips. Appl. Biochem. Biotechnol. 167, 81e99. Kayser, A., Weber, J., Hecht, V., Rinas, U., 2005. Metabolic flux analysis of Escherichia coli in glucose-limited continuous culture. I. Growthrate-dependent metabolic efficiency at steady state. Microbiology 151, 693e706. Keller, F.A., Hamilton, J.E., Nguyen, Q.A., 2003. Microbial pretreatment of biomass: potential for reducing severity of thermochemical biomass pretreatment. Appl. Biochem. Biotechnol. 105e108, 27e41. Kim, Y., Hendrickson, R., Mosier, N.S., Ladisch, M.R., 2009. Liquid hot water pretreatment of cellulosic biomass. Methods Mol. Biol. 581, 93e102. Kirk, T., Tien, M., Johnrud, S.C., Eriksson, K.E., 1986. Lignin-degrading activity of Phanerochaete chrysosporium Burds: comparison of cellulase-negative and other strains. Enzyme Microb. Technol. 8, 75e80. Kirk, T.K., Moore, W.E., 1972. Removing lignin from wood with whiterot fungi and digestibility from resulting wood. Wood Fiber 4, 72e79. Knoshaug, E.P., Zhang, M., 2009. Butanol tolerance in a selection of microorganisms. Appl. Biochem. Biotechnol. 153, 13e20. Kosa, M., Ragauskas, A.J., 2012. Bioconversion of lignin model compounds with oleaginous Rhodococci. Appl. Microbiol. Biotechnol. 93, 891e900. Kramer, C., Kreisel, G., Fahr, K., Kassbohrer, J., Schlosser, D., 2004. Degradation of 2-fluorophenol by the brown-rot fungus Gloeophyllum striatum: evidence for the involvement of extracellular Fenton chemistry. Appl. Microbiol. Biotechnol. 64, 387e395. Kuhar, S., Nair, L.M., Kuhad, R.C., 2008. Pretreatment of lignocellulosic material with fungi capable of higher lignin degradation and lower carbohydrate degradation improves substrate acid hydrolysis and the eventual conversion to ethanol. Can. J. Microbiol. 54, 305e313. Kumar, P., Barrett, D.M., Delwiche, M.J., Stroeve, P., 2009. Methods for pretreatment of lignocellulosic biomass for efficient hydrolysis and biofuel production. Ind. Eng. Chem. Res. 48, 3713e3729.
90 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS Kumar, R., Wyman, C.E., 2009. Effects of cellulase and xylanase enzymes on the deconstruction of solids from pretreatment of poplar by leading technologies. Biotechnol. Prog. 25, 302e314. Kwon, Y.J., Wang, F., Liu, C.Z., 2011. Deep-bed solid state fermentation of sweet sorghum stalk to ethanol by thermotolerant Issatchenkia orientalis IPE 100. Bioresour. Technol. 102, 11262e11265. Lamed, R., Setter, E., Bayer, E.A., 1983. Characterization of a cellulosebinding, cellulase-containing complex in Clostridium thermocellum. J. Bacteriol. 156, 828e836. Larsson, S., Reimann, A., Nilvebrant, N.-O., Jönsson, L.J., 1999. Comparison of different methods for the detoxification of lignocellulose hydrolyzates of spruce. Appl. Biochem. Biotechnol. 77, 91e103. Lee, J.W., Gwak, K.S., Park, J.Y., Park, M.J., Choi, D.H., Kwon, M., Choi, I.G., 2007. Biological pretreatment of softwood Pinus densiflora by three white rot fungi. J. Microbiol. 45, 485e491. Lee, S.K., Chou, H., Ham, T.S., Lee, T.S., Keasling, J.D., 2008a. Metabolic engineering of microorganisms for biofuels production: from bugs to synthetic biology to fuels. Curr. Opin. Biotechnol. 19, 556e563. Lee, S.Y., Park, J.H., Jang, S.H., Nielsen, L.K., Kim, J., Jung, K.S., 2008b. Fermentative butanol production by Clostridia. Biotechnol. Bioeng. 101, 209e228. Liang, Y.S., Yuan, X.Z., Zeng, G.M., Hu, C.L., Zhong, H., Huang, D.L., Tang, L., Zhao, J.J., 2010. Biodelignification of rice straw by Phanerochaete chrysosporium in the presence of dirhamnolipid. Biodegradation 21, 615e624. Lin, C.L., Tang, Y.L., Lin, S.M., 2011. Efficient bioconversion of compactin to pravastatin by the quinoline-degrading microorganism Pseudonocardia carboxydivorans isolated from petroleumcontaminated soil. Bioresour. Technol. 102, 10187e10193. Lin, H., Cheng, W., Ding, H.T., Chen, X.J., Zhou, Q.F., Zhao, Y.H., 2010. Direct microbial conversion of wheat straw into lipid by a cellulolytic fungus of Aspergillus oryzae A-4 in solid-state fermentation. Bioresour. Technol. 101, 7556e7562. Liu, S., Bischoff, K.M., Hughes, S.R., Leathers, T.D., Price, N.P., Qureshi, N., Rich, J.O., 2009. Conversion of biomass hydrolysates and other substrates to ethanol and other chemicals by Lactobacillus buchneri. Lett. Appl. Microbiol. 48, 337e342. Liu, Y., Koh, C.M., Ji, L., 2011. Bioconversion of crude glycerol to glycolipids in Ustilago maydis. Bioresour. Technol. 102, 3927e3933. Liu, Z.L., 2006. Genomic adaptation of ethanologenic yeast to biomass conversion inhibitors. Appl. Microbiol. Biotechnol. 73, 27e36. Loguercio-Leite, C., Michels, J., Baltazar, J.M., 2008. New records of lignocellulolytic basidiomycetes (Fungi): Parque Estadual da Serra do Tabuleiro (P.E.S.T.), Santa Catarina, Brazil. Biotemas 21, 7e14. Lynd, L., Wyman, C.E., Gerngross, T.U., 1999. Biocommodity engineering. Biotechnol. Prog. 15, 777e793. Lynd, L.R., van Zyl, W.H., McBride, J.E., Laser, M., 2005. Consolidated bioprocessing of cellulosic biomass: an update. Curr. Opin. Biotechnol. 16, 577e583. Lynd, L.R., Weimer, P.J., van Zyl, W.H., Pretorius, I.S., 2002. Microbial cellulose utilization: fundamentals and biotechnology. Microbiol. Mol. Biol. Rev. 66, 506e577. Lyr, H., 1960. Formation of ectoenzymes by wood-rotting and woodinhabiting fungi on different culture media. Part V. A complex medium as carbon source. Arch. Microbiol. 35, 258e278. Ma, H., Liu, W.W., Chen, X., Wu, Y.J., Yu, Z.L., 2009. Enhanced enzymatic saccharification of rice straw by microwave pretreatment. Bioresour. Technol. 100, 1279e1284. Madhavan, A., Srivastava, A., Kondo, A., Bisaria, V.S., 2012. Bioconversion of lignocellulose-derived sugars to ethanol by engineered Saccharomyces cerevisiae. Crit. Rev. Biotechnol. 32, 22e48. Madzak, C., Mimmi, M.C., Caminade, E., Brault, A., Baumberger, S., Briozzo, P., Mougin, C., Jolivalt, C., 2006. Shifting the optimal pH of activity for a laccase from the fungus Trametes versicolor by structure-based mutagenesis. Protein Eng. Des. Sel. 19, 77e84. Mahalaxmi, S., Jackson, C., Williford, C., Burandt, C.L., 2010. Estimation of treatment time for microbial preprocessing of biomass. Appl. Biochem. Biotechnol. 162, 1414e1422. Maki, M., Leung, K.T., Qin, W., 2009. The prospects of cellulaseproducing bacteria for the bioconversion of lignocellulosic biomass. Int. J. Biol. Sci. 5, 500e516. Martinez, D., Challacombe, J., Morgenstern, I., Hibbett, D., Schmoll, M., Kubicek, C.P., Ferreira, P., Ruiz-Duenas, F.J., Martinez, A.T., Kersten, P., et al., 2009. Genome, transcriptome, and secretome analysis of wood decay fungus Postia placenta supports unique mechanisms of lignocellulose conversion. Proc. Natl. Acad. Sci. USA 106, 1954e1959. McMillan, J.D., 1994. Pretreatment of lignocellulosic biomass. In: Himmel, J.O.B.M.E., Overend, R.P. (Eds.), Enzymatic Conversion of Biomass for Fuels Production. ACS, Washington, USA, pp. 292e324. McMillan, J.D., 1997. Bioethanol production: status and prospects. Renew. Energ 10, 295e302. McMillan, J.D., Newman, M.M., Templeton, D.W., Mohagheghi, A., Hatakka, 1999. Simultaneous saccharification and cofermentation of dilute-acid pretreated yellow poplar hardwood to ethanol using xylose-fermenting Zymomonas mobilis. Appl. Biochem. Biotechnol. 77-79, 649e665. Mistry, F.R., Cooney, C.L., 1989. Production of ethanol by Clostridium thermosaccharolyticum: II. A quantitative model describing product distributions. Biotechnol. Bioeng. 34, 1305e1320. Monrroy, M., Ortega, I., Ramirez, M., Baeza, J., Freer, J., 2011. Structural change in wood by brown rot fungi and effect on enzymatic hydrolysis. Enzyme Microb. Technol. 49, 472e477. Moon, T.S., Yoon, S.H., Lanza, A.M., Roy-Mayhew, J.D., Prather, K.L., 2009. Production of glucaric acid from a synthetic pathway in recombinant Escherichia coli. Appl. Environ. Microbiol. 75, 589e595. Mosier, N., Wyman, C., Dale, B., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., 2005. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96, 673e686. Muller, H.W., Trosch, W., 1986. Screening of white-rot fungi for biological pretreatment of wheat straw for biogas production. Appl. Microbiol. Biotechnol. 24, 180e185. Ng, T.K., Ben-Bassat, A., Zeikus, J.G., 1981. Ethanol production by thermophilic bacteria: fermentation of cellulosic substrates by cocultures of Clostridium thermocellum and Clostridium thermohydrosulfuricum. Appl. Environ. Microbiol. 41, 1337e1343. Nilsson, T., Daniel, G.F., Kirk, T.K., Obst, J.R., 1989. Chemistry and microscopy of wood decay by some higher ascomycetes. Holzforschung 43, 11e18. Nilsson, T., Singh, A.P., Daniel, G.F., 1992. Ultrastructure of the attack of Eusideroxylon zwageri wood by tunnelling bacteria. Holzforschung 46, 361e367. Nimura, Y., Tsujiyama, S., Ueno, M., 2010. Bioconversion of cinnamic acid derivatives by Schizophyllum commune. J. Gen. Appl. Microbiol. 56, 381e387. Ohgren, K., Bura, R., Saddler, J., Zacchi, G., 2007. Effect of hemicellulose and lignin removal on enzymatic hydrolysis of steam pretreated corn stover. Bioresour. Technol. 98, 2503e2510. Ohta, K., Beall, D.S., Mejia, J.P., Shanmugam, K.T., Ingram, L.O., 1991. Genetic improvement of Escherichia coli for ethanol production: chromosomal integration of Zymomonas mobilis genes encoding pyruvate decarboxylase and alcohol dehydrogenase II. Appl. Environ. Microbiol. 57, 893e900. Okano, K., Kitagawa, M., Sasaki, Y., Watanabe, T., 2005. Conversion of Japanese red cedar (Cryptomeria japonica) into a feed for ruminants by white-rot basidiomycetes. Anim. Feed Sci. Technol. 120, 235e243.
REFERENCES Olofsson, K., Bertilsson, M., Liden, G., 2008. A short review on SSF - an interesting process option for ethanol production from lignocellulosic feedstocks. Biotechnol. Biofuels 1, 7. Ooshima, H., Aso, K., Harano, Y., Yamamoto., T., 1984. Microwave treatment of cellulosic materials for their enzymatic-hydrolysis. Biotechnol. Lett. 6, 289e294. Ou, M.S., Mohammed, N., Ingram, L.O., Shanmugam, K.T., 2009. Thermophilic Bacillus coagulans requires less cellulases for simultaneous saccharification and fermentation of cellulose to products than mesophilic microbial biocatalysts. Appl. Biochem. Biotechnol. 155, 379e385. Ozkan, P., Sariyar, B., Utkur, F.O., Akman, U., Hortacsu, A., 2005. Metabolic flux analysis of recombinant protein overproduction in Escherichia coli. Biochem. Eng. J. 22, 167e195. Palmqvist, E., Hahn-Hägerdal, B., 2000. Fermentation of lignocellulosic hydrolysates. II: inhibitors and mechanisms of inhibition. Bioresour. Technol. 74, 25e33. Parekh, S., Wayman, M., 1986. Fermentation of cellobiose and wood sugars to ethanol by Candida shehatae and Pichia stipitis. Biotechnol. Lett. 8, 597e600. Patil, K.R., Akesson, M., Nielsen, J., 2004. Use of genome-scale microbial models for metabolic engineering. Curr. Opin. Biotechnol. 15, 64e69. Percival Zhang, Y.H., Himmel, M.E., Mielenz, J.R., 2006. Outlook for cellulase improvement: screening and selection strategies. Biotechnol. Adv. 24, 452e481. Pinto, P.A., Dias, A.A., Fraga, I., Marques, G., Rodrigues, M.A., Colaco, J., Sampaio, A., Bezerra, R.M., 2012. Influence of ligninolytic enzymes on straw saccharification during fungal pretreatment. Bioresour. Technol. 111, 261e267. Piskur, B., Bajc, M., Robek, R., Humar, M., Sinjur, I., Kadunc, A., Oven, P., Rep, G., Al Sayegh Petkovsek, S., Kraigher, H., et al., 2011. Influence of Pleurotus ostreatus inoculation on wood degradation and fungal colonization. Bioresour. Technol. 102, 10611e10617. Playne, M.J., 1984. Increased digestibility of bagasses by pretreatment with alkalis and steam explosion. Biotechnol. Bioeng. 26, 426e433. Prakash, G., Varma, A.J., Prabhune, A., Shouche, Y., Rao, M., 2011. Microbial production of xylitol from D-xylose and sugarcane bagasse hemicellulose using newly isolated thermotolerant yeast Debaryomyces hansenii. Bioresour. Technol. 102, 3304e3308. Qi, W., Chen, C.L., Wang, J.Y., 2011. Reducing sugar-producing bacteria from guts of Tenebrio molitor Linnaeus (yellow mealworm) for lignocellulosic waste minimization. Microbes Environ. 26, 354e359. Rabinovich, M.L., Bolobova, A.V., Vasil’chenko, L.G., 2004. Fungal decomposition of natural aromatic structures and xenobiotics: a review. Appl. Biochem. Microbiol. 40, 1e17. Ramachandriya, K.D., Wilkins, M.R., Delorme, M.J., Zhu, X., Kundiyana, D.K., Atiyeh, H.K., Huhnke, R.L., 2011. Reduction of acetone to isopropanol using producer gas fermenting microbes. Biotechnol. Bioeng. 108, 2330e2338. Raman, B., McKeown, C.K., Rodriguez, M., Brown Jr., S.D., Mielenz, J.R., 2011. Transcriptomic analysis of Clostridium thermocellum ATCC 27405 cellulose fermentation. BMC Microbiol. 11, 134. Rao, D., Gullapalli, P., Yoshihara, A., Jenkinson, S.F., Morimoto, K., Takata, G., Akimitsu, K., Tajima, S., Fleet, G.W., Izumori, K., 2008. Direct production of L-tagatose from L-psicose by Enterobacter aerogenes 230S. J. Biosci. Bioeng. 106, 473e480. Ray, M.J., Leak, D.J., Spanu, P.D., Murphy, R.J., 2010. Brown rot fungal early stage decay mechanism as a biological pretreatment for softwood biomass in biofuel production. Biomass Bioenergy 34, 1257e1262. Richards, D.B., 1954. Physical changes in decaying wood. J. For. 52, 260e265. Richter, H., Qureshi, N., Heger, S., Dien, B., Cotta, M.A., Angenent, L.T., 2012. Prolonged conversion of n-butyrate to n-butanol with Clostridium saccharoperbutylacetonicum in a two-stage continuous culture with in-situ product removal. Biotechnol. Bioeng. 109, 913e921. 91 Roberts, S.B., Gowen, C.M., Brooks, J.P., Fong, S.S., 2010. Genome-scale metabolic analysis of Clostridium thermocellum for bioethanol production. BMC Syst. Biol. 4, 31. Rogers, C.J., Blanford, C.F., Giddens, S.R., Skamnioti, P., Armstrong, F.A., Gurr, S.J., 2009. Designer laccases: a vogue for high-potential fungal enzymes? Trends Biotechnol. 28, 63e72. Rogers, P.L., Lee, K.J., Skotnicki, M.L., Tribe, D.E., 1982. Ethanol production by Zymomonas mobilis. Adv. Biochem. Eng. 23, 37e84. Ruel, K., Barnoud, F., Eriksson, K.E., 1981. Micromorphological and ultrastructural aspects of spruce wood degradation by wild-type Sporotrichum pulverulentum and its cellulase-less mutant Ce144. Holzforschung 35, 157e171. Rumbold, K., van Buijsen, H.J., Gray, V.M., van Groenestijn, J.W., Overkamp, K.M., Slomp, R.S., van der Werf, M.J., Punt, P.J., 2010. Microbial renewable feedstock utilization: a substrate-oriented approach. Bioeng. Bugs 1, 359e366. Rumbold, K., van Buijsen, H.J., Overkamp, K.M., van Groenestijn, J.W., Punt, P.J., van der Werf, M.J., 2009. Microbial production host selection for converting second-generation feedstocks into bioproducts. Microb. Cell Fact 8, 64. Saddler, J.N., Ramos, L.P., Breuil, C., 1993. Steam pretreatment of lignocellulosic residues. In: Saddler, J.N. (Ed.), Bioconversion of Forest and Agricultural Plant Wastes. C.A.B. International, Wallingford, UK, pp. 73e92. Sakon, J., Adney, W.S., Himmel, M.E., Thomas, S.R., Karplus, P.A., 1996. Crystal structure of thermostable family 5 endocellulase E1 from Acidothermus cellulolyticus in complex with cellotetraose. Biochem. 35, 10648e10660. Salgado, J.M., Rodriguez, N., Cortes, S., Dominguez, J.M., 2012. Effect of nutrient supplementation of crude or detoxified concentrated distilled grape marc hemicellulosic hydrolysates on the xylitol production by Debaryomyces hansenii. Prep. Biochem. Biotechnol. 42, 1e14. Salvachua, D., Prieto, A., Lopez-Abelairas, M., Lu-Chau, T., Martinez, A.T., Martinez, M.J., 2011. Fungal pretreatment: an alternative in second-generation ethanol from wheat straw. Bioresour. Technol. 102, 7500e7506. Sanchez, C., 2009. Lignocellulosic residues: biodegradation and bioconversion by fungi. Biotechnol. Adv. 27, 185e194. Saritha, M., Arora, A., Lata, 2012a. Biological pretreatment of lignocellulosic substrates for enhanced delignification and enzymatic digestibility. Indian J. Microbiol. 52, 122e130. Saritha, M., Arora, A., Nain, L., 2012b. Pretreatment of paddy straw with Trametes hirsuta for improved enzymatic saccharification. Bioresour. Technol. 104, 459e465. Sasmal, S., Goud, V.V., Mohanty, K., 2012. Ultrasound assisted lime pretreatment of lignocellulosic biomass toward bioethanol production. Energy Fuels 26, 3777e3784. Sawada, T., Nakamura, Y., Kobayashi, F., Kuwahara, M., Watanabe, T., 1995. Effects of fungal pretreatment and steam explosion pretreatment on enzymatic saccharification of plant biomass. Biotechnol. Bioeng. 48, 719e724. Schilling, J.S., Ai, J., Blanchette, R.A., Duncan, S.M., Filley, T.R., Tschirner, U.W., 2012. Lignocellulose modifications by brown rot fungi and their effects, as pretreatments, on cellulolysis. Bioresour. Technol. 116, 147e154. Schilling, J.S., Tewalt, J.P., Duncan, S.M., 2009. Synergy between pretreatment lignocellulose modifications and saccharification efficiency in two brown rot fungal systems. Appl. Microbiol. Biotechnol. 84, 465e475. Schulein, M., 2000. Protein engineering of cellulases. Biochem. Biophys. Acta. 1543, 239e252. Schwarz, W.H., 2001. The cellulosome and cellulose degradation by anaerobic bacteria. Appl. Microbiol. Biotechnol. 56, 634e649. Sheridan, C., 2009. Making green. Nat. Biotechnol. 27, 1074e1076.
92 5. BIOFUELS AND BIOPRODUCTS PRODUCED THROUGH MICROBIAL CONVERSION OF BIOMASS Shi, J., Chinn, M.S., Sharma-Shivappa, R., 2008. Microbial pretreatment of cotton stalks by solid state cultivation of Phanerochaete chrysosporium for bioethanol production. Bioresour. Technol. 99, 556e6564. Shin, H.D., McClendon, S., Vo, T., Chen, R.R., 2010. Escherichia coli binary culture engineered for direct fermentation of hemicellulose to a biofuel. Appl. Environ. Microbiol. 76, 8150e8159. Shoham, Y., Lamed, R., Bayer, E.A., 1999. The cellulosome concept as an efficient microbial strategy for the degradation of insoluble polysaccharides. Trends Microbiol. 7, 275e281. Shrivastava, B., Nandal, P., Sharma, A., Jain, K.K., Khasa, Y.P., Das, T.K., Mani, V., Kewalramani, N.J., Kundu, S.S., Kuhad, R.C., 2012. Solid state bioconversion of wheat straw into digestible and nutritive ruminant feed by Ganoderma sp. rckk02. Bioresour. Technol. 107, 347e351. Singh, A.P., Butcher, J.A., 1991. Bacterial degradation of wood cell walls: a review of degradation patterns. J. Inst. Wood Sci. 12, 143e157. Singh, A.P., Butcher, S.A., 1985. Degradation of CCA-treated Pinus radiata posts by erosion bacteria. J. Inst. Wood Sci. 10, 140e144. Singh, A.P., Nilsson, T., Daniel, G.F., 1990. Bacterial attack of Pinus sylvestris wood under near-anaerobic conditions. J. Inst. Wood Sci. 11, 237e249. Singh, P., Sulaiman, O., Hashim, R., Rupani, P.F., Peng, L.C., 2010. Biopulping of lignocellulosic material using different fungal species: a review. Rev. Environ. Sci. Biotechnol. 9, 141e151. Singh, P., Suman, A., Tiwari, P., Arya, N., Gaur, A., Shrivastava, A.K., 2008. Biological pretreatment of sugarcane trash for its conversion to fermentable sugars. World J. Microbiol. Biotechnol. 24, 667e673. Steen, E.J., Kang, Y., Bokinsky, G., Hu, Z., Schirmer, A., McClure, A., Del Cardayre, S.B., Keasling, J.D., 2010. Microbial production of fatty-acid-derived fuels and chemicals from plant biomass. Nature 463, 559e562. Steenbakkers, P.J., Harhangi, H.R., Bosscher, M.W., van der Hooft, M.M., Keltjens, J.T., van der Drift, C., Vogels, G.D., op den Camp, H.J., 2003. Beta-Glucosidase in cellulosome of the anaerobic fungus Piromyces sp. strain E2 is a family 3 glycoside hydrolase. Biochem. J. 370, 963e970. Stenberg, K., Tengborg, C., Galbe, M., Zacchi, G., 1998. Optimisation of steam pretreatment of SO2-impregnated mixed softwoods for ethanol production. J. Chem. Technol. Biotechnol. 71, 299e308. Stephanopoulos, G., 1999. Metabolic fluxes and metabolic engineering. Metab. Eng. 1, 1e11. Strohl, W.R., 2001. Biochemical engineering of natural product biosynthesis pathways. Metab. Eng. 3, 4e14. Sun, F.-h., Jiang, L., Yue-xiang, Y., Zhi-ying, Y., Liu, X.-F., 2011. Effect of biological pretreatment with Trametes hirsuta yj9 on enzymatic hydrolysis of corn stover. Int. Biodeterior. Biodegrad. 65, 931e938. Sun, Y., Cheng, J., 2002. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresour. Technol. 83, 1e11. Surwase, S.N., Patil, S.A., Jadhav, S.B., Jadhav, J.P., 2012. Optimization of L-DOPA production by Brevundimonas sp. SGJ using response surface methodology. Microb. Biotechnol. 5, 731e737. Suzuki, M.R., Hunt, C.G., Houtman, C.J., Dalebroux, Z.D., Hammel, K.E., 2006. Fungal hydroquinones contribute to brown rot of wood. Environ. Microbiol. 8, 2214e2223. Taherzadeh, M.J., Karimi, K., 2007. Enzymatic-based hydrolysis processes for ethanol from lignocellulosic materials: a review. Bioresources 2, 707e738. Taherzadeh, M.J., Karimi, K., 2008. Pretreatment of lignocellulosic wastes to improve ethanol and biogas production: a review. Int. J. Mol. Sci. 9, 1621e1651. Tammali, R., Seenayya, G., Reddy, G., 2003. Fermentation of cellulose to acetic acid by Clostridium lentocellum SG6: induction of sporulation and effect of buffering agent on acetic acid production. Lett. Appl. Microbiol. 37, 304e308. Taniguchi, M., Suzuki, H., Watanabe, D., Sakai, K., Hoshino, K., Tanaka, T., 2005. Evaluation of pretreatment with Pleurotus ostreatus for enzymatic hydrolysis of rice straw. J. Biosci. Bioeng. 100, 637e643. Tassinari, T., Macy, C., Spano, L., 1980. Energy requirements and process design considerations in compression-milling pretreatment of cellulosic wastes for enzymatic hydrolysis. Biotechnol. Bioeng. 22, 1689e1705. Taylor, M.P., Eley, K.L., Martin, S., Tuffin, M.I., Burton, S.G., Cowan, D.A., 2009. Thermophilic ethanologenesis: future prospects for second-generation bioethanol production. Trends Biotechnol. 27, 398e405. Teixeira, L.C., Linden, J.C., Schroeder, H.A., 2000. Simultaneous saccharification and cofermentation of peracetic acid-pretreated biomass. Appl. Biochem. Biotechnol. 84-86, 111e127. ten Have, R., Teunissen, P.J., 2001. Oxidative mechanisms involved in lignin degradation by white-rot fungi. Chem. Rev. 101, 3397e3413. Tewalt, J., Schilling, J., 2010. Assessment of saccharification efficacy in the cellulase system of the brown rot fungus Gloeophyllum trabeum. Appl. Microbiol. Biotechnol. 86, 1785e1793. Teymouri, F., Laureano-Perez, L., Alizadeh, H., Dale, B.E., 2004. Ammonia fiber explosion treatment of corn stover. Appl. Biochem. Biotechnol. 113-116, 951e963. Teymouri, F., Laureano-Perez, L., Alizadeh, H., Dale, B.E., 2005. Optimization of the ammonia fiber explosion (AFEX) treatment parameters for enzymatic hydrolysis of corn stover. Bioresour. Technol. 96, 2014e2018. Thiru, M., Sankh, S., Rangaswamy, V., 2011. Process for biodiesel production from Cryptococcus curvatus. Bioresour. Technol. 102, 10436e10440. Tomme, P., Van Tilbeurgh, H., Pettersson, G., Van Damme, J., Vandekerckhove, J., Knowles, J., Teeri, T., Claeyssens, M., 1988. Studies of the cellulolytic system of Trichoderma reesei QM 9414. Analysis of domain function in two cellobiohydrolases by limited proteolysis. Eur. J. Biochem. 170, 575e581. Tran, D.T., Yeh, K.L., Chen, C.L., Chang, J.S., 2012. Enzymatic transesterification of microalgal oil from Chlorella vulgaris ESP-31 for biodiesel synthesis using immobilized Burkholderia lipase. Bioresour. Technol. 108, 119e127. Trinh, C.T., Srienc, F., 2009. Metabolic engineering of Escherichia coli for efficient conversion of glycerol to ethanol. Appl. Environ. Microbiol. 75, 6696e6705. Vaidya, A., Singh, T., 2012. Pre-treatment of Pinus radiata substrates by basidiomycetes fungi to enhance enzymatic hydrolysis. Biotechnol. Lett. 34, 1263e1267. van Zyl, W.H., den Haan, R., McBride, J.E., 2007. Consolidated bioprocessing for bioethanol production using Saccharomyces cerevisiae. Adv. Biochem. Eng. Biotechnol. 108, 205e235. Verma, S., Ray, A.K., De, B.K., 2010. Bioconversion of heptanal to heptanol by Saccharomyces cerevisiae. Yeast 27, 269e275. Wan, C., Li, Y., 2010. Microbial pretreatment of corn stover with Ceriporiopsis subvermispora for enzymatic hydrolysis and ethanol production. Bioresour. Technol. 101, 6398e6403. Wan, C., Li, Y., 2011. Effectiveness of microbial pretreatment by Ceriporiopsis subvermispora on different biomass feedstocks. Bioresour. Technol. 102, 7507e7512. Wang, J., Sun, Q., Gao, P., Wang, J.F., Xu, C., Sun, Q.L., 2010. Bioconversion of glycyrrhizinic acid in liquorice into 18-beta-glycyrrhetinic acid by Aspergillus parasiticus speare BGB. Prikl. Biokhim. Mikrobiol. 46, 462e466. Wang, W., Yuan, T., Wang, K., Cui, B., Dai, Y., 2012. Combination of biological pretreatment with liquid hot water pretreatment to
REFERENCES enhance enzymatic hydrolysis of Populus tomentosa. Bioresour. Technol. 107, 282e286. Warnick, T.A., Methe, B.A., Leschine, S.B., 2002. Clostridium phytofermentans sp. nov., a cellulolytic mesophile from forest soil. Int. J. Syst. Evol. Microbiol. 52, 1155e1160. Wawrik, B., Mendivelso, M., Parisi, V.A., Suflita, J.M., Davidova, I.A., Marks, C.R., Van Nostrand, J.D., Liang, Y., Zhou, J., Huizinga, B.J., et al., 2012. Field and laboratory studies on the bioconversion of coal to methane in the San Juan Basin. FEMS Microbiol. Ecol. 81, 26e42. Weber, C., Farwick, A., Benisch, F., Brat, D., Dietz, H., Subtil, T., Boles, E., 2010. Trends and challenges in the microbial production of lignocellulosic bioalcohol fuels. Appl. Microbiol. Biotechnol. 87, 1303e1315. Wilcox, W.W., 1978. Review of literature on the effects of early stages of decay on wood strength. Wood Fiber 9, 252e257. Wilson, D.B., 2004. Studies of Thermobifida fusca plant cell wall degrading enzymes. Chem. Rec. 4, 72e82. Wither, S.G., 2001. Mechanism of glycosyl transferase and hydrolases. Carbohydr. Polym. 44, 325e337. Wood, T.M., Saddler, J.N., 1988. Increasing the availability of cellulose in biomass materials. Methods Enzymol. 160, 3e11. Wu, J., Zhang, X., Wan, J., Ma, F., Tang, Y., 2011. Production of fiberboard using corn stalk pretreated with white-rot fungus Trametes hirsute by hot pressing without adhesive. Bioresour. Technol. 102, 11258e11261. Wu, Z., Lee, Y.Y., 1998. Nonisothermal simultaneous saccharification and fermentation for direct conversion of lignocellulosic biomass to ethanol. Appl. Biochem. Biotechnol. 7072, 479e492. Xu, Q., Singh, A., Himmel, M.E., 2009. Perspectives and new directions for the production of bioethanol using consolidated bioprocessing of lignocellulose. Curr. Opin. Biotechnol. 20, 364e371. Yaghoubi, K., Pazouki, M., Shojaosadati, S.A., 2008. Variable optimization for biopulping of agricultural residues by Ceriporiopsis subvermispora. Bioresour. Technol. 99, 4321e4328. Yamagishi, K., Kimura, T., Watanabe, T., 2011. Treatment of rice straw with selected Cyathus stercoreus strains to improve enzymatic saccharification. Bioresour. Technol. 102, 6937e6943. Yanase, S., Hasunuma, T., Yamada, R., Tanaka, T., Ogino, C., Fukuda, H., Kondo, A., 2010a. Direct ethanol production from cellulosic materials at high temperature using the thermotolerant yeast Kluyveromyces marxianus displaying cellulolytic enzymes. Appl. Microbiol. Biotechnol. 88, 381e388. Yanase, S., Yamada, R., Kaneko, S., Noda, H., Hasunuma, T., Tanaka, T., Ogino, C., Fukuda, H., Kondo, A., 2010b. Ethanol production from cellulosic materials using cellulase-expressing yeast. Biotechnol. J. 5, 449e455. Yang, H.H., Effland, M.J., Kirk, T.K., 1980. Factors influencing fungal degradation of lignin in a representative lignocellulosic, thermomechanical pulp. Biotechnol. Bioeng. 22, 5e77. 93 Yomano, L.P., York, S.W., Ingram, L.O., 1998. Isolation and characterization of ethanol-tolerant mutants of Escherichia coli KO11 for fuel ethanol production. J. Ind. Microbiol. Biotechnol. 20, 132e138. Yomano, L.P., York, S.W., Zhou, S., Shanmugam, K.T., Ingram, L.O., 2008. Re-engineering Escherichia coli for ethanol production. Biotechnol. Lett. 30, 2097e2103. Yuan, X., Li, P., Wang, H., Wang, X., Cheng, X., Cui, Z., 2011. Enhancing the anaerobic digestion of corn stalks using composite microbial pretreatment. J. Microbiol. Biotechnol. 21, 746e752. Zeng, Y., Yang, X., Yu, H., Zhang, X., Ma, F., 2011. Comparative studies on thermochemical characterization of corn stover pretreated by white-rot and brown-rot fungi. J. Agric. Food Chem. 59, 9965e9971. Zhang, J., Geng, A., Yao, C., Lu, Y., Li, Q., 2012a. Xylitol production from D-xylose and horticultural waste hemicellulosic hydrolysate by a new isolate of Candida athensensis SB18. Bioresour. Technol. 105, 134e141. Zhang, Q., Zhang, W., Lin, C., Xu, X., Shen, Z., 2012b. Expression of an Acidothermus cellulolyticus endoglucanase in transgenic rice seeds. Protein Expression Purif. 82, 279e283. Zhang, Y., Song, L., Gao, Q., Yu, S.M., Li, L., Gao, N.F., 2012c. The twostep biotransformation of monosodium glutamate to GABA by Lactobacillus brevis growing and resting cells. Appl. Microbiol. Biotechnol. 94, 1619e1627. Zhang, Y., Zhu, Y., Zhu, Y., Li, Y., 2009. The importance of engineering physiological functionality into microbes. Trends Biotechnol. 12, 664e672. Zhang, Y.H., Lynd, L.R., 2005. Cellulose utilization by Clostridium thermocellum: bioenergetics and hydrolysis product assimilation. Proc. Natl. Acad. Sci. USA 102, 7321e7325. Zhao, L., Cao, G.L., Wang, A.J., Ren, H.Y., Dong, D., Liu, Z.N., Guan, X.Y., Xu, C.J., Ren, N.Q., 2012. Fungal pretreatment of cornstalk with Phanerochaete chrysosporium for enhancing enzymatic saccharification and hydrogen production. Bioresour. Technol. 114, 365e369. Zhao, X., Cheng, K., Liu, D., 2009. Organosolv pretreatment of lignocellulosic biomass for enzymatic hydrolysis. Appl. Microbiol. Biotechnol. 82, 815e827. Zheng, Y., Lin, H.-M., Wen, J., Cao, N., Yu, X., Tsao, G.T., 1995. Supercritical carbon dioxide explosion as a pretreatment for cellulose hydrolysis. Biotechnol. Lett. 17, 845e850. Zheng, Y., Lin, H., Tsao, G.T., 1998. Pretreatment for cellulose hydrolysis by carbon dioxide explosion. Biotechnol. Prog. 14, 890e896. Zhong, W., Zhang, Z., Luo, Y., Sun, S., Qiao, W., Xiao, M., 2011. Effect of biological pretreatments in enhancing corn straw biogas production. Bioresour. Technol. 102, 11177e11182. Zhu, Y., Lee, Y.Y., Elander, R.T., 2007. Conversion of aqueous ammonia-treated corn stover to lactic acid by simultaneous saccharification and cofermentation. Appl. Biochem. Biotechnol. 137-140, 721e738.
C H A P T E R 6 Databases for Bioenergy-Related Enzymes Yanbin Yin Department of Biological Sciences, Northern Illinois University, DeKalb, IL, USA email: yyin@niu.edu O U T L I N E Plant Biomass 95 Bioenergy-Related Enzymes and Regulation 96 Databases and Web Servers CAZy Database CAT and dbCAN FOLy Database 98 98 101 102 Purdue Cell Wall Genomics and UC-Riverside Cell Wall Navigator Databases Plant Coexpression Network Databases: PlaNet and ATTED 102 Future Perspectives 103 References 103 PLANT BIOMASS Note that unlike celluloses, hemicelluloses and pectins both refer to a collection of complex polysaccharides mostly with side chains. Hemicelluloses contain four major groups: xyloglucans, mannans, xylans and mixed-linkage glucans, while pectins contain three major groups: galacturonans, rhamnogalacturonan I and rhamnogalaturonan II. Each of the groups of hemicelluloses and pectins do not refer to a single type of polysaccharides; they often refer to polysaccharides with the same backbone structure (sugars and linkages) while with different side chains. Due to this reason, all these biopolymers are cross-linked and interwoven (Somerville et al., 2004; Himmel et al., 2007) to form very complex and heterogeneous cell wall structures. Particularly celluloses are wrapped by hemicelluloses and buried in a lignin network and not accessible to enzymes so that the degradation efficiency is very low if no costly chemical pretreatment is applied. Although celluloses are simple polymer of glucose linked by beta-1,4,-glucosidic bond, the complexity of chemical compositions of hemicelluloses and pectins is remarkably high (Somerville et al., 2004). The reasons are as follows: (1) there are 14 different monosaccharaides (sugar units) found in hemicelluloses and pectins (Pauly and Keegstra, 2008b); (2) the possible glycosidic The major components of plant biomass are carbohydrate-rich cell walls, composed of different biopolymers such as polysaccharides and lignins as well as some minor wall structural glycoproteins (Somerville et al., 2004). Biomass used for biofuel production is primarily derived from secondary cell walls. For example, wood cells from poplar trees contain a thin layer of primary cell walls and multiple layers of much thicker and tougher secondary cell walls. All plant cells of different tissues have primary cell walls while only in developed cells (stopped growing) secondary cell walls appear (Cosgrove, 2005). The chemical compositions in primary and secondary cell walls differ significantly (Mohnen et al., 2008). The primary cell wall contains no lignins and the polysaccharides include celluloses, hemicelluloses (primarily xyloglucans and mannans in dicots and xylans in monocots) and pectins. However, in the secondary cell walls, there are higher percentage of celluloses, different hemicellulosic polysaccharides (primarily xylans in both dicots and monocots) and lignins. For example, wood secondary cell walls contain 35e50% celluloses, 25e30% hemicelluloses (mostly xylans) and 15e30% lignins (Himmel et al., 2007). Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00006-1 102 95 Copyright Ó 2014 Elsevier B.V. All rights reserved.
96 6. DATABASES FOR BIOENERGY-RELATED ENZYMES linkages formed between two sugars are extremely diverse as theoretically they can be connected between any hydroxyl group of two sugars and (3) sugars in the polysaccharides can be further modified by, e.g. methylation, acetylation or esterification. Lignins, however, are complex heterogeneous phenolic polymers and chemically very distinct from polysaccharides. They are formed by three major monomers: hydroxyphenyl (H), guaiacyl (G) and syringyl (S) units, which are derived from coumaryl alcohol, coniferyl alcohol and sinapyl alcohol, respectively (Boerjan et al., 2003). All the biopolymers in plant cell walls are cross-linked and interwoven (Somerville et al., 2004; Himmel et al., 2007) to form very complex and heterogeneous structures, which is believed to enable cell walls recalcitrant to enzymatic degradation. Besides, cell wall compositions and structures also vary from tissue to tissue. The reason is because the cell wall biosynthetic enzymes are differentially regulated and expressed in different tissues. Furthermore, different plants, especially those of distant evolutionary clades, have very distinct cell wall biopolymer compositions. For example, grasses generally have significantly higher percentage of xylans than trees (Pauly and Keegstra, 2008a). For biofuel production, polysaccharides especially celluloses are favored as their degradation releases fermentable sugars. Lignins, however, are phenolic polymers and chemically distinct from polysaccharides, not giving rise to sugars and meant to be removed in the biofuel production. In order to develop transgenic plants with modified cell wall compositions (i.e. higher cellulose and lower lignin content), we need a better understanding of how plant cell wall polysaccharides and lignins are synthesized. BIOENERGY-RELATED ENZYMES AND REGULATION Because of the overwhelming complexity of cell wall polymers in terms of their chemical compositions, linkages and structures, plant biomass formation and microbial degradation involve a surprisingly large number of genes in plant and microbes, respectively. For example, presumably every different glycosidic bond in the polysaccharides will be formed using a different enzyme. It is estimated that w10% of genes in Arabidopsis genome are involved in cell wall synthesis and modification (Yong et al., 2005), which account for w2000 genes encoding enzymes for sugar and lignin precursor synthesis, polysaccharide and lignin synthesis and modification, lignin-polysaccharide cross-linking, transcription factors (TFs)and signaling proteins, etc. The most important enzymes are clearly those involved in polysaccharide synthesis and lignin synthesis. To form polysaccharides, glycosyltransferases (GTs) take the activated sugar donors, nucleoside diphosphate sugars (NDP-sugars), as the substrates to build glycosidic bonds between two sugars. Except for celluloses, other cell wall polysaccharides are mostly synthesized in Golgi apparatus, where GTs, NDP-sugar biosynthetic enzymes and sugar transporters are located and work together. Glycoside hydrolases (GHs), on the other hand, are used to break glycosidic bonds through hydrolysis reactions to release sugars from polysaccharides. In plants this is often used to modify existing polysaccharides, e.g. when plant cells are growing, while in microbes GHs are the most critical enzymes degrading plant biomass. Clearly not all GH and GT enzymes are involved in cell wall polysaccharide metabolism, as many of them are involved in metabolism of storage polysaccharides, glycoproteins, glycolipids and other glycol-conjugates that are not relevant to plant cell walls. All GT and GH enzymes are categorized by the CAZy (Carbohydrate Active enZyme) database (CAZyDB) (Cantarel et al., 2009), which provides a general classification scheme for all carbohydrate active enzymes (CAZymes) and is widely accepted by the carbohydrate research community. So far there are a limited number of enzymes biochemically or genetically characterized to be involved in plant cell wall synthesis or modification, many of which belong to some large GH and GT families. For example, the GT2 family is known to include cellulose synthases and some hemicellulose backbone synthases (Lerouxel et al., 2006), such as mannan synthases (Dhugga et al., 2004; Liepman et al., 2005), putative xyloglucan synthases (Cocuron et al., 2007), and mixed linkage glucan synthases (Burton et al., 2006). With respect to the synthesis of xylan, the most abundant hemicellulose, proteins of GT43, GT47 and GT8 are likely to be involved (Zhong et al., 2005; Brown et al., 2007; Lee et al., 2007; Pena et al., 2007; Persson et al., 2007; York and O’Neill, 2008; Brown et al., 2009; Wu et al., 2009). Some of these known cell wall-related CAZyme families are included in Purdue’s Cell Wall genomics database (Yong et al., 2005) and UC-Riverside’s (UCR) Cell Wall Navigator database (Girke et al., 2004), and more families are discussed in the literature or to be characterized in terms of their roles in biomass-related polysaccharide formation and degradation. For example, a few recent papers (Scheller and Ulvskov, 2010; Driouich et al., 2012) and a book (Ulvskov, 2011) updated our knowledge about the GT family members involved in cell wall synthesis: GT2, 8, 31, 34, 37, 43, 47, 61, 64, 75, 77, while there must be more GT families not included and to be identified as cell wall related (CWR), e.g. GT92 (Liwanag et al., 2012).
BIOENERGY-RELATED ENZYMES AND REGULATION Lignins are complex heterogeneous polymers with lots of aromatic rings. The monolignol synthesis pathway that starts from phenylalanine to synthesize G, S and H units has been relatively well known, with about 10 gene families characterized encoding most of the enzymes in the pathway (Humphreys and Chapple, 2002; Boerjan et al., 2003; Vanholme et al., 2008; Xu et al., 2009; Zhong and Ye, 2009; Li and Chapple, 2010; Weng and Chapple, 2010; Carpita, 2012). All these lignin synthesis-related enzymes have been extensively reviewed in the literature and are included in Purdue’s Cell Wall genomics database. Transporting the units to the outside of the cell and assembling them into lignin polymers are less understood but some candidate transporters and two major enzyme families, peroxidase and laccase, are suggested (McCaig et al., 2005; Liu et al., 2011; Zhang et al., 2011; Alejandro et al., 2012; Carpita, 2012; Handford et al., 2012; Sibout and Hofte, 2012). As with all other metabolic pathways, biomass formation and degradation are also under strict regulation. However, compared to enzymatic activities, regulatory mechanism is even more difficult to elucidate. In plants, only a handful of TFs are known to regulate cell wall synthesis. The most studied process is the regulation of lignin biosynthesis (Zhong and Ye, 2009; Zhong et al., 2010; Zhao and Dixon, 2011; Wang and Dixon, 2012). TF families NAC, WRKY, and MYB among a few others have been shown to directly or indirectly control the monolignol synthesis. Some of the TF family members are global regulators that regulate the entire secondary cell wall synthesis, including the synthesis of celluloses and xylans, suggesting that the different biopolymers in biomass are not synthesized independently but in a coordinated way. On the other hand, genetically modifying the regulation of cell wall biosynthesis represents a very promising way to improve the desired traits of bioenergy crops. For example, Wang et al. showed that a mutation found in a WRKY TF could rewire the regulatory network of secondary cell wall synthesis and improve 50% of the biomass production in Arabidopsis (Wang et al., 2010). Similarly, micro ribonucleic acids (miRNAs) are also excellent targets for controlling the regulation of cell wall synthesis (Fu et al., 2012), which is less discussed in the literature. Clearly looking for novel transcription regulators, either TFs and miRNAs, and further building the regulatory network of cell wall synthesis is the ultimate goal for the elucidation of the mechanism of biomass formation. Recently a few plant journals published special issues on plant cell wall researches: Plant Physiology (McCann and Rose, 2010), Current Opinion in Plant Biology (Pauly and Keegstra, 2008b), Frontiers in Plant Science (Debolt and Estevez, 2012) and Molecular Plant (has a cell wall biology category). Particularly, a number of review articles published in these special issues and a 97 few book chapters (Table 6.1) gave overviews of latest progress in a specific area of cell wall research and are very useful for pointing to the original research papers reporting the characterization of specific CWR genes. In terms of degradation, cell wall polysaccharides are degraded by microbial GHs and other CAZymes that are defined and categorized in CAZyDB. Lignins are mostly degraded by microbes too particularly by certain fungi (Dashtban et al., 2009). Enzymes involved in the degradation include fungal laccases and peroxidases, which are categorized in the FOLy (fungal oxidative enzymes) database (Levasseur et al., 2008). Note that these two families are not restricted to fungi. Instead they both belong to large protein families having many homologs in various organisms such as plants, animals and bacteria, bearing slightly different biochemical activities (Welinder, 1992). As mentioned above, these enzyme families are also used for lignin polymerization TABLE 6.1 Selected Publications for CWR Genes Category Publications Lignins (Humphreys and Chapple, 2002; Boerjan et al., 2003; Vanholme et al., 2008; Xu et al., 2009; Zhong and Ye, 2009; Li and Chapple, 2010; Weng and Chapple, 2010; Carpita, 2012) Celluloses (Somerville, 2006; Joshi and Mansfield, 2007; Gu and Somerville, 2010; Endler and Persson, 2011; Lei et al., 2012) Xylans (York and O’Neill, 2008; Carpita, 2012; Doering et al., 2012) Hemicelluloses (Sandhu et al., 2009; Scheller and Ulvskov, 2010; Carpita, 2012; Driouich et al., 2012) Pectins (Mohnen, 2008; Harholt et al., 2010; Driouich et al., 2012) NDP-sugars (Reiter and; Vanzin, 2001; Seifert, 2004; Reiter, 2008; Bar-Peled and O’Neill, 2011) TFs* (Zhong and Ye, 2009; Zhong et al., 2010; Zhao and Dixon, 2011; Carpita, 2012; Wang and Dixon, 2012) Transporters (Liu et al., 2011; Zhang et al., 2011; Alejandro et al., 2012; Carpita, 2012; Handford et al., 2012; Sibout and Hofte, 2012) miRNAs (Sun; Sun et al.; Fu et al., 2012) GTs* (Yokoyama and Nishitani, 2004; Scheller and Ulvskov, 2010; Driouich et al., 2012; Harholt et al., 2012) Other CAZymes (Yokoyama and Nishitani, 2004; Minic, 2008) DUF* families (Carpita, 2012; Hansen et al., 2012; Harholt et al., 2012) Bioinformatics (Egelund et al., 2004; Yong et al., 2005; Penning et al., 2009; Michel et al., 2010) * TF, transcription factor; GT, glycosyltranferase; DUF, domain of undefined function.
98 6. DATABASES FOR BIOENERGY-RELATED ENZYMES in plants. There is also increasing evidence to show that such enzymes are also used for lignin degradation in bacteria (Claus, 2003; Li et al., 2009; Bugg et al., 2011a, 2011b). Notably many cell wall biosynthesis-related gene families are also rooted in bacteria (Royo et al., 2000; Nobles and Brown, 2004; Emiliani et al., 2009; Yin et al., 2009; Weng and Chapple, 2010; Yin et al., 2010, 2011; Popper et al., 2011). In other words, although carbohydrate and lignin-rich plant cell walls are almost unique to plants, the biosynthetic machinery has evolved from ancient gene families that were already present in early prokaryotes. On the degradation side, microbes are responsible for breaking down biomass, while plants also contain homologs of many microbial degrading enzymes such GHs and peroxidases. Obviously plants also inherited these enzymes for different purposes: modify existing polysaccharide or complete the lignification process. Similarly, less is known about the regulation of enzymes for the polysaccharide and lignin degradation in microbes than the enzymes themselves. As opposed to plants, microbes involved in biomass degradation are more taxonomically distributed spanning from eukaryotic fungi (e.g. Neurospora crassa) to prokaryotic bacteria (Clostridium thermocellum). As a result, the regulation systems in these divergent organisms are often not very conserved, e.g. many of the TFs found in fungi are not present in bacteria and vice versa. Furthermore, there are numerous model microbes used for bioenergy research and the transcription regulators regulating cellulases, hemicellulases or ligninases are highly dispersed in the literature, e.g. (Aro et al., 2005; Portnoy et al., 2011; Coradetti et al., 2012; Sun et al., 2012). All these make the curation and annotation of the regulators and targeting cis elements to be very difficult. Recently, global gene expression data (e.g. microarray) and other omics data have been generated to help study the regulation of biomass degradation (Nataf et al., 2010; Raman et al., 2011; Riederer et al., 2011; Yang et al., 2012), which represents the future trend of understanding biofuel production at the systems biology level. Similar to regulators for cell wall synthesis, there is a lack of web-based bioinformatics databases to include the regulatory genes for bioenergy-related degradation enzymes. DATABASES AND WEB SERVERS Our knowledge about cell wall biosynthesis and biodegradation has steadily increased in the past decades, although there is still a long way to go to fully understand these extremely complex processes. So far our knowledge is at best fragmented and characterized genes are often from various organisms and dispersed in the literature. Therefore, dedicated, centralized and frequently updated databases of bioenergy-related genes are crucial to guide the development of transgenic biofuel crops and the annotation of newly sequenced metagenomes/genomes to look for novel enzymes. Like other biology research areas, bioenergy research has also been benefiting from bioinformatics. Table 6.2 provides a list of bioinformatics databases and webbased resources developed specifically for bioenergyrelated enzymes as well as some general bioinformatics web resources that are valuable for bioenergy research. Here we offer a summary of a few selected resources that are particularly useful. CAZy Database For bioenergy research, CAZymes are obviously the most important enzymes. The CAZyDB team started to classify and annotate CAZyme proteins from GenBank, UniProt and PDB to protein families since the early 1990s. It is the original database that defined over 300 CAZyme protein families throughout the past 20 years and the most comprehensive database providing high-quality manual annotation by extracting associated knowledge from the literature (Cantarel et al., 2009). Its Web site regularly updates every few weeks, mainly by assigning new proteins in public databases to existing CAZyme families by sequence similarity search or creating new families if there are new biochemically characterized CAZyme proteins (supported by published papers) that do not belong to existing CAZyme families. Sometimes the functional annotation information (e.g. known activities) for some families is also updated if relevant literature came out. Currently the database comprises five classes of protein families: in addition to GHs and GTs, there are three other classes, carbohydrate esterases (CEs), polysaccharide lyases (PLs) and carbohydrate binding modules (CBMs). As aforementioned, GTs are used for building polysaccharides or glycolconjugates, while GH for breaking them. CE and PL are also used for breaking carbohydrate molecules while using different mechanisms or cutting different chemical bonds. CBMs, as indicated by names, are structural modules used for recognizing and binding different sugars. Currently the five classes contain 94 GT families, 131 GH families, 16 CE families, 22 PL families and 64 CBM families. Each class also has an unclassified family, meaning proteins are annotated to belong to a certain class but are not able to be assigned to any existing families in that class. Each family is named with the class name followed by a sequential number, e.g. GT2. Note such name does not indicate any biochemical activity of each family. The reason is that these families are defined solely based on sequence similarity: there are many cases that one family contains
99 DATABASES AND WEB SERVERS TABLE 6.2 Bioenergy-Related Databases General Plant biomass formation Database Description References CAZyDB General carbohydrate active enzyme database; a classification scheme with five classes (GH, GT, CE, PL and CBM) and over 300 families; supported by the biochemical literature; links to proteins in GenBank, UniProt and structures in PDB; subfamilies for selected families; Enzyme commission annotation and biochemically curated function annotation (Cantarel et al., 2009) CAT CAZyme analysis toolkit; allow CAZyme annotation of user submitted data using BLAST (basic local alignment search tool) and Pfam-based search (Park et al., 2010) dbCAN Web server for automated CAZyme annotation; allow submission of predicted protein sequences from newly sequenced genomes and metagenomes and return a table and graphical diagram to show the matched CAZyme domains; users could also download HMMs representing all CAZyme domains to perform annotation locally by running HMMER3 (hmmer.org) (Yin et al., 2012) Survey of databases for cell wall synthesis A paper reviewed nine public databases (Cao et al., 2010) XTH World Xyloglucosyl transferase gene nomenclature; gene structure and literature in Arabidopsis, rice and tomato MAIZEWALL Cell wall gene catalogue, expression data and literature in maize (Guillaumie et al., 2007) coreCarb PlaNet family tool; Arabidopsis CAZyme proteins; sequence, expression, regulon (coexpressed genes), phylogeny data (Mutwil et al., 2009) GolgiP Web server for prediction of Golgi localized proteins (Chou et al., 2010) Purdue cell wall genomics General classification scheme for plant cell wall synthesis; with mutants and spectrotype information; also including lignin synthesis and modification and NDP-sugar synthesis genes and signaling proteins etc.; phylogeny for gene families in Arabidopsis, rice, maize and sorghum; literature (Yong et al., 2005) UC-Riverside cell wall navigator Similar classification scheme for plant cell wall synthesis proteins to Purdue cell wall database but not including lignin-related and signaling proteins; including sequence, literature and microarray expression data; primarily for Arabidopsis and rice, but also generally from UniProt (Girke et al., 2004) Stanford cellulose Designed for the CesA (cellulose synthase) superfamily and homologs; deprecated web site (Richmond and Somerville, 2000) Rice GT Part of the rice phylogenomics database; rice GT protein phylogeny, sequence, expression, mutants, ortholog, BLAST (Cao et al., 2008) Csl families Web site supplemental to (Yin et al., 2009); protein sequences, alignments and phylogeny of the CesA superfamily in fully sequenced plant and algal genomes (Yin et al., 2009) pDAWG CAZymes in fully sequenced plant and algal genomes based on search against CAZyDB; phylogeny, predicted subcellular localization and proteineprotein interaction data; BLAST (Mao et al., 2009) PPI for cell wall Proteineprotein interaction graphs for cell wall-related proteins in Arabidopsis (Zhou et al., 2010a) PlaNet General coexpression database for seven plant organisms and comparison among them; highest reciprocal rank based coexpression and clustering using Heuristic Cluster Chiseling Algorithm (HCCA, Mutwil et al.); queried gene-centered display of coexpression network (Mutwil et al., 2011) (Continued)
100 6. DATABASES FOR BIOENERGY-RELATED ENZYMES TABLE 6.2 Bioenergy-Related Databasesdcont’d Biomass degradation Misc Database Description References Cell wall coexpression database Biclustering coexpression analysis of cell wall-related genes from Purdue cell wall genomics database; coexpression modules and graphs generated using Cytoscape (Shannon et al., 2003); cisregulatory elements identified in promoter regions of genes of a same module (Wang et al., 2012) ATTED General coexpression database and predicted cis-regulatory elements for Arabidopsis and rice; mutual rank based; also identified conserved coexpression links and referred to proteineprotein interaction data (Obayashi et al., 2009) GAS db Glycosyl hydrolase AnnotationS (GAS) database; GH data identified from UniProt and JGI metagenomes based on CAZyDB and Pfam search; featured with the graphical domain diagrams and comparison between two selected bacteria (Zhou et al., 2010b) FOLyDB Fungal enzymes for lignin degradation; 10 families of lignin oxidases and auxiliary enzymes; proteins from GenBank, UniProt and PDB (Levasseur et al., 2008) PeroxiBase General peroxidase database including peroxidases (EC 1.11.1.x) from over 1000 organisms; lignin-related peroxidases are a subset of the database (Fawal et al., 2012) LccED General laccase database and their homologs in the multicopper oxidase superfamily (Sirim et al., 2011) Biofuel feedstock genomic resource (BFGR) Database of 54 plant organisms with sequenced genomes or significant amount of EST (expressed sequence tag) data; integrated data including expression, ortholog and paralog, pathway prediction, and functional information (Childs et al., 2012) BESC-KB Knowledgebase for the Bioenergy Science Center of DOE; a web portal to a number computational tools and databases dedicated for bioenergy research and developed within the center (Syed et al., 2012) Pathway-genome database of poplar Populus trichocarpa metabolic pathways generated automatically through the Pathway Tool; currently the NDP-sugar biosynthetic pathways were manually curated by experts (Nag et al., 2012) JGI IMG/M Joint Genome Institute’s integrated microbial genomes and metagenomes web site (Markowitz et al., 2012) Phytozome JGI’s plant genome web site; currently most sequenced plant genomes are available in this web site (Goodstein et al., 2012) proteins characterized with different biochemical activities. Recent efforts from the CAZyDB team suggest that further classification of family into subfamilies could be useful as subfamily may contain proteins with the same activity (Stam et al., 2006; Lombard et al., 2010; Aspeborg et al., 2012). CAZyDB’s annotation also evolved in the past 20 years. Among the 327 CAZyme families as of December 2012, there are 10 depreciated families; they were created during the life course of CAZyDB but later were deleted since they were shown not related to carbohydrate metabolism or due to some other reasons. However, to keep the existing nomenclature system unchanged, these family names remain in the system but indicated to be deleted on the Web pages for these families. Other examples include CE10 family, whose Web page was not updated since 2002 because after the family was created it was shown that most CE10 family members do not take carbohydrates as substrate; CBM33 was thought to be a carbohydrate active binding module but later shown likely to be an oxidase family. For a decade, CAZyDB provides an HTML page for each family to list member proteins and associated functional information. In recent updates, CAZyDB added a Web page for each genome, providing a list of GenBank protein accession numbers of that genome together with the CAZyme family assignment for each protein, which is termed “CAZyome” of an organism. So far, there are almost 2400 genomes spanning from eukaryotes to prokaryotes and viruses annotated in CAZyDB. It is said that such genome-scale annotation of CAZyme proteins is done semiautomatically (Coutinho and
DATABASES AND WEB SERVERS Henrissat, 2011). A backend automated domain modulebased search is performed first and then manual curation will be conducted to remove false positives or include false negatives. Obviously this process is rather accurate and of high quality but time consuming because it is done manually and requires expert knowledge. Indeed such genome annotation can only be done by the collaboration with the CAZyDB team, which is often invisible to and out of the control of the users, e.g. people who sequenced the genome. Over the past 10 years, the CAZyDB team has done expert CAZyme annotation for dozens of genome sequencing projects that led to a lot of collaborative genome annotation papers. CAT and dbCAN With more and more bioenergy-related genomes of plants and microbes as well as environmental metagenomes sequenced, there is an urgent need for automated CAZyme annotation. Although such annotation will not reach a quality as accurate as the expert annotation from CAZyDB, it is expected to be much faster and users can control the annotation at their will. Moreover, nowadays all newly sequenced genomes are relying on generic protein domain/family databases such as Pfam (Finn et al., 2006), InterPro (Hunter et al., 2009), and conserved domain database (CDD, Marchler-Bauer et al., 2009) for automatic genome annotation. Clearly annotation from these databases is often too general and too far from the exact function; the precisely actual function still needs to be determined by experimental approaches. However, most genome annotators are still interested in such genome-scale annotation, as it can give them a quick summary about what the genome encodes, how large the gene families are and how that compares to other genomes. In fact, even CAZyDB’s manual annotation (assign proteins to existing CAZyme families) on newly sequenced genomes is unlikely to be 100% correct. Considering every new genome contains a high percentage of proteins that are not experimentally studied, the manual curation is still largely based upon additional bioinformatics analysis such as BLAST search against public protein sequence databases (e.g. UniProt (Bairoch et al., 2005)) and domain databases (e.g. Pfam) and inspection of top matches. With these in mind, automated CAZyme annotation is still very useful, e.g. particularly for a quick and general overview of how many CAZymes and what CAZymes a newly sequenced genome has. Using the annotated CAZyme proteins and classification scheme in CAZyDB as the foundation, two bioinformatics efforts have been published since 2010, both supporting automated CAZyme annotation, given a protein sequence dataset predicted from a genome/metagenome. The CAZyme 101 Analysis Toolkit (CAT) (Park et al., 2010) allows a BLAST search against CAZyme proteins annotated by CAZyDB and also a Pfam domain-based search. The simple BLAST search suffers from the inability to accurately annotate the prevalent multidomain CAZyme proteins. The Pfam domain-based search can solve this problem. These Pfam domains are either given by CAZyDB in the CAZyme family Web pages or identified to correspond to CAZyme family using an association rule built by CAT. However, there are only 142 (46%) of over 300 CAZyme families linked to Pfam domains by CAZyDB. In fact, many of the Pfam domains were originally created after CAZyDB. We recently developed dbCAN (Yin et al., 2012) to define a signature domain model for all CAZyme families. Aside from the 142 CAZyme families annotated with a Pfam domain, we managed to associate other CAZyme families to functional domains in a broader and general protein domain database CDD . This way, we were able to find a CDD domain for 248 CAZyme families. For the remaining families, we performed a literature curation by reading relevant biochemical papers that are linked to these families by CAZyDB. In the end, we extracted the domain regions in all the member proteins annotated in CAZyDB and built a multiple sequence alignment (MSA) for each of the CAZyme family. These MSAs were further processed and represented by hidden Markov models (HMMs), statistical models widely used in the bioinformatics field to represent protein sequence alignments, e.g. by Pfam. As of June 2012, dbCAN has 320 HMMs representing 317 CAZyme families and three cellulosome modules. We provide all these HMMs freely to the public so that they can run domain-based tool hmmscan of the HMMER 3.0 package (hmmer.org) to annotate their genomes/metagenomes in a local computer, exactly the way that people perform Pfam, InterPro or CDD annotation. To help users who do not know how to run hmmscan on a Linux PC, we offer the web server (http://csbl.bmb. uga.edu/dbCAN/annotate.php) so that people can submit their sequences for annotation on the web. The 320 CAZyme family-specific HMMs are our key contribution to the carbohydrate research community and ideally should be included in the general protein domain/family database such as Pfam in the future. In addition to the Web server, dbCAN also provides a database where precomputed CAZyme homologs in a number of protein databases are showed on the Web. Particularly, starting from the 320 dbCAN HMMs, we scanned public metagenome datasets such as NCBIenv-nr, CAMERA (Seshadri et al., 2007), JGI metagenomes (Markowitz et al., 2012), human gut metagenomes (Meta-HIT) (Qin et al., 2010) and cow rumen metagenomes (Hess et al., 2011) as well as plant (Goodstein et al., 2012), bacterial and fungal genomes. Tests on
102 6. DATABASES FOR BIOENERGY-RELATED ENZYMES Arabidopsis thaliana (plant) and C. thermocellum (bacteria) using CAZyDB as the positive set suggest that the automated CAZyme annotation achieved a fairly good accuracy (A. thaliana: sensitivity ¼ 96.3%, precision ¼ 78.8% and average ¼ 87.6%; C. thermocellum: sensitivity ¼ 99.3% and precision ¼ 89.4%). Particularly the sensitivity is over 95% for both organisms, meaning dbCAN annotation tends not to lose true CAZyme proteins. category to support their classification. Proteins of the two databases are primarily from model organisms such as Arabidopsis and rice. However, it is easy to use the sequences as query to search for their homologs and even orthologs in other plants. The protein accessions and classification complied by the two databases serve as an excellent overview of the current achievement made by the entire plant cell wall community in terms of our latest understanding of cell wall synthesis. FOLy Database Inspired by CAZyDB, Levasseur et al. developed a new database named FOLy, for the classification of ligninases in fungi (Levasseur et al., 2008), as these enzymes are critical for breaking down lignins in the biomass but are not included in CAZyDB. Similar to CAZyDB, FOLyDB started from biochemically characterized proteins or structures to recruit homologs from GenBank, UniProt and PDB databases. Based on sequence similarity, three lignin oxidase families and seven lignin degrading auxiliary enzyme families were created, each containing biochemically characterized proteins together with their sequence homologs. Similarly, FOLyDB is featured with expert manual curation of continuingly published literature to include more characterized proteins in order to create new families and populate the database. Like CAZyDB, it is not designed for automated genome annotation but BLAST and Pfam domain-based search against annotated proteins in FOLyDB has been widely used to annotate newly sequenced genomes for ligninases. Purdue Cell Wall Genomics and UC-Riverside Cell Wall Navigator Databases Unlike the above general protein family databases, Purdue’s Cell Wall Genomics (Yong et al., 2005) and UCR’s Cell Wall Navigator (Girke et al., 2004) databases are specifically designed for plant cell wall biosynthesis. As opposed to sequence similarity-based classification, both databases categorize proteins based on their physiological roles in cell wall synthesis. In UCR’s database, there are five categories: monosaccharide synthesis, polysaccharide synthesis, reassembly, structural proteins and glycoprotein synthesis, basically all centered on carbohydrate molecules in cell walls. However, Purdue’s database is even broader with six categories: pathways for substrate generation, polysaccharide synthases, secretion and targeting pathways, assembly/architecture and growth, differentiation and secondary wall formation and signaling and response pathways. Particularly, Purdue’s database also includes lignin synthesis and polymerization proteins as well as signaling proteins involved in cell wall synthesis. Both databases also provide references and the literature associated with each Plant Coexpression Network Databases: PlaNet and ATTED As we mentioned earlier, regulation of biomass formation and degradation is also extremely important. Studying the regulation has been benefited a lot from coexpression analysis of microarray data and recently on high-throughput RNA sequencing data. There are numerous tools, Web based or stand alone, allowing for coexpression analysis with user-submitted genes as query or by general browsing. Many online tools even offer prebuilt coexpression networks, with nodes in the network graphs representing genes and edges representing coexpression relationships. These coexpression networks are very informative and insightful, in terms of suggesting candidate genes involving in the same metabolic pathways or potential regulators of genes of interest to the users (Ruprecht and Persson, 2012). Although there are many such tools available, in plant cell wall field PlaNet (Mutwil et al., 2011) family tools (coreCarb (Mutwil et al., 2009), AraNet, GeneCat) stand out as they were developed by researchers of the cell wall field, have a Web-based interface and have been shown to be effective in suggesting new genes for cell wall synthesis (Mutwil et al., 2009; Ruprecht et al., 2011). PlaNet placed query genes in network graphs of three levels: (1) the coexpressed node vicinity network, containing the query gene and genes coexpressed two steps away and the links among them; (2) a larger coexpression cluster containing the query gene and genes coexpressed, which resulted from running a heuristic clustering algorithm; and (3) the largest meta-network with nodes now representing all coexpression clusters instead of individual genes. PlaNet also has a module called NetworkComparer, allowing a comparative analysis of gene expression networks across seven plant organisms. Such comparative coexpression analysis has recently become very popular as it can help deal with missing data in a single species, reduce false positives identified as coexpressed in a single species, and enable to study the conservation of coexpression network from an evolution perspective (Movahedi et al., 2012). An earlier tool ATTED-II (Obayashi et al., 2009) is also well known, which was developed by plant biologists since 2003 as a database for Arabidopsis tissue-specific
REFERENCES (ATTED) expression. Compared to PlaNet, ATTED-II provides richer annotation and a lot of useful links to external resources in the query gene browse page. It also has a nicer network graph with less genes (top certain amount of genes for better visualization) and gene function information (biologically meaningful gene names) and labeled TFs. Besides, ATTED-II also predicted cis-regulatory elements in the upstream regions of coexpression genes. However, ATTED-II currently includes only two plants: Arabidopsis and rice, and the gene locus page is only available for Arabidopsis. FUTURE PERSPECTIVES As a perspective for the future development of bioenergy-related databases, we ask: what do we need from newly developed databases? Nucleic Acids Research publishes a prestigious annual Database Special Issue since 20 years ago. Most databases published there for a particular class of proteins such as plant TFs (Guo et al., 2008), peroxidases (Fawal et al., 2012), transporters (Saier et al., 2006) and peptidases (Rawlings et al., 2012), all provide the following data or functionalities: (1) a general classification of targeted protein families, manually collected references, a list of characterized proteins curated from the literature and/or predicted member proteins; (2) secondary data derived from further in-depth bioinformatics analysis, such as computer-based functional annotation (e.g. Gene Ontology or protein domain annotation), sequence alignment, phylogenetic trees, predicted protein structures etc.; (3) simple Web-based BLAST search against the sequence database and text search using keywords; (4) long-term maintenance to update regularly with new data; and (5) plenty of documentation such as help, FAQ or tutorial pages. These could be considered as criteria for a good protein family database. Although the plant biomass formation-related databases listed in Table 6.2 are all very useful, none of them have sufficiently integrated various functional omics data. Biologists working on one model plant often want to take advantage of these data to study their interested genes, e.g. investigate fully sequenced plant genomes to look for orthologs, or transcriptome data (microarray and RNA-seq) for expression profiles or look for coexpressed genes and go to the upstream regions for candidate cis-regulatory motifs; all these analyses have to be done using individual bioinformatics tools or servers, which often requires expert knowledge to run or to interpret the results. In addition, many of the databases are outdated and none of them have included all CWR genes. For example, Purdue’s database is an excellent resource, but it does not include many of the 103 newly characterized CWR genes such as TF family NAC, WRKY, MYB members shown to control lignin synthesis; many of the newly characterized CAZyme families such as GT43, 61, 75, transporters for NDPsugars and monolignols; miRNAs; DUF (domain of unknown function) families etc. It includes neither much annotation data nor any search functionalities. Therefore, the future plant CWR gene databases should aim to include all experimentally characterized CWR genes from any organisms, associated sequences and functional descriptions collected from the published literature, e.g. those listed in Table 6.1. Such characterized gene list could be highly useful for annotating sequenced bioenergy plants such as switchgrass, poplar, maize, sorghum and Eucalyptus grandis. The CWR gene repertories for these organisms will be highly valuable for the bioenergy research community as people are trying to select candidate CWR genes to knock down or overexpress for developing transgenic plants in these model organisms. Gene families for CWR genes and other extensive secondary bioinformatics data should also be included in the databases, particularly phylogeny (used to identify orthologs from homologs), predicted cis-regulatory element, conserved coexpression network modules of known CWR genes, expression profiling, coexpressed gene list including noncoding RNAs, genomic location, gene neighborhood, epigenomics, proteineprotein interactions, structures, subcellular locations, single-nucleotide polymorphism, indels, etc. Similar databases should also be developed for plant CAZymes and include the above bioinformatics-derived data types. The reason is that CAZyDB now only covered 2 (A. thaliana and Oryza sativa) of over 40 sequenced plant and green algal genomes, not to mention there are more incomplete genomes and transcriptomes (ESTs and RNA-seq data). References Alejandro, S., Lee, Y., Tohge, T., Sudre, D., Osorio, S., Park, J., Bovet, L., Lee, Y., Geldner, N., Fernie, A.R., Martinoia, E., 2012. AtABCG29 is a monolignol transporter involved in lignin biosynthesis. Curr. Biol. 22, 1207e1212. Aro, N., Pakula, T., Penttila, M., 2005. Transcriptional regulation of plant cell wall degradation by filamentous fungi. FEMS Microbiol. Rev. 29, 719e739. Aspeborg, H., Coutinho, P.M., Wang, Y., Brumer 3rd, H., Henrissat, B., 2012. Evolution, substrate specificity and subfamily classification of glycoside hydrolase family 5 (GH5). BMC Evol. Biol. 12, 186. Bairoch, A., Apweiler, R., Wu, C.H., Barker, W.C., Boeckmann, B., Ferro, S., Gasteiger, E., Huang, H., Lopez, R., Magrane, M., Martin, M.J., Natale, D.A., O’Donovan, C., Redaschi, N., Yeh, L.S., 2005. The universal protein resource (UniProt). Nucleic Acids Res. 33, D154eD159. Bar-Peled, M., O’Neill, M.A., 2011. Plant nucleotide sugar formation, interconversion, and salvage by sugar recycling. Annu. Rev. Plant Biol.
104 6. DATABASES FOR BIOENERGY-RELATED ENZYMES Boerjan, W., Ralph, J., Baucher, M., 2003. Lignin biosynthesis. Annu. Rev. Plant Biol. 54, 519e546. Brown, D.M., Goubet, F., Wong, V.W., Goodacre, R., Stephens, E., Dupree, P., Turner, S.R., 2007. Comparison of five xylan synthesis mutants reveals new insight into the mechanisms of xylan synthesis. Plant J. 52, 1154e1168. Brown, D.M., Zhang, Z., Stephens, E., Dupree, P., Turner, S.R., 2009. Characterization of IRX10 and IRX10-like reveals an essential role in glucuronoxylan biosynthesis in Arabidopsis. Plant J. 57, 732e746. Bugg, T.D., Ahmad, M., Hardiman, E.M., Rahmanpour, R., 2011a. Pathways for degradation of lignin in bacteria and fungi. Nat. Prod. Rep. 28, 1883e1896. Bugg, T.D., Ahmad, M., Hardiman, E.M., Singh, R., 2011b. The emerging role for bacteria in lignin degradation and bio-product formation. Curr. Opin. Biotechnol. 22, 394e400. Burton, R.A., Wilson, S.M., Hrmova, M., Harvey, A.J., Shirley, N.J., Stone, B.A., Newbigin, E.J., Bacic, A., Fincher, G.B., 2006. Cellulose synthase-like CslF genes mediate the synthesis of cell wall (1,3;1,4)beta-D-glucans. Science 311, 1940e1942. Cantarel, B.L., Coutinho, P.M., Rancurel, C., Bernard, T., Lombard, V., Henrissat, B., 2009. The carbohydrate-active enZymes database (CAZy): an expert resource for glycogenomics. Nucleic Acids Res. 37, D233eD238. Cao, P.J., Bartley, L.E., Jung, K.H., Ronald, P.C., 2008. Construction of a rice glycosyltransferase phylogenomic database and identification of rice-diverged glycosyltransferases. Mol. Plant 1, 858e877. Cao, P.J., Jung, K.H., Ronald, P.C., 2010. A survey of databases for analysis of plant cell wall-related enzymes. BioEnergy Res. 3, 108e114. Carpita, N.C., 2012. Progress in the biological synthesis of the plant cell wall: new ideas for improving biomass for bioenergy. Curr. Opin. Biotechnol. 23, 330e337. Childs, K.L., Konganti, K., Buell, C.R., 2012. The biofuel feedstock genomics resource: a web-based portal and database to enable functional genomics of plant biofuel feedstock species. Database (Oxford) bar061. Chou, W.C., Yin, Y., Xu, Y., 2010. GolgiP: prediction of golgi-resident proteins in plants. Bioinformatics 26, 2464e2465. Claus, H., 2003. Laccases and their occurrence in prokaryotes. Arch. Microbiol. 179, 145e150. Cocuron, J.C., Lerouxel, O., Drakakaki, G., Alonso, A.P., Liepman, A.H., Keegstra, K., Raikhel, N., Wilkerson, C.G., 2007. A gene from the cellulose synthase-like C family encodes a beta-1,4 glucan synthase. Proc. Natl. Acad. Sci. USA 104, 8550e8555. Coradetti, S.T., Craig, J.P., Xiong, Y., Shock, T., Tian, C., Glass, N.L., 2012. Conserved and essential transcription factors for cellulase gene expression in ascomycete fungi. Proc. Natl. Acad. Sci. USA 109, 7397e7402. Cosgrove, D.J., 2005. Growth of the plant cell wall. Nat. Rev. Mol. Cell Biol. 6, 850e861. Coutinho, P.M., Henrissat, B., 2011. Annotating carbohydrate-active enzymes in plant genomes: present challenges. In: Ulvskov, P. (Ed.), Annual Plant Reviews: Plant Polysaccharides, Biosynthesis and Bioengineering. Wiley-Blackwell, Oxford, UK, pp. 93e107. Dashtban, M., Schraft, H., Qin, W., 2009. Fungal bioconversion of lignocellulosic residues; opportunities & perspectives. Int. J. Biol. Sci. 5, 578e595. Debolt, S., Estevez, J.M., 2012. Current challenges in plant cell walls: editorial overview. Front. Plant Sci. 3, 232. Dhugga, K.S., Barreiro, R., Whitten, B., Stecca, K., Hazebroek, J., Randhawa, G.S., Dolan, M., Kinney, A.J., Tomes, D., Nichols, S., Anderson, P., 2004. Guar seed beta-mannan synthase is a member of the cellulose synthase super gene family. Science 303, 363e366. Doering, A., Lathe, R., Persson, S., 2012. An update on xylan synthesis. Mol. Plant 5, 769e771. Driouich, A., Follet-Gueye, M.L., Bernard, S., Kousar, S., Chevalier, L., Vicre-Gibouin, M., Lerouxel, O., 2012. Golgi-mediated synthesis and secretion of matrix polysaccharides of the primary cell wall of higher plants. Front. Plant Sci. 3, 79. Egelund, J., Skjot, M., Geshi, N., Ulvskov, P., Petersen, B.L., 2004. A complementary bioinformatics approach to identify potential plant cell wall glycosyltransferase-encoding genes. Plant Physiol. 136, 2609e2620. Emiliani, G., Fondi, M., Fani, R., Gribaldo, S., 2009. A horizontal gene transfer at the origin of phenylpropanoid metabolism: a key adaptation of plants to land. Biol. Direct 4. Endler, A., Persson, S., 2011. Cellulose synthases and synthesis in Arabidopsis. Mol. Plant 4, 199e211. Fawal, N., Li, Q., Savelli, B., Brette, M., Passaia, G., Fabre, M., Mathe, C., Dunand, C., 2012. PeroxiBase: a database for large-scale evolutionary analysis of peroxidases. Nucleic Acids Res. Finn, R.D., Mistry, J., Schuster-Bockler, B., Griffiths-Jones, S., Hollich, V., Lassmann, T., Moxon, S., Marshall, M., Khanna, A., Durbin, R., Eddy, S.R., Sonnhammer, E.L., Bateman, A., 2006. Pfam: clans, web tools and services. Nucleic Acids Res. 34, D247eD251. Fu, C., Sunkar, R., Zhou, C., Shen, H., Zhang, J.Y., Matts, J., Wolf, J., Mann, D.G., Stewart Jr., C.N., Tang, Y., Wang, Z.Y., 2012. Overexpression of miR156 in switchgrass (Panicum virgatum L.) results in various morphological alterations and leads to improved biomass production. Plant Biotechnol. J. 10, 443e452. Girke, T., Lauricha, J., Tran, H., Keegstra, K., Raikhel, N., 2004. The cell wall navigator database. A systems-based approach to organismunrestricted mining of protein families involved in cell wall metabolism. Plant Physiol. 136, 3003e3008. Goodstein, D.M., Shu, S., Howson, R., Neupane, R., Hayes, R.D., Fazo, J., Mitros, T., Dirks, W., Hellsten, U., Putnam, N., Rokhsar, D.S., 2012. Phytozome: a comparative platform for green plant genomics. Nucleic Acids Res. 40, D1178eD1186. Gu, Y., Somerville, C., 2010. Cellulose synthase interacting protein: a new factor in cellulose synthesis. Plant Signaling Behav. 5, 1571e1574. Guillaumie, S., San-Clemente, H., Deswarte, C., Martinez, Y., Lapierre, C., Murigneux, A., Barriere, Y., Pichon, M., Goffner, D., 2007. MAIZEWALL. Database and developmental gene expression profiling of cell wall biosynthesis and assembly in maize. Plant Physiol. 143, 339e363. Guo, A.Y., Chen, X., Gao, G., Zhang, H., Zhu, Q.H., Liu, X.C., Zhong, Y.F., Gu, X., He, K., Luo, J., 2008. PlantTFDB: a comprehensive plant transcription factor database. Nucleic Acids Res. 36, D966eD969. Handford, M., Rodriguez-Furlan, C., Marchant, L., Segura, M., Gomez, D., Alvarez-Buylla, E., Xiong, G.Y., Pauly, M., Orellana, A., 2012. Arabidopsis thaliana AtUTr7 encodes a golgi-localized UDP-glucose/UDP-galactose transporter that affects lateral root emergence. Mol. Plant 5, 1263e1280. Hansen, S.F., Harholt, J., Oikawa, A., Scheller, H.V., 2012. Plant glycosyltransferases beyond CAZy: a perspective on DUF families. Front. Plant Sci. 3, 59. Harholt, J., Sorensen, I., Fangel, J., Roberts, A., Willats, W.G., Scheller, H.V., Petersen, B.L., Banks, J.A., Ulvskov, P., 2012. The glycosyltransferase repertoire of the spikemoss Selaginella moellendorffii and a comparative study of its cell wall. PLoS One 7, e35846. Harholt, J., Suttangkakul, A., Vibe Scheller, H., 2010. Biosynthesis of pectin. Plant Physiol. 153, 384e395. Hess, M., Sczyrba, A., Egan, R., Kim, T.W., Chokhawala, H., Schroth, G., Luo, S., Clark, D.S., Chen, F., Zhang, T., Mackie, R.I., Pennacchio, L.A., Tringe, S.G., Visel, A., Woyke, T., Wang, Z., Rubin, E.M., 2011. Metagenomic discovery of biomass-degrading genes and genomes from cow rumen. Science 331, 463e467.
REFERENCES Himmel, M.E., Ding, S.Y., Johnson, D.K., Adney, W.S., Nimlos, M.R., Brady, J.W., Foust, T.D., 2007. Biomass recalcitrance: engineering plants and enzymes for biofuels production. Science 315, 804e807. Humphreys, J.M., Chapple, C., 2002. Rewriting the lignin roadmap. Curr. Opin. Plant Biol. 5, 224e229. Hunter, S., Apweiler, R., Attwood, T.K., Bairoch, A., Bateman, A., Binns, D., Bork, P., Das, U., Daugherty, L., Duquenne, L., Finn, R.D., Gough, J., Haft, D., Hulo, N., Kahn, D., Kelly, E., Laugraud, A., Letunic, I., Lonsdale, D., Lopez, R., Madera, M., Maslen, J., McAnulla, C., McDowall, J., Mistry, J., Mitchell, A., Mulder, N., Natale, D., Orengo, C., Quinn, A.F., Selengut, J.D., Sigrist, C.J., Thimma, M., Thomas, P.D., Valentin, F., Wilson, D., Wu, C.H., Yeats, C., 2009. InterPro: the integrative protein signature database. Nucleic Acids Res. 37, D211eD215. Joshi, C.P., Mansfield, S.D., 2007. The cellulose paradoxesimple molecule, complex biosynthesis. Curr. Opin. Plant Biol. 10, 220e226. Lee, C., O’Neill, M.A., Tsumuraya, Y., Darvill, A.G., Ye, Z.H., 2007. The irregular xylem9 mutant is deficient in xylan xylosyltransferase activity. Plant Cell Physiol. 48, 1624e1634. Lei, L., Li, S., Gu, Y., 2012. Cellulose synthase complexes: composition and regulation. Front. Plant Sci. 3, 75. Lerouxel, O., Cavalier, D.M., Liepman, A.H., Keegstra, K., 2006. Biosynthesis of plant cell wall polysaccharides - a complex process. Curr. Opin. Plant Biol. 9, 621e630. Levasseur, A., Piumi, F., Coutinho, P.M., Rancurel, C., Asther, M., Delattre, M., Henrissat, B., Pontarotti, P., Asther, M., Record, E., 2008. FOLy: an integrated database for the classification and functional annotation of fungal oxidoreductases potentially involved in the degradation of lignin and related aromatic compounds. Fungal Genet. Biol. 45, 638e645. Li, J., Yuan, H., Yang, J., 2009. Bacteria and lignin degradation. Front. Biol. China 4, 29e38. Li, X., Chapple, C., 2010. Understanding lignification: challenges beyond monolignol biosynthesis. Plant Physiol. 154, 449e452. Liepman, A.H., Wilkerson, C.G., Keegstra, K., 2005. Expression of cellulose synthase-like (Csl) genes in insect cells reveals that CslA family members encode mannan synthases. Proc. Natl. Acad. Sci. USA 102, 2221e2226. Liu, C.J., Miao, Y.C., Zhang, K.W., 2011. Sequestration and transport of lignin monomeric precursors. Molecules 16, 710e727. Liwanag, A.J., Ebert, B., Verhertbruggen, Y., Rennie, E.A., Rautengarten, C., Oikawa, A., Andersen, M.C., Clausen, M.H., Scheller, H.V., 2012. Pectin biosynthesis: GALS1 in Arabidopsis thaliana is a beta-1,4-galactan beta-1,4-galactosyltransferase. Plant Cell. Lombard, V., Bernard, T., Rancurel, C., Brumer, H., Coutinho, P.M., Henrissat, B., 2010. A hierarchical classification of polysaccharide lyases for glycogenomics. Biochem. J. 432, 437e444. Mao, F.L., Yin, Y.B., Zhou, F.F., Chou, W.C., Zhou, C., Chen, H.L., Xu, Y., 2009. pDAWG: an integrated database for plant cell wall genes. BioEnergy Res. 2, 209e216. Marchler-Bauer, A., Anderson, J.B., Chitsaz, F., Derbyshire, M.K., DeWeese-Scott, C., Fong, J.H., Geer, L.Y., Geer, R.C., Gonzales, N.R., Gwadz, M., He, S., Hurwitz, D.I., Jackson, J.D., Ke, Z., Lanczycki, C.J., Liebert, C.A., Liu, C., Lu, F., Lu, S., Marchler, G.H., Mullokandov, M., Song, J.S., Tasneem, A., Thanki, N., Yamashita, R.A., Zhang, D., Zhang, N., Bryant, S.H., 2009. CDD: specific functional annotation with the conserved domain database. Nucleic Acids Res. 37, D205eD210. Markowitz, V.M., Chen, I.M., Chu, K., Szeto, E., Palaniappan, K., Grechkin, Y., Ratner, A., Jacob, B., Pati, A., Huntemann, M., Liolios, K., Pagani, I., Anderson, I., Mavromatis, K., Ivanova, N.N., Kyrpides, N.C., 2012. IMG/M: the integrated metagenome data management and comparative analysis system. Nucleic Acids Res. 40, D123eD129. 105 McCaig, B.C., Meagher, R.B., Dean, J.F., 2005. Gene structure and molecular analysis of the laccase-like multicopper oxidase (LMCO) gene family in Arabidopsis thaliana. Planta 221, 619e636. McCann, M., Rose, J., 2010. Blueprints for building plant cell walls. Plant Physiol. 153, 365. Michel, G., Tonon, T., Scornet, D., Cock, J.M., Kloareg, B., 2010. The cell wall polysaccharide metabolism of the brown alga Ectocarpus siliculosus. Insights into the evolution of extracellular matrix polysaccharides in Eukaryotes. New Phytol. 188, 82e97. Minic, Z., 2008. Physiological roles of plant glycoside hydrolases. Planta 227, 723e740. Mohnen, D., 2008. Pectin structure and biosynthesis. Curr. Opin. Plant Biol. 11, 266e277. Mohnen, D., Bar-Peled, M., Somerville, C.R., 2008. Cell wall polysaccharide synthesis. In: Himmel, M.E. (Ed.), Biomass Recalcitrance: Deconstructing the Plant Cell Wall for Bioenergy. Blackwell Publishing, pp. 94e159. Movahedi, S., Van Bel, M., Heyndrickx, K.S., Vandepoele, K., 2012. Comparative co-expression analysis in plant biology. Plant Cell Environ. 35, 1787e1798. Mutwil, M., Klie, S., Tohge, T., Giorgi, F.M., Wilkins, O., Campbell, M.M., Fernie, A.R., Usadel, B., Nikoloski, Z., Persson, S., 2011. PlaNet: combined sequence and expression comparisons across plant networks derived from seven species. Plant Cell 23, 895e910. Mutwil, M., Ruprecht, C., Giorgi, F.M., Bringmann, M., Usadel, B., Persson, S., 2009. Transcriptional wiring of cell wall-related genes in Arabidopsis. Mol. Plant 2, 1015e1024. Mutwil, M., Usadel, B., Schutte, M., Loraine, A., Ebenhoh, O., Persson, S. Assembly of an interactive correlation network for the Arabidopsis genome using a novel heuristic clustering algorithm. Plant Physiol. 152, 29e43. Nag, A., Karpinets, T.V., Chang, C.H., Bar-Peled, M., 2012. Enhancing a pathway-genome database (PGDB) to capture subcellular localization of metabolites and enzymes: the nucleotide-sugar biosynthetic pathways of Populus trichocarpa. Database (Oxford). bas013. Nataf, Y., Bahari, L., Kahel-Raifer, H., Borovok, I., Lamed, R., Bayer, E.A., Sonenshein, A.L., Shoham, Y., 2010. Clostridium thermocellum cellulosomal genes are regulated by extracytoplasmic polysaccharides via alternative sigma factors. Proc. Natl. Acad. Sci. USA 107, 18646e18651. Nobles, D.R., Brown, R.M., 2004. The pivotal role of cyanobacteria in the evolution of cellulose synthases and cellulose synthase-like proteins. Cellulose 11, 437e448. Obayashi, T., Hayashi, S., Saeki, M., Ohta, H., Kinoshita, K., 2009. ATTED-II provides coexpressed gene networks for Arabidopsis. Nucleic Acids Res. 37, D987eD991. Park, B.H., Karpinets, T.V., Syed, M.H., Leuze, M.R., Uberbacher, E.C., 2010. CAZymes analysis toolkit (CAT): web service for searching and analyzing carbohydrate-active enzymes in a newly sequenced organism using CAZy database. Glycobiology 20, 1574e1584. Pauly, M., Keegstra, K., 2008a. Cell-wall carbohydrates and their modification as a resource for biofuels. Plant J. 54, 559e568. Pauly, M., Keegstra, K., 2008b. Physiology and metabolism ’Tear down this wall’. Curr. Opin. Plant Biol. 11, 233e235. Pena, M.J., Zhong, R., Zhou, G.K., Richardson, E.A., O’Neill, M.A., Darvill, A.G., York, W.S., Ye, Z.H., 2007. Arabidopsis irregular xylem8 and irregular xylem9: implications for the complexity of glucuronoxylan biosynthesis. Plant Cell 19, 549e563. Penning, B.W., Hunter 3rd, C.T., Tayengwa, R., Eveland, A.L., Dugard, C.K., Olek, A.T., Vermerris, W., Koch, K.E., McCarty, D.R., Davis, M.F., Thomas, S.R., McCann, M.C., Carpita, N.C., 2009. Genetic resources for maize cell wall biology. Plant Physiol. 151, 1703e1728. Persson, S., Caffall, K.H., Freshour, G., Hilley, M.T., Bauer, S., Poindexter, P., Hahn, M.G., Mohnen, D., Somerville, C., 2007. The
106 6. DATABASES FOR BIOENERGY-RELATED ENZYMES Arabidopsis irregular xylem8 mutant is deficient in glucuronoxylan and homogalacturonan, which are essential for secondary cell wall integrity. Plant Cell 19, 237e255. Popper, Z., Michel, G., Herve, C., Domozych, D.S., Willats, W.G., Tuohy, M.G., Kloareg, B., Stengel, D.B., 2011. Evolution and diversity of plant cell walls: from algae to flowering plants. Annu. Rev. Plant Biol. 62, 567e590. Portnoy, T., Margeot, A., Seidl-Seiboth, V., Le Crom, S., Ben Chaabane, F., Linke, R., Seiboth, B., Kubicek, C.P., 2011. Differential regulation of the cellulase transcription factors XYR1, ACE2, and ACE1 in Trichoderma reesei strains producing high and low levels of cellulase. Eukaryotic Cell 10, 262e271. Qin, J., Li, R., Raes, J., Arumugam, M., Burgdorf, K.S., Manichanh, C., Nielsen, T., Pons, N., Levenez, F., Yamada, T., Mende, D.R., Li, J., Xu, J., Li, S., Li, D., Cao, J., Wang, B., Liang, H., Zheng, H., Xie, Y., Tap, J., Lepage, P., Bertalan, M., Batto, J.M., Hansen, T., Le Paslier, D., Linneberg, A., Nielsen, H.B., Pelletier, E., Renault, P., SicheritzPonten, T., Turner, K., Zhu, H., Yu, C., Jian, M., Zhou, Y., Li, Y., Zhang, X., Qin, N., Yang, H., Wang, J., Brunak, S., Dore, J., Guarner, F., Kristiansen, K., Pedersen, O., Parkhill, J., Weissenbach, J., Bork, P., Ehrlich, S.D., 2010. A human gut microbial gene catalogue established by metagenomic sequencing. Nature 464, 59e65. Raman, B., McKeown, C.K., Rodriguez Jr., M., Brown, S.D., Mielenz, J.R., 2011. Transcriptomic analysis of Clostridium thermocellum ATCC 27405 cellulose fermentation. BMC Microbiol. 11, 134. Rawlings, N.D., Barrett, A.J., Bateman, A., 2012. MEROPS: the database of proteolytic enzymes, their substrates and inhibitors. Nucleic Acids Res. 40, D343eD350. Reiter, W.D., 2008. Biochemical genetics of nucleotide sugar interconversion reactions. Curr. Opin. Plant Biol. 11, 236e243. Reiter, W.D., Vanzin, G.F., 2001. Molecular genetics of nucleotide sugar interconversion pathways in plants. Plant Mol. Biol. 47, 95e113. Richmond, T.A., Somerville, C.R., 2000. The cellulose synthase superfamily. Plant Physiol. 124, 495e498. Riederer, A., Takasuka, T.E., Makino, S., Stevenson, D.M., Bukhman, Y.V., Elsen, N.L., Fox, B.G., 2011. Global gene expression patterns in Clostridium thermocellum as determined by microarray analysis of chemostat cultures on cellulose or cellobiose. Appl. Environ. Microbiol. 77, 1243e1253. Royo, J., Gimez, E., Hueros, G., 2000. CMP-KDO synthetase: a plant gene borrowed from gram-negative eubacteria. Trends Genet. 16, 432e433. Ruprecht, C., Mutwil, M., Saxe, F., Eder, M., Nikoloski, Z., Persson, S., 2011. Large-scale co-expression approach to dissect secondary cell wall formation across plant species. Front. Plant Sci. 2, 23. Ruprecht, C., Persson, S., 2012. Co-expression of cell-wall related genes: new tools and insights. Front. Plant Sci. 3, 83. Saier Jr., M.H., Tran, C.V., Barabote, R.D., 2006. TCDB: the transporter classification database for membrane transport protein analyses and information. Nucleic Acids Res. 34, D181eD186. Sandhu, A.P., Randhawa, G.S., Dhugga, K.S., 2009. Plant cell wall matrix polysaccharide biosynthesis. Mol. Plant 2, 840e850. Scheller, H.V., Ulvskov, P., 2010. Hemicelluloses. Annu. Rev. Plant Biol. 61, 263e289. Seifert, G.J., 2004. Nucleotide sugar interconversions and cell wall biosynthesis: how to bring the inside to the outside. Curr. Opin. Plant Biol. 7, 277e284. Seshadri, R., Kravitz, S.A., Smarr, L., Gilna, P., Frazier, M., 2007. CAMERA: a community resource for metagenomics. PLoS Biol. 5, e75. Shannon, P., Markiel, A., Ozier, O., Baliga, N.S., Wang, J.T., Ramage, D., Amin, N., Schwikowski, B., Ideker, T., 2003. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498e2504. Sibout, R., Hofte, H., 2012. Plant cell biology: the ABC of monolignol transport. Curr. Biol. 22, R533eR535. Sirim, D., Wagner, F., Wang, L., Schmid, R.D., Pleiss, J., 2011. The laccase engineering database: a classification and analysis system for laccases and related multicopper oxidases. Database (Oxford) bar006. Somerville, C., 2006. Cellulose synthesis in higher plants. Annu. Rev. Cell Dev. Biol. 22, 53e78. Somerville, C., Bauer, S., Brininstool, G., Facette, M., Hamann, T., Milne, J., Osborne, E., Paredez, A., Persson, S., Raab, T., Vorwerk, S., Youngs, H., 2004. Toward a systems approach to understanding plant cell walls. Science 306, 2206e2211. Stam, M.R., Danchin, E.G., Rancurel, C., Coutinho, P.M., Henrissat, B., 2006. Dividing the large glycoside hydrolase family 13 into subfamilies: towards improved functional annotations of alphaamylase-related proteins. Protein Eng. Des. Sel 19, 555e562. Sun, G. MicroRNAs and their diverse functions in plants. Plant Mol. Biol. 80, 17e36. Sun, J., Tian, C., Diamond, S., Glass, N.L., 2012. Deciphering transcriptional regulatory mechanisms associated with hemicellulose degradation in Neurospora crassa. Eukaryotic Cell 11, 482e493. Sun, Y.H., Shi, R., Zhang, X.H., Chiang, V.L., Sederoff, R.R. MicroRNAs in trees. Plant Mol. Biol. 80, 37e53. Syed, M.H., Karpinets, T.V., Parang, M., Leuze, M.R., Park, B.H., Hyatt, D., Brown, S.D., Moulton, S., Galloway, M.D., Uberbacher, E.C., 2012. BESC knowledgebase public portal. Bioinformatics 28, 750e751. Ulvskov, P., 2011. In: Ulvskov, P. (Ed.), Annual Plant Reviews: Plant Polysaccharides, Biosynthesis and Bioengineering. Wiley-Blackwell, Oxford, UK. Vanholme, R., Morreel, K., Ralph, J., Boerjan, W., 2008. Lignin engineering. Curr. Opin. Plant Biol. 11, 278e285. Wang, H., Avci, U., Nakashima, J., Hahn, M.G., Chen, F., Dixon, R.A., 2010. Mutation of WRKY transcription factors initiates pith secondary wall formation and increases stem biomass in dicotyledonous plants. Proc. Natl. Acad. Sci. USA 107, 22338e22343. Wang, H.Z., Dixon, R.A., 2012. On-off switches for secondary cell wall biosynthesis. Mol. Plant 5, 297e303. Wang, S., Yin, Y., Ma, Q., Tang, X., Hao, D., Xu, Y., 2012. Genome-scale identification of cell-wall related genes in Arabidopsis based on coexpression network analysis. BMC Plant Biol. 12, 138. Welinder, K.G., 1992. Superfamily of plant, fungal and bacterial peroxidases. Curr. Opin. Struct. Biol. 2, 388e393. Weng, J.K., Chapple, C., 2010. The origin and evolution of lignin biosynthesis. New Phytol. 187, 273e285. Wu, A.M., Rihouey, C., Seveno, M., Hornblad, E., Singh, S.K., Matsunaga, T., Ishii, T., Lerouge, P., Marchant, A., 2009. The Arabidopsis IRX10 and IRX10-LIKE glycosyltransferases are critical for glucuronoxylan biosynthesis during secondary cell wall formation. Plant J. 57, 718e731. Xu, Z., Zhang, D., Hu, J., Zhou, X., Ye, X., Reichel, K.L., Stewart, N.R., Syrenne, R.D., Yang, X., Gao, P., Shi, W., Doeppke, C., Sykes, R.W., Burris, J.N., Bozell, J.J., Cheng, M.Z., Hayes, D.G., Labbe, N., Davis, M., Stewart Jr., C.N., Yuan, J.S., 2009. Comparative genome analysis of lignin biosynthesis gene families across the plant kingdom. BMC Bioinform. 10 (Suppl. 11), S3. Yang, S., Giannone, R.J., Dice, L., Yang, Z.K., Engle, N.L., Tschaplinski, T.J., Hettich, R.L., Brown, S.D., 2012. Clostridium thermocellum ATCC27405 transcriptomic, metabolomic and proteomic profiles after ethanol stress. BMC Genomics 13, 336. Yin, Y., Chen, H., Hahn, M.G., Mohnen, D., Xu, Y., 2010. Evolution and function of the plant cell wall synthesis-related glycosyltransferase family 8. Plant Physiol. 153, 1729e1746. Yin, Y., Huang, J., Gu, X., Bar-Peled, M., Xu, Y., 2011. Evolution of plant nucleotide-sugar interconversion enzymes. PLoS One 6, e27995. Yin, Y., Huang, J., Xu, Y., 2009. The cellulose synthase superfamily in fully sequenced plants and algae. BMC Plant Biol. 9, 99.
REFERENCES Yin, Y.B., Mao, X.Z., Yang, J.C., Chen, X., Mao, F.L., Xu, Y., 2012. dbCAN: a web resource for automated carbohydrate-active enzyme annotation. Nucleic Acids Res. 40, W445eW451. Yokoyama, R., Nishitani, K., 2004. Genomic basis for cell-wall diversity in plants. A comparative approach to gene families in rice and Arabidopsis. Plant Cell Physiol. 45, 1111e1121. Yong, W., Link, B., O’Malley, R., Tewari, J., Hunter, C.T., Lu, C.A., Li, X., Bleecker, A.B., Koch, K.E., McCann, M.C., McCarty, D.R., Patterson, S.E., Reiter, W.D., Staiger, C., Thomas, S.R., Vermerris, W., Carpita, N.C., 2005. Genomics of plant cell wall biogenesis. Planta 221, 747e751. York, W.S., O’Neill, M.A., 2008. Biochemical control of xylan biosynthesis - which end is up? Curr. Opin. Plant Biol. 11, 258e265. Zhang, B., Liu, X., Qian, Q., Liu, L., Dong, G., Xiong, G., Zeng, D., Zhou, Y., 2011. Golgi nucleotide sugar transporter modulates cell wall biosynthesis and plant growth in rice. Proc. Natl. Acad. Sci. USA 108, 5110e5115. 107 Zhao, Q., Dixon, R.A., 2011. Transcriptional networks for lignin biosyn-x thesis: more complex than we thought? Trends Plant Sci. 16, 227e233. Zhong, R., Lee, C., Ye, Z.H., 2010. Evolutionary conservation of the transcriptional network regulating secondary cell wall biosynthesis. Trends Plant Sci. 15, 625e632. Zhong, R., Pena, M.J., Zhou, G.K., Nairn, C.J., Wood-Jones, A., Richardson, E.A., Morrison 3rd, W.H., Darvill, A.G., York, W.S., Ye, Z.H., 2005. Arabidopsis fragile fiber8, which encodes a putative glucuronyltransferase, is essential for normal secondary wall synthesis. Plant Cell 17, 3390e3408. Zhong, R., Ye, Z.H., 2009. Transcriptional regulation of lignin biosynthesis. Plant Signaling Behav. 4, 1028e1034. Zhou, C., Yin, Y., Dam, P., Xu, Y., 2010a. Identification of novel proteins involved in plant cell-wall synthesis based on protein-protein interaction data. J. Proteome Res. 9, 5025e5037. Zhou, F., Chen, H., Xu, Y., 2010b. GASdb: a large-scale and comparative exploration database of glycosyl hydrolysis systems. BMC Microbiol. 10, 69.
C H A P T E R 7 Isobutanol Production from Bioenergy Crops Thaddeus Chukwuemeka Ezeji 1,*, Nasib Qureshi 2, Victor Ujor 1 1 The Ohio State University, Department of Animal Sciences and Ohio State Agricultural Research and Development Center (OARDC), Wooster, OH, USA, 2United States Department of Agriculture,a National Center for Agricultural Utilization Research, ARS, Bioenergy Research, Peoria, IL, USA *Corresponding author email: ezeji.1@osu.edu O U T L I N E Background/Introduction 109 Keto Acid Pathways for Higher Alcohol Production Feasibility of Using Bioenergy Crops as Sustainable 114 Feedstocks for Isobutanol Production 110 Biochemistry of Isobutanol Fermentation 112 Technologies That have been Developed for Simultaneous Butanol Fermentation and Recovery 115 Metabolic Engineering of Microorganisms for Isobutanol Production 113 BACKGROUND/INTRODUCTION Isobutanol (Inte‘rnational Union of Pure and Applied Chemistry nomenclature: 2-methylpropan-1-ol) is a branched four-carbon alcohol [(CH3)2CHCH2OH], with a boiling point of 108  C, a melting point of 108  C, and a relative density of 0.806 at 15  C (Budavari, 1996). It is also known as isobutyl alcohol or 2-methyl1-propanol. Isobutanol has a vapor pressure of 10.43 mm Hg or 13.9 hPa at 25  C (Daubert and Danner, 1985) and a water solubility of 85.0 g/l at 25  C (Valvani et al., 1981). These properties reveal that isobutanol is lighter than, and also soluble in water. While isobutanol is produced industrially via carbonylation (incorporation of carbon monoxide into organic/inorganic compounds) of propylene, isobutanol can be produced biologically via fermentation of glucose with a potential to use lignocellulosic biomass. Isobutanol is naturally Conclusion and Future Perspective 116 References 116 produced in low amounts by Saccharomyces cerevisiae as a degradation product of valine. The first report of biological production of isobutanol was by Dickinson et al. (1998) who demonstrated that S. cerevisiae was able to produce isobutanol using 13C-labeled valine as substrate. It was hypothesized that the product of valine transamination, a-ketoisovalerate, had four potential routes to isobutanol, which include (1) catalysis of a-ketoisovalerate by branched-chain a-keto acid dehydrogenase to produce isobutyryl-CoA and subsequently isobutanol; (2) catalysis of a-ketoisovalerate to isobutanol by pyruvate decarboxylase (PDC); (3) reduction of a-ketoisovalerate to a-hydroxyisovalerate by a-ketoisovalerate reductase; and (4) use of the PDC-like enzyme encoded by YDL080c to produce isobutanol. Given the fact that riddance of branched-chain a-keto acid dehydrogenase activity in an lpd1 disruption mutant did not prevent the formation of isobutanol, S. cerevisiae cell a Mention of trade names or commercial products in this article is solely for the purpose of providing scientific information and does not imply recommendation or endorsement by the United States Department of Agriculture. USDA is an equal opportunity provider and employer. Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00007-3 109 Copyright Ó 2014 Elsevier B.V. All rights reserved.
110 7. ISOBUTANOL PRODUCTION FROM BIOENERGY CROPS homogenates could not convert a-hydroxyisovalerate to isobutanol, and a strain with a disrupted PDC-like gene, YDL080c, produced wild-type levels of isobutanol, hence, routes 1, 3, and 4 were eliminated in S. cerevisiae. Notably, elimination of PDC activity in a pdc1 pdc5 pdc6 triple mutant abolished isobutanol production thus, buttressing the notion that this route is the right channel to isobutanol biosynthesis. As a feedstock chemical, isobutanol is used for the production of isobutyl acetate, which is subsequently used for the production of lacquers. It also finds use as a direct solvent, production of amino resins, isobutyl amines, and acrylate and methyl acrylate esters. The largest use for isobutanol is for the production of zinc dialkyldithiophosphates, an additive for lube oils, greases and hydraulic fluids, in which it functions as an antiwear and antioxidant additive. The second most significant applications of isobutanol are in the production of isobutyl acetate and as a solvent, primarily for surface coatings and adhesives (Bizziari et al., 2002). Recent advances in liquid biofuel technology (Mariano et al., 2011, 2012; Atsumi et al., 2008), depletion of petroleum reserves, global population growth, environmental and energy security concerns, have revived research efforts aimed at producing environmentally friendly liquid fuel chemicals. Indeed, global population is projected to reach 8.92 American billion by 2050 and world energy use may increase 53% by 2035. Consequently, there is an exigent need to source for or develop new fuels to fill potential shortfalls and, possibly replace our fast depleting petroleum reserves. Between 1980 and 2010, efforts have been focused on engineering microorganisms to make production of ethanol from biomass more efficient for use as a biofuel. Compared to ethanol, longer chain alcohols (e.g. n-propanol, n-butanol and isobutanol) have greater energy content, lower vapor pressure, and lower hygroscopicity, which make them superior alternatives to ethanol as a biofuel (Ladisch, 1991; Ezeji et al., 2005). Isobutanol has the potential to substitute gasoline or serve as a gasoline supplement and can be produced from domestically abundant biomass sources including lignocellulosic biomass. Lignocellulosic biomass, which may contain xylan, arabinan, galactan, glucuronic, acetic, ferulic, and coumaric acids, is the most abundant renewable resource on the planet (Koukiekolo et al., 2005) and has great potential as a substrate for isobutanol production (Higashide et al., 2011). Substrate cost has long been recognized to have significant influence on biofuel price and has been identified as a major factor affecting economic viability of n-butanol production by fermentation (Qureshi and Blaschek, 2000). Production of isobutanol from low-cost lignocellulosic biomass which does not compete with food crops may be critical for cost-effective fermentative production of isobutanol. Whereas majority of producing microorganisms including S. cerevisiae use glucose as preferred substrate for growth and alcohol production, recent advances in genetic engineering have made it possible to metabolically engineer these microorganisms to expand their substrate range to include pentose sugar components of lignocellulosic biomass hydrolysates such as xylose and arabinose, as in the case of ethanologenic microorganisms such as Escherichia coli (Dien et al., 1999, 2000), Zymomonas mobilis (Zhang et al., 1995; Deanda et al., 1996), and S. cerevisiae (Jin et al., 2005; Wisselink et al., 2007; Garcia Sanchez et al., 2010). This chapter describes the biochemistry of isobutanol production from biomass, latest developments in isobutanol production technology, and efforts directed toward development of more efficient and cost-effective processes for isobutanol production from biomass. KETO ACID PATHWAYS FOR HIGHER ALCOHOL PRODUCTION Keto acids are organic acids with ketone functional group on the second carbon, typically known as a-carbon, and are present in microorganisms as intermediate products of amino acids production pathways, and degradation of biosynthesized amino acids to alcohols is what is commonly known as Ehrlich pathway (Figure 7.1). Different strains of S. cerevisiae and yeasts belonging to genera such as Endomycopsis, Candida, and Hansenula are known to produce higher alcohols via keto acid pathways (Singh and Kunkee, 1976; Cronk et al., 1979). In addition to ethanol and CO2 production, S. cerevisiae produces a variety of relatively lowmolecular weight flavor compounds such as alcohols, diacetyl, esters, organic acids, organic sulfides, and carbonyl compounds during fermentation (Ter Schure Amino acid Transamination: e.g. 2-ketoglutaric acid L -glutamic acid α-Keto acid (e.g. L -glutamic acid ) Decarboxylation: removal of CO2 Aldehyde (e.g. isobutyraldehyde) NAD(P)H Dehydrogenase NAD(P)+ Alcohol (e.g. isobutanol) FIGURE 7.1 Simplified biochemistry of branched chain higher alcohol production from amino acids.
111 KETO ACID PATHWAYS FOR HIGHER ALCOHOL PRODUCTION TABLE 7.1 Intermediate Products from Potential Amino Acid Degradation (Ehrlich Pathway) Amino acid a-Keto Acid Fusel Aldehyde Fusel Alcohol Fusel Acid Isoleucine: HO2CCH(NH2) CH(CH3)CH2CH3 a-Ketomethylvalerate Methylvaleraldehyde Amyl alcohol Methylvaleric acid Leucine: HO2CCH(NH2) CH2CH(CH3)2 a-Ketoisocaproate Isoamylaldehyde Isoamyl alcohol Isovaleric acid Methionine: HO2CCH(NH2) CH2CH2SCH3 a-Keto-g-(methylthio) butyrate Methional Methionol 3-(Methylthio) propionic acid Phenylalanine: C6H5CH2CH(NH2) COOH Phenylpyruvate 2-Phenylacetaldehyde 2-Phenylethanol 2-Phenylacetate Tryptophan: C11H12N2O2 3-Indole pyruvate 3-Indole acetaldehyde Tryptophol 2-(Indole-3-yl) acetic acid Tyrosine: C9H11NO3 p-Hydroxyphenylpyruvate p-Hydroxyphenylacetaldehyde p-Hydroxyphenylethanol p-Hydroxyphenylacetic acid Valine: HO2CCH(NH2) CH(CH3)2 a-Ketoisovalerate Isobutyraldehyde or isovaleraldehyde or isobutanal Isobutanol Isobutyric acid et al., 1998; Hazelwood et al., 2008). These compounds are formed via the Ehrlich pathway involving branchedchain amino acids such as isoleucine, leucine, methionine, phenylalanine, tryptophan, tyrosine, and valine, and biocatalysts such as transaminase, decarboxylase, and alcohol dehydrogenase (ADH) (Figure 7.1, Table 7.1 Ryan and Kohlhaw, 1974). This pathway is prevalent in yeast and is especially active when yeast is cultivated in growth medium whose carbon source is solely amino acids. Notably, catabolism of isoleucine, leucine, methionine, phenylalanine, tryptophan, tyrosine, and valine by S. cerevisiae via Ehrlich pathway generates 2-methylbutanol (active amyl alcohol), 3-methylbutanol (isoamyl alcohol), methionol, 2-phenylethanol, tryptophol, p-hydroxyphenyl ethanol, and isobutanol, respectively (Figure 7.1; Table 7.1). These relatively long-chain alcohols are often referred to as fusel oils or fusel alcohols. During ethanolic fermentation by S. cerevisiae, small quantities of these alcohols (fusel oil) are produced (Singh and Kunkee, 1976). Whereas this mixture of alcohols may contribute flavor and body to wines, it can produce an off-flavor in wines when the acceptable concentration threshold is exceeded. Indeed, the catabolism of amino acids in S. cerevisiae and its regulation has been studied extensively (Ter Schure et al., 1998; Hazelwood et al., 2008; Dickinson et al., 1998, 2000). By transamination of amino acids to a-keto acids followed by decarboxylation of a-keto acids to aldehydes, these aldehydes can undergo reduction reaction to produce alcohols (Figure 7.1; Table 7.1). Scientists are exploiting this biosynthetic pathway to take advantage of the amino acid biosynthesis capability of producing microorganisms such as E. coli and S. cerevisiae to produce fusel alcohol, of which isobutanol appears to be the most attractive. In particular, the specificity of decarboxylases has been suggested to be an important factor influencing the composition of fusel alcohols (Harrison and Collins, 1968; Suomalainen and Keranen, 1967). Furthermore, the amount of fusel alcohols produced by different yeasts and specific ADH activities with the corresponding alcohols as substrates was found to be related as well (Singh and Kunkee, 1976). In recent years, metabolic engineering strategy using heterologous hosts such as E. coli and Clostridium cellulolyticum to produce higher alcohols from glucose and cellulose, respectively, is under investigation (Atsumi et al., 2008; Higashide et al., 2011). Depending on the source, 2-ketoacid decarboxylase (KDC, encoded by the kivd gene) and ADH (encoded by the adh2 gene), which play critical roles in fusel oil production, may have broad substrate specificities toward the catalysis of 2-ketoacids and generation of isobutanol (Figures 7.1 and 7.2). When these two genes, kivd from Lactococcus lactis and adh2 from S. cerevisiae, were cloned and overexpressed in E. coli, approximately six long-chain alcohols including 1-propanol, 1-butanol, isobutanol, 2-methyl-1-butanol, 3-methyl1-butanol, and 2-phenylethanol were produced (Atsumi et al., 2008). This strategy exploits the presence of a highly active amino acid biosynthetic pathway in the host microorganism, keto acid pathway, and the ability
112 7. ISOBUTANOL PRODUCTION FROM BIOENERGY CROPS Glucose Acetyl-CoA PDA1 Pyruvate 2-acetolactate ILV2 NAD(P)H 2-acetolactate ILV5 NAD(P)+ NAD(P)H ILV5 2,3-dihydroxy-isovalerate NAD(P)+ Mitochondria ILV2 Pyruvate FIGURE 7.2 Schematic diagram depicting pathways leading to valine and isobutanol biosynthesis in S. cerevisiae. Genes encoding enzymes that catalyze each step are indicated and are as follows: ADH2 (alcohol dehydrogenase), Bat1 and Bat2 (branched chain amino acid aminotransferases), ILV2 (acetolactate synthase), ILV3 (dihydroxyacid dehydratase), ILV5 (acetohydroxyacid reductoisomerase), Kdc/kivd/Pdc, 2-ketoacid decarboxylase (pyruvate decarboxylase); and PDA (pyruvate dehydrogenase complex). (For color version of this figure, the reader is referred to the online version of this book.) ILV3 2,3-dihydroxy-isovalerate 2-ketoisovalerate ILV3 2-ketoisovalerate PDC6, 5, 1 (kivd) L-glutamate BAT1 2-oxoglutarate L -Valine Isobutyraldehyde ADH2 NAD(P)H Isobutanol L-glutamate BAT2 2-oxoglutarate NAD(P)+ 2-ketoisovalerate PDC6, 5, 1 (kivd) Isobutyraldehyde ADH2 Cytosol NAD(P)H NAD(P)+ Isobutanol of the host to reroute its 2-ketoacid intermediates for alcohol synthesis (Atsumi et al., 2008). The amount of individual alcohol produced is compared with the level of its corresponding ketoacid. For example, when alsS from Bacillus subtilis and ilvCD from E. coli were overexpressed in E. coli, the resulting strain accumulated remarkable amounts of 2-ketoisovalerate (KIV) in the fermentation broth. Furthermore, when kivd and adh2 were co-expressed in this recombinant strain, approximately 22 g/l isobutanol was produced over the course of 112 h of fermentation (Atsumi et al., 2008). Notably, AlsS of B. subtilis, kivd from L. lactis, adh2 from S. cerevisiae, and YqhD (nicotinamide adenine dinucleotide phosphate (NADPH)-dependent ADH) from E. coli have high affinity for pyruvate and 2-ketoacids, 2-ketoacids, isobutyraldehyde, and isobutyraldehyde, respectively. BIOCHEMISTRY OF ISOBUTANOL FERMENTATION In the presence of excess sugar, yeast, especially S. cerevisiae, has a strong tendency to undergo alcoholic fermentation, even when oxygen is available in excess (Van Diken and Scheffers, 1986). In the biochemical pathway for carbohydrate metabolism in most yeasts, two modes of disaccharide metabolism exist. While extracellular hydrolysis of sucrose to glucose and fructose followed by transport of these monosaccharides into the cell is the most common method for sucrose metabolism in yeast, transport of disaccharides by protonesugar symport followed by intracellular hydrolysis occurs in maltose and lactose metabolism (Weusthuis et al., 1994). However, hydrolysis of sucrose can occur either intracellularly or extracellularly in S. cerevisiae (Santos et al., 1982), followed by the phosphorylation of glucose to glucose-6-phosphate, which is subsequently catabolized to pyruvate via the EmbdeneMeyerhofeParnas pathway (Figure 7.2). Although most of the synthesized pyruvate is decarboxylated to acetaldehyde (ethanal) by PDC followed by the reduction of acetaldehyde to ethanol by ADH, a small proportion of the pyruvate is converted to fusel alcohols such as isobutanol (Figures 7.1 and 7.2; Table 7.1). KIV is an important precursor for valine biosynthesis, which is also shared by isobutanol production. KIV biosynthesis is initiated by the condensation of two pyruvate molecules to 2-acetolactate, which is catalyzed by acetolactate synthase (ILV2 þ ILV6; Figure 7.2). Notably, ILV6 is the regulatory subunit of acetolactate synthase and an enhancer of ILV2 catalytic activity
METABOLIC ENGINEERING OF MICROORGANISMS FOR ISOBUTANOL PRODUCTION (Chen et al., 2011). The 2-acetolactate is reduced to 2,3-dihydroxyisovalerate via catalysis by acetohydroxyacid reductoisomerase (ILV5), the precursor for KIV biosynthesis (Figure 7.2; Velasco et al., 1993). Thus, KIV is produced through catalysis of 2,3-dihydroxyisovalerate by dihydroxyacid dehydratase (ILV3). Further, the bidirectional conversion between KIV and valine is catalyzed by aminotransferases (Bat1 and Bat2). While aminotransferase Bat1 is present in the mitochondrial matrix of S. cerevisiae, aminotransferase Bat2 is present in the cytosol (Kispal et al., 1996; Chen et al., 2011). Next, KIV is decarboxylated by PDC, a KDC, to isobutyraldehyde and subsequently, reduced to isobutanol by ADHs (Figure 7.2). METABOLIC ENGINEERING OF MICROORGANISMS FOR ISOBUTANOL PRODUCTION In an effort to metabolically engineer microorganisms for efficient isobutanol production, various researchers sought to understand isobutanol production at both the molecular and protein levels. A few years before 2013, a number of studies directed at finding molecular and biochemical bases for isobutanol synthesis have been conducted (Table 7.2). Using a prokaryote (E. coli) as host, Atsumi et al. (2008) demonstrated that KIV can serve as a precursor for efficient isobutanol production from glucose in a biosynthetic pathway consisting of 2-KDC from L. lactis and ADH2 from S. cerevisiae with broad-range substrate specificity in combination with the expression of the alsS gene (encoding AHAS (acetohydroxy acid synthase)) from B. subtilis and the ilvCD gene from E. coli. A follow-up study on the role of different ADHs on isobutanol production with E. coli showed that chromosomally encoded YqhD is the major isobutyraldehyde-converting enzyme, and ADH2 from S. cerevisiae contributes only to a minor extent to isobutanol production in E. coli (Atsumi et al., 2010). This supposition was made after yqhD gene was deleted from the genome of E. coli; the generated recombinant E. coli strain (yqhD deficient) accumulated isobutyraldehyde during fermentation and experienced 80% reduction in isobutanol production. Using an a eukaryote (S. cerevisiae) and working on the assumption that inefficient production of isobutanol from glucose may be due to limited supply of KIV, a precursor for the valine biosynthesis pathway (Figure 7.2), Lee et al. (2012) added exogenous KIV (0.5 g/l) into the growth medium. As expected, supplementation of the medium with KIV improved the production of isobutanol, which suggests that the endogenous pathway for producing KIV in S. cerevisiae (and potentially, other producing microorganisms) is the limiting step in the isobutanol production pathway. Consequently, this finding made KIV biosynthesis a rational target for metabolic engineering 113 toward designing more robust isobutanol-producing strains. Development of hyper-isobutanol-producing strains has typically followed one of the three approaches: (1) identification of rate-limiting steps in Ehrlich/2-ketoacid biosynthetic pathways and overexpression of KIV biosynthetic genes; (2) improvement of heterologous expression of enzymes of the isobutanol pathway by codon optimization; and (3) removal of feedback inhibition and deletion of other competitive pathways, especially competition for pyruvate. Pursuant to the first strategy, Lee et al. (2012) screened and identified a 2-KDC exhibiting a relatively higher activity on KIV through in vitro activity assays of KDC using crude extracts of transformants overexpressing KDCs from various microorganisms. The highest KDC activity with KIV was observed from the transformant expressing kivd from L. lactis subsp. lactis KACC13877. Subsequently, Chen et al. (2011) evaluated the effect of overexpressing the genes, ILV2, ILV3, ILV5, ILV6, and BAT2, involved in valine metabolism, in different combinations in S. cerevisiae, on isobutanol production. Following cultivation of the ILV2, ILV3, and ILV5 overexpressing strain (ILV235_XCY561) and the reference strain (CEN.PK113-5D) in mineral glucose medium supplemented with uracil in fermentors under anaerobic conditions, the recombinant strain ILV235_XCY561 produced 0.97  0.14 mg isobutanol/g glucose, which was sixfold higher than the control strain (Chen et al., 2011), hence attesting to the fact that overexpression of the genes ILV2, ILV3, and ILV5 may have led to a higher concentration of KIV, which resulted in higher isobutanol production. In parallel, Atsumi et al. (2009) engineered a cyanobacterium, Synechococcus elongatus, by expressing a KDC gene (kivd) from L. lactis in this cyanobacterium using an expression cassette under the control of the isopropyl-b-D-thiogalactoside-inducible promoter Ptrc and integration into neutral site I (Bustos and Golden, 1992) by homologous recombination (Golden et al., 1987), and strain SA578 was generated. To improve flux toward KIV, alsS gene from B. subtilis and the ilvC and ilvD genes from E. coli were integrated into neutral site II (Andersson et al., 2000) in the genome of strain SA578 to generate strain SA590, which produced high levels of isobutyraldehyde upon cultivation in a Roux culture bottle at 30  C (Atsumi et al., 2009, Figure 7.2). Given the fact that isobutyraldehyde can easily undergo a reduction reaction to produce isobutanol, Atsumi et al. (2009) evaluated feasibility of using ADHs (ADH2 from S. cerevisiae, YqhD from E. coli, and AdhA from L. lactis) and Kivd from L. lactis (strain SA590) to achieve this reduction reaction. These genes (ADH2, YqhD, and AdhA) were integrated downstream of kivd individually, hence, strains SA413, SA561 and SA562 were generated, respectively. Whereas YqhD, an NADPH-dependent enzyme, was the most active one in S. elongatus, AdhA and ADH2 were nicotinamide adenine
114 7. ISOBUTANOL PRODUCTION FROM BIOENERGY CROPS TABLE 7.2 Isobutanol Levels Produced by Genetically Modified Microorganisms during Fermentation Species Isobutanol Titer (g/l) Modified or Overexpressed Genes Escherichia coli 21.2 Carbon Source References Random mutagenesis and selection Hexose and pentose* Smith and Liao (2011) Clostridium cellulolyticum 0.66 alsS, ilvC, ilvD, kivd, yqhD Cellulose Higashide et al. (2011) Saccharomyces cerevisiae 0.00136 xyla, xks1, tal1, aro10, adh2, ilv2, ilv5, ilv3 Hexose and pentose Brat and Boles (2013) Saccharomyces cerevisiae 0.143 kdc (kivd), adh (adh6) Hexose and pentose* Kondo et al. (2012) Saccharomyces cerevisiae 0.151 kivd, Ilv2, ilv3, ilv5 Hexose and pentose* Lee et al. (2012) Corynebacterium glutamicum 2.2 alsS, ilvC, ilvD, kivd Hexose and pentose* Smith et al. (2010) Corynebacterium glutamicum 2.6 alsS, ilvC, ilvD, kivd, adhA Hexose and pentose* Smith et al. (2010) Synechococcus elongatus 7942 0.018 kivd, yqhD Carbon dioxide Atsumi et al. (2009) Synechococcus elongatus 7942 0.45 alsS, ilvC, kivd, yqhD Carbon dioxide Atsumi et al. (2009) Synechocystis sp. strain PCC 6803 0.114 kivd, adhA Carbon dioxide and hexose Varman et al. (2013) * Recombinant strains capable of using pentose sugars as carbon source are available. Abbreviations: adhA, alcohol dehydrogenase from Lactococcus lactis; aro10, phenylpyruvate pyruvate decarboxylase; alsS, acetolactate synthase from Bacillus subtilis; ilv2, acetolactate synthase; ilv3 (ilvD), dihydroxyacid dehydratase; ilv5 (ilvC), acetohydroxyacid reductoisomerase; Kdc/kivd/Pdc, 2-ketoacid decarboxylase (pyruvate decarboxylase); yqhD, alcohol dehydrogenase/aldehyde reductase; xyla, xylose isomerase from Clostridium phytofermentans; xks1, xylulokinase; tal1, transaldolase. dinucleotide-dependent, and generated strain SA579, which produced 450 mg/l isobutanol. The second strategy derives from differences in codon bias among different microorganisms, based on their individual transfer RNA content and requisite expression levels of specific proteins in each microorganism (Ikemura, 1985; Percudani et al., 1997). Given the fact that codons at the beginning of an open reading frame play a critical role in protein expression (Vervoort et al., 2000), codon bias influences heterologous expression of foreign proteins to a great extent. For instance, heterologously expressed protein levels in E. coli (Atsumi et al., 2010) and S. cerevisiae (Brat and Boles, 2013) were improved by codon optimization, especially at the 50 end of the coding sequence. To realize the full potential of heterologously overexpressed genes in producing microorganisms with respect to efficient isobutanol production, six genes including adhE (ADH), ldhA (lactate dehydrogenase), frd (fumarate reductase), fnr (encodes redox-sensing transcription regulator, which partakes in the regulation of lactate synthesis), pta (phosphate acetyltransferase), and pflB (pyruvate formate lyase) that are involved in byproduct formation in E. coli were deleted following overexpression of AlsS (B. subtilis), IlvC (E. coli), and IlvD (E. coli) (Atsumi et al., 2008). These deletions may have increased the level of pyruvate available for the valine biosynthesis pathway. As a consequence, the isobutanol strain (JCL260/pSA55/pSA69) produced more than 22 g/l in 112 h (Atsumi et al., 2008). In a similar study, Kondo et al. (2012) overexpressed 2-KDC and ADH in S. cerevisiae to enhance the endogenous activity of the Ehrlich pathway followed by overexpression of Ilv2, which catalyzes the first step in the valine synthetic pathway and deletion of the PDC1 gene encoding a major PDC with the intent of reducing ethanol flux via pyruvate. As a result, S. cerevisiae YTD306 was generated. Upon cultivation of S. cerevisiae YTD306 along with modification of culture conditions, a 13-fold increase in isobutanol titer was produced (from 11 mg/l to 143 mg/l) when compared with the control (Table 7.2, Kondo et al., 2012). The strategy described here, in which amino acid biosynthetic and 2-ketoacid degradation pathways were exploited for isobutanol production, represents a new paradigm for biofuel production. Indeed, this paradigm employs non-CoA-mediated chemistry and uses only pyruvate as a precursor, unlike ethanol and butanol production by native alcohols producing microorganisms that are CoA-dependent (Ezeji et al., 2010; Atsumi et al., 2008). FEASIBILITY OF USING BIOENERGY CROPS AS SUSTAINABLE FEEDSTOCKS FOR ISOBUTANOL PRODUCTION Perennial grasses (lignocellulosic biomass) such as switchgrass (Panicum virgatum), Miscanthus, and Napier grass (Pennisetum purpureum) have been gaining attention recently for use in biofuel production because of
TECHNOLOGIES THAT HAVE BEEN DEVELOPED FOR SIMULTANEOUS BUTANOL FERMENTATION AND RECOVERY their low energy requirement for production in the US and Europe, and high biomass yield (Khanna et al., 2008). Miscanthus species most often used in biomass research is the sterile hybrid Miscanthus x giganteus, a hardy and fast growing C4 grass that is cultivated via rhizomes (Lewandowski et al., 2000). Yields per acre vary depending on where the crop is grown. The typical yield is 4e10 tons/acre, but yields have been known to reach 16 tons/acre in southern Europe (Lewandowski et al., 2000). Recent research on M. x giganteus and switchgrass at the University of Illinois has produced an average yield of 12 tons/acre and a maximum of 24.7 tons/acre for M. x giganteus and approximately 5 tons/acre for switchgrass (Heaton et al., 2008). The harvestable biomass of Miscanthus is 190% greater than that of corn and could produce 742 more gallons of ethanol per acre (Heaton et al., 2008) or 600 more gallons of butanol per acre. Napier grass, which belongs to sugarcane family and a native to Africa, is now found in most tropical and subtropical regions of the world (Pennisetum purpureum, 2013). It has a high moisture content of 70e80% and reaches maturation following 8 months of plantation/rationing. Approximately two-thirds of Napier grass biomass (dry weight) is composed of sugars: glucan (38.43%), xylan (20.20%), galactan (2.02%), arabinan (2.73%) and mannan (0.23%), and lignin accounted for 20.93% of the lignocellulosic material, with ash (7.75%) and extractives (1.76%) comprising the remaining fraction (Takara and Khanal 2011). Napier grass has a rapid and dense growth, which have attracted the attention of researchers as a potentially ideal source for lignocellulosic biomass. Napier grass is capable of producing 42 dry tons/acre/year, approximately double the biomass yields of sugarcane and switchgrass (McLaughlin and Kszos, 2005; Takara and Khanal, 2011). Nonetheless, one of the key steps in the lignocellulosic biomass-to-fermentable sugars conversion is pretreatment. The goal of pretreatment is to disrupt the biomass structure and disentangle ligninecarbohydrate complex such that enzymatic hydrolysis of the carbohydrate fraction of the lignocellulosic biomass-to-simple sugars can be achieved more rapidly and with greater yield (Mosier et al., 2005; Ezeji and Blaschek, 2010). Economic analysis of the current pretreatment methods has shown that the relatively high costs of biofuel (ethanol) production from lignocellulosic biomass arise mainly from costs associated with three factors: (a) harsh pretreatment conditions (high temperature, high pressure, use of acids or bases, long residence time, and so on, allowing for inhibitor formation); (b) overuse of expensive enzymes; and (c) recovery of end products (low ethanol concentration in beer; Eggeman and Elander, 2005; Ezeji and Blaschek, 2010). Technologies that lead to improvement in any of these areas will help to make isobutanol production using energy crops as 115 feedstock more cost-effective. Moreover, energy crops can be genetically modified to improve biomass yield (per acre per year) without the risk of compromising grain yield or quality along with reducing their recalcitrance to efficient deconstruction to monomeric sugars. While producing microorganisms have not been shown to directly utilize lignocellulosic biomass as a carbon source for isobutanol production, Higashide et al. (2011) recently demonstrated the first production of isobutanol from crystalline cellulose using C. cellulolyticum. This breakthrough was accomplished after a couple of attempts. First, the activities of the first three enzymes in the isobutanol production pathway were examined by transforming plasmids expressing alsS or alsS ilvCD into C. cellulolyticum and no C. cellulolyticum alsS or alsS ilvCD transformants were obtained. Realizing that alsS and alsS ilvCD transformants could not be obtained, a second attempt wherein genes encoding B. subtilis a-acetolactate synthase, E. coli acetohydroxyacid isomeroreductase, E. coli dihydroxyacid dehydratase, L. lactis KDC, and E. coli and L. lactis ADHs (alsS, ilvCD, kivd, and adhA, complete isobutanol production pathway genes) were expressed in C. cellulolyticum (Higashide et al., 2011). Despite a mutation in alsS, the alsS ilvCD kivd adhA strain produced 140 and 420 mg/l isobutanol from cellobiose and cellulose, respectively. When plasmids expressing kivd yqhD alsS ilvCD, in which alsS was the third gene in the operon, was constructed and transformed into C. cellulolyticum, 364 and 660 mg/l isobutanol were produced from cellobiose and cellulose, respectively (Table 7.2). Given the fact that isobutanol production technology has been changing at a rapid pace, this accomplishment in which cellulose is used as a carbon source is significant because it opens the frontier for utilizing lignocellulosic biomass such as energy crops for isobutanol production. TECHNOLOGIES THAT HAVE BEEN DEVELOPED FOR SIMULTANEOUS BUTANOL FERMENTATION AND RECOVERY Production of isobutanol by fermentation is looking attractive owing to two main reasons: (1) the higher tolerance of producing microorganisms to isobutanol, usually 36.9e51.9 g/l as compared to n-butanol (called butanol in the later sections of this chapter), which is 20e30 g/l in selected hyper-producing strains and (2) having a lower boiling point (108  C vs 118  C) than butanol, which comparatively may be economical to recover. The above titer values for the isobutanol are without simultaneous product recovery (Baez et al., 2011). However, in this report, it was mentioned that in situ recovery by gas stripping improves isobutanol
116 7. ISOBUTANOL PRODUCTION FROM BIOENERGY CROPS production. To the authors’ knowledge, this is the only report where isobutanol fermentation and recovery were performed simultaneously. Fermentative production of isobutanol or butanol can be economically achieved in two ways: (1) by developing the high-tolerant or high-producer strain, which also offers some benefits during the recovery process, and (2) using energy-efficient process engineering techniques to simultaneously remove the toxic product. Interestingly, the first approach has been reported for isobutanol fermentation (Baez et al., 2011) with great success. For butanol producing strains, numerous attempts have been made to improve performance; however, success has been limited, with maximum titer stagnating around 21 g/l (total acetone butanol ethanol (ABE) 32.6 g/l) (Chen and Blaschek, 1999). Nevertheless, butanol has drawn significant amount of attention from the process engineering point of view. One of the main focuses has been the development of integrated process technologies where fermentation and simultaneous product recovery have been integrated. The reader is directed to a couple of following reports, where much higher production of ABE than in a batch system has been achieved. Employing such integrated systems (gas stripping and perstraction), cumulative ABE production from 461 g/l to 698 g/l (Ezeji et al., 2013; Jeon and Lee, 1989) has been achieved compared to 21 g/l butanol or 51.9 g/l isobutanol. It should also be noted that several simultaneous product recovery systems such as adsorption, liquideliquid extraction, pervaporation, ionic liquid extraction, and reverse osmosis have been investigated (Qureshi et al., 2013). Also, other advances have been made for butanol production from agricultural residues such as wheat straw (Qureshi et al., 2007; Qureshi et al., 2008a), barley straw (Qureshi et al., 2010a), corn stover (Parekh et al., 1988; Qureshi et al., 2010b), switchgrass (Qureshi et al., 2010b), and distillers dry grains and soluble (Ezeji and Blaschek, 2008). CONCLUSION AND FUTURE PERSPECTIVE Given the higher microbial tolerance of isobutanol and its greater volatility in comparison to butanol, it is likely that simultaneous product recovery using gas stripping, perstraction, and/or pervaporation would achieve even higher production levels than reported for butanol, thus benefiting economics of isobutanol production process. To make this biofuel even more attractive, recent advances in fermentation of lignocellulosic biomass such as separate hydrolysis of lignocellulosic biomass, combined with fermentation, and product recovery separate hydrolysis, fermentation, and recovery (SHFR), and simultaneous saccharification, fermentation, and recovery (SSFR) should be applied (Qureshi et al., 2013). In a process where wheat straw was used to produce butanol by SSFR, 192 g/l ABE was produced from 430 g/l of lignocellulosic sugars (Qureshi et al., 2008b). At this stage, engineering producing microorganisms by applying similar method already reported by Higashide et al. (2011) to utilize pentose sugars such as arabinose and xylose as substrates for the production of isobutanol is looking promising. Owing to the fact that butanol producing cultures have the potential to be tolerant to isobutanol, it is reasonable to express isobutanol-producing genes in solventogenic Clostridium species. Such an undertaking would have two advantages: (1) ability of the developed strain to utilize pentose sugars that accounts for about 30e40% of carbohydrates present in lignocellulosic biomass and (2) the developed strain may produce higher titers of isobutanol than yeast or E. coli. References Andersson, C.R., Tsinoremas, N.F., Shelton, J., Lebedeva, N.V., Yarrow, J., Min, H., Golden, S.S., 2000. Application of bioluminescence to the study of circadian rhythms in cyanobacteria. Methods Enzymol. 305, 527e542. Atsumi, S., Hanai, T., Liao, J.C., 2008. Non-fermentative pathways for synthesis of branched-chain higher alcohols as biofuels. Nature 451, 86e89. Atsumi, S., Higashide, W., Liao, J.C., 2009. Direct photosynthetic recycling of carbon dioxide to isobutyraldehyde. Nat. Biotechnol. 27, 1177e1180. Atsumi, S., Wu, T.-Y., Eckl, E.-M., Hawkins, S.D., Buelter, T., Liao, J.C., 2010. Engineering the isobutanol biosynthetic pathway in Escherichia coli by comparison of three aldehyde reductase/ alcohol dehydrogenase genes. Appl. Microbiol. Biotechnol. 85, 651e657. Baez, A., Cho, K.M., Liao, J.C., 2011. High-flux isobutanol production using engineered Escherichia coli: a bioreactor study with in situ product removal. Appl. Microbiol. Biotechnol. 90, 1681e1690. Brat, D., Boles, E., 2013. Isobutanol production from D-xylose by recombinant Saccharomyces cerevisiae. FEMS Yeast Res. 13, 241e244. Budavari, S., 1996. The Merck Index. An Encyclopedia of chemicals, Drugs, and Biologicals. Merck Research Laboratories, Division of Merck & Co., Inc, Whitehouse Station, NJ. Bustos, S.A., Golden, S.S., 1992. Light-regulated expression of the psbD gene family in Synechococcus sp. strain PCC 7942: evidence for the role of duplicated psbD genes in cyanobacteria. Mol. Gen. Genet. 232, 221e230. Bizziari, S.N., Gubler, R., Kishi, A., 2002. CEH Marketing Research Report for Plasticizer Alcohols. Chen, C.-K., Blaschek, H.P., 1999. Acetate enhances solvent production and prevents degeneration in Clostridium beijerinckii BA101. Appl. Microbiol. Biotechnol. 52, 170e173. Chen, X., Nielsen, K.F., Borodina, I., Kielland-Brandt, M.C., Karhumaa, K., 2011. Increased isobutanol production in Saccharomyces cerevisiae by overexpression of genes in valine metabolism. Biotechnol. Biofuels 4, 21. Cronk, T.C., Mattick, L.R., Steinkraus, K.H., Hackler, L.R., 1979. Production of higher alcohols during Indonesian tape ketan fermentation. Appl. Environ. Microbiol. 37, 892e896.
REFERENCES Daubert, T.E., Danner, R.P., 1985. Data Compilation Tables of Properties of Pure Compounds. 1985. Design Institute for Physical Property Data, American Institute of Chemical Engineers. Deanda, K., Zhang, M., Eddy, C., Picataggio, S., 1996. Development of an arabinose-fermenting Zymomonas mobilis strain by metabolic pathway engineering. Appl. Environ. Microbiol. 62, 4465e4470. Dickinson, J.R., Harrison, S.J., Hewlins, M.J., 1998. An investigation of the metabolism of valine to isobutyl alcohol in Saccharomyces cerevisiae. J. Biol. Chem. 273, 25751e25756. Dickinson, J.R., Harrison, S.J., Dickinson, J.A., Hewlins, M.J., 2000. An investigation of the metabolism of isoleucine to active amyl alcohol in Saccharomyces cerevisiae. J. Biol. Chem. 275, 10937e10942. Dien, B.S., Iten, L.B., Bothast, R.J., 1999. Conversion of corn fiber to ethanol by recombinant E. coli strain FBR3. J. Ind. Microbiol. Biotechnol. 22, 575e581. Dien, B.S., Nichols, N.N., OBryan, P.J., Bothast, R.J., 2000. Development of new ethanologenic Escherichia coli strains for fermentation of lignocellulosic biomass. Appl. Biochem. Biotechnol. 84, 181e196. Eggeman, T., Elander, R.T., 2005. Process and economic analysis of pretreatment technologies. Bioresour. Technol. 96, 2019e2025. Ezeji, T.C., Blaschek, H.P., 2008. Fermentation of dried distillers’ grains and soluble (DDGS) hydrolysates to solvents and value-added products by solventogenic clostridia. Bioresour. Technol. 99, 5232e5242. Ezeji, T.C., Blaschek, H.P., 2010. Butanol production from lignocellulosic biomass. In: Blaschek, H.P., Ezeji, T.C., Scheffran, J. (Eds.), Biofuels from Agricultural Wastes and Byproducts. WileyBlackwell, Ames, IA, pp. 19e37. Ezeji, T.C., Karcher, P.M., Qureshi, N., Blaschek, H.P., 2005. Improving performance of a gas stripping-based recovery system to remove butanol from Clostridium beijerinckii fermentation. Bioprocess Biosyst. Eng. 27, 207e214. Ezeji, T., Milne, C., Price, N.D., Blaschek, H.P., 2010. Achievements and perspectives to overcome the poor solvent resistance in acetone and butanol-producing microorganisms. Appl. Microbiol. Biotechnol. 85, 1697e1712. Ezeji, T.C., Qureshi, N., Blaschek, H.P., 2013. Microbial production of a biofuel (acetone-butanol-ethanol) in a continuous bioreactor: impact of bleed and simultaneous product removal. Bioprocess Biosyst. Eng. 36, 109e116. Garcia Sanchez, R., Karhumaa, K., Fonseca, C., Sànchez Nogué, V., Almeida, J.R.M., Larsson, C.U., Bengtsson, O., Bettiga, M., HahnHägerdal, B., Gorwa-Grauslund, M.F., 2010. Improved xylose and arabinose utilization by an industrial recombinant Saccharomyces cerevisiae strain using evolutionary engineering. Biotechnol. Biofuels 3, 13. Golden, S.S., Brusslan, J., Haselkorn, R., 1987. Genetic engineering of the cyanobacterial chromosome. Methods Enzymol. 153, 215e231. Harrison, G.A.F., Collins, E., 1968. Gas-chromatographic determination of individual a-keto acids in beer. Proc. Am. Soc. Brew. Chem. 1968, 101e105. Hazelwood, L.A., Daran, J.M., van Maris, A.J., Pronk, J.T., Dickinson, J.R., 2008. The Ehrlich pathway for fusel alcohol production: a century of research on Saccharomyces cerevisiae metabolism. Appl. Environ. Microbiol. 74, 2259e2266. Heaton, E.A., Dohleman, F.G., Long, S.P., 2008. Meeting US biofuel goals with less land: the potential of Miscanthus. Global Change Bio. 14, 2000e2014. Higashide, W., Li, Y., Yang, Y., Liao, J.C., 2011. Metabolic engineering of Clostridium cellulolyticum for production of isobutanol from cellulose. Appl. Environ. Microbiol. 77, 2727e2733. Ikemura, T., 1985. Codon usage and tRNA content in unicellular and multicellular organisms. Mol. Biol. Evol. 2, 13e14. Jeon, Y.J., Lee, Y.Y., 1989. In situ product separation in butanol fermentation by membrane-assisted extraction. Enzyme Microb. Technol. 11, 575e582. 117 Jin, Y.-S., Alper, H., Yang, Y.-T., Stephanopoulos, G., 2005. Improvement of xylose uptake and ethanol production in recombinant Saccharomyces cerevisiae through an inverse metabolic engineering approach. Appl. Environ. Microbiol. 71, 8249e8256. Khanna, M., Dhungana, B., Clifton-Brown, J., 2008. Costs of producing Miscanthus and switchgrass for bioenergy in Illinois. Biomass Bioenergy 32, 82e493. Kispal, G., Steiner, H., Court, D.A., Rolinski, B., Lill, R., 1996. Mitochondrial and cytosolic branched-chain amino acid transaminases from yeast, homologs of the myc oncogene-regulated Eca39 protein. J. Biol. Chem. 271, 24458e24464. Kondo, T., Tezuka, H., Ishii, J., Matsuda, F., Ogino, C., Kondo, A., 2012. Genetic engineering to enhance the Ehrlich pathway and alter carbon flux for increased isobutanol production from glucose by Saccharomyces cerevisiae. J. Biotechnol. 159, 32e37. Koukiekolo, R., Cho, H.-Y., Kosugi, A., Inui, M., Yukawa, H., Doi, R.Y., 2005. Degradation of corn fiber by Clostridium cellulovorans cellulases and hemicellulases and contribution of scaffolding protein CbpA. Appl. Environ. Microbiol. 71, 3504e3511. Ladisch, M.R., 1991. Fermentation-derived butanol and scenarios for its uses in energy-related applications. Enzyme Microb. Technol. 13, 280e283. Lee, W.-H., Seo, S.-O., Bae, Y.-H., Nan, H., Jin, Y.-S., Seo, J.-H., 2012. Isobutanol production in engineered Saccharomyces cerevisiae by overexpression of 2-ketoisovalerate decarboxylase and valine biosynthetic enzymes. Bioprocess Biosyst. Eng. 35, 1467e1475. Lewandowski, I., Clifton-Brown, J.C., Scurlock, J.M.O., Huisman, W., 2000. Miscanthus: European experience with a novel energy crop. Biomass Bioenergy 19, 209e227. Mariano, A.P., Qureshi, N., Filho, R.M., Ezeji, T.C., 2011. Bioproduction of butanol in bioreactors: new insights from simultaneous in situ butanol recovery to eliminate product toxicity. Biotechnol. Bioeng. 108, 1757e1765. Mariano, A.P., Qureshi, N., Filho, R.M., Ezeji, T.C., 2012. Assessment of in situ butanol recovery by vacuum during acetone butanol ethanol (ABE) fermentation. J. Chem. Technol. Biotechnol. 87, 334e340. McLaughlin, S.B., Kszos, L.A., 2005. Development of switchgrass (Panicum virgatum) as a bioenergy feedstock in the United States. Biomass Bioenergy 28, 515e535. Mosier, N., Wyman, C., Dale, B.E., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., 2005. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96, 673e686. Parekh, S.R., Parekh, R.S., Wayman, M., 1988. Ethanol and butanol production by fermentation of enzymatically saccharified SO2-prehydrolysed lignocellulosics. Enzyme Microb. Technol. 10, 660e668. Pennisetum purpureum, 2013. Pacific Island Ecosystems at Risk (PIER). Available at: <http://www.hear.org/pier/species/pennisetum_ purpureum.htm> (accessed May, 2013). Percudani, R., Pavesi, A., Ottonello, S., 1997. Transfer RNA gene redundancy and translational selection in Saccharomyces cerevisiae. J. Mol. Biol. 268, 322e330. Qureshi, N., Blaschek, H.P., 2000. Economics of butanol fermentation using hyper-butanol producing Clostridium beijerinckii BA101. Trans IChemE 78, Part C, 139e144. Qureshi, N., Saha, B.C., Cotta, M.A., 2007. Butanol production from wheat straw hydrolysate using Clostridium beijerinckii. Bioprocess Biosyst. Eng. 30, 419e427. Qureshi, N., Saha, B.C., Hector, R.E., Hughes, S.R., Cotta, M.A., 2008a. Butanol production from wheat straw by simultaneous saccharification and fermentation using Clostridium beijerinckii: part Idbatch fermentation. Biomass Bioenergy 32, 168e175. Qureshi, N., Saha, B.C., Cotta, M.A., 2008b. Butanol production from wheat straw by simultaneous saccharification and fermentation
118 7. ISOBUTANOL PRODUCTION FROM BIOENERGY CROPS using Clostridium beijerinckii: part IIdfed-batch fermentation. Biomass Bioenergy 32, 785e791. Qureshi, N., Saha, B.C., Dien, B., Hector, R., Cotta, M.A., 2010a. Production of butanol (a biofuel) from agricultural residues: part Iduse of barley straw hydrolysate. Biomass Bioenergy 34, 559e565. Qureshi, N., Saha, B.C., Hector, R.E., Dien, B., Hughes, S.R., Liu, S., Iten, L., Bowman, M.J., Sarath, G., Cotta, M.A., 2010b. Production of butanol (a biofuel) from agricultural residues: part IIduse of corn stover and switchgrass hydrolysates. Biomass Bioenergy 34, 566e571. Qureshi, N., Liu, S., Ezeji, T.C., 2013. Cellulosic butanol production from agricultural biomass and residues: recent advances in technology. In: Lee, J. (Ed.), Advanced Biofuels and Bioproducts. Springer Science þ Business Media, New York, pp. 247e265. Ryan, E.D., Kohlhaw, G.B., 1974. Subcellular localization of isoleucinevaline biosynthetic enzymes in yeast. J. Bacteriol. 120, 631e637. Santos, E.,L., Rodriguez, M., Elorza, V., Sentandreu, R., 1982. Uptake of sucrose by Saccharomyces cerevisiae. Arch. Biochem. Biophys. 216, 652e660. Singh, R., Kunkee, R., 1976. Alcohol dehydrogenase activities of wine yeasts in relation to higher alcohol formation. Appl. Environ. Microbiol. 32, 666e670. Smith, K.M., Liao, J.C., 2011. An evolutionary strategy for isobutanol production strain development in Escherichia coli. Metab. Eng. 13, 674e681. Smith, K.M., Cho, K.-M., Liao, J.C., 2010. Engineering Corynebacterium glutamicum for isobutanol production. Appl. Microbiol. Biotechnol. 87, 1045e1055. Suomalainen, H., Keranen, A.J.A., 1967. Keto acids formed by Baker’s yeast. J. Inst. Brew. 73, 477e484. Takara, D., Khanal, S.K., 2011. Green processing of tropical banagrass into biofuel and biobased products: an innovative biorefinery approach. Bioresour. Technol. 102, 1587e1592. Ter Schure, E.G., Flikweert, M.T., van Dijken, J.P., Pronk, J.T., Verrips, C.T., 1998. Pyruvate decarboxylase catalyzes decarboxylation of branched-chain 2-oxo acids but is not essential for fusel alcohol production by Saccharomyces cerevisiae. Appl. Environ. Microbiol. 64, 1303e1307. Valvani, S.C., Yalkowsky, S.H., Rosemand, T.J., 1981. Solubility and Partitioning. IV. Aqueous Solubility and Octanol-Water Partition Coefficients of Liquid Non-electrolytes. J. Pharm. Sci. 70, 502e507. Van Diken, J.P., Scheffers, W.A., 1986. Redox balances in the metabolism of sugars by yeasts. FEMS Microbiol. Rev. 32, 199e224. Varman, A.M., Xiao, Y., Pakrasi, H.B., Tang, Y.J., 2013. Metabolic engineering of Synechocystis sp. strain PCC 6803 for isobutanol production. Appl. Environ. Microbiol. 79, 908e914. Velasco, J.A., Cansado, J., Pena, M.C., Kawakami, T., Laborda, J., Notario, V., 1993. Cloning of the dihydroxyacid dehydrataseencoding gene (ILV3) from Saccharomyces cerevisiae. Gene 137, 179e185. Vervoort, E.B., van Ravestein, A., van Peij, N.N.M.E., Heikoop, J.C., van Haastert, P.J.M., Verheijden, G.F., Linskens, M.H.K., 2000. Optimizing heterologous expression in Dictyostelium: importance of 5’ codon adaptation. Nucleic Acids Res. 28, 2069e2074. Weusthuis, R.A., Pronk, J.T., van den Broek, P.J.A., van Dijken, J.P., 1994. Chemostat cultivation as a tool for studies on sugar transport in yeasts. Microbiol. Rev. 58, 616e630. Wisselink, H.W., Toirkens, M.J., del Rosario Franco Berriel, M., Winkler, A.A., van Dijken, J.P., Pronk, J.T., van Maris, A.J.A., 2007. Engineering of Saccharomyces cerevisiae for efficient anaerobic alcoholic fermentation of L-arabinose. Appl. Environ. Microbiol. 73, 4881e4891. Zhang, M., Eddy, C., Deanda, K., Finkestein, M., Picataggio, S., 1995. Metabolic engineering of a pentose metabolism pathway in ethanologenic Zymomonas mobilis. Science 267, 240e243.
C H A P T E R 8 Lipase-Catalyzed Biodiesel Production: Technical Challenges Rama Raju Baadhe 1, Ravichandra Potumarthi 2,*, Vijai K. Gupta 3 1 Department of Biotechnology, National Institute of Technology, Warangal, Andhra Pradesh, India, 2 Department of Chemical Engineering, Monash University, Clayton, Victoria, Australia, 3 Molecular Glycobiotechnology Group, Department of Biochemistry, School of Natural Sciences, National University of Ireland Galway, Galway, Ireland *Corresponding author email: ravichandra.potumarthi@monash.edu; pravichandra@gmail.com O U T L I N E Introduction 119 Chemistry of Biodiesel 120 Transesterification 120 Disadvantages of Chemical Transesterification 120 Advantages of Using Lipases in Biodiesel Production 121 Historical Background of Lipase 121 Animal Oils/Fats Waste Oils/Fats Algae Oils 123 123 124 Choice of Enzyme 124 Molar Ratio (Alcohol/Oil) 124 Temperature 124 Water Content 126 Acyl Acceptors 126 Lipase-Catalyzed Transesterification Done in Two Approaches 121 Solvents 126 Advantages of Immobilized Lipase 122 Reactor System 126 Technical Challenges 123 Conclusions 127 Feedstock Vegetable Oils 123 123 References 127 INTRODUCTION World’s commercial primary energy needs are mostly supplied through fossil fuels and accounts about 87% of total energy source (OPEC, 2011, 2012). Primary energy demand by 2035 increases to 54% and still fossil fuels contributes 82% of the global total by 2035 (OPEC, 2012). All fossil-fuel sources are finite and if the crude oil consumption continued at current usage rates, it will last Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00008-5 only for 54.2 years (British Petroleum Statistical Review, 2012). Projected demand for oil reaches 110 mb/day by 2035. Among the fossil fuel, diesel fuels have an essential function in the industrial, transportation and agricultural sectors in developing countries. Gradual depletion of crude oil and emission of greenhouse gases in to the environment triggers the alarm for suitable alternative fuels for use in diesel engines (Ganesan et al., 2009). Biodiesel is one of the attractive and 119 Copyright Ó 2014 Elsevier B.V. All rights reserved.
120 8. LIPASE-CATALYZED BIODIESEL PRODUCTION: TECHNICAL CHALLENGES alternative fuels along with bioethanol. Biodiesel or fatty acid methyl esters (FAMEs) are mono-alkyl esters of long-chain fatty acids, derived from transesterification of triglycerides (plant or animal or algal origin). It can be used directly in its pure form or as a blend with conventional diesel fuel in diesel engines (Ma and Hanna, 1999). This fuel is biodegradable and nontoxic and has low emission profiles when compared to petroleum diesel (Krawczyk, 1996). But the cost of biodiesel, however, is the main obstacle to commercialize the product. There are four primary ways of making biodiesel: direct use and blending (Ma and Hanna, 1999), micro emulsions (Schwab et al., 1987), thermal cracking (pyrolysis) (Sonntag, 1979) and transesterification. However, the first three have some limitations and drawbacks in case of physiochemical properties of biodiesel (Schwab et al., 1987). Transesterification is the well-known method and involves conversion of oils or fat to FAMEs or fatty acid ethyl esters in the presence of a catalyst such as acid, base or lipase (Bisen et al., 2010). The conventional method for producing biodiesel involves acid and base catalysts to form fatty acid alkyl esters. Processing expenses and environmental concerns associated with biodiesel production and difficulties connected with by-products recovery have led to the search for alternative production methods and alternative sources (Bisen et al., 2010). Enzyme-mediated transesterification can be a moderate alternative to produce biodiesel in its pure form, which also makes its separation easy against the by-product (glycerol). But still due to the cost of enzyme, commercialization of biodiesel has not come to reality. Though there are many attempts made for biodiesel production through enzyme-mediated method (Ranganathan et al., 2008; Sanchez and Vasudevan, 2006; Lai et al., 2005; Noureddini et al., 2005; De Oliveira et al., 2004; Xu et al., 2004; Sha et al., 2003; Belafi-Bako et al., 2002; Iso et al., 2001; Fukuda et al., 2001; Freedman et al., 1984), profitable commercial production was not achieved for industrial utilization. Recombinant DNA and protein engineering technologies improved the quantities and catalytic efficiency of lipase (Akoh et al., 2007). There are several technical challenges that need to be addressed to make biodiesel production profitable. Some of them associated with enzyme transesterification process. In this chapter, some of the technical challenges 3 CH2–O–CO–R 2 H–C–O–CO–R′ 1 CH2–O–CO–R′′ 1,2,3,-O-tri-acyl -glycerol R′, Rand R′′ are saturated or unsaturated chains FIGURE 8.1 Chemical structure of triacylglycerol. Source: Bisen et al. (2010). involved in the lipase-catalyzed biodiesel production were discussed. CHEMISTRY OF BIODIESEL Chemically biodiesel is defined as mono-alkyl (methyl or ethyl) esters of triacylglycerol. All vegetable oils, algal lipids and animal fats (triacylglycerol/triglyceride molecules) consist of a three-carbon chain forms the glycerol backbone, which consists of three long fatty acid chains (Figure 8.1). Amounts of each fatty acid present in molecules determine the properties of triacylglycerol (Knothe, 2001). TRANSESTERIFICATION In chemical transesterification process, fatty acid reacts with any alcohol and forms mono-alkyl ester (biodiesel) in the presence of a catalyst (acid, base and enzyme). General reaction scheme of biodiesel production is shown in Figure 8.2. The reaction has two inputs: triacylglycerol and the alcoholdcommonly ethanol or methanol is used (Meher et al., 2004). DISADVANTAGES OF CHEMICAL TRANSESTERIFICATION • Requires high reaction temperatures. • Soap formation: in the base-catalyzed transesterification process, free fatty acid (FFAs) level of feedstock should be less, otherwise it will result in too much soaps formation. • Recovery of by-product: purification of glycerol is very difficult. FIGURE 8.2 The scheme of transesterification reaction: (R ¼ CH3) alcohol is methanol.
LIPASE-CATALYZED TRANSESTERIFICATION DONE IN TWO APPROACHES • Pretreatment step needed: FFAs level of the feedstocks should not exceed 3 wt%, beyond which it has to undergo pretreatment steps before transesterification (Leung et al., 2010). • Yield of methyl esters: yields of the methyl esters are lower compared to enzymatic transesterification. • Purification of methyl esters: purification of methyl esters requires repeated washing which increases process operational cost. • Less active: since alkali catalysts (NaOH and KOH) are inexpensive, they are preferred but activity is less (Demirbas, 2008). • Energy consumption: alkali-catalyzed transesterification needed large energy consumption during downstream biodiesel refining process (Madras and Kolluru, 2004). • Corrosion: when H2SO4 is used as catalyst, it leads to corrosion of the reactor and huge wastewater generated during neutralization of mineral acid (Atadashi et al., 2013). • Use of homogenous catalysts makes biodiesel product separation difficulty and recovery of catalyst cumbersome (Atadashi et al., 2013). • Acid-catalyzed transesterification reaction needs higher alcohol-to-oil molar ratios (Atadashi et al., 2013). • In base-catalyzed transesterification reaction, large amount of catalyst is needed. Difficulties arise during chemical catalysis can be overcome by enzyme-mediated (biocatalysts) transesterification and they are becoming increasingly important in biodiesel preparation due to their ability to beat chemical catalysts. Lipases (E.C.3.1.1.3) are widely considered as biocatalysts to catalyze transesterification and esterification reactions. ADVANTAGES OF USING LIPASES IN BIODIESEL PRODUCTION Advantages of using lipases in biodiesel production are the following: • Ability to work under different media environments includes biphasic system and monophasic system (aqueous and nonaqueous) (Mittelbach, 1990; Linko et al., 1994; Mukesh et al., 1994). • They can be produced in bulk and are sturdy and adaptable enzymes. • Separation is not necessary if transesterification process with lipase carried out in a packed-bed reactor. • Proficiencies like short-chain, alcohol-tolerant and higher thermo stability of lipase make it very appropriate for use in biodiesel production (Ghaly et al., 2010). 121 • Thermal stability of lipases makes it possible to run the transesterification process at elevated temperatures which allows (1) increased solubility of lipids and other hydrophobic substrates in water; (2) higher diffusion rates; (3) decreased substrate viscosities; (4) increased reactant solubilities; (5) faster reaction rates; and (6) reduced risk of microbial contamination. HISTORICAL BACKGROUND OF LIPASE Over 300 years ago, triglycerides hydrolyzing enzymes have been studied well. Nearly 70 years ago, lipase’s catalysis ability and synthesis ability have been known. In 1856, Claude Bernard first revealed that pancreatic juice contains an enzyme (lipase) that hydrolyzed the insoluble oil droplets and converted them in to soluble products. In 1901 activity of microbial lipases has been observed in Bacillus prodigiosus, Bacillus pyocyaneus and Bacillus fluorescens. Serratia marcescens, Pseudomonas aeruginosa and Pseudomonas fluorescens are the best-studied lipase-producing bacteria (Fariha et al., 2006). Lipase-catalyzed biodiesel production was reported first by Mittelbach (1990). Depending upon the specificity, lipases are divided into three groups: (1) 1,3-specific, (2) fatty acid-specific, and (3) nonspecific. Among these three, 1,3-specific lipases discharge fatty acids from positions 1 and 3 of a glyceride and hydrolyze ester bonds in these positions (Antczak et al., 2009; Ribeiro et al., 2011). LIPASE-CATALYZED TRANSESTERIFICATION DONE IN TWO APPROACHES 1. Extracellular/free lipase lipases (i.e. recovered and purified from the cultivation broth): For industrial level production of extracellular lipases, bacteria, yeast and fungi are preferred. Lipases from different sources are able to catalyze the same reaction (Table 8.1). Bacterial and fungal lipases are mostly used in biodiesel production and recently, Streptomyces sp. was proved as an effective lipase producing microbe and the enzyme produced was appropriate for biodiesel production (Cho et al., 2012). Use of free or extracellular enzymes was limited due to their price. Cost of the enzyme was increased due to their specific separation and purification techniques. Extracellular lipases are soluble enzymes and they are dispersed in the solution and can move freely during the catalytic reaction, thus difficult to handle and reuse (Iso et al., 2001). One auspicious approach to overcome this
122 8. LIPASE-CATALYZED BIODIESEL PRODUCTION: TECHNICAL CHALLENGES TABLE 8.1 Different Microbial Sources for Lipases Used in Biodiesel Production Lipase Producing Microorganisms References Pseudomonas fluorescens Iso et al. (2001) Pseudomonas cepacia Noureddini et al. (2005) Candida antarctica Nelson et al. (1996) Rhizopus delemar Nelson et al. (1996) Rhizopus oryzae Ghamgui et al. (2004) Mucor miehei Nelson et al. (1996) Geotrichum candidum Nelson et al. (1996) Candida rugosa Shimada et al. (2002); Ma et al. (2002); Chowdary and Prapulla (2002) Rhizomucor miehei Soumanou and Bornscheuer (2003a,b) Thermomyces lanuginosa Iso et al. (2001); Xu et al. (2003); Soumanou and Bornscheuer (2003a,b); Du et al. (2003) Aspergillus niger Haas et al. (2002) Pseudomonas cepacia Noureddini et al. (2005) Chromobacterium viscosum Yahya et al. (1998) Photobacterium lipolyticum Yahya et al. (1998) Streptomyces sp. Yahya et al. (1998); Cho et al. (2012) TABLE 2 difficulty is to immobilize the enzyme in a way that can be separated and reused later by using simple separation methods like centrifugation and filtration (Iso et al., 2001; Cao, 2005). 2. Intracellular/immobilized lipases; i.e. lipases remain either inside or attached to the cell wall. In this case, enzyme is immobilized (naturally) directly or together with the whole cell (intracellular). This strategy eliminates downstream operations and promises the recycling of enzymes. Alternatively lipases can be immobilized synthetically by different mechanisms. Immobilization restricts movement of the enzyme and constrains its location to an inert support or a carrier (Cao, 2005). Various methods like adsorption, covalent bonding, entrapment, and cross-linking are available for enzyme immobilization. The choice of method and support material is a protuberant factor for obtaining an efficient lipase (Table 8.2) (Sevil et al., 2012). ADVANTAGES OF IMMOBILIZED LIPASE 1. Enzyme becomes more stable. 2. Immobilization of enzyme increases surface area of biocatalyst. 3. Option of regeneration and reuse of the immobilized lipase. 4. Protection from solvent inhibition. Comparison of Different Immobilization Methods Characteristics Entrapment Covalent Bonding Cross-Linking Adsorption Protection from microbes Yes No Possible No Immobilized process Difficult Difficult Difficult Easy Interaction Strong Strong Strong Weak Nature of immobilization Irreversible Irreversible Irreversible Reversible Recovery of lipase activity High Low Moderate Low Enzyme leakage Yes No No Yes Regeneration of immobilized lipase Impossible Impossible Impossible Possible Immobilized cost Low High Moderate Low Immobilization efficiency depend on Nature of the polymer Surface -CHO groups of support and surface amino acid residues of the enzyme Interaction between enzyme and the carrier pH, temperature and ionic strength. Enzyme stability High High High Low Active site Not effected Effected Effected Not effected Mass transfer limitation Yes No No No
FEEDSTOCK 5. Separation of product and enzyme is easier. 6. Avoids contamination of enzyme or whole cell. 7. Rigid external support expected to increase optimal temperature, thereby fasten reaction rate. TECHNICAL CHALLENGES Yield of biodiesel through lipase catalysis is effected by (1) feedstock quality, (2) choice of enzyme (extracellular or intracellular), (3) molar ratio (alcohol/oil), (4) temperature, (5) water content, (6) acyl acceptors, (7) solvent and (8) reactor system. FEEDSTOCK One of the main barriers for commercialization of biodiesel production is choice and availability of feedstock, which comprise nearly 80% of the overall biodiesel production cost. Diverse kinds of feedstocks are available such as edible and nonedible vegetable oil, animal fats, waste oil, microbial oil and microalgae oil and they can be used for enzyme-catalyzed transesterification (Sevil et al., 2012). Vegetable Oils Vegetable oils are well-known for their high heat content and they are alternative fuels for diesel engines. High viscosity restricts their consumption directly in diesel engines, which leads to many problems (Koh and Ghazi, 2011; Singh and Singh, 2010). Most widely used edible vegetable oils in enzymatic transesterification are soybean (Wenlei and Ning, 2010; Du et al., 2003), sunflower (Karout and Pierre, 2009), palm (Talukder et al., 2011; Matassoli et al., 2009), corn (Mata et al., 2012), cottonseed (Chattopadhyay et al., 2011), canola (Jang et al., 2012) and olive (Sanchez and Vasudevan, 2006). Higher quality of edible oil is good feedstock to produce biodiesel by enzymatic transesterification. However, major concern is the economic viability of biodiesel since refined vegetable oils are expensive. Also, use of high-value edible vegetable oil as biodiesel feedstocks has caused food crisis. Furthermore, percentage of oil and yield per hectare are effective parameters in selecting potential renewable feedstock for biodiesel production (Nielsen et al., 2008). Hence, in order to make biodiesel production more economical, low-cost and nonedible oils need to be preferred. Babassu (Orbignya martiana), Jatropha curcas (Linnaeus), neem (Azadirachta indica), polanga (Calophyllum inophyllum), karanja (Pongamia pinnata), rubber seed tree (Hevea brasiliensis), mahua (Madhuca indica and Madhuca longifolia), tobacco (Nicotiana tabacum), etc. are most widely used nonedible oil sources for biodiesel 123 production. Biodiesel produced from these nonedible oils meets key specifications of biodiesel as per the standard organization requirements (Mohibbe et al., 2005). All these low-cost feedstocks contain large amount of FFA which leads to undesirable soap formation during traditional base-catalyzed transesterification. However, high free-acid content is not a problem in enzyme transesterification. Animal Oils/Fats Animal oils differ from vegetable oil in their fatty acid composition. Vegetable oils have high content of unsaturated fatty acids (mainly oleic and linoleic acid), while animal fat has higher proportion of saturated fatty acids. Commonly used animal fats for biodiesel production via enzymatic route contains lard (Jike et al., 2007), lamb meet, beef tallow, chicken fat and animal fat mix (Vivian Feddern et al., 2011). Waste animal fats from animal processing industries and slaughter houses are also a good source for animal fats; however it is a decent alternative instead of their direct dispose in to environment. Their favorable features like noncorrosive nature, high cetane number, and renewable nature makes them a good source for biodiesel production. But their relatively high FFA (5e30%) and water content led to soap formation in chemical transesterification process and their saturated fats prone them to oxidation and crystallization at high temperatures (Huynh et al., 2011). Removal of contaminants is another problem from animal fats, which generally contain phospholipids, or gums, and cause insoluble precipitates when they come into contact with water. Gums are removed by adding water and citric or phosphoric acid to the animal fats followed by centrifugal separation of precipitates. Phospholipids get separated with glycerin during processing, or by water washing/ion exchange separation. Removal of sulfur contents is also a serious issue. Beef tallow and some chicken fat contain around 100 ppm of sulfur. Vacuum distillation is the only reliable technique for reducing the sulfur level to permissible levels (15 ppm) (Farm energy, 2012). Waste Oils/Fats Used edible oils generally recycled as animal feed or used as a raw material for lubricant and paints and the rest discharged into the environment (Watanabe et al., 2001). To eliminate environment and human health risk caused by waste oils/fats (Chen et al., 2006) and to lower biodiesel production cost, usage of waste oils/fats for biodiesel production is recommended (Watanabe et al., 2001). Waste cooking oil, animal fats, yellow grease, brown grease and waste from vegetable oil refining industries are major sources of waste oil for biodiesel
124 8. LIPASE-CATALYZED BIODIESEL PRODUCTION: TECHNICAL CHALLENGES production (Huynh et al., 2011). Waste oils are rich in high percentage of FFA and high water content, so lipasemediated transesterification is a promising method for production of biodiesel with high yields (Huynh et al., 2011). Novozym 435 is capable of converting used olive oils (Sanchez and Vasudevan, 2006). Novozym 435 is capable of converting used olive oils in to biodiesel (Sanchez and Vasudevan 2006). Algae Oils Usage of algae oil for biodiesel production has considerable interest because of their availability costs compared to edible oils and animal fats. High photosynthetic rate, rapid growth rate, and high productivity make algae a good renewable source for oil/fats. High lipid content (20e40%), tolerance to water, and smaller land usage up to 132 times less compared to terrestrial oil crops make them more prominent choice for oil source (Karatay and Donmez, 2011). Usage of algae oil could reduce the food scarcity problem caused by bioenergy crops (Mata et al., 2010). However, technological development is needed to improve the microalgae oil extraction processes. CHOICE OF ENZYME Lipases from bacteria and fungi are the most commonly used for transesterification. In general at reaction temperatures 30e50  C, the best enzymes will show conversions above 90%. Catalytic reaction time, alcohol type and enzyme condition (free enzyme or immobilized) are also the crucial parameters for selecting enzyme (Table 8.3). Immobilized Pseudomonas cepacia lipases converts the jatropha oil into FAME in 8 h with ethanol; for the same free enzyme, it took 90 h for transesterifying soybean oil with methanol. MOLAR RATIO (ALCOHOL/OIL) According to the stoichiometric ratio of transesterification, it requires 3 mol of alcohol and 1 mol of triglyceride to yield 3 mol of fatty acid alkyl esters and 1 mol of glycerol. Since it is a reversible reaction, traditional transesterification process biodiesel yield always improved due to excess amount of alcohol over fatty acids in triglycerides (Sevil et al., 2012). But for enzyme-catalyzed transesterification, methanol exhibits negative effects on enzyme activity thus decrease the production yield. High molar ratio of alcohol to triglycerides increases the glycerol solubility and effects the separation of glycerol. Glycerol in solution, drives the equilibrium back to the left, lowering the yield of alkyl esters. Apart from this, highly hydrophilic nature of alcohol eliminates water layer essential for the enzymes and inactivates them. Gradual mixing of alcohol is a potential approach for molar ratio optimization in solvent-free systems. Most of the studies show molar ratio 9:1, which seems to be the most appropriate. Addition of cosolvents, such as n-hexane and tetrahydrofuran, slightly discharges this problem and increases the reaction rates, but a more complicated process should apply for separation of biodiesel (Fan, 2012; Meher et al., 2006). A two-step reaction system was reported to avoid the inactivation of lipase by addition of excess amounts of methanol (Watanabe et al., 2007). Moreno-Pirajan done experiments on different molar ratios of methanol and ethanol and his studies with various alcohol types and palm oil with 10.4 M ratios by using Candida rugosa lipase yielded 85 mol% of methyl esters (Moreno-Pirajan and Giraldo, 2011). Zaidi et al. 2002 reported inhibition coefficient of alcohol increased from 0.034 mol/l to 0.42 mol/l, when number of carbon atoms increased from 1 (methanol) to 18 (oleyl alcohol), respectively. Dizge and Keskinler (2008) conducted experiments at different molar ratios of methanol to canola oil and proved that up to certain molar ratio (1:6), there is an increase in ester production; upon increasing molar ratio, it has shown the negative effect by decreasing the formation of esters due to enzyme inactivation. Finally amount of alcohol needed varies suggestively depending on the origin of the lipase and triglycerides. Optimization of molar ratio is a big technical challenge and there is much scope in this area for designing the molar ratios for different types of lipids and enzyme (Sevil et al., 2012). TEMPERATURE Transesterification can occur at different temperatures varying from 25  C to 60  C depending on the oil used and many studies reported the effect of temperature on transesterification which influences reaction rate and yield of esters. Generally high temperatures increase the ester yields (Freedman et al., 1984). However, increased temperature beyond optimum point promotes the denaturation and higher thermal deactivation of enzyme, thus decreases catalytic activity (Sevil et al., 2012). In batch process, optimal temperature range was 40e50  C whereas in repeated batch process, it lost its activity at 40  C (Kose et al., 2002). Various research groups worked to find out the effect of temperature on biodiesel production with immobilized enzymes. Iso et al.’s (2001) studies on the transesterification reaction using free and immobilized lipase produced by P. fluorescens at a temperature range from 40  C to 70  C revealed that conversion rate was the highest at 60  C,
TABLE 8.3 Different Lipase-Catalyzed Biodiesel Production Process Conditions Oil/Fat Alcohol Yield (%) Pseudomonas fluorescens Soybean oil Methanol 90 35 Kaieda et al. (2001) Immobilized Candida antarctica lipase B (Novozym 435) Jatropha oil Ethyl acetate 91.3 e Modi et al. (2007) Candida antarctica lipase Cottonseed oil Methanol 97 e Royon et al. (2007) Candida antarctica Tallow Methanol 74 e Lee et al. (2002) Immobilized Mucor miehei lipase (Lipozyme IM-20) Mowrah, mango, kernel, sal C4eC18:1 alcohols 86.8e99.2 e De et al. (1999) Candida antarctica lipase (lipase SP-435) Sunflower Secondary alcohols 61.2e83.8 e Mittelbach, (1990) Rhizopus miehei lipase (Lipozyme IM-60) Methanol 19.4 e Rhizopus miehei lipase (Lipozyme IM-60) Ethanol 65.5 e Pseudomonas fluorescens lipase Methanol 3 e References Candida rugosa Waste ABE Methanol, ethanol, 1-propanol, 3,1-butanol, iso-butanol, isoamylalcohol, and n-octanol e e Noureddini et al. (2005) Immobilized Candida sp. 99e125 Salad Methanol e 40 Nie et al. (2006) Novozym 435, Lipozyme TL IM and Lipozyme RM IM Soybean Ethanol e 25 Hernandez-Martin and Otero (2008) Candida sp. 99e125 Waste cooking Methanol e 40e50 Chen et al. (2009) Novozym 435 Rapeseed Methanol e 40 Jeong and Park (2008) Novozym 435 Tung and palm Methanol and ethanol e 55 Wang et al. (2011a,b) Novozym 435 Cottonseed Dimethyl carbonate as organic solvent e 50 Su et al. (2007) Novozym 435 Canalo Methanol e 38 Chang et al. (2005) Novozym 435 Olive Methanol e 40 Sanchez and Vasudevan (2006) Novozym 435 Soybean T-amyl e 40 Zheng et al. (2009) Novozym 435 Sunflower Methanol e 45 Ognjanovic et al. (2009) Novozym 435 Stillingia Methanol e 40 Liu et al. (2009) TEMPERATURE Lipase Optimum Temperature ( C) 125
126 8. LIPASE-CATALYZED BIODIESEL PRODUCTION: TECHNICAL CHALLENGES whereas free lipase activity highly decreased at 70  C and immobilized enzyme activity remained active. This work reveals the fact that increase in thermal stability of enzyme is due to immobilization enzyme and rigid external carrier provides temperature resistance for lipase molecule. The studies about the effect of temperature for lipase transesterification are shown in Table 8.3. WATER CONTENT Water content plays a key role in enzymatic transesterification as it is vital to sustain the threedimensional conformations of enzyme catalytic site. Presence of an oilewater interface creates a favorable environment for the conformation of active site (AlZuhair et al., 2006, 2003). Water interacts with the enzyme hydrophilic groups located on surface, and changes the conformation of hydrogen bond interactions inside enzyme, leading to transformation of lipase active (Gao et al., 2006). Generally lipase activity increases with increase in water content up to 15% (w/w of oils). Beyond 15%, the conversion rate decreased slightly. But 20% of water content also efficiently catalyzed alcoholysis using lipases from Rhizopus delemar and Rhizomucor miehei (Tweddell et al., 1998). About 5% of initial water content was suggested as optimum for biodiesel production from jatropha oil using various lipases (Shah and Gupta, 2007). Thus, the optimum level of initial water (moisture) is based on the type of biocatalyst and reaction conditions. ACYL ACCEPTORS Generally transesterification reactions are conducted using straight- and branched-chain alcohols. Because of abundant availability and low cost, methanol is the widely used short-chain alcohol acyl acceptor for biodiesel production (Fan, 2012). The negative effect of methanol on enzyme activity alleviates by stepwise addition of alcohols. Ethanol, n-butanol and i-butanol, n-amylalcohol and i-amylalcohol, and n-propanol were also used during transesterification. But increase in C number of the alcohols has not significantly influenced fatty acid ester contents and shown the negative effect (Soumanou and Bornscheuer, 2003a,b). Also, it is generally believed that primary alcohols are more suitable than secondary alcohols and alcohols with less than eight carbon atoms can be used under the conditions that gave the highest conversion of the oils to FAME. Methyl acetate had no negative effect on enzymatic activity. No changes were detected in lipase activity even after being continuously used for 100 batches (Sulaiman, 2007). Recently, ethyl acetate, methyl acetate, butyl acetate, vinyl acetate and dimethyl carbonate (DMC) are considered as novel acyl acceptors. The work revealed by Er-Zheng et al. (2007) proved that DMC gives two- to threefold higher conversion than those of conventional acyl acceptors (methanol and methyl acetate) and is also ecofriendly, neutral, odorless, cheap, noncorrosive, nontoxic, and exhibits good solvent properties. SOLVENTS In the above section (molar ratio) discussed that excess amount of alcohols increases FAME yield. In order to increase the solubility of alcohol (not the enzyme), solvents are used and they alleviate negative effect of methanol on the catalyst and precede the transesterification. Enzyme should be insoluble in solvent; otherwise, it will not be active (Kanerva et al., 1990; Antczak et al., 2009). Various hydrophilic and hydrophobic organic solvents such as cyclohexane, n-hexane, tert-butanol, petroleum ether, isooctane and 1,4-dioxane are mainly studied organic solvents in enzymatic biodiesel production. If organic solvent is used as medium, overall alcohol is added at the beginning of the reaction. In solvent-free reaction medium, alcohol is added stepwise to prevent enzyme activity with high alcohol concentration (Sevil et al., 2012). REACTOR SYSTEM Development of enzymatic biodiesel production at commercial scale is dependent on the reactor systems. Various reactors, including batch reactors, packed-bed reactors and supercritical reactors, are studied for biodiesel production. Most of the studies have done on batch reactors and packed-bed reactors. Batch reactors are simple to use in the laboratory. But shear stress caused by stirrer would disrupt the enzyme life (Tan et al., 2010). Batch operation is a laborious process and is not suitable for automation (Chen et al., 2010). Packed-bed reactors are continuous and are a good alternative for batch reactors to lower the shear stresses (long-term enzyme stability) and to make the process economical (Wang et al., 2010). In addition, this system offers high bed volume and is simple to scale up (Hama et al., 2011). Because of its continuous mode, stepwise addition of alcohol is possible in order to reduce the inactivation of the enzyme caused by excess alcohol. Lipase inhibition due to the cloggage by glycerol accumulation inside the reactor is a major challenge (Xu et al., 2012). This can be resolved using more than one column in the reactor. Yoshida et al. (2012) developed a reactor in which a reactant solution is pumped through a column containing immobilized recombinant
REFERENCES Aspergillus oryzae and the effluent from the column is recycled into the same column with a stepwise addition of methanol. This reactor system gave better lipase activity up to five cycles with 96.1% FAME content. Wang et al. (2011b) developed a four-packed-bed reactor in order to provide longer residence time to the reaction mixture in the reactor and to lower lipase inhibition by product accumulation. A single-packed-bed reactor and the four-packed-bed reactor were used to produce biodiesel by using refined soybean oil with P. cepacia lipase. Over 88% conversion rate and great stability were achieved with the four-packed-bed reactor compared to single-packed-bed reactor (Wang et al., 2010). This process improved the reaction efficiency and additionally, the cost of biodiesel production can be reduced by effective recycling of enzyme (Fjerbaek et al., 2009). Supercritical reactors are also investigated for enzymatic biodiesel production. D. Oliveira and J.V. Oliveira (2001) produced biodiesel from palm kernel oil in the presence of Novozym 435 and Lipozyme IM in supercritical carbon dioxide. Lipozyme IM showed better conversion (77.5%). But the problem is high pressure (beyond 200 bar) used in this process. Study by Taher et al. (2011) has given only 49.2% conversion rate with lamb-meat fat in supercritical carbon dioxide by Novozym 435. Supercritical reactors could not be commercialized due to low conversion rate and high cost of the system. Subsequently, a technically improved packed-bed reactor system with high transesterification efficiency is a good alternative for industrial scale-up of enzymatic biodiesel production in an economic way. CONCLUSIONS Energy crisis and environmental concerns raised the necessity for the new biofuels. Biodiesel is a clean alternative to fossil fuel. A green approach for biodiesel production through enzymatic biodiesel production has gained a lot of attention due to the drawbacks of chemical methods. Promising enzymatic processes are established for biodiesel production. The main obstacle for the industrialization of enzymatic process would be overall cost of production. Production cost could be reduced by increasing the productivity or by increasing the catalytic efficiency of lipases. Immobilization and genetic engineering methods appear to be an attractive way to obtain more active, stable, and reusable lipases in different reaction systems. Operational parameters like water content, temperature, solvent, acyl acceptors, and so on plays key role in transesterification process. Along with all these technical operational conditions, novel bioreactor designing has also promising challenges 127 in order to make biodiesel a great potential commercial fuel in future. References Akoh, C.C., Chang, S.W., Lee, G.C., Shaw, J.F., 2007. Enzymatic approach to biodiesel production. J. Agric. Food Chem. 55, 8995e9005. Al-Zuhair, S., Jayaraman, K.V., Krishnan, S., Chan, W.H., 2006. The effect of fatty acid concentration and water content on the production of biodiesel by lipase. Biochem. Eng. J. 30, 212e217. Al-Zuhair, S., Hasan, M., Ramachandran, K.B., 2003. Kinetic hydrolysis of palm oil using lipase. Process Biochem. 38, 1155e1163. Antczak, M.S., Kubiak, A., Antczak, T., Bielecki, S., 2009. Enzymatic biodiesel synthesisdkey factors affecting efficiency of the process. Renewable Energy 34, 1185e1194. Atadashi, I.M., Aroua, M.K., Abdul Aziz, A.R., Sulaiman, N.M.N., 2013. The effects of catalysts in biodiesel production: a review. J. Ind. Eng. Chem. 19, 14e26. Belafi-Bako, K., Kovacs, F., Gubicza, L., Hancsok, J., 2002. Enzymatic biodiesel production from sunflower oil by Candida antarctica lipase in a solvent-free system. Biocatal. Biotransform. 20, 437e439. Bisen, P.S., Sanodiya, B.S., Thakur, G.S., Bhagel, R.K., Prasad, G.B.K.S., 2010. Biodiesel production with special emphasis on lipase catalysed transesterification. Biotechnol. Lett. 32, 1019e1030. British Petroleum Statistical Review of World Energy, 2012. (Accessed on 22 August, 2012) available on http://www.bp.com/ sectionbodycopy.do?categoryId¼7500&contentId¼7068481. Cao, L., 2005. Immobilised enzymes: science or art? Curr. Opin. Chem. Bio. 9, 217e226. Chang, H.M., Liao, H.F., Lee, C.C., Shieh, C.J., 2005. Optimised synthesis of lipase-catalyzed biodiesel by Novozym 435. J. Chem. Technol. Biotechnol. 80, 307e312. Chattopadhyay, S., Karemore, A., Das, S., Deysarkar, A., Sen, R., 2011. Biocatalytic production of biodiesel from cottonseed oil: standardization of process parameters and comparison of fuel characteristics. Appl. Energy 88, 1251e1256. Chen, G., Ying, M., Li, W., 2006. Enzymatic conversion of waste cooking oils in to alternative fuel-biodiesel. Appl. Biochem. Biotechnol. 129e132, 911e921. Chen, Y.H., Huang, Y.H., Lin, R.H., Shang, N.C., 2010. A continuousflow biodiesel production process using a rotating packed bed. Bioresour. Technol. 101, 668e673. Chen, Y., Xiao, B., Chang, J., Fu, Y., Lv, P., Wang, X., 2009. Synthesis of biodiesel from waste cooking oil using immobilized lipase in fixed bed reactor. Energy Convers. Manage. 50, 668e673. Cho, S.S., Park, D.J., Simkhada, J.R., Hong, J.H., Sohng, J.K., Lee, O.H., Yoo, J.C., 2012. A neutral lipase applicable in biodiesel production from a newly isolated Streptomyces sp. CS326. Bioprocess Biosyst. Eng. 35, 227e234. Chowdary, G.V., Prapulla, S.G., 2002. The influence of water activity on the lipase catalyzed synthesis of butyl butyrate by transesterification. Process Biochem. 38, 393e397. De Oliveira, D., Di Luccio, M., Faccio, C., Rosa, C.D., Bender, J.P., Lipke, N., Menoncin, S., Amroginski, C., De Oliveira, J.V., 2004. Optimization of enzymatic production of biodiesel from castor oil in organic solvent medium. Appl. Biochem. Biotechnol. 113e116, 771e780. De, B.K., Bhattacharyya, D.K., Bandhu, C., 1999. Enzymatic synthesis of fatty alcohol esters by alcoholysis. J. Am. Oil Chem. Soc. 76, 451e453. Demirbas, A., 2008. A Realistic Fuel for Alternative Diesel Engines. Human press, Springer, Verlag, London. p. 208. Dizge, N., Keskinler, B., 2008. Enzymatic production of biodiesel from canola oil using immobilized lipase. Biomass Bioenergy 32, 1274e1278.
128 8. LIPASE-CATALYZED BIODIESEL PRODUCTION: TECHNICAL CHALLENGES Du, W., Xu, Y., Liu, D., 2003. Lipase-catalysed transesterification of soya bean oil for biodiesel production during continuous batch operation. Biotechnol. Appl. Biochem. 38, 103e106. Er-Zheng, S., Zhang, M.J., Zhang, J.G., Gao, J.F., Wei, D.Z., 2007. Lipase-catalyzed irreversible transesterification of vegetable oils for fatty acid methyl esters production with dimethyl carbonate as the acyl acceptor. Biochem. Eng. J. 36, 167e173. Fan, X., 2012. Enzymatic biodiesel productiondthe way of the future. Lipid Technol. 24, 31e32. Fariha, H., Shah, A.A., Hameed, A., 2006. Industrial applications of microbial lipases. Enzyme Microb. Technol. 39, 235e251. Farm energy, 2012. Animal fats for biodiesel production. (Accessed on 20 February, 2013) available on http://www.extension.org/pages/ 30256/animal-fats-for-biodiesel-production. Fjerbaek, L., Christensen, K.V., Norddahl, B., 2009. A review of the current state of biodiesel production using enzymatic transesterification. Biotechnol. Bioeng. 102, 1298e1315. Freedman, B., Pryde, E.H., Mounts, T.L., 1984. Variables affecting the yields of fatty esters from transesterified vegetable oils. J. Am. Oil Chem. Soc. 61, 1638e1643. Fukuda, H., Kondo, A., Noda, H., 2001. Biodiesel fuel production by transesterification of oils. J. Biosci. Bioeng. 92, 405e416. Ganesan, D., Rajendran, A., Thangavelu, V., 2009. An overview on the recent advances in the transesterification of vegetable oil for biodiesel production using chemical and biocatalyst. Rev. Environ. Sci. Biotechnol. 8, 367e394. Gao, Y., Tan, W.T., Nie, K.L., Wang, F., 2006. Immobilization of lipase on macroporous resin and its application in synthesis of biodiesel in low aqueous media. Chin. J. Biotechnol. 22, 114e118. Ghaly, A.E., Dave, D., Brooks, M.S., Budge, S., 2010. Production of biodiesel by enzymatic transesterification: review. Am. J. Biochem. Biotechnol. 6, 54e76. Ghamgui, H., Karra-Chaa bouni, M., Gargouri, Y., 2004. 1-Butyl oleate synthesis by immobilized lipase from Rhizopus oryzae: a comparative study between n-hexane and solvent free system. Enzyme Microb. Technol. 35, 335e363. Haas, M.J., Piazza, G.J., Foglia, T.A., 2002. Enzymatic approaches to the production of biodiesel fuels. In: Kuo, T.M., Gardner, H.W. (Eds.), Lipid Biotechnology. Marcel Dekker Inc., New York, pp. 587e598. Hama, S., Tamalampudi, S., Yoshida, A., Tamadani, N., Kuratani, N., Nodaa, H., Fukuda, H., Kondo, A., 2011. Enzymatic packed-bed reactor integrated with glycerol-separating system for solventfree production of biodiesel fuel. Biochem. Eng. J. 55, 66e71. Hernandez-Martin, E., Otero, C., 2008. Different enzyme requirements for the synthesis of biodiesel: Novozym 435 and Lipozyme TL IM. Bioresour. Technol. 99, 277e286. Huynh, L.H., Kasim, N.S., Ju, Y.H., 2011. Biodiesel production from waste oils. In: Pandey, A.K., Larroche, C., Ricke, S.C., Dussap, C.G., Gnansounou, E. (Eds.), Biofuels: Alternative Feedstocks and Conversion Processes. Academic press, Elsevier Inc, USA, pp. 375e392. Iso, M., Chen, B., Eguchi, M., Kudo, T., Shrestha, S., 2001. Production of biodiesel fuel from triglycerides and alcohol using immobilized lipase. J. Mol. Catal. B: Enzym. 16, 53e58. Jang, M.G., Kim, D.K., Park, S.C., Lee, J.S., Kim, S.W., 2012. Biodiesel production from crude canola oil by two-step enzymatic processes. Renewable Energy 42, 99e104. Jeong, G.T., Park, D.H., 2008. Lipase-catalyzed transesterification of rapeseed oil for biodiesel production with tert-butanol. Appl. Biochem. Biotechnol. 148, 131e139. Jike, L., Nie, K., Xie, F., Wang, F., Tan, T., 2007. Enzymatic synthesis of fatty acid methyl esters from lard with immobilized Candida sp. 99e125. Process Biochem. 42, 1367e1370. Kaieda, M., Samukawa, T., Kondo, A., Fukuda, H., 2001. Effect of methanol and water contents on production of biodiesel fuel from plant oil catalyzed by various lipases in a solvent-free system. J. Biosci. Bioeng. 91, 12e15. Kanerva, L.T., Vihanto, J., Halme, M.H., Loponen, J.M., Euranto, E.K., 1990. Solvent effects in lipase-catalysed transesterification reactions. Acta Chem. Scand. 44, 1032e1035. Karatay, S.E., Donmez, G., 2011. Microbial oil production from thermophile cyanobacteria for biodiesel production. Appl. Energy 88, 3632e3635. Karout, A., Pierre, A.C., 2009. Partial transesterification of sunflower oil with ethanol by a silica fiber reinforced aerogel encapsulated lipase. J. Sol-Gel Sci. Technol. 52, 276e286. Knothe, G., 2001. Historical perspectives on vegetable oil-based diesel fuels. Inform 12, 1103e1107. Koh, M.Y., Ghazi, T.I.M., 2011. A review of biodiesel production from Jatropha curcas L. Oil. Renewable Sustainable Energy Rev. 15, 2240e2251. Kose, O., Tooter, M., Aksoy, H.A., 2002. Immobilized Candida Antarctica lipase catalyzed alcoholysis of cotton seed oil in a solvent-free medium. Bioresour. Technol. 83, 125e129. Krawczyk, T., 1996. Biodieseldalternative fuel makes inroads but hurdles remain. Inform 7, 801e829. Lai, C.C., Zullaikah, S., Vali, S.R., Ju, Y.H., 2005. Lipase-catalyzed production of biodiesel from rice bran oil. J. Chem. Technol. Biotechnol. 80, 331e337. Lee, K.T., Foglia, T.A., Chang, K.S., 2002. Production of alkyl ester as biodiesel from fractionated lard and restaurant grease. J. Am. Oil Chem. Soc. 79, 191e195. Leung, D.Y.C., Wu, X., Leung, M.K.H., 2010. A review on biodiesel production using catalyzed transesterification. Appl. Energy 87, 1083e1095. Linko, Y.Y., Lamsa, M., Huhtala, A., Linko, P., 1994. Lipase catalysed transesterification of rapeseed oil and 2-ethyl-1-hexanol. J. Am. Oil Chem. Soc. 71, 1411e1414. Liu, Y., Xin, H., Yan, Y.J., 2009. Physicochemical properties of stillingia oil: feasibility for biodiesel production by enzyme transesterification. Ind. Crops Prod. 30, 431e436. Ma, L., Persson, M., Adlercreutz, P., 2002. Water activity dependence of lipase catalysis in organic media explains successful transesterification reactions. Enzyme Microb. Technol. 31, 1024e1029. Ma, F., Hanna, F.A., 1999. Biodiesel production: a review. Bioresour. Technol. 70, 1e15. Madras, G., Kolluru, C.R., 2004. Synthesis of biodiesel in supercritical fluids. Fuel 83, 2029e2033. Mata, T.M., Martins, A.A., Caetano, S.N., 2010. Microalgae for biodiesel production and other applications: a review. Renewable Sustainable Energy Rev. 14, 217e232. Mata, T.M., Sousa, I.R.B.G., Vieira, S.S., Caetano, N.S., 2012. Biodiesel production from corn oil via enzymatic catalysis with ethanol. Energy Fuels 26, 3034e3041. Matassoli, A.L.F., Corrêa, I.N.S., Portilho, M.F., Veloso, C.O., Langone, M.A.P., 2009. Enzymatic synthesis of biodiesel via alcoholysis of palm oil. Appl. Biochem. Biotechnol. 155, 347e355. Meher, L.C., Sagar, D.V., Naik, S.N., 2006. Technical aspects of biodiesel production by transesterificationda review. Renewable Sustainable Energy Rev. 10, 248e268. Mittelbach, M., 1990. Lipase catalyzed alcoholysis of sunflower oil. J. Am. Oil Chem. Soc. 67, 168e170. Modi, M.K., Reddy, J.R.C., Rao, B.V.S.K., Prasad, R.B.N., 2007. Lipasecatalyzed mediated conversion of vegetable oils into biodiesel using ethyl acetate as acyl. Bioresour. Technol. 98, 1260e1264. Mohibbe, A.M., Amtul, W., Nahar, N.M., 2005. Prospects and potential of fatty acid methyl esters of some non-traditional seed oils for use as biodiesel in India. Biomass Bioenergy 29, 293e302. Moreno-Pirajan, J.C., Giraldo, L., 2011. Study of immobilized Candida rugosa lipase for biodiesel fuel production from palm oil by flow micro calorimetry. Arab. J. Chem. 4, 55e62.
REFERENCES Mukesh, D., Banerji, A.A., Bevinakatti, H.S., 1994. A note on transesterifications of vegetable oils catalysed by lipase in a packed tubular reactor. Indian Chem. Eng., Sect. A 36, 193e196. Nelson, L.A., Foglia, T.A., Marmer, W.N., 1996. Lipase-catalysed production of biodiesel. J. Am. Oil Chem. Soc. 73, 1191e1195. Nie, K., Xie, F., Wang, F., Tan, T., 2006. Lipase catalyzed methanolysis to produce biodiesel: optimization of the biodiesel production. J. Mol. Catal. B: Enzym. 43, 142e147. Nielsen, P.M., Brask, J., Fjerbaek, L., 2008. Enzymatic biodiesel production: technical and economical considerations. Eur. J. Lipid Sci. Technol. 110, 692e700. Noureddini, H., Gao, X., Philkana, R.S., 2005. Immobilized Pseudomonas cepacia lipase for biodiesel fuel production from soybean oil. Bioresour. Technol. 96, 769e777. Ognjanovic, N., Bezbradica, D., Knezevic-Jugovic, Z., 2009. Enzymatic conversion of sunflower oil to biodiesel in a solvent-free system: process optimization and the immobilized system stability. Bioresour. Technol. 100, 5146e5154. Oliveira, D., Oliveira, J.V., 2001. Enzymatic alcoholysis of palm kernel oil in n-hexane and SC CO2. J. Supercrit. Fluids 19 (2), 141e148. OPEC: Organization of the Petroleum Exporting Countries, 2011. World Oil Outlook. available on. http://www.opec.org/opec_ web/static_files_project/media/downloads/publications/WOO_ 2011.pdf (accessed on 21.12.2012). OPEC: Organization of the Petroleum Exporting Countries, 2012. World Oil Outlook. available on. http://www.opec.org/opec_ web/static_files_project/media/downloads/publications/ WOO2012.pdf (accessed on 12.07.2012). Ranganathan, S.V., Narasimhan, S.L., Muthukumar, K., 2008. An overview of enzymatic production of biodiesel. Bioresour. Technol 99, 3975e3981. Ribeiro, B.D., De Castro, A.M., Coelho, M.A.Z., Freire, D.M.G., 2011. Production and use of lipases in bioenergy: a review from the feedstocks to biodiesel production. Enzyme Res. 2011, 1e16. Royon, D., Daz, M., Ellenrieder, G., Locatelli, S., 2007. Enzymatic production of biodiesel from cottonseed oil using n-butanol as a solvent. Bioresour. Technol. 98, 648e653. Sanchez, F., Vasudevan, P.T., 2006. Enzyme catalyzed production of biodiesel from olive oil. Appl. Biochem. Biotechnol. 135, 1e14. Schwab, A.W., Bagby, M.O., Freedman, B., 1987. Preparation and properties of diesel fuels from vegetable oils. Fuel 66, 1372e1378. Sevil, Y., Terzio glu, P., Özçimen, D., 2012. Lipase applications in biodiesel production. In: Fang, Z. (Ed.), BiodieseldFeedstocks, Production and Applications. InTech, Rijeka, Croatia, pp. 210e250. Sha, S., Sharma, S., Gupta, M.N., 2003. Mini review: enzymatic transesterification for biodiesel production. Indian J. Biochem. Biophys. 40, 392e399. Shah, S., Gupta, M.N., 2007. Lipase catalyzed preparation of biodiesel from jatropha oil in a solvent free system. Process Biochem. 42, 409e414. Shimada, Y., Watanabe, Y., Sugihara, A., Tominaga, Y., 2002. Enzymatic alcoholysis for biodiesel fuel production and application of the reaction to oil processing. J. Mol. Catal. B: Enzym. 17, 133e142. Singh, S.P., Singh, D., 2010. Biodiesel production through the use of different sources and characterization of oils and their esters as the substitute of biodiesel: a review. Renewable Sustainable Energy Rev. 12, 200e216. Sonntag, N.O.V., 1979. Reactions of fats and fatty acids. In: Swern, D., Bailey, A.E. (Eds.), Industrial Oil and Fat Products, fourth ed. Wiley, New York, USA, p. 99. Soumanou, M.M., Bornscheuer, U.T., 2003a. Improvement in lipasecatalyzed synthesis of fatty acid methyl esters from sunflower oil. Enzyme Microb. Technol. 33, 97e103. Soumanou, M.M., Bornscheuer, U.T., 2003b. Lipase-catalyzed alcoholysis of vegetable oils. Eur. J. Lipid Sci. Technol. 105, 656e660. 129 Su, E.Z., Zhang, M.J., Zhang, J.G., Gao, J.F., Wei, D.Z., 2007. Lipasecatalyzed irreversible transesterification of vegetable oils for fatty acid methyl esters production with dimethyl carbonate as the acyl acceptor. Biochem. Eng. J. 36, 167e173. Sulaiman, A.Z., 2007. Production of biodiesel: possibilities and challenges. Biofuels, Bioprod. Biorefin. 1, 57e66. Taher, H., Al-Zuhair, S., AlMarzouqui, A., Hashim, I., 2011. Extracted fat from lamb meat by supercritical CO2 as feedstock for biodiesel production. Biochem. Eng. J. 55, 23e31. Talukder, M.M.R., Das, P., Fang, T.S., Wu, J.C., 2011. Enhanced enzymatic transesterification of palm oil to biodiesel. Biochemical. Eng. J. 55, 119e122. Tan, T., Lu, J., Nie, K., Deng, L., Wang, F., 2010. Biodiesel production with immobilized lipase: a review. Biotechnol. Adv. 28, 628e634. Tweddell, R.J., Kermasha, S., Combes, D., Marty, A., 1998. Esterification, interesterification activities of lipase from Rhizopus niveus and Mucor miehei in three different types of organic media: a comparative study. Enzyme Microb. Technol. 22, 439e445. Vivian, F., Junior, A.C., De Pra, M.C., de Abreu, P.G., dos Santos Filho, J.I., Higarashi, M.M., Sulenta, M., Coldebella, A., 2011. Animal fat wastes for biodiesel production. In: Stoytcheva, M., Gisela, M. (Eds.), BiodieseldFeedstocks and Processing Technologies. InTech, Rijeka, Croatia, pp. 45e70. Wang, Y.N., Chen, M.H., Ko, C.H., Lu, P.J., Chern, J.M., Wu, C.H., Chang, F.C., 2011a. Lipase Catalyzed Transesterification of Tung and Palm Oil for Biodiesel. World Renewable Energy Congress; Sweden, 8e13 May. Wang, X., Liu, X., Zhao, C., Ding, Y., Xu, P., 2011b. Biodiesel production in packed-bed reactors using lipaseenanoparticle biocomposite. Bioresour. Technol. 102, 6352e6355. Watanabe, Y., Pinsirodom, P., Nagao, T., Asao, Y., Kobayashi, T., Nishida, Y., Takagi, Y., Shimada, Y., 2007. Conversion of acid oil byproduced in vegetable oil refining to biodiesel fuel by immobilized Candida antarctica lipase. J. Mol. Catal. B: Enzym. 44, 99e105. Watanabe, Y., Shimada, Y., Sugihara, A., Tominaga, Y., 2001. Enzymatic conversion of waste edible oil to biodiesel fuel in a fixed-bed bioreactor. J9830. J. Am. Oil Chem. Soc. 78, 703e707. Wenlei, X., Ning, M., 2010. Enzymatic transesterification of soybean oil by using immobilized lipase on magnetic nano-particles. Biomass Bioenergy 34, 890e896. Xu, Y., Nordblad, M., Woodley, J.M., 2012. A two-stage enzymatic ethanol-based biodiesel production in a packed bed reactor. J. Biotechnol. 162, 407e414. Xu, Y., Du, W., Liu, D., Zeng, J., 2003. A novel enzymatic route for biodiesel production from renewable oils in a solvent-free medium. Biotechnol. Lett. 25, 1239e1241. Xu, Y.Y., Du, W., Zeng, J., Liu, D.H., 2004. Conversion of soybean oil to biodiesel fuel using Lipozyme TL 1M in a solvent-free medium. Biocatal. Biotransform. 22, 45e48. Yahya, A.R.M., Anderson, W.A., Moo-Young, M., 1998. Ester synthesis in lipase catalysed reactions. Enzyme Microb. Technol. 23, 438e450. Yoshida, A., Hama, S., Tamadani, N., Fukuda, H., Kondo, A., 2012. Improved performance of a packed-bed reactor for biodiesel production through whole-cell biocatalysis employing a high-lipaseexpression system. Biochem. Eng. J. 63, 76e80. Zaidi, A., Gainer, J.L., Carta, G., Mrani, A., Kadiri, T., Belarbi, Y., Mir, A., 2002. Esterification of fatty acids using nylon-immobilized lipase in n-hexane: kinetic parameters and chain-length effects. J. Biotechnol. 93, 209e216. Zhang, B., Weng, Y., Xu, H., Mao, Z., 2012. Enzyme immobilization for biodiesel production. Appl. Microbiol. Biotechnol. 93, 61e70. Zheng, Y., Quan, J., Ning, X., Zhu, L.M., Jiang, B., He, Z.Y., 2009. Lipase-catalyzed transesterification of soybean oil for biodiesel production in tert-amyl alcohol. World J. Microbiol. Biotechnol. 25, 41e46.
C H A P T E R 9 Bioelectrochemistry of Microbial Fuel Cells and their Potential Applications in Bioenergy Minghua Zhou 1,*, Jie Yang 1, Hongyu Wang 1, Tao Jin 1, Daniel J. Hassett 2, Tingyue Gu 3,* 1 Key Laboratory of Pollution Processes and Environmental Criteria (Ministry of Education), College of Environmental Science and Engineering, Nankai University, Tianjin, China, 2Department of Molecular Genetics, Biochemistry and Microbiology, University of Cincinnati, College of Medicine, Cincinnati, OH, USA, 3 Department of Chemical and Biomolecular Engineering, Ohio University, Athens, OH, USA *Corresponding author email: gu@ohio.edu;zhoumh@nankai.edu.cn O U T L I N E Introduction 132 Bioelectrochemistry of MFC Electrode Reactions in MFC Anode Reaction Cathode Reaction Electron Transfer Methods DET for Anodic Biofilms MET for Anodic Biofilms Electrogens in Biofilms for MFCs Biocathodes Electron Transfer for Biocathodes DET for Biocathodes MET for Biocathodes 132 132 132 133 133 134 134 135 136 137 137 137 Biofilm Electrochemistry for Enhanced MFC Performance: A Molecular Biology Perspective Bacterial Metabolism: How to Power MFCs through Respiratory/Anaerobic Fluxes Mediator-Less Factors Affecting MFC Performance TFP (or “Nanowires”): Geobacter and Shewanella Species as Model Organisms Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00009-7 139 139 139 Cytochromes (Cell-Bound) Brief Synopsis of the S. oneidensis MR-1 Bioelectrochemical Machinery in Reverse: Potential Role in the Biosynthesis of Biofuels in MFCs Mediators for Accelerated Electron Transfer in Biofilms Flavins Phenazines 140 142 142 142 143 MFCs for Wastewater Treatment with Concomitant Electricity Production 143 MFC Reactor Designs 143 Substrates Used in MFCs 145 Simple Biodegradable Organics 145 Wastewater Types 146 Lignocellulosic Biomass 146 Summary and Perspectives 147 References 147 139 131 Copyright Ó 2014 Elsevier B.V. All rights reserved.
132 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY INTRODUCTION Currently, the energy sources utilized in our society are mainly fossil fuels such as oil, natural gas and coal (Makarieva et al., 2008). However, their supplies are limited and nonrenewable (Logan, 2009). When fossil fuels are combusted, their carbon, sulfur and nitrogen contents are converted into carbon oxides, sulfur oxides and nitric oxides, respectively, resulting in greenhouse gas emission and environmental pollution (e.g. acid rain). With dwindling oil reserves, global warming signs and worsening air pollution in many countries, more efforts are devoted to the use of renewable energy such as solar, wind and bioenergy. Bioenergy is a sustainable alternative to fossil fuels as part of an integrated energy solution to alleviate the worldwide energy crisis and environmental pollution problems (Srikanth and Venkata, 2012). Recently, microbial fuel cells (MFCs) have been intensively investigated in many academic labs as a potential technology for bioenergy production from organic carbon sources such as wastewater, sludge and some lignocellulosic biomass (Allen and Bennetto, 1993; Lovley, 2006b; Rhoads et al., 2005). In a typical MFC, the microbes forming the anodic biofilm oxidize the substrates (organic material) by anaerobic respiration (Bond and Lovley, 2003; Logan et al., 2006) and release electrons (e) and protons (Hþ) (Srikanth and Venkata, 2012). The electrons are transferred to the anode and then reach the cathode via an external circuit. Simultaneously, protons in the solution travel through a proton exchange membrane (PEM) and reach the cathode where electrons are used to reduce oxygen. In this fashion, electricity is generated by converting the energy stored in the chemical bonds in the organic matter (or feedstock) provided to each system (Choi et al., 2003a; Gil et al., 2003; Huang et al., 2011b; Moon et al., 2006; Osman et al., 2010). Thus, MFCs produce bioelectricity directly instead of a biofuel in the process of degrading organic matter in the wastewater (Chaudhuri and Lovley, 2003; Oh and Logan, 2005; Park and Zeikus, 2000). Bioelectricity production by an MFC was first reported by Petter (1911). Not much research was done on MFCs until 1980s when mediators were found to improve MFC power density greatly. However, externally supplied mediators such as methyl viologen, neutral red, and thionine are not sustainable. They are expensive and toxic, limiting their uses to academic research (Du et al., 2007). In recent years, some microorganisms such as Shewanella putrefaciens (Kim et al., 2002), Rhodoferax ferrireducens (Chaudhuri and Lovley, 2003), and Geobacteraceae sulfurreducens (Bond and Lovley, 2003) have been found to transfer electrons from the cytoplasm where metabolic respiration occurs to an external electrode surface (anode), resulting in the development of mediator-less MFCs. Intensive research efforts from 1990 to 2010 have improved MFC power densities by several orders of magnitude to up to several watts per square meter (anode area) under optimal laboratory conditions. Recently, Tong et al. (2012) compared the power densities between MFCs and conventional fuel cells and found that MFCs were still behind by three orders of magnitude. It is unrealistic to expect MFCs to catch up with chemical fuel cells because the latter uses pure hydrogen, ethanol or other high-energy-density fuels rather than wastes. However, it is still necessary to improve MFC power densities much further to make the MFC power generation practically useful. Major breakthroughs are needed in biofilm engineering, materials for electrodes and reactor configuration to achieve far better bioelectrochemical performance and to lower the currently rather high costs in MFC construction, maintenance and operation (Zhou et al., 2012). This chapter addresses various bioelectrochemical issues in MFC operation for the improvement of MFC performance. BIOELECTROCHEMISTRY OF MFC Electrode Reactions in MFC A typical MFC reactor contains an anodic chamber, a cathodic chamber and a PEM partitioning the two chambers. Figure 9.1 shows a dual-chamber MFC. Anode Reaction The microbes in the anodic chamber oxidize substrates such as glucose, acetate and some refractory organics. For example, glucose is oxidized as follows to generate electrons, protons and carbon dioxide (Pham et al., 2006): C6 H12 O6 þ 6H2 O/6CO2 þ 24Hþ þ 24e (9.1) Resistance e e– – PEM Anode Cathode H 2O CO2 e– e– Substrate H+ O2 H+ Air FIGURE 9.1 Schematic diagram of a microbial fuel cell. (For color version of this figure, the reader is referred to the online version of this book.)
BIOELECTROCHEMISTRY OF MFC Because electrons cannot “swim” in an aqueous solution, the oxidation reaction must occur in a biofilm that is capable of transferring electrons to the anode. In the absence of a suitable oxidant in the anodic chamber to absorb the electrons, electrons will be transferred to the anode by the biofilm (Zhao et al., 2009). The electrons reach the cathode via an external circuit linking the anode and the cathode, where they are used to reduce an oxidant such as oxygen (Figure 9.1). A load is placed on the external circuit to harvest the electricity. To maintain electroneutrality, protons must carry an equal amount of positive charges from the anodic chamber to the cathodic chamber usually through a PEM. Inefficient proton migration will result in accumulation of protons that causes acidity in the anodic chamber (Xu et al., 2012). In the anodic chamber, anaerobic conditions are very important to guarantee the substrate oxidation by the microbes through anaerobic respiration (Liu et al., 2005b; Logan et al., 2006). Oxygen leaked into the anodic chamber from outside air or through diffusion from the cathodic chamber (Figure 9.1) would reduce Coulombic efficiency of the MFC by directly oxidizing the organic matter in the anodic chamber. In this case, energy will be released as low-grade heat instead of electricity. A PEM plays an important role of preventing oxygen diffusion from the cathodic chamber to the anodic chamber (Li et al., 2011), while allowing positive charges to go through it via a proton exchange process. If nonoxygen oxidants such as sulfate and nitrate are present in sufficient quantities in the anodic chamber feed stream, the biofilm on the anode must not be able to catalyze their reduction because it would divert the electrons released from oxidation of organic matters for the local reduction of sulfate or nitrate. A buffer solution that usually contains NH4Cl, NaH2PO4, Na2HPO4, KCl, and so on is often used to enhance the proton transfer in laboratory MFC investigations (Liu et al., 2011). The presence of a buffer solution increases the conductivity, thus reducing internal resistance of the MFC (Liu et al., 2005a). Cathode Reaction The cathode reaction has a major impact on MFC performance. The electrons coming from the anode via the external circuit, the protons coming from the anodic chamber via the PEM and the electron acceptors (e.g. O2) will react with the help of catalysts on the cathode (Pham et al., 2006): 24Hþ þ 24e þ 6O2 / 12H2 O (9.2) Reactions (9.1) and (9.2) form a thermodynamically favorable redox reaction, that is, the aerobic oxidation of glucose. However, a thermodynamically favorable reaction may not proceed at an appreciable rate if the kinetics is too slow. In an MFC, anode and cathode reactions almost always require catalysis. For the anode 133 reaction, a biofilm is required to catalyze organic carbon oxidation and electron transfer. For the cathode, oxygen reduction rate is very slow without catalysis. The cathodic reaction efficiency depends on the concentration and type of electron acceptors, proton concentration, electrode structure and its catalytic ability (Zhou et al., 2012). In order to improve electricity generation, a good catalytic cathode is crucial since the catalysts can reduce the activation energy and thus greatly increase the reaction rate. Currently, for oxygen reduction, platinum (Pt) appears to be most effective. However, it is extremely expensive and, thus, unrealistic for most practical applications even when only Pt coating is used. Some alternative catalysts have been explored such as MnOx, CoTMPP, PbO2, iron(II) phthalocyanine (FePc) and recently the biocathode (Roche and Scott, 2008; Zhou et al., 2011). Oxygen is the most popular acceptor because of its high standard potential (0.818 mV), low cost and environmental “friendliness”. However, the rate of oxygen reduction is very low on the cathode surface, resulting in a high overpotential, which is one of the most important limiting factors in MFCs (Gil et al., 2003). Potassium ferricyanide (K3[Fe(CN)6]) can overcome this handicap (Logan et al., 2006; Nevin et al., 2008; Park and Zeikus, 2003). However, the regeneration of K3[Fe(CN)6]is a problem because it usually is not sufficiently oxidized by oxygen. It needs to be replenished periodically (Franks and Nevin, 2010). In addition, K3[Fe(CN)6] can diffuse into the anodic chamber through the PEM, thereby influencing the desired anaerobic conditions of anodic chamber (Logan et al., 2006). Potassium permanganate is also used as an acceptor, and the power density was reported to be higher than that with K3[Fe(CN)6] and oxygen (You et al., 2006). In practice, wastewater streams are low-grade energy sources that are pale in comparison to pure fuels such as hydrogen or ethanol as a fuel. This inherently means that a large volume of water must be treated to harvest a sufficient amount of electricity. This makes all externally added soluble catalysts impractical, limiting them to academic investigations. To overcome the requirement for catalysis by oxygen oxidation on the cathode, biocathodes have been explored (Biocathodes Section). Various biofilms have been tested on cathodes to biocatalyze oxygen or a nonoxygen oxidant such as nitrate and perchlorate (Shea et al., 2008; Srikanth et al., 2012; Zhang and Angelidaki, 2012). Electron Transfer Methods How the electrons released by organic carbon oxidation in the bacterial anaerobic cytoplasm are transferred by the biofilm to the anode surface is an important factor in MFC performance (Neto et al., 2010). Major advances
134 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY have been made between 2000 and 2010 in understanding the electron transfer mechanisms by electrogens. There are two primary mechanisms: one is the direct electron transfer (DET) and the other, mediated electron transfer (MET). DET for Anodic Biofilms DET occurs via a direct physical contact between the microbial cell wall and the anode surface, or via a pilus that links the two. Gene expression studies (Holmes et al., 2008) and electrochemical analysis (Busalmen et al., 2008) have demonstrated that there are active sites for cytochrome proteins on the outer cell surface (Franks and Nevin, 2010; Zhou et al., 2012). When the microbes contact the anode surface, the cytochromes can transfer the electrons from the inside of the microbial cell wall to the outer cell wall and then to the anode surface (Rinaldi et al., 2008) (Figure 9.2(a)). Shewanella putrefaciens (Kim et al., 2002), R. ferrireducens (Chaudhuri and Lovley, 2003) and G. sulfurreducens (Bond and Lovley, 2003) use such cytochromes to achieve electron transfer. One major disadvantage for this mechanism is that only a monolayer of sessile cells in a biofilm can transfer electrons to the anode surface. This explains why the power and current densities of MFCs relying on this kind of DET are lower, sometimes by several orders of magnitude, than that of MFCs with MET (Schröder, 2007) because MET can utilize more than one monolayer. Recently, some researchers have observed that some microbial strains (such as Shewanella oneidensis and G. sulfurreducens; Logan and Regan, 2006; Torres et al., 2010) can produce pili (conductive nanowires) to form physical conductive connections between the cell wall and the anode surface while the microbial cell wall is at a short distance from the anode (Reguera et al., 2005; Rinaldi et al., 2008). An extensive pilus network would allow several layers of sessile cells to donate electrons to the anode, thus multiplying the MFC power output. Summers et al. found a whole cell aggregate (a) Substrate H+ A n o d e (b) H+ Substrate e– CO2 e– H+ Cytochrome CO2 Conductive pili or filament FIGURE 9.2 Different DET methods: (a) direct cell wall-electrode contact (b) conductive pili or filament linkage. (For color version of this figure, the reader is referred to the online version of this book.) consisting of G. sulfurreducens and Geobacter metallireducens is conductive when the coculture was grown on ethanol. The c-type cytochrome OmcS of G. sulfurreducens was suspected to play a key role in accepting electrons from G. metallireducens (Summers et al., 2010). This overcomes the inability of G. sulfurreducens to use H2 for interspecies electron transfer. The ability of interspecies electron exchange suggests that a nonelectrogenic species in a synergistic biofilm consortium may contribute to electricity generation as long as its electrons can be taken up by an electrogenic species through interspecies electron transfer. In addition to H2, other molecules such as formate may act as an electron shuttle for biofilm communities (Morita et al., 2011). An exciting new discovery by Pfeffer et al. (2012) provides new hope for greatly enhancing electron transfer in microorganisms. They found that some filamentous bacteria in marine sediments are capable of transferring electrons over centimeter-long distances via conductive filaments that are 200 nM or wider in diameter. This distance is far greater than the much thinner pili could achieve. This indicates that potentially many more layers of sessile cells could be networked via the conductive filaments than pili could achieve. Figure 9.2 is a schematic illustration of DET via direct cell walleelectrode contact and via a pili or conductive filament linkage. MET for Anodic Biofilms MET THROUGH EXOGENOUS REDOX MEDIATORS Some microbes such as Escherichia coli, Pseudomonas sp., Proteus and Bacillus (Lovley, 2006a) cannot directly transfer electrons to the anode and must rely on mediators (Lovley, 2006a). When the oxidized mediators reach the surface of the microbes, they penetrate the cell membrane of the microbes, and they are reduced by electrons. The reduced mediators pass through the cell membrane again and reach the anode surface where then they are reoxidized (losing the electrons). In this fashion, electrons are transferred to the anode while the oxidized mediators enter the microbes again, thereby continuing the redox cycle (Figure 9.3(a)) (Neto et al., 2010; Rabaey et al., 2005b). Properties of good exogenous mediators should be the ability to (a) cross cell membranes with ease; (b) receive electrons from electron donors without interfering with other metabolic processes; (c) deliver electrons inside the cytoplasm for oxidation reactions and regenerate at rapid rates; (d) have good solubility and stability in both oxidized and reduced forms; (e) have no cytotoxicity; and (f) not be consumed by microbes in the biofilm as a nutrient (Bao and Wu, 2004). These mediators include thionine, neutral red, 2-hydroxy-1,4-naphthoquinone, phenazines, quinines,
BIOELECTROCHEMISTRY OF MFC (a) CO2 Med red (b) A n o d e e– Substrate H2 CO2 Substrate 135 anodic chamber, thus improving the MFC performance (Osman et al., 2010). However, in continuous flow MFCs for wastewater treatment, the secondary metabolites can be insufficient due to diluted concentrations as a result of flow (Lee et al., 2003; Rabaey et al., 2005c), thus resulting in the decline of the performance after the flow starts (Lovley, 2006a; Osman et al., 2010). MET THROUGH PRIMARY METABOLITES H+ Med ox e– H+ Electrocatalyst FIGURE 9.3 The mechanism of MET: (a) exogenous or secondary metabolites and (b) primary metabolites. (For color version of this figure, the reader is referred to the online version of this book.) Fe(III) ethylenediaminetetraacetic acid, methylene blue, phenothiazines, phenoxazines and others (Choi et al., 2003b; Lovley, 2006a; McKinlay and Zeikus, 2004; Newman and Kolter, 2000; Osman et al., 2010; Park and Zeikus, 2000). However, these mediators are unsuitable for practical applications because they are costly and most of them are toxic and recalcitrant, harmful to the environment (Erable et al., 2010a; Lovley, 2006a). MET THROUGH THE SECONDARY METABOLITES Researchers have found that some microbes can transfer electrons without DET in the absence of exogenous redox mediators. These microbes such as S. putrefaciens, S. oneidensis, G. sulfurreducens, Pseudomonas aeruginosa, and Clostridium butyricum can produce their own mediators (Angenent et al., 2004; Erable et al., 2010a; Fitzgerald et al., 2012; Newman and Kolter, 2000; Rabaey et al., 2005a). The presence of these microbes in the mixed cultures enhances electron transfer. These mediators mainly include phenazine derivatives like pyocyanine and 2-amino-3-carboxy-1,4-naphthoquinone (Osman et al., 2010). In practical applications, the secondary metabolites (endogenous redox mediators) may be very important to MFCs because they can transfer the electron without the exogenous redox mediators (Schröder, 2007). The mechanism of electron transfer by the secondary metabolites is similar to that of the exogenous electrochemical redox mediators (Figure 9.3(a)). The secondary metabolites can be reused, and one metabolite molecule can transfer thousands of electrons (Schröder, 2007). So a small amount of the secondary metabolites can singlehandedly enhance the rate of electron transfer and thus increasing power density and improve the MFC performance without introducing costly exogenous mediators. In batch-mode operations, these microbes are very suitable because the mediators will accumulate in the The other endogenous redox mediators are primary metabolites. Some microbes can produce fermentation products such as hydrogen (H2), hydrogen sulfide (H2S), alcohols and ammonia (Erable et al., 2010a). When these primary metabolites reach the surface of the anode, they are oxidized, and the released electrons will be further transferred to the anode surface. There are two types of anaerobic metabolism that can produce primary redox metabolites: one is anaerobic respiration, and the other is fermentation. Some microbes such as Proteus vulgaris, E. coli, P. aeruginosa and Desulfovibrio desulfuricans can produce sulfide which may serve as the mediator to transfer electrons (Bullen et al., 2006; Schröder, 2007):  þ  Cytoplasm : SO2 4 þ 9H þ 8e /HS þ 4H2 O (9.3) þ  Anode : HS þ 4H2 O/SO2 4 þ 9H þ 8e (9.4) This process relies on sulfate reducing bacteria (SRB) that cannot metabolize carbohydrates. A fermentation process can produce small organic acids and alcohols that can be used in anaerobic respiration (Schröder, 2007). Many SRB degrade the substrates incompletely and this lowers the MFC power output. Electrode poisoning by sulfide due to its easy absorption on the electrode surface is also a major drawback (Reimers et al., 2006; Ryckelynck et al., 2005). Fermentation also produces primary metabolites such as hydrogen, ethanol and formate. They can be oxidized directly by electrolysis on an anode such as platinum or tungsten carbide (Rosenbaum et al., 2006). For example, through electrocatalysis, the molecular hydrogen near and on an anode surface would be oxidized to Hþ, accompanying the electron transfer (Figure 9.3(b)). Molecular hydrogen is known to be used as an electron carrier used by hydrogenasepositive microbes such as some SRB in microbiologically influenced corrosion (Gu, 2012). Thus, it contributes to power generation. Electrogens in Biofilms for MFCs The microbial species in a biofilm covering an anode are important because they determine the mode of electron transfer and the mechanism of electricity generation as well as what forms of organic material can be
136 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY utilized in the feed stream. Theoretically, a myriad of microorganisms may be useful for MFCs, but most of them have no direct electrochemical activity and thus cannot transfer electrons directly from the cytoplasm to the anode, i.e. they are not electrogenic. However, many microorganisms with the addition of a soluble redox mediator can act as electron transfer intermediates to transfer electrons. Table 9.1 shows the microbial species and the electron transfer mechanism in the anodic chamber that can perform such processes. In MFCs, mixed cultures usually possess higher electron transfer efficiency than the pure culture because its specificity to the microbe is very strong and its growth rate is relatively slow (Hassan et al., 2012). Mixed cultures are often found to perform better than pure stains. This is because a synergistic biofilm consortium contains various syntrophic species, with each organism contributing specific roles. A consortium can adapt to substrate variations in wastewater and harsh environmental conditions because generally a biofilm consortium is far more robust metabolically than a pure-culture biofilm. The consortium is able to self-select the most efficient electron transfer mechanism if several are available. Biocathodes Biocathodes use biofilms as catalysts to improve the cathode reaction, avoiding using precious metal catalysts. Another unique advantage of biocathodes is that oxidants other than oxygen can be used, including sulfate, nitrate, carbon dioxide, Hþ, Fe(III), Cr(VI), U(VI), Mn(IV), tetrachloroethene, fumarate, perchlorate, and trichloroethene (Huang et al., 2011c). In addition, the sustainability of MFC may be improved with the elimination of problems such as sulfur poisoning of Pt and the requirement for electron mediators in the cathodic chamber (He and Angenent, 2006). There are two types of biocathodes: aerobic and anaerobic. Aerobic biocathodes reduce oxygen (electron acceptor). The biofilm on the cathode surface can catalyze the oxidation of transition metal compounds, such as Fe(II) and Mn (II), releasing the electrons to oxygen. MFCs with aerobic biocathodes can produce higher power density than that of anaerobic biocathodes (Srikanth and Venkata, 2012). The use of a biocathode also means that an MFC can potentially be used to treat an additional wastewater stream in the cathodic chamber. It may be a wastewater stream containing sulfate or nitrate that can come from agricultural runoff (Srikanth and Venkata, 2012). However, the accumulation of microbial metabolites in the cathode chamber can inhibit microbial activities. In addition, metabolites which act as electron donors for bacteria can also compete against the cathode, and therefore reduce the MFC performance (Hamid et al., 2008). TABLE 9.1 The Microbial Species in the Anodic Chamber Microbe Electron Transfer References Escherichia coli K12 MET Erable et al. (2010a) Clostridium beijerinckii MET Erable et al. (2010a) Clostridium butyricum MET Erable et al. (2010a) Proteus vulgaris MET Kim et al. (2000a,b); Thurston et al. (1985) Shewanella putrefaciens MET/DET Kim et al. (1999, 2002) Geothrix fermentans MET Bond and Lovley (2005) Pseudomonas aeruginosa MET Rabaey et al. (2005a) Shewanella oneidensis MET/DET Biffinger et al. (2008, 2007); Hou et al. (2009); Manohar et al. (2008); Qian et al. (2009); Ringeisen et al. (2006) Desulfuromonas acetoxidans DET Bond et al. (2002) Geobacter sulfurreducens DET Holmes et al. (2004) Geobacter metallireducens DET Holmes et al. (2004); Min et al. (2005) Rhodoferax ferrireducens DET Holmes et al. (2004) Desulfobulbus propionicus DET Holmes et al. (2004) Aeromonas hydrophila MET Pham et al. (2003) Clostridium butyricum DET Niessen et al. (2004) Hansenula anomala DET Prasad et al. (2007) Rhodopseudomonas palustris MET Xing et al. (2008); Zhou et al. (2012) Enterococcus faecium MET Rabaey et al. (2005a) Desulfovibrio desulfuricans DET Cooney et al. (1996) Erwinia dissolvens MET Vega and Fernandez (1987) Escherichia coli MET/DET McKinlay and Zeikus (2004); Schröder et al. (2003) Desulfovibrio vulgaris MET Tsujimura et al. (2001) Shewanella putrefaciens IR-1 DET Schröder (2007) Shewanella putrefaciens MR-1 DET Schröder (2007) Shewanella putrefaciens SR-1 DET Schröder (2007) Aeromonas hydrophila PA 3 DET Schröder (2007) Clostridium sp. EG 3 DET Schröder (2007)
BIOELECTROCHEMISTRY OF MFC 137 Furthermore, Zhou et al. (2013) indicated that the voltage output for the combined redox reaction involving Eqn (9.3) and the oxidation of an organic carbon such as acetate may be too small for MFC after subtracting various overpotentials. systems, the mixed culture biocathodes also can transfer electrons via DET. When nitrate, carbon dioxide or trichloroethene is used as the electron acceptor, the DET in the mixed culture biocathode improves the power generation (Aulenta et al., 2010; Cao et al., 2009; Clauwaert et al., 2007a). Electron Transfer for Biocathodes MET for Biocathodes There are numerous investigations on the mechanisms of electron transfer to the anode by the microbes, while the reports about electron transfer to a biocathode are rather limited (Lovley, 2008). The two electron transfer directions are opposite. The biocathode is an electron donor while the anode is an electron acceptor. Despite this difference, biocathodes use the same electron transfer mechanisms, DET and MET (Rosenbaum et al., 2011), because biofilm electron transfer can be bidirectional. MET THROUGH EXOGENOUS REDOX MEDIATORS DET for Biocathodes Similar to DET for anodes, DET for biocathodes also requires physical contact of the microbial cell wall with the electrode surface. At the site of direct contact, the electrons directly transfer to the outer cell membrane-bound redox macromolecules (such as c-type cytochromes) from the electrode (Figure 9.4(a); Huang et al., 2011c). However, this kind of DET can only utilize a monolayer of sessile cells on the cathode, thus limiting the biocathode performance. With an increase in biofilm thickness, the power generation decreased due to mass transfer resistance to oxidant diffusion from the bulk fluid to the cathode surface (Behera et al., 2010). Geobacter species and mixed cultures that use nitrate, fumarate, tetrachloroethene, O2, CO2, U(VI)/U(IV), and so on as an electron acceptor generally transfer the electrons via DET (Table 9.2). On biocathodes, most of the microbes are found to be Gram negative although some Gram-positive microbes exhibit the DET mechanism in cyclic voltammetry (Huang et al., 2011c). Compared with the pure culture (a) Oxidized acceptor C a t h o d e (b) Oxidized acceptor Med red e– Reduced acceptor Med ox Reduced acceptor FIGURE 9.4 The mechanism of electron transfer in biocathode: (a) DET and (b) MET. (For color version of this figure, the reader is referred to the online version of this book.) Similar to bioanodes, the same exogenous mediators including neutral red, methyl viologen and the anthraquinone-2,6-disulfonate can be used for biocathodes (Hatch and Finneran, 2008; Park and Zeikus, 1999; Steinbusch et al., 2010) to enhance MFC performance significantly. When mediators are added into the cathode chamber, they are reduced by the electrons donated by the cathode. The reduced mediators reach the microbial cell wall and then transfer the electrons through the wall while the mediators are oxidized. Subsequently, the oxidized mediators diffuse back to the cathodic surface for reuse. This cyclic process is illustrated in Figure 9.4(b). Usually one mediator molecule can accomplish thousands of cycles. These mediators are relatively short-lived and costly, making their use unsustainable. Just like their use for bioanodes, these exogenous mediators are used only in laboratory investigations of MFC mechanisms for academic purposes. Pili can also be used by microbes to transfer extracellular electrons to the cytoplasm (Zhou et al., 2013). In manganese-oxidizing bacteria, manganese (IV) plays an important role in the electron transfer. This mechanism is similar to the exogenous mediator MnO2 on the biocathode surface. It is first reduced to MnOOH by the electrons donated from the cathode and then Mn2þ is released. Finally, with the help of manganeseoxidizing bacteria, Mn2þ was oxidized by dissolved oxygen to regenerate MnO2 (Nguyen et al., 2007). The power density can be improved by two orders of magnitude, compared with the abiotic cathode (Rhoads et al., 2005), making it attractive for potential practical applications. MET THROUGH SELF-EXCRETED REDOX MEDIATORS Apart from exogenous redox mediators, some microbes can excrete metabolites that are redox active. For example, Pseudomonas spp. can produce phenazines (Venkataraman et al., 2010) and S. oneidensis can produce flavins (Marsili et al., 2008). These mediators can be used by biocathodes indirectly. In the presence of these mediators, the rate of electron transfer is enhanced. These mediators are more easily utilized by other microbes than their producers (Rosenbaum et al., 2011). Therefore, in biocathodes, the self-excreted mediators play an important role in a synergistic biofilm consortium covering a cathode. Their mechanism of electron transfer is
138 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY TABLE 9.2 The Microbial Species in the Biocathode Chamber Species Electron Transfer Oxidant/End Product References Burkholderia cepacia DET O2/H2O Cournet et al. (2010) Brevundimonas diminuta DET O2/H2O Branhamella catarrhalis DET O2/H2O Bacillus subtilis DET O2/H2O Acinetobacter sp. DET O2/H2O Shigella flexneri DET O2/H2O Escherichia coli DET O2/H2O Enterobacter cloacae DET O2/H2O Pseudomonas aeruginosa DET O2/H2O Pseudomonas fluorescens DET O2/H2O Kingella denitrificans DET O2/H2O Staphylococcus carnosus DET O2/H2O Kingella kingae DET O2/H2O Micrococcus luteus DET O2/H2O Geobacter sulfurreducens DET Fumarate/succinate Dumas et al. (2008) Mixed culture DET O2/H2O Aldrovandi et al. (2009) Leptothrix discophora SP-6 MET O2/H2O Nguyen et al. (2007) Hydrogenophilic methanogenic culture DET CO2/CH4 Hþ/H2 Villano et al. (2010) Mixed culture DET CO2/CH2O* Cao et al. (2009) Geobacter sulfurreducens DET U(VI)/U(IV) Gregory and Lovley (2005) Shewanella putrefaciens MET O2/H2O Freguia et al. (2010) Acinetobacter calcoaceticus MET O2/H2O Mixed culture DET O2/H2O Rabaey et al. (2008) Mixed culture DET O2/H2O Erable et al. (2010b) Mixed culture MET O2/H2O Clauwaert et al. (2007b) DET  ClO 4 /Cl Shea et al. (2008) MET  ClO 4 /Cl Thrash et al. (2007) Azospira suillum MET  ClO 4 /Cl Mixed culture DET Cr(VI)/Cr(III) Huang et al. (2011a); Tandukar et al. (2009) Mixed culture DET TCE/cis-DCE Aulenta et al. (2010) Mixed culture Dechloromonas agitata Desulfovibrio vulgaris MET þ H /H2 þ Lojou et al. (2002) DET H /H2 Jeremiasse et al. (2010) DET NO 3 /N2 Clauwaert et al. (2007a) Mixed culture DET NO 3 /N2 Lefebvre et al. (2008) Mixed culture MET Acetate/ethanol Steinbusch et al. (2010) Mixed culture Mixed culture * CH2O represents the approximate formula of biomass.
BIOFILM ELECTROCHEMISTRY FOR ENHANCED MFC PERFORMANCE: A MOLECULAR BIOLOGY PERSPECTIVE similar to that used by exogenous redox mediators. Table 9.2 shows some reported microbial species for biocathodes. BIOFILM ELECTROCHEMISTRY FOR ENHANCED MFC PERFORMANCE: A MOLECULAR BIOLOGY PERSPECTIVE Mechanisms by which bacteria generate power in MFCs have been intensively investigated in recent years. How might we make “super-bug” electrogens using the power of genetic manipulation? There are many factors that we may ponder. The MFC literature is rife with biological, biochemical and biophysical aspects of highly electrogenic bacteria including members of the genera Geobacter, Shewanella, and Rhodoferax species, and many other Gram-positive and -negative electrogenic bacteria (Huang et al., 2012; Guo et al., 2012). Many of these organisms can exist and/or thrive in a myriad of different niches, including those involving significant variations in temperature, pH, osmolarity, pollutants, biocides, and metabolizable/nonmetabolizable carbon sources. As such, many studies involving MFCs house single species bacteria, bacterial “consortia,” media or feedstock, anode/cathode materials, MFC design and design material, flow rates, Coulombic efficiency and other MFC parameters that can impact power density measurements. Bacterial Metabolism: How to Power MFCs through Respiratory/Anaerobic Fluxes Many bacteria are metabolically versatile organisms, and can utilize nearly every carbon-containing compound produced in nature. As stated earlier, they are capable of aerobic and/or anaerobic respiration as well as fermentation. Aerobic respiration requires molecular oxygen (O2) while the latter can use alternative electron 2 acceptors including but are not limited to NO 3 , SO4 , 3þ Fe , dimethyl sulfoxide (DMSO), trimethylamine N-oxide (TMAO), and CO2. A member of the tricarboxylic acid cycle (TCA) cycle, fumarate, can also be used. Interestingly, forms of electron acceptors including oxides of iron and manganese (Shi et al., 2007), vanadium, selenium, tellurium and toxic metals including chromium, arsenic and cobalt can also be used by some organisms. Thus, depending on the organism(s), the goal of genetically unmodified bacteria is to couple oxidation of organic matter and reduction of terminal electron acceptor (in most cases, the anode of the MFC). However, in MFCs, the electron acceptor in most cases, with the exception of biocathodes, and bioanode/biocathode MFCs, is the anodic surface. This can occur in what are commonly termed mediator-dependent or mediator-less MFCs (discussed below). Many facultative 139 and obligatory anaerobic bacteria can undergo an anaerobic process known as fermentation that does not require functional cytochromes, respiratory chains and produces far less energy in the form of adenosine triphosphate (ATP) than, for example, glucose respiration in E. coli. Mediator-Less Factors Affecting MFC Performance Many studies have been conducted in the past w7 years to either isolate superior unknown electrogens or improve the electrogenic properties of existing organisms possessing such capacity. The power of various molecular genetics tools (mutations, deletions, gene transfer, overexpression, etc.) is the central force underlying the discovery of such strains. However, surfacelocalized factors such as bacterial type IV pili (TFP) represent a major mediator-less protein that contributes significantly to some of the more extensively studied electrogenic bacteria. TFP (or “Nanowires”): Geobacter and Shewanella Species as Model Organisms TFP have been found to be critical for the transfer of electrons generated metabolically to metal oxides (e.g. iron oxides; Reguera et al., 2005) that represent just one component of an MFC anode. These are the extendable (fully extended outside the cell, followed by retraction, degradation, and new pilus synthesis) proteinaceous appendages (for fundamental structures of three TFP of P. aeruginosa, N. gonorrhoeae and V. cholerae, see Figure 9.5(aec) from Craig et al., 2004). Pili are often essential for optimal biofilm formation in many bacterial genera (Zechner et al., 2012), a requirement for mediator-less current on the MFC electrode(s) surface. Geobacter members are capable of reducing oxide of either insoluble iron (Fe3þ) or manganese (Mn4þ) that are directly coupled with organic carbon oxidation. The pilus extends from many bacteria to bind to and retract from surfaces for biofilm formation and dispersion in some bacteria (O’Toole and Kolter, 1998) and is capable of a “grappling hook” retraction mechanism, followed by degradation, new pilus synthesis and extension, followed again by the complete retractione degradationesynthesiseextension loop. The pili of electrogenic G. sulfurreducens have been termed “nanowires” due to their highly conductive properties (Reguera et al., 2005) that appear to differ markedly from similar members of the same genus (e.g. G. metallireducens). These “nanowires” have also been described in S. oneidensis (Gorby et al., 2006) and are likely in many other bacteria. The electrogenic importance of the pilus was proven in a deletion mutant strain of G. sulfurreducens pilA that generated nearly 10-fold less the power density than that of wild-type, pilusþ bacteria (Reguera et al., 2006).
140 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY FIGURE 9.5 Superimposition of the ribbon structures of (a) P. aeruginosa PAK pilin, (b) N. gonorrhoeae pilin and (c) V. cholerae TcpA. Details of the superimposition parameters are based upon the color of each pilin where PAK pilin is in blue, gonococcal pilin is in white, and TcpA is in red. Source: Figure from Craig et al. (2004) with permission from Nature Publishing Group. (For interpretation of the references to color in this figure legend, the reader is referred to the online version of this book.) These results were independent of anode composition, whether it is inexpensive, highly reproducible conductive graphite, or gold, an expensive yet sometimes problematic (reproducibility issues) anodic material (Richter et al., 2008). A longer isoform of PilA is critical for optimal power generation than a shorter PilA (Richter et al., 2012). Many genes, including those involved in flagella and pilus biosynthesis in G. sulfurreducens (and even bacterial human opportunistic pathogens, Proteus mirabilis and P. aeruginosa (Totten et al., 1990; Zhao et al., 1999)) are controlled by the nitrogen sigma factor, RpoN (Leang et al., 2009), as identified by microarray analyses. Thus, predictably strains lacking RpoN are not electrogenic when compared to wild-type bacteria. Yi et al. (2009) demonstrated indirect evidence that T4P in KN400 strain of G. sulfurreducens not only formed more robust biofilms but also provided superior power generation (KN400 current (7.6 A/m2) and power (3.9 W/m2); wild-type DL1-(1.4 A/m2 and 0.5 W/m2)). An excellent review by Lovley et al. (2011) has shed light on the unique processes involved in G. sulfurreducens metabolism and how such unique metabolic properties lends to its reputation as a highly electrogenic organism. Gorby et al. (2006) have shown that S. oneidensis MR-1 produce conductive pili in response to a reduction in or a lack of a terminal electron acceptor. Those researchers linked electron carrier proteins (c-type decaheme cytochromes MtrC and OmcA, see below) as well as mutations in the type II secretion pathway, where there are often periplasm protein modifications (e.g. disulfide bond formation) within Gram-negative bacteria. Thus, despite possessing pili, bacteria lacking specific cytochromes possessed reduced electrogenic properties. Yi et al. (2009) isolated a mutant of G. sulfurreducens DL1 (KN400 strain) that was more effective in current production than wild-type bacteria. The paradoxical results were manifested with KN400 forming thinner biofilms, increased current production, great nanowire production, flagellum production, far less outersurface c-type cytochromes and, above all, lower MFC internal resistance. Recently, however, an artificial matrix termed a conductive artificial biofilm (CAB) was developed that allows for adherence and nearly 11-fold increased conductive properties of Shewanella biofilm bacteria (Yu et al., 2011). Cytochromes (Cell-Bound) Redox properties of some bacterial cytochromes (either membrane-bound or soluble cytochromes (e.g. cytochrome c) electron carriers) have been connected with the conductive properties of pili (described above). Typically, these are critical for normal respiratory functions in both prokaryotic and eukaryotic cells. In recent years, electrogenesis by metal-oxidizing Shewanella and Geobacter species as described above are facilitated by the production of pili and flagella, yet cytochromes have also emerged as one of the major drivers of the electrogenic process. This is due, in part, to the organisms harboring such compounds transport and cellular localization of these redox-active cytochromes to the surface or near-surface of the aforementioned organisms (Figure 9.6). Thus, the surface (e.g. an iron oxide (Fe3þ) anode) has to be readily accessible to component(s) of the respiratory pathway of such organisms for optimal electrogenesis to occur. Using S. oneidensis as a model organism for examining the role of cytochromes in the electrogenic process, there are at least
BIOFILM ELECTROCHEMISTRY FOR ENHANCED MFC PERFORMANCE: A MOLECULAR BIOLOGY PERSPECTIVE mtrD (a) mtrE mtrF omcA mtrC mtrA mtrB mtrC mtrA mtrB mtrC mtrA 141 S. oneidensis MR-1 A. hydrophila cymA mtrD Fe(III) reducers mtrE mtrF undA mtrG orfA omcA mtrB F. balearica mtrK2 mtrK3 mtrK1 cymA1 mtrC mtrA mtrB mtrJ mtrI orfC cymA 2orfB mtrH R. ferrireducens mtoD mtoA (b) mtoB mtoC D. aromatica RCB mtoC mtoD mtoA mtoB G. capsiferriformans ES-2 Fe(II) oxidizers mtoD mtoA mtoB cymA pioB pioC S. lithotrophicus ES-1 pioA R. palustris TIE-1 FIGURE 9.6 (a) Genetic organization of iron-reducing vs (b) iron-oxidizing bacteria. Note close proximity of the Mtr (metal reducing) loci in the iron-reducing and Mto (metal oxidizing) cluster organisms. Source: Figure from Shi et al. (2012a) under STM Permission Guidelines. (For color version of this figure, the reader is referred to the online version of this book.) 10 gene products involved in iron reduction that are critical for some features of electrogenesis in this organism, an event that has been studied by many research groups for more than two decades (Arnold et al., 1990). Conveniently, most of the genes (especially mtr genes) involved in the process of iron oxide reduction and electrogenesis in MFCs are located in close proximity on the S. oneidensis genome (Figure 9.6). Figure 9.6 lists the organisms that are also iron-oxidizing bacteria for Fe(III) oxide Chelator Fe(II) Chelator–Fe(III) OmcA MtrC OmcA Flavinsox FlavinsRE MtrB OM MtrA PS e– CymA IM quinol quinone quinol quinone comparative purposes. Of the loci involved in metal reduction, these include mtrDEF, outer membrane cytochrome (omcA), followed by the mtrCAB genes. Figure 9.7 is a simplified recent schematic diagram of the mechanism of precisely how this process functions in S. oneidensis, elegantly described by Shi et al. (2012b). Prior to this exhaustive process of mechanistic functionality, the first genes found to be required for iron and manganese oxide reduction were performed FIGURE 9.7 Model of the decaheme reduction of insoluble iron by S. oneidensis OmcA (outer membrane cytochrome A), MtrABC (redox-active metal reducing proteins) incorporating the influence of inner membrane menaquinone pool as well as CymA (tetraheme cytochrome). Source: Figure from Shi et al. (2012b); courtesy of Frontiers Editorial Office in Switzerland. (For color version of this figure, the reader is referred to the online version of this book.)
142 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY in S. putrefaciens using transposon mutagenesis in 1998 (Beliaev and Saffarini, 1998). MtrB was found to be an outer membrane cytochrome while the upstream locus, mtrA, encodes a periplasmic decaheme cytochrome. MtrC of both S. putrefaciens and S. oneidensis is also an outer membrane cytochrome with apparent terminal iron reductase activity (Beliaev et al., 2001; Hartshorne et al., 2007). Shi et al. (2006) demonstrated that OcmA, yet another decaheme cytochrome, binds under acidic conditions to MtrC and, in fact, these form a highaffinity protein complex with one another. MtrF, MtrD and MtrE appear to be homologs of MtrCAB, yet one set cannot replace the other functionally, although some components can coaggregate. The outlier is that the mtrFDE loci appear to be highly expressed in biofilms and the Mtr system, in general, is required for optimal biofilm formation (Coursolle et al., 2010), similar to aggregation of bacteria on a conductive surface such as an iron oxide anode in MFCs. In summary, the order of electron flow for optimal electrogenesis of S. oneidensis is the following: cytoplasmic membrane-bound menaquinone, periplasmic tetraheme CymA with electron flowing through the b-barrel of MtrB to the decaheme cytochromes MtrA/ F and finally to two other decaheme proteins, MtrC and OmcA (Figure 9.7). The final destination for electrons prior to reduction of iron oxides is MtrC. Thus, again, it is intuitive that the genes encoding those proteins involved in iron reduction are localized in the following order on the S. oneidensis genome, mtrDEFeomcAemtrCAB, respectively. Similar to the discovered mechanism of proton pumping in the F1F0-ATP synthase using bacteriorhodopsin in membrane vesicles, scientists have proven that the protein complex of MtrCAB conducts electron when embedded within membrane vesicles (Hartshorne et al., 2009). In 2010, Tai et al. (2010) assessed potential networks of transcriptional regulation between chemotactic and electron transport properties and found that previously unknown roles of genes including cheA (a chemotaxis gene), mgtE-1 (an Mg2þ transport gene) and SO4572 (a triheme cytochrome gene). More recently, Leang et al. (2010) showed that the redox-cytochrome OmcS of G. sulfurreducens actually binds to the conductive pili, thereby contributing to their electrogenic properties. However, using a whole-cell cyclic voltammetric analysis of various mutant strains including (DmtrC/ DomcA), transmembrane pili (DpilM-Q, DmshH-Q, and DpilM-Q/DmshH-Q) and flagella (Dflg), Carmona et al. (2011) demonstrated that even with such mutations in place, often there are “by-pass” mechanisms of electron transfer, still allowing for some level of electrogenesis using cyclic voltammetric techniques. A synopsis of these results is shown in schematic form in Figure 9.8. Brief Synopsis of the S. oneidensis MR-1 Bioelectrochemical Machinery in Reverse: Potential Role in the Biosynthesis of Biofuels in MFCs The multiple proteins and other factors involved in bacterial electrogenesis in MFCs are complex. A process termed electron flow reversal, or, better put, electron diversion, is critical for a nonelectrogenic process for the purpose of generating single or multiple compounds of value. Ross et al. (2011) have helped simplifying many features of this process in their 2011 publication. Obviously, the goal of scientists working with electrogenic bacteria is to maximize their power density while wasting the energy harness in the carbon skeletons they consume for sustenance. From the above information collectively, it appears that TFP and cytochromes involved in the Mtr respiratory pathway facilitate the transfer of direct current in the form of electrons to one or more electrodes. Figure 9.9 helps simplifying what is currently understood of these systems and other drivers that will be discussed below. This process is also clearly dependent upon the carbon sources (or feedstock in more complex, multisubstrate systems). In that study, multiple isogenic mutants were created that (1) lacked the periplasmic fumarate reductase (FccA) and thus could not reduce fumarate using electrons derived from electrodes, a process adversely affected by nearly 90% by (2) deletion of mtrB, or worse, (3) the periplasmic cytochrome, MtrA, and prevention of menaquinone biosynthesis. Mediators for Accelerated Electron Transfer in Biofilms Flavins In addition to the mediator-less form of electrogenesis discussed above (TFP and soluble/insoluble cytochromes), S. oneidensis is an electrogenic bacterium that can also secrete extracellular, soluble, and redox-active mediators. Mediators can accept electrons in the anaerobic extracellular milieu or directly from the bacterial cell surface, and, due to their lower redox potential (Eo0 in V), they donate electrons directly to the anodic surface. One group of the S. oneidensis mediators is flavins. In addition to pili and cytochromes, S. oneidensis produces extracellular flavins that contribute to the electrogenic process and can actually be reduced, in part, via the Mtr/Omc system. Such flavins include riboflavin (vitamin B2) (Figure 9.10), flavin (isoalloxazine from which flavins are derived), and flavin mononuclotide (FMN). Thus, cells that are unbound to the metal oxide surface are still capable of reducing it, although the reduction process also requires reduced members of MtrC and OmcA (Figure 9.7).
143 MFCS FOR WASTEWATER TREATMENT WITH CONCOMITANT ELECTRICITY PRODUCTION Flavinox e– OmcA (a) Anode Anode FlavinRed e– OmcA MtrC Flavinox Anode MtrA e– e– MET-1 CymA FlavinRed Anode e– MtrC OM MtrB MtrB e– MtrA MtrA Periplasm e– e– DET-1 OM MtrB MtrB MET-1 CymA MtrA Periplasm DET-1 IM IM Wild-type, ΔpilM-Q, ΔmshH-Q, ΔpilM-Q/ ΔmshH-Q and Δflg ΔmtrC/ΔomcA IM e– DET-2 Wild-type, ΔmtrC/ΔomcA ΔmshH-Q and Δflg (c) FlavinRed e– Anode e– MET-2 e– OM OM e– MET-3 DET-2 IM IM msh-type pilus MET-2 pil-type pilus OM e– Flavinox FlavinRed e– pil-type pilus Flavinox Anode (b) Anode e– DET-3 msh-type pilus Anode Wild-type, ΔmtrC/ΔomcA ΔpilM-Q and Δflg ΔpilM-Q and ΔpilM-Q/ ΔmshH-Q e– MET-3 OM e– DET-3 IM ΔmshH-Q and ΔpilM-Q/ ΔmshH-Q (d) Anode Anode Anode Thick biofilm Wild-type, ΔmtrC/ΔomcA and ΔpilM-Q Δflg ΔmshH-Q and ΔpilM-Q/ ΔmshH-Q Poor biofilm formation FIGURE 9.8 DET and MET electron transfer pathways utilized by S. oneidensis and selected mutant strains. (a) Electron transfer via the cytochrome pool. Transmembrane pilus electron transfer via (b) pil-type pilus and via (c) msh-type pilus, and (d) biofilm formation behaviour. OM: Outer membrane and IM: Inner membrane. Source: Figure from Carmona et al. (2011) with permission from Elsevier Phenazines Phenazines are tricyclic, redox-active compounds that are produced by a number of species of the genus Pseudomonas. Pseudomonas aeruginosa, depending upon the mutations acquired in a specific microniche, can produce, or in the case of mutations within negative regulators or modulators such as RpoS, actually overproduce redox-active 1-hydroxyphenazine, pyorubrin, or pyocyanin (Figure 9.11). The process of phenazine biosynthesis in these organisms was highlighted by Mavrodi et al. (2001) where the entire pathway is based upon anthranilate synthesis and genes beginning with the acronym phn (for phenazine). A classic demonstration of the electrogenic contribution of P. aeruginosa phenazines to the electrogenic properties of this organism was shown using the power of classical bacterial genetics. Using a mutant approach, Rabaey et al. (2005a) showed that pyocyanin, pyorubrin, and 1-hydroxyphenazine could act as excellent mediators in MFCs. Bacteria that could not produce these mediators possessed reduced electrogenic properties relative to those genetically incapable of producing them, or organisms whose media were amended with known quantities of each phenazine. Given the known electrochemical potential of each of the aforementioned mediators, it is not surprising that power density was greatest in MFCs containing pyocyanin > pyorubrin (aeruginosin A; Rabaey et al., 2005a) > 1-hydroxyphenazine. Supportive of these results were those of Luo et al. (2009b) in the isolation of strain RE7. MFCS FOR WASTEWATER TREATMENT WITH CONCOMITANT ELECTRICITY PRODUCTION MFC Reactor Designs There are many different types of MFC bioreactors. They include single-chamber, dual-chamber, multichamber, membrane-less, multianode, multicathode and so on. Many MFC reactors were discussed by Du
144 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY (a) Cytoplasmic membrane Periplasmic space Outer membrane Extracellular space e– MtrA MtrC soluble mediator Electrode direct MtrB CymA MQ MQH2 CymA FccA fumarate succinate (b) CymA –300 mV MtrA:MtrC FccA Potential window of electrode linked fumarate reduction V vs SHE –200 mV –100 mV 0 mV fum/suc +100 mV FIGURE 9.9 (a) Mechanism of electron reversal (or inward electron flux) in S. oneidensis and the gene products involved in this process. (b) Fumarate reduction in S. oneidensis: windows of redox and midpoint (deep-red lines) potentials for each electron carrier. Source: Figure from Ross et al. (2011) with permission from Public Library of Science. (For interpretation of the references to color in this figure legend, the reader is referred to the online version of this book.) et al. (2007). More recently, Zhou et al. (2013) reviewed some new MFC reactors and their combinations, including MFCs operated as microbial electrolysis cells (MECs) to produce bio-products such as hydrogen and methane. It should be pointed out that improvement in MFC reactor design must consider cost and maintenance. Complicated designs are not only costly but also prone to biofouling, causing maintenance and sustainability problems. A simplistic tubular MFC reactor with convective axial flow was proposed (Zhou et al., 2013). To reduce cost and fouling, no membrane was used. To prevent oxygen back-diffusion into the anodic region, a substantial flow rate from the anode to the cathode is required. This means that the biofilm has to be highly efficient in the digestion of organic matter in wastewater streams. This type of design will become attractive only when robust “super-bug” biofilms are successfully engineered.
MFCS FOR WASTEWATER TREATMENT WITH CONCOMITANT ELECTRICITY PRODUCTION 145 FIGURE 9.10 (a). The riboflavin (vitamin B2) biosynthetic genes of S. oneidensis. (b). Structures of riboflavin, isoalloxazine, and flavin mononucleotide. Source: Figure courtesy of Dr Jeff Gralnick of University of Minnesota. (For color version of this figure, the reader is referred to the online version of this book.) the metabolic pathways to utilize high-grade organic carbons such as cellulose, hemicellulose, various hexoses and phenylpropane moieties (components of lignin). Most of the electrogenic microbes capable of DET feed only on low-grade organic carbons such as VFAs and alcohols. Only a few organisms such as R. ferrireducens (Chaudhuri and Lovley, 2003; Schröder, 2007) utilize glucose, while Geobacter and Shewanella strains cannot (Lovley, 2006a). This limits MFC power output because high-grade organic carbons are unutilized. Simple Biodegradable Organics FIGURE 9.11 Microcentrifuge tubes containing chloroformextractable pyocyanin (blue bottom layer and the “merlot” colored) and water-soluble pyorubrin layer (top). The tube on the left is derived from a lasI rhlI mutant that is incapable of quorum sensing and, as such, is incapable of producing pigments, while that on the right is from rpoS mutant bacteria that overproduce both pyocyanin and pyorubrin. Source: Suh et al. (1999). (For interpretation of the references to color in this figure legend, the reader is referred to the online version of this book.) Substrates Used in MFCs In MFCs, the substrates greatly impact their performances such as power density and Coulombic efficiency (Pant et al., 2010). The substrates range from the simple volatile fatty acids (VFAs) to complex compounds such as lignocellulosic biomass. Anaerobes evolved when the earth’s atmosphere was still anaerobic long before aerobes evolved. Many of them lack Acetate and glucose are two common substrates in laboratory studies. Compared to the recalcitrant substrates, they are far more readily utilized by microbes for energy generation. Thus, they are usually used as the carbon source for microbes used in MFCs. Acetate has an advantage that at normal temperatures, it is not a good nutrient for fermentation and methanogenesis. In contrast, glucose is a fermentable sugar that can be consumed by the processes of fermentation and methanogenesis (Pant et al., 2010). Thus, the Coulombic efficiency of acetate is usually higher than glucose. However, glucose can be used to promote the microbial diversity of a biofilm consortium. When glucose was used as the substrate, a maximum power density of 216 mW m2 was achieved (Rabaey et al., 2003), while it reached 506 mW m2 for acetate (Liu et al., 2005b). Some other simple substrates such as butyrate have also been used as the substrate in MFCs.
146 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY Wastewater Types Various wastewaters have been tested as substrates for MFCs because they contain many different kinds of organic carbon molecules. They are attractive for use in MFCs because the organic carbons are otherwise wasted. As shown in Table 9.3, the output power density is dependent on the wastewater quality (high COD values) and the MFC reactor structure. For example, a maximum power density of 528 mW m2 for brewery wastewater was obtained (Feng et al., 2008), while an average power density of 72 mW m2 was achieved for domestic wastewater (Sharma and Kundu, 2010). Some biorefractory wastewaters such as dye, leachates and pharmaceutical wastewater have also been tested for MFC power generation. A landfill leachate containing heavy metals, dissolved organic matters and other matters achieved a maximum power density of 1.38 mW m2 (Greenman et al., 2009). A maximum power density of 9.1 W m3 was achieved when using phenol as the sole carbon source. While glucose was added as a supplement, the maximum power density increased to 28.3 W m3 (Luo et al., 2009a). In addition, some refractory compounds such as pyridine, quinoline and indole were also used as substrates for MFCs (Hu et al., 2011). Lignocellulosic Biomass Lignocellulosic biomass includes corn stover, straw, wheat stover, algae and others. The primary components in lignocellulosic biomass are cellulose, hemicellulose and lignin. Compositions differ for different types of biomass. Lignocellulosic biomass is considered unfermentable because most microbes cannot degrade it TABLE 9.3 An Updated List of Substrates Used in MFCs Substrates Reactor Style Pmax (mW mL2) References Glucose Dual chamber 283 Rahimnejad et al. (2011) Cheese whey Dual chamber 42 Stamatelatou et al. (2011) Food waste Single chamber 207.2** Kannaiah and Venkata (2011) Palm oil mill effluent with acetate Dual chamber 622 Jong et al. (2011) Dairy wastewater Single chamber 5.7* Ayyaru and Dharmalingam (2011) Leachates Single chamber 20.9 Va’zquez-Larios et al. (2011) Composite food waste Single chamber 107.89 Goud et al. (2011) Pharmaceutical wastewater Single chamber 177.36 Velvizhi and Venkata (2011) Azo dye Single-chamber e Sun et al. (2011) Human feces wastewater Dual chamber 70.8 Du et al. (2011) Synthetic penicillin wastewater with glucose Single chamber 101.2* Wen et al. (2011) Paper wastewater Single chamber 125*** Velasquez et al. (2011) Dairy wastewater Single chamber 25*** Brewery and bakery wastewaters Single chamber 10*** Distillery wastewater Single chamber 245.34 Mohanakrishna et al. (2012) Sewage sludge Tubular MFC 73 Yuan et al. (2012) Primary clarifier effluent Single chamber 13 Ishii et al. (2012) Alcohol distillery wastewater Dual chamber 1000 Ha et al. (2012) Agriculture wastewater Single chamber 13 Nimje et al. (2012) Domestic wastewater Single chamber 42 Paper wastewater Single chamber 8 Food/dairy wastewater Single chamber 15 Bad wine Dual chamber 3.82* 3 * In W m . ** Calculated from power and current densities. *** In mA m2. Rengasamy and Berchmans (2012)
REFERENCES without pretreatment and lignin is optimally degraded under aerobic conditions via several dioxygenasetype enzymes, although some anaerobic bacteria can degrade it, albeit slowly. Pretreatment methods include mechanical, hydrothermal, biological, chemical, ammonia or supercritical CO2 explosion and ionic liquid extraction (Gu, 2013). An MFC using corn stover after steam-explosion pretreatment as the substrate achieved a maximum power density of 861 mW m2 (Zuo et al., 2006). MFCs fed with Chlorella vulgaris and Ulva lactuca powders achieved maximum power densities of 0.98 W m2 (277 W m3) and 0.76 W m2 (215 W m3), respectively (Velasquez-Orta et al., 2009). Cellulose is relative easy to utilize by MFCs compared with lignocellulosic biomass. A maximum power density of 272 mW m2 was achieved using carboxymethyl cellulose as substrate in an MFC (Rezaei et al., 2009). This means that it is possible to utilize the tissue paper (cellulose) in municipal wastewater as substrate. Table 9.3 shows the list of substrates used for MFCs studied until 2013. SUMMARY AND PERSPECTIVES This chapter discusses the operating principles of MFCs and various aspects in bioelectrochemistry in MFC research. Although tremendous advances have been made around 2013 in academic MFC research including a much better understanding of biofilm electrochemistry and better reactor designs, major technological hurdles remain for practical MFC applications beyond powering sensor devices. It is unreasonable to expect MFCs to reach power densities on par with those from chemical fuel cells because MFCs are powered by lowenergy-density fuels such as dilute organic matter in wastewaters. However, it is still necessary to increase MFC power generation to what would be considered a useful level (e.g. to offset part of the energy input in wastewater treatment), much higher than what has been achieved. Various approaches have been attempted to increase MFC performance including improved reactor designs, electrode and membrane materials, feedstock selection and modification, introduction of exogenous mediators, and utilization of secreted endogenous mediators. Unfortunately, many of the improvements come with inherent cost increases with little hope for practical applications. Some MFC researchers have come to realize that a breakthrough in biofilm engineering should be explored. Recent discoveries such as interspecies electron transfer, conductive cell aggregates and long-distance conductive filaments provide new hope for means to engineer robust “super-bug” biofilms with greatly enhanced electron transfer capacity and a 147 voracious appetite for complex organic matter digestion. The dawn of a new era for MFC research might be in sight and the synergistic involvement of biochemical and environmental engineers, microbiologists and molecular biologists may soon bear fruit in this exciting field of practical research. References Aldrovandi, A., Marsili, E., Stante, L., Paganin, P., Tabacchioni, S., Giordano, A., 2009. Sustainable power production in a membraneless and mediator-less synthetic wastewater microbial fuel cell. Bioresour. Technol. 100, 3252e3260. Allen, R.M., Bennetto, H.P., 1993. Microbial fuel-cells electricity production from carbohydrates. Appl. Biochem. Biotech. 39/40, 27e40. Angenent, L.T., Karim, K., Dahhan, M.H.A., Wrenn, B.A., Rosa, D.E., 2004. Production of bioenergy and biochemicals from industrial and agricultural wastewater. Trends Biotechnol. 22, 477e485. Arnold, R.G., Hoffmann, M.R., Dichristina, T.J., Picardal, F.W., 1990. Regulation of dissimilatory Fe(III) reduction activity in Shewanella putrefaciens. Appl. Environ. Microbiol. 56, 2811e2817. Aulenta, F., Reale, P., Canosa, A., Rossetti, S., Panero, S., Majone, M., 2010. Characterization of an electro-active biocathode capable of dechlorinating trichloroethene and cis-dichloroethene to ethene. Biosens. Bioelectron. 25, 1796e1802. Ayyaru, S., Dharmalingam, S., 2011. Development of MFC using sulphonated polyether ether ketone (SPEEK) membrane for electricity generation from waste water. Bioresour. Technol. 102, 11167e11171. Bao, Y., Wu, X., 2004. Progress in research for biofuel cell. Electrochemistry 10, 1e8. Behera, M., Jana, P.S., Ghangrekar, M.M., 2010. Performance evaluation of low cost microbial fuel cell fabricated using earthen pot with biotic and abiotic cathode. Bioresour. Technol. 101, 1183e1189. Beliaev, A.S., Saffarini, D.A., 1998. Shewanella putrefaciens mtrB encodes an outer membrane protein required for Fe(III) and Mn(IV) reduction. J. Bacteriol. 180, 6292e6297. Beliaev, A.S., Saffarini, D.A., McLaughlin, J.L., Hunnicutt, D., 2001. MtrC, an outer membrane decahaem c cytochrome required for metal reduction in Shewanella putrefaciens MR-1. Mol. Microbiol. 39, 722e730. Biffinger, J.C., Pietron, J., Ray, R., 2007. A biofilm enhanced miniature microbial fuel cell using Shewanella oneidensis DSP10 and oxygen reduction cathodes. Biosens. Bioelectron. 22, 1672e1679. Biffinger, J.C., Pietron, J., Bretschger, O., 2008. The influence of acidity on microbial fuel cell containing Shewanella oneidensis. Biosens. Bioelectron. 24, 906e911. Bond, D.R., Lovley, D.R., 2003. Electricity production by Geobacter sulfurreducens attached to electrodes. Appl. Environ. Microbiol. 69, 1548e1555. Bond, D.R., Lovley, D.R., 2005. Evidence for involvement of an electron shuttle in electricity generation by Geothrix fermentans. Appl. Environ. Microbiol. 71, 2186e2189. Bond, D.R., E.Holmes, D., tender, L.M., Lovely, D.R., 2002. Electrodereducing microorganisms that harvest energy from marine sediments. Science 295, 483e485. Bullen, R.A., Arnot, T.C., Lakeman, J.B., Walsh, F.C., 2006. Biofuel cells and their development. Biosens. Bioelectron. 21, 2015e2045. Busalmen, J.P., Nunez, A.E., Berna, A., Feliu, J.M., 2008. C-type cytochromes wire electricity-producing bacteria to electrodes. Angew. Chem. Int. Edit. 47, 4874e4877. Cao, X., Huang, X., Liang, P., Boon, N., Fan, M., Zhang, L., Zhang, X., 2009. A completely anoxic microbial fuel cell using a photobiocathode for cathodic carbon dioxide reduction. Energy Environ. Sci. 2, 498.
148 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY Carmona, M.A.A., Harnisch, F., Fitzgerald, L.A., Biffinger, J.C., Ringeisen, B.R., Schröder, U., 2011. Cyclic voltammetric analysis of the electron transfer of Shewanella oneidensis MR-1 and nanofilament and cytochrome knock-out mutants. Bioelectrochemistry 81, 74e80. Chaudhuri, S.K., Lovley, D.R., 2003. Electricity generation by direct oxidation of glucose in mediatorless microbial fuel cells. Nat. Biotechnol. 21, 1229e1232. Choi, Y., Jung, E., Kim, S., Jung, S., 2003a. Membrane fluidity sensoring microbial fuel cell. Bioelectrochemistry 59, 121e127. Choi, Y., Kim, N., Kim, S., Jung, S., 2003b. Dynamic behaviors of redox mediators within the hydrophobic layers as an important factor for effective microbial fuel cell operation. B. Korean Chem. Soc. 24, 437e440. Clauwaert, P., Rabaey, K., Aelterman, P., Schamphelaire, L.D., Pham, H.T., Boeckx, P., Boon, N., Verstraete, W., 2007a. Biological denitrification in microbial fuel cells. Environ. Sci. Technol. 41, 3354e3360. Clauwaert, P., Ha, D.V.D., Boon, N., Verbeken, K., Verhaege, M., Rabaey, K., Verstraete, W., 2007b. Open air biocathode enables effective electricity generation with microbial fuel cells. Environ. Sci. Technol. 41, 7564e7569. Cooney, M.J., Roschi, E., Marison, I.W., Comninellis, C., Stockar, U.v., 1996. Physiologic studies with the sulfate-reducing bacterium Desulfovibrio desulfuricans evaluation for use in a biofuel cell. Enzyme Microb. Tech. 18, 358e365. Cournet, A., Délia, M.L., Bergel, A., Roques, C., Bergé, M., 2010. Electrochemical reduction of oxygen catalyzed by a wide range of bacteria including gram-positive. Electrochem. Commun. 12, 505e508. Coursolle, D., Baron, D.B., Bond, D.R., Gralnick, J.A., 2010. The Mtr respiratory pathway is essential for reducing flavins and electrodes in Shewanella oneidensis. J. Bacteriol. 192, 467e474. Craig, L., Pique, M.E., Tainer, J.A., 2004. Type IV pilus structure and bacterial pathogenicity. Nat. Rev. Microbiol. 2, 363e378. Du, Z., Li, H., Gu, T., 2007. A state of the art review on microbial fuel cells: a promising technology for wastewater treatment and bioenergy. Biotechnol. Adv. 25, 464e482. Du, F., Li, Z., Yang, S., Xie, B., Liu, H., 2011. Electricity generation directly using human feces wastewater for life support system. Acta Astronaut. 68, 1537e1547. Dumas, C., Basseguy, R., Bergel, A., 2008. Microbial electrocatalysis with Geobacter sulfurreducens biofilm on stainless steel cathodes. Electrochim. Acta 53, 2494e2500. Erable, B., Duteanu, N.M., Ghangrekar, M.M., Dumas, C., Scott, K., 2010a. Application of electro-active biofilms. Biofouling 26, 57e71. Erable, B., Vandecandelaere, I., Faimali, M., Delia, M.L., Etcheverry, L., Vandamme, P., Bergel, A., 2010b. Marine aerobic biofilm as biocathode catalyst. Bioelectrochemistry 78, 51e56. Feng, Y., Wang, X., Logan, B.E., Lee, H., 2008. Brewery wastewater treatment using air-cathode microbial fuel cells. Appl. Microbiol. Biotechnol. 78, 873e880. Fitzgerald, L.A., Petersen, E.R., Ray, R.I., Little, B.J., Cooper, C.J., Howard, E.C., Ringeisen, B.R., Biffinger, J.C., 2012. Shewanella oneidensis MR-1 Msh pilin proteins are involved in extracellular electron transfer in microbial fuel cells. Process Biochem. 47, 170e174. Franks, A.E., Nevin, K.P., 2010. Microbial fuel cells, a current review. Energies 3, 899e919. Freguia, S., Tsujimura, S., Kano, K., 2010. Electron transfer pathways in microbial oxygen biocathodes. Electrochim. Acta 55, 813e818. Gil, G.C., Chang, I.S., Kim, B.H., Kim, M., Jang, J.K., Park, H.S., Kim, H.J., 2003. Operational parameters affecting the performance of a mediator-less microbial fuel cell. Biosens. Bioelectron. 18, 327e334. Gorby, Y.A., Yanina, S., McLean, J.S., Rosso, K.M., Moyles, D., Dohnalkova, A., Beveridge, T.J., Chang, I.S., Kim, B.H., Kim, K.S., Culley, D.E., Reed, S.B., Romine, M.F., Saffarini, D.A., Hill, E.A., Shi, L., Elias, D.A., Kennedy, D.W., Pinchuk, G., Watanabe, K., Ishii, S., Logan, B., Nealson, K.H., Fredrickson, J.K., 2006. Electrically conductive bacterial nanowires produced by Shewanella oneidensis strain MR-1 and other microorganisms. Proc. Natl. Acad. Sci. USA 103, 11358e11363. Goud, R.K., Babu, P.S., Mohan, S.V., 2011. Canteen based composite food waste as potential anodic fuel for bioelectricity generation in single chambered microbial fuel cell (MFC): bio-electrochemical evaluation under increasing substrate loading condition. Int. J. Hydrogen Energy 36, 6210e6218. Greenman, J., Gálvez, A., Giusti, L., Ieropoulos, I., 2009. Electricity from landfill leachate using microbial fuel cells: comparison with a biological aerated filter. Enzyme Microb. Tech. 44, 112e119. Gregory, K.B., Lovley, D.R., 2005. Remediation and recovery of uranium from contaminated subsurface environments with electrodes. Environ. Sci. Technol. 39, 8943e8947. Gu, T., 2012. New understandings of biocorrosion mechanisms and their classifications. J. Chem. Technol. Biotechnol. 4, 3e6. Gu, T., 2013. Green Biomass Pretreatment for Biofuels Production. Springer, Berlin-New York. Guo, K., Hassett, D.J., Gu, T., 2012. Microbial fuel cells: electricity generation from organic wastes by microbes. Advances in microbial fuel cells for potential energy production from organic feed streams. In: Arora, R. (Ed.), Microbial Biotechnology: Energy and Environment. CAB International, Oxon, United Kingdom, ISBN 978-1845939564. (Chapter 12). Ha, P.T., Lee, T.K., Rittmann, B.E., Park, J., Chang, I.S., 2012. Treatment of alcohol distillery wastewater using a bacteroidetes-dominant thermophilic microbial fuel cell. Environ. Sci. Technol. 46, 3022e3030. Hamid, R.Y., Carver, S.M., Christy, A.D., Tuovinen, O.H., 2008. Cathodic limitations in microbial fuel cells: an overview. J. Power Sources 180, 683e694. Hartshorne, R.S., Jepson, B.N., Clarke, T.A., Field, S.J., Fredrickson, J., Zachara, J., Shi, L., Butt, J.N., Richardson, D.J., 2007. Characterization of Shewanella oneidensis MtrC: a cell-surface decaheme cytochrome involved in respiratory electron transport to extracellular electron acceptors. J. Biol. Inorg. Chem. 12, 1083e1094. Hartshorne, R.S., Reardon, C.L., Ross, D., Nuester, J., Clarke, T.A., Gates, A.J., Mills, P.C., Fredrickson, J.K., Zachara, J.M., Shi, L., Beliaev, A.S., Marshall, M.J., Tien, M., Brantley, S., Butt, J.N., Richardson, D.J., 2009. Characterization of an electron conduit between bacteria and the extracellular environment. Proc. Natl. Acad. Sci. USA 106, 22169e22174. Hassan, S.H.A., Kim, Y.S., Oh, S.E., 2012. Power generation from cellulose using mixed and pure cultures of cellulose-degrading bacteria in a microbial fuel cell. Enzyme Microb. Technol. 51, 269e273. Hatch, J.L., Finneran, K.T., 2008. Influence of reduced electron shuttling compounds on biological H2 production in the fermentative pure culture Clostridium beijerinckii. Curr. Microbiol. 56, 268e273. He, Z., Angenent, L.T., 2006. Application of bacterial biocathodes in microbial fuel cells. J. Electroanal. Chem. 18, 2009e2015. Holmes, D.E., Bond, D.R., O’Neil, R.A., Reimers, C.E., Tender, L.R., Lovley, D.R., 2004. Microbial communities associated with electrodes harvesting electricity from a variety of aquatic sediments. Microb. Ecol. 48, 178e190. Holmes, D.E., Mester, T., O’Neil, R.A., Perpetua, L.A., Larrahondo, M.J., Glaven, R., Sharma, M.L., Ward, J.E., Nevin, K.P., Lovley, D.R., 2008. Genes for two multicopper proteins required for Fe(III) oxide reduction in Geobacter sulfurreducens have different expression patterns both in the subsurface and on energyharvesting electrodes. Microbiology 154, 1422e1435. Hou, H., Li, L., Cho, Y., 2009. Microfabricated microbial fuel cell arrays reveal electrochemically active microbes. PLoS One 4, 6570.
REFERENCES Hu, W.J., Niu, C.G., Wang, Y., Zeng, G.M., Wu, Z., 2011. Nitrogenous heterocyclic compounds degradation in the microbial fuel cells. Process Saf. Environ. Prot. 89, 133e140. Huang, L., Chai, X., Cheng, S., Chen, G., 2011a. Evaluation of carbonbased materials in tubular biocathode microbial fuel cells in terms of hexavalent chromium reduction and electricity generation. Chem. Eng. J. 166, 652e661. Huang, L., Cheng, S., Chen, G., 2011b. Bioelectrochemical systems for efficient recalcitrant wastes treatment. J. Chem. Technol. Biotechnol. 86, 481e491. Huang, L., Regan, J.M., Quan, X., 2011c. Electron transfer mechanisms, new applications, and performance of biocathode microbial fuel cells. Bioresour. Technol. 102, 316e323. Huang, L., Cheng, S., Hassett, D.J., Gu, T., 2012. Wastewater treatment with concomitant bioenergy production using microbial fuel cells. In: Sharma, S.K., Sanghi, R. (Eds.), Water Treatment and Pollution Prevention: Advances in Research. Springer-Verlag, Berlin-New York, pp. 405e452. (Chapter 18). Ishii, S.i., Suzuki, S., Norden-Krichmar, T.M., Nealson, K.H., Sekiguchi, Y., Gorby, Y.A., Bretschger, O., 2012. Functionally stable and phylogenetically diverse microbial enrichments from microbial fuel cells during wastewater treatment. PLoS One 7, 1e12. Jeremiasse, A.W., Hamelers, H.V., Buisman, C.J., 2010. Microbial electrolysis cell with a microbial biocathode. Bioelectrochemistry 78, 39e43. Jong, B.C., Liew, P.W.Y., Juri, M.L., Kim, B.H., Mohd Dzomir, A.Z., Leo, K.W., Awang, M.R., 2011. Performance and microbial diversity of palm oil mill effluent microbial fuel cell. Lett. Appl. Microbiol. 53, 660e667. Kannaiah, G.R., Venkata, M.S., 2011. Pre-fermentation of waste as a strategy to enhance the performance of single chambered microbial fuel cell (MFC). Int. J. Hydrogen Energy 36, 13753e13762. Kim, B.H., Ikeda, T., Park, H.S., Kim, H.J., Hyun, M.S., Kano, K., Takagi, K., Tatsumi, H., 1999. Electrochemical activity of an Fe(III)reducing bacterium, Shewanella putrefaciens IR-1, in the presence of alternative electron acceptors. Biotechnol. Technol. 13, 475e478. Kim, N., Choi, Y., Jung, S., 2000a. Development of microbial fuel cells using Proteus vulgaris. B. Korean Chem. Soc. 21, 44e48. Kim, N., Choi, Y., Jung, S., 2000b. Effect of initial carbon sources on the performance of microbial fuel cells containing Proteus vulgaris. Biotechnol. Bioeng. 70, 109e114. Kim, H.J., Park, H.S., Hyun, M.S., Chang, I.S., Kim, M., Kima, B.H., 2002. A mediator-less microbial fuel cell using a metal reducing bacterium, Shewanella putrefaciens. Enzyme Microb. Technol. 30, 145e152. Leang, C., Krushkal, J., Ueki, T., Puljic, M., Sun, J., Juarez, K., Nunez, C., Reguera, G., DiDonato, R., Postier, B., Adkins, R.M., Lovley, D.R., 2009. Genome-wide analysis of the RpoN regulon in Geobacter sulfurreducens. BMC Genomics 10, 331. Leang, C., Qian, X., Mester, T., Lovley, D.R., 2010. Alignment of the c-type cytochrome OmcS along pili of Geobacter sulfurreducens. Appl. Environ. Microbiol. 76, 4080e4084. Lee, J., Phung, N.T., Chang, I.S., Kim, B.H., Sung, H.C., 2003. Use of acetate for enrichment of electrochemically active microorganisms and their 16S rDNA analyses. FEMS Microbiol. Lett. 223, 185e191. Lefebvre, O., Mamun, A.A., Ng, H.Y., 2008. A microbial fuel cell equipped with a biocathode for organic removal and denitrification. Water Sci. Technol. 58, 881e885. Li, W.W., Sheng, G.P., Liu, X.W., Yu, H.Q., 2011. Recent advances in the separators for microbial fuel cells. Bioresour. Technol. 102, 244e252. Liu, H., Cheng, S., Logan, B.E., 2005a. Power generation in fedbatch microbial fuel cells as a function of ionic strength, temperature, and reactor configuration. Environ. Sci. Technol. 39, 5488e5493. 149 Liu, H., Cheng, S., Logan, B.E., 2005b. Production of electricity from acetate or butyrate using a single-chamber microbial fuel Cell. Environ. Sci. Technol. 39, 658e662. Liu, G., Yates, M.D., Cheng, S., Call, D.F., Sun, D., Logan, B.E., 2011. Examination of microbial fuel cell start-up times with domestic wastewater and additional amendments. Bioresour. Technol. 102, 7301e7306. Logan, B.E., 2009. Exoelectrogenic bacteria that power microbial fuel cells. Nat. Rev. Microbiol. 7, 375e381. Logan, B.E., Regan, J.M., 2006. Electricity-producing bacterial communities in microbial fuel cells. Trends Microbiol. 14, 512e518. Logan, B.E., Hamelers, B., Rozendal, R., Schröder, U., Keller, J., Freguia, S., Aelterman, P., Verstraete, W., Rabaey, K., 2006. Microbial fuel cells: methodology and technology. Environ. Sci. Technol. 40, 5181e5192. Lojou, E., Durand, M.C., Dolla, A., Bianco, P., 2002. Hydrogenase activity control at Desulfovibrio vulgaris cell. J. Electroanal. Chem. 14, 913e922. Lovley, D.R., 2006a. Bug juice: harvesting electricity with microorganisms. Nat. Rev. Microbiol. 4, 497e508. Lovley, D.R., 2006b. Microbial fuel cells: novel microbial physiologies and engineering approaches. Curr. Opin. Biotechnol. 17, 327e332. Lovley, D.R., 2008. The microbe electric: conversion of organic matter to electricity. Curr. Opin. Biotechnol. 19, 564e571. Lovley, D.R., Ueki, T., Zhang, T., Malvankar, N.S., Shrestha, P.M., Flanagan, K.A., Aklujkar, M., Butler, J.E., Giloteaux, L., Rotaru, A.E., Holmes, D.E., Franks, A.E., Orellana, R., Risso, C., Nevin, K.P., 2011. Geobacter: the microbe electric’s physiology, ecology, and practical applications. Adv. Microb. Physiol. 59, 1e100. Luo, H., Liu, G., Zhang, R., Jin, S., 2009a. Phenol degradation in microbial fuel cells. Chem. Eng. J. 147, 259e264. Luo, H.P., Liu, G.l., Zhang, R.D., Cao, L.X., 2009b. Isolation and characterization of electrochemical active bacterial Pseudomonas aeruginosa strain RE7. Huan Jing Ke Xue 30, 2118e2123. Makarieva, A.M., Gorshkov, V.G., Li, B.L., 2008. Energy budget of the biosphere and civilization: rethinking environmental security of global renewable and non-renewable resources. Ecol. Complex. 5, 281e288. Manohar, A.K., Bretschger, O., Nealson, K.H., 2008. The use of electrochemical impedance spectroscopy (EIS) in the evaluation of the electrochemical properties of a microbial fuel cell. Bioelectrochemistry 72, 149e154. Marsili, E., Baron, D.B., Shikhare, I.D., Coursolle, D., Gralnick, J.A., Bond, D.R., 2008. Shewanella secretes flavins that mediate extracellular electron transfer. Microbiology 105, 3968e3973. Mavrodi, D.V., Bonsall, R.F., Delaney, S.M., Soule, M.J., Phillips, G., Thomashow, L.S., 2001. Functional analysis of genes for biosynthesis of pyocyanin and phenazine-1-carboxamide from Pseudomonas aeruginosa PAO1. J. Bacteriol. 183, 6454e6465. McKinlay, J.B., Zeikus, J.G., 2004. Extracellular iron reduction is mediated in part by neutral red and hydrogenase in Escherichia coli. Appl. Environ. Microbiol. 70, 3467e3474. Min, B., Cheng, S., Logan, B., 2005. Electricity generation using membrane and salt bridge microbial fuel cells. Water Res. 39, 1675e1686. Mohanakrishna, G., Mohan, S.K., Mohan, S.V., 2012. Carbon based nanotubes and nanopowder as impregnated electrode structures for enhanced power generation: evaluation with real field wastewater. Appl. Energy 95, 31e37. Moon, H., Chang, I.S., Kim, B.H., 2006. Continuous electricity production from artificial wastewater using a mediator-less microbial fuel cell. Bioresour. Technol. 97, 621e627. Morita, M., Malvankar, N.S., Franks, A.E., Summers, Z.M., Giloteaux, L., Rotaru, A.E., Rotaru, C., Lovley, D.R., 2011. Potential
150 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY for direct interspecies electron transfer in methanogenic wastewater digester aggregates. MBio. 2, e00159e11. Neto, S.A., Forti, J.C., Andrade, A.R., 2010. An overview of enzymatic biofuel cells. Electrocatalysis 1, 87e94. Nevin, K.P., Richter, H., Covalla, S.F., Johnson, J.P., Woodard, T.L., Orloff, A.L., Jia, H., Zhang, M., Lovley, D.R., 2008. Power output and columbic efficiencies from biofilms of Geobacter sulfurreducens comparable to mixed community microbial fuel cells. Environ. Microbiol. 10, 2505e2514. Newman, D.K., Kolter, R., 2000. A role for excreted quinones in extracellular electron transfer. Nature 450, 94e97. Nguyen, T.A., Lu, Y., Yang, X., Shi, X., 2007. Carbon and steel surfaces modified by Leptothrix discophora SP-6: characterization and implications. Environ. Sci. Technol. 41, 7987e7996. Niessen, J., Schröder, U., Scholz, F., 2004. Exploiting complex carbohydrates for microbial electricity generation-a bacterial fuel cell operating on starch. Electrochem. Commun. 6, 955e958. Nimje, V.R., Chen, C.Y., Chen, H.R., Chen, C.C., Huang, Y.M., Tseng, M.J., Cheng, K.C., Chang, Y.F., 2012. Comparative bioelectricity production from various wastewaters in microbial fuel cells using mixed cultures and a pure strain of Shewanella oneidensis. Bioresour. Technol. 104, 315e323. O’Toole, G.A., Kolter, R., 1998. Flagellar and twitching motility are necessary for Pseudomonas aeruginosa biofilm development. Mol. Microbiol. 30, 295e304. Oh, S.E., Logan, B.E., 2005. Hydrogen and electricity production from a food processing wastewater using fermentation and microbial fuel cell technologies. Water Res. 39, 4673e4682. Osman, M.H., Shah, A.A., Walsh, F.C., 2010. Recent progress and continuing challenges in bio-fuel cells. Part II: microbial. Biosens. Bioelectron. 26, 953e963. Pant, D., Van Bogaert, G., Diels, L., Vanbroekhoven, K., 2010. A review of the substrates used in microbial fuel cells (MFCs) for sustainable energy production. Bioresour. Technol. 101, 1533e1543. Park, D.H., Zeikus, J.G., 1999. Utilization of electrically reduced neutral red by Actinobacillus succinogenes: physiological function of neutral red in membrane driven fumarate reduction and energy conservation. J. Bacteriol. 181, 2403e2410. Park, D.H., Zeikus, J.G., 2000. Electricity generation in microbial fuel cells using neutral red as an electronophore. Appl. Environ. Microbiol. 66, 1292e1297. Park, D.H., Zeikus, J.G., 2003. Improved fuel cell and electrode designs for producing electricity from microbial degradation. Biotechnol. Bioeng. 81, 348e355. Petter, M.C., 1911. Electrical effects accompanying the decomposition of organic compounds. Proc. R Soc. Lond. B Biol. Sci. 84, 260e276. Pfeffer, C., Larsen, S., Song, J., Dong, M., Besenbacher, F., Meyer, R.L., Kjeldsen, K.U., Schreiber, L., Gorby, Y.A., El-Naggar, M.Y., Leung, K.M., Schramm, A., Risgaard-Petersen, N., Nielsen, L.P., 2012. Filamentous bacteria transport electrons over centimetre distances. Nature 491, 218e221. Pham, C.A., Jung, S.J., Phung, N.T., Lee, J., Chang, I.S., Kim, B.H., Yi, H., Chun, J., 2003. A novel electrochemically active and Fe(III)reducing bacterium phylogenetically related to Aeromonas hydrophila, isolated from a microbial fuel cell. FEMS Microbiol. Lett. 223, 129e134. Pham, T.H., Rabaey, K., Aelterman, P., Clauwaert, P., De Schamphelaire, L., Boon, N., Verstraete, W., 2006. Microbial fuel cells in relation to conventional anaerobic digestion technology. Eng. Life Sci. 6, 285e292. Prasad, D., Arun, S., Murugesan, M., Padmanaban, S., Satyanarayanan, R.S., Berchmans, S., Yegnaraman, V., 2007. Direct electron transfer with yeast cells and construction of a mediatorless microbial fuel cell. Biosens. Bioelectron. 22, 2604e2610. Qian, F., Baum, M., Gu, Q., Morse, D.E., 2009. A 1.5 mL microbial fuel cell for on-chip bioelectricity generation. Lab Chip 9, 3076e3081. Rabaey, K., Lissens, G., Siciliano, S.D., Verstraete, W., 2003. A microbial fuel cell capable of converting glucose to electricity at high rate and efficiency. Biotechnol. Lett. 25, 1531e1535. Rabaey, K., Boon, N., Hofte, M., Verstraete, W., 2005a. Microbial phenazine production enhances electron transfer in biofuel cells. Environ. Sci. Technol. 39, 3401e3408. Rabaey, K., Lissens, G., Verstraete, W., 2005b. Microbial fuel cells performances and perspectives. In: Biofuels for Fuel Cells: Renewable Energy from Biomass Fermentation. IWA Publishing, London. Rabaey, K., Ossieur, W., Verhaege, M., Verstraete, W., 2005c. Continuous microbial fuel cells convert carbohydrates to electricity. Water Sci. Technol. 52, 515e523. Rabaey, K., Read, S.T., Clauwaert, P., Freguia, S., Bond, P.L., Blackall, L.L., Keller, J., 2008. Cathodic oxygen reduction catalyzed by bacteria in microbial fuel cells. ISME J. 2, 519e527. Rahimnejad, M., Ghoreyshi, A.A., Najafpour, G., Jafary, T., 2011. Power generation from organic substrate in batch and continuous flow microbial fuel cell operations. Appl. Energy 88, 3999e4004. Reguera, G., McCarthy, K.D., Mehta, T., Nicoll, J.S., Tuominen, M.T., Lovley, D.R., 2005. Extracellular electron transfer via microbial nanowires. Nature 435, 1098e1101. Reguera, G., Nevin, K.P., Nicoll, J.S., Covalla, S.F., Woodard, T.L., Lovley, D.R., 2006. Biofilm and nanowire production leads to increased current in Geobacter sulfurreducens fuel cells. Appl. Environ. Microbiol. 72, 7345e7348. Reimers, C.E., Girguis, P., Stecher, H.A., Tender, L.M., Ryckekynck, N., Whaling, P., 2006. Microbial fuel cell energy from an ocean cold seep. Geobiology 4, 123e136. Rengasamy, K., Berchmans, S., 2012. Simultaneous degradation of bad wine and electricity generation with the aid of the coexisting biocatalysts Acetobacter aceti and Gluconobacter roseus. Bioresour. Technol. 104, 388e393. Rezaei, F., Richard, T.L., Logan, B.E., 2009. Analysis of chitin particle size on maximum power generation, power longevity, and Coulombic efficiency in solidesubstrate microbial fuel cells. J. Power Sources 192, 304e309. Rhoads, A., Beyenal, H., Lewandowski, Z., 2005. Microbial fuel cell using anaerobic respiration as an anodic reaction and biomineralized manganese as a cathodic reactant. Environ. Sci. Technol. 39, 4666e4671. Richter, H., McCarthy, K., Nevin, K.P., Johnson, J.P., Rotello, V.M., Lovley, D.R., 2008. Electricity generation by Geobacter sulfurreducens attached to gold electrodes. Langmuir 24, 4376e4379. Richter, L.V., Sandler, S.J., Weis, R.M., 2012. Two isoforms of Geobacter sulfurreducens PilA have distinct roles in pilus biogenesis, cytochrome localization, extracellular electron transfer, and biofilm formation. J. Bacteriol. 194, 2551e2563. Rinaldi, A., Mecheri, B., Garavaglia, V., Licoccia, S., Nardo, P.D., Traversa, E., 2008. Engineering materials and biology to boost performance of microbial fuel cells: a critical review. Energy Environ. Sci. 1, 417. Ringeisen, B.R., Henderson, E., Wu, P.K., Pietron, J., Ray, R., Little, B., Biffinger, J.C., Jones-Meehan, J.M., 2006. High power density from a miniature microbial fuel cell using Shewanella oneidensis DSP10. Environ. Sci. Technol. 40, 2629e2634. Roche, I., Scott, K., 2008. Carbon-supported manganese oxide nanoparticles as electrocatalysts for oxygen reduction reaction (orr) in neutral solution. J. Appl. Electrochem. 39, 197e204. Rosenbaum, M., Zhao, F., Schröder, U., Scholz, F., 2006. Interfacing electrocatalysis and biocatalysis with tungsten carbide: a high-
REFERENCES performance, noble-metal-free microbial fuel cell. Angew. Chem. Int. Edit. 118, 6810e6813. Rosenbaum, M., Aulenta, F., Villano, M., Angenent, L.T., 2011. Cathodes as electron donors for microbial metabolism: which extracellular electron transfer mechanisms are involved? Bioresour. Technol. 102, 324e333. Ross, D.E., Flynn, J.M., Baron, D.B., Gralnick, J.A., Bond, D.R., 2011. Towards electrosynthesis in Shewanella: energetics of reversing the mtr pathway for reductive metabolism. PLoS One 6, 16649. Ryckelynck, N., Stecher, H.A., Reimers, C.E., 2005. Understanding the anodic mechanism of a seafloor fuel cell: Interactions between geochemistry and microbial activity. Biogeochemistry 76, 113e139. Schröder, U., 2007. Anodic electron transfer mechanisms in microbial fuel cells and their energy efficiency. Phys. Chem. Chem. Phys. 9, 2619e2629. Schröder, U., Niessen, J., Scholz, F., 2003. A generation of microbial fuel cells with current outputs boosted by more than one order of magnitude. Angew. Chem. Int. Edit. 42, 2880e2883. Sharma, V., Kundu, P.P., 2010. Biocatalysts in microbial fuel cells. Enzyme Microb. Tech. 47, 179e188. Shea, C., Clauwaert, P., Verstraete, W., 2008. Adapting a denitrifying biocathode for perchlorate. Water Sci. Technol. 58, 1941e1946. Shi, L., Chen, B., Wang, Z., Elias, D.A., Mayer, M.U., Gorby, Y.A., Ni, S., Lower, B.H., Kennedy, D.W., Wunschel, D.S., Mottaz, H.M., Marshall, M.J., Hill, E.A., Beliaev, A.S., Zachara, J.M., Fredrickson, J.K., Squier, T.C., 2006. Isolation of a high-affinity functional protein complex between OmcA and MtrC: two outer membrane decaheme c-type cytochromes of Shewanella oneidensis MR-1. J. Bacteriol. 188, 4705e4714. Shi, L., Squier, T.C., Zachara, J.M., Fredrickson, J.K., 2007. Respiration of metal (hydr)oxides by Shewanella and Geobacter: a key role for multihaem c-type cytochromes. Mol. Microbiol. 65, 12e20. Shi, L., Rosso, K.M., Zachara, J.M., Fredrickson, J.K., 2012a. Mtr extracellular electron-transfer pathways in Fe (III)-reducing or Fe (II)-oxidizing bacteria: a genomic perspective. Biochem. Soc. Trans. 40, 1261e1267. Shi, L., Rosso, K.M., Clarke, T.A., Richardson, D.J., Zachara, J.M., Fredrickson, J.K., 2012b. Molecular underpinnings of Fe(III) oxide reduction by Shewanella oneidensis MR-1. Front. Microbiol. 3, 50. Srikanth, S., Reddy, M.V., Mohan, S.V., 2012. Microaerophilic microenvironment at biocathode enhances electrogenesis with simultaneous synthesis of polyhydroxyalkanoates (PHA) in bioelectrochemical system (BES). Bioresour. Technol. 125, 291e299. Srikanth, S., Venkata, M.S., 2012. Change in electrogenic activity of the microbial fuel cell (MFC) with the function of biocathode microenvironment as terminal electron accepting condition: influence on overpotentials and bio-electro kinetics. Bioresour. Technol. 119C, 241e251. Stamatelatou, K., Antonopoulou, G., Tremouli, A., Lyberatos, G., 2011. Production of gaseous biofuels and electricity from cheese whey. Ind. Eng. Chem. Res. 50, 639e644. Steinbusch, K.J., Hamelers, H.V., Schaap, J.D., Kampman, C., Buisman, C.J., 2010. Bioelectrochemical ethanol production through mediated acetate reduction by mixed cultures. Environ. Sci. Technol. 44, 513e517. Suh, S.J., Silo-Suh, L., Woods, D.E., Hassett, D.J., West, S.E., Ohman, D.E., 1999. Effect of rpoS mutation on the stress response and expression of virulence factors in Pseudomonas aeruginosa. J. Bacteriol. 181, 3890e3897. Summers, Z.M., Fogarty, H.E., Leang, C., Franks, A.E., Malvankar, N.S., Lovley, D.R., 2010. Direct exchange of electrons within aggregates of an evolved syntrophic coculture of anaerobic bacteria. Science 330, 1413e1415. 151 Sun, J., Hu, Y., Hou, B., 2011. Electrochemical characterization of the bioanode during simultaneous azo dye decolorization and bioelectricity generation in an air-cathode single chambered microbial fuel cell. Electrochim. Acta 56, 6874e6879. Tai, S.K., Wu, G., Yuan, S., Li, K.C., 2010. Genome-wide expression links the electron transfer pathway of Shewanella oneidensis to chemotaxis. BMC Genomics 11, 319. Tandukar, M., Huber, S.J., Onodera, T., Pavlostathis, S.G., 2009. Biological chromium(VI) reduction in the cathode of a microbial fuel cell. Environ. Sci. Technol. 43, 8159e8165. Thrash, J.C., Trump, J.I., Weber, K.A., Miller, E., Achenbach, L.A., Coates, J.D., 2007. Electrochemical stimulation of microbial perchlorate reduction. Environ. Sci. Technol. 41, 1740e1746. Thurston, C.F., Bennetto, H.P., Delaney, G.M., 1985. Glucose metabolism in a microbial fuel cell stoichiometry of product formation in a thionine-mediated Proteus vulgaris fuel cell and its relation to coulombic yields. J. Gen. Appl. Microbiol. 131, 1393e1404. Tong, M., Du, Z., Gu, T., 2012. Converting low-grade biomass to produce energy using bio-fuel cells. In: Zhou, Y. (Ed.), Eco- and Renewable Energy Materials. Springer, Berlin-New York, ISBN 978-3-642-33496-2, pp. 73e97. Torres, C.I., Marcus, A.K., Lee, H.S., Parameswaran, P., Brown, K.R., Rittmann, B.E., 2010. A kinetic perspective on extracellular electron transfer by anode-respiring bacteria. FEMS Microbiol. Rev. 34, 3e17. Totten, P.A., Lara, J.C., Lory, S., 1990. The rpoN gene product of Pseudomonas aeruginosa is required for expression of diverse genes, including the flagellin gene. J. Bacteriol. 172, 389e396. Tsujimura, S., Fujita, M., Tatsumi, H., Kano, K., Ikeda, T., 2001. Bioelectrocatalysis-based dihydrogen/dioxygen fuel cell operating at physiological pH. Phys. Chem. Chem. Phys. 3, 1331e1335. Va’zquez-Larios, A.L., Solorza-Feria, O., Va’zquez-Huerta, G., EsparzaGarcı’a, F., Rinderknecht-Seijas, N., Poggi-Varaldo, H.c.M., 2011. Effects of architectural changes and inoculum type on internal resistance of a microbial fuel cell designed for the treatment of leachates from the dark hydrogenogenic fermentation of organic solid wastes. Int. J. Hydrogen Energy 36, 6199e6209. Vega, C.A., Fernandez, I., 1987. Mediating effect of ferric chelate compounds in microbial fuel-cells with Lactobacillus plantarum, Streptococcus lactis, and Erwinia dissolvens. Bioelectrochemistry 17, 217e222. Velasquez, O.S.B., Head, I.M., Curtis, T.P., Scott, K., 2011. Factors affecting current production in microbial fuel cells using different industrial wastewaters. Bioresour. Technol. 102, 5105e5112. Velasquez-Orta, S.B., Curtis, T.P., Logan, B.E., 2009. Energy from algae using microbial fuel cells. Biotechnol. Bioeng. 103, 1068e1076. Velvizhi, G., Venkata, M.S., 2011. Biocatalyst behavior under selfinduced electrogenic microenvironment in comparison with anaerobic treatment: evaluation with pharmaceutical wastewater for multi-pollutant removal. Bioresour. Technol. 102, 10784e10793. Venkataraman, A., Rosenbaum, M., Arends, J.B.A., Halitschke, R., Angenent, L.T., 2010. Quorum sensing regulates electric current generation of Pseudomonas aeruginosa PA14 in bioelectrochemical systems. Electrochem. Commun. 12, 459e462. Villano, M., Aulenta, F., Ciucci, C., Ferri, T., Giuliano, A., Majone, M., 2010. Bioelectrochemical reduction of CO2 to CH4 via direct and indirect extracellular electron transfer by a hydrogenophilic methanogenic culture. Bioresour. Technol. 101, 3085e3090. Wen, Q., Kong, F., Zheng, H., Cao, D., Ren, Y., Yin, J., 2011. Electricity generation from synthetic penicillin wastewater in an air-cathode single chamber microbial fuel cell. Chem. Eng. J. 168, 572e576. Xing, D.F., Zuo, Y., Cheng, S., 2008. Electricity generation by Rhodopseudomonas palustris DX-1. Environ. Sci. Technol. 42, 4146e4151.
152 9. BIOELECTROCHEMISTRY OF MICROBIAL FUEL CELLS AND THEIR POTENTIAL APPLICATIONS IN BIOENERGY Xu, J., Sheng, G.P., Luo, H.W., Li, W.W., Wang, L.F., Yu, H.Q., 2012. Fouling of proton exchange membrane (PEM) deteriorates the performance of microbial fuel cell. Water Res. 46, 1817e1824. Yi, H., Nevin, K.P., Kim, B.C., Franks, A.E., Klimes, A., Tender, L.M., Lovley, D.R., 2009. Selection of a variant of Geobacter sulfurreducens with enhanced capacity for current production in microbial fuel cells. Biosens. Bioelectron. 24, 3498e3503. You, S., Zhao, Q., Zhang, J., Jiang, J., Zhao, S., 2006. A microbial fuel cell using permanganate as the cathodic electron acceptor. J. Power Sources 162, 1409e1415. Yu, Y.Y., Chen, H.L., Yong, Y.C., Kim, D.H., Song, H., 2011. Conductive artificial biofilm dramatically enhances bioelectricity production in Shewanella-inoculated microbial fuel cells. Chem. Commun. 47, 12825e12827. Yuan, Y., Chen, Q., Zhou, S., Zhuang, L., Hu, P., 2012. Improved electricity production from sewage sludge under alkaline conditions in an insert-type air-cathode microbial fuel cell. J. Chem. Technol. Biotechnol. 87, 80e86. Zechner, E.L., Lang, S., Schildbach, J.F., 2012. Assembly and mechanisms of bacterial type IV secretion machines. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 1073e1087. Zhang, Y., Angelidaki, I., 2012. Bioelectrode-based approach for enhancing nitrate and nitrite removal and electricity generation from eutrophic lakes. Water Res. 46, 6445e6453. Zhao, H., Li, X., Johnson, D.E., Mobley, H.L., 1999. Identification of protease and rpoN-associated genes of uropathogenic Proteus mirabilis by negative selection in a mouse model of ascending urinary tract infection. Microbiology 145 (Pt 1), 185e195. Zhao, F., Slade, R.C., Varcoe, J.R., 2009. Techniques for the study and development of microbial fuel cells: an electrochemical perspective. Chem. Soc. Rev. 38, 1926e1939. Zhou, M., Chi, M., Luo, J., He, H., Jin, T., 2011. An overview of electrode materials in microbial fuel cells. J. Power Sources 196, 4427e4435. Zhou, M., Jin, T., Wu, Z., Chi, M., Gu, T., 2012. Microbial fuel cells for bioenergy and bioproducts. In: Gopalakrishnan, K., Leeuwen, J.v., Brown, R. (Eds.), Sustainable Bioenergy and Bioproducts. Springer-Verlag, Berlin-New York, pp. 131e172. Zhou, M., Wang, H., Hassett, D.J., Gu, T., 2013. Recent advances in microbial fuel cells (MFCs) and microbial electrolysis cells (MECs) for wastewater treatment, bioenergy and bioproducts. J. Chem. Technol. Biotechnol. 88, 508e518. Zuo, Y., Maness, P.C., Logan, B.E., 2006. Electricity production from steam-exploded corn stover biomass. Energy Fuel 20, 1716e1721.
C H A P T E R 10 Second-Generation Biofuel from High-Efficiency Algal-Derived Biocrude Rhykka Connelly UT Algae Science and Technology Facility, University of Texas at Austin, Austin, TX, USA email: r.connelly@cem.utexas.edu O U T L I N E Introduction 153 Biodiesel 158 Microalgal Biofuel History 154 Production of Biodiesel from Microalgae 159 Microalgae Biomass/Biofuel ProductiondCultivation Comparison of Biodiesel to Petrodiesel 160 155 Bioethanol 161 Phototrophic Microalgae 155 Bioethanol Production Process 161 Heterotrophic Microalgae 155 Biomethane 164 Nutrients 156 Biohydrogen 165 Contamination 156 Biocrude 166 Mixing 156 Properties of Subcritical Water 166 Culture Techniques 156 Hydrothermal Catalytic Liquefaction 167 Open-Pond Culture 157 HTL Summary and Outlook 167 Photobioreactors 157 Conclusions 167 Processing Microalgal Biomass for Biofuels 158 References 168 Microalgal Biomass to Biofuels 158 INTRODUCTION First-generation, or conventional, biofuels are derived from sugars, starches, or vegetable oils from traditional agricultural crops and waste oils. Given firstgeneration biofuels’ impact on agricultural crop demand and prices, alternative feedstocks have been sought out. Microalgae have since been identified as a viable second-generation biofuels feedstock Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00010-3 (Figure 10.1). The advantages of using microalgae for biofuel production in comparison with other available feedstocks have been extensively reported. There are an estimated 100,000 microalgae species, each with specific properties that allow them to exist in nearly every environment on Earth, including arid climates that do not sustain most agricultural crops. Therefore, microalgal production systems need not displace other traditional land-based crops intended for human 153 Copyright Ó 2014 Elsevier B.V. All rights reserved.
154 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE FIGURE 10.1 The progression from first- to second-generation biofuels. (For color version of this figure, the reader is referred to the online version of this book.) or livestock consumption, which in turn greatly reduces the impact to the food distribution chain. Further, microalgae may be harvested multiple times a year, which greatly increases yearly production yields. The cultivation of microalgae for biofuels production can also be coupled with other beneficial production schemes to improve net income and positively address environmental concerns. Some possibilities currently being investigated include the following: strategies, microalgae intended for biofuel production can potentially revolutionize a large number of biotechnology areas concurrently, including pharmaceuticals, cosmetics, nutrition and food additives, aquaculture, and pollution prevention.  3 • Reclamation of nutrients such as NHþ 4 , NO3 , PO4 , and others from wastewater, which reduces costs associated with cultivating the algae and treating wastewater (Zhu, 2013; Batten, 2013). • Utilization of waste CO2 from industrial flue gases, which reduces greenhouse gas emissions while producing biofuel (González-López, 2012). • Cultivation and extraction of value-added metabolites within microalgae intended for biofuel production. In this scenario, the value-added metabolite is extracted prior to, or during, the biofuel production stream. Commercially relevant products include a large range of fine chemicals and bulk products, such as polyunsaturated omega fatty acids, antioxidants, high-value bioactive compounds, natural dyes, sugars, and proteins (Mimouni, 2012; Skjånes et al., 2013). • After oil and target metabolite extraction, the processed algal biomass can be used as a nutrient-rich livestock feed, or used as sustainable organic fertilizer due to its high N:P ratio (Mulbry, 2005; Stamey, 2012). Beginning in the 1950s, Golueke et al. (1957) conducted early work on the anaerobic digestion of microalgal biomass for the production of methane fuel. The energy crisis in 1973 prompted the formation of The National Renewable Energy Laboratory (NREL) under the Jimmy Carter Administration. From 1978 to 1996, NREL conducted the most authoritative study to date on the development of biofuels from algae (Sheehan et al., 1998). The study concluded that under controlled conditions, algae are capable of producing 40 times the amount of oil for biodiesel per unit area of land when compared to terrestrial oilseed crops such as soy and canola, and that the use of wastewater as a nutrient source for algae propagation was the most practical approach for near-term production of algal biodiesel (Sheehan et al., 1998; Oswald, 2003). Despite the promise of cost-effective fuel production from microalgae, interest in renewable energy quickly waned as the energy crisis subsided and fuel prices fell. The recent world-wide escalation in oil prices has renewed interest in microalgae as a biofuels feedstock. Since the original NREL study, other groups also have conducted analyses of full-scale algae-to-biodiesel production (Benemann et al., 1982; Weissman and Goebel, Because of this variety of value-added biological derivatives, coupled with environmental sustaining MICROALGAL BIOFUEL HISTORY
HETEROTROPHIC MICROALGAE 1987; Beal, 2012a). Although these and other studies have indicated a great potential for profitable biofuel from microalgae, they also highlighted the need for system improvements, in both cultivation management and processing schemes to improve yields and reduce costs in order to be competitive with fossil fuels. For example, even when robust algae growth was achieved, inefficient processing techniques such as biomass centrifugation and drying followed by solvent extraction made recovery of biofuels cost-prohibitive. To overcome this barrier, changes to the system have been introduced, including processing techniques that eliminate the need for expensive dewatering regimens such as centrifugation and drying of the harvested biomass prior to oil extraction with solvents. One suggested path forward is a solventless wet stream process whereby microalgae are concentrated using pH-driven flocculation using inexpensive lime, followed by rupturing of the cells by pulsed electric field, and ultimate recovery of released lipids by cross-flow filtration. When coupled with waste streams for CO2 and nutrients, this process has a positive return on investment (Beal, 2012b). Another suggested path forward toward practical biofuel extraction from microalgae is the use of hydrothermal liquefaction (HTL) processing. This method eliminates the need for solvents to break open algae cells, instead relying on heat and pressure to remove the water from the biomass. An ancillary benefit of the HTL method is that in addition to lipids, other organic metabolites such as carbohydrates, proteins, and nucleic acids can likewise be converted to biocrude during the HTL process. Thus, a cultivation strategy needs only to focus on the production of biomass rather than inducing the accumulation of lipids at the expense of cellular proliferation. Ultimately, cultivation and processing strategies should be firmly supported by realtime analysis of fuel precursors such as lipids that can be converted to biodiesel, carbohydrates that can be converted to bioethanol, and the organic biomass that can be converted to biocrude. Detailed analytical feedback is necessary to optimize growth conditions to maximize specific biofuel precursors. MICROALGAE BIOMASS/BIOFUEL PRODUCTIONdCULTIVATION The intended final biofuel product defines successful microalgae cultivation. If biodiesel is the final product, algal strains should be selected and cultured to produce maximal saturated fatty acids. If biocrude is the desired product, high organic content, or a simple abundance of biomass, is required. Whatever the target product, successful cultivation requires specific environmental conditions to drive the production of specific fuel precursors. Major parameters that influence biomass production 155 include adequate light (wavelength and intensity), temperature, CO2 concentration, nutrient composition, salinity, contaminants, and mixing conditions. PHOTOTROPHIC MICROALGAE Phototrophic microalgae use carbon dioxide (carbon source), sunlight (energy source), and nutrients to proliferate. Two properties of light energy are important for algal growth and metabolism: quality of the light spectrum and quantity of the light photons. As phototrophs, light-harvesting pigments (chlorophyll and carotenoids) absorb light at specific wavelengths to drive the photosynthetic process. Light absorption, however, is hindered both by light scattering through increasing depths of the culture medium and by mutual shading as the culture increases in density. Antenna structures of microalgae are excessively efficient at harvesting light energy, absorbing all the photons that hit them even though only a fraction of those photons are used for photosynthesis. This deprives nearby algae from absorbing photons and consequently leads to low productivity. Aggressive mixing of the culture mitigates some of these effects, but cannot completely overcome the light penetration limitations inherent in a photosynthetic system. HETEROTROPHIC MICROALGAE Several wild-type and genetically modified species of microalgae have been reported capable of growing phototrophically, heterotrophically or both (mixotrophically). Unlike phototrophic algae that require light energy, heterotrophic algae have no such requirement. Instead, these algae utilize organic carbons supplied in the media to drive cellular proliferation and lipid accumulation. Without the limitations imposed by inefficient light harvesting due to mutual shading and light scattering in the medium, the densities of heterotrophic cultures can far exceed the densities achieved in phototrophic systems. Increased densities can translate to higher biofuel precursor yields. For example, when Chlorella protothecoides was grown heterotrophically using an organic carbon source, oil accumulation far exceeded that seen in corresponding autotrophic cells (Miao and Wu, 2004). Hence, heterotrophic production has several advantages over phototrophic systems including increased densities that eliminate the need for dewatering, and increased process control that facilitates the maintenance and rapid growth of monocultures and the creation of a consistent product. The primary limitation for commercial-scale heterotrophic production of biofuel oils in microalgae is the cost of
156 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE the organic carbon source. Sugars such as glucose and acetate have been utilized as the primary carbon source at the bench scale, but become cost-prohibitive at production scale. It is therefore unsurprising that increased efforts to identify microalgal species that can thrive on waste sugars, such as bagasse or cellulosic waste, are underway. NUTRIENTS To maximize biomass production and the accumulation of fuel precursors, algal cultures must be supplied with various concentrations of macronutrients, vitamins, and trace elements depending on species requirements. While there are limited reports on optimal levels of nutrients required for mass algal cultures, it is generally accepted that required macronutrients are nitrogen and phosphorus (Brzezinski, 1985; Harrison and Berges, 2005). Trace elements such as cobalt, copper, molybdenum, zinc, and nickel are likewise necessary, and in some species are considered to be effective in hydrogen production (Ramachandran and Mitsui, 1984). There appears to be no consensus on the optimal ratios for these nutrients, even for specific species grown successively in the same system. Therefore, nutrients are often added in excess to avoid nutrient limitations (Richmond, 1999; Sanchez et al., 1999; Acien Fernandez et al., 2001). One strategy to reduce costs associated with adding excess nutrients involves culturing microalgae in reclaimed water or wastewater blends. The use of algae to absorb nutrients in the wastewater processing stream has been widely employed by water treatment facilities (Megharaj et al., 1992; Tredici et al., 1992; Nurdogan and Oswald, 1995; Kaya and Picard, 1995; Craggs et al., 1995). The green microalga Scenedesmus obliquus has demonstrated vitality in urban wastewaters, registering growth rates similar to those reported for a complete synthetic medium. This freshwater alga tolerates a wide range of temperature and pH, making it versatile for water purification (Kessler, 1991). Similar findings for other algal species continue to emerge, along with the energy return on investment analyses that confirm the utility of coupling scaled algal (EROI) production with nutrient reclamation from waste streams, resulting in decreased costs for both algal growth and water treatment (Beal, 2012b). CONTAMINATION Another barrier to the large-scale production of algae biofuels is the maintenance of axenic or nearly axenic cultures. In particular, cultivation systems that are open to the environment (e.g. open ponds) are easily susceptible to contamination by unwanted species if extreme care is not taken. A new open pond is typically inoculated with the desired strain of microalgae with the hope that the algae will aggressively proliferate and dominate the pond flora. Over time, it is likely that undesired species will be introduced, which may graze on the algae or outcompete the inoculated species and lead to severely reduced yields. Once a competitor has taken residence in a pond, it is extremely difficult to eradicate (Schenk et al., 2008). It is therefore crucial to aggressively monitor cultures to identify and eradicate contaminates as soon as possible. A number of strategies have been employed to minimize culture contaminations. Cultivating algal extremophiles that tolerate and outcompete invasive species in particular environments (e.g. pH and salinity) facilitates open-pond production. High bicarbonate concentrations allow Spirulina to be grown in open ponds with few invasive algae, and high-saline environments allow marine algae like Dunaliella salina to be grown in “relative pure cultures” (Anderson, 2005). Another popular strategy involves shortening the longevity of the culture; cultures are scaled and harvested before major contamination can occur (Benemann, 2008). Cultivation of microalgae in closed photobioreactors (PBRs) offers another level of protection against predators. Occasionally, cultures can be treated with antibiotics and antifungals to eliminate bacteria and fungi, but this practice can lead to microbial resistance and render the treatment ineffective. Predator ciliates can be treated with dioctyl sulfosuccinate, which is used to eliminate ciliates in the udders of milking cows (Abou Akkada, 1968) with minimal harm done to the algae. MIXING At high algae concentrations, a thin top layer of cells absorbs nearly all lightdthis phenomenon can be avoided by proper mixing. Mixing must sufficiently keep algae cells in suspension, aid distribution of CO2 and O2, and provide uniform exposure of light to all cells. Mixing also decreases the boundary layer around cells, which facilitates increased uptake of metabolic products (Molina Grima et al., 1999). CULTURE TECHNIQUES The choice of cultivation systems is an important aspect that significantly affects the efficiency and costeffectiveness of a microalgal biofuel production process (Lee, 2001; Pulz, 2001; Carvalho et al., 2006). A wide variation exists among the microalgal cultivation systems for the production of biomass. Raceways, PBRs, and fermenters, which are the three most widely used microalgae culture systems, will be discussed below.
157 PHOTOBIOREACTORS OPEN-POND CULTURE Large-scale cultivation of microalgae in outdoor open-pond systems is well documented (Benemann and Oswald, 1996; Borowitzka, 2005). Open ponds most closely resemble the natural milieu of microalgae. Indeed, ponds can be natural bodies of water, excavated ditches that are unlined or lined with impermeable materials, or they can be constructed above ground with walls (Figure 10.2). Despite a certain variability in shape, the most common technical design for open-pond systems is raceway cultivators driven by paddle wheels and usually operating at water depths of 15e20 cm (Figure 10.1). At these water depths, biomass concentrations of up to 1000 mg/l and productivities of 60e100 mg/(l/day), i.e. 10e25 g/(m2/day) are possible. Similar in design are the circular ponds commonly seen in Asia and the Ukraine (Becker, 2007). Such circular ponds usually have the provision of a centrally located rotating arm (similar to those used in wastewater treatment) for mixing and may have productivities ranging between 8.5 and 21 g/m2 day (Benemann and Oswald, 1996). On the other hand, thin-layer, inclined ponds consist of slightly inclined shallow trays and may achieve productivities up to 31 g/m2 day (Doucha and Livansky, 2006). Because these ponds are open to the environment, they are most suitable for algal species that can tolerate extreme environmental conditions (salinity, pH, nutrient loads, etc.) to the exclusion of invasive species. Such algal species include fast growers such as Chlorella, Spirulina, and Dunaliella, which thrive in highly alkaline or saline environments (Chisti, 2007). Limitations to successful scale-up of microalgae in open-pond systems include contamination, evaporation, limited species suitability, low-volumetric productivities, and the need for large land area. PHOTOBIOREACTORS The problems associated with open systems have encouraged the development of closed system PBRs. PBRs can be located indoors under supplemental illumination or outdoors utilizing natural sunlight. Various types of PBRs have been designed depending on growers’ needs; these include tubular PBRs, vertical bubble columns and airlift reactors, combined bubble column and inclined tubular reactors, helical PBRs, and flat-plate PBRs (Tredici and Zittelli, 1998; Sanchez et al., 1999; Berzin, 2005; Ugwu et al., 2005) (Figure 10.3). Closed PBRs allow for tighter regulation and control of nearly all the biotechnologically important parameters FIGURE 10.2 (a) Open-pond production systems at Seambiotic in TelAviv, Israel and (b) Cyantotech in Kona, Hawaii. (For color version of this figure, the reader is referred to the online version of this book.) FIGURE 10.3 (a) Horizontal photobioreactors used in the biomass production plant in Klötze, Saxony-Anhalt (ÓBioprodukte Prof. Steinberg GmbH) and (b) vertical photobioreactors used at the University of Texas. (For color version of this figure, the reader is referred to the online version of this book.)
158 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE and confer the following fundamental benefits: a reduced contamination risk, reduced CO2 losses, reproducible cultivation conditions, controllable hydrodynamics, and temperature (Pulz, 1992). However, widespread implementation has been hampered by the high capital costs associated with PBRs. PROCESSING MICROALGAL BIOMASS FOR BIOFUELS There are several methods to process microalgae into biofuel products. Figure 10.4 shows some of the more common approaches to (1) harvest/dewater microalgae, (2) release fuel precursors by compromising the integrity of the algae, followed by (3) conversion of fuel precursors to biofuel products. Many algal species can be preconcentrated by simply allowing unmixed cells to settle by gravity. Additional concentration can be achieved by flocculation, centrifugation, microfiltration, and drying. Freshly clarified media can be recycled back to the growth environment, although there are limited data regarding the number of times growth media can be recycled. Concentrated wet algal cells may be subsequently compromised by passage through a pulsed electric field, mechanical bead-milling, sonication, or enzymatic degradation (Beal, 2012a). Solvent extraction of the biomass generally requires that the biomass is dried as an initial step. Once fuel precursors are exposed, they may be converted as fuel products by specific fuel conversion approaches. These methods are discussed in detail below. FIGURE 10.4 MICROALGAL BIOMASS TO BIOFUELS Microalgae can provide several different types of renewable biofuels, and numerous options exist for the conversion of components of microalgal biomass to biofuel. These include methane produced by anaerobic digestion of the algal biomass (Spolaore et al., 2006); biodiesel derived from microalgal oil (Roessler et al., 1990; Sawayama et al., 1995; Dunahay et al., 1996; Sheehan et al., 1998; Banerjee et al., 2002; Gavrilescu and Chisti, 2005); biohydrogen (Ghirardi et al., 2000; Akkerman et al., 2002; Melis, 2002; Fedorov et al., 2005; Kapdan and Kargi, 2006); and biocrude derived from organics comprising microalgae. An important distinction to note is whether extracted compounds, whole biomass, or both will be converted to biofuel. Microalgae that contain high-value bioproducts (e.g. carotenoids, sulfated polysaccharides, and phycobilliproteins) may undergo a two-phase extraction scheme where the value-added product is fractionated from the biofuel production stream prior to conversion of lipids to biodiesel and carbohydrates to bioethanol. Alternatively, the remaining organic fraction of the biomass can be converted to biocrude by HTL. BIODIESEL Biodiesel is derived from fatty acyl lipids from plant and animal sources. Table 10.1 shows the average oil yield per hectare from various crops. Using the average oil yield per acre, the footprint needed to meet 50% of the U.S. transport fuel needs is calculated. For example, Microalgal biomass-to-biofuel processing pathway choices. (For color version of this figure, the reader is referred to the online version of this book.)
159 PRODUCTION OF BIODIESEL FROM MICROALGAE TABLE 10.1 Comparison of Biodiesel Feedstocks: Oil Yields vs Land Area Necessary to Meet 50% of Current Transportation Fuel Demand Crop Oil yield (l/ha) Land Area Needed (M ha) Percentage of Existing US Agricultural Area Corn 172 1540 846 Soybean 446 594 326 Canola 1190 223 122 Jatropha 1892 140 77 Coconut 2689 99 54 Oil palm 5950 45 24 Microalgae (30% oil by wt) 58,700 4.5 2.5 Microalgae (70% oil by wt) 136,900 2 1.1 Source: Christi, 2007. the high-yielding crop oil palm requires a 45 Mha cropping area, or 24% of the existing agricultural footprint in the US to meet only 50% of the current transport fuel needs. Given the large agricultural footprint required, it is clear that land-based oilseed crops cannot realistically satisfy current demand. Lipid-rich microalgae, however, hold more promise as a sustainable feedstock that can significantly contribute toward demand. Under controlled conditions, the footprint required to produce an order of magnitude higher oil yields requires an order of magnitude smaller cropping area compared to oil palm, assuming an oil content of 30% in the microalgae. A caveat to these numbers is that the microalgal oil yields given in Table 10.1 are based on experimentally demonstrated biomass productivity in PBRs. Demonstrated biodiesel yields on a larger scale have been much smaller. The large-scale cultivation of lipid-rich microalgae remains a significant challenge in the algae biofuel industry and thus still under intense investigation. PRODUCTION OF BIODIESEL FROM MICROALGAE Biodiesel is derived from plant and animal lipids. Lipids are subdivided in two main classes based on their chemical characteristics: polar and nonpolar (neutral) lipids. Neutral lipids include the tri- and diglycerides, waxes, and isoprenoid-type lipids. Monoglycerides divide neutral lipids from polar lipids. Polar lipids include phospholipids (e.g. phosphatidylinositol and phosphatidylethanolamine), free fatty acids, and glycerol. Desirable feedstocks for biodiesel production are composed of a higher proportion of saturated fatty acyl neutral, rather than polar lipids. Compared to animal fats and other seed-based oils, many microalgal species have been reported to contain a relatively greater proportion of polar lipids to neutral lipids (triglycerides) and the predominance of long-chain polyunsaturated fatty acids (greater than C18). However, several species of microalgae have been shown to produce various lipids, hydrocarbons, and other complex oils suitable for biodiesel production (Banerjee et al., 2002; Guschina and Harwood, 2006). To accurately predict yields from microalgae, it is critical to understand the lipid composition of the feedstock. The fluorescence probe Nile Red is often used to monitor neutral lipid composition within microalgae. However, Nile Red cannot provide information regarding carbon chain length or saturation of fatty acids. Gas chromatography is often utilized for the identification of specific fatty acids and the separation, identification and quantification of specific lipid classes by High-performance liquid chromatographyeevaporative light scattering detection (HPLC-ELSD) has recently been described (Jones et al., 2012). An informed realtime understanding of the lipid composition of the culture may lead to better cultivation practices, which can drive the accumulation of desirable lipids and ultimately higher biodiesel yields. The oil to biodiesel conversion process is termed transesterification (Figure 10.5). During transesterification, an alcohol (e.g. methanol and ethanol) is reacted with vegetable oil (fatty acid) in the presence of catalyst. Catalysts include alkalis (e.g. KOH and NaOH) or acids (e.g. H2SO4) to produce fatty acid methyl esters (FAME) or fatty acid ethyl esters and glycerol. Generally, methanol is preferred for transesterification because it is less expensive than ethanol. Transesterification requires 3 mol of alcohol for every 1 mol of triglyceride to produce 1 mol of glycerol and 3 mol of methyl esters. This
160 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE FIGURE 10.5 The transesterification of triglyceride to 3 mol each of fatty acid methyl esters and glycerol. reaction is reversible in nature and eventually arrives at equilibrium (Fukuda et al., 2001). The produced biodiesel is immiscible and thus easily separated from glycerol by phase partitioning the biodiesel in a nonpolar solvent such as hexane or heptane. The solvent is later recovered by distillation. Transesterification is an inexpensive way of transforming the large, branched molecular structure of the vegetable oils into smaller, straight-chain molecules of the type required in regular diesel combustion engines. Using microalgae as a feedstock, biodiesel can be produced from extracted algal oils or by direct conversion of the biomass. The production of biodiesel from extracted microalgal oil proceeds as described above. For direct conversion of the biomass to biodiesel, the microalgae are first concentrated to a paste-like consistency. The cells are then incubated in methanol or ethanol in the presence of a strong acid or base at an elevated temperature. In this process, fatty acids derived from not only triglycerides but also diglycerides and free fatty acids are transesterified to biodiesel. The remaining residue contains starch and proteins, which can further be processed into ethanol, animal feed, or used as a feedstock in an anaerobic fermenter. TABLE 10.2 COMPARISON OF BIODIESEL TO PETRODIESEL Biodiesel is a proven fuel. The conversion of vegetable oil to biodiesel was first described as early as 1853 by Patrick Duffy, many years before the first diesel engine became functional (Duffy, 1853). Rudolf Diesel’s engine was built several years later, running for the first time on August 10, 1893 using nothing but peanut oil feedstock. In a 1912 speech, Diesel said, “the use of vegetable oils for engine fuels may seem insignificant today but such oils may become, in the course of time, as important as petroleum and the coal-tar products of the present time.” Fossil fuel-derived petrodiesel is produced from the fractional distillation of fossil fuel crude oil. It contains w75% saturated hydrocarbons and 25% aromatic hydrocarbons (including naphthalenes and alkylbenzenes). Compared to petrodiesel, biodiesel molecules are comprised almost entirely FAME saturated, or monosaturated, hydrocarbons and w5% aromatic compounds. Table 10.2 shows a comparison between the properties of biodiesel to petrodiesel. Biodiesel has a higher lubricity and thus better lubricating properties Fuel Properties of Biodiesel and Petrodiesel Property Biodiesel Petrodiesel Production process Chemical reaction Reaction þ fractionation Cetane number 51e62 44e49 Oxygen 10e12% free oxygen Very low Aromatics 5% 18e22% Sulfur None 0.05%  Flash point 300e400 F 125  F Lubricity Much greater than diesel. Comparable to oil lubricants Low-sulphur fuel has low lubricity factor Biodegradability Biodegrades readily Poor biodegradability Toxicity Essentially nontoxic Highly toxic
BIOETHANOL PRODUCTION PROCESS than fossil diesel, which reduces wear on fuel systems and engine components. Biodiesel likewise has higher cetane ratings than today’s lower sulfur diesel fuels. The cetane number is a measure of a fuel’s ignition delay, or the time period between the start of injection and the first identifiable pressure increase during combustion of the fuel; the higher the cetane number the more easily the fuel will combust. Therefore higher cetane biodiesel should cause an engine to run more smoothly and quietly. Biodiesel’s higher flash point makes biofuel vehicles much safer in accidents than those powered by petrodiesel or gasoline. Biodiesel is biodegradable and nontoxic and also contains little to no sulfur, which makes it a much cleaner burning fuel compared to petrodiesel (Hai et al., 2000; Anderson et al., 2002; Hoekema et al., 2002; Choi et al., 2003; Grima et al., 2003; Zijffers et al., 2008; Brindley et al., 2011). Biodiesel has higher oxygen content than petrodiesel, which can also reduce pollution emissions. However, this benefit is offset by the fact that biodiesel is more likely to oxidize (react with oxygen), producing contaminants (gumming/sludge) that will plug fuel filters, leave deposits on injectors and cause injector pump problems. Further, continuous oxidization leads to the fuel becoming more acidic, which in turn causes corrosion on the components in the injection system. It will also dissolve fossil-diesel sludge built up over time and send it through fuel lines, plugging fuel filters. Biodiesel cloud or gel point is higher than pump diesel, meaning that it tends to gel at low temperatures more readily which can lead to poor cold starting. Clearly, there are both benefits and drawbacks for using biodiesel in today’s automobile engines. BIOETHANOL First-generation bioethanol is usually produced by alcoholic fermentation of starch (e.g. corn and wheat) or sugar (e.g. sugarcane, sugar beet and sweet sorghum). Second-generation bioethanol feedstocks include lignocellulosic grasses, woody biomass, and algae. Bioethanol is an already well-established fuel in Brazil and the USA (Goldemberg, 2007). Owing to mandates enacted by the Brazilian government in 1976, all light-duty fleet vehicles are required to operate using a blend of gasoline and bioethanol fluctuating between 10% and 25%, or E10eE25. In 2003, the Brazilian car manufacturing industry introduced flexible-fuel vehicles that can run on any proportion of gasoline (E20eE25 blend) and hydrous ethanol (E100) (Horta Nogueira, 2004). Sales reached an impressive 92.3% share of all new cars and light-vehicle sales for 2009, and overall bioethanol production reached 5.5 billion U.S. liquid gallons. 161 Although the vast majority of bioethanol is produced by fermentation of corn glucose in the United States or sugarcane sucrose in Brazil (Rosillo-Calle and Cortez, 1998), bioethanol can be derived from any material that contains sugars, including microalgae. Unlike land-based food crops, the production of bioethanol from microalgae does not divert agricultural foods away from grocer’s shelves. This is especially true for corn and corn products, which serve as base ingredients of many processed foods. Further, microalgae can be cultivated in areas nonsuitable for traditional agricultural crops and can be harvested many times a year. Therefore, in the U.S., microalgae are generally thought to be the only practical alternative to current bioethanol crops such as corn and soybean (Chisti, 2007; Hu et al., 2008; Singh and Gu, 2010). Matsumoto et al. (2003) screened several strains of marine microalgae with high-carbohydrate content and identified a total of 76 strains with a carbohydrate content ranging from 33% to 53% . It has been estimated that approximately 46e140 kl of ethanol/ha year can be produced from microalgae (Mussatto, 2010). This yield is several orders of magnitude higher than yields obtained from other bioethanol feedstocks (Table 10.3). BIOETHANOL PRODUCTION PROCESS Monomeric sugars can be converted to ethanol directly, while starches and cellulose first must be hydrolyzed to fermentable sugars either enzymatically or chemically (Bashir and Lee, 1994). Like most biofuels processes, bioethanol production from microalgae begins with the concentration of algae. The algae are then further dried and ground to a powder. In the next step of the process, the algae mass is hydrolyzed and Saccharomyces cerevisiae yeast is added to the biomass to begin the fermentation process. The resulting fermented mash contains about 11e15% ethanol by volume as well as the nonfermentable solids from algae and yeast cells. Ethanol is then distilled off the mash at w96% strength. Despite widespread knowledge of this fermentation process, the details of the conversion process of algal celluloses-to-bioethanol are only partially understood. Celluloses comprise a large fraction of algal cells walls. These molecules are tightly packed and enzymatic access is often limited without a pretreatment step (Figure 10.6). Many authors have reported that it is essential to introduce a pretreatment stage to release and convert the complex carbohydrates entrapped in the cell wall into simple sugars necessary for yeast fermentation. Cellulose can be made more accessible by the addition of an acid (Figure 10.7). Arantes and Saddler (2010) have suggested a model where prior to hydrolysis of cellulose to
162 TABLE 10.3 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE Comparison of Bioethanol Feedstocks Feedstock Productivity (dry mg/ha year) %Fermentable Carbohydrate Corn 7* 80{{ Switchgrass {{ 3.6e15* Woody biomass 70e85 xx 15{{ 5.6 1.05 2.8e11.5 0.4e1.8 7e18.7 4e7.7 12 {{ 10e22 Lignin Productivity (dry mg/ha year) {{ 76.4 x %Lignin Fermentable Carbohydrate Productivity (dry mg/ha year) { {{ 25e35 { Chlorella sp. 127.8e262.8 33.4 0 42.7e87.8 0 Tetraselmis suecia 38*e139.4** 11e47* 0* 4.2e65.5 0 Arthrospira sp. 27e70* 15e50* 0* 4.1e35 0 * Dismukes et al., 2008. x Ragauskas et al., 2006. { Kristensen, 1990. ** Zittelli et al., 1999. xx Chisti, 2007. {{ Sanchez et al., 1999. FIGURE 10.6 The microalgal bioethanol production process. (For color version of this figure, the reader is referred to the online version of this book.) FIGURE 10.7 Acid-driven hydrolysis of cellulose.
BIOETHANOL PRODUCTION PROCESS 163 FIGURE 10.8 Theoretical breakdown of cellulose into monomeric units of glucose. Source: Arantes and Saddler (2010). (For color version of this figure, the reader is referred to the online version of this book.) monomeric units, cellulases must adsorb onto the surface of the insoluble cellulose (Figure 10.8). The action of the cellulases serves to loosen tightly packed fibrous cellulosic networks and create additional access to cellulose chains buried within the fibrils. Then the synergistic action of exo- and endoglucanases cleave accessible molecules to form soluble cello-oligosaccharides, or oligomers of 6 sugar units. These oligosaccharides are quickly hydrolyzed to primarily cellobiose, or two glucose molecules linked by a b (1/4) bond. Cellobiose hydrolyzation to glucose monomers is usually completed by the extraneous addition of b-glucosidase. Once glucose monomers have been rendered, bioethanol from microalgal biomass can be produced through two distinct pathways: direct dark fermentation or yeast fermentation of saccharified biomass. Whereas direct dark fermentation yields are typically much lower, the yeast fermentation process is a very wellestablished, relatively high-yield, low-energy-intensive process. Because microalgae can be harvested multiple times a year, some species have been shown to theoretically yield an order of magnitude more bioethanol compared to a land-based crop such as corn (Table 10.3). Further, using microalgae as a raw material is strongly advantageous as algae sugars may be derived from multiple sourcesdfrom intracellular starches and from the cellulosic cell wall. Nevertheless, to achieve higher yields, it is still necessary to screen for high starch-producing algal strains coupled with identifying mechanisms and culture conditions for inducing maximal accumulation of intracellular starches. In comparison to terrestrial feedstocks that contain lignin, certain species of microalgae and cyanobacteria have high potentiality for bioethanol production due to their high productivity rates, high biomass fermentable carbohydrate content, and lack of lignin. Lignin is a recalcitrant substance (i.e. not easily degraded) present in the cell walls of terrestrial biomass that cannot be converted to bioethanoldits processing is a major impediment for bioethanol production (Ragauskas et al., 2006). Microalgae’s potential can be highlighted by the fact that 75% of algal complex carbohydrates can be
164 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE hydrolyzed into a fermentable hexose monomer, and the fermentation yield of bioethanol is w80% of the theoretical optimal value (Huntley and Redalje, 2007). Harun et al. (2009) have shown that the blue-green Chlorococum sp. produces a maximum bioethanol concentration of 3.83 g/l obtained from 10 g/l samples that are preextracted for lipids versus those that remain as dried intact cells. This indicates that microalgae can be used for the production of both lipid-based biofuels and ethanol biofuels from the same biomass as a means to increase their overall economic value (Jones and Mayfield, 2012). The microalgae Chlorella vulgaris and Porphyridium sp., particularly, have been considered as promising feedstocks for bioethanol production because they can accumulate up to 37% and 54% (dry weight) of starch, respectively. The potential for simple, low-cost methods of bioethanol production from microalgae and cyanobacteria are real. The next phase of biofuel research should develop improved methodologies to increase intracellular ethanol production efficiencies. BIOMETHANE Biomethane (CH4) production from microalgal biomass is of interest because the efficiency of algal biomass production per hectare is estimated to be 5e30 times greater than that of the terrestrial crop plants (Sheehan et al., 1998). Golueke and Oswald (1959) published one of the first feasibility studies using microalgae for CH4 production and concluded that the process was feasible (Golueke and Oswald, 1959). There are two well-established methods of CH4 production: (1) harvest of an algal polyculture from a wastewater treatment pond, or (2) axenic growth of specific algae at a bench scale (Asinari Di San Marzano et al., 1982; Yen and Brune, 2007). The digestion process is described in Figure 10.9. It begins with bacterial hydrolysis of the algal biomass. Organic polymers, such as lipids, carbohydrates, and proteins, are first broken down to soluble derivatives, which are further fractionated into carbon dioxide, hydrogen, ammonia, and organic acids by acidogenic bacteria. Acetogenic bacteria then convert these resulting organic acids into acetic acid, along with additional ammonia, hydrogen, and carbon dioxide. Finally, methanogens convert these products to methane and carbon dioxide. Regardless of operating conditions and species, the proportion of methane in the biogas produced for the majority of studies falls in the range 69e75%. Anaerobic digestion is an effective process for biological oxygen demand removal, but it is not effective for nutrient removal. Thus, there is a need for further treatment of effluent from anaerobic digesters before it can be discharged into the environment. The nutrient-rich digestate also produced can be used as fertilizer. This process of converting microalgae to CH4 is dependent on several key metrics, namely (1) pH, (2) retention time, (3) mixing, (4) composition of the biomass and (5) composition of the surrounding milieu. One of the most important factors influencing CH4 biogas production from algal biomass has been reported to be pH. At high pH, due to high alkalinity from NH3 release, the gas production will shift toward CH4. The oxidation state of the biomass also affects biogas quality, which in turn drives the proportion of methane released (Sialve et al., 2009). Due to lowered content of sulfated amino acids, the microalgal biomass digestion releases a lower amount of hydrogen sulfide than do other types of organic substrates (Becker, 1988). The composition of the microalgal feedstock also affects biomethane yields. The relatively high lipid, starch, and protein contents and the absence of lignin make microalgae an ideal candidate for efficient biomethane production via fermentation in biogas plants. Theoretically, higher FIGURE 10.9 Anaerobic digestion process of microalgae. (For color version of this figure, the reader is referred to the online version of this book.)
165 BIOHYDROGEN Hydrogen is seen as one of the most promising fuels for the future owing to the fact that it is renewable and liberates large amounts of energy per unit weight without evolving CO2 when combusted. Biohydrogen production has several advantages over hydrogen production by photoelectrochemical or thermochemical processes. For example, whereas electrochemical hydrogen production requires the use of solar batteries with high energy requirements to split water and form the hydrogen product, biohydrogen production by photosynthetic microorganisms only requires simple PBRs with low energy requirements. A select group of green algae (including Chlamydomonas reinhardtii) and cyanobacteria offer an alternative route to renewable H2 production (Levin et al., 2004; Sakurai and Masukawa, 2007). Cyanobacteria are able to diverge the electrons emerging from the two primary reactions of oxygenic photosynthesis directly into the production of H2, making them attractive for the production of renewable H2 from solar energy and water. Cyanobacteria utilize two enzymatic pathways for H2 production, either nitrogenases or bidirectional hydrogenases (Angermayr et al., 2009). Nitrogenases require ATP, whereas bidirectional hydrogenases do not require ATP for H2 production, hence making them more efficient and favorable for H2 production with a much higher turnover. The fundamental aspects of cyanobacterial hydrogenases, and their more applied potential use as future producers of renewable H2 from sun and water, are receiving increased international attention. At the same time, significant progress is being made in the understanding of the molecular regulation of the genes encoding both the enzymes and the accessory proteins H2 O/2Hþ þ 2e þ 1 2 O2 2Hþ þ 2e /H2 H2 Combustion H2 þ 1 2 O2 /H2 O þ 285:8 kJ=mol = BIOHYDROGEN needed for the correct assembly of an active hydrogenase. With the increasing interest of both scientific and public communities in clean and renewable energy sources, and consequent funding opportunities, rapid progress will likely be made in the fundamental understanding of the regulation of cyanobacterial hydrogenases at both genetic and proteomic levels. Bandyopadhyay et al. (2010) have described Cyanothece sp. ATCC 51142, a unicellular, diazotrophic cyanobacterium with capacity to generate high levels of hydrogen under aerobic conditions. Wild-type Cyanothece sp. 51142 can produce hydrogen at rates as high as 465 mmol/mg of chlorophyll/h in the presence of glycerol. Authors also report that hydrogen production in this strain is mediated by an efficient nitrogenase system, which can be manipulated to convert solar energy into hydrogen at rates that are several fold higher, compared to other previously described wild-type hydrogen-producing photosynthetic microbes. These strains have evolved the ability to use solar energy to produce H2 from water (Esquı́vel, 2011; Levin et al., 1961). The theoretical conversion efficiency from light to H2 is calculated to be as high as w10% (Levin et al., 1961). Photosystem II (PSII) drives the first stage of the process (Figure 10.10), by splitting H2O into protons (H2), electrons (e), and O2. H2 Production = cellular lipid contents will result in higher methane yields. Thus lipid-rich microalgae make attractive substrates for anaerobic digestion, as they have a higher gas production potential when compared to carbohydrates and proteins (Li et al., 2002; Cirne et al., 2007). The hydraulic and solid retention time is another key metric in the anaerobic process. The hydraulic and solid retention time is a measure of the average length of the time that a soluble compound remains in a constructed bioreactor. Retention times should be sufficiently high to allow active bacterial populations (e.g. methanogens) to remain in the reactor yet not limit hydrolysis, which is considered to be the rate-limiting step in the overall conversion of complex substrates to methane. Moreover, optimal loading rates and hydraulic retention times must be enhanced to ensure efficient conversion of organic matter, and will depend on algal substrate composition and accessibility. Normally, the photosynthetic light reactions and the Calvin cycle produce carbohydrates that fuel mitochondrial respiration and cell growth. Under anaerobic conditions, however, mitochondrial oxidative phosphorylation is largely inhibited, which leads some organisms (e.g. Chlamydomonas reinhardtii) to reroute the energy stored in FIGURE 10.10 Biohydrogen production by microalgal respiration. (For color version of this figure, the reader is referred to the online version of this book.)
166 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE carbohydrates to a chloroplast hydrogenase (HydA), likely using an NAD(P)HPQ e transfer mechanism, to facilitate ATP production via photophosphorylation. Thus, hydrogenase reacts with Hþ (from the medium) and e (from reduced ferredoxin) to produce H2 gas that is subsequently excreted from the cell. The combustion of the recovered H2 yields only heat and H2O and thus is a model green technology. Several renewable energy laboratories have concluded that production efficiencies must be improved from 0.2% photon to H2 conversion efficiency at 20 W/m2 illumination to w7e10% at 230 W/m2 illumination (day light) to make the process economically viable. Through extensive preliminary work, the efficiency of this process has been enhanced to w1.0% from light to H2 and 2% to biomass. The H2 gas produced in such mutants has a purity of w90e95% and typical yields are 500 ml H2 for a 1 l culture (10 days; 110 W illumination). Without further purification, the H2 gas can used to power a small-scale fuel cell car. In addition to work with Chlamydomonas, a large number of unicellular, filamentous, freshwater, and marine cyanobacterial species have been reported to produce large quantities of biohydrogen. Among other species, Anabaena azollae, Anabaena cylindrica, Anabaena variabilis, Arthrospira (Spirulina) platensis, Cyanothece, Gloeocapsa alpicola, and Nostoc muscorum have been reported to produce high levels of hydrogen gas (Jeffries et al., 1978; Aoyama et al., 1997; Antal and Lindblad, 2005). In particular, Anabaena sp. is reported to produce relatively large quantities of biohydrogen. Among these species, nitrogen-starved A. cylindrica cells produce the highest concentration of biohydrogen (30 ml H2/l/h) (Margheri et al., 1990). These cyanobacterial strains use two sets of enzymes to generate hydrogen gas. The first enzyme is nitrogenase, and it is found in the heterocysts of filamentous cyanobacteria when grown under nitrogen-limiting conditions. Hydrogen is produced as a by-product of fixation of nitrogen into ammonia. The reaction consumes 16 ATP for fixation of 1 mol of N2, and results in formation of 1 mol of H2. The other hydrogen-metabolizing or hydrogen-producing enzymes in cyanobacteria are hydrogenases, which occur as two distinct types in different cyanobacterial species. The first type is uptake hydrogenase (encoded by hupSL), which has the ability to oxidize hydrogen via oxyhydrogenation or the Knallgas reaction. The other type of hydrogenase is reversible or bidirectional hydrogenase (encoded by hoxFUYH), and it is capable of uptake and production of hydrogen (Schmitz et al., 1995; Tamagnini et al., 2002). Hydrogen is an important fuel source and is widely applied in fuel cells, coal liquefaction, upgrading of heavy oils, and several other operations. Hydrogen can be produced biologically by various means, including the steam reformation of bio-oils, dark- and photofermentation of organic materials, and photolysis of water catalyzed by special microalgal species (Kapdan and Kargi, 2006; Ran et al., 2006; Wang et al., 2008). BIOCRUDE In addition to direct combustion, there is growing attention to conversion of biomass into liquid energy carriers. Applying more traditional biofuel production processes (e.g. lipid extraction followed by transesterification, fast pyrolysis and gasification) to algal biomass requires that the algae be dried prior to use. Unless access to waste heat is available, the energy required to first concentrate the biomass to a paste followed by complete drying far exceeds to energy value of the produced biocrude. An alternative production pathway called hydrothermal liquefaction (HTL) bypasses the drying step and converts the algal biomass into a hydrocarbonbased biocrude fuel in the aqueous phase. A simple comparison of the enthalpies of liquid water at 350  C and water vapor at 50  C (i.e. drying the biomass) indicates that processing in liquid water saves 921 kJ/kg. PROPERTIES OF SUBCRITICAL WATER In HTL, water is an important reactant and catalyst, and thus the biomass can be directly converted without an energy-consuming drying step, as in the case of pyrolysis (Bridgwater, 2004). As hot compressed liquid water approaches its thermodynamic critical point (Tc ¼ 373.95  C, Pc ¼ 22.064 MPa), its dielectric constant decreases due to a decrease in hydrogen bonding between water molecules (Figure 10.11). At these conditions, water is still in a liquid state, and has a range of exotic properties very different from liquid water at room temperature. Among them is increased solubility of hydrophobic organic compounds, such as free fatty acids (Holliday et al., 1997). Subcritical water can also sustain acid and base ions simultaneously and promotes radical-driven chemistry. These properties make subcritical water an excellent medium for fast, homogeneous and efficient conversions of algal organics to biocrude. But this technology is not without challengesdthe solubility of some salts in the reacting medium decreases significantly leading to excess precipitate in the system. Salts present in the HTL process are typically subdivided into two categories: Type I and Type II. Type 1 salts, such as NaCl, still exhibit relatively high solubility at subcritical conditions. Type 2 salts such as Na2SO4, on the other hand, have very limited solubility at these conditions (Hodes, 2004). If Type II salts are present in the
167 CONCLUSIONS FIGURE 10.11 The critical point of water. (For color version of this figure, the reader is referred to the online version of this book.) reaction medium, the decreased solubility can lead to what’s known as “shock precipitate” which can adsorb onto the walls of processing equipment causing fouling and eventually blockage. Technologies designed to remove or reduce salts from the production stream are currently being evaluated (Marrone, 2004). processing of microalgae, heterogeneous catalysts may provide a more attractive option than homogeneous catalysts because heterogeneous catalysts can be more easily separated from the reaction products. Further, the yields of HTL biocrude using heterogeneous catalyst have been reported to be as high as 71% (Zhang et al., 2013). HYDROTHERMAL CATALYTIC LIQUEFACTION HTL SUMMARY AND OUTLOOK The principal role of HTL is to fractionate organic macromolecules into simpler molecular units that can then be further upgraded to produce specific liquid fuels. The HTL environment promotes the hydrolytic cleavage of ester linkages in lipids, peptide linkages in proteins, and glycosidic ether linkages in carbohydrates. The speed and efficiency of these cleavage reactions can be improved by the addition of catalysts to the reaction medium. Catalysts are generally classified as homogeneous and heterogeneous. In chemistry, homogeneous catalysis is a sequence of reactions that occur when a catalyst is codissolved in the same phase as the reactants. The most reported homogeneous catalyst for HTL processing of microalgae is Na2CO3 (Tekin, 2013; Zhang et al., 2013). While it has been reported that the addition of Na2CO3 to the HTL process increases the overall biocrude yield from microalgae, others have reported that Na2CO3 negatively impacts yields derived from lipids or proteins, but improves yields of precursors derived from carbohydrates (Biller et al. 2011). The effects of other homogeneous catalysts (e.g. KOH, HCOOH, and CH3COOH) on HTL of microalgae have been examined and ordered according to effectiveness Na2CO3 > CH3COOH > KOH > HCOOH. For HTL Though only a limited amount of work has been done to date, it is clear that hydrothermal catalytic conversion of algae can produce hydrocarbons for liquid biofuels. Thus, there is tremendous potential for this field and the outlook is bright. The majority of the work to date on producing liquid fuels from hydrothermal conversion of aquatic biomass has focused on homogeneous catalysis by metal salts or alkali. More recent studies, however, are beginning to examine heterogeneous catalysts due to advantages in separation and selectivity of the catalyst. More work is needed to identify better heterogeneous catalysts for these applications. In particular, the development of nonprecious metal-based catalysts would provide a major advance. CONCLUSIONS Microalgae are a promising source of clean, renewable biofuel. Not only can it be grown and produced on a large scale, it can be grown in virtually every part of the world including locations that are considered to be otherwise unsuitable to agricultural production and thus lie dormant. However, several challenges remain to its full execution: (1) the successful production of
168 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE feedstock on a large scale; (2) the development of processing methods that are cost-effective and leave intact the desired molecular end-products; and (3) a richer understanding of microalgal chemistry and product accumulation during both growth and processing phases. Whether employing open ponds or PBRs for biomass production, cultures must be carefully monitored to maintain the desired composition of the culture. Factors such as nutrient loads, mixing and light source, and contaminants all drive the production of biomass and thus biofuel precursors. There is a growing trend toward processing microalgae directly from the aqueous stream, eliminating costly drying steps and conserving water. As such, HTL is an emerging process that converts biomass to biocrude in hot, compressed water, thereby eliminating the need for drying or organic solvents. Further, all organic components serve as the feedstock for the HTL process rather than discreet components, such as lipids for biodiesel or ethanol for bioethanol. Biohydrogen is another provocative fuel derived from microalgae. Whatever the feedstock and biofuel process, additional improvements to each of the technologies are required to make the production of renewable fuels from microalgae cost-effective. These improvements can only result from systems using real-time analytical feedback to inform growth and processing and from innovations derived from a multidisciplinary approach. References S&T, January, 2003. The Addition of Ethanol from Wheat to GH Senius. S & T Consultants, Delta, BC. Abou Akkada, A.R., Bartley, E.E., Berube, R., Fina, L.R., Meyer, R.M., Henricks, D., Julius, F., 1968. Simple method to remove completely ciliate protozoa of adult ruminants. Appl. Environ. Microbiol. 16 (10), 1475e1477. Acien Fernandez, F.G., Fernandez Sevilla, J.M., Sanchez Perez, J.A., Molina Grima, E., Chisti, Y., 2001. Airlift-driven external-loop tubular photobioreactors for outdoor production of microalgae: assessment of design and performance. Chem. Eng. Sci. 56, 2721e2732. Akkerman, I., Janssen, M., Rocha, J., Wijffels, R.H., 2002. Photobiological hydrogen production: photochemical efficiency and bioreactor design. Int. J. Hydrogen Energy 27, 1195e1208. Anderson, R., 2005. Algal Culturing Techniques, first ed. Elsevier, Burlington, MA. Anderson, G.A., Kommareddy, A., Schipull, M.A., 2002. Photobioreactor Design. Paper No. MBSK02e216: An ASAE/CSAE Meeting Presentation, Canada. Angermayr, S.A., Hellingwerf, K.J., Lindblad, P., de Mattos, M.J., 2009. Energy biotechnology with cyanobacteria. Curr Opin Biotechnol. 20, 257e263. Antal, T.K., Lindblad, P., 2005. Production of H2 by sulphur-deprived cells of the unicellular cyanobacteria Gloeocapsa alpicola and Synechocystis sp. PCC 6803 during dark incubation with methane or at various extracellular pH. J. App Microbiol. 98, 114e120. Aoyama, K., Uemura, I., Miyake, J., Asada, Y., 1997. Fermentative metabolism to produce hydrogen gas and organic compounds in a cyanobacterium, Spirulina platensis. J. Ferment. Bioeng. 83, 17e20. Arantes, V., Saddler, J.N., 2010. Access to cellulose limits the efficiency of enzymatic hydrolysis: the role of amorphogenesis. Biotechnol. for biofuels 3, 1e11. Asinari Di San Marzano, C.M., Legros, A., Naveau, H.P., Nyns, E.J., 1982. Biomethanation of the marine algae Tetraselmis. Int. J. Sustainable Energy 1, 263e272. Bandyopadhyay, S.S., Navid, M.H., Ghosh, T., Schnitzler, P., Ray, B., 2011. Structural features and in vitro antiviral activities of sulfated polysaccharides from Sphacelaria indica. Phytochemistry 72, 276e283. Banerjee, A., Sharma, R., Chisti, Y., Banerjee, U.C., 2002. Botryococcus braunii: A renewable source of hydrocarbons and other chemicals. Crit. Rev. Biotechnol. 22, 245e279. Bashir, S., Lee, S., 1994. Fuel ethanol production from agricultural lignocellulose feedstocks e a review. Fuel Science Technol. Int’l 12, 1427e1473. Batten, D., Beer, T., Freischmidt, G., Grant, T., Liffman, K., Paterson, D., Priestley, T., Rye, L., Threlfall, G., 2013. Using wastewater and highrate algal ponds for nutrient removal and the production of bioenergy and biofuels. Water Sci. Technol. 67 (4). Beal, C.M., Hebner, R.E., Romanovicz, D.K., Connelly, R.L., 2012a. Progression of lipid profile and cell structure in a research-scale production pathway for algal biocrude. Renewable Energy 50, 86e93. Beal, C.M., Stillwell, A.S., King, C.W., Cohen, S., Berberoglu, H., Bhattarai, R.P., Connelly, R.L., Webber, M.E., Hebner, R.E., 2012b. Energy return on investment for algal biofuel production coupled with wastewater treatment. Water Environ. Res. 84 (9), 692e710. Becker, E.W., 1988. Microalgae for human and animal consumption. In: Borowitzka, M.A., Borowitzka, L.J. (Eds.), Microalgal Biotechnology. Cambridge University Press, Oxford, pp. 222e249. Becker, E.W., 2007. Micro-algae as a source of protein. Biotechnol. Adv. 25, 207e210. Benemann, J.R., 2008. Open Ponds and Closed Photobioreactors e Comparative Economics (Slide Presentation). Paper presented at the Fifth Annual World Congress on Industrial Biotechnology & Bioprocessing, April 27e30, Chicago, Illinois. Benemann, J.R., Oswald, W.J., 1996. Systems and Economic Analysis of Microalgae Ponds for Conversion of Carbon Dioxide to Biomass (Final Report: Grant No. DE-FG22-93PC93204). Pittsburgh Energy Technology Center, Pittsburgh, PA (U.S. Department of Energy). Benemann, J.R., Goebel, R.P., Weissman, J.C., Augenstein, D.C., 1982. Microalgae as a Source of Liquid Fuels (Final Technical Report, Contract Deacos 81 ER 30014). U.S. Department of Energy. Berzin, I., Nov. 24, 2005. Photobioreactor and Process for Biomass Production and Mitigation of Pollutants in Flue Gases. United States Patent Application. Pub no. US2005/0260553 A1, USA, Publication Date. Biller, P., Riley, R., Ross, A.B., 2011. Catalytic hydrothermal processing of microalgae: decomposition and upgrading of lipids. Bioresour Technol. 102, 4841e4848. Borowitzka, M.A., 2005. Culturing microalgae in outdoor ponds. In: Anderson, R.A. (Ed.), Algal Culturing Techniques. Elsevier, Burlington, MA, pp. 205e218. Bridgwater, A.V., Maniatis, K., 2004. The production of biofuels by the thermochemical processing of biomass. In: Archer, M.D., Barber, J. (Eds.), Molecular to Global Photosynthesis. IC Press, London, pp. 521e612. Brindley, C., Fernandez, F.G.A., Fernandez-Sevilla, J.M., 2011. Analysis of light regime in continuous light distributions in photobioreactors. Bioresour. Technol. 102, 3138e3148. Brzezinski, M.A., 1985. The SieCeN ratio of marine diatoms: interspecific variability and the effect of some environmental variables. J. Phycol. 21, 34e357.
REFERENCES Carvalho, A.P., Meireles, L.A., Malcata, F.X., 2006. Microalgal reactors: A review of enclosed system designs and performances. Biotechnol. Prog. 22, 1490e1506. Chisti, Y., 2007. Biodiesel from microalgae. Biotechnol. Adv. 25, 294e306. Choi, S.L., Suh, I.S., Lee, C.G., 2003. Lumostatic operation of bubble column photobioreactors for Haematococcus pluvialis cultures using a specific light uptake rate as a control parameter. Enzyme Microb. Technol. 33, 403e409. Cirne, D.G., Paloumet, X., Bjornsson, L., Alves, M.M., Mattiasson, B., 2007. Anaerobic digestion of lipid-rich wastedeffects of lipid concentration. Renewable Energy 32, 965e975. Craggs, R.J., Smith, V.J., McAuley, P.J., 1995. Wastewater nutrient removal by marine micro-algae cultured under ambient conditions in mini-ponds. Wat. Sci. Technol. 31, 151e160. Dismukes, G.C.D., Carrieri, N., Bennette, G., Ananyey, M., 2008. Aquatic phototrophs: efficient alternatives to land-based crops for biofuels. Curr. Opin. Biotechnol. 19, 235e240. Doucha, J., Livansky, K., 2006. Productivity, CO2/O2 exchange and hydraulics in outdoor open high density microalgal (Chlorella sp.) photobioreactors operated in a middle and southern European climate. J. Appl. Phycol. 18, 811e826. Duffy, P., 1853. XXV. On the constitution of stearine. Q. J. Chem. Soc. London 5, 303. Dunahay, T.G., Jarvis, E.E., Dais, S.S., Roessler, P.G., 1996. Manipulation of microalgal lipid production using genetic engineering. Appl. Biochem. Biotechnol. 57, 223e231. Esquı́vel, M.G., Amaro, H.M., Pinto, T.S., Fevereiro, P.S., Malcata, F.X., 2011. Efficient H2 production via Chlamydomonas reinhardtii. Trends Biotechnol. 29 (12), 595e600. Fedorov, A.S., Kosourov, S., Ghirardi, M.L., Seibert, M., 2005. Continuous H2 photoproduction by Chlamydomonas reinhardtii using a novel two-stage, sulfate-limited chemostat system. Appl Biochem Biotechnol. 121e124, 403e412. Fukuda, H., Kondo, A., Noda, H., 2001. Biodiesel fuel production by transesterification of oils. J. Biosci. Bioeng. 92, 405e416. Gavrilescu, M., Chisti, Y., 2005. Biotechnologyda sustainable alternative for chemical industry. Biotechnol. Adv. 23, 471e499. Ghirardi, M.L., Zhang, J.P., Lee, J.W., Flynn, T., Seiber, M., Greenbaum, E., 2000. Microalgae: a green source of renewable H2. Trends Biotechnol. 18, 506e511. Goldemberg, J., 2007. Ethanol for a sustainable energy future. Science 315, 808e810. Golueke, C.G., Oswald, W.J., 1959. Biological conversion of light energy to the chemical energy of methane. Appl. Microbiol. 7, 219e227. Golueke, C.G., Oswald, W.J., Gotaas, H.B., 1957. Anaerobic digestion of algae. Appl. Microbiol. 5, 47e55. González-López, C.V., Acién Fernández, F.G., Fernández-Sevilla, J.M., Sánchez Fernández, J.F., Molina Grima, E., 2012. Development of a process for efficient use of CO2 from flue gases in the production of photosynthetic microorganisms. Biotechnol. Bioeng. 109 (7), 1637e1650. Grima, E.M., Belarbia, E.H., Fernandez, F.G.A., Medina, A.R., Chisti, Y., 2003. Recovery of microalgal biomass and metabolites: process options and economics. Biotechnol. Adv. 20, 491e515. Guschina, I.A., Harwood, J.L., 2006. Lipids and lipid metabolism in eukaryotic algae. Prog. Lipid Res. 45, 160e186. Hai, T., Ahlers, H., Gorenflo, V., Steinbuchel, A., 2000. Axenic cultivation of anoxygenic phototrophic bacteria, cyanobacteria and microalgae in a new closed tubular glass photobioreactor. Appl. Microbiol. Biotechnol. 53, 383e389. Harrison, P., Berges, J., 2005. Marine culture media. In: Andersen, R.A. (Ed.), Algal Culturing Techniques. Elsevier, Burlington, MA, pp. 21e33. 169 Harun, R., Danquah, M.K., Forde, G.M., 2009. Microalgal biomass as a fermentation feedstock for bioethanol production. J. Chem. Technol. Biotechnol. 85, 199e203. Hodes, M., Marrone, P.A., Hong, G.T., Smith, K.A., Tester, J.W., 2004. Salt precipitation and scale control in supercritical water oxidation part A: fundamentals and research. J. Supercrit Fluids 29, 265e288. Hoekema, S., Bijmans, M., Janssen, M., Tramper, J., Wijffles, R.H., 2002. A pneumatically agitated flat-panel photobioreactor with gas re-circulation: anaerobic photoheterotrophic cultivation of a purple non-sulfur bacterium. Int. J. Hydrogen Energy 27, 1331e1338. Holliday, R.L., King, J.W., List, G.R., 1997. Hydrolysis of vegetable oils in sub- and supercritical water. Ind. Eng. Chem. Res. 36, 932e935. Horta Nogueira, L., 2004. Perspectivas de un Programa de Biocombustibles en America Central: Proyecto uso Sustentable de Hidrocarburos. Comision Economica para America Latina y el Caribe (CEPAL). Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M., Darzins, A., 2008. Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances. Plant J. 54, 621e639. Huntley, M., Redalje, D., 2007. CO2 mitigation and renewable oil from photosynthetic microbes: A new appraisal. Mitigation Adapt. Strategy Global Change 12, 573e608. Jeffries, T.W., Timourian, H., Ward, R.L., 1978. Hydrogen production by Anabaena cylindrica: effects of varying ammonium and ferric ions, pH and light. Appl. Environ. Microbiol. 35, 704e710. Jones, J., Manning, S., Montoya, M., Keller, K., Poenie, M., 2012. Extraction of algal lipids and their analysis by HPLC and mass spectrometry. J. Am. Oil Chem. Soc. 89 (8), 1371e1381. Jones, C.S., Mayfield, S.P., 2012. Algae biofuels: versatility for the future of bioenergy. Curr Opin Biotechnol. 23, 346e351. Kapdan, I.K., Kargi, F., 2006. Bio-hydrogen production from waste materials. Enzym. Microb. Technol. 38, 569e582. Kaya, V.M., Picard, G., 1995. The viability of Scenedesmus bicellularis cells immobilized on alginate screens following nutrient starvation in air at 100% relative humidity. Biotechnol. Bioeng. 46, 459e464. Kessler, E., 1991. Scenedesmus: problems of a highly variable genus of green algae. Botanica Acta 104, 169e171. Kristensen, E., 1990. Characterization of biogenic organic matter by stepwise thermogravimetry (STG). Biogeochemistry 9, 135e159. Lee, Y.K., 2001. Microalgal mass culture systems and methods: their limitations and potential. J. Appl. Phycol. 13, 307e315. Levin, G.V., Clendenning, J.R., Gibor, A., Bogar, F.D., 1961. Harvesting of algae by froth flotation. Appl. Microbiol. 10, 169e175. Levin, D.B., Lawrence, P., Murry, L., 2004. Biohydrogen production: prospects and limitations to practical application. Int. J. Hydrogen Energy 29, 173e185. Li, Y.Y., Sasaki, H., Yamashita, K., Seki, K., Kamigochi, I., 2002. Highrate methane fermentation of lipid-rich food wastes by a highsolids co-digestion process. Water Sci. Technol. 45, 143e150. Margheri, M.C., Tredici, M.R., Allotta, G., Vagnoli, L., 1990. Heterotrophic metabolism and regulation of uptake hydrogenase activity in symbiotic cyanobacteria. In: Polsinelli, M., Materassi, R., Vincenzini, M. (Eds.), Developments in Plant and Soil Sciences-Bio Nitrogen Fixation. Kluwer Academic Publications, Dordrecht, pp. 481e486. Marrone, P.A., Hodes, M., Smith, K.A., Tester, J.W., 2004. Salt precipitation and scale control in supercritical water oxidation part B: Commercial/full-scale applications. J. Supercrit. Fluids 29, 289e312. Matsumoto, M., Hiroko, Y., Nobukazu, S., Hiroshi, O., Tadashi, M., 2003. Saccharification of marine microalgae using marine bacteria for ethanol production. Appl. Biochem. Biotechnol. 105, 247e254.
170 10. SECOND-GENERATION BIOFUEL FROM HIGH-EFFICIENCY ALGAL-DERIVED BIOCRUDE Megharaj, M., Peardon, H.W., Venkateswarlu, K., 1992. Removal of nitrogen and phosphorus by immobilized cells Chlorella vulgaris and Scenedesmus bijugatus isolated from soil. Enz. Microb. Technol. 14, 656e658. Melis, A., 2002. Green alga hydrogen production: Progress, challenges and prospects. Int. J. Hydrogen Energy 27, 1217e1228. Miao, X.L., Wu, Q.Y., 2004. High yield bio-oil production from fast pyrolysis by metabolic controlling of Chlorella protothecoides. J. Biotechnol. 110, 85e93. Mimouni, V., Ulmann, L., Pasquet, V., Mathieu, M., Picot, L., Bougaran, G., Cadoret, J.P., Morant-Manceau, A., Schoefs, B., 2012. The potential of microalgae for the production of bioactive molecules of pharmaceutical interest. Curr. Pharm. Biotechnol. 13 (15), 2733e2750. Molina Grima, E., Acien Fernandez, F.G., Garcia Camacho, F., Chisti, Y., 1999. Photobioreactors: light regime, mass transfer, and scale-up. J. Biotechnol. 70, 231e248. Mulbry, W., Westhead,.EK., Pizarro, C., Sikora, L., 2005. Recycling of manure nutrients: use of algal biomass from dairy manure treatment as a slow release fertilizer. Bioresour. Technol. 96 (4), 451e458. Mussatto, S.I., Dragone, G., Guimarães, P., Silva, J.P., Carneiro, L.M., Roberto, I.C., Vicente, A., Domingues, L., Teixeira, J.A., 2010. Technological trends, global market, and challenges of bio-ethanol production. Biotechnol. Adv. 28, 817e830. Nurdogan, Y., Oswald, W.J., 1995. Enhanced nutrient removal in high rate ponds. Water Sci. Technol. 31, 33e43. Oswald, W.J., 2003. My sixty years in applied algology. J. Appl. Phycol. 15, 99e106. Pulz, O., 1992. Cultivation Techniques for Microalgae in Open and Closed Ponds. Proceedings of the First European workshop on microalgal biotechnology. Potsdam, p. 61. Pulz, O., 2001. Photobioreactors: Production systems for phototrophic microorganisms. Appl. Microbiol. Biotechnol. 57, 287e293. Ragauskas, A.J., Williams, C.K., Davison, B.H., Britovsek, G., Cairney, J., Eckert, C.A., Frederick Jr., W.J., Hallett, J.P., Leak, D.J., Liotta, C.L., Mielenz, J.R., Murphy, R., Templer, R., Tschaplinski, T., 2006. The path forward for biofuels and biomaterials. Science 311, 484e489. Ramachandran, S., Mitsui, A., 1984. Recycling of hydrogen production system using an immobilized blue green algae Oscillatoria sp. Miami BG7, solar energy and sea water. In: Abstracts of the VII International Biotechnology Symposium, pp. 183e184. Ran, J.-H., Wei, X.-X., Wang, X.-Q., 2006. Molecular phylogeny and biogeography of Picea (Pinaceae): implications for phylogeographical studies using cytoplasmic haplotypes. Mol. Phylogenet. Evol. 41, 405e419. Richmond, A., 1999. Physiological principles and modes of cultivation in mass production of photoautotrophic microalgae. In: Cohen, Z. (Ed.), Chemicals from Microalgae. Taylor & Francis, London, UK, pp. 353e386. Roessler, P.G., 1990. Environmental control of glycerolipid metabolism in microalgae: commercial implications and future research directions. J. Phycol. 26, 393e399. Rosillo-Calle, F., Cortez, L., 1998. Towards proalcohol II: a review of the Brazilian bioethanol programme. Biomass Bioenergy 14, 115e124. Sakurai, H., Masukawa, H., 2007. Promoting R & D in photobiological hydrogen production utilizing mariculture-raised cyanobacteria. Mar. Biotechnol 9, 128e145. Sanchez, E.J., Novotny, L., Xie, X.S., 1999. Near-field fluorescence microscopy based on two-photon excitation with metal tips. Phys. Rev. Lett. 82, 4014e4017. Sawayama, S., Inoue, S., Yokoyama, S., 1995. Phylogenetic position of Botryococcus braunii (Chlorophyceae) based on small subunit ribosomal RNA sequence data. J. Phycol. 31, 419e420. Schenk, P., Thomas-Hall, S., Stephens, E., Marx, U., Mussgnug, J., Posten, C., Kruse, O., Hankamer, B., 2008. Second generation biofuels: high-efficiency microalgae for biodiesel production. Bioenergy Res. 1, 20e43. Schmitz, O., Boison, G., Hilscher, R., Hundeshagen, B., Zimmer, W., Lottspeich, F., Bothe, H., 1995. Molecular biological analysis of a bidirectional hydrogenase from cyanobacteria. Eur. J. Biochem. 233, 266e276. Sheehan, J., Dunahay, T., Benemann, J., Roessler, P., 1998. A Look Back at the U.S. Department of Energy’s Aquatic Species Program e Biodiesel from Algae. National Renewable Energy Laboratory. NREL/TP-580e24190. Sialve, B., Bernet, N., Bernard, O., 2009. Anaerobic digestion of microalgae as a necessary step to make microalgal biodiesel sustainable. Biotechnol. Adv. 27, 409e416. Singh, J., Gu, S., 2010. Commercialization potential of microalgae for biofuels production. Renewable and Sustainable Energy Reviews 14, 2596e2610. Skjånes, K., Rebours, C., Lindblad, P., 2013. Potential for green microalgae to produce hydrogen, pharmaceuticals and other highvalue products in a combined process. Crit. Rev. Biotechnol. 33 (2), 172e215. Spolaore, P., Joannis-Cassan, C., Duran, E., Isambert, A., 2006. Commercial applications of microalgae. J. Biosci. Bioeng. 101, 87e96. Stamey, J.A., Shepherd, D.M., de Veth, M.J., Corl,., B.A., 2012. Use of algae or algal oil rich in n-3 fatty acids as a feed supplement for dairy cattle. J. Dairy Sci. 95 (9), 5269e5275. Tamagnini, P., Axelsson, R., Lindberg, P., Oxelfelt, F., Wunschiers, R., Lindblad, P., 2002. Hydrogenases and hydrogen metabolism of cyanobacteria. Microbiol. Mol. Biol. Rev. 66, 1e20. Tekin, K., Karagoz, S., 2013. Non-catalytic and catalytic hydrothermal liquefaction of biomass. Res. Chem. Intermed. 39 (2), 485e498. Tredici, M.R., Margheri, M.C., Zittelli, G.C., Biagiolini, S., Capolino, E., Natali, M., 1992. Nitrogen and phosphorus reclamation from municipal wastewater through an artificial food-chain system. Bioresource Technol. 42, 247e253. Tredici, M.R., Zitelli, G.C., 1998. Efficiency of sunlight utilization: tubular versus flat photobioreactors. Biotechnol. Bioeng. 587, 187e197. Ugwu, C.U., Ogbonna, J.C., Tanaka, H., 2005. Characterization of light utilization and biomass yields of Chlorella sorokiniana in inclined outdoor tubular photobioreactors equipped with static mixers. Process Biochem. 40, 3406e3411. Wang, B., Li, Y., Wu, N., Lan, C.Q., 2008. CO2 bio-mitigation using microalgae. Appl. Microbiol. Biotechnol. 79, 707e718. Weissman, J.C., Goebel, R.P., 1987. Design and Analysis of Microalgal Open Pond Systems for the Purpose of Producing Fuels. SERI/ STR-231-2840. Solar Energy Research Institute, Golden, Colorado. Yen, H.W., Brune, D.E., 2007. Anaerobic co-digestion of algal sludge and waste paper to produce methane. Bioresour. Technol. 98, 130e134. Zhang, J., Chen, W.T., Zhang, P., Luo, Z., Zhang, Y., 2013. Hydrothermal liquefaction of Chlorella pyrenoidosa in sub- and supercritical ethanol with heterogeneous catalysts. Bioresour Technol. 133, 389e397. Zhu, L., Wang, Z., Shu, Q., Takala, J., Hiltunen, E., Feng, P., Yuan, Z., 2013. Nutrient removal and biodiesel production by integration of freshwater algae cultivation with piggery wastewater treatment. Water Res. 47 (13), 4294e4302. Zijffers, J.W.F., Salim, S., Janssen, M., Tramper, J., Wijffels, R.H., 2008. Capturing sunlight into a photobioreactor: ray tracing simulations of the propagation of light from capture to distribution into the reactor. Chem. Eng. J. 145, 316e327. Zittelli, C.G., Lavista, F., Bastianini, A., Rodolfi, L., Vincenzini, M., Tredici, M.R., 1999. Production of eicosapentaenoic acid by Nannochloropsis sp. cultures in outdoor tubular photobioreactors. J Biotechnol. 70, 299e312.
C H A P T E R 11 Microalgae: The Tiny Microbes with a Big Impact Shovon Mandal 1, Nirupama Mallick 2,* 1 Section of Ecology, Behavior and Evolution, University of California, San Diego, CA, USA, Agricultural and Food Engineering Department, Indian Institute of Technology, Kharagpur, West Bengal, India *Corresponding author email: nm@agfe.iitkgp.ernet.in 2 O U T L I N E Fatty Acid Methyl Esters and Fuel Properties Renewable Energy 171 Petroleum Fuel Scenario in India 172 Biodiesel 172 Microalgae: Viable Feedstocks for Biodiesel 173 Waste Utilization for Biodiesel Production: A Case Study with Scenedesmus obliquus in a Recirculatory Aquaculture System 179 Selection of Potent Strains 173 Concluding Remarks 181 Genetic Engineering Approach 175 References 181 Microalgal Biodiesel Production 177 RENEWABLE ENERGY Energy is an important currency for human society. The world population growth and rapid economic progresses are expected to result in considerable increase in the demand for energy. In the reference scenario, the International Energy Agency has projected an increase in energy need by 55%, between 2005 and 2030, at an average annual rate of 1.8% (IEA, 2007). Driven by such increasing demand, and the dwindling fuel production, the cost of petroleum fuel has gone up sky high in recent times, which can jeopardize the economic progresses of a nation. Despite the fuel crisis, increasing concentrations of CO2 and other heattrapping greenhouse gases (GHGs) in the atmosphere, primarily due to the combustion of fossil fuels, is clearly the prime reason for rapid warming of the Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00011-5 179 planet (Shay, 1993). The use of renewable energy is largely motivated from the standpoint of global energy crisis and environmental issues. Renewable energy is a form of energy that is produced from natural sources like sunlight, wind, hydropower, geothermal and biomass, which can be naturally replenished. Currently, renewable energy supplies only w18% of the world’s energy consumption (Kumar et al., 2010). Most of these renewable energy sources (hydropower, wind, solar and geothermal) target the electricity market, while the majority of world energy consumption (about two-thirds) is derived from liquid fuels (Campbell, 2008; Hankamer et al., 2007). This has stimulated recent interest to explore alternative sources for petroleumbased fuels and much of the attention has been focused on biomass-derived liquid fuels or biofuels (Haag, 2007; Schneider, 2006). 171 Copyright Ó 2014 Elsevier B.V. All rights reserved.
172 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT PETROLEUM FUEL SCENARIO IN INDIA India ranks seventh as the world’s energy producer accounting for about 2.5% of the world’s total annual energy production, and world’s fifth largest energy consumer with about 3.5% of the global primary energy demand (IEA, 2007; Planning Commission, Govt. of India, 2007). Despite being among the largest energy producer, India is a net importer of energy, largely due to huge imbalance between energy consumption and production. About 30% of India’s total primary energy need is being met by petroleum oil, of which 76% is imported. India’s transportation fuel requirements are unique in the world. India consumes almost five times more diesel fuel than gasoline, whereas all other countries in the world use more gasoline than diesel fuel (Khan et al., 2009). Thus, search for alternatives to diesel fuel is of special importance in India. Bioalcohols are unsuitable substitutes for diesel engines, because of their low cetane numbers (CNs) along with poor energy content per unit biomass (Bhattacharyya and Reddy, 1994; Rao and Gopalkrishnan, 1991). Therefore, biodiesel is the only option to fulfill the requirements in future. BIODIESEL Biodiesel is chemically monoalkyl esters of longchain fatty acids derived from vegetable oils or animal fats. The history of using vegetable oil as an alternative fuel dates back to 1900, when Rudolph Diesel used peanut oil as fuel in the World Exhibition in Paris. It was found that vegetable oils, in general, have acceptable CNs and calorific values comparable with the conventional diesel. However, the major problem with the direct use of vegetable oils as fuel of compression ignition engine is their high viscosity, which interferes with the fuel injection and atomization contributing to incomplete combustion, nozzle clogging, injector coking, severe engine deposits, ring sticking and gum formation leading to engine failure (Knothe, 2005; Meher et al., 2006; Singh and Rastogi, 2009). Therefore, vegetable oils need to be modified to bring their combustion-related properties closer to those of diesel fuel. One possible method to overcome the problem of high viscosity of vegetable oils is their chemical modification to esters, what is nowadays called as “biodiesel”. Biodiesel has emerged as the most suitable alternative to petroleum diesel fuel owing to its ecofriendly characteristics and renewability (Krawczyk, 1996). It burns in conventional diesel engines with or without any modifications while reducing pollution (100% less sulfur dioxide, 37% less unburned hydrocarbons, 46% less carbon monoxide, and 84% less particulate matter) in comparison to the conventional diesel fuel (McMillen et al., 2005). The basic feedstocks for the production of first-generation biodiesel were mainly edible vegetable oils like soybean, rapeseed, sunflower and safflower. The use of first-generation biodiesel has generated a lot of controversy, mainly due to their impact on global food markets and food security for diverting food away from the human food chain. The second-generation biodiesel was produced by using nonedible oil sources like used frying oil, grease, tallow, lard, karanja, jatropha and mahua oils (Alcantara et al., 2000; Francis and Becker 2002; Canakci and Gerpen, 1999; Dorado et al., 2002; Ghadge and Raheman, 2006; Mittelbach, 1990). Nevertheless, the cost of biodiesel production is still a major obstacle for large-scale commercial exploitation, mainly due to the high feed cost of vegetable oils (Lang et al., 2001). Moreover, the first- as well as the second-generation biodiesel based on terrestrial plants initiate land clearing and potentially compete with net food production (Chisti, 2008; Marsh, 2009). The focus of researchers has now been shifted to the next generation biodiesel. The third-generation biodiesel is both promising and different; it is based on simple microscopic organisms that live in water and grow hydroponically, i.e. microalgae. The possibilities of biodiesel production from edible oil resources in India is almost impossible, as primary need is to first meet the demand of edible oil that is being imported. India accounts for 9.3% of the world’s total oil seed production and contributes as the fourth largest edible oil producing country. Even then, about 46% of edible oil is imported for catering the domestic needs (Jain and Sharma, 2010). So the nonedible oil resources like Jatropha, pongamia, mahua, etc. seem to be the only possibility for biodiesel production in the country. The Government of India has duly realized the importance of biodiesel and introduced a nationwide program under the National Biodiesel Mission in 2003 with the aim of achieving a target of meeting 13.4 Mt of biodiesel (@ 20% blending) from Jatropha curcas by 2012, and to achieve the target about 27 billion of planting materials are required to be planted over 11.2 million hectares of land (Planning Commission, Govt. of India, 2003). At the current rate of consumption, if all petroleumderived transport fuel is to be replaced with biodiesel from Jatropha oil, Jatropha would need to be grown over an area of 384 million hectares, which is more than 100% of the geographic area of India (Khan et al., 2009). Therefore, India must find additional, reliable, cost-effective and sustainable feedstock for biodiesel production. In this context, biodiesel from microalgae seems to be a suitable substitute for diesel fuel in the long run.
SELECTION OF POTENT STRAINS MICROALGAE: VIABLE FEEDSTOCKS FOR BIODIESEL Microalgae are a diverse group of photosynthetic organisms whose systematics is based on the kinds and combinations of photosynthetic pigments present in different species. They can grow in diverse environmental conditions, and are able to produce a wide range of chemical products with applications in feed, food, nutritional, cosmetic and pharmaceutical industries. These are primitive organisms with a simple cellular structure and a large surface to volume body ratio, which gives them the ability to take up a large amount of nutrients. While the mechanism of photosynthesis in microalgae is similar to that of higher plants, they have the ability to capture solar energy with an efficiency of 10e50 times higher than that of terrestrial plants (Li et al., 2008). Moreover, because the cells grow in aqueous suspension, they have more efficient access to water, CO2 and other nutrients. For these reasons, microalgae are capable of producing more amount of oil per unit area of land in comparison to that of all other known oil-producing crops (Chisti, 2007; Haag, 2007). The per hectare yield of microalgal oil has been projected to be 58,700e136,900 l/year depending upon the oil content of algae, which is about 10e20 times higher than the best oil producing crop, i.e. palm (5950 l/ha year, Chisti, 2007). The most acclaimed energy crop, i.e. Jatropha has been estimated to produce only 1892 l/ha year. More importantly, due to being aquatic in nature, algae do not compete for arable land for their cultivation; they can be grown in freshwater or saline, and salt concentrations up to twice that of seawater can be used effectively for few species (Aresta et al., 2005; Brown and Zeiler, 1993). The utilization of wastewaters that are rich in nitrogen and phosphorus may bring about remarkable advantages by providing N and P nutrients for growing microalgae, while removing N and P from the wastewaters (Mallick, 2002). This implies that algae need not compete with other users for freshwater (Campbell, 2008). On top of these advantages, microalgae grow even better when fed with extra carbon dioxide, the main GHG. If so, these tiny organisms can fix CO2 from power stations and other industrial plants, thereby cleaning up the greenhouse problem. Each ton of algae produced consumes about 1.8 ton of CO2 (Chisti, 2007). Thus, the integrated efforts to cleanup industrial flue gas with microalgal culture by combining it with wastewater treatment will significantly enhance the environmental and economical benefits of the technology for biodiesel production by minimizing the additional cost of nutrients and saving the precious freshwater resources. 173 SELECTION OF POTENT STRAINS Realizing the oil-yielding potentialities with much faster growth rate and efficient CO2 fixation, microalgae appear to be the best option as a renewable source of biodiesel that has the potentiality to completely replace the petroleum diesel fuel. However, the lipid content in the selected microalga/strain is required to be high; otherwise the economic performance would be hard to achieve. Each species of microalga produces different ratios of lipids, carbohydrates and proteins. Nevertheless, these tiny organisms have the ability to manipulate their metabolism by simple manipulations of the chemical composition of the culture medium (Behrens and Kyle, 1996); thus, high lipid productivity can be achieved. Physiological stresses such as nutrient limitation/deficiency, salt stress and high light intensity have been employed for directing metabolic fluxes to lipid biosynthesis of microalgae. Many reports are available, where attempts have been made to raise the lipid pool of various microalgal species. Table 11.1 summarizes those studies. Exceptionally, an oil content of 86% of dry cell weight (dcw) was reported in the brown resting state colonies of Botryococcus braunii, while the green active state colonies were found to account for 17% only (Brown et al., 1969). However, the major obstacle in focusing B. braunii as an industrial organism for biodiesel production is its poor growth rate (Dayananda et al., 2007). Nitrogen limitation/deficiency has been found to raise the lipid content of a number of microalgal species profoundly. For instance, Piorreck and Pohl (1984) reported an increased lipid pool from 12% to 53% (dcw) in Chlorella vulgaris under nitrogen-limited condition. Unlike the green algae, the blue-green algae viz. Anacystis nidulans and Oscillatoria rubescens contained the same quantities of lipid at different nitrogen concentrations. It was observed by Illman et al. (2000) that four species of Chlorella (Chlorella emersonii, Chlorella minutissima, C. vulgaris and Chlorella pyrenoidosa) could accumulate lipid up to 63, 57, 40 and 23% (dcw), respectively, in low N-medium. These values in control vessels were, respectively, 29%, 31%, 18% and 11% in the above order. In the same year, Takagi et al. (2000) observed an increase in intracellular lipid pool up to 51% (dcw) against 31% control in 3% CO2-purged cultures of Nannochloris sp. UTEX LB1999 grown in continuous low nitrate (0.9 mM)-fed medium. Chlorella protothecoides also showed a rise in lipid pool from 15% to 55% (dcw), when grown heterotrophically with glucose (1%) under reduced nitrogen concentration (Miao and Wu, 2004). Similarly, C. protothecoides depicted a lipid pool of 55% (dcw) when grown heterotrophically with corn powder hydrolysate under nitrogen limitation (Xu et al., 2006).
174 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT TABLE 11.1 A List of Studies on Increased Lipid Accumulation in Microalgae under Various Specific Conditions Microalga Growth Condition Lipid Content as Percent of Dry Cell Weight Botryococcus braunii Brown resting state 86 (17*) Brown et al. (1969) Chlorella vulgaris Nitrogen limitation 53 (12*) Piorreck and Pohl (1984) Chlorella emersonii Nitrogen limitation 63 (29*) Illman et al. (2000) Chlorella minutissima 57 (31*) Chlorella vulgaris 40 (18*) Chlorella pyrenoidosa 23 (11*) References Nannochloris sp. UTEX LB1999 Nitrogen limitation 51 (31*) Takagi et al. (2000) Chlorella protothecoides Heterotrophy with 0.1% glucose under nitrogen limitation 55 (15*) Miao and Wu (2004) Heterotrophy with corn powder hydrolysate under nitrogen limitation 55 (15*) Xu et al. (2006) Dunaliella sp. 1 M NaCl 71 (64*) Takagi et al. (2006) Chlorella sp. Heterotrophy with 1% sucrose 33 (15*) Rattanapoltee et al. (2008) Scenedesmus obliquus Nitrogen and phosphorus limitations in presence of thiosulphate 58 (13*) Mandal and Mallick (2009) Neochloris oleoabundans Nitrogen deficiency 56 (29*) Gouveia and Oliveira (2009) Nannochloropsis oculata NCTU-3 2% CO2 50 (31*) Chiu et al. (2009) Nannochloropsis sp. F&M-M24 Nitrogen deficiency 60 (31*) Rodolfi et al. (2009) Phosphorus deficiency 50 (31*) Nannochloropsis oculata Nitrogen limitation 15 (8*) Chlorella vulgaris Converti et al. (2009) 16 (6*) Choricystis minor Nitrogen and phosphorus deficiencies 60 (27*) Sobczuk and Chisti (2010) Haematococcus pluvialis High light intensity 35 (15*) Damiani et al. (2010) High light intensity under nitrogen deficiency 33 (15*) Chlorella protothecoides Heterotrophy with sweet sorghum hydrolysate under nitrogen limitation 50 (15*) Gao et al. (2010) Chlorella zofingiensis Nitrogen limitation 55 (27*) Feng et al. (2011)a Isochrysis zhangjiangensis High nitrogen (0.9%) supplementation 53 (41*) Feng et al. (2011)b Dunaliella tertiolecta Nitrogen deficiency 26 (12*) Jiang et al. (2012) Thalassiosira pseudonana Chlorella vulgaris * Lipid content of control culture. 20 (13*) Nitrogen, phosphorus and iron limitations 57 (8*) Mallick et al. (2012)
GENETIC ENGINEERING APPROACH Gao et al. (2010) used sweet sorghum hydrolysate instead of corn powder for C. protothecoides culture, and lipid yield of 50% (dcw) was recorded. Nitrogen limitation/starvation also enhanced the lipid content in Neochloris oleoabundans, Nannochloropsis oculata, C. vulgaris, Chlorella zofingiensis, Dunaliella tertiolecta and Thalassiosira pseudonana (Converti et al., 2009; Feng et al., 2011a; Gouveia and Oliveira, 2009; Jiang et al., 2012). However, the marine microalga Isochrysis zhangjiangensis was found to accumulate lipid under high nitrate concentration, rather than limitation or depletion (Feng et al., 2011b). Limitation of phosphate was also found to enhance lipid accumulation in Ankistrodesmus falcutus and Monodus subterraneus (Kilham et al., 1997; Khozin-Goldberg and Cohen, 2006). Rodolfi et al. (2009) screened 30 microalgal strains for lipid production, among which the marine genus Nannochloropsis sp. F&M-M24 emerged as the best candidate for oil production (50% under phosphorus deficiency against 31% control). Sobczuk and Chisti (2010) observed an increase in intracellular lipid content up to 60% (dcw) against 27% control in Choricystis minor under simultaneous nitrate and phosphate deficiencies. In Scenedemus obliquus, lipid accumulation up to 58% (dcw) was recorded when subjected to simultaneous nitrate and phosphate limitations in presence of sodium thiosulphate (against 13% under control condition, Mandal and Mallick, 2009). Simultaneous nitrate, phosphate and iron limitations have also been reported to stimulate lipid accumulation in a microalga C. vulgaris (57% against 8% control, Mallick et al., 2012). In addition to nutrient limitations/deficiencies, other stress conditions may also enhance lipid accumulation in microalgae. Takagi et al. (2006) studied the effect of NaCl on accumulation of lipids and triacylglycerides in the marine microalga Dunaliella sp. Increase in initial NaCl concentration from 0.5 M (seawater) to 1.0 M resulted in a higher intracellular lipid accumulation (71% dcw). Damiani et al. (2010) studied the effects of continuous high light intensity (300 mmol photons/ m2 s) on lipid accumulation in Haematococcus pluvialis grown under nitrogen-sufficient and nitrogen-deprived conditions. A lipid yield of 33e35% was recorded under the high light intensity as compared to 15% yield in control cultures. Nitrogen deprivation was, however, not found to raise the lipid content of H. pluvialis cultures. Nutrient limitations/deficiencies or physiological stresses required for accumulation of lipids in microalgal cells is associated with reduced cell division (Ratledge, 2002). The overall lipid productivity is therefore compromised due to the low biomass productivity. For instance, Scragg et al. (2002) studied the energy recovery from C. vulgaris and C. emersonii grown in complete Watanabe medium and also in low nitrogen 175 medium. The results showed that the low nitrogen medium, although induced higher lipid accumulation in both the test algae with high calorific values, the overall energy recovery was lower in comparison to Watanabe’s medium. A commonly suggested counter measure is to use a two-stage cultivation strategy, dedicating the first stage for cell growth/division in nutrient sufficient medium, and the second stage for lipid accumulation under nutrient starvation or other physiological stresses. To get maximal biomass and lipid yield, CO2 can also be utilized. Chiu et al. (2009) reported an increased accumulation of lipid (from 31% to 50% dcw) in the stationary phase cultures of N. oculata NCTU-3 grown under 2% CO2 aeration. GENETIC ENGINEERING APPROACH High oil-yielding transgenic microalgae could be a promising source for biodiesel production. However, the biotechnological processes based on transgenic microalgae are still in infancy. In manipulation of genetically modified algae for high oil content, acetyl-CoA carboxylase (ACCase) was first isolated from the diatom Cyclotella cryptica by Roessler (1990), and then successfully transformed into the diatoms C. cryptica and Navicula saprophila (Dunahay et al., 1995, 1996; Sheehan et al., 1998). A plasmid was constructed that contained acc1 gene driven by the cauliflower mosaic virus 35S ribosomal gene promoter (CaMV35S) and the selectable marker nptII from Escherichia coli. Introduction of plasmids into the diatoms was mediated by microprojectile bombardment. The acc1 was overexpressed with the enzyme activity enhanced by threefold. These experiments demonstrated that ACCase could be transformed efficiently into microalgae, although no significant increase in lipid accumulation was observed in the transgenic diatoms (Dunahay et al., 1995, 1996). Recently, diacylglycerol acyltransferases (DGATs) homologous genes have been identified in the genome of Chlamydomonas reinhardtii and were overexpressed in the same microalga (Russa et al., 2012). This resulted in an enhanced mRNA expression level of DGAT genes, but did not boost the intracellular triacylglycerol (TAG) synthesis. Thus, till date, there is no success story with respect to lipid overproduction in microalgae using the genetic engineering approach. Extensive studies have also been carried out on enhancement of lipid production using genetic engineering approaches in different bacterial and plant species, which may provide valuable background for future studies with microalgae. Some of these studies are summarized in Table 11.2. The cytosolic ACCase from Arabidopsis sp. was overexpressed in Brassica napus (rapeseed) plastids. The fatty acid content of the
176 TABLE 11.2 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT A List on Trials to Enhance Lipid Biosynthesis in Transgenic Organisms Gene (Enzyme) Source Species Receiver Species Result References acc1 (ACCase) Cyclotella cryptica Cyclotella cryptica 3-fold rise in ACCase activity, no change in lipid content Dunahay et al. (1995, 1996) Navicula saprophila 3-fold rise in ACCase activity, no change in lipid content acc1 (ACCase) Arabidopsis sp. Brassica napus 2-fold rise in plastid ACCase, 6% rise in fatty acid content Roesler et al. (1997) LPAT Saccharomyces cerevisiae Brassica napus 6-fold rise in oil content Zou et al. (1997) accA, accB, accC, accD (ACCase) E. coli E. coli 6-fold rise in fatty acid synthesis Davis et al. (2000) are1 and are2 (DGAT) Arabidopsis thaliana Saccharomyces cerevisiae 9-fold rise in TAG content Bouvier-Nave et al. (2000) DGAT Arabidopsis sp. Arabidopsis sp. 70% rise in lipid content Jako et al. (2001) acc1 (ACCase) Arabidopsis sp. Solanum tuberosum 5-fold rise in TAG content Klaus et al. (2004) acs (ACS) E. coil E. coli 9-fold rise in ACS activity Lin et al. (2006) malEMt and malEMc (malic enzyme, ME) Mortierella alpina and Mucor circinelloides M. circinelloides 2.5-fold rise in lipid accumulation Zhang et al. (2007) fadD, ACCase, thioesterase (TE) E. coil E. coli 20-fold rise in fatty acid synthesis Lu et al. (2008) wri1 Brassica napus Arabidopsis thaliana 40% rise in oil content Liu et al. (2010) Acyl-ACP thioesterase Diploknema butyracea, Ricinus communis and Jatropha curcas E. coli 0.2e2.0 g/l free fatty acid yield Zhang et al. (2011) DGAT Chlamydomonas reinhardtii C. reinhardtii 29-Fold rise in mRNA level, no change in TAG Russa et al. (2012) Source: Modified from Courchesne et al. (2009) recombinant was 6% higher than that of the control (Roesler et al., 1997). In prokaryotes like E. coli, overexpression of four ACCase subunits resulted in sixfold rise in the rate of fatty acid synthesis (Davis et al., 2000), confirming that the ACCase-catalyzed committing step was indeed the rate-limiting step for fatty acid biosynthesis in this strain. Nevertheless, Klaus et al. (2004) achieved an increase in fatty acid synthesis and a more than fivefold rise in the amount of TAG in Solanum tuberosum (potato) by overexpressing the ACCase from Arabidopsis in the amyloplasts of potato tubers. Transformation of rape seed with a putative sn-2-acyltransferase gene from Saccharomyces cerevisiae was carried out by Zou et al. (1997), leading to overexpression of seed lysophosphatidate acid acyl-transferase (LPAT) activity. This enzyme is involved in TAG formation and its overexpression led to profound rise in oil content from 8% to 48% on seed dry weight basis. However, it was cautioned that the steady state level of diacylglycerol could be perturbed by an increase in LPAT activity in the developing seeds. Transformations of S. cerevisiae with the Arabidopsis DGAT were performed by Bouvier-Nave et al. (2000). About 600-fold rise in DGAT activity in the transformed S. cerevisiae was observed, which led to a ninefold increase in TAG accumulation. DGAT gene has also been overexpressed in the plant Arabidopsis and it was shown that the oil content was enhanced in correlation with the DGAT activity (Jako et al., 2001). All these results suggest that the reaction catalyzed by ACCase, LPAT and DGAT are important rate-limiting steps in lipid biosynthesis. A few enzymes that are not directly involved in lipid metabolism have also been demonstrated to influence the rate of lipid accumulation. For instance, it was observed by Lin et al. (2006) that by overexpressing the acs gene in E. coli, the acetyl-CoA synthase activity was increased by ninefold, leading to a significant increase in the assimilation of acetate from the medium, which can contribute to lipid biosynthesis. The genes coding for malic enzyme from Mucor circinelloides (malEMt) and from Mortierella alpina (malEMc), respectively, were overexpressed in M. circinelloides which led to a 2.5-fold increase in lipid accumulation (Zhang et al., 2007). Lu et al. (2008) reported a 20-fold enhancement
MICROALGAL BIODIESEL PRODUCTION of fatty acid productivity of E. coli by combining four targeted genotypic changes: deletion of the fadD gene encoding the first enzyme in fatty acid degradation, overexpression of the genes encoding the endogenous ACCase, and overexpression of both an endogenous thioesterase (TE) as well as a heterologous plant TE. Overexpression of wri1 gene from B. napus in transgenic Arabidopsis thaliana resulted into 40% increased seed oil content (Liu et al., 2010). Zhang et al. (2011) studied the effects of the overexpression of different acyl-ACP TE genes from Diploknema butyracea, Ricinus communis and J. curcas on free fatty acid contents of E. coli. The strain carrying the acyl-ACP TE gene from D. butyracea produced approximately 0.2 g/l of free fatty acid while the strains carrying acyl-ACP TE genes from R. communis and J. curcas produced the free fatty acid at a high level of more than 2.0 g/l. MICROALGAL BIODIESEL PRODUCTION Microalgal biodiesel production is relatively new and not very well explored. Some reports are available, where attempts have been made to produce biodiesel from algae (Table 11.3). Miao and Wu (2006) reported that lipid extracted from the heterotrophically grown microalga, C. protothecoides, transformed into biodiesel with a yield of 63% under 1:1 weight ratio of H2SO4 to oil, and 56:1 molar ratio of methanol to oil at 30  C for a reaction time of 4 h. Xu et al. (2006) characterized the biodiesel obtained from the C. protothecoides oil by acid-catalyzed transesterification. The most abundant fatty acid methyl ester (FAME) in C. prothecoides biodiesel was methyl oleate (61% of total FAME) followed by methyl linoleate (17%) and methyl palmitate (13%). Subsequently, Li et al. (2007) showed that it was feasible to grow C. protothecoides in a commercial-scale bioreactor. Using 75% immobilized lipase, these researchers claimed w98% conversion could be obtained in 12 h when the reaction condition with respect to solvent type, water content and pH were optimized. Hossain and Salleh (2008) studied biodiesel production from Oedogonium and Spirogyra species using NaOH as catalyst. Algal oil and biodiesel production was higher in Oedogonium sp. than in Spirogyra sp. Umdu et al. (2009) studied the effects of Al2O3 supported CaO and MgO catalysts in the transesterification of lipid of N. oculata. These researchers found that pure CaO and MgO were not active, and CaO-Al2O3 catalyst showed the highest activity. Biodiesel yield was increased up to 98% from 23% under CaO-Al2O3 catalyzed reaction when methanol: lipid ratio was increased from 6:1 to 30:1. Lipid extracted from N. oleoabundans was found to have an adequate fatty acid profile and iodine value according to the biodiesel specifications of European 177 Standards (EN, Gouveia et al. 2009). Converti et al. (2009) analyzed the FAMEs in biodiesel produced from N. oculata and C. vulgaris. The most abundant composition was methyl palmitate, which was 62% and 66%, respectively, in N. oculata and C. vulgaris biodiesel. However, the concentration of linolenic acid (18%) in N. oculata could not meet the requirement of European legislation for biodiesel. Johnson and Wen (2009) prepared biodiesel from the microalga Schizochytrium limacinum by direct transesterification of algal biomass. Parameters such as free glycerol, total glycerol, acid number, soap content, corrosiveness to copper, flash point and viscosity met the American Society for Testing and Materials (ASTM) and European standards, while the water and sediment content, as well as the sulfur content did not pass the standards. Damiani et al. (2010) studied biodiesel production from H. pluvialis using potassium hydroxide as the catalyst. The major constituent of H. pluvialis biodiesel was palmitic acid followed by linoleic, oleic and linolenic acid methyl esters. The iodine value was within the limit established by European standards. Chinnasamy et al. (2010) produced biodiesel by a two-step transesterification process (acid-catalyzed followed by base-catalyzed) from a consortium of 15 native algae cultivated in carpet industry wastewater. Algal methyl esters were predominated by linolenic, linoleic, palmitic and oleic acids. The biodiesel was found to contain 0.0155% and 0.0001% bound and free glycerin, respectively, and met the ASTM and European standard specifications. Patil et al. (2012) optimized the direct conversion of wet Nannochlopsis sp. biomass to biodiesel under supercritical methanol treatment, without using any catalyst. In the supercritical state, at high pressure and temperature, the methanol molecules enabled simultaneous extraction and transesterification of lipids in wet algal biomass. The abundant FAME in Nannochlopsis sp. biodiesel was methyl oleate (37%) followed by methyl palmitolate (32%) and methyl palmitate (8%). Velasquez-Orta et al. (2012) compared in situ transesterification of C. vulgaris with acid as well as alkaline catalysts, in which the oil extraction step was eliminated. FAME yield reached a maximum of 77.6% after 45 min using a catalyst (NaOH) ratio of 0.15:1 and solvent ratio of 600:1 at 60  C under constant stirring rate of 380 rpm. However, with sulfuric acid as catalyst FAME yield reached up to 96.9% with catalyst : oil ratio of 0.35:1 for a reaction time of 20 h. Recently, Mallick et al. (2012) characterized the biodiesel obtained from the C. vulgaris oil by acid-catalyzed transesterification. The fuel properties (density, viscosity, acid value, iodine value, calorific value, cetane index, ash and water contents) of C. vulgaris biodiesel are comparable with the international (ASTM and EN) and Indian standards (IS).
TABLE 11.3 Attempts on Biodiesel Production from Microalgae Major Ester Physical Property References Chlorella protothecoides H2SO4-catalyzed (63%) NC Density: 0.86 kg/l, viscosity: 5.2 cSt, flash point: 115  C, acid value: 0.37 mg KOH/g, heating value: 41 MJ/kg Miao and Wu (2006) H2SO4-catalyzed (63%) Methyl oleate: 61%, methyl linoleate: 17%, methyl palmitate: 13% Density: 0.86 kg/ l, viscosity: 5.2 cSt, flash point: 115  C, solidifying point: 12  C, acid value: 0.37 mg KOH/g Xu et al. (2006) Lipase-catalyzed (98%) Methyl oleate: 65%, methyl linoleate: 18%, methyl palmitate: 10% NC Li et al. (2007) Oedogonium sp. NaOH-catalyzed (95%) NC NC Hossain and Salleh (2008) Spirogyra sp. NaOH-catalyzed (93%) Nannochloropsis oculata Heterogeneous catalyst (Al2O3supported CaO & MgO) (98%) NC NC Umdu et al. (2009) Neochloris oleoabundans BF3-catalyzed (NR) Methyl oleate: 38%, methyl palmitate: 17%, methyl stearate: 14%, methyl linolenate: 8% Iodine value: 72 g I2/100 g Gouveia et al. (2009) Nannochloropsis oculata Acid-catalyzed (NR) Methyl palmitate: 62%, methyl linolenate: 18%, methyl linoleate: 12%, methyl oleate: 6% NC Converti et al. (2009) Chlorella vulgaris Methyl palmitate: 66%, methyl linolenate: 12%, methyl linoleate: 11%, methyl oleate: 7% Schizochytrium limacinum H2SO4-catalyzed (66%) Methyl palmitate: 57%, methyl ester of C22: 6:30% Viscosity: 3.87 cSt, flash point: 204  C, moisture content: 0.11%, acid value: 0.11 mg KOH/g, total glycerin: 0.097%, free glycerin: 0.003%, Johnson and Wen (2009) Haematococcus pluvialis KOH-catalyzed (NR) Methyl palmitate: 23%, methyl linoleate: 20%, methyl oleate: 19%, methyl linolenate: 16% Iodine value: 111 g I2/100 g Damiani et al. (2010) A consortium of 15 native microalgae Acid-catalyzed followed by base-catalyzed (64%) Methyl linolenate: 28%, methyl linoleate: 20%, methyl palmitate: 16%, methyl oleate: 12% Bound glycerin: 0.0155%, free glycerin: 0.0001% Chinnasamy et al. (2010) Nannochloropsis sp. Supercritical methanol Methyl oleate: 37%, methyl palmitoleate: 23%, methyl palmitate: 8% NC Patil et al. (2012) Chlorella vulgaris Alkaline in situ (78%) Methyl linolenate: 22%, methyl oleate: 21%, methyl stearate: 11% NC Velasquez-Orta et al. (2012) Chlorella vulgaris HCl-catalyzed (NR) Methyl palmitate: 62%, methyl oleate: 20%, methyl linoleate: 10% Density: 0.88 kg/l, viscosity: 4.5 cSt, calorific value: 38.4 MJ/kg, iodine value: 56.2 g I2/ 100 g, acid value: 0.6 mg KOH/g, cetane index: 54.7, ash content: 0.01%, water content: 0.03% Mallick et al. (2012) NR, not reported; NC, not characterized. 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT Name of the Alga 178 Properties of Biodiesel Transesterification Process with % Conversion
WASTE UTILIZATION FOR BIODIESEL PRODUCTION: A CASE STUDY WITH SCENEDESMUS OBLIQUUS FATTY ACID METHYL ESTERS AND FUEL PROPERTIES As stated before, biodiesel is the best substitute for petrodiesel due to its fuel properties, which are very close to those of diesel. Diesel is a mixture of C15 to C18 hydrocarbons obtained from crude oil in the distillation range of 250e350  C. It contains only carbon and hydrogen atoms, which are arranged in straight or branched chain structures, as well as aromatic configurations. Diesel may contain both saturated and unsaturated hydrocarbons. Biodiesel rather has a different chemical structure than the conventional diesel fuel. It is monoalkyl esters of long-chain fatty acids derived from various types of vegetable oils. The fatty acids are of C12 to C24, with over 90% of them being between C16 and C18. Fuel properties of biodiesel that are influenced by the fatty acid profile and, in turn, by the structural features of various fatty acid esters is CN, which ultimately affects the exhaust emission, heat of combustion, cold flow, oxidative stability, viscosity, and lubricity. Structural features of a fatty acid ester molecule that influence the physical and fuel properties are chain length and degree of unsaturation (Knothe, 2005). Since biodiesel is produced in quite differently scaled plants from vegetable oils of varying origin and quality, it is necessary to install a standardization of fuel quality to guarantee engine performance without any difficulties. CN is widely used as diesel fuel quality parameter related to the ignition delay time and combustion quality. The higher the CN is the better are the ignition properties (Meher et al., 2006). High CNs ensure good cold start properties and minimize the formation of white smoke. The longer the fatty acid carbon chains and the more saturated the molecules are, the higher are the CNs (Bajpai and Tyagi, 2006). According to Knothe and Dunn. (2003), high CNs are observed for esters of saturated fatty acids such as palmitic and stearic acids. The oxidation stability decreased with increase in the contents of polyunsaturated FAMEs (Ramos et al., 2009). The limitation of unsaturated fatty acids is also necessary due to the fact that heating of higher unsaturated fatty acids results in polymerization of glycerides. This can lead to the formation of heavy deposits in the machines (Mittelbach, 1996). One of the major problems associated with the use of biodiesel is its poor cold flow property, indicated by relatively high cloud point and pour point. Saturated fatty acids have significantly higher melting points and crystallize even at room temperature. Thus biodiesel produced from the sources with high amounts of saturated fats would show higher cloud points and pour points. Viscosity also increases with the increasing degree of saturation and chain length (Knothe, 2005). Unsaturated fatty acids exhibit better lubricity than 179 saturated ones (Kenesey and Ecker, 2003). Heat of combustion increases with the chain length and decreases with unsaturation (Goering et al., 1982). The increase in heat content results from a gross increase in number of carbon and hydrogen as well as increase in the ratio of these elements relative to oxygen. Therefore no single fatty acid could fulfill every fuel properties. Rather, a very good compromise can be reached by considering a fuel rich in the monounsaturated fatty acids, such as oleate or palmitoleate, and low in both saturated and polyunsaturated fatty acids (Durrett et al., 2008). WASTE UTILIZATION FOR BIODIESEL PRODUCTION: A CASE STUDY WITH SCENEDESMUS OBLIQUUS IN A RECIRCULATORY AQUACULTURE SYSTEM Nowadays, waste disposal is a worldwide problem. In agricultural countries like India, waste discharges from agriculture, agrobased industries and city sewages are the main sources of water pollution. Conventional wastewater treatment systems do not seem to be the definitive solution to pollution and eutrophication problems. The major drawbacks are cost and lack of nutrient recycling (Eisenberg et al., 1981). Secondary sewage treatment plants are specifically designed to control the quantity of organic compounds in wastewaters. Other pollutants including nitrogen and phosphorus are only slightly affected by this type of treatment (Gates and Borchardt, 1964). Owing to the ability to use nitrogen and phosphorus for growth, algae can successfully be cultivated in such type of wastewaters (Mallick, 2002). This has been evolved from the early work of Oswald (Oswald et al., 1953) using microalgae in tertiary treatment of municipal wastewaters. The widely used microalgae cultures for nutrient removal are Chlorella (González et al., 1997; Lee and Lee, 2001), Scenedesmus (Martinez et al., 1999, 2000) and Spirulina (Olguı́n et al., 2003). Nutrient removal efficiency of Nannochloris sp. (Jimenez-Perez et al., 2004), B. brauinii (An et al., 2003), and Phormidium sp. (Dumas et al., 1998; Laliberte et al., 1997) has also been investigated. One of the well-known algae-based bioprocesses for wastewater treatment is high-rate algal ponds (Cromar et al., 1996; Deviller et al., 2004). Recently, corrugated raceways (Craggs et al., 1997; Olguı́n et al., 2003), triangular photobioreactors (Dumas et al., 1998), and tubular photobioreactors (Briassoulis et al., 2010; Molina et al., 2000) have been developed for nutrient removal. Among agroindustries, a large quantity of wastewater is generated from intensified aquaculture practices. The main source of potentially polluting waste in fish culture is feed derived, mainly unconsumed and undigested feed and fish excreta. Discharging these
180 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT effluents directly into water resources causes eutrophication of the receiving waters. Qian et al. (1996) reported the collapse of a prawn industry in China due to outbreak of pathogenic bacteria caused by high nutrient load. A few studies have shown the efficiency of algae biofilters in removing nitrogen from fish effluents (Cohen and Neori, 1991; Jimenez del Rio et al., 1996; Schuenhoff et al., 2003). These works are based on the use of seaweeds of the genera Ulva and Gracilaria to treat effluent water from aquaculture. Recently, we intend to explore an integrated approach to produce biodiesel with simultaneous waste recycling by a green microalga S. obliquus with three types of wastes, viz. poultry litter (PL), fish pond discharge (FPD), and municipal secondary settling tank discharge. Our initial trial under laboratory batch culture conditions (Mandal and Mallick, 2011) encouraged us to conduct a small-scale field experiment in a recirculatory aquaculture system (RAS) using FPD and PL with the same microalga (Mandal and Mallick, 2012). Figure 11.1 presents a schematic diagram of RAS, developed at Agricultural and Food Engineering Department, Indian Institute of Technology Kharagpur, West Bengal, India. The effluent from a fish pond was pumped into a settling tank for removal of large solids. After 24 h, the supernatant was siphoned to an inclined plate settler for removal of fine solids. To have a clear picture of the inclined plate settler readers are requested to refer Sarkar et al. (2007). The effluent was then entered into fiber-reinforced plastic tanks (length 125 cm, breadth 60 cm, depth 45 cm) for culturing the test microalga. FPD has a very high load of solid particles in suspension, which contributes to increase in turbidity. Experiments carried on with sedimented and nonsedimented FPD showed that the nutrient removal efficiency of S. obliquus was higher in the sedimented one. Further experiments with sedimented FPD demonstrated that biomass and lipid yield was maximum at 15 cm culture depth with stirring. In seasonal variation study, the maximum algal biomass and lipid productivity was recorded during summer when sunshine hour was relatively large. During the summer season, when S. obliquus cultures pregrown in FPD supplemented with 5 g PL/l were transferred to the optimized conditions to maximize the lipid accumulation (to have details on optimized condition readers are requested to refer Mandal and Mallick, 2009), lipid yield was raised by more than sixfold (up to 780 mg/l, Mandal and Mallick, 2012). During rainy and winter seasons, comparable lipid yield was recorded by providing artificial lights for few hours. Thus an areal lipid productivity of 14,000 l/ha year (approximately) has been projected assuming 11 cultivation cycles per year, leaving the rest of the period for cleaning and maintenance of the system (Mandal and FP : fish pond CP : centrifugal pump ST : settling tank IPS : inclined plate settler ACT : algae culture tank RWT : remediated water tank ST RWT FIGURE 11.1 Diagrammatic representation of recirculatory aquaculture system (RAS).
REFERENCES Mallick, 2012). Nevertheless, this value is w8 times higher than that of Jatropha, one of the most acclaimed energy crops (Khan et al., 2009). 181 a project (Studies on Microalgal Triacylglycerols as a Source of Biodiesel) to continue research efforts in this exciting and imminent field. References CONCLUDING REMARKS Chisti (2007) envisioned a lipid productivity of 58,700e136,900 l/ha year, considering lipid content of 30e70% of dry biomass. However, to attain this, increasing the volumetric and areal production rates should be the focus (Grobbelaar, 2012). Grobbelaar (2009) projected the upper limits of biomass productivity of about 200 g (dcw)/m2 day. Six decades of worldwide research on outdoor mass cultivation of microalgae, however, have demonstrated only a diminutive fraction of this, where the highest value being recorded was 30 g (dcw)/m2 day (Lee, 2001). As opined by Grobbelaar (2012), with the available strains, an average long-term rate close to 50 g (dcw)/m2 day could be attainable by optimizing various conditions, such as culture depth, mixing, nutrients and CO2 supply, temperature and light, and controlling predators, pathogens and alien microalgal invasion in open raceways. This equates to an annual productivity of about 150 tons of algal biomass per hectare. Thus, maximum lipid productivity would vary between 50,000 and 84,000 l/ha year, considering lipid content of 30e50% of dry biomass. For the last few years, there has been a worldwide impetus to achieve commercial-scale production of biodiesel from microalgae. In October 28, 2010, US Naval base in Norfolk, Virginia, completed a successful test by running a 15 m gunboat with 50:50 mix of algaebased fuel and diesel. However, the cost of this mix was $112 for liter. In March 2012, the US Navy put another milestone by sailing a fleet ship w1200 miles on “Soladiesel”, an algae-based fuel blend. Recently, researchers at Brookhaven National Laboratory, USA, have announced the development of a new process that could significantly lower the cost. Nevertheless, cost of producing microalgal biodiesel can be reduced further by using a biorefinery-based production strategy, like a petroleum refinery, where each and every component is used to produce valuable products. There is much to be researched in this exciting and upcoming field, and the only thing we can say for certain is that the best method/technology of biofuel production will survive and rise above the others. It is our job as researchers to find this, and be proud to be a part of this endeavor. Acknowledgments Nirupama Mallick is thankful to NFBSFARA, Indian Council of Agricultural Research, New Delhi, India, for financial support in the form of Alcantara, R., Amores, J., Canoira, L., Fidalgo, E., Franco, M.J., Navarro, A., 2000. Catalytic production of biodiesel from soybean oil, used frying oil and tallow. Biomass Bioenergy 18, 515e527. An, J.Y., Sim, S.J., Lee, J.S., Kim, B.K., 2003. Hydrocarbon production from secondarily treated piggery wastewater by the green algae Botryococcus braunii. J. Appl. Phycol. 15, 185e191. Aresta, M., Dibenedetto, A., Carone, M., Colonna, T., Fagale, C., 2005. Production of biodiesel from macroalgae by supercritical CO2 extraction and thermochemical liquifaction. Environ. Chem. Lett. 3, 136e139. Bajpai, D., Tyagi, V.K., 2006. Biodiesel: source, production, composition, properties and its benefits. J. Oleo Sci. 55, 487e502. Behrens, P.W., Kyle, D.J., 1996. Microalgae as a source of fatty acids. J. Food Lipids 3, 259e272. Bhattacharyya, S., Reddy, C.S., 1994. Vegetable oils as fuel for internal combustion engine: a review. J. Agric. Eng. Res. 57, 157e166. Bouvier-Nave, P., Benveniste, P., Oelkers, P., Sturley, S.L., Schaller, H., 2000. Expression in yeast and tobacco of plant cDNAs encoding acyl CoA: diacylglycerol acyltransferase. Eur. J. Biochem. 267, 85e96. Briassoulis, D., Panagakis, P., Chionidis, M., Tzenos, D., Lalos, A., Tsinos, C., Berberidis, K., Jacobsen, A., 2010. An experimental helical-tubular photobioreactor for continuous production of Nannochloropsis sp. Bioresour. Technol. 101, 6768e6777. Brown, A.C., Knights, B.A., Conway, E., 1969. Hydrocarbon content and its relationship to physiological state in the green alga Botryococcus braunii. Phytochemicals 8, 543e547. Brown, L., Zeiler, K.G., 1993. Aquatic biomass and carbon dioxide trapping. Energy Convers. Manage. 34, 1005e1013. Campbell, M.N., 2008. Biodiesel: algae as a renewable source for liquid fuel. Guelph Eng. J. 1, 1916-1107. Canakci, M., Gerpan, J.V., 1999. Biodiesel production via acid catalysis. Trans. ASAE 42, 1203e1210. Chinnasamy, S., Bhatnagar, A., Hunt, R.W., Das, D.C., 2010. Microalgae cultivation in a wastewater dominated by carpet mill effluents for biofuel applications. Bioresour. Technol. 101, 3097e3105. Chisti, Y., 2008. Biodiesel from microalgae beats bioethanol. Trends Biotechnol. 26, 126e131. Chisti, Y., 2007. Biodiesel from microalgae. Biotechnol. Adv. 25, 294e306. Chiu, S.Y., Kao, C.Y., Tsai, M.T., Ong, S.C., Chen, C.H., Lin, C.S., 2009. Lipid accumulation and CO2 utilization of Nannochloropsis oculata in response to CO2 aeration. Bioresour. Technol. 100, 833e838. Cohen, I., Neori, A., 1991. Ulva lactuca biofilters for marine fishpond effluents: 1. ammonia uptake kinetics and nitrogen content. Bot. Mar. 34, 475e482. Converti, A., Casazza, A.A., Ortiz, E.Y., Perego, P., Del Borghi, M., 2009. Effect of temperature and nitrogen concentration on the growth and lipid content of Nannochloropsis oculata and Chlorella vulgaris for biodiesel production. Chem. Eng. Process 48, 1146e1151. Courchesne, N.M.D., Parisien, A., Wang, B., Lan, C.Q., 2009. Enhancement of lipid production using biochemical, genetic and transcription factor engineering approaches. J. Biotechnol. 141, 31e41. Craggs, R.J., McAuley, P.J., Smith, V.J., 1997. Wastewater nutrient removal by marine microalgae grown on corrugated raceway. Water Res. 31, 1701e1707. Cromar, N.J., Fallowfield, H.J., Martin, N.J., 1996. Influence of environmental parameters on biomass production and nutrient
182 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT removal in a high rate algal pond operated by continuous culture. Water Sci. Technol. 34, 133e140. Damiani, M.C., Cecilia, A.P., Diana, C., Patricia, I.L., 2010. Lipid analysis in Haematococcus pluvialis to assess its potential use as a biodiesel feedstock. Bioresour. Technol. 101, 3801e3807. Davis, M.S., Solbiati, J., Cronan, J.E., 2000. Overproduction of acetylCoA carboxylase activity increases the rate of fatty acid biosynthesis in Escherichia coli. J. Biol. Chem. 275, 28593e28598. Dayananda, C., Sarada, R., Usha Rani, M., Shamala, T.R., Ravishankar, G.A., 2007. Autotrophic cultivation of Botryococcus braunii for the production of hydrocarbons and exopolysaccharides in various media. Biomass Bioenergy 31, 87e93. Deviller, G., Aliaume, C., Nava, M.A.F., Casellas, C., Blancheton, J.P., 2004. High-rate algal pond treatment for water reuse in an integrated marine fish recirculating system: effect on water quality and sea bass growth. Aquaculture 235, 331e344. Dorado, M.P., Ballesteros, E., Almeida, J.A., Schellert, C., Lohrlein, H.P., Krause, R., 2002. An alkali-catalyzed transesterification process for high free fatty acid waste oils. Trans. ASAE 45, 525e529. Dumas, A., Laliberté, G., Lessard, P., de la Noüe, J., 1998. Biotreatment of fish farm effluents using the cyanobacterium Phormidium bohneri. Aqua. Eng. 17, 57e68. Dunahay, T.G., Jarvis, E.E., Roessler, P.G., 1995. Genetic transformation of the diatoms Cyclotella cryptica and Navicula saprophila. J. Phycol. 31, 1004e1012. Dunahay, T.G., Jarvis, E.E., Dais, S.S., Roessler, P.G., 1996. Manipulation of microalgal lipid production using genetic engineering. Appl. Biochem. Biotechnol. 57, 223e231. Durrett, T., Benning, C., Ohlrogge, J., 2008. Plant triacylglycerols as feedstocks for the production of biofuels. Plant J. 54, 593e607. Eisenberg, M., Koopman, B., Benemann, J.R., Oswald, W.L., 1981. Algal bioflocculation and energy conservation in microalgal sewage ponds. Biotechnol. Bioeng. Symp. 11, 429e448. Feng, D., Chen, Z., Xue, S., Zhang, W., 2011b. Increased lipid production of the marine oleaginous microalgae Isochrysis zhangjiangensis (Chrysophyta) by nitrogen supplement. Bioresour. Technol. 102, 6710e6716. Feng, P., Deng, Z., Hu, Z., Fan, L., 2011a. Lipid accumulation and growth of Chlorella zofingiensis in flat plate photobioreactors outdoors. Bioresour. Technol. 102, 10577e10584. Francis, G., Becker, K., 2002. Biodiesel from Jatropha plantations on degraded land. University of Hohenheim, Stuttgart, Germany p 9. http://www.youmanitas.nl/pdf/Bio-diesel.pdf. Gao, C., Zhai, Y., Ding, Y., Wu, Q., 2010. Application of sweet sorghum for biodiesel production by heterotrophic microalga Chlorella protothecoides. Appl. Energy 87, 756e761. Gates, W.E., Borchardt, J.A., 1964. Nitrogen and phosphorus extraction from domestic waste water treatment plant effluents by controlled algal culture. Res. J. Water Pollut. Control Fed. 36, 443e462. Ghadge, S.V., Raheman, H., 2006. Process optimization for biodiesel production from mahua (Madhuca indica) oil using response surface methodology. Bioresour. Technol. 97, 379e384. Goering, C.E., Schwab, A.W., Daugherty, M.J., Pryde, E.H., Heakin, A.J., 1982. Fuel properties of eleven vegetable oils. Trans. ASAE 25, 1472e1477. González, L.E., Cañizaresb, R.O., Baenaa, S., 1997. Efficiency of ammonia and phosphorus removal from a colombian agroindustrial wastewater by the microalgae Chlorella vulgaris and Scenedesmus dimorphus. Bioresour. Technol. 60, 259e262. Gouveia, L., Oliveira, A.C., 2009. Microalgae as a raw material for biofuels production. J. Ind. Microbiol. Biotechnol. 36, 269e274. Gouveia, L., Marques, A.E., da Silva, T.L., Reis, A., 2009. Neochloris oleoabundans UTEX #1185: a suitable renewable lipid source for biofuel production. J. Ind. Microbiol. Biotechnol. 36, 821e826. Grobbelaar, J.U., 2009. Upper limits of photosynthetic productivity and problems of scaling. J. Appl. Phycol. 21, 519e522. Grobbelaar, J.U., 2012. Microalgae mass culture: the constraints of scaling-up. J. Appl. Phycol. 24, 315e318. Haag, A.L., 2007. Algae bloom again. Nature 447, 520e521. Hankamer, B., Lehr, F., Rupprecht, J., Mussgnug, J.H., Posten, C., Kruse, O., 2007. Photosynthetic biomass and H2 production by green algae: from bioengineering to bioreactor scale-up. Physiol. Plant. 131, 10e21. Hossain, A.B.M.S., Salleh, A., 2008. Biodiesel fuel production from algae as renewable energy. Am. J. Biochem. Biotechnol. 4, 250e254. Illman, A.M., Scragg, A.H., Shales, S.W., 2000. Increase in Chlorella strains calorific values when grown in low nitrogen medium. Enzyme Microb. Technol. 27, 631e635. International Energy Agency, 2007. World Energy Outlook. Executive Summary, China and India insight. Paris, France, pp.14. Jain, S., Sharma, M.P., 2010. Prospects of biodiesel from Jatropha in India: a review. Renewable Sustainable Energy Rev. 14, 763e771. Jako, C., Kumar, A., Wei, Y., Zou, J., Barton, D.L., Giblin, E.M., Covello, P.S., Taylor, D.C., 2001. Seed-specific over-expression of an Arabidopsis cDNA encoding a diacylglycerol acyltransferase enhances seed oil content and seed weight. Plant Physiol. 126, 861e874. Jiang, Y., Yoshida, T., Quigg, A., 2012. Photosynthetic performance, lipid production and biomass composition in response to nitrogen limitation in marine microalgae. Plant Physiol. Biochem. 54, 70e77. Jimenez del Rio, M., Ramazanov, Z., Garcia-Reina, G., 1996. Ulva rigida (Ulvales, Chlorophyta) tank culture as biofilters for dissolved inorganic nitrogen from fishpond effluents. Hydrobiologia 326, 61e66. Jimenez-Perez, M.V., Sanches-Castillo, P., Romera, O., FernandezMoreno, D., Perez-Martinez, C., 2004. Growth and nutrient removal in free and immobilized planktonic green algae isolated from pig manure. Enzyme Microb. Technol. 34, 392e398. Johnson, M.B., Wen, Z., 2009. Production of biodiesel fuel from the microalga Schizochytrium limacinum by direct transesterification of algal biomass. Energy Fuels 23, 5179e5183. Kenesey, E., Ecker, A., 2003. Oxygen bond to improve the lubricity of fuel. Tribol. Schmierungstech. 50, 21e26. Khan, S.A., Rashmi, M.Z., Hussain, S., Prasad, Banerjee, U.C., 2009. Prospects of biodiesel production from microalgae in India. Renewable Sustainable Energy Rev. 13, 2361e2372. Khozin-Goldberg, I., Cohen, Z., 2006. The effect of phosphate starvation on the lipid and fatty acid composition of the fresh water eustigmatophyte Monodus subterraneus. Phytochemicals 67, 696e701. Kilham, S.S., Kreeger, D.A., Gulden, C.E., Lynn, S.G., 1997. Effects of nutrients limitation on biochemical constituents of Ankistrodesmus falcatus. Freshwater Biol. 38, 591e596. Klaus, D., Ohlrogge, J.B., Neuhaus, H.E., Dormann, P., 2004. Increased fatty acid production in potato by engineering of acetyl-CoA carboxylase. Planta 219, 389e396. Knothe, G., 2005. Dependence of biodiesel fuel properties on the structure of fatty acid alkyl esters. Fuel Process. Technol. 86, 1059e1070. Knothe, G., Dunn, R.O., 2003. Oxidative stability of biodiesel/jet fuel blends by oil stability index (OSI) analysis. J. Am. Oil Chem. Soc. 80, 1047e1048. Krawczyk, T., 1996. Biodiesel alternative fuel makes inroads but hurdles remain. Inform 7, 801e829. Kumar, A., Kumar, K., Kaushik, N., Sharma, S., Mishra, S., 2010. Renewable energy in India: current status and future potentials. Renewable Sustainable Energy Rev. 14, 2434e2442. Laliberte, G., Lessard, P., de la Noüe, J., Sylvestre, S., 1997. Effect of phosphorus addition on nutrient removal from wastewater with the cyanobacterium Phormidium bohneri. Bioresour. Technol. 59, 227e233.
REFERENCES Lang, X., Dalai, A.K., Bakhshi, N.N., Reany, M.J., Hertz, P.B., 2001. Preparation and characterization of biodiesel from various bio-oils. Bioresour. Technol. 80, 53e62. Lee, Y.-K., 2001. Microalgal mass culture systems and methods: their limitation and potential. J. Appl. Phycol. 13, 307e315. Lee, K., Lee, C.G., 2001. Effect of light/dark cycles on wastewater treatments by microalgae. Biotechnol. Bioprocess Eng. 6, 194e199. Li, X., Xu, H., Wu, Q., 2007. Large-scale biodiesel production from microalga Chlorella protothecoides through heterotrophic cultivation in bioreactors. Biotechnol. Bioeng. 98, 764e771. Li, Y., Horsman, M., Wu, N., Lan, C.Q., Dubois-Calero, N., 2008. Biofuels from microalgae. Biotechnol. Prog. 24, 815e820. Lin, H., Castro, N.M., Bennett, G.N., San, S.Y., 2006. Acetyl-CoA synthetase overexpression in Escherichia coli demonstrates more efficient acetate assimilation and lower acetate accumulation: a potential tool in metabolic engineering. Appl. Microbiol. Biotechnol. 71, 870e874. Liu, J., Hua, W., Zhan, G., Wei, F., Wang, X., Liu, G., Wang, H., 2010. Increasing seed mass and oil content in transgenic Arabidopsis by the overexpression of wri1-like gene from Brassica napus. Plant Physiol. Biochem. 48, 9e15. Lu, X., Vora, H., Khosla, S., 2008. Over production of free fatty acids in E. coli: implications for biodiesel production. Metab. Eng. 10, 333e339. Mallick, N., 2002. Biotechnological potential of immobilized algae for wastewater N, P and metal removal: a review. BioMetals 15, 377e390. Mallick, N., Mandal, S., Singh, A.K., Bishai, M., Dash, A., 2012. Green microalga Chlorella vulgaris as a potential feedstock for biodiesel. J. Chem. Technol. Biotechnol. 87, 137e145. Mandal, S., Mallick, N., 2009. Microalga Scenedesmus obliquus as a potential source for biodiesel production. Appl. Microbiol. Biotechnol. 84, 281e291. Mandal, S., Mallick, N., 2011. Waste utilization and biodiesel production by the green microalga Scenedesmus obliquus. Appl. Environ. Microbiol. 77, 374e377. Mandal, S., Mallick, N., 2012. Biodiesel production by the green microalga Scenedesmus obliquus in a recirculatory aquaculture system. Appl. Environ. Microbiol. 78, 5929e5934. Marsh, G., 2009. Small wonders: biomass from algae. Renewable Energy Focus 9, 74e78. Martinez, M.E., Castillo, J.M., El Yousfi, F., 1999. Photoautotrophic consumption of phosphorus by Scenedesmus obliquus in a continuous culture. Influence of light intensity. Process Biochem. 34, 811e818. Martinez, M.E., Sanchez, S., Jimenez, J.M., El Yousfi, F., Munoz, L., 2000. Nitrogen and phosphorus removal from urban wastewater by the microalga Scenedesmus obliquus. Bioresour. Technol. 73, 263e272. McMillen, S., Shaw, P., Jolly, N., Goulding, B., Finkle, V., 2005. Biodiesel: Fuel for Thought, Fuel for Connecticut’s Future. Connecticut Center for Economic Analysis Report. University of Connecticut, pp. 56. Meher, L.C., Vidya Sagar, D., Naik, S.N., 2006. Technical aspects of biodiesel production by transesterification - a review. Renewable Sustainable Energy Rev. 10, 248e268. Miao, X., Wu, Q., 2004. High yield bio-oil production from fast pyrolysis by metabolic controlling of Chlorella protothecoides. J. Biotechnol. 110, 85e93. Miao, X., Wu, Q., 2006. Biodiesel production from heterotrophic microalgal oil. Bioresour. Technol. 97, 841e846. Mittelbach, M., 1990. Lipase-catalyzed alcoholysis of sunflower oil. J. Am. Oil Chem. Soc. 67, 168e170. Mittelbach, M., 1996. Diesel fuel derived from vegetable oils, VI: specifications and quality control of biodiesel. Bioresour. Technol. 27, 435e437. 183 Molina, E., Fernández, A.F.G., Camacho, G.F., Rubio, C.F., Chisti, Y., 2000. Scale-up of tubular photobioreactors. J. Appl. Phycol. 12, 355e368. Olguı́n, E.J., Galicia, S., Mercado, G., Perez, T., 2003. Annual productivity of Spirulina (Arthrospira) and nutrient removal in a pig wastewater recycle process under tropical conditions. J. Appl. Phycol. 15, 249e257. Oswald, W.J., Gotaas, H.B., Ludwig, H.F., Lynch, W., 1953. Algae symbiosis in oxidation ponds: III. phototosynthetic oxygenation. Sewage Ind. Wastes 25, 692e704. Patil, P.D., Gude, V.G., Mannarswamy, A., Deng, S., Cooke, P., Munson-McGee, S., Rhodes, I., Lammers, P., Nirmalakhandan, N., 2012. Optimization of direct conversion of wet algae to biodiesel under supercritical methanol conditions. Bioresour. Technol. 101, 118e122. Piorreck, M., Pohl, P., 1984. Formation of biomass, total protein, chlorophylls, lipids and fatty acids in green and blue-green algae during one growth phase. Phytochemicals 23, 217e223. Planning Commission, 2003. Report of the Committee on Development of Biofuel. Government of India, New Delhi, India, pp. 140. Planning Commission, 2007. Eleventh Five Year Plan 2007e2012, Energy Sector. Government of India, New Delhi, India, pp. 65. Qian, P., Wu, C.Y., Wu, M., Xie, Y.K., 1996. Integrated cultivation of the red alga Kappaphycus alvarezii and the pearl oyster Pinctada martensi. Aquaculture 147, 21e35. Ramos, M.J., Fernández, C.M., Casas, A., Rodrı́guez, L., Pérez, Á., 2009. Influence of fatty acid composition of raw materials on biodiesel properties. Bioresour. Technol. 100, 261e268. Rao, P.P., Gopalkrishnan, K.V., 1991. Vegetable oils and their methyl esters as fuels diesel engines. Indian J. Technol. 29, 292e297. Ratledge, C., 2002. Regulation of lipid accumulation in oleaginous micro-organisms. Biochem. Soc. Trans. 30, 1047e1050. Rattanapoltee, P., Chulalaksananukul, W., James, A.E., Kaewkannetra, P., 2008. Comparison of autotrophic and heterotrophic cultivations of microalgae as a raw material for biodiesel production. J. Biotechnol. 136, S412. Rodolfi, L., Zittelli, G., Bassi, N., Padovani, G., Bonini, N., Bonini, G., Tredici, M.R., 2009. Microalgae for oil: strain selection, induction of lipid synthesis and outdoor mass cultivation in a low-cost photobioreactor. Biotechnol. Bioeng. 102, 100e112. Roesler, K., Shintani, D., Savage, L., Boddupalli, S., Ohlrogge, J., 1997. Targeting of the Arabidopsis homomeric acetyl-coenzyme A carboxylase to plastids of rapeseeds. Plant Physiol. 113, 75e81. Roessler, P.G., 1990. Purification and characterization of acetyl-CoA carboxylase from the diatom Cyclotella cryptica. Plant Physiol. 92, 73e78. Russa, L.M., Bogen, C., Uhmeyer, A., Doebbe, A., Filippone, E., Kruse, O., Mussgnug, J.H., 2012. Functional analysis of three type2 DGAT homologue genes for triacylglycerol production in the green microalga Chlamydomonas reinhardtii. J. Biotechnol. 162, 13e20. Sarkar, S., Kamilya, D., Mal, B.C., 2007. Effect of geometric and process variables on the performance of inclined plate settlers in treating aquacultural waste. Water Res. 41, 993e1000. Schneider, D., 2006. Grow your own? Would the widespread adoption of biomass-derived transportation fuels really help the environment. Am. Sci. 94, 408e409. Schuenhoff, A., Shpigel, M., Lupatsch, I., Ashkenazi, A., Msuya, F.E., Neori, A., 2003. A semi-recirculating, integrated system for the culture of fish and seaweed. Aquaculture 221, 167e181. Scragg, A.H., Illman, A.M., Carden, A., Shales, S.W., 2002. Growth of microalgae with increased calorific values in a tubular bioreactor. Biomass Bioenergy 23, 67e73. Shay, E., 1993. Diesel fuel from vegetable oils: status and opportunities. Biomass Bioenergy 4, 227e242.
184 11. MICROALGAE: THE TINY MICROBES WITH A BIG IMPACT Sheehan, J., Dunahay, T., Benemann, J., Roessler, P., 1998. A Look Back at the U.S. Department of Energy’s Aquatic Species Program Biodiesel from Algae. National Renewable Energy Laboratory Report. U.S. Department of Energy, pp. 323. Singh, I., Rastogi, V., 2009. Performance analysis of a modified 4-stroke engine using biodiesel fuel for irrigation purpose. Int. J. Environ. Sci. 4, 229e242. Sobczuk, T.M., Chisti, Y., 2010. Potential fuel oils from the microalga Choricystis minor. J. Chem. Technol. Biotechnol. 85, 100e108. Takagi, M., Karseno, Yoshida, T., 2006. Effect of salt concentration on intracellular accumulation of lipids and triacylglyceride in marine microalgae Dunaliella cells. J. Biosci. Bioeng. 101, 223e226. Takagi, M., Watanabe, K., Yamaberi, K., Yoshida, T., 2000. Limited feeding of potassium nitrate for intracellular lipid and triglyceride accumulation of Nannochloris sp. UTEX LB1999. Appl. Microbiol. Biotechnol. 54, 112e117. Umdu, E.S., Tuncer, M., Seker, E., 2009. Transesterification of Nannochloropsis oculata microalga’s lipid to biodiesel on Al2O3 supported CaO and MgO catalysts. Bioresour. Technol. 100, 2828e2831. Velasquez-Orta, S.B., Lee, J.G.M., Harvey, A., 2012. Alkaline in situ transesterification of Chlorella vulgaris. Fuel 94, 544e550. Xu, H., Miao, X., Wu, Q., 2006. High quality biodiesel production from a microalgae Chlorella protothecoides by heterotrophic growth in fermenters. J. Biotechnol. 126, 499e507. Zhang, Y., Adams, I.P., Ratledge, C., 2007. Malic enzyme: the controlling activity for lipid production? Overexpression of malic enzyme in Mucor circinelloides leads to a 2.5-fold increase in lipid accumulation. Microbiology 153, 2013e2025. Zhang, Z., Li, M., Agarwal, A., San, K.-Y., 2011. Efficient free fatty acid production in Escherichia coli using plant acyl-ACP thioesterases. Metab. Eng. 13, 713e722. Zou, J., Katavic, V., Giblin, E.M., Barton, D.L., MacKenzie, S.L., Keller, W.A., Hu, X., Taylor, D.C., 1997. Modification of seed oil content and acyl composition in the Brassicaceae by expression of a yeast sn-2 acyltransferase gene. Plant Cell 9, 909e923.
C H A P T E R 12 Biobased Fats (Lipids) and Oils from Biomass as a Source of Bioenergy Ciarán John Forde, Marie Meaney, John Bosco Carrigan, Clive Mills, Susan Boland, Alan Hernon* AER BIO, National Institute for Bioprocessing Research & Training (NIBRT), Blackrock, Co. Dublin, Ireland *Corresponding author email: alan.hernon@aer-bio.com O U T L I N E Introduction 185 Sources of Biolipids Plant-Derived Biolipids Edible Lipids Nonedible Lipids Waste Edible oil Animal-Derived Biolipids Microalgae and Other Oleaginous MicroorganismsDerived Biolipids 186 186 186 187 187 188 Supply and Projected/Purrent Volume 190 Energy Balance 192 Processing of Biolipids and Properties of Biolipid-Derived Biofuels Extraction Steam Distillation Maceration (Solvent Extraction) Enzymatic Hydrolytic Maceration Expression (Cold Pressing) 189 193 193 193 193 193 194 INTRODUCTION Biolipids have been an important source of energy since prehistoric times. While the term “biofuel” is now often synonymously used with “biodiesel”, the first biofuels used were wood or other plant materials, which were burnt to provide heat, light, protection from Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00012-7 Hot Continuous Extraction (Soxhlet) Countercurrent Extraction Ultrasound Extraction (Sonication) Supercritical Fluid Extraction 194 194 194 194 Properties of Pure Plant Oil Degumming Alkaline Neutralization Winterization Bleaching Transesterification 195 195 195 195 196 196 Properties of Biodiesel 196 Biomass to Liquid Fuels (Bio-oil) Gasification Cleaning Process Synthesis 197 197 197 197 Conclusion 198 References 198 predators and for cooking. The earliest lamps recorded were made using plant material that was soaked with animal fat, such as lard. Later lamps, which used oils, were introduced in the eighteenth century, with early lamp fuels being oils from fish, whale and a variety of nut and other plant sources. Whale oil was much sought after for a lamp fuel as it produced a cleaner flame with 185 Copyright Ó 2014 Elsevier B.V. All rights reserved.
186 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY less odor and smoke. Another source of light was candles, which were made from tallow and other oils rendered from animal waste. These fuels are known as primary biofuels, fuels that are used without any significant processing in contrast to secondary biofuels where significant processing is required before the raw products can be used as fuels. As they were discovered, coal, gas and petroleum products (kerosene in particular) slowly replaced tallow and other animal-based fuels. Similarly, the use of biolipids as transport fuel is not novel; in fact, in 1900 when Rudolf Diesel showcased his internal combustion engine at The Exposition Universelle in Paris it was fuelled by peanut oil (Stauffer and Byron, 2007). However, advancements in the use of petroleum as fuel at the turn of the century resulted in the selection of this abundant, cheap and efficient hydrocarbon as the fuel of choice for transport. It was not until the oil crisis of 1973 when oil became expensive and the security of supply became paramount that biolipids were investigated again; however, this interest was short lived as the supply of crude oil from the Organization of Arab Petroleum Exporting counties was restored in 1974. Now over 100 years after Diesel’s invention we are almost completely dependent on this finite, expensive and polluting hydrocarbon (petroleum) as a transport fuel. Consequently, the use of petroleum-based products has resulted in a significant number of environmental issues including global warming via the greenhouse gas (GHG) effect. Also, in an era when it is generally accepted that we have reached peak oil production and it is projected that the demand for transport fuel will increase globally by 39% in the next 10 years interest in the use of biolipids as fuel has reached new heights. Recent years have seen significant research, investment and advances in sustainable energy technologies such as solar, wind, geothermal, tidal and hydroelectrical. It should be noted, however, that these energy sources, along with nuclear power, relate to the generation of electricity. Currently electricity only accounts for about 33% of the world energy market, whereas liquid fuels account for the remaining 67% of global energy consumption. These figures, along with the finite nature of crude oil stocks, illustrate the need to drastically increase the production of sustainable liquid fuels (Schenk et al., 2008). Alternative liquid fuel sources are continually being sought (Bereczky, 2012; Singh and Singh, 2010) and while the obvious solution is to revert to the use of vegetable oil used in 1900, there are several problems with that approach. Most notably is the need to use arable land to feed the world’s exponentially growing population. Land use for the production of liquid biofuels has become a hotly debated topic since 2007 when a combination of poor harvests and allocation of vast quantities of land for the production of biofuel (mostly corn ethanol) resulted in a spike in world food prices (Tenenbaum, 2008). The ease in supply of food to the world market in 2007/2008 acted at an indicator to what will happen in the future as the world’s population increases beyond 8 billion people and we struggle to meet the nutritional needs of humankind. It will simply be impossible to grow enough terrestrial crops to meet the worlds nutritional and energy needs. It is therefore necessary to explore the use of biolipids from all sources including lipids from plant, animal and microalgae sources. Recovering lipids from waste products like recovered vegetable oil and beef tallow will also have a role to play in meeting our insatiable demand for energy. Therefore, it is important to judicially select biolipids that require the minimum land usage (maximizing ton of oil per hectare) and lipids with good fuel properties, as discussed below. In addition, the energy consumed in growing and recovering the biolipid is also an important consideration when selecting a biomass for the production of biofuel. SOURCES OF BIOLIPIDS Biolipids can be derived from plant, animal, oleaginous microorganisms and algal sources. The composition of biolipids derived from each of these sources differs greatly and has varying degrees of suitability to the biofuel production industry. The major lipids produced from each of these sources are listed below and the degree of suitability to the production of biofuel production is discussed. Plant-Derived Biolipids In 2007, 95% of world biodiesel was produced via edible plant oils, which were supplied by the agricultural industry, with the vast majority supplied by rapeseed oil, 84% (Food and Agriculture Organization, 2008). Overall, plant lipids are divided into three major categories: edible, nonedible and waste vegetable oils described below. Edible Lipids The main edible oils used for biofuel production are rapeseed, palm and soy bean oils. Edible oils have the disadvantage of competing directly with food production. The use of edible oils for the production of biodiesel competes directly with the use of land for the production of food and without proper planning results in reduced food production (Gui et al., 2008). However, the productivity from edible oils is high in terms of oil yield and the quality of the resulting biofuel. The oil yield from palm is the highest of the commonly grown edible oil crops at 5 tons per hectare while rapeseed
SOURCES OF BIOLIPIDS produces 1 ton per hectare and soy bean 0.52 tons per hectare. A high lipid yield is vital for the economical production of biofuel from these plants. Although the productivity from palm oil is particularly high its use as a biofuel is limited as it is the world’s most commonly used edible lipid and thus competition for the oil between the food and biofuel industry would result in an increase in the price of this oil (Lam et al., 2009). In terms of the suitability for biofuel, palm oil has a high degree of saturation and thus is not the most suitable for biofuel production with the resulting fuel having poor cold flow properties. However, the cold flow properties of a lipid can be altered by the use of cold filtration (Kerschbaum et al., 2008) or alternatively the use of alcohols such as ethanol, isopropanol or isobutanol, which results in the production of fatty alkyl esters with lower freezing points and therefore improved cold flow properties (Dunn, 2009). There are also some environmental and ecological concerns surrounding palm oil production, with the clearing of rain forests to make way for palm plantations. The plantation costs of edible oil crops are relatively low with the exception of palm oil, which has a higher cost; however, this is offset by the high oil yield from the crop. The overall estimated energy balance of rapeseed and soybean is similar at 3.7 and 3.4, respectively, while palm oil is significantly higher at 9.6 due to the high yields (Food and Agriculture Organization, 2008). Currently rapeseed oil is the most commonly used plant oil used in biodiesel production because it makes an excellent biofuel with excellent cold flow properties. The main disadvantage of using rapeseed oil is the growth of rapes is difficult and unsustainable as it must be part of a one in five rotation due to the large quantity of nutrients required for the growth of the organism and the buildup of pathogens and disease in the environment targeting rapeseed if grown annually. Nonedible Lipids Nonedible oils that may be used in biofuel production include Jatropha, Pongamia, jojoba, linseed and cotton seed oil. Nonedible oils are not suitable for human consumption due to the presence of toxic compounds in the oils, for example, curicin present in Jatropha oils is a toxic lectin. Biofuels from nonedible lipids have many advantages over the edible alternative including the ability of these organisms to grow in harsh nutrientand moisture-limiting conditions and the reduction in carbon emissions. Nonedible oils are generally more cost-effective as they do not have applications in food production and thus are lower value oils, containing low sulfur concentrations and low aromatic compound concentrations and the lipids produced are biodegradable (No, 2011). A disadvantage of using nonedible oils 187 is the large amounts of free fatty acids (FFAs) that cannot be converted into biodiesel using an alkaline catalyst (Demirbas et al., 2011). Jatropha is one of the most widely used nonedible oils due to the high potential yield of 0.5e12 tons per hectare per year; the yield is highly effected by the conditions in which it is produced, and the ability of the organism to grow in harsh environmental conditions of low water availability and low nutrient content (Francis et al., 2005). The oil produced by Jatropha has good cold flow properties due to the composition of the oil. The Jatropha plant is a small tree and produces seeds with high lipid content. In addition to the drought resistance within the plant it is also pest tolerant and unpalatable to animals and grows rapidly with a lifetime of 30 years; each of these factors makes it a suitable choice for the production of biofuels. The ability of the plant to grow in harsh conditions led to Jatropha being considered a revolutionary plant that could provide the solution for the production of large volume of lipids without competing with the food industry. However, when grown in marginal lands studies revealed that the number of seeds produced by the plant was quite low and although the tree is capable of growing in low nutrient conditions, the lipid production is low (Pandey et al., 2012). Therefore, the economic returns of Jatropha grown on marginal lands is low; however, growing the crop in developing areas with poor land may be a viable method of production of oil on a small scale. The energy balance from the crop is also low if only the seeds are used for the production of biofuel; however, the value is increased if all components, for example, the husks are also utilized (Prueksakorn and Gheewala, 2006). Waste Edible oil Waste edible oil (WEO) is the waste product of cooking or frying foods. The disposal of WEO is difficult and thus the use of WEO as a biofuel would both alleviate the problem of disposal in addition to providing a renewable source of biodiesel. WEO has a high volume of FFAs, 0.5e15% in comparison with the 0.5% content of refined virgin vegetable oil, which cannot be converted to biodiesel using an alkaline catalyst as the FFAs undergo a saponification reaction with the catalyst thus reducing efficiency and yield (Knothe et al., 2005). The problem may be overcome by using a supercritical methanol transesterification for the transesterification process rather than an alkaline catalyst (Kusdiana and Saka, 2004). The volume of WEOs available is quite high with approximately 1 million tons produced in Europe each year while 10 million tons are produced annually in the United States (Gui et al., 2008). WEO is available two to three times cheaper than virgin vegetable oils (Phan and Phan, 2008) and the high
188 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY volume of WEO available means it a viable method for biodiesel production. WEO has a higher estimated energy balance than rapeseed and soybean of 5.8; however, the value is lower than that of palm oil at 9.5 (Food and Agriculture Organization, 2008). Animal-Derived Biolipids As outlined above many biological sources can be used for the generation of biofuels (Demirbas et al., 2011; Vasudevan et al., 2005); however, one source of biomass for the production of biodiesel that is often overlooked is the waste fat from animals (e.g. (Ali et al., 2012; Duku et al., 2011; Feddern, 2011; Panneerselvam et al., 2011; Wisniewski Jr et al., 2010)). Generally three broad categories of waste animal fats are describeddtallow and related raw fats from processing industries, yellow grease from waste cooking oil used to cook, for example, chicken, and brown grease that is obtained from traps used to prevent waste fats and oils being released into the environment. Animal fats can be sourced as room temperature solids or semisolids from a variety of animals and include tallow and suet (cattle and mutton), lard (pigs), schmaltz (poultry especially chicken and goose), duck, fish oil and dairy products (milk, butter) (Jayasinghe and Hawboldt, 2011; Kerihuel et al., 2005; Mrad et al., 2012; Panneerselvam et al., 2011; Wisniewski Jr et al., 2010). It is also possible to reclaim waste animal fats from wastewater (Awad et al., 2012a). Many of the properties of animal fats used put to specific uses have been known for a long time (Andés, 1898; Shahidi and Zhong, 2005). A significant percentage of waste animal fat can be converted to biodiesel using similar techniques to those used for plant oils, the main process being transesterification, described later (Prosková et al., 2009). The triglycerides in animal fats are saturated, compared to unsaturated plant triglycerides, and this has some implications when used as biodiesel. In particular the cloud point, the temperature at which the oils solidify, is higher for animal fats. However, when used as additives to other sources of diesel, for example, 5% or 20% biodiesel (B5 or B20 blends), the high cloud point does not affect the blend overall. Production of biodiesel from waste animal fats has been shown using a variety of methods including a novel, integrated method in which fat from lamb meat is continuously extracted by supercritical CO2 followed by enzymatic production of biodiesel (Schenk et al., 2008). Feedstocks containing high levels of FFAs require an additional preproduction step to convert the FFAs into esters, which can subsequently be converted into biodiesel. Waste sources that contain high levels of FFAs require a separate step (acid catalyzed pretreatment) before the base catalyzed reactions can be used to provide maximal yields of biodiesel (Canakci and Van Gerpen, 2001; Knothe et al., 2005; Popescu and Ionel, 2011). Multistep processes using waste restaurant oil and animal (pig) fat containing high levels of FFAs can achieve high yields of biodiesel of up to 80% by volume on a small scale (Math et al., 2010). Other high FFA content oils, including used cooking oils, rendered animal fat and some inedible plant oils (Mathiyazhagan et al., 2011) can be processed in a similar fashion (Bakir and Fadhil, 2011). The feasibility and sustainability of using waste animal fats as feedstocks for biofuel production has been the subject of many studies in many areas, for example, general studies (Demirbas, 2009; Nigam and Singh, 2011), Australia (Puri et al., 2012), Ghana (Duku et al., 2011), the United States (Groschen, 2002), Brazil (Aranda et al., 2009), Ireland (Thamsiriroj and Murphy, 2010) and Hungary (Lakó et al., 2008). In addition, the use of animal fats from waste tissue may also have environmental benefits, such as being considered as a waste management process and as a fuel source that does not compete with food resources (e.g. soybean), the food versus fuel debate. Table 12.1 shows typical values reported for triglycerides in several animal fats in comparison to values for soy, a commonly used plantderived feedstock. In all cases, waste animal fats contain high levels of the fatty acids that are capable of being converted to methyl esters by transesterification reactions to produce usable biodiesel. From a sustainability point of view an estimate of the total annual US production of animal fats as compared to plantderived oils is shown in Table 12.2. Vegetable oils tend to be produced for human consumption, whereas animal fats form part of a wide group of animal by-products that are rendered into many products that may be used in part for human consumption (e.g. production of gelatin). All animal byproducts, including fats, are coded and classified (Alakangas et al., 2011) according to their intended use and animal fats not intended for human consumption are controlled in the European Union by Regulation (EC) No 1069/2009 and related legislation. Similarly, TABLE 12.1 Percentages of Fatty Acids in Animal Fats Fatty Acid Beef Tallow Pork Lard Chicken Fat Whale Soy Myristic 14:0 1.4e6.3 0.5e2.5 1 Palmitic 16:0 20e37 20e32 25 Palmitoleic 16:1 0.7e8.8 Stearic 18:0 1.7e5 4e8 e 7e12 w10 8 7e18 e 1e3 6e40 5e24 6 Oleic 18:1 26e50 35e62 41 Linoleic 18:2 0.5e5 3e16 18 w5 28e32 w20 1e2 w50
SOURCES OF BIOLIPIDS TABLE 12.2 Total Annual Production of US Fats and Oils Vegetable Oil Production (billion pounds per year) Canola 1.04 Corn 2.49 Cottonseed 0.617 Soybean Sunflower Total Vegetable Oil 19.61 0.731 24.49 Animal Fats (billion pounds per year) Edible Tallow 1.859 Inedible Tallow 3.299 Lard & Grease 1.63 Yellow Grease 1.40 Poultry Fat 1.42 Total Animal Fat 9.61 Source: U.S. Department of Agriculture, 2010; U.S. Census Bureau, 2010. the storage of animal fats for use as fuels also needs to be addressed. The storage of raw animal fat under unsuitable conditions can lead to oxidation and other undesirable chemical and microbial processes that can affect the quality of the final biodiesel product. The stability of the final biodiesel:diesel blend can also be affected by longterm storage under unsuitable conditions, and additives such as antioxidants might be added to improve stability (Geller et al., 2008; Jain and Sharma, 2010). With the advent of Bovine spongiform encephalopathy (BSE) and more specifically Transmissible spongiform encephalopathies (TSE), there is a greater need to monitor human health issues when using waste animal fats for the production of biofuel, at all stages of the production process. The rendering industry recognizes that safe product (fats) can only be supplied if certain standards are adhered to (Woodgate and Van Der Veen, 2004). The raw materials could well have microbial contamination including pathogenic bacteria and possibly prion material (Baribeau et al., 2005; Brown et al., 2007; Bruederle et al., 2008; Greene et al., 2007). There is also concern that prions will survive the rendering process itself (Bruederle et al., 2008). These concerns have in part led to the publication of guidelines for the safe handling and use of biodiesels (National Renewable Energy Laboratory, 2009). Many trials of waste animal fat biodiesel-powered engines have been published (Darunde Dhiraj and Deshmukh Mangesh, 2012; Kleinová et al., 2011; Panneerselvam et al., 2011; Varuvel et al., 2012). One trial using public transport buses (Proc, 2006) showed that the biodiesel does not have any harmful effects on the engines at B5 and B20 mixes and also shows 189 environmental benefit by way of reduced exhaust pollutants. However, there are other potential health and environmental issues in using animal fats as a feedstock for biodiesel production (Greene et al., 2007) and the production of safe biodiesel is in part dependent on a safe feedstock (Woodgate and Van Der Veen, 2004). Finally, the processes involved (e.g. rendering, cleanup, transesterification, etc.) in the production of biodiesel will generate waste that also needs to be assessed (Ellis, 2007). Microalgae and Other Oleaginous Microorganisms-Derived Biolipids Microalgae are a heterogeneous group of organisms consisting of both prokaryotes such as cyanobacteria and eukaryotes such as diatoms (Bacillariophyta), dinoflagelates (Dinophyta), green algae (Chlorophyta), yellow-green algae (Xanthophyta), and red algae (Rhodophyta) (Brennan and Owende, 2010; Hu et al., 2008). Similarly, other oleaginous microorganisms are defined as microorganisms with lipid content in excess of 20%. The number of bacteria that produce lipids that could be used for biodiesel production is very small. As a result, bacteria are mainly used for special lipid production such as Docosahexaenoic acid (DHA). Many yeasts and fungi also produce high quantities of lipid. Yeasts with high lipid content include Candida curvata (58%), Cryptococcus albidus (65%), Lipomyces strakeyi (64%) and Rhodotorula glutinous (72%). Oleaginous fungi include Aspergillus oryzae (57%), Mortierella isabellina (86%), Humicola lanuginose (75%) and Mortierella vinacea (66%) (Meng et al., 2009). In terms of microalgae, species are generally unicellular organisms but there are also a number of simple multicellular organisms that occur as colonial or filamentous groups of cells. Microalgae are capable of autotrophic, heterotrophic and mixotrophic growth. Microalgae populate a wide variety of ecological niches due to a wide range of tolerance for various growth conditions such as availability of nutrients, salinity, pH and temperature (Brennan and Owende, 2010; Gong and Jiang, 2011; Schenk et al., 2008). Currently, microalgae contribute very little biolipid to the overall bioenergy market as full-scale commercialization has yet to be realized. Despite this fact, microalgae remain the feedstock with the greatest potential for supplying future demand for bioenergy in the form of liquid fuels. The idea of using microalgae as a source of biolipids for biofuel is not a new one, however. For example, the Aquatic Species Program was launched in 1978 by what is now known as the National Renewable Energy Laboratory (NREL) with its main focus being, “the production of biodiesel from high lipid-content algae grown in ponds, utilising waste CO2from coal fired power plants” (Sheehan et al., 1998). Over 3000 microalgae strains were initially
190 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY collected, 300 of which were eventually identified as oil rich. When the program was officially closed in 1998 the conclusions were that no “fundamental engineering and economic issues” were identified that would hamper the feasibility of large-scale microalgae culture. The authors noted, however, that total biomass and algal lipids produced were still below “theoretical potential, and the requirements for economic viability” (Sheehan et al., 1998). The economic viability was, of course, based on a time when oil prices in the United States were among their all-time lowest at less than $20 per barrel (adjusted for inflation). Today the average oil price is approximately $100 per barrel and this, along with increased pressure to reduce GHG emissions as well as significant technical advances, has made microalgae-derived biofuels even more relevant to meet current bioenergy demands. SUPPLY AND PROJECTED/PURRENT VOLUME Growing microalgae for biolipid production usually involves a lag phase of growth followed by a stationary phase induced by some sort of “stress” This “stress”, often nitrogen depletion, induces a switch in the metabolism of the microalgae, which encourages the production of storage lipids in the form of triacylglycerides (TAGs) rather than cell division (Meng et al., 2009; Widjaja et al., 2009). Currently microalgae can be grown at industrial scale autotrophically in open raceway ponds (Sapphire Energy, 2013) or closed photobioreactor (PBR) systems (Solix BioSystems, 2013). In addition, many microalgae species have the ability to grow heterotrophically, in closed fermenters, given a suitable carbon source (Solazyme Inc., 2013). Open culture systems, such as race way ponds, are significantly lower cost in terms of capital expenditure. They require greater land area than closed systems and are more prone to contamination by invasive species. Water loss due to evaporation can also be a significant problem when compared to closed systems (Chisti, 2007; Pulz, 2001; Sheehan et al., 1998). Closed systems, on the other hand, such as PBRs or fermenters are by their nature closed and thus less likely to be contaminated. Nutrient concentration can be more easily controlled and water loss through evaporation is negligible. However, some have argued that loss of cooling water, used to control temperature, negates any savings made from using a closed culture system. The tighter control over culture conditions facilitated by a closed culture system, along with more sterile cultures, results in PBRs producing much greater levels of microalgae biomass, when compared to raceway ponds. However, the increased production capability must be offset against the much larger capital cost involved in commissioning and maintaining a closed culture system (Carvalho et al., 2006; Pulz, 2001; Ugwu et al., 2008). Hybrid systems have also been proposed whereby a closed system is used for the log phase production of biomass and the nutrient depleted lag phase is allowed to occur in large raceway ponds. It is hoped that the relatively concentrated inoculation of the raceway ponds will not allow any invasive species to become established (Greenwell et al., 2010; Huntley and Redalje, 2007; Rodolfi et al., 2008). Microalgae present significant potential as a source of biolipids for bioenergy over more traditional sources of biolipids such as palm, soya or Jatropha for a number of reasons. Firstly, the oil content of microalgae as a percentage of the dry weight, shown in Table 12.3, is generally in the range of 20e70%, although levels above 40% are rarely observed (Borowitzka, 1988). Similarly, the potential yield of biolipids and derived biodiesel from microalgae per area far outweighs that of any current oilseed crop. For example, one of the best available studies of large-scale algae cultivation produced 0.1 g/l day or 20e23 g dry weight/m2 day. A conservative lipid content of 30% could therefore yield 24,000 l biodiesel/ha year (Moheimani and Borowitzka, 2006; Schenk et al., 2008). This compares extremely favorably with both Jatropha (1892 l biodiesel/ha year) and oil palm (5950 l biodiesel/ha year) (Schenk et al., 2008). The high potential yield of biodiesel from microalgaederived biolipids is due to a number of factors including the growth rate of microalgae (Scott et al., 2010) all year round production capability (Schenk et al., 2008) and the higher photon conversion efficiency compared to terrestrial plants (Melis, 2009). Unlike algae-derived biofuels, first-generation biofuels directly competed with food crops for arable land sparking the “Food vs Fuel” debate (Gui et al., 2008). Although second-generation fuel crops such as Jatropha can grow on marginal land (Francis et al., 2005), microalgae are capable of growing on nonarable land ensuring competition for land with food crops is significantly reduced. Similarly, in terms of other resource demands, 1 kg of algae biomass requires 1.83 kg of CO2 to grow (Chisti, 2007) and much research has investigated the potential of industrial flue gases as a source of this CO2 (Bilanovic et al., 2009). This possibility of both sequestering excess CO2 from flue gases that would otherwise be released into the atmosphere, while also increasing the growth rate of microalgae to be used for bioenergy, offers both environmental and economic advantages (Pires et al., 2012; Yun et al., 1997). More recently, the apparent “peak phosphorus” problem has been identified whereby phosphorus will become a limiting resource in agriculture. As a result, the potential industrial scale culture of microalgae, which requires a phosphorus
SUPPLY AND PROJECTED/PURRENT VOLUME TABLE 12.3 191 Lipid Content and Biomass Productivity of Biofuel Relevant Algae Species Algae Species Lipid Content (% Dry Weight) Biomass Productivity (g/l day) Botryococcus braunii 25e75 Chlorella protothecoides 15e58 Chlorella emersonii 63 (Gouveia and Oliveira, 2009) Chlorella minutissima 57 (Gouveia and Oliveira, 2009) Chlorella protothecoides 55 (Gouveia and Oliveira, 2009) Chlorella sorokiana 22 (Gouveia and Oliveira, 2009) Chlorella sorokiniana 19e22 Chlorella sp. 28e32 (Chisti, 2007) Chlorella vulgaris 56 (Gouveia and Oliveira, 2009) Chlorococcum sp. 19 Crypthecodinium cohnii 20 (Chisti, 2007) Cylindrotheca sp. 16e37 (Chisti, 2007) Dunaliella bioculata 8 (Gouveia and Oliveira, 2009) Dunaliella primolecta 23 (Chisti, 2007) Dunaliella salina 6e25 (Gong and Jiang, 2011; Gouveia and Oliveira, 2009) Ellipsoidion sp. 27 Isochrysis sp. 25e33 (Chisti, 2007) Monallanthus salina 20 (Chisti, 2007) Nannochloris sp. 20e35 0.038e0.061 (Chisti, 2007) Nannochloropsis oculata 22e30 0.084e0.142 (Gong and Jiang, 2011) Nannochloropsis sp. 31e68 Neochloris oleoabundans 29e65 Nitzschia sp. 45e47 Pavlova lutheri 36 0.05 (Gong and Jiang, 2011) Pavlova salina 31 0.049 (Gong and Jiang, 2011) Phaeodactylum tricornutum 18e57 0.045 (Chisti, 2007; Gong and Jiang, 2011) Scenedesmus dimorphus 16e40 (Gouveia and Oliveira, 2009) Scenedesmus obliquus 35e55 (Gouveia and Oliveira, 2009) Scenedesmus sp. 20e21 Schizochytrium sp. 50e77 (Chisti, 2007) Spirulina maxima 4e9 (Gouveia and Oliveira, 2009) Tetraselmis sueica 15e23 (Chisti, 2007) References (Chisti, 2007) 1.214 (Gong and Jiang, 2011) 0.045 (Gong and Jiang, 2011) 0.054 (Gong and Jiang, 2011) 0.047 (Gong and Jiang, 2011) (Chisti, 2007; Gong and Jiang, 2011) 0.090e0.134 (Chisti, 2007; Gong and Jiang, 2011; Gouveia and Oliveira, 2009) (Chisti, 2007) 0.041e0.054 and nitrogen source for growth, would also be affected (Cordell et al., 2009). Both phosphorus and nitrogen are available in plentiful supply within waste water streams (Sawayama et al., 1995; Yun et al., 1997). (Gong and Jiang, 2011) Commercial harvesting of algae blooms from wastewater has already been demonstrated in New Zealand (Aquaflow, 2013) and the use of wastewater streams as a nutrient source in large-scale cultivation of microalgae
192 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY has been well studied and implemented. Similarly, in terms of water usage, microalgae cultivation, particularly in closed cultivation systems, demonstrates significant water savings when compared to traditional biofuel crops. Many microalgae species are also capable of growing in brackish water most notably Dunaliella salina (Weldy and Huesemann, 2007). ENERGY BALANCE Any credible source of bioenergy should not only be economically viable but also environmentally sustainable. The economic and environmental impacts of any source of bioenergy, including biolipids from microalgae, will usually be measured in terms of energy return on energy investment (EROI) and/or GHG emissions. These economic and environmental impacts of biofuels and microalgae biofuels in general have been hotly debated in recent years. A number of life cycle analyses (LCAs) have been undertaken with seemingly conflicting results (Benemann et al., 2012; Liu et al., 2011; Resurreccion et al., 2012; Sun et al., 2011). Similar disparities arose in the case of second-generation biofuels such as corn ethanol before the introduction of the Energy and Resources Group (ERG) Bioenergy Meta-Model (Farrell et al., 2006). The results of reported LCA analyses are hindered by the lack of fully integrated commercial-scale microalgae to bioenergy systems from which to obtain accurate measurements. Estimates are based on projections from laboratoryand pilot-scale tests, as well as some commercial data. Despite these facts an overall meta-analysis concluded that algae-based biodiesel would result in energy consumption and GHG emissions on par with terrestrial alternatives (Liu et al., 2011). In this study the authors consider a microalga-based bioenergy system whereby CO2 and nitrogen for microalgae cultivation are recycled from waste streams and the microalgae coproducts are used for further bioenergy production in the form of methane. This concept of an integrated “biorefinery” has been proposed previously (Borowitzka, 1995, 1999; Chisti, 2007; Martı́n and Grossmann, 2012). As alluded above, the “biorefinery” concept envisages the main inputs into the cultivation process such as carbon, nitrogen and phosphorus being supplied through various waste streams. Similarly, the microalgae product resulting from cultivation could be fully “refined” into a number of outputs including biolipids for bioenergy, biolipids for nutraceutical applications, proteins for animal feeds, sugars for bioethanol production, etc. At present, where fully commercial scale cultivation of microalgae and conversion to fuel alone is still not economically feasible, the “biorefinery” concept appears to offer the best short to medium term path to scale-up. In addition to the potential economic and environmental advantages of using microalgae-derived biolipids, the properties of the resulting biodiesel product are also worth considering. As detailed later in this chapter, biodiesel is produced by transesterification of the biolipids from an appropriate feedstock. Much like the plant- and animal-based biolipids discussed previously, the profile of the microalgae-derived biolipids that undergo transesterification will ultimately determine the quality of the biodiesel product. This profile will include the level of polyunsaturated fatty acids (PUFAs), the level of FFAs and the level of TAGs. Although the lipid profile of microalgae varies among species and even among the same species under different conditions of growth, approximately 80% of the lipid content of microalgae, in general, will be made up of storage lipids in the form of TAGs. TAGs are made up of three fatty acid chains, usually with a chain length of C14 to C22 for microalgae-derived biolipids, joined to glycerol through three ester bonds (Scott et al., 2010). These TAGs can be easily transesterified in the presence of methanol, as described later in the chapter, to fatty acid methyl esters (FAMEs), which make up biodiesel. The presence of FFAs, however, results in the formation of soaps during transesterification in the presence of a base catalyst such as NaOH. This increases the downstream processing required to produce a finished biodiesel product. Similarly, the presence of PUFAs in biolipids derived from some microalgae species can cause tar formation resulting from fatty acid chains cross-linking (Burton et al., 2009). A high PUFA content could also mean that a biodiesel product would not pass European standards for biodiesel (EN14214), which demand the content of FAMEs with four or more double bonds to be below 1% mol (Knothe et al., 2005). Other properties that have been considered with regard to other feedstocks mentioned in this chapter include the cloud point, the cetane number and the oxidation stability of the biodiesel fuel. It has been suggested that biodiesel from microalgae oils may face significant performance problems regarding cold flow and oxidative stability in particular (Knothe, 2011); however, exceptions to this observation may apply to some microalgae such as Trichosporon capitatum. Also, in a recent study, biodiesel derived from the microalgae Chaetoceros gracilis was found to generate similar torque and power to soy-derived biodiesel. In terms of emissions, the C. gracilis-derived biodiesel also produced less CO, NOx and hydrocarbons than petroleum diesel (Wahlen et al., 2012). It is clear that the potential for algae to supply a sustainable source of biolipid for transportation fuel and other forms of bioenergy is not in doubt. However, there remain technical, economic and environmental challenges to be overcome. In a recent report by the National
PROCESSING OF BIOLIPIDS AND PROPERTIES OF BIOLIPID-DERIVED BIOFUELS Research Council in the United States entitled, “Sustainable development of algal biofuels” a number of sustainability concerns were highlighted. These included EROI; GHG emissions and resource usage such as land, water, nitrogen, phosphorus, and carbon dioxide (National Research Council, 2012). None of these concerns, however, were considered a “definitive barrier to sustainable development of algal biofuels”. This is because a number of strategies have already been implemented to tackle these challenges. As mentioned previously the use of wastewater streams can drastically reduce resource usage and GHG emissions as well as greatly increase EROI. Current projects, at industrial scale, such as Sapphire Energy’s “Green Crude Farm” (Sapphire Energy, 2013) aim to have a capacity of 1 million gallons per year of finished biofuel product. It is predicted that this will result in a 60e70% reduction in GHG emissions compared to traditional fossil crude oil, which, if achieved, will make the potential of microalgaederived biofuel a very definite reality. PROCESSING OF BIOLIPIDS AND PROPERTIES OF BIOLIPID-DERIVED BIOFUELS Independent of the biomass source, biolipids can be used in various ways as a source of bioenergy. There are a number of basic steps involved in processing biolipids to biofuel. These can include some or all of lipid extraction, degumming, neutralization, winterization, bleaching and transesterification. The sources of biomass and how they are produced have been described previously in this chapter and the first processing step will usually involve efficient extraction of the biolipid from the biomass. Following extraction, some biolipids can be used in their pure form as pure plant oils (PPOs). Other biolipids are further processed, usually into biodiesel. Here the extraction step is followed by purification and stabilization of the biolipid and the conversion to biodiesel. The various steps involved in processing biolipids are described below, beginning with extraction, along with the fuel properties of both PPO and biodiesel. Extraction Extraction is a process consisting of the separation of a specific substance from a complex matrix. In the context of extraction lipids from biomass, the purpose is to use standardized extraction procedures to isolate the biomolecules of interest, i.e. lipids, concurrent to rejection of the remaining inert biomass. This is most commonly achieved by using a selective solvent known as menstruum (Handa, 2008), or by solventless physical 193 extraction means. The resultant lipid may be ready for use in the form of fluid extracts, it may be further processed into a variety of biofuel and nutraceutical products, or it may be fractionated to isolate individual chemical entities or a combination of the above as proposed by the “biorefinery” concept discussed previously. The most common biolipid extraction procedures are summarized below. Steam Distillation Steam distillation is a process that is commonly applied to the extraction of essential oils (Gutierrez et al., 2009). Plant material is placed into a still where pressurized steam penetrates the plant material causing internal lipid vacuoles to rupture. Upon exposure to the surrounding environment, the lipid evaporates to form a mixture of easily separable vapors (essential oil and water). The vapors condense and the distillate (separated into two immiscible layers) is collected in a graduated tube connected to a condenser. The aqueous phase is recirculated into the flask, while the volatile oil is collected separately. The main disadvantage associated with steam distillation is that thermolabile components risk being degraded (Sarker et al., 2005). A combination of solvents and steam distillation is often used to improve the final product of a biodiesel production process. Maceration (Solvent Extraction) Maceration is used for creating extracts and resins in a simple yet well-established procedure. Whole or coarsely powdered biomass is placed in intimate contact with a suitable extractant in a closed vessel. The mixture is allowed to stand at room temperature for a defined period of time, typically at least 3 days, with frequent agitation (using mechanical shakers or mixers) to ensure homogeneity (Sarker et al., 2005). The organic phase is separated from the solids by either filtration, decantation or in some cases centrifugation and the remaining solid material is pressed to ensure efficient solvent recovery. The recovered liquid phases are combined and clarified for further processing. This process can be repeated several times to achieve maximum lipid recovery. The main disadvantage associated with maceration is that the process can be quite onerous, potentially taking from a few days up to several weeks (Takahashi et al., 2001). Enzymatic Hydrolytic Maceration Certain plant materials require enzymatic maceration prior to lipid release as their volatile components are glycosidically bound. Enzymes can be either
194 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY endogenous or exogenous to the biomass. For example, methyl salicylate (wintergreen oil) is an organic ester that is naturally produced by many species of plants. The plant leaves contain the precursor gaultherin and the enzyme primeverosidase; when the leaves are macerated in warm water, the endogenous enzyme acts on the gaultherin and liberates free methyl salicylate and primeverose (Handa, 2008). In the case of the exogenous addition of enzymes, recent advances in the field of algal lipids have demonstrated the addition of complex mixtures of enzymes to selectively degrade cell walls in a cascade of hydrolytic reactions. Released lipids are isolated and collected for further processing (Liang et al., 2012). Expression (Cold Pressing) Expression or cold pressing is commonly used in the production of essential and food oils. The term expression refers to any physical process in which the essential oil glands in the biomass are crushed or broken to release the oil. The resulting oilewater emulsion is typically separated by centrifugation. Traditionally, cold pressing was conducted by hand; however, for largescale commercialization, this is impractical. Thus, with the advancement of industrialization, a number of machines have been designed to achieve the same results on commercial scale. It is important to note that oils extracted using this method have a relatively short shelf life (Martı́nez et al., 2008). Hot Continuous Extraction (Soxhlet) In this method, finely ground biomass is placed in a porous bag or “thimble” made of strong cellulose, which is placed in the extraction chamber of a Soxhlet apparatus. The menstruum is heated, and the condensed extractant drips into the thimble containing the biomass, ensuring intimate continuous contact with the biomass. When the level of liquid in the extraction chamber reaches overflow, the liquid contents siphon into the heating chamber. This process is continuous and is carried out until complete extraction is achieved (Morrison and Coventry, 2006). The advantage of this method is that large amounts of lipid can be extracted with a much smaller quantity of solvent. Countercurrent Extraction Counter-current extraction is a process whereby wet raw material is pulverized using toothed disc disintegrators to produce slurry in a semicontinuous stream. As the pulverization of the biomass is in aqueous media, the heat generated during comminution is counterbalanced by the slurry water, preserving thermolabile compounds. The slurry stream is moved in one direction within the cylindrical extractor where it comes into discreet contact with a suitable menstruum (Vishwakarma, 2010). Complete extraction is possible when the quantities of solvent and material and their respective flow rates are optimized. The quantity of solvent required is generally minimal and as the process is most often conducted at room temperature, the threat to thermodegradation of volatile compounds is negated (Handa, 2008). Ultrasound Extraction (Sonication) The use of sonication is an emerging technology that is gaining widespread industrial acceptance due to recent advances in the scalability of the technology (Awad et al., 2012b; Dolatowski et al., 2007). In the context of lipid extraction from biomass, ultrasound technology is used to increase the permeability of biomass cell walls by generating cavitation events. These events are created by the use of high frequencies (20e2000 kHz) to generate a microbubble in solution; the intensity of the waves leads to the eventual collapse of the bubble generating extreme localized pressure and temperature events in close proximity to the biomass. These cavitation events assist in the rupturing of the cell walls to release the intercellular constituents into the surrounding environment. Once the biomolecules of interest are released from the biomass they can be recovered using conventional techniques. One disadvantage of using ultrasonics in the occurrence of sonolysis, i.e. the occasional but deleterious effect that when high power (typically greater than 20 kHz) is applied in aqueous media it can lead to the formation of free radicals and hydrogen peroxide. These are generated at the interfacial double layer established during cavitations, which subsequently diffuse into solution (O’Donnell et al., 2010). Supercritical Fluid Extraction Another technology in the extraction space is supercritical fluid extraction (SFE) whereby a solvent is subjected to temperature and pressure conditions to adjust the properties to those intermediate to a gas and liquid in a dedicated reactor setup. This in turn effects the solubilization of solutes in a matrix (Wenclawiak, 1992). The main supercritical solvent employed is carbon dioxide. Carbon dioxide (critical conditions: T ¼ 30.9  C and P ¼ 73.8 bar) is cheap, environmentally friendly and has generally recognized as safe status from the US Food and Drug Administration. Supercritical CO2 (SC-CO2) is also attractive because of its high diffusivity combined with its easily tunable solvent strength (Herrero et al., 2010).
PROPERTIES OF PURE PLANT OIL However, due to its chemical nature, it possesses several polarity limitations. As mentioned previously, solvent polarity is particularly important when extracting polar solutes and when strong matrix interactions are present. To augment the process, organic solvents are commonly added to the carbon dioxide extracting fluid to alleviate the polarity limitations (Handa, 2008). CO2 is gaseous at room temperature and pressure, which makes recovery very simple and provides solvent-free products, i.e. once the liquid depressurizes, the CO2 returns to a gaseous state, and only the extracted products remain. SFE using CO2 can be operated at low temperatures, which allows the extraction and integrity preservation of thermolabile compounds (Mendiola et al., 2007). PROPERTIES OF PURE PLANT OIL Following extraction from biomass, biolipids can be used as pure oil (generally plant) or can be converted to biodiesel by a process known as transesterification, described later. However, the use of PPO as a fuel requires the modification of diesel engines unlike biodiesel, which, particularly when blended with petroleum diesel, can be used in unmodified diesel engines. These engine modifications are needed as PPO is more than 10 times as viscous as biodiesel. As a result, it has a tendency to gum up in cold weather, which can be somewhat overcome by blending with traditional fossil diesel. Nevertheless, it has some advantages: with a flash point of over 300  C, storage and transport are simplified. According to the VwVwS (Verwaltungsvorschrift wassergefährdende Stoffe), which is the national German regulation on water hazard classification, PPO is not designated as even a hazard to water given that it is biodegradable. In an unmodified engine, poor atomization of the fuel will lead to coking of the injectors and accumulation of soot deposits. Modification is designed to preheat fuel or involves installation of a two-tank system. In the latter, the engine is started with diesel and only changes to PPO when the operating temperature has been reached. It must switch back to diesel before being turned off, to flush out the remainder of the PPO in order to ready the engine for the next operation. Other options exist, such as the specialist engine developed by Ludwig Elsbett in the 1970s. The fuel emissions of PPO are also much lower in sulfur emissions when compared to the fossil equivalent. For a detailed overview see (Russo et al., 2012). After extraction, if the biolipid is not to be used as PPO, or other pure oil, it needs to be further processed into a more useable biofuel, usually biodiesel. Here the biolipid goes through a series of processing steps beginning with degumming. 195 Degumming Following extraction and regardless of the process described above, the end product will generally be a rather impure biolipid that contains undesirable contents such as FFAs, tocopherols, waxes and possibly phosphatides. The latter, if not removed before storage, will produce a thick gum over time. Gums are formed when the biolipid absorbs water, which causes some of the phosphatides (such as phosphocholine) to become hydrated and thereby lipid insoluble. Accordingly, hydrating the gums and removing the hydrated gums from the oil before storage can prevent the formation of a gum deposit. This treatment is called water degumming and involves the addition of water at 60e90  C before the phase is separated. An optimum temperature is sought, as it must not be so high as to increase the solubility of phosphatides in oil. A temperature that is too low will increase the viscosity, making phase separation more difficult. It is never applied to fruit oils like olive oil and palm oil, since these oils have already had considerable water contact during their production. The removal of nonhydratable phosphatides (such as phosphatidic acid) requires the addition of an acid, usually citric or phosphoric, which will form a sludge that can be easily removed (Dijkstra and Van Opstal, 1989). This addition of acid is proportional to the amount of phosphorous already contained in the sample. In addition, this acid also reduces any iron salts and decreases chlorophyll contamination. Enzymatic degumming focuses on the use of lipases, which convert nonhydratable lipids to more hydratable forms. Although the process has been tried at a larger scale for 20 years, it has not made the advancement toward widespread use (Dijkstra, 2010; Yang et al., 2008). Alkaline Neutralization As mentioned previously, the presence of FFAs in the biolipid is detrimental to oil quality and function, including biodiesel production. Removal typically involves the reaction of these FFAs with an alkaline solution. In the edible oil industry, usually only caustic soda is used for this reaction, but potassium hydroxide is also used by some producers. The acidity of the FFA comes from the Hþ of the carboxyl group. This Hþ of the functional group of the stearic acid reacts with the OH group of the caustic soda (NaOH) to produce soap and water. In addition to the removal of FFAs, other undesirable nonglyceride materials are also removed in this fashion such as phenol, oxidized fatty compounds, heavy metals and phospholipids. Winterization Most biolipids do not need dewaxing, as they contain little or no waxes. Only biolipids of higher melting
196 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY temperatures, such as sunflower oil and rice bran oil, give a hazy appearance during winter season due to precipitation of dissolved waxes. Hence, they require being dewaxed. This is carried out by chilling the oil to 10e15  C, followed by filtration of precipitated solids. The oil thus treated has a sparkling appearance, even in winter temperatures. Bleaching Oil bleaching, which is performed in order to prepare a sufficiently light-colored product of enhanced appearance and improved stability, is usually achieved by treating the crude or refined oil with powdered absorbent. These absorbents usually contain a calcium montmorillonite (fuller’s earth) or natural hydrated aluminum silicate (bentonite). Adsorption of color bodies, trace metals and oxidation products, as well as residual soaps and phospholipids remaining after washing neutralized oils takes place, if possible. Acidactivated clays are the major adsorbent used, although active carbons and synthetic silicas are also applied industrially with more specific goals. Thus, active carbons are used specifically to eliminate polycyclic aromatic hydrocarbons from some oils, especially fish oils and pomace oils, while synthetic silicas are quite efficient in adsorbing secondary oxidation products, phospholipids and soaps (León-Camacho et al., 2003). There are a number of good sources of material with more detailed descriptions of each process found online at the Lipid Library (Hardwood and Weselake, 2013), in “Proceedings of the World Conference on Oilseed Technology and Utilization” (Applewhite, 1993) and finally in, Edible Oil Processing (Hamm and Hamilton, 2000). Transesterification Despite being energetically favorable, the direct use of plant or other biolipids in fuel engines is problematic as described earlier. Briefly, due to high viscosity (over 10 times higher than diesel fuel) and low volatility, they do not burn efficiently and can form deposits in the fuel injector of diesel engines. Furthermore, acrolein (a highly toxic substance) is formed through thermal decomposition of glycerol. Different ways have been considered to reduce the high viscosity of plant and other biolipids, but the principal method is to engage in chemical transesterification to produce biodiesel, which could be used in the common diesel engine with minor modification. As mentioned previously, biolipids consist primarily of triglycerides, which are three hydrocarbon chains connected by glycerol. The bonds are hydrolyzed to allow the formation of FFAs, which are mixed and reacted with methanol or ethanol to form methyl (or ethyl) fatty acid esters. The use of methanol (methanolysis) is widespread and considered advantageous, as it is cheaper than ethanol (although in Brazil, ethanol 90 is plentiful) and has less azeotrophic qualities (Encinar et al., 2007). The same reaction using ethanol is more complicated as it requires a water-free alcohol, as well as a biolipid with low water content, in order to obtain good glycerol separation. Methanolysis can happen by heating 80e90% methanol with a small amount of catalyst. The received biodiesel after methanolysis is FAME and with ethanol to form fatty acid ethyl ester. The use of ethanolysis reaction using bioethanol has been discussed as being possibly more environmentally favorable as it would involve the use of a nonfossil fuel. Apart from this, ethanol is less toxic and slightly increases the cetane number of the biofuel. Although transesterification can proceed in the absence of catalysts, the reaction proceeds much too slowly to be economically viable and thus typically requires an acidic or alkaline catalysis. Among the most commonly used alkaline catalysts in the biodiesel industry are potassium hydroxide (KOH) and sodium hydroxide (NaOH) flakes, which are inexpensive, easy to handle and can be transported and stored easily. For this reason, they are preferred by smaller producers. Alkyl oxide solutions of sodium methoxide (NaOCH3) or potassium methoxide (KOCH3) in methanol, which are now commercially available, are the preferred catalysts for large continuous-flow production processes. In the transesterification process, the effective species of catalysis is the methoxide radicals (CH3O) and the activity of a catalyst depends upon the amount of methoxide radicals (Komers et al., 2001a,b). For sodium or potassium hydroxide, the methoxide ion is prepared in situ by reacting methanol with hydroxide, a reaction that will also produce water that remains in the system. Hydrolysis of triglycerides and alkyl esters may occur due to the presence of this water, which further leads to the formation of FFAs and thus to a soap. Saponification may also occur if a strong base, e.g. NaOH or KOH, is present in the system by reacting with esters and triglycerides directly. All these problems can be avoided completely if sodium and potassium methoxide solutions, which can be prepared water-free, are applied (Singh et al., 2006). PROPERTIES OF BIODIESEL Untreated biodiesel blend stocks, generated by transesterification, generally exhibit poor oxidation stability, which can result in long-term storage problems. Biodiesel has many similar fuel economies to fossil diesel. Although it has about 10% less energy content
BIOMASS TO LIQUID FUELS (BIO-OIL) per volume, its cetane number and lubricating effect are higher, which is advantageous (Rutz and Janssen, 2007). The higher oxygen content leads to better combustion and fewer pollutants, particularly sulfur oxides. Biodiesel is produced in a pure form (100% biodiesel blend stock, referred to as “B100” or “neat biodiesel”) and is typically blended with petroleum-based diesel fuel. Such biodiesel blends are designated as BXX, where XX represents the percentage by volume of pure biodiesel contained in the blend (e.g. “B5” or “B20”). According to a “Technical Statement on the Use of Biodiesel in Compression Ignition Engines” released in 2009 by the Truck and Engine Manufactures Association, neat biodiesel and higher percentage biodiesel blends can cause a variety of engine performance problems. These include fuel filter plugging, injector coking, piston ring sticking and breaking, elastomer seal swelling and hardening/cracking, and severe engine lubricant degradation and dilution. The report goes on to state that when converting from petroleumbased diesel to a biodiesel blend, residual fuel system deposits may accumulate in fuel filters due to the high solvency of the fuel. Thus, more frequent filter service may be required until the fuel system deposits are stabilized. More information on biofuel handling can be found in "Biodiesel Handling and Use Guide: Fourth edition (Revised)" published by NREL (National Renewable Energy Laboratory, 2009). BIOMASS TO LIQUID FUELS (BIO-OIL) While the focus of this chapter has been on biolipids it is important to note that any biomass can be converted to “bio-oil” via a high-temperature process known as pyrolÒ ysis. This “bio-oil” also known as Synfuel or Sunfuel is currently only produced on a small scale and it very much belongs to the second-generation biofuels, as it is a way of generating fuel from a range of biomass including straw, wood or other materials high in lignin, which are difficult to convert to bioethanol. The potential for mass production remains enormous. The production of this biomass to liquid or BtL fuel can vary in complexity and can vary depending on the individual needs, but it essentially comprises the following steps. Gasification Gasification is a form of incomplete combustion in which a fuel is burnt in an oxygen-deficient atmosphere. An energy-rich gas, consisting principally of methane, CO and hydrogen, is formed but heat release is minimized. Thus an energy-rich fuel (biomass) is converted into an energy-rich gas. There are differing processes for gasification. For example, a description of the 197 Ò Carbo-V process first developed by Chloren Industries but now owned by Linde Engineering GmbH was outlined in the Biofuel Technology Handbook (Rutz and Janssen, 2007). This involves low-temperature gasification, where low-temperature pyrolysis with air or oxygen at 400e500  C allows the continuous production of a gas containing both tars (volatile component) and char (carbon solids). This is followed by a hightemperature gasification, where the gas is further oxidized (again hypostoichiometrically) in a combustion chamber. The third part involves blowing the pulverized char into the hot gasification medium. Pulverized char and gasification medium react endothermically in the gasification reactor and are converted into a raw synthesis gas. Other gasification processes can be found, such as the recently developed BioliqÒ , which was formed by Lurgi AG (Frankfurt Germany) with Karlsruhe Institute of Technology (Karlsruhe Germeny). Cleaning Process After gasification, it is usual to have many impurities and thus cleaning remains one of the most important and most technical challenges. Remaining tars tend to be refractory and difficult to remove by thermal or physical processes. Generally, the impurities in biosyngas produced from the gasifier can be grouped into three types: (1) organic impurities, such as tars, benzene, toluene, and xylenes; (2) inorganic impurities, such as O2, NH3, HCN, H2S, Carbonyl sulfide (COS), and HCl; and (3) other impurities, such as soot and dust. Both thermal cracking, which involves the addition of steam and oxygen at 200e1000  C, and catalytic cracking at lower temperatures is possible, as is low-temperature scrubbing with an oil-based medium may all encompass the process. A multicontaminant syngas treatment process created by Southern Research Institute, Birmingham, Alabama, USA, uses a candle filter, which can be catalytic, closely coupled with the gasifier. A variety of sorbents is injected into the gasifier or between the gasifier and filter to remove various contaminants (e.g. alkali metals, sulfur species, and halides) both by reaction in the gas phase and on the filter cake. Catalysts may be incorporated into the candle filter or the filter may be coated with a catalyst to crack tar and ammonia depending on the operating temperature of the candle filter. An outline of the process can be seen in Figure 12.1. Synthesis Two methods are available for this production step, but the Fischer-Tropsch (FT) synthesis is the most widely known. It was developed at the Kaiser-Wilhelm Institute for Research on Coal (Mühlheim/Ruhr) in 1925. In Germany, coal to liquid fuels have been
198 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY FIGURE 12.1 Biobased fats (lipids) and oils from biomass as a source of bioenergy. Integration of a catalytic filter into a gasifier for combined particle separation and tar removal from biomass gasification gas. Source: Courtesy of Southern Research Institute, Birmingham, Alabama, USA. (For color version of this figure, the reader is referred to the online version of this book.) produced with the help of FT synthesis since 1938. During the process, CO and H2, with the aid of a catalyst, will form hydrocarbons. A variety of catalysts exist, but the most common are usually transition metals such as cobalt. In the case of biomass, however, an iron catalyst is often favored (Hu et al., 2012). The other process is the methanol-to-gasolineÒ method, in which the syngas is first transformed into methanol as an intermediate state. In a following step, fuels can be obtained from this compound. Finally, after separating the produced liquid hydrocarbons into heavy, medium and light fractions, these hydrocarbons are refined and blended to achieve the desired fuel properties. CONCLUSION The search for a sustainable supply of fuel that does not contribute to global warming has consumed environmental scientists for decades. While it is unlikely there is a “silver bullet” solution to the pending energy crisis the use of biolipids has enormous potential to meet a large proportion of the global transport fuel requirements. Similarly, no Single lipid source is produced in sufficient quantities to impact on the world’s fuel supplies; therefore, a combination of all biolipids outlined above will be required if biolipids are to be a realistic alternative to petroleum-based fuels. While plant-derived biolipids currently dominate the liquid bioenergy markets, microalgae remain the most promising source of biolipids in the future. The limited land usage requirement and efficient carbon fixing capabilities of microalgae make them the ideal choice as a source of biolipids; however, there are a number of stumbling blocks to be overcome before algal biofuels are a commercial reality. These include the challenge of growing algae at industrial scale to meet the increasing demand for liquid transport fuel, the energy input involved in harvesting and dewatering algae and finally the cost and environmental impact of efficiently extracting biolipids from algae. These challenges are far from insurmountable, however, and each challenge is being tackled by numerous academic institutions and increasingly, by large, multinational energy, food and industrial chemical companies. This concerted effort with regard to algae biofuels, coupled with the more established plant- and animal-based biofuel industries can supply a significant portion of the world’s energy needs in the future. References Alakangas, E., Nikolaisen, L., Sikkema, R., Junginger, M., 2011. Combined nomenclatures (CN Codes) for biomass fuels. Available from: www.vtt.fi/inf/julkaisut/muut/2011/D2-4-EUBIONETIII_ CN_code_report.pdf (accessed 15.02.13.). Ali, A., Ahmad, F., Farhan, M., Ahmad, M., 2012. Biodiesel production from residual animal fat using various catalysts. Pak. J. Sci. 64, 282e286. Andés, L.E., 1898. Animal Fats and Oils: Their Practical Production, Purification and Uses for a Great Variety of Purposes, Their Properties, Falsification and Examination; a Handbook for Manufacturers of Oil-and Fat-products, Soap and Candle Makers, Agriculturists, Tanners, Etc. Scott, Greenwood, London, 240.
REFERENCES Applewhite, T., 1993. Proceedings of the World Conference on Oilseed Technology and Utilization. Amer Oil Chemists Society, USA. pp. 507. Aquaflow, 2013. Blenheim municipal wastewater. Available from: www.aquaflowgroup.com/projects/blenheim-municipalwastewater (accessed 10.01.13.). Aranda, D.A.G., da Silva, C.C.C.M., Detoni, C., 2009. Current processes in Brazilian biodiesel production. Int. Rev. Chem. Eng. 1, 603e608. Awad, S., Paraschiv, M., Varuvel, E.G., Tazerout, M., 2012a. Optimization of biodiesel production from animal fat residue in wastewater using response surface methodology. Bioresour. Technol. 129, 315e320. Awad, T., Moharram, H., Shaltout, O., Asker, D., Youssef, M., 2012b. Applications of ultrasound in analysis, processing and quality control of food: a review. Food Res. Int. 48, 410e427. Bakir, E.T., Fadhil, A.B., 2011. Production of biodiesel from chicken frying oil. Pak. J. Anal. Environ. Chem. 12, 95e101. Baribeau, A.M., R. Bradley, P. Brown, J.J. Goodwin, U. Kihm, E. Lotero, D. O’Connor, M. Schuppers and D. Taylor. 2005. Biodiesel from specified risk material tallow: an appraisal of TSE risks and their reduction. Available from: www.iea-amf.vtt.fi/pdf/ annex30_vol1.pdf (accessed 15.02.13.). Benemann, J., Woertz, I., Lundquist, T., 2012. Life cycle assessment for microalgae oil production. Disrupt. Sci. Technol. 1, 68e78. Bereczky, Á., 2012. Alternative fuels and technologies for compression ignition internal combustion engines. J. KONES Powertrain Transp. 19, 43e51. Bilanovic, D., Andargatchew, A., Kroeger, T., Shelef, G., 2009. Freshwater and marine microalgae sequestering of CO2 at different C and N concentrations e response surface methodology analysis. Energy Convers. Manage. 50, 262e267. Borowitzka, M.A., 1988. Fats, oils and hydrocarbons. In: Borowitzka, L.J., Borowitzka, M.A. (Eds.), Micro-Algal Biotechnology. Cambridge University Press, Cambridge, pp. 257e287. Borowitzka, M.A., 1995. Microalgae as sources of pharmaceuticals and other biologically active compounds. J. Appl. Phycol. 7, 3e15. Borowitzka, M.A., 1999. Commercial production of microalgae: ponds, tanks, and fermenters. Prog. Ind. Microbiol. 35, 313e321. Brennan, L., Owende, P., 2010. Biofuels from microalgae - a review of technologies for production, processing, and extractions of biofuels and co-products. Renewable Sustainable Energy Rev. 14, 557e577. Brown, R.S., Keller, B.O., Oleschuk, R.D., 2007. Detection of prion proteins and TSE infectivity in the rendering and biodiesel manufacture processes. Available from: www.iea-amf.vtt.fi/pdf/ annex30_vol2.pdf (accessed 15.02.13.). Bruederle, C.E., Hnasko, R.M., Kraemer, T., Garcia, R.A., Haas, M.J., Marmer, W.N., Carter, J.M., 2008. Prion infected meat-and-bone meal is still infectious after biodiesel production. PloS One. 3, 1e8. Burton, T., Lyons, H., Lerat, Y., Stanley, M., Rasmussen, M.B., 2009. A review of the potential of marine algae as a source of biofuel in Ireland. Available from: www.seai.ie/Publications/Renewables_ Publications_/Bioenergy/Algaereport.pdf (accessed 21.02.13.). Canakci, M., Van Gerpen, J., 2001. Biodiesel production from oils and fats with high free fatty acids. Trans. ASME. 44, 1429e1436. Carvalho, A.P., Meireles, L.A., Malcata, F.X., 2006. Microalgal reactors: a review of enclosed system designs and performances. Biotechnol. Prog. 22, 1490e1506. Chisti, Y., 2007. Biodiesel from microalgae. Biotechnol. Adv. 25, 294e 306. Cordell, D., Drangert, J.O., White, S., 2009. The story of phosphorus: global food security and food for thought. Global Environ. Change 19, 292e305. Darunde Dhiraj, S., Deshmukh Mangesh, M., 2012. Biodiesel production from animal fats and its impact on the diesel engine 199 with ethanol-diesel blends: a review. Int. J. Emerging Technol. Adv. Eng. 2, 179e185. Demirbas, A., 2009. Political, economic and environmental impacts of biofuels: a review. Appl. Energy. 86, S108eS117. Demirbas, M.F., Balat, M., Balat, H., 2011. Biowastes-to-biofuels. Energy Convers. Manage. 52, 1815e1828. Dijkstra, A.J., 2010. Enzymatic degumming. Eur. J. Lipid Sci. Technol. 112, 1178e1189. Dijkstra, A.J., Van Opstal, M., 1989. The total degumming process. J. Am. Oil Chem. Soc. 66, 1002e1009. Dolatowski, Z.J., Stadnik, J., Stasiak, D., 2007. Applications of ultrasound in food technology. Acta Sci. Pol. Technol. Aliment. 6, 89e 99. Duku, M.H., Gu, S., Hagan, E.B., 2011. A comprehensive review of biomass resources and biofuels potential in Ghana. Renewable Sustainable Energy Rev. 15, 404e415. Dunn, R.O., 2009. Cold-flow properties of soybean oil fatty acid monoalkyl ester admixtures. Energy Fuels 23, 4082e4091. Ellis, M., 2007. Air Quality and Biodiesel Production. Nebraska Department of Environmental Quality. Available from: www. deq.state.ne.us/Publica.nsf/c4afc76e4e077e11862568770059b73f/ 9bf9daf0963eb8728625729f00575cbd/$FILE/06-272.pdf (accessed 14.02.13.). Encinar, J., González, J., Rodrı́guez-Reinares, A., 2007. Ethanolysis of used frying oil. Biodiesel preparation and characterization. Fuel Process. Technol. 88, 513e522. Farrell, A.E., Plevin, R.J., Turner, B.T., Jones, A.D., O’Hare, M., Kammen, D.M., 2006. Ethanol can contribute to energy and environmental goals. Science 311, 506e508. Feddern, V.J., De Praá, A.C., de Abreu, M.C., dos Santos Filho, P.G., Higarashi, J.I., Sulenta, M.M., Coldebella, A., 2011. Animal fat wastes for biodiesel production. In: Montero, M.S.A.G. (Ed.), Biodiesel - Feedstock and Processing Technologies. InTech, pp. 45e70. Food and Agriculture Organization, 2008. The state of food and agriculture 2008 biofuels: prospects, risks and opportunities. Available from: www.fao.org/docrep/011/i0100e/i0100e00.htm (accessed 14.02.13.). Francis, G., Edinger, R., Becker, K., 2005. A concept for simultaneous wasteland reclamation, fuel production, and socioeconomic development in degraded areas in India: need, potential and perspectives of Jatropha plantations. Nat. Resour. Forum 29, 12e24. Geller, D.P., Adams, T.T., Goodrum, J.W., Pendergrass, J., 2008. Storage stability of poultry fat and diesel fuel mixtures: specific gravity and viscosity. Fuel 87, 92e102. Gong, Y., Jiang, M., 2011. Biodiesel production with microalgae as feedstock: from strains to biodiesel. Biotechnol. Lett. 33, 1269e 1284. Gouveia, L., Oliveira, A.C., 2009. Microalgae as a raw material for biofuels production. J. Ind. Microbiol. Biotechnol. 36, 269e274. Greene, A.K., Dawson, P.L., Nixon, D., Atkins, J.R., Pearl, G.G., 2007. Safety of animal fats for biodiesel production: a critical review of the Literature. Available from: www.iea-amf.vtt.fi/pdf/annex30_ vol3.pdf (accessed 15.02.13.). Greenwell, H., Laurens, L., Shields, R., Lovitt, R., Flynn, K., 2010. Placing microalgae on the biofuels priority list: a review of the technological challenges. J. R. Soc., Interface 7, 703e726. Groschen, R.. Overview of the Feasibility of Biodiesel from Waste/ Recycled Greases and Animal Fats. Minnesota Department of Agriculture. Available from: www.mda.state.mn.us/news/publications/ renewable/wastefatsfeasability.pdf (accessed 08.02.13.). Gui, M.M., Lee, K.T., Bhatia, S., 2008. Feasibility of edible oil vs nonedible oil vs waste edible oil as biodiesel feedstock. Energy 33, 1646e1653.
200 12. BIOBASED FATS (LIPIDS) AND OILS FROM BIOMASS AS A SOURCE OF BIOENERGY Gutierrez, J., Barry-Ryan, C., Bourke, P., 2009. Antimicrobial activity of plant essential oils using food model media: efficacy, synergistic potential and interactions with food components. Food Microbiol. 26, 142e150. Hamm, W., Hamilton, R.R.J., 2000. Edible Oil Processing. Sheffield Academic Press, Sheffield. p. 281. Handa, S.S., 2008. An overview of extraction techniques for medicinal and aromatic plants. In: Handa, S.S., Khanuja, S.P.S., Longo, G., Rakesh, D.D. (Eds.), Extraction Technologies for Medicinal and Aromatic Plants. International Centre for Science and High Technology (ICS), Trieste, pp. 21e52. Hardwood, J.L., Weselake, R.J., 2013. AOCS lipid library. Available from: www.lipidlibrary.com (accessed 10.01.13.). Herrero, M., Mendiola, J.A., Cifuentes, A., Ibáñez, E., 2010. Supercritical fluid extraction: recent advances and applications. J. Chromatogr. A 1217, 2495e2511. Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M., Darzins, A., 2008. Microalgal triacylglycerols as feedstocks for biofuel production: perspectives and advances. Plant J. 54, 621e639. Hu, J., Yu, F., Lu, Y., 2012. Application of FischereTropsch synthesis in biomass to liquid conversion. Catalysts 2, 303e326. Huntley, M.E., Redalje, D.G., 2007. CO2 mitigation and renewable oil from photosynthetic microbes: a new appraisal. Mitigation Adapt. Strategies Global Change 12, 573e608. Jain, S., Sharma, M., 2010. Stability of biodiesel and its blends: a review. Renewable Sustainable Energy Rev. 14, 667e678. Jayasinghe, P., Hawboldt, K., 2011. A review of bio-oils from waste biomass: focus on fish processing waste. Renewable Sustainable Energy Rev. 16, 798e821. Kerihuel, A., Senthil Kumar, M., Bellettre, J., Tazerout, M., 2005. Use of animal fats as CI engine fuel by making stable emulsions with water and methanol. Fuel 84, 1713e1716. Kerschbaum, S., Rinke, G., Schubert, K., 2008. Winterization of biodiesel by micro process engineering. Fuel 87, 2590e2597. Kleinová, A., Vailing, I., Lábaj, J., Mikulec, J., Cvengros, J., 2011. Vegetable oils and animal fats as alternative fuels for diesel engines with dual fuel operation. Fuel Process. Technol. 92, 1980e1986. Knothe, G., 2011. A technical evaluation of biodiesel from vegetable oils vs algae. Will algae-derived biodiesel perform? Green Chem. 13, 3048e3065. Knothe, G., Van Gerpen, J.H., Krahl, 2005. The Biodiesel Handbook, vol. 1. AOCS press, Champaign, IL. USA. p. 304. Komers, K., Machek, J., Stloukal, R., 2001a. Biodiesel from rapeseed oil, methanol and KOH. 2. Composition of solution of KOH in methanol as reaction partner of oil. Eur. J. Lipid Sci. Technol. 103, 359e362. Komers, K., Stloukal, R., Machek, J., Skopal, F., 2001b. Biodiesel from rapeseed oil, methanol and KOH. 3. Analysis of composition of actual reaction mixture. Eur. J. Lipid Sci. Technol. 103, 363e371. Kusdiana, D., Saka, S., 2004. Effects of water on biodiesel fuel production by supercritical methanol treatment. Bioresour. Technol. 91, 289e295. Lakó, J., Hancsók, J., Yuzhakova, T., Marton, G., Utasi, A., Rédey, Á, Popa, V., 2008. Biomassea source of chemicals and energy for sustainable development. Environ. Eng. Manage. J. 7, 499e509. Lam, M.K., Tan, K.T., Lee, K.T., Mohamed, R., 2009. Malaysian palm oil: surviving the food versus fuel dispute for a sustainable future. Renewable Sustainable Energy Rev. 13, 1456e1464. León-Camacho, M., Viera-Alcaide, I., Ruiz-Méndez, M., 2003. Elimination of polycyclic aromatic hydrocarbons by bleaching of olive pomace oil. Eur. J. Lipid Sci. Technol. 105, 9e16. Liang, K., Zhang, Q., Cong, W., 2012. Enzyme-assisted aqueous extraction of lipid from microalgae. J. Agric. Food Chem. 60, 11771e11776. Liu, X., Clarens, A.F., Colosi, L.M., 2011. Algae biodiesel has potential despite inconclusive results to date. Bioresour. Technol. 104, 803e 806. Martı́n, M., Grossmann, I.E., 2012. On the systematic synthesis of sustainable biorefineries. Ind. Eng. Chem. Res. 52, 3044e3064. Martı́nez, M.L., Mattea, M.A., Maestri, D.M., 2008. Pressing and supercritical carbon dioxide extraction of walnut oil. J. Food Eng. 88, 399e404. Math, M., Kumar, S.P., Chetty, S.V., 2010. Optimization of biodiesel production from oils and fats with high free fatty acids. Indian J. Sci. Technol. 3, 318e321. Mathiyazhagan, M., Ganapathi, A., Jaganath, B., Renganayaki, N., Sasireka, N., 2011. Production of biodiesel from non-edible plant oils having high FFA content. Int. J. Chem. Environ. Eng. 2, 119e 122. Melis, A., 2009. Solar energy conversion efficiencies in photosynthesis: minimizing the chlorophyll antennae to maximize efficiency. Plant Sci. 177, 272e280. Mendiola, J.A., Herrero, M., Cifuentes, A., Ibañez, E., 2007. Use of compressed fluids for sample preparation: food applications. J. Chromatogr., A 1152, 234e246. Meng, X., Yang, J., Xu, X., Zhang, L., Nie, Q., Xian, M., 2009. Biodiesel production from oleaginous microorganisms. Renewable Energy 34, 1e5. Moheimani, N.R., Borowitzka, M.A., 2006. The long-term culture of the coccolithophore Pleurochrysis carterae (Haptophyta) in outdoor raceway ponds. J. Appl. Phycol. 18, 703e712. Morrison, W., Coventry, A., 2006. Extraction of lipids from cereal starches with hot aqueous alcohols. Starch-Stärke. 37, 83e87. Mrad, N., Varuvel, E.G., Tazerout, M., Aloui, F., 2012. Effects of biofuel from fish oil industrial residueediesel blends in diesel engine. Energy 44, 955e963. National Renewable Energy Laboratory, 2009. Biodiesel Handling and Use Guide. Revised, Fourth ed. National Renewable Energy Laboratory, Washington, D.C., USA. National Research Council, 2012. Sustainable Development of Algal Biofuels in the United States. The National Academies Press, Washington, D.C. USA. p. 232. Nigam, P.S., Singh, A., 2011. Production of liquid biofuels from renewable resources. Prog. Energy Combust. Sci. 37, 52e68. No, S.Y., 2011. Inedible vegetable oils and their derivatives for alternative diesel fuels in CI engines: a review. Renewable Sustainable Energy Rev. 15, 131e149. O’Donnell, C., Tiwari, B., Bourke, P., Cullen, P., 2010. Effect of ultrasonic processing on food enzymes of industrial importance. Trends Food Sci. Technol. 21, 358e367. Pandey, V.C., Singh, K., Singh, J.S., Kumar, A., Singh, B., Singh, R.P., 2012. Jatropha curcas: a potential biofuel plant for sustainable environmental development. Renewable Sustainable Energy Rev. 16, 2870e2883. Panneerselvam, S.I., Parthiban, R., Miranda, L.R., 2011. Poultry fatda cheap and viable source for biodiesel production. Proc. Int. Conf. Environ. Sci. Technol. (ICEST 2011) 6, 371e374. Phan, A.N., Phan, T.M., 2008. Biodiesel production from waste cooking oils. Fuel 87, 3490e3496. Pires, J.C.M., Alvim-Ferraz, M.C.M., Martins, F.G., Simões, M., 2012. Carbon dioxide capture from flue gases using microalgae: engineering aspects and biorefinery concept. Renewable Sustainable Energy Rev. 16, 3043e3053. Popescu, F., Ionel, I., 2011. Waste animal fats with high FFA as a renewable energy source for biodiesel production - concept, experimental production and impact evaluation on air quality. In: Manzanera., M. (Ed.), Alternative Fuel. Intech, pp. 93e108. Proc, K.B., Barnitt, R., Hayes, R.R., Ratcliff, M., McCormick, R.L., 2006. 100,000-mile evaluation of transit buses operated on biodiesel
REFERENCES blends (B20). In: Paper presented at Society of Automotive Engineers Powertrain and Fluid Systems, Toronto, Canada.  Prosková, A., Kopicová, Z., Kucera, J., Skarková, L., 2009. Acid catalysed transesterification of animal waste fat. Res. Agric. Eng. 55, 24e28. Prueksakorn, K., Gheewala, T.E., 2006. Energy and greenhouse gas implications of biodiesel production from Jatropha curcas L. In: Paper presented at Proceedings of the Second Joint International Conference on Sustainable Energy and Environment, Bangkok, Thailand. Pulz, O., 2001. Photobioreactors: production systems for phototrophic microorganisms. Appl. Microbiol. Biotechnol. 57, 287e293. Puri, M., Abraham, R.E., Barrow, C.J., 2012. Biofuel production: prospects, challenges and feedstock in Australia. Renewable Sustainable Energy Rev. 16, 6022e6031. Resurreccion, E.P., Colosi, L.M., White, M.A., Clarens, A.F., 2012. Comparison of algae cultivation methods for bioenergy production using a combined life cycle assessment and life cycle costing approach. Bioresour. Technol. 126, 298e306. Rodolfi, L., Chini Zittelli, G., Bassi, N., Padovani, G., Biondi, N., Bonini, G., Tredici, M.R., 2008. Microalgae for oil: strain selection, induction of lipid synthesis and outdoor mass cultivation in a lowcost photobioreactor. Biotechnol. Bioeng. 102, 100e112. Russo, D., Dassisti, M., Lawlor, V., Olabi, A., 2012. State of the art of biofuels from pure plant oil. Renewable Sustainable Energy Rev. 16, 4056e4070. Rutz, D., Janssen, R., 2007. Biofuel Technology Handbook, vol. 1. WIP Renewable Energies, 149. Sapphire Energy, 2013. Green crude farm. Available from: www. sapphireenergy.com/locations/green-crude-farm.html (accessed 10.01.13.). Sarker, S.D., Latif, Z., Gray, A.I., 2005. Natural Products Isolation, second ed. Springer, New York, USA. p. 515. Sawayama, S., Inoue, S., Dote, Y., Yokoyama, S.Y., 1995. CO2 fixation and oil production through microalga. Energy Convers. Manage. 36, 729e731. Schenk, P.M., Thomas-Hall, S.R., Stephens, E., Marx, U.C., Mussgnug, J.H., Posten, C., Kruse, O., Hankamer, B., 2008. Second generation biofuels: high-efficiency microalgae for biodiesel production. BioEnergy Res. 1, 20e43. Scott, S.A., Davey, M.P., Dennis, J.S., Horst, I., Howe, C.J., Lea-Smith, D.J., Smith, A.G., 2010. Biodiesel from algae: challenges and prospects. Curr. Opin. Biotechnol. 21, 277e286. Shahidi, F., Zhong, Y., 2005. Marine mammal oils. In: Shahidi, F. (Ed.), Bailey’s Industrial Oil and Fat Products. John Wiley & Sons, Inc., USA. pp. 259e278. Sheehan, J., Dunahay, T., Benemann, J., Roessler, P., 1998. A Look Back at the US Department of Energy’s Aquatic Species Program: Biodiesel from Algae. National Renewable Energy Laboratory. Available from: www.nrel.gov/biomass/pdfs/24190.pdf (accessed 10.01.13.). Singh, S., Singh, D., 2010. Biodiesel production through the use of different sources and characterization of oils and their esters as the substitute of diesel: a review. Renewable Sustainable Energy Rev. 14, 200e216. Singh, A., He, B., Thompson, J., Van Gerpen, J., 2006. Process optimization of biodiesel production using alkaline catalysts. Appl. Eng. Agric. 22, 597e600. 201 Solazyme Inc. 2013. Biotechnology that creates renewable oils from microalgae. Available from: http://solazyme.com/technology (accessed 10.01.13.) Solix BioSystems, 2013. Available from: www.solixbiofuels.com (accessed 10.01.13.). Stauffer, E., Byron, D., 2007. Alternative fuels in fire debris analysis: biodiesel basics. J. Forensic Sci. 52, 371e379. Sun, A., Davis, R., Starbuck, M., Ben-Amotz, A., Pate, R., Pienkos, P.T., 2011. Comparative cost analysis of algal oil production for biofuels. Energy 36, 5169e5179. Takahashi, H., Hirata, S., Minami, H., Fukuyama, Y., 2001. Triterpene and flavanone glycoside from Rhododendron simsii. Phytochemistry 56, 875e879. Tenenbaum, D.J., 2008. Food vs fuel: diversion of crops could cause more hunger. Environ. Health Perspect. 116, A254eA257. Thamsiriroj, T., Murphy, J., 2010. How much of the target for biofuels can be met by biodiesel generated from residues in Ireland? Fuel. 89, 3579e3589. Ugwu, C., Aoyagi, H., Uchiyama, H., 2008. Photobioreactors for mass cultivation of algae. Bioresour. Technol. 99, 4021e4028. U.S. Census Bureau, 2010. Current industrial reports - M311K - fats and oils: production, consumption, and stocks. Available from: http://www.census.gov/cir/www/311/m311k.html (accessed 10.01.13.). U.S. Department of Agriculture, 2010. Oil Crops Yearbook. Available from: usda.mannlib.cornell.edu/MannUsda/viewDocumentInfo. do?documentID¼1290 (accessed 10.01.13.). Varuvel, E.G., Mrad, N., Tazerout, M., Aloui, F., 2012. Experimental analysis of biofuel as an alternative fuel for diesel engines. Appl. Energy. 94, 224e231. Vasudevan, P., Sharma, S., Kumar, A., 2005. Liquid fuel from biomass: an overview. J. Sci. Ind. Res. 64, 822e831. Vishwakarma, S., 2010. Countercurrent extraction of 2, 3-Butanediol. Int. J. Chem. Eng. Appl. 1, 147e150. Wahlen, B.D., Morgan, M.R., McCurdy, A.T., Willis, R.M., Morgan, M.D., Dye, D.J., Bugbee, B., Wood, B.D., Seefeldt, L.C., 2012. Biodiesel from microalgae, yeast, and bacteria: engine performance and exhaust emissions. Energy Fuels 27, 220e228. Weldy, C.S., Huesemann, M.H., 2007. Lipid production by Dunaliella salina in batch culture: effects of nitrogen limitation and light intensity. J. Undergrad. Chem. Res. 7, 115e122. Wenclawiak, B., 1992. Analysis with Supercritical Fluids: Extraction and Chromatography. Springer-Verlag, Munich, Germany. Widjaja, A., Chien, C.C., Ju, Y.H., 2009. Study of increasing lipid production from fresh water microalgae Chlorella vulgaris. J. Taiwan Inst. Chem. Eng. 40, 13e20. Wisniewski Jr, A., Wiggers, V.R., Simionatto, E.L., Meier, H.F., Barros, A.A.C., Madureira, L.A.S., 2010. Biofuels from waste fish oil pyrolysis: chemical composition. Fuel 89, 563e568. Woodgate, S., Van Der Veen, J., 2004. The role of fat processing and rendering in the European Union animal production industry. Biotechnol. Agron. Soc. Environ. 8, 283e294. Yang, B., Zhou, R., Yang, J.G., Wang, Y.H., Wang, W.F., 2008. Insight into the enzymatic degumming process of soybean oil. J. Am. Oil Chem. Soc. 85, 421e425. Yun, Y.S., Lee, S.B., Park, J.M., Lee, C.I., Yang, J.W., 1997. Carbon dioxide fixation by algal cultivation using wastewater nutrients. J. Chem. Technol. Biotechnol. 69, 451e455.
C H A P T E R 13 Use of Volatile Solids from Biomass for Energy Production W.J. Oosterkamp Oosterkamp Oosterbeek Octooien, The Netherlands email: willemjan@oosterkamp.org O U T L I N E Introduction 204 Biodegradability 204 Addition of Macro- and Micronutrients 204 Digestion Systems Family-Size Biogas Plant Wet Digesters Scum Layer Digester Solid Biomass Digester 211 211 211 211 212 Addition of Microbes 205 Increase in Solids Content in Wet Digesters 212 Addition of Enzymes 206 Loading and Unloading of Digesters 212 Pretreatments Biological Pretreatment with Enzymes Chemical Pretreatment Hot Water Treatment Mechanical Pretreatment 207 Treatment of Digestate in Wet Digesters 212 Use of Methane 213 Chemical Conversion of Volatile Solids Combustion Gasification 213 213 213 Longer Retention Times 207 Energy Crops 207 Thermal Conversion of Volatile Solids Slow Pyrolysis Flash Pyrolysis 214 214 214 Food Processing Residues Rice Husks Bagasse Coffee Husks and Mucilage 207 207 207 208 Crop Residues 209 Discussion Maximum Methane Yield Nutrient Recycling Soil Fertility Digesters 214 214 214 214 214 Spent Bedding 209 Conclusions 214 Kitchen and Garden Waste 209 References 215 Aquatic Weeds 209 Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00013-9 207 207 207 207 203 Copyright Ó 2014 Elsevier B.V. All rights reserved.
204 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION INTRODUCTION All-renewable energy resources are required to reduce dependency on fossil fuels from politically unstable regions. Biomass is one such renewable energy resource. Farm and food processing residues are preferred but, where economic, energy plants can be used. Biomass as such cannot replace fossil fuels. Such materials have to be converted into gas, liquid or electricity. Biological volatilizing (anaerobic digestion) converts organic by-products and residues into methane and carbon dioxide, an energy source that can be used for cooking, the production of electricity and as transportation fuel. In Asia there are over 10 million family-size anaerobic digestion plants utilizing manure and some straw. The biogas is used for cooking. There are significant health advantages in using biogas, compared to the local alternative of the burning of cattle manure, leaves and wood inside the houses. There are a few thousand centralized biogas plants in Europe that use manure with a whole range of easily digestible residues. Other biogas plants in Europe use sludge from wastewater cleanup plants. They convert the biogas into electricity and heat. Carbon dioxide is removed from the biogas in a number of recent plants; the gas is compressed and injected into the natural gas grid. The digestate, after the production of biogas, should be used as an organic fertilizer. This will recycle the macro elements nitrogen, potassium, phosphorus and carbon to the soil. Recycling of carbon is essential for high soil productivity and will reverse the trend of lowering of crop yields (Hossain, 2001). The energy content of the animal residues (mostly manure) produced worldwide is equivalent to an average power of 50e150 W per person (9e25 EJ/a). The energy content of crop residues (mostly straw) is also 50e150 W per person (Hoogwijk et al., 2003). Worldwide energy consumption is 2.5 kW per person (500 EJ/a). Oil production worldwide is 1 kW per person (80 million barrels a day). Biogas from straw and manure can replace about 10e30% of the world oil production. This substitution can be doubled by the use of forest residues. only part of it can be depolymerized into soluble components. Anaerobic digestion is a complex process that is slow compared to chemical processes. Chynoweth et al. (1987) have published on the processes involved in the anaerobic digestion of biomass. Hydrolytic bacteria break down the cellulose and hemicellulose into organic acids and neutral compounds. Hydrogen producing bacteria convert the acids into hydrogen. Homoacetogenic bacteria convert hydrogen into acetic acid. Methanogenic bacteria convert acetic acid into methane. A by-product in these conversions is carbon dioxide. Anaerobic biodegradation potential assay is performed by mixing the material with digestate from an operating digester or by mixing the material with a defined nutrient medium according to Owen et al. (1979). The methane produced is measured at different times. Chandler et al. (1980) made a correlation based on 15 different lingocellulosic materials. yCH4 ¼ a  ðb  c  li Þ (13.1) where yCH4 is the methane yield in l/kg volatile solids (VS) a ¼ 440 l/kg is the conversion between methane yield and VS reduction (Jerger et al., 1982). b ¼ 0.83 fitted constant. c ¼ 2.8 fitted constant. li is lignin fraction of the VS This correlation gives a standard deviation of 80 l/kg VS for straws and woody biomass (Table 13.1). A different correlation was developed for straws and woody biomass.  (13.2) y CH4 ¼ a  ð1  li Þ  1  edt d ¼ f * (1  g * li) is exponential factor. f ¼ 0.025 fitted constant. g ¼ 3 fitted constant. This correlation assumes that biodegradation can be described as a first-order process. Shielding of cellulose and hemicellulose by lignin is reflected in the exponential factor. This shielding eventually breaks down. This correlation performs better with a standard deviation of 32 l/kg VS. BIODEGRADABILITY ADDITION OF MACRO- AND MICRONUTRIENTS Agricultural waste materials like straw and the solid fraction of manure are lignocellulosic materials. These materials are strong, flexible and protected against decay. They consist of cellulose, hemicellulose and lignin. Lignin cannot be converted into biogas, and Solubilization of lignocellulosic materials is inhibited by free fatty acids produced during hydrolyzation and subsequent acidogenesis. Increasing the number of methanogenic bacteria reduces the concentration of free fatty acids. Macronutrients nitrogen and phosphate
205 ADDITION OF MICROBES TABLE 13.1 Substrate Biodegradability Lignin (%) Test Period (days) Methane Yield (l/kg) Calculated Methane Yield Chandler et al., 1980 Maize I 8 70 313 298 286 Amon et al., 2003 Maize II 6 70 326 315 312 Amon et al., 2003 Maize III 4 70 287 332 338 Amon et al., 2003 Rice Straw 10 40 240 200 260 Somayai et al., 1994 Barley 13 100 300 299 221 Moeller et al., 2006 Wheat Straw 17 100 270 258 169 Moeller et al., 2006 Maize straw 16 70 200 221 182 Mo et al., 2011 Willow 14 139 289 328 208 Lehtomaeki 2006 Oat Straw 21 150 320 261 117 Lehtomaeki 2006 Rapeseed Straw 20 154 240 277 130 Lehtomaeki 2006 Cotton Wood 25 120 140 174 65 Jerger et al., 1982 Hybrid Poplar 26 120 130 157 52 Jerger et al., 1982 Sycamore 26 120 190 157 52 Jerger et al., 1982 Black Alder 28 120 70 121 26 Jerger et al., 1982 Lignin data for rice straw, barley straw, wheat straw and maize straw are from Godin et al. (2010). The others are from the authors of the biodegradability tests. Godin et al. give for oats straw a lignin value of 0.14. and the micronutrients S, Ca, Mg, Fe, Ni, Co, Mo, Zn, Mn, Se and Cu (Demirel et al.., 2011) are required for the multiplication of methanogenic bacteria. Scherer et al. (1983) determined the chemical composition of a number of methanogenic bacteria (Table 13.2). Scherer (2011) advises also on the minimum concentration of ions in a digester. The basal medium of GuengoerDemirci et al. (2004) and the recommendations of Speece (1987) are similar. Lebuhn et al. (2010) formulated a special cocktail. They commented that for maize silage Co should be added at 0.1 mg/kg VS and sodium at 30 mg/kg VS. Speece (1987) has a recommendation for Se. Se was not limiting in the tests by Lebuhn et al. (2010). Straws, husks, bagasse and woody biomass are generally deficient in macro- and micronutrients. The same holds for cattle manure in Asia as cattle feed mostly on rice straw. Manures in Europe and North America have an excess of nitrogen. Jerger et al., 1982 found a 60% increase in methane yield in half the time in batch-fed anaerobic potential assays with extra-micronutrients (Table 13.3). They also added NH4CL and KH2PO4 to reduce the C/N ratio to 15 and the C/P ratio to 75. Similar results have been obtained by Komatsu et al. (2007). They obtained a methane yield of 280 l/kg VS with sewage sludge and rice straw at a hydraulic retention time of 20 days in a continuously operating digester at 36  C. Somayaji et al. (1994) had 240 l/kg VS in 40 days for rice straw. The addition of micronutrients has an effect of 10e70% on the methane production. Sewage water cleanup sludges are a source of macro and micronutrients. Average primary sewage sludge has the right concentration for Ni and Mo. Co is an order of magnitude too low. In some sludges the concentrations of Fe, Co and Ni are too low (Speece, 1988). Concentrations for Ca, Fe, Zn, Mn and Cu are an order of magnitude too high for its use in agriculture (Wolf et al., 2005). Optimum nutrient conditions are cost-effective. Industrial fertilizers should be used, lacking organic sources of nitrogen and phosphate. For each kilogram of dry lignocellulosic biomass a maximum of 40 g of urea and 20 g of phosphate are required. Human urine is a good source of nitrogen and phosphate. Human feces are also good but require storage for more than 100 days in order to prevent the spread of illnesses. Ecosan toilets (Terefe and Edström, 1999) separate urine and feces, so that urine can be directly used and feces stored for the required period. ADDITION OF MICROBES Op den Camp et al. (1991) describe an acidogenic reactor with rumen-derived bacteria. A hydraulic retention time of 12 h and a solids retention time of 72 h resulted in a methane yield of 440 l/kg VS for cellulose and 120 l/kg VS for barley and rye straw. The
206 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION TABLE 13.2 Nutrients Methanogenic Bacteria Scherer et al., 1983 (g/kg) Speece, 1987 (mg/l) Lebuhn et al., 2010 (mg/l) Scherer (2011) Ions (mg/l) Sewage Sludge Basal Medium Guengor-Demirci and Demirer, 2004 (mg/l) Loeffen et al. Kelly et al., 2005 1984 (g/kg) (g/kg) N 100 800 33 80 P 20 19 23 40 K 25 130 3 S 10 40 16 Ca 3 15 39 Mg 3 4 4 Fe 2 10 0.300 11 11 Ni 0.120 0.1 0.160 2 0.12 0.080 Co 0.065 5 0.220 0.5 2.4 0.004 Mo 0.040 0.1 0.115 0.5 0.26 0.030 Zn 0.340 0.001 0.24 1.740 Mn 0.015 0.260 0.14 0.260 Cu 0.085 0.19 0.850 Se 0.1 0.010 0.05 1.5 0.6 0.01 Author Scherer et al., 1983 Concentrations for methanogenic bacteria Guengor-Demirci et al., 2004 Basal medium Kelly et al., 1984 Sewage sludge; median values of 200 sludge samples in the United States. Values vary by an order of magnitude from sample to sample Loeffen et al. 2005 Sewage sludge data from the Netherlands 2002 TABLE 13.3 Effect of Micronutrients on Wood Substrates Methane Yield (l/kg VS) Woody Biomass Without Micro Elements 120 days With Micro Elements 60 days Cotton Wood 140 200 Hybrid Poplar 130 270 Sycamore 190 220 Black Alder 70 135 methane was produced in a second reactor separated from the first by a filter with 0.03 mm pore size. The liquid without the free fatty acids was recycled to the first reactor. Soluble lignin products (humic acids) inhibited further degradation of the straws. The German company Ares Technology is performing tests at pilot plant scale. Typical conversion yields are around 50% (Strecker, 2012). Weiss et al. (2009) isolated and multiplied hemicellulytic bacteria. These were immobilized on trace metal activated zeolite. Digestion of second-stage sludge from a biogas plant gave a methane yield of 215 l/kg VS after 34 days (35  C) and 150 l/kg VS for the control. ADDITION OF ENZYMES Some European companies (Telschow, 2006; Chollet, 2011) advertise the application of special enzyme combinations in biogas digesters. A 30% faster digestion or a 10% higher biogas yield is reported. Water cleanup secondary sludge is a source of enzymes. The secondary sludge consists mainly of bacteria and the intracellular liquid of these bacteria contains lyses enzymes.
207 FOOD PROCESSING RESIDUES PRETREATMENTS Biological Pretreatment with Enzymes Shredded straws, bagasse and husks are seasonal products and need to be stored before being used as substrate in a digester. Storage with silage can be used to improve the biodigestability of the substrate. Methane yield for maize silage increased from 290 l/kg VS to 330 l/kg VS using the enzyme mixture Microfern (Bossuwe, 2011). Methane yield increased from 145 l/kg VS (fresh reed) to 200 l/kg VS (reed silage prepared with the enzyme mixture Methaplus; Helbig, 2009). Komatsu et al. (2007) report an increase in methane yield from 280 l/kg VS to 310 l/kg VS for rice straw soaked in a solution of an unspecified enzyme codigested with sewage sludge. Chemical Pretreatment Lime (calcium hydroxide) is a relatively cheap chemical and calcium improves the fertility of the soil. In its production about 0.8 kWh/kg high-temperature thermal energy is used. Gunnerson et al.; (1987) advise to compost straw with lime, water and dung. In this method a fraction of VS is lost. Raju et al., 2010 demonstrated an increase of 60% in biogas production using a pretreatment at 0.015 kg Ca(OH)2 per kilogram VS. The pretreatment with 1.5% CaOH is equivalent to an increase in retention time from 32 to 100 days (Moeller et al., 2006). Klopfenstein (1978) found for hemicellulose and cellulose an increase of 80% and 20%, respectively, for sodium hydroxide using corncobs as substrate. The yield increase was only 25% using calcium hydroxide both for hemicellulose and cellulose. Pretreatment with a minimum amount of dilute acids at 50e100  C dissolves the hemicellulose and leaves a solid residue that is highly porous (Tsao, 1987). German biogas tanks have an acid pretreatment (Sauter, 2012). Lebuhn et al. (2010) report technical difficulties with the acid pretreatment and no increase in methane yield. Schober et al. (2006) and Busch et al. (2006) describe an aerated percolation reactor followed by a methanogenese reactor. They report shorter retention times for kitchen and garden waste and maize silage compared to wet systems. Hot Water Treatment Raju et al. (2010) obtained a 40% increase in methane yield using a 15 min pretreatment of wheat and rapeseed straw at 75  C. Mechanical Pretreatment Jerger et al., 1983 found an increase in the methane yield from 270 l/kg VS for particles of hybrid poplar <8 mm to 310 l/kg VS for particles <0.8 mm. The duration of the tests was 90 days. Slotyuk (Oechsner, 2012) found an increase from 230 l/kg VS for 10 mm wheat straw particles to 300 l/kg VS for 1 mm particles. The duration of the tests was 35 days (Table 13.4). LONGER RETENTION TIMES Doubling of the retention times increases the gas yield with 30e50% (Table 13.5). It is unfortunate that the tests were not done at optimum nutrient concentrations. Calculated yields for shorter retention times using Eqns (1) and (2) are compared with measured yields in Table 13.6. The standard deviation between measurement and calculation is 35 l/kg VS (similar to the correlation for longer retention times) (Table 13.7). ENERGY CROPS About 7% of the land used for agriculture in Germany is planted with maize destined for methane production. There are a number of other energy crops with higher production costs (Boese, 2010). Some of these crops have a higher methane yield per hectare (Table 13.8). The humus content of the soils will decrease when only maize is planted as crop (Willms et al., 2009). FOOD PROCESSING RESIDUES Rice Husks The production of rice husks is about 100 million tons per year. Only a fraction of it is used as animal bedding or as fuel for energy production. In Asia briquettes are produced from rice husks. These are expensive to produce, due to the silicon content of the husks. Hill et al. (1981) obtained 110 l/kg VS at a retention time of 17 days. Pretreatment with 8% NaOH gave a methane yield of 200 l/kg VS (Vevekanandan et al., 2011). Bagasse Bagasse is the pressed stalks from sugarcane. It is washed to remove nearly all the sugar in the stalks and leaves the factory at 50% humidity with 5% sugar remaining. World production is 140 million tons (dry weight). Most of the bagasse is used as fuel in the sugar
208 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION TABLE 13.4 Effect of Nutrients Author Process Temperature Test Duration Control Methane (l/kg VS) Test Methane (l/kg VS) RICE STRAW Bardiya et al., 1999 25 ppm Fe Batch Ambient 40 days 63 120 Bardiya et al., 1999 Single dose Ni Batch Ambient 40 days 63 110 20 mM FeSO4 Continuously fed 37  C 16 days 128 180 CATTLE MANURE Preeti Rao et al., 1993  Guengoer-Demirci et al., 2004 Basal medium Batch 35 C 90 days 260 290 Lar et al., 2010 Ca 1.5 g/l Batch 35  C Lar et al., 2010 Fe 0.4 g/l Lehtomaeki, 2006 Ali et al., 2010 25% Jatropha press cake Haque et al., 2006 160 185 50 days 190 225  20 days 150 215 95 115 35 C Batch 30% barley straw 50 days  35 C Continuously fed  27 C Batch  2% dry matter urine Batch 28e35 C 40 days 195 260 10% w/w Glauconite Continuously fed 55  C 15 days 67 90 165 days 250 300 280 380 270 370 SWINE MANURE Hansen et al.,1999  35 C Ahn et al., 2006 3% CaCl Batch Moeller et al., 2006 30% wheat straw Continuously fed 0.2 mg/l Se Continuously fed FOOD WASTE Zhang et al., 2010 TABLE 13.5 35  C Effect of Retention Time on Methane Yield (l/kg VS) Author Material Process Temperature 15 (days) 30 (days) Hansen et al., 1999 Swine manure Continuously fed 55  C 67 180 Shyam 2001 Cattle manure Continuously fed Moeller et al., 2006 Wheat Straw Torres-Castillo et al., 1995 Barley straw Torres-Castillo et al., 1995 Lehtomaeki 2006 Lehtomaeki 2006 Lehtomaeki 2006 50 (days) 100 (days) 22  C 100 160 35  C 190 270 Batch leachate recycling 35  C 145 195 Barley straw Batch leachate recycling 25  C Oats straw Batch 35  C Rapeseed straw Willow Batch Batch factory and some is made into paper and fiberboard. The factories have an excess of bagasse and this is stored in the open air. The stacks produce methane and open fires are common giving off soot and polluting the air. Bagasse has a low biodegradability of 120 l/kg VS (Table 13.8). 160 140 240 300  145 165 200  155 35 C 35 C 200 (days) 240 270 Coffee Husks and Mucilage The production of coffee husks and mucilage is 5 million tons per year. They are dumped near the factories causing methane emissions and degrading the environment. They have a high lignin content of 21%
209 AQUATIC WEEDS TABLE 13.6 Measured and Calculated Methane Yield for Shorter Retention Times Methane (l/kg VS) Lignin days Yield Calculated Wheat Straw 17 50 190 165 Moeller et al., 2006 Barley Straw 13 50 145 205 Torres-Castillo et al., 1995 Oats Straw 14 50 240 195 Lehtomaeki 2006 Oats Straw 14 30 140 195 Lehtomaeki 2006 Rapeseed Straw 20 50 165 140 Lehtomaeki 2006 Rapeseed Straw 20 30 145 91 Lehtomaeki 2006 Willow 14 30 155 135 Lehtomaeki 2006 (Zoca et al., 2012). Padua Ferreira et al. (2011) gives a value of 28%. Biodigestability is 70 l/kg (Frederiks, 2012). Methane capture can be economic using a cheap biodigester and long retention times. The coffee factories can use the methane for electricity production. Coffee farmers can install small digesters and use the methane for cooking. TABLE 13.7 Methane Yield, Costs and Humus Gain for Various Energy Crops Methane Yield (m3/ Methane Costs ha) (V/m3) Humus Gain (kg/ha a) Reye (Grains) 2400 0.39 550 Meadows 2900 0.40 Reye (Whole Plants) 3200 0.34 Grass 3400 0.42 800 Maize Silage 4300 0.30 950 Reye (Whole Plants) and Grass (Intermediate Crop) 4800 0.35 Barley (Whole Plants) and Sorghum (Intermediate Crop) 4800 0.39 Energy Beets 4800 0.42 Rye (Green) and Maize 5200 0.38 CROP RESIDUES The residue from grain crops amounts to nearly 3000 million tons per year (Table 13.8). The most important are maize, wheat and rice. A fraction of this is used as animal fodder or animal bedding. Animal fodder turns into manure and bedding becomes a bedding manure mixture. The fraction that is left in the fields can be used for anaerobic digestion. During threshing the straws are deposited in the fields in swaths and can be picked up after a few days. Losses are between 20% and 50%. The costs for collecting the residues are high. Schmaltschinski (2008) estimates 75 V per ton of shredded straw on the truck at the field in Germany. Mo et al. (2011) report a value of 50 V in Poland. Baled straw sells for 140 V in the Netherlands. Substrate costs are 0.30 V/m3 methane for shredded straw and 0.50 V/m3 for baled straw at 100 days retention times (Table 13.8). SPENT BEDDING Straw used as animal bedding can be collected and digested. It is then an organic fertilizer. The costs for this are mainly transport costs. The use of spent bedding means that all straw can contribute to methane production. Barsega et al. (1994) digested cattle bedding from wheat straw. Spent bedding of rice husks from piggery housings gave less than 20% reduction in VS (Tait et al., 2009). Weathering in a pig shed improved the biodegradability. Bonilla et al. (1985) successfully digested spent poultry litter of rice chaff. Spent wheat bedding from ducks gave a methane yield of 310 l/kg VS (Buisonje, 2009). KITCHEN AND GARDEN WASTE Around 100 kg/person kitchen and garden waste are separately collected in Germany (Kern et al., 2008). VS are around 40%. Extrapolation to the world population would give 150 million tons of VS per year. AQUATIC WEEDS 860 Sewage from most urban areas in tropical countries is discharged untreated in rivers and lakes. This leads to eutrofication. Fast growing aquatic weeds use the nutrients and form thick mats hindering fishing and navigation. Lake Victoria is a prime example. Biological control of the waterweeds is seen as one method to reduce the problems, but does not address the root cause of
210 TABLE 13.8 Substrate Availability Availability (109 kg/a) Rice Husks 100 Uses Fuel, animal bedding Pretreatment 4% NaOH Test Duration (days) Temperature ( C) CH4 Yield (l/kg VS) Author 17 35 140 Hills et al., 1981 40 Ambient 200 Vevekanandan et al., 2011 Joseph et al., 2009 Bagasse 200 Fuel, paper, board 60 35 120 Coffee Husks 10 Fuel 45 30 80 Frederiks, 2012 Wheat Straw 800 Fuel, animal bedding 1.5% CaOH 32 37 260 Raju et al., 2010 None 32 37 170 Raju et al., 2010 None 100 37 270 Moeller et al., 2006 100 37 300 Moeller et al., 2006 Ambient 240 Somayaji et al., 1994 200 Mo et al., 2011 35 240 Lehtomaeki 2006 Barley Straw 200 Animal bedding Rice Straw 700 Animal fodder 40 Maize Straw 1400 Animal fodder 70 Rapeseed Straw 150 Animal bedding 150 Poultry Litter/Rice Straw 52 Ambient 230 Bonilla et al., 1985 Duck Litter/Wheat Straw 50 30 310 Buisonje 2009 Cattle Bedding/Wheat Straw 27 35 270 Barsega et al., 1994 250 Vandevivere et al., 2003 240 Vaidyanathan et al., 1985 Kitchen/Garden Waste 150 Compost Aquatic Weeds 100 Compost 90 30 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION Material
211 DIGESTION SYSTEMS TABLE 13.9 Methane Yield and Test Conditions Water Hyacinth Test Duration (days) Methane Yield (l/kg VS) 125 190 Digested water hyacinth 90 240 Swine manure 60 60 10 Fresh rumen residue 65 250 20e30 10 Fresh rumen residue 95 200 25e35 20 30 20 Author T ( C) Wolverton, 1979 25 Vaidyanathan et al., 1985 29 30 29 Moorhead et al, 1993 35 5 13 Almoustapha et al., 2008 30e40 Almoustapha et al., 2008 Ofoefule et al., 2009 VS (Kg/m3) C/N ratio Seed 17 18 eutrofication. Controlling of one type of waterweed will give others the opportunity to become a pest. There are a number of instances where aquatic weeds are used as an organic fertilizer (Jandl, 2010), but this practice is not widespread. Aquatic weeds are digested in Luzira prison Kampala, Uganda, (Lindsay et al., 2000) and at the Songhai agricultural training centre in Porto Novo Benin (Jandl, 2010). Harvesting of aquatic weeds is expensive. Antunuassi et al. (2002) calculate a cost of 15,000 V/ha or 0.15 V/kg VS. Veitch (2007) suggest significant cost reductions using outboard motor-powered launches with a rake and a land-based backhoe. Anaerobic digestion of whole plants is not common. There are a number of tests with chopped (10e60 mm) water hyacinth (Table 13.9). Gas yields are high for the experiments of Wolverton et al. (1979), Vaidyanathan et al. (1985), and Almoustapha et al. (2008). Swine manure is not well suited as a seed for water hyacinth digestion as it has little biogas bacteria and this explains the low yield of Moorhead et al. (1993). Duration of the tests by Ofoefule et al. (2009) is too low. Moorhead et al., 1993 have done tests with ground (1.6 mm) water hyacinth and chopped water hyacinth (12.6 mm), resulting in a 15% lower gas yield for the chopped water hyacinth. The results for the digestion of whole plants will be also lower than for ground water hyacinth. DIGESTION SYSTEMS Family-Size Biogas Plant Straws and most other biomass have a tendency to float on the water and form a scum layer. Tests in India have shown that these types of materials will generate biogas when they have been in contact with digester fluid for at least three days. This can be achieved by filling the digester with a quantity of wet compost. The compost will float. New material is added from the bottom of the digester. This material will push the compost upward but will stay under the digester liquid, as the compost being wet is relatively heavy. The resulting digested material forms a thick mat. The digested material can be removed once per year after opening the plant (Oosterkamp, 2003). A plant suitable to convert maize stalks and other biomass into biogas is of the water jacket floating drum plant type. The water jacket reduces the emission of methane. In fixed dome plants 10% of the methane production is lost. The main modification is the use of a large 0.3 m diameter inlet pipe. Through this pipe the waste biomass can be pushed down into the digester proper. The floating drum can be removed and digested material taken out from the top. Wet Digesters Most digester systems in Europe are manure based (wet) and/or use maize silage for their feed. Some of these wet systems add about 10% straws (VS). Two or more sequentially linked digesters give about 15% more gas yield than a single continuously stirred digester (Angelidaki et al., 2005). The wet systems have the disadvantage that solids (10e15%) need to be kept in suspension by an impeller. With a high percentage of grasses and straw scum layers are formed that need to be removed mechanically after opening the digester. Scum Layer Digester Sauter (2012) circumvents this problem by spraying digestate on top of the scum layer using water canons. The patent of Rossow (2011) claims that the scum layer can be 5 m in thickness. It is to be expected that digested particles with a high lignin content will be washed down, sink to the bottom and can be collected there using a pump. The input material is loaded into the digester through an input shaft with its exit below the lower level of the scum layer in the digester. It will be acidified for a few days in the first step of the
212 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION digestion process. The acids will be transformed into methane and carbon dioxide in the other parts of the digester. Schoenberg and Linke (2012) tested a 45 l scum layer digester with whole plants (Silphium perfoliatum) as substrate. A loading rate of 8 kg/(m3d) was possible with a methane yield of 215 l/kg VS. Solid Biomass Digester VandeVivere et al. (2003) gave an overview of anaerobic digesters for solid biomass. These “dry” (20e40% solids) systems are used for the digestion of kitchen and garden waste and the mechanically sorted fraction of municipal waste, but can be used for most types of solid farm and food processing by-products. The Kompogas (VandeVivere et al., 2003) system consists of a cylindrical plug flow reactor in which the fermenting wastes are mixed and moved by paddles. Temperature is 55  C and retention time 15 days. Capacity is limited to 10,000 tons/a due to the maximum dimensions of the axis with paddles. Capacity of these systems in Europe and Qatar is more than 1 million tons/a. Methane yields are 200e300 l/kg VS (VandeVivere et al., 2003). There are some batch systems with leachate recirculation. Methane yield is up to 40% less due to channeling of the leachate in the substrate (Vandevivere). Mussoline et al., 2012 obtained 175 l/kg VS for rice straw and swine manure in a digestion period of 1 year. Temperatures ranged from 15 to 35  C. The cell with 6500 m3 capacity was filled with cylindrical bales. Packing density was 100 kg/m3. Storage of straw for 6 months is expensive (25 V/ton). In Western Europe storage cum digester tanks can be used. The idea is to fill these tanks in the period Julye October with shredded straw and a fraction of old digestate together with macro- and micronutrients. Mussoline (2012) demonstrated this concept during one year with rice straw bales and swine manure. Daily power generation was directly correlated to the digester temperature. German tanks with concrete tops are 120 V/m3 and with foil tops 70 V/m3. At a packing density of 10% this amounts to an investment of 700e1200 V/ton straw stored and digested. INCREASE IN SOLIDS CONTENT IN WET DIGESTERS A reduction in digestion plant size can be obtained by increasing the solids content. The amount of work in transporting and spreading the digestate is also reduced. Ong et al. (2000) obtained an increase in gas yield compared to a continuously fed and stirred reactor in a continuously fed nonstirred reactor with the outlet in the middle. It seems that some solids removal from the bottom has to take place, as accumulation of inert solids will reduce the effective volume of the reactor. Shyam (2001) demonstrated that cow manure can be digested at 18% total solids. The method increases the gas yield with 40% using practically the same equipment as before. LOADING AND UNLOADING OF DIGESTERS Manure-based (wet) digesters use pumps for loading. Straw is mixed with digestate and pumped in the digester. An alternative is using an auger and pushing the straw to below the liquid level in the digester. Heavy solids are removed from the bottom of the digester using a rotating rake and pump. Straw disintegrates during digestion and is pumped off together with the digested manure. In the scum layer plant straw can be loaded with front loaders using an inlet shaft with an outlet to below the liquid level. Shredded straw can also be blown into the inlet shaft. Solids are removed in the same way as in wet digesters. The remaining digestate is pumped out. Plug flow dry digesters are loaded and unloaded with augers. Leachate recirculation digesters are operated in batch mode. The digesters are opened and the digestate is unloaded with front loaders. Capacity of these digesters is relatively low due to the loading procedure. TREATMENT OF DIGESTATE IN WET DIGESTERS A screw press or a centrifuge can separate the digestate into a solid and a liquid fraction. Recycling of the solid fraction will increase the solid content of the digesters further. The recycled solid fraction will yield between 30 and 100 l/kg VS depending on the pH and protein content (Balsari et al., 2010). Disposal of the liquid fraction into a sewer or into surface waters requires the removal of phosphate and nitrogen. Phosphate can be separated by the addition of magnesium and precipitation of magnesium ammonium phosphate (struvite). The production of ammonium requires about 30 MJ/kg. The ammonium can be recovered by adding extra phosphate and magnesium to the effluent. Tuerker and Celen (2007) give a cost for chemicals of 7.7 $/kg N removed (price level of 2001) for magnesium chloride and phosphoric acid and sodium hydroxide. About a third of the cost is for the sodium hydroxide
CHEMICAL CONVERSION OF VOLATILE SOLIDS necessary for the adjustment of pH. Struvite is not a conventional fertilizer and the price is quite speculative, but it should be at least the value of the phosphate 2.9 $/kg N. Dvorak et al., 2011 heat the liquid fraction of digestate up to 70  C and use aeration for the removal of CO2. This results in an increase in the pH to nearly 10 and a shift of the ammonia ions to gaseous ammonia. The diet of dairy cows is rich in calcium. Consequently the phosphate in the digestate is in the form of insoluble calcium phosphate particles. CO2 bubbles keep these particles in suspension. Elimination of the CO2 results in settling of the particles in a quiescent tank. The system removes 70% of ammonia and 80% of phosphate. End products are a solution (35%) of ammonium sulfate and phosphatecontaining solids. Karakashev et al. (2005) came to the conclusion that microfiltration is unsuited for treatment of digested pig manure due to membrane clogging. They developed a method at laboratory scale to clean the supernatant after decanting-centrifuging. It involves a up-flow anaerobic sludge blanket reactor, precipitation of magnesium-ammonia-phosphate (struvite) by adding magnesium oxide, partial aeration and ammonium removal by anaerobic ammonia-oxidizing bacteria. Phosphate and nitrogen can also be concentrated by removing water from the liquid fraction. Waste heat (50e70  C) from electricity generation removes only fraction of the water in a single pass. Up to three passes are possible. Distillation with vapor recompression has been tried (Melse et al. 2005). The electric energy consumption was about 0.3 MJe/kg water removed. Technical and economical reasons led to abandonment of the process. The Biorek (Preez et al., 2005) process uses a two-step filtration and reverse osmosis process to increase the solids content. One project in the Netherlands was stopped due to operational difficulties. The 20 MWe biogas plant in Penkum, Germany, uses a decanter followed by a swinging screen for the removal of solids and evaporation and reverse osmosis for the removal of salts (Herbes, 2010). USE OF METHANE At present most methane from anaerobic digestion is used for cooking and in gas or dual fuel engines (diesel engines where part of the fuel is substituted by biogas) to generate electricity with an overall efficiency of 30e50%. In Western Europe part of the excess heat is used in residences and factories. In a few instances fuel cells are used to generate electricity and hightemperature process heat. It is, however, better to remove the carbon dioxide and to inject the gas into 213 the natural gas grid. It can then be used to generate electricity in 60% efficient combined cycle plants In Europe and in several states of the United States there are requirements to gradually introduce biomassderived fuels in the transport sector. Approximately 5 million cars currently run on compressed natural gas and could run on compressed methane from anaerobic digestion. An alternative is to liquefy the gas and use the liquefied gas as biofuel in vehicles. This is done in Snurrevarden (Norway) and Gasum (Finland). Anaerobic digestion and the use of compressed methane is more energy efficient than the hydrolysis of cellulose and hemicellulose to sugars and conversion of these sugars into alcohol. This alcohol has to be distilled in order for it to be used as a transportation fuel. In areas where there is no natural gas infrastructure, methane in high-pressure bottles can replace bottled liquefied petroleum gas (LPG or propane). Energy densities of 20% of that of LPG bottles can be reached at a pressure of 4 MPa using bottles filled with activated carbon. CHEMICAL CONVERSION OF VOLATILE SOLIDS Combustion It is estimated that around 3 billion people worldwide rely on wood, stubble, dung and leaves for cooking fuel. Burning biomass fuels on open fires and in inefficient stoves releases many harmful pollutants. These pollutants result in excess respiratory illnesses and death in women and children. Known as a “silent killer”, over 1.6 million children die annually throughout the developing world from the consequences of exposure to biomass fuel smoke (Edelstein et al., 2008). Improved stoves reduce the fuel consumption and indoor pollution by 50% (Ravindranath et al., 1997; Halim, 2008). Co-combustion of solid biomass and coal is reviewed by Cremers (2009). Combustionesteam cycle. This combustion of solid biomass and the use of a steam cycle is not very energy efficient (32%, Yang et al., 2006). The maximum temperature is limited as potassium and calcium together with silicon form at high temperatures glasslike deposits on the furnace walls. Corrosion problems occur in strawbased furnaces at 500  C (Hansen et al., 2000). Gasification Gasifying cook stoves are described by Field (2012). A high-pressure liquid ash gasifyer has an efficiency
214 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION of 50%. These are large installations with capacities of over 200 MWe. At present only 15% biomass is cogasified with coal (Drift, 2008). THERMAL CONVERSION OF VOLATILE SOLIDS Slow Pyrolysis Only a few percent of chlorine and potassium are carried in the pyrolysis gas and bio-oil at a pyrolysis temperature of less than 700  C (Jensen et al., 1999). About 40% of the energy is at a temperature of 700  C in the gas fraction. Fifty percent of the energy remains in the biochar in slow pyrolysis at 400  C. Total efficiency is 20% (Halwachs, 2010). Flash Pyrolysis In flash pyrolysis the production of liquid bio-oil is maximized using high temperatures and high heating rates. The liquid fraction of fast pyrolysis products contains a high concentration of potassium and calcium and its use in diesel engines and gas turbines is not prudent. DISCUSSION Maximum Methane Yield Jerger et al. (1982) demonstrated that a correct ratio of macronutrients and minimum values of micronutrients offer a way to increase the methane yield at low costs. Tests on the biodegradability should be made with optimal concentrations of macronutrients and sufficient micronutrients. Some companies offer the service to test for these conditions in biogas plants. It would be wise for Indian cattle manure plants to formulate packages with the required macro- and micronutrients. Komatsu et al. (2007) showed that primary sewage sludge is an excellent medium of both macro and micronutrients. Secondary sludge not only has methanogenic bacteria but also contains enzymes that break down straws, bagasse and husks. They used 65% VS sewage sludge and demonstrated that high concentrations of micronutrients are not harmful. The codigestion of lignocellulosic materials with sewage sludge is attractive. The proteins in the sludge are difficult to digest. Codigestion reduces the VS more than digestion the sludge and lignocellulosic materials separately (Rashed et al., 2008). Longer retention times will increase the methane yield further given optimal nutrient conditions. This requires more or larger digester tanks. The tanks form only 20% of the total investment in a biogas installation and retention time can be increased fivefold when the methane yield is doubled for the same investment per cubic meter methane. The size for concrete tanks is limited to around 3000 m3 and multiple tanks are normally required. Much work has been done on mechanical, chemical or biological pretreatment of straws, bagasse and husks. These are effective giving up to 50% more methane. No comparison has been made with longer retention times in terms of investments and operational costs. Enzyme addition when making silage and boiling water addition during loading of straw and husks are the next most cost-effective treatments. This involves no extra handling of the substrates. Nutrient Recycling Nutrient recycling back to the soil is possible with anaerobic digestion. Most of the nitrogen and phosphate is in the liquid fraction of the digestate, when the digestate is separated in a solid and liquid fraction. The liquid fraction could be spread at nearby fields. Lignocellulosic biomass should be codigested with sewage sludge, where nutrient recycling is not economic. Soil Fertility Soil fertility is enhanced by humus. Humus is in turn produced by the degradation of lignin. The lignin remains in the digestate in anaerobic digestion and can be used as “fertilizer”. A similar effect is obtained by the biochar from the pyrolysis of lignocellulosic biomass (terra preta). No comparison has been made on investment costs and operational costs between pyrolysis and anaerobic digestion. Digesters Solid biomasses need to be digested at optimum nutrient conditions and long retention times for maximum methane yields. The wet scum layer digester is one option. The maximum solids concentration in the systems without giving operational difficulties has not been established. Batch systems with leachate recycling are also an option. Substrate handling is minimized. The disadvantage is their small size and the danger of explosions during opening of the digesters. CONCLUSIONS Anaerobic digestion of food processing and crop residues can contribute to reduce the dependency on fossil fuels. Energy recovery in methane is 70% with digestion periods between 100 and 150 days, depending on the
REFERENCES lignin content of the substrates and temperature of the digestion. Digestion times can be reduced with optimum concentrations of macro and micronutrients. Food processing residues are a cheaper substrate for anaerobic digestion than crop residues, as collection has already been paid for. Systems using an auger or paddles to transport the substrate inside the digester tank have a significant higher methane yield than batch systems with leachate recycling. Their investment warrants only their use for kitchen and garden wastes, for which a gate fee is paid. Losses in batch systems may be reduced by longer retention times. Storage cum digester tanks with leachate recycling will reduce substrate handling to a minimum. References Ahn, J.H., Do, T.H., Kim, S.D., Hwang, S., 2006. The effect of calcium on the anaerobic digestion treating swine waste water. Biochem. Eng. J. 30, 33e38. Ali, N., Kurchania, A.K., Babel, S., 2010. Bio-methanisation of Jatropha curcas defatted waste. J. of Eng. Tech. Res. 2 (3), 038e043. Almoustapha, O., Kenfack, S., Millogo-Rasolodimby, J., 2008. Biogas production using water hyacinth to meet collective energy needs in a Sahelian country. Fact Rep. 1, 72e79. Amon, Th., Kryvoruchko, V., Amon, B., Zollitsch, W., Mayer, K., Buga, S., Amid, A., November 2003. Biogaserzeugung aus mais e einfluss der inhaltstoffe uaf das specifische methanbildungsvermoegen von fruhe- bis spaetreifen maisorten. In: Rueckenbauer (Ed.), Arbeitstagung 2003 der Vereinigung Oestereichischer Planzenzuechter und Saatgutkaufleute zum thema Hybridmais, Zuechtung und Verwertung. HBLFA, Raumberg-Gumpenstein, Austria, pp. 25e27. (November). Angelidaki, A., Ellegaard, L., 2005. Biogas reactors in series may yield up to 15% more biogas. Bioenergy Res. 9, 2e3. Antunuassi, U.R., Velini, E.D., Martins, D., 2002. Remocao mechanica de plantas aquaticas: analise economica e operacional. Planta Daninha 20, 35e42. Balsari, P., Gioelli, F., Menardo, S., Paschetta, E., 2010. The (re)use of mechanical separated solid fraction of digested or not digested slurry in anaerobic digestion plants. In: Marques dos Santos, C.S.C., Ferreira, C.L. (Eds.), Proceedings Fourteenth Ramiran Conference Lisboa, Portugal 13e15 September 2010. Universidade Technical de Lisboa, Portugal, pp. 1e4. Bardiya, N., Gaur, A.C., 1999. Iron supplementation enhances biogas generation. Bio Energy News 1, 6e19. Barsega, U., Egger, K., Wellinger, A., 1994. Biogas aus festmist, entwicklung einer kontinuierlich betriebene biogasanlage zur vergaerung von strohreichen mist. FAT Berichte 451, 1e12. Boese, S., 2010. Kosten sind nicht alles. Praxisnahe 4, 8e9. Bonilla Garcia, S., Georg, R., Hoffmann, R., Ulrich, G., 1985. Fermentation of poultry excrement/rice chaff mixtures. Gate 4, 1e3. Bossuwe, M., 2011. Voorbehandeling van Mais bij Inkuiling voor Toename van de Biogasproductie. HoWest Kortrijk, Belgium. pp. 1e121. Buisonje, F., 2009. Stromest van vleeseenden. V-Focus 4, 36e37. Busch, G., Sieber, M., 2006. Zweistufiges fest-fluessig-biogasverfahren mit offener hydrolyse eein neues technologisches konzept fuer die biogasgewinnung aus nachwachsender rohstoffe und bioerfuegbare abfaellen. Forum der Forschung 19, 63e68. Chandler, J.A., Jewell, J.M., Van Soest, P.J., Robertson, J.B., 1980. Predicting Methane Fermentation Biodegradability. Solar Energy Research Institute, Golden CO, USA. pp. 1e234. 215 Chollet, J.D., 2011. Application de biocatþ dans le Digesteur de Kompogas de Germanier Encore Cyclage S.A. Lavigny Morges. Citadel Biocat, Moreg, Switserland. pp. 1e4. Chynoweth, D.P., Isaacson, R., 1987. Anaerobic Digestion of Biomass. Elsevier Applied Science, London, United Kingdom. pp. 1e279. Cremers, M.F.G., 2009. Technical Status of Biomass Co-Firing. Kema, Arnhem, Netherlands. pp. 1e43. Demirel, B., Scherer, P., 2011. Trace element requirement of agricultural biogas digesters during biological conversion of renewable biomass to methane. Biomass Bioenergy 35, 992e998. Drift van der, A., 2008. Status of Biomass Gasification. ECN Petten, Netherlands, 1e24. Dvorak, S.W., Chen, S., Frear, C., VanLoo, B.J., Zhao, Q., 2011. Nutrient recovery systems and method. Patent application WO/2011/156767. Edelstein, M., Pitchforth, E., Asres, G., Silverman, M., Kulkarni, N., 2008. Awareness of health effects of cooking smoke among women in the Gondar Region of Ethiopia: a pilot survey. BMC Int. Health Hum. Rights 8 (10), 1e4. Field, J., 2012. Gasifying Cookstoves Database. Colorado State University Fort Collins, USA. pp. 1e24. Frederiks, B., 2012. Biogas Tests with Euphorbia Trirucali, Sugar Filter Mud, Coffee Husk, Banana Skin, Gras Manure and Maize Mixture. Fact Foundation, Wageningen, Netherlands. pp. 1e3. Godin, B., Ghysel, F., Agneesens, R., Schmitt, T., Gofflot, S., Lamaudiere, S., Sinnaeve, G., Goffart, J.-P., Gerin, P.A., Stilmant, D., Delcarte, J., 2010. Determination de la cellulose, des hemicelluluse, de la lignine et des cendres dan diverses cultures lignocelluloses dediees a la production de bioethanol de deuxieme generation. Biotechnol. Agron. Soc. Environ. 14, 561e566. Guengoer-Demirci, G., Demirer, G.N., 2004. Effect of initial COD concentration, nutrient addition, temperature and microbial acclimation on anaerobic treatability of broiler and cattle manure. Bioresour Technol. 93 (2), 109e117. Gunnerson, C.G., Stuckey, D.C., 1987. Integrated resource recovery: Anaerobic digestion; Principles and practices for biogas systems. UNDP, New York, USA., 1e154. Halim, S.A., 2008. A Technical Manual for Improved Cooking Stoves. Village education resource center, Dhaka, Bangladesh. pp. 1e68. Halwachs, M., 2010. Rotary Kiln Pyrolysis e First Results of a 3 MW Plant. Technical University Wien, Austria. pp. 1e21. Hansen, L.A., Nielsen, H.P., Frandsen, F.J., Dam-Johansen, K., Horlyck, S., Karlsson, A., 2000. Influence of deposit formation on corrosion at a straw fired boiler. Fuel Process. Technol. 63 (1e3), 189e209. Haque, M.S., Haque, M.N., 2006. Studies on the effect of urine on biogas production. Bangladesh J. Sci. Ind. Res. 41 (1), 23e32. Helbig, S., 2009. Biogas Production from Common Reed. Mariental Desert Foundation of Namibia, Windhoek, Namibia. pp. 1e46. Herbes, C., 2010. Biomethane Production on Industrial Scale. Nawaro bioenergie, Leipzig, Germany. pp. 1e19. Hills, D.J., Roberts, D.W., 1981. Anaerobic digestion of dairy manure and field crop residues. Agric. Wastes. 3, 179e189. Hoogwijk, M., Faaij, A., Van den Broek, R., Berndes, G., Gielen, D., Turkenburg, W., 2003. Exploration of the ranges of the global potential of biomass for energy. Biomass Bioenergy 25, 119e133. Hossain, M.Z., 2001. Farmers view on soil organic matter depletion and its management in Bangladesh. Nutrient Cycling in Agroecosystems 61, 197e204. Jandl, O.M., 2010. Barriers for the Employment of Floating Invasive Weeds for Biogas Production in Local Communities in West African Developing Countries. Thesis Technical University Eindhoven, Netherlands. pp. 1e66. Jensen, P.A., Dam-Johansen, K., 1999. Release of potassium and chlorine during straw pyrolysis. In: Overend, R.P., Chornet, E.
216 13. USE OF VOLATILE SOLIDS FROM BIOMASS FOR ENERGY PRODUCTION (Eds.), Fourth Biomass Conference of the Americas, Oakland, USA. Elsevier Science, Oxford, UK, pp. 1169e1176. Jerger, D.E., Conrad, J.R., Fannin, K.F., Cynoweth, D.P., 1982. Biogasification of woody biomass. In: White, J.W., McGrew, W., Sutton, M.R. (Eds.), Energy from Biomass and Wastes VI. Institute of Gas Technology, Chicago Il, USA, pp. 341e372. Jerger, D.E., Dolenc, D.A., Chynoweth, D.P., 1983. Biogasification of woody biomass following physical and chemical pretreatment. In: Klass, D.L., Eliott (Eds.), Seventh Symposium on Energy from Biomass and Wastes Orlando. Institute of Gas technology, Chicago, Ill. USA, pp. 233e234. Joseph, O., Rouez, M., Meltivier-Pignon, H., Bayard, R., Emmanuel, E., Gourdon, R., 2009. Adsorption of heavy metals on sugarcane bagasse: improvement of adsorption capacities due to anaerobic degradation of the biosorbent. Environ. Technol. 30, 1371e1379. Karakashev, D., Batstone, D.J., Angelidaki, I., 2005. Influence of environmental conditions on methanogenic compositions in anearobic biogas reactors. Appl. Environ. Microbiol. 71 (1), 331e338. Kelly, W.D., Martens, D.C., Reneau, R.B., Simpson, T.W., 1984. Agricultural Use of Sewage Sludge: A Literature Review. Bulletin 143. Virginia Water Resources Research Centre, Gloucester Point Va, USA, pp 1e47. Kern, M., Siepenkothen, J., 2008. Potenziale fuer die erzeugung von biogas in der deutschen abfalwirtschaft. In: ThomeKozmiensky, K.J., Beckmann, M., Versteyl, A. (Eds.), Berliner Abfallwirtschafts- und Energiekonferenz Ersatzbrennstoffe und Biogas, pp. 495e505. Berlin. Klopfenstein, T., 1978. Chemical treatment of crop residues. J. Anim. Sci. 46, 841e848. Komatsu, T., Kudo, K., Himeno, S., 2007. Anaerobic digestion of sewage sludge and rice straw. In: LeBlanc, R.J., Laughton, P.J., Rajesh, T. (Eds.), Moving Forward Waste Water Biosolids Sustainability Technical, Managerial, and Public Synergy June 24e27, 2007. Moncton, Greater Moncton Sewerage Commission Moncton, NB Canada, pp. 495e501. Lar, J.S., Li, R., Li, X., 2010. The influence of calcium and iron supplementation on the methane yield of biogas treating dairy manure. Energy Sources Part A 32, 1651e1658. Lebuhn, M., Andrade, D., Bauer, C., Gronauer, A., 2010. Intensivierung des anaeroben Biomassenabbaus zur Methanproduktion aus Nachwachsende Rohstoffen. LfL Tier und Technik Freising, Germany. pp. 1e132. Lehtomaeki, A., 2006. Biogas Production from Energy Crops and Crop Residues. Dissertation. University of Jyvaeskylae Finland. pp. 1e85. Lindsey, K., Hirt, H.-M., 2000. Use Water Hyacinth: a Practical Handbook of Uses for Water Hyacinth from across the World. Aname, Winnenden Germany. pp. 1e114. Loeffen, P., Geraats, B., 2005. Toekomstige Kwantiteit en Kwaliteit van ZuiveringsslibStowa Rapport 2005e2006. Stowa, Amersfoort, Netherlands. pp. 1e96. Melse, R.W., Verdoes, N., 2005. Evaluation of four farm-scale systems for the treatment of liquid pig manure. Biosyst. Eng. 92 (1), 47e57. Mo, Z., Polarski, K., 2011. Preliminary comparison of biogas productivity from maize silage and maize straw silage. J. Res. Appl. Agric. Eng. 56 (2), 108e110. Moeller, H.B., Nielsen, A.M., 2006. Straw and energy crops in biogas plants. BioEnergy Res. 14, 4e5. Moorhead, K.K., Norstedt, R.A., 1993. Batch anaerobic digestion of water hyacinth: effect of particle size, plant nitrogen content and inoculum volume. Bioresour. Technol. 44, 71e76. Mussoline, W.A., Esposito, G., Giordano, A., 2012. Optimization of Anaerobic Digestion of Rice Straw Inoculated with Piggery Waste. University of Paris-Est, France. pp. 1e2. Oechsner, H., 2012. Forschungsinitiative und eprojecte zur biogasforschung. In: Kranert, M. (Ed.), 8e Biogastag Baden Wuertenberg 13 Maerz 2012 Hohenheim. Oldenbourg Industrie Verlag, Essen Germany, pp. 1e144. Ofoefule, A.U., Uzodinma, E.O., Onukwuli, O.D., 2009. Comparative study of the effect of different pre-treatment methods on biogas yield from water hyacinth (Eichhornia crassipies). Int. J. Phys. Sci. 4 (8), 535e539. Ong, H.K., Greenfield, P.F., Pullammanappallil, P.C., 2000. An operational strategy for improved biomethanation of cattle manure slurry in an unmixed single stage digester. Bioresour. Technol. 73, 87e89. Oosterkamp, W.J., 2003. Werkwijze voor het vergisten van biomassa. Patent Nl 1023558. Bureau voor Industrieel Eigendom, Rijswijk, Netherlands, pp 1e7. Op den Camp, H.J.M., Gijzen, H.J., 1991. The Rudad process for enhanced degradation of solid waste materials. In: Martin, A.M. (Ed.), Biological Degradation of Waste. Elsevier Science Publishers, Barking, England, pp. 281e306. Owen, W.F., Stucley, D.C., Healy jr, J.B., Young, L.Y., McCarty, P.L., 1979. Bioassay for monitoring biochemical methane potential and anaerobic toxicity. Water Res. 13 (6), 485e492. Padua Ferreira, R.V. de, Sakata, S.K., Bellini, M.H., Marumo, J.T., 2011. Biosorption of Am-241 and CS-137 by Radioactieve Liquid Waste by Coffee Husk. IPEN, Sao Paulo, SP, Brazil, 1e5. Preeti Rao, P., Seenayya, G., 1993. Improvement of methanogenesis from cow dung and poultry litter waste digesters by addition of iron. World J. Microbiol. Biotechnol. 10, 211e214. Preez, J., du, Norddahl, B., Christensen, K., 2005. The biorek concept: a hybrid membrane reactor concept for very strong wastewater. Desalination 183, 407e415. Raju, C.S., Ward, A.J., Moeller, H.B., 2010. The effect of thermochemical pre-treatment on the ultimate biogas potential of straw. In: Marques dos Santos, C.S.C., Ferreira, C.L. (Eds.), Proceedings Fourteenth Ramiran Conference Lisboa, Portugal 13e15 September 201. Universidade Technical de Lisboa, Portugal, pp. 1e4. Rashed, E.M., Fouad, H.N., Awadalla, M.S., 2008. Utilization of agricultural residues in improving sludge digester efficiency. J. Appl. Sci. Res. 4 (12), 2127e2133. Ravindranath, N.H., Ramakrishna, J., 1997. Energy options for cooking in India. Energy Policy 25 (1), 63e75. Rossow, N., 2011. Verfahren und vorrichtung zur erhoehung der besiedlungsdichte von methanbildende bakterienstaemme in biogasfermentern. European patent EP2314666 A1. Sauter, 2012. Personal communication. Scherer, P., 2011. Wirkungsweise von spurenelementen in der biovergasungskette. In: Binder, W. (Ed.), Fachtagung Spurenelementen in Biogasanlagen. Energie Agentur, Goetingen, Germany, pp. 1e47. Scherer, P., Lippert, H., Wollf, G., 1983. Composition of the minor elements and trace elements of 10 methanogenic bacteria determined by inductively coupled plasma emission spectrometry. Biol. Trace Elem. Res. 5 (3), 149e163. Schmaltschinski, T., 2008. Optimierung des Strohtransportes zu einem Saege- und Spanplattenwerk Abslussbericht zum Forschungsproject (11D/07Keil). Institut Fuer Forstbenutzung und Forstliche Arbeitswissenschaft, Freiburg, Germany. pp. 1e21. Schober, G., Wellinger, A., Widmer, C. 2006. Bau und Betrieb einer Perkolationsanlage im Pilotmassstab zur Aufbereitung von Bioabfalle Projekt Nr. 38714. Nova Energie, Aadorf, Switserland, pp. 1e34. Schoenberg, M., Linke, B., 2012. Prozessoptimierung in der zweiphasigen/zweistufigen vergaerung festerbiomassen. Bornimer Agrartechnischen Ber. 79, 34e44. Shyam, M., 2001. A biogas plant for the digestion of fresh undiluted cattle dung. Boiling Point 47, 33e35. Somayaji, D., Khanna, S., 1994. Biomethanation of rice and wheat straw. World J. Microbiol. Biotechnol. 10, 521e523.
REFERENCES Speece, R.E., 1987. Nutrient requirements. In: Chynoweth, D.P., Isaacson, R. (Eds.), Anaerobic Digestion of Biomass. Elsevier Applied Science London, United Kingdom, pp. 109e127. Speece, R.E., 1988. A survey of municipal anaerobic sludge digesters and diagnostic activity assays. Water Res. 22 (3), 365e372. Strecker, M., 2012. In: Willingmann, A. (Ed.), Das Ruminotec verfahren zur Gewinnung von Biogas aus Primaer Cellulosehaltigen Reststoffen, pp. 1e55. Jahrestag des BWK 24 October, 2012. Regiona, Wernigrode, Germany. Tait, S., Tarnis, J., Edgeton, B., Batstone, D.J., 2009. Anaerobic digestion of spent bedding from deep litter piggery housing. Bioresour. Technol. 100 (7), 2210e2218. Telschow, D. 2006. Untersuchung des Einflusses von Enzymen auf die Biogasbildung thesis. Bauhaus Universitaet, Weimar, Germany. Terefe, A., Edström, G., 1999. Ecosan-ecological sanitation. In: Pickford, J. (Ed.), Twenty fifth WEDC Conference Integrated Development for Water Supply and Sanitation, pp. 16e17. Addis Ababa, Ethiopia; 1999 WEDC Loughborough University Leicestershire LE11 3TU, UK. Torres-Castillo, R., Llabres-Luengo, P., Mata-Alvarez, J., 1995. Temperature effect on anaerobic digestion of bedding straw in a one-phase system at different inoculum concentration. Agric., Ecosyst. Environ. 54, 55e66. Tsao, G.T., 1987. Pre-/posttreatment. In anaerobic digestion of biomass. In: Chynoweth, D.P., Isaacson, R. (Eds.), Anaerobic Digestion of Biomass. Elsevier Applied Science, London, United Kingdom, pp. 91e107. Tuerker, M., Celen, I., 2007. Removal of ammonia as struvite from anaerobic digester effluents and recycling of magnesium and phosphate. Bioresour. Technol. 98, 1529e1534. Vaidyanathan, S., Kavadia, K.M., Schroff, K.C., Mahajan, S.P., 1985. Biogas production in batch and semicontinuous digesters using water hyacinth. Biotechnol. Bioeng. 28, 905e908. Vandevivere, P., Baere, L., Verstraete, W., 2003. Types of anaerobic digesters for solid wastes. In: Mata-Alvarez, J. (Ed.), 217 Biomethanization of the Organic Fraction of Municipal Solid Wastes. IWA publishing, London, pp. 336e367. Veitch, V.D., Burrows, Hudson, 2007. Trialling different low cost methods of water hyacinth removal in tropical coastal wetlands. In: Wilson, A.L., Dehaan, R.I., Watts, R.J., Page, K.J., Bowmar, K.H., Curtis, A. (Eds.), Proceedings of the Fifth Australian Stream Management Conference: Australian Rivers Making a Difference. Charles Stuart University, Thurgoona, NSW, Australia, pp. 407e412. Vevekanandan, S., Kamaraj, G., 2011. Investigation on cow dung as co-substrate with pretreated sodium hydroxide on rice chaff for efficient biogas production. Int. J. Sci. Adv. Technol. 1 (4), 76e80. Weiss, S., Tauber, M., Somitscht, M., Meincke, R., Mueller, H., Berg, G., Guebitz, G.M., 2009. Enhancement of biogas production by addition of hemicellulolytic bacteria immobilised o activated zeolite. Water Research 44 (6), 1970e1980. Willms, M., Deumlich, D., Hufnagel, J., Reinicke, F., Wagner, B., Buttlar, C., 2009. Anbauverfahren fuer energiepflanzen e auswirkungen auf boden und umwelt. In: Schuette, A. (Ed.), Tagungsband zum KTBL/FNR biogas-Kongress von 15 bis 16 September 2009, Weimar. Fachagentur fuer nachwachsend Rohstoffe, Guelzow-Pruesen Germany, pp. 148e162. Wolf, P de, Heeres, E., Postma, J., 2005. Compost voor de Biologische Kringloop. Praktijkonderzoek plant en omgeving, Wageningen, Netherland, 1e22. Wolverton, B.C., McDonald, R.C., 1979. Energy from Aquatic Waste Water Treatment Systems. NASA, United States. Yang, Y.B., Shariff, V., Swithenbank, J., 2006. Simulation of a Large-Scale Straw-fired Furnace. University of Sheffield England. pp. 1e18. Zhang, Y., Walker, M., Banks, C.J., 2010. Optimising Processes for the Stable Operation of Food Waste Digestion Defra WR1208. University of Southampton, England. pp. 1e111. Zoca, S., Rosolem, C., 2012. Potential Benefits of Coffee Cherry Applications to Agricultural Soil. Okalahoma State University Stilwater, O.K., USA, pp 1e3.
C H A P T E R 14 Biorefinery Systems: An Overview Maria Gavrilescu Department of Environmental Engineering and Management, Gheorghe Asachi Technical University of Iasi, Iasi, Romania; Academy of Romanian Scientists, Bucharest, Romania email: mgav@tuiasi.ro O U T L I N E IntroductiondBiorefinery, Concepts and Emerging Opportunities for Sustainable Economy 219 Short History of Biorefineries and Bio-Based Products 221 Biomass Feedstock 221 Structure of Biorefinery Concept 224 INTRODUCTIONdBIOREFINERY, CONCEPTS AND EMERGING OPPORTUNITIES FOR SUSTAINABLE ECONOMY In a continuous developing world, the industrial system needs to sustain an increasing Earth population, which poses a high pressure on planet biocapacity. It is generally recognized that the current industrial system which generates products and services required by the society is not sustainable: Earth continues to be depleted of its resources at such rates that need to be diminished. In addition, the current production efficiency is <10%, while 90% of the material resources used in the production process end up as waste, with high impacts in the environment. Furthermore, climate changes require significant minimization in the current greenhouse gases emissions, by using new technologies able to bridge the gap between the economic growth and environmental sustainability as well as alternative sources of energy (WEF, 2010). The rapid increase in energy requirement run in parallel to the technological development, so as R&D activities are encouraged to study new Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00014-0 Biorefinery Platforms 227 Biorefinery Eco-Efficiency 231 Concluding Remarks and Perspectives 236 References 239 and biorenewable energy resources (Demirbas, 2010). Recent data show that 25% of the population from industrialized countries consumes about 75% of the world natural resources, and controls about 88% of the total production, 80% of trade and 94% of industrial products of the entire world (Hens and Quynh, 2008). These pressures on the planet biocapacity associated to the unprecedented change in the climate and trends in fossil fuel depletion generate premises to study and evaluate the ways of producing and process the natural resources in an efficient and sustainable way to ensure an environmentally friendly and eco-efficient development (EEA, 1999; PP, 2012). Designing for sustainability of smart production systems should be based on raw materials renewability, correlated with the rate of use. Therefore the design of production processes, products and services needs to be developed on a broader perspective having economic, technical, social and environmental efficiency, effectiveness, and performance as red lines in minimizing the impact on humans and environment (Centi and Perathoner, 2009; Yuan et al., 2013). The biological feedstock is an essential part of the smart economy, since the preservation and management 219 Copyright Ó 2014 Elsevier B.V. All rights reserved.
220 14. BIOREFINERY SYSTEMS: AN OVERVIEW of various resources are fundamental strategic lines to foster sustainability tasks in the twentieth century (Kamm et al., 2006). In this context, biotechnology becomes more and more a sustainable alternative to various industrial sectors, in particular chemical industry since some shifts from production of goods and services from fossil to bio-based materials are essential (Gavrilescu and Chisti, 2005). This would also constitute the seeds of a new synergy of biological physical, chemical and technical sciences, as was pointed by Kamm et al. (2006). So, biomass has been and continues to be seen as the most promising carbon neutral source of energy able to mitigate greenhouse emissions because the quantity of CO2 released during combustion is the same as that absorbed by plants during photosynthesis (Demirbas, 2010; Khan et al., 2009). Biomass is also considered environmentally sustainable since it is derived prevalent from industrial waste or resulted from agriculture and forestry management. An integrated production of food, feed, chemicals, materials, fuels, energy, and goods is possible by the development of technologies for processing biogenic raw materials biomass, with those for the production of intermediate and final products (Kamm, 2013). These final products, termed as “bio-based products” include three categories: biofuels (biodiesel and bioethanol); bioenergy (heat and power); and bio-based chemicals (Figure 14.1). All are produced by “a biorefinery that integrates the biomass conversion processes” (WEF, 2010). Therefore, the biorefinery economy means a perspective in which the fossil fuels are replaced with biorenewables. This integration under the coverage of biorefineries registered notable accomplishments in research and development over the world, in several specific directions. First, Plant raw material Crop residues Sugar crops Woody & herbaceous crops FIGURE 14.1 Preprocessing the development of knowledge in the area of biomass resources in parallel with the improvement of systems sustainability which develop, harvest and process biomass resources, contributes in a high degree to biorefinery establishment as a new and efficient way in saving resources and improving the global ecological and carbon foot prints (Kamm, 2013). This would shift the actual promising state for the industrial biorefinery (with most second generation plants), toward largescale commercial production (WEF, 2010). Furthermore, the development of a multitude of products from biobased materials may be possible due to the increased efficiency and performance of process and technologies conversion and distribution of these materials and products (Kamm, 2013). In this context, biorefineries can be seen as “facilities that convert biomass-biological materials from living or recently living organisms-into fuels, energy, chemicals and materials” (WEF, 2010). The International Energy Agency (IEA) defines biorefineries in the IEA Bioenergy Task 42 on Biorefineries, as “the sustainable processing of biomass into a spectrum of marketable products and energy” (de Jong et al., 2009). In this context, the similarity to the spectrum of light in the image provided by the mentioned work is significant in terms of biorefineries tasks (Figure 14.2). In accordance with the concept of sustainability, a new and more general definition was proposed by the “Biorefinery Euroview” project. It defines biorefineries as “. integrated bio-based industries, using a variety of different technologies to produce chemicals, biofuels, food and feed ingredients, biomaterials (including fibers) and power from biomass raw materials” (Biorefinery Euroview, 2008). Final processing Feeds & foods Functional unit Products to replace petroleum based or petroleum dependent products Recycle or disposal Recycled within product system or to other product system Compost pile or landfill Integrated production of bio-based products from biogenic raw materials with specific technologies for biomass processing (Kim and Dale, 2004). Source: Reproduced with the permission of the author of PowerPoint presentation, Prof. Bruce E. Dale. (For color version of this figure, the reader is referred to the online version of this book.)
221 BIOMASS FEEDSTOCK Heat power fuels chemicals Biomass Biorefinery Food feed materials FIGURE 14.2 The spectrum of sustainable products and energy resulted through biomass processing (de Jong et al., 2009). Source: Reproduced with the permission of IEA Bioenergy Task Leader, Dr. Ed de Jong. (For color version of this figure, the reader is referred to the online version of this book.) Consequently, the biorefinery concept includes the management of all sustainability issues, concerning economic, environmental and societal components entailing a valuable strategy focused on the green economy era, being expected to play a significant role in supporting economical and social development (PP, 2012; Gavrilescu and Chisti, 2005). It should be noted that all these themes are in a tight connection with the development of regulations in the frame of bio-based products, and of market environment as well, and will contribute to overcome the differences in the status quo of industrial commercialization. These fundamental trends generated a very large interest in bio-based products, so as these products become parts of strategic plans of numerous stakeholders in various industries (WEF, 2010). Since biorefineries can generate multiple products, they offer the advantages of maximization the values derived from biomass feedstock (Xin et al., 2011; WEF, 2010). SHORT HISTORY OF BIOREFINERIES AND BIO-BASED PRODUCTS The industrial conversion of renewable resources has a quite long history, lasting since 6000 BC, in particular on the utilization of sugar cane (Demirbas, 2010; Kamm et al., 2006). However, proofs on the production of ethanol by distillation were found in China, in the form of dried residues of 9000 years old. Also, the ancient Egyptians used to produce alcohol by fermentations from vegetal materials (Demirbas, 2010). An analysis of biorefineries history should entail various aspects of wood saccharification, sugar production, synthesis of various bio-based products (furfural, lipids, lactic acid and many others), energy sources, and integrated processes (Kamm et al., 2006; Demirbas, 2010; de Jong and Marcotullio, 2010; Martin and Grossmann, 2012). Therefore, the topic branches of biorefineries which process renewable materials became well known and applied worldwide. These developments were more evident since the nineteenth and the beginning of the twentieth century, distinctively in the pulp and paper industry, where wood is the main raw material and the derived wastes gave rise to various solutions for the exploitation of valuable components they include (Rodsrud et al., 2012). Also, the food industry was a sector with high potential of waste valorization and recovery. Moreover, the increase in environmental concerns, especially related to the use of fossil fuels, has asked for sustainable solutions to limit the greenhouse gas effects and resources depletion. Table 14.1 provides a short outline of biorefinery evolution, based on data existent in various sources (Demirbas, 2010; Kamm et al., 2006; Rodsrud et al., 2012). However, some voices claim that the concept of “biorefinery” appeared in the 1990s as reaction to some trends of industry such as the need to use biomass resources in a more balanced way from both economic and environmental perspectives; an emergent concern in the promotion of low-quality lignocellulosic biomass to valuable products; an increased attention to the production of starch for energy applications; a need to develop extra high-value products and expand product combinations to face global competition; and to exploit an excess of biomass (especially in the pulp and paper industry) (Alakangas and Mäkinen, 2008; Berntsson et al., 2012). BIOMASS FEEDSTOCK Biorefineries process a bio-based feedstock input, analogous to the petroleum refineries, where a variety of different products may result, such as fuel, power, or chemicals (WEF, 2013). Although biorefineries use a large variety of different raw materials and conversion technologies, a clear alternative to fossil-based products does not exist still today (WEF, 2013). However, four classes of feedstocks are established (Demirbas, 2010): • First generation which entails edible biomass (starchrich, oily plants) to produce bioalcohols, vegetable oil, biodiesel, biosyngas, and biogas. • Second generation which uses biomass in the form of nonfood sources and crops (residual nonfood parts of crops, solid waste, wheat straw, etc.) to produce bioalcohols, biooil, biohydrogen, bio-Fischere Tropsch diesel. • Third generation which includes algae to produce vegetable oil and biodiesel. • Fourth generation which uses vegetable oil and biodiesel to produce biogasoline. A more detailed presentation is done in Table 14.2. The option to choose one or more of the four different
222 TABLE 14.1 14. BIOREFINERY SYSTEMS: AN OVERVIEW Short History of Biorefineries and Bio-Based Products Key Moment Place and Actors Innovations and Activities References 9000 BC China e Discovery of the art of distillation, which increases the concentration of alcohol in fermented solutions Demirbas, 2009 6000 BC Asia e Utilization of sugar cane Demirbas, 2010 Fifteenth Century American plantations e Export of sugar cane James et al., 1989 1748 Andreas Sigismund Margraff, German scientist e Key initiator of the modern sugar industry e Research on the isolation of crystalline sugar from different roots and beet Kamm et al., 2006, Burton and Cox, 1998 1780 Carl Wilhem Scheele e Discovery of lactic acid Benninga, 1990 1801 Cunern/Schlesien Poland e The first sugar refinery based on sugar bet F.C. Achard Paulik, 2011 Pennington and Baker, 1990 Early Nineteenth Century Samule Morey e First tested ethanol in internal combustion engine Lee and Lavore, 2013 1806 Napoleon Bonaparte e Economic continental blockade to limit overseas trade in cane sugar starch hydrolysis became of interest for the economy Brown, 2009 Harris, 1919 Paulik, 2011 1811 G.S.C. Kirchoff German pharmacist e Conversion of potato starch into “grape sugar“ e The starting point of starch industry Kamm et al., 2006, Paulik, 2011 van der Maarel et al., 2002 1812 Weimar, Germany J.W. Döbereiner e The first starch sugar plant was established Jentoft, 2003 Kamm et al., 2006 1819 H. Braconnot, French plant chemist e Treatment of wood with concentrated H2SO4 results in sugar (glucose) Binder and Raines, 2010 Jeffries and Lindblad, 2009; Paulik, 2011 1831 Döbereiner e First report on the production and separation of furfural by bran distillation with diluted acid de Jong and Marcotullio, 2010 Yang et al., 2011 1835 J.J. Berzelius, Swedish Professor e Development of enzymatic hydrolysis of starch to sugar (“catalysis”) Buchholz et al., 2005 Cheeptham and Lal, 2012 1839 A. Payen e Cellulose was obtained by wood treatment with nitric acid and subsequent treatment with a sodium hydroxide solution (“les cellules”) Kamm et al., 2006 Paulik, 2011 1840 G.J. Mulder e Synthesis of levulinic acid by heating fructose with hydrochloride Kamm et al., 2006 Paulik, 2011 1845 G. Fowners e Proposed the name of “furfurol” changed in “furfural” due to aldehyde function Kamm et al., 2006 1854 M.A.C Mellier e Disintegration of cellulose pulp from straw with caustic soda and steam Hofmann, 1873 Jeffries and Lindblad, 2009 Kamm et al., 2006 1855 G.F. Melsens e Wood conversion to sugar with dilute acid e Development of two approach on wood hydrolysis e Hydrolysis with concentrated acid at low temperature; hydrolysis with diluted acid at high temperature Kamm et al., 2006 Kupiainen, 2012 1863 B.C. Tilghman e The first patent for cellulose production by use of calcium bisulphite Gao et al., 2013
223 BIOMASS FEEDSTOCK TABLE 14.1 Short History of Biorefineries and Bio-Based Productsdcont’d Key Moment Place and Actors Innovations and Activities References 1866 B.C. Tilghman and brother (paper mill Harding and Sons) e Start of the first industrial experiment for the production of pulp from wood and hydrogen sulphite Antonsson, 2008; de Sa, 2004 1872 C.D. Ekman e Production of cellulose sulfate using magnesium sulfate as cracking agent Kamm et al., 2006 1874 W. Haarman F. Tiemann e Vanillin synthesis from cambial juice of coniferous wood Kamm et al., 2006 Paulik, 2011 1875 Company Haarman and Reimer e Coniferindthe first precursor for the production of vanillin was isolated, oxidized to glucovanillin and cleaved into glucose and vanillin e Industrial vanillin production e The first industrial utilization of lignin Kamm et al., 2006 Paulik, 2011 Wolfrom, 1970 1878 A. Mitscherlich e Improved the sulfite pulp process by fermentation of sugar from waste liquor to ethanol e Applied procedure to obtain paper glue from the waste liquor Kamm et al., 2006 Sindall, 1906 Watt, 1890 1895 A. Boehringer e Industrial lactic acid fermentation Benninga, 1990 The End of the Nineteenth Century e Ethanol was used in farm machinery and introduced in the automobile market Lee and Lavore, 2013 1900 e Development of pulp and paper mils (5200 worldwide) Kamm et al., 2006 Paulik, 2011 1901 A. Classen e The first commercial process of wood saccharification (German Patent 130980) with sulfuric acid Kamm et al., 2006 Hajny, 1981 1902 W. Normann e Liquid plant oils are converting into tempered fat by augmentation of hydrogen e Hydration of liquid catalytic (Ni), resulting tempered stearic acid Kamm et al., 2006 WEF, 2010 1909 M. Ewen G. Tomlinson e The first commercial process of wood working with dilute sulfuric acid (US Patent 938208) Kamm et al., 2006 Lloyd and Harris, 1955 Otulugbu, 2012 1893e1912 Company Boehringer-Ingelheim e The pioneer of industrial biotechnology Bio Deutschland, 2012 Interbelic Period Friedrich Bergius e Development technologically viable processes for wood saccharification e Ethanol production from the fermentation of wood sugar Kamm et al., 2006 Schobert, 2013 1920 Quaker Oats company e Development of furfural production from pentoses Marcus, 2005 RIRDC, 2006 1925 W.J. Hale, H. Dow, C.H. Herty e Chemurgy was founded in USA, having as an objective the utilization of agricultural resources in industry Kamm et al., 2006 1927 American Maraton Corporation e Development of commercial products from the organic solids in the spent sulfite liquor from pulp and paper manufacture as leather tanning agents and dispensing agents Kamm et al., 2006 WEF, 2010 1932 W.H. Carothers Van Natta e Discovery and developing a polyester made from lactic acid Huijser, 2009 Kobayashi, 2010 (Continued )
224 TABLE 14.1 14. BIOREFINERY SYSTEMS: AN OVERVIEW Short History of Biorefineries and Bio-Based Productsdcont’d Key Moment Place and Actors Innovations and Activities References 1934 Cedar rapids, Iowa e Furfural production was established as an industrial process Kamm et al., 2006 Peters, 1937 1940 A.E. Staley Dectur Illinois e Commercial production of levulinic acid in autoclaves e Utilization of hexoses from low cost cellulose production was experimented for the production of levulinic acid Kamm et al., 2006 Kitano et al., 1975 1941 Henry Ford e A car 100% biosynthetic composite material made from cellulose meal, soy meal, formaldehyde resin, with methanol as fuel produced from cannabis Kamm et al., 2006 1990s Company nature works e Commercialization of the poly(lactic) acid made from lactic acid Vink et al., 2003 alternatives to replace the fossil fuels-based products with biomass-based products depends on, among others, the costs involved (Sanders et al., 2005; van Ree and Annavelink, 2007). There are different paths for biomass utilization (Table 14.3) (Wagemann, 2012): • integral unmodified or modified biomass, without component separation; • various individual components of biomass; • biomass components in a complete way/form at various location; • the whole biomass in its complete forms. However, any classification is generic only based on a too large generalization and provides little information on the intimacy of involved processes as well as of the possibility to apply various technological processes to different feedstocks (Cherubini et al., 2009). No classification criterion allows the combination of different biorefinery systems by linking different technologies involved in both energy-driven biorefinery systems and material-driven biorefinery systems. Cherubini et al. (2009) mentioned some examples in this regard: “if the carbohydrate fraction of a lignocellulosic feedstock is used to produce cellulose and xylose, the system is classified as a lignocellulosic feedstock biorefinery; but can also be classified as a forest-based biorefinery and, if the lignin fraction is pyrolyzed, the same biorefinery is also suitable for classification as a two-platform concept biorefinery”. STRUCTURE OF BIOREFINERY CONCEPT The biorefinery is more than a fixed technology since it includes a collection of unitary processes, by several different routes from feedstocks to products (Xiu et al., 2011). Figure 14.3 shows the structural scheme of biorefinery concepts, including process types with the unitary processes and the primary products and intermediates, as well as secondary products (Hackl and Harvey, 2010). The economic viability of bio-based products preparation involves different processes and methods: physical, chemical, biological, and thermal. Table 14.3 describes shortly some of these processes and methods. However, a clear set of criteria to classify the different biorefinery concepts is still missing. van Ree and Annevelink (2007) considered a classification based on the following: • Raw material input, resulting in some classes of biorefineries, like Green, Whole Crop, Lignocellulosic, Feedstock, and Marine Biorefineries. • Technologies applied for biomass processing: Two Platform Concept, Thermo, Chemical Biorefineries. • Products resulted (main, intermediate): Syngas, Sugar, Lignin Platforms. Due to the complexity of this structure, process integration is the most sustainable approach to ensure the system efficiency and products quality. In an integrated configuration, biorefinery systems are structured in various ways by considering the use of raw materials, the environmentally sound character, and the degree of integration as follows (van Ree and Annevelink, 2007, Martin and Grossmann, 2012, Wagemann, 2012): • Lignocellulosic feedstock biorefinery is based on the processing of lignocellulosic-rich biomass sources in three steps (Figure 14.4): cellulose (sugar raw material); hemicelluloses (polyses); and lignin. These
TABLE 14.2 Feedstock Generation First Generation Feedstock Type and Characteristics Bio-Based Products Actual Situation Crops (sugar/starch rich) e Sugar cane: the most preferred, due to case of production but it is restricted to certain location (15e18% sugar) e Sugar beet (15e18% sugar) e Sweet sorghum, green grass (30% fibers, 20% proteins, aminoacids, 15% polysaccharides, 9% mono/disaccharides, 3% oligosaccharides) e corn, wheat, cassava: can be hydrolyzed enzymatically and resulted sugars are then fermented and processed in fuels and chemicals e Ethanol e Bio-based chemicals e Proteins for animals e Over 400 operational first generation refineries in the world e Risk of an excessive consumption of food crop competition between food and biorefinery e Risk of deforestation on long term e Excessive use of fertilizers and pesticides to improve production levels Demirbas, 2009; Kamm et al., 2006; McKillip et al., 2001; WEF, 2010; van Ree and Annevelink, 2007; Lee and Lavore, 2013; Sanders et al., 2005 Vegetable oil from oily plants e Palm (PPO) (40e50% oil) e Soybean (PPO) (11e22% oil) e Sunflower seed (PPO) (30e50% oil) (All are limited by the agricultural capacity of a country, and land-use change limitations e Jatropha (30e37% oil) e Waste vegetable oil e Biodiesel (glycerol as by-product) e The process itself relies on extracting the oils and converting them into biodiesel by breaking the bonds linking the long chain fatty acids to glycerol, replacing it with methanol (transesterification) e Pure plant oil (PPO) e Waste vegetable oil (WVO) e 40e50%yield WEF, 2010; Demirbas, 2009; Kamm et al., 2006; Lee and Lavore, 2013; Sanders et al., 2005 Lignocellulosic biomass (cellulose, hemicellulose, and lignin) e Inedible plants e Homogeneous biomass: wood chips and waste (16e24% lignin; 25e35% hemicellulose, 43e47% cellulose) e Quasihomogeneous biomass: straw (8e14% lignin, 24e30% hemicellulose, 31e41% cellulose) e Whole plants (25e31% lignin, 25e29% hemicellulose, 40e44% cellulose) e Switchgrass, miscanthus, short rotation poplar (exclude the direct land-use change) e Nonhomogeneous biomass, including low-value feedstock as municipal solid wastes e Biofuels (bioethanol) e Bio-based chemicals e Biomass to liquid (BTL) e Such biomass is generally more complex to convert and its production is dependent on new technologies significant breakthroughs in enzymatic conversion process e Significant and more abundant than first generation e Reduced dependence on food crops e Improvement of energy and environmental cycles e Improvement and costs of production e The conversion process is done according to two different approaches referred to as “thermo” or “bio” pathways Demirbas, 2009; Lee and Lavore, 2013; Sanders et al., 2005; Zhang et al., 2007; Speight, 2011 Category References STRUCTURE OF BIOREFINERY CONCEPT Second Generation Categories of Feedstock Utilized in Biorefineries for Various Bio-Based Products (Continued) 225
226 TABLE 14.2 Feedstock Generation Categories of Feedstock Utilized in Biorefineries for Various Bio-Based Productsdcont’d Feedstock Type and Characteristics Bio-Based Products Microalgae e Unicellular photo and heterotrophic organisms e Contain storage lipids (as triacylglycerols, transformed in biodiesel through transesterification) e Oil contents of some microalgae can exceed 80% of dry weight (20 times than traditional feedstock) e Safe biodegradable, not competing with arable land e Are highly productive, quick to cultivate, require CO2, sunlight and water for growing e Needs improvements in algal biology to achieve high growth rates, high liquid content, and ease extraction e Developments in photobioreactor engineering e Biofuels (ecologically friendly) Jatropha curcas tree e Contains 27e40% inedible oil, ehich can be converted to biodiesel e Positive effects on the environment and GHG emissions e Possible to be cultivated on wasteland or degraded ground. Lestari et al., 2010; Speight, 2011; WEF, 2010 Third Generation Agricultural, forestry, petrochemical and urban waste and residues e Agricultural waste and residues e Forest waste and residues e Municipal waste (paper cardboard, town cleaning) e Sludges (wet or dry biosolids) e Residual biomass from process industries e Could be gasification or combustion resulting in energy, heat, clan synthesis gas de Jong et al., 2009; Kamm et al., 2006; Star-COLIBRI, 2011; WEF, 2010 Fourth Generation Vegetable oil e Hydrolytic conversion e Deoxygenation Category References Lee and Lavore, 2013; Chisti, 2007; Chisti, 2008; Ribeiro et al., 2007; Speight, 2011 Al-Zuhair, 2007; Demirbas, 2009; Cherubini, 2010 14. BIOREFINERY SYSTEMS: AN OVERVIEW e Biogasoline e Bio jet fuel e Biodiesel Actual Situation
227 BIOREFINERY PLATFORMS TABLE 14.3 Biomass Utilization Paths (Wagemann, 2012) Biomass Utilization Examples Use of unmodified biomass without component separation for chemicals/materials or bioenergy Wood for wood-based raw materials or sawing products Wood used as fuel Insulating materials made of natural fibers Linseed oil as solvent Use of individual biomass components for chemicals/ materials and/or bioenergy Vegetable oil from rape or as component of lacquers/dyes Starch from cereal crops for the production of bioethanol or for the production of paper starch Sugar from sugar beet used as a fermentation raw material Complete utilization of the biomass components for chemicals/materials and/or bioenergy at various location Biogas from corn for local generation of electricity and heat respectively for biomethane as feed-in into grid for use in different locations Palm oil generation aboard, its transportation to Europe, and its domestic processing Complete, integrated utilization of the biomass components for chemicals/materials and/or bioenergy in one (networked) location for chemicals/materials and bioenergy Biorefinery concepts using a platform for the integrated production of a spectrum of products Source: Adapted with the permission of the coordinator of “Biorefineries Roadmap as part of the German Federal Government action plans for the material and energetic utilization of renewable raw materials” brochure on behalf of The Federal Government, Professor Kurt Wagemann. processing steps result in feeds, chemicals, biopolymers and other biomaterials. All residues are incinerated for the cogeneration of heat and power (van Ree and Annevelink, 2007). • Whole crop biorefinery uses raw materials (cereals, maize, and wheat) in the form of grain, flour (meal), and straw (combination of ears, leaves, chaff and nodes), based on dry or wet milling biomass. Their processing results in feeds, chemicals and biomaterials (Figure 14.5). • Green biorefineries use “nature wet” (fresh) biomass (green grass, clover, alfalfa, and immature cereals), resulting in a fiber-rich press cake and a nutrient-rich press juice (Figure 14.6). • Thermochemical biorefinery (TCBR) entails the biomass refining into a large portfolio of value-added products, by applying several technologies such as pyrolysis, gasification, torrefaction, and hydrothermal upgrading. The resulting products could be introduced into the existing infrastructures and substituting fossil fuels (de Wild, 2011; Martin and Grossmann, 2012). A particular concept derived from TCBR and developed by de Wild (2011) relies to Staged Catalytic Biorefinery Concept, which offers the possibility to process biomass in different sequential technological steps, with reducing the severity of the processing conditions using suitable catalysts, and to separate diverse products at different stages. • Marine biorefinery (MBR) is based on marine crops, i.e. microalgae (diatoms; green, golden, and blue/green algae) and macroalgae (brown, red and green seaweeds), and their derived products (Bowles, 2007; van Ree and Annevelink, 2007; Martin and Grossmann, 2012). Depending on the materials resulted after primary refinery steps, the leading procedures applied for further transformation and the integration degree of these above mentioned biorefinery systems could be included in various biorefinery platforms: biochemical, thermochemical, and microorganism platforms (Cherubini et al., 2009; Kammm et al., 2006; WEF, 2010) (Table 14.4.) In this context, the biorefinery is “an explicitly integrative, multifunctional overall concept that biomass as a diverse source of raw materials for the sustainable generation of a spectrum of different intermediates and products (chemicals, materials, bioenergy/biofuels), allowing the fullest possible use of all raw material components. The coproducts can also be food and/or feed. These objectives necessitate the integration of a range of different methods and technologies” (Wagemann, 2012). The integration and multifunctionality in biorefineries can be performed at four levels raw material, process, product, and industry (Martin and Grossmann, 2012; Wagemann, 2012) (Figure 14.7). BIOREFINERY PLATFORMS Biorefineries can produce chemicals and feels from biomass on several integrated platforms (Figure 14.8) (WEF, 2010). 1. The biochemical (sugar) platform, based on the biochemical conversion of biomass, focusing on sugar fermentation, and including steps dedicated to products separation and purification. 2. The thermochemical platform, based on the thermochemical conversion of biomass focusing on the gasification of carbonaceous materials and lignocellulosic biomass. 3. The microorganism platform, focusing on algae biomass cultivated in raceway type ponds or in photobioreactors.
228 14. BIOREFINERY SYSTEMS: AN OVERVIEW Primary process Primary products/ intermediates Secondary products Gasification Product gas Syngas, SNG,FT-fuels, MeOH, olefins etc. Pyrolysis Pyrolysis oil Fuel, biochemicals Torrefaction Torrefied biomass Biofuel with improved properties Combustion Steam/heat Heat and electricity Anaerobic digestion Biogas Biomethane, methanol, olefins Fermentation Ethanol Fuel, ethylene, ETBE, ethylamines Enzymatic hydrolysis Fermentable sugars Ethanol,fuel, ethylene, ethylamines Acid hydrolysis Cellulose, hemicelulose, lignin Fermentable sugars, biofuel, ethanol Supercritical conversion of biomass Cellulose, hemicelulose, lignin Fermentable sugars, biofuel, ethanol Solvent extraction Cellulose, hemicelulose, lignin, polysaccharides Ethanol, extractives, waxes Separation E.g. lignin Biofuel, heat, electricity, materials Drying and pelletising Biofuel Heat and electricity Extraction Vegetable oils, organic acids, extracts Fuels, oleochemicals Process type Thermochemical conversion Biological conversion Biomass Chemical conversion Mechanical conversion FIGURE 14.3 Schematic structure of biomass conversion processes and potential products (Hackl and Harvey, 2010). Source: Reproduced with the permission of authors, Dr. Roman Hackl and Prof. Simon Harvey, Department of Energy and Environment, Division of Heat and Power Technology Chalmers University of Technology. (For color version of this figure, the reader is referred to the online version of this book.) The biorefinery concept considered by the National Renewable Energy Laboratory is based on two different primary platforms integrating various routes included in the biorefinery structure (NREL, 2009): • The biochemical (sugar) platform performs the biomass breakdown into sugars based on chemical and biological processes: • If lignin is the result of pretreatment and enzymatic hydrolysis, two steps can be involved in its further transformation: e lignin upgrading, to etherified gasoline; e lignin pulping to high quality paper. • If aqueous sugars result after pretreatment and enzymatic hydrolysis, they are involved in fermentation processes, resulting in ethanol, butanol, and hydrogen. • The thermochemical platform is based on the biomass conversion onto synthesis gas through gasification, pyrolysis or hydrothermal conversion. • Gasification results in syngas, which can be further transformed in alkanes, methanol or hydrogen by FischereTropsch, catalysis, wateregas shift processes.
BIOREFINERY PLATFORMS Lignocellulosic feedstock biorefinery (LCFBR) Cellulose Hemicellulose Lignin Glucose polymer Pentoses, hexoses Phenol polymer Residues Cogeneration heat and power, extractions Fuels chemicals biomaterials FIGURE 14.4 Scheme of integrated biorefinery process of lignocellulosic feedstock biorefinery (LCFBR) type. Source: Adapted from Kamm et al. (2006) and Wagemann (2012). (For color version of this figure, the reader is referred to the online version of this book.) FIGURE 14.5 Scheme of integrated biorefinery process of whole crop biorefinery (WCBR) type. Source: Adapted from Kamm et al. (2006) and Wagemann (2012). (For color version of this figure, the reader is referred to the online version of this book.) FIGURE 14.6 Scheme of integrated biorefinery process of green biorefineries type. Source: Adapted from Kamm et al. (2006) and Wagemann (2012). (For color version of this figure, the reader is referred to the online version of this book.) 229 • Pyrolysis and hydrothermal conversion result in biooil, which is further transformed during the following processes: e upgrading, when liquid fuel results; e catalytic reforming, resulting in hydrogen; e extraction, when various chemicals are obtained; e cross-linking resulting in various (bio) materials. The third platformdmicroorganism platformdhas been included in the biorefineries structure by the National Renewable Energy Laboratory (WEF, 2010). This structure demonstrates that various processes can occur in a complex biorefinery, similar to a conventional oil refinery. This similarity was also graphically demonstrated by Kamm et al. (2006) (Figure 1.3). There are also some unclassified biorefineries, which include (de Jong and van Ree, 2006) side and waste streams, MBR, most generation III biorefineries, and consortia of different industries. They are expected to play a significant role in the future, since the classic concept of biorefinery is tightly linked with the progress of agriculture, the efficiency and availability of food and feed production, with major consequences for the prime arable land (PP, 2012). Considering these problems, it is essential to promote integrated biorefinery models, which would be able to surpass the challenges addressing retaining and recycling of phosphorous, finding new sources of soil organic carbon, maintaining biodiversity by adequate measures (PP, 2012; Star-COLIBRI, 2011). Besides, a new and challenging development began to be focused on the integrated valorization of organic waste streams, such as agrofood by-products, effluents, resulting in new value-added chemicals, biofuels, biomaterial, and water (PP, 2012; Liu et al., 2010; Visvanathan, 2010; Laufenberg et al., 2003). This way, the integration of biorefinery platforms would be able to generate the synergism, as the underlying concept of industrial ecology. By closing material cycles and cascade utilization and recycling, it would be ensured a multilevel, explicitly integrative, multifunctional incorporation of raw materials, processes, and products, belonging to various industrial systems, simultaneously with preventing resource loss by source reduction and waste minimization along the entire biorefinery value chain. A full overview of the platforms, products, feedstocks and conversion processes is given in Figure 14.9 (de Jong and Marcotullio, 2010). Moreover, the eco-efficiency would become the leading concept governing the full system, since processes for biomass treatment and conversion should be resource efficient in terms of materials and energy use and long lifetime of goods and products, along with consumption of auxiliaries, and should avoid adverse
230 14. BIOREFINERY SYSTEMS: AN OVERVIEW TABLE 14.4 Conversion Processes, Methods and Techniques Employed by Biorefineries to Transform the Raw Biomass into Commercial Products Conversion Process Description Products References Fermentation of Sugar/Starch Crops • Raw materials Bioethanol e Starch crops Butyric acid e Lignocellulosic materials Butanol • Pretreatment e Starch hydrolyzed enzymatically to deliver sugar solutions (b-glucosidase, endocellulase, and exocellulase) • Treatment e Microbial fermentation to produce bioethanol (S. cerevisiae, P. skittles, Z. mobiles, E. coli) or butanol (C. tyrobutyricum, C. acetobutylicum) Chen, 2011; Hodge et al., 2008; Ramey, 1998; Sharara et al., 2012; WEF, 2010 Fermentation of Lignocellulosic Biomass • Raw materials e Lignocellulosic biomass • Pretreatment e Mechanical e Chemical or thermal treatment to separate the cellulosic and hemicellulosic material from the nonfermentable lignin • Treatment e Enzymatic hydrolysis of cellulosic and hemicellulosic components e Fermentation of sugars Cherubini, 2010; FitzPatrick et al., 2010; Knauf and Moniruzzaman, 2004; WEF, 2010 Transesterification of Triglycerides • Raw materials e Plant or algal oil • Treatment e Triglycerides are treated with methanol in the presence of a dedicated catalyst to deliver fatty acids and methyl esters Biodiesel Asakuma et al., 2009; Chisti, 2007; Meher et al., 2006 Gasification to Syngas e Breakdown of carbonaceous materials into H2 and CO (syngas) through thermal decomposition in the presence of a limited quantity of oxygen e H2 þ CO mixture can be further converted by partial oxidation at elevated temperature as FischereTropsch reactions e Physical and chemical properties of feedstocks affects the quality of syngas and process efficiency (moisture, ash minerals) e The use of catalysts can improve the thermochemical conversion by facilitating a preferred reaction mechanism e Syngas (H2 þ CO) can be used to e Direct combustion in boilers, turbines or internal combustion engines e Produce hydrogen by separation e Produce chemical products (ammoma) by chemical synthesis e Produce liquid fuels through the FischereTropsch process Devi et al., 2003; Göransson et al., 2011; Huang and Ramaswamy, 2013; Hughes and Larson, 1998; Pryadarsan et al., 2004; Van der Drift et al., 2001; WEF, 2010 Fast Pyrolysis e Thermal decomposition of biomass in a biooil (in the absence of oxygen), which can be further converted through hydrogenation or gasification into certain hydrocarbon e Reduced costs, compared to gasification of solid biomass About 100 chemicals species biooil (biocrude) with 44e47% carbon, 6e7% hydrogen, and 46e48% oxygen) WEF, 2010; Sharara et al.; Evans and Milne, 1987; Bridgwater et al., 1999
231 BIOREFINERY ECO-EFFICIENCY TABLE 14.4 Conversion Processes, Methods and Techniques Employed by Biorefineries to Transform the Raw Biomass into Commercial Productsdcont’d Conversion Process Description Products References FischereTropsch Synthesis e Catalytic conversion of sugars into liquid hydrocarbons (C1eC50) e The process in selective depending on temperature, pressure and catalysts Synthetic fuel Demirbas, 2010; Lappas and Heracleous, 2011; NREL, 2009 Hydrogenation e Hydrotreatment of biooils, resulting hydrotreated renewable jet fuels (HRJ) e Removes oxygen and others impurities from organic oils (extracted directly from feedstocks with high oil content or produced by pyrolysis) HRJdhydrotreated renewable jet fuels, with similar properties as kerosene Conversion of Syngas to Methane (SNG) e Thermal gasification and particular FischereTropsch reaction SNGdsynthetic natural gas (a good substitute of the natural gas) Martin and Grossmann, 2012 Aerobic Digestion e Conversion of biodegradable waste or energy crops into a gaseous fuel biogas e Conversion efficiency is about 70% Biogas (50% methane) Martin and Grossmann, 2012 Catalytic Thermochemical Conversion e Increases the yield and optimize the composition of output products of thermochemical conversion e Helps in overcoming the problematic qualities of biooil (thermal and temporal instability) e Catalyst can be incorporated during or after the production process, or in both stages (activated alumina, silicate, Y-zeolite, ZSM-5) Pyrolysis oil (biooil) which is a chemical intermediate or directly as liquid fuel Carlson et al., 2009; de Wild, 2011; Sharara et al., 2012; Zhang et al., 2009 environmental impacts and risks (Gavrilescu, 2011; Wagemann, 2012). Therefore, we can discuss about a multilevel integration of the biorefinery concept along the supplying chain: raw materialseprocesseseproductse industrial platforms (Figure 14.10). BIOREFINERY ECO-EFFICIENCY Biorefineries offer numerous business opportunities (WEF, 2010). It is expected that the global market of biofuels will increase from almost $80,000 million in 2011 to $185,000 million in 2021. Moreover, it is estimated that almost 136,000 million L of biofuels will be consumed in 2020 in USA, which would require over 500 commercial-scale cellulosic ethanol refineries, with capital requirements of $168,000 million (Solecki et al., 2012). Other estimations showed that the actual revenue potentials due to biomass conversion are of $80,000 million for biofuels, $10,000e15,000 million for biobased chemicals, including bioplastics (WEF, 2010). There are important revenue potentials along the entire biomass value chain, associated with the most relevant steps, as follows (WEF, 2010): agricultural inputs, $15,000 million; biomass production, $89,000 million; biomass trading, $30,000 million; biorefinering inputs, $10,000 million; biorefinering chemicals and downstream chemistry, $6,000 million; biorefinering fuels, $80,000 million; biomass power and heat, $65,000 million. Today a major challenge is associated to the production of biofuels and bio-based chemicals in an ecoefficient manner. Since the production, the products
232 14. BIOREFINERY SYSTEMS: AN OVERVIEW FIGURE 14.7 Levels of integration and multifunctionality already realized in biorefineries. Source: Adapted upon Wagemann (2012). Adapted with the permission of the coordinator of “Biorefineries Roadmap as part of the German Federal Government action plans for the material and energetic utilization of renewable raw materials” brochure on behalf of The Federal Government, Professor Kurt Wagemann. (For color version of this figure, the reader is referred to the online version of this book.) Biorefinery platforms Biochemical platform Crops (sugar) starch rich Biobased products Biomaterials Vegetable oily plants Thermochemical platform Feedstock Lignocellulosic biomass Biomass Fuels and energy Microorganism platform Microalgae Oily trees Combined heat and power Biochemicals • Composite materials • Dyes and pigments • Detergents and cleaners • Adhesives • Oils and inks • Etc. • Lignin, biogas, cake • Ethanol, methanol, fuel oil • Syngas, methane, • Hydrogen • Agricultural chemicals • Activated carbon • Specialty chemicals • Industrial surfactants • Fatty acids • Acetic acid • Etc FIGURE 14.8 Integration of three biorefinery-integrated platforms. Source: Adapted from WEF (2010) and Kamm et al. (2006). (For color version of this figure, the reader is referred to the online version of this book.) and by-products are quite numerous and diverse; a simple approach for the estimation of production economics would be always opportune, so as to offer valuable information about the relative feasibility of various production alternatives and routes. For example, Melin and Hurne (2011) developed an algorithm to find “the production route with the minimum production costs for a biofuel or a chemical, for each raw material, when the process and the economic parameters occur in a known range”. Other several studies have estimated biofuel production costs from corn stove through gasification and FischereTropsch routes (Demirbas, 2010;
233 BIOREFINERY ECO-EFFICIENCY Organic residues and others Starch crops Grasses Fractionation and/or pressing Separation Grain Sugar crops Lignocellulosic crops Lignocellulosic residues Oil based residues Oil crops Straw Straw Pretreatment Pressing Lignin Fiber Separation Gasification Organic solution Oil Pyrolysis, HTU Hydrolysis Syngas Extraction Anaerobic digestion Pyrolytic liquid C5 sugars C6 sugars Separation Water gas shift Hydrogenation Biogas Fermentation Methanisation Chemical reaction Chemical reaction Upgrading Combustion Steam reforming Water electrolysis H2 Chemical reaction Legend Feedstock Platform Material products Chemical process Mechanical/ Physical process Estherification Thermochemical process Biochemical processes Bio-methane Energy products Link among biorefinery pathways Bio-H2 Fertilizer Organic acids & extracts Biomaterials Synthetic liquid biofuels (FT) Bioethanol Chemicals & polymers Glycerine Food Electricity and heat Animal feed Biodiesel FIGURE 14.9 Full network of the platforms, products, feedstocks and conversion processes (de Jong et al., 2009). Source: Reproduced with the permission of IEA Bioenergy Task Leader, Dr. Ed de Jong. (For color version of this figure, the reader is referred to the online version of this book.) Batsi et al., 2012; Swanson et al., 2010). The objective was to compare capital investment costs and production costs for various biorefinery scenarios. Building a bio-based economy must be able not only to solve the current economic difficulties but also to generate an economic system with minimal impact to the environment. Even though regarded as similar to petroleum refinery, a comparison of the biorefinery and petrochemical value chains show some similarities but also a large number of differences. Both result in complex product trees, but one of the most relevant differences consists in compositions of fossil raw materials and biogenic raw materials (Kamm et al., 2006; Wagemann, 2012). Table 14.5 illustrates some similarities and differences between two value chains. Consequently, for decision-making process, it is necessary to develop a methodology to drive decisions on biorefinery, with a focus on product design and process. Transition to a biorefinery economy could involve significant investments in infrastructure to produce, store and sell biorefinery products to customers (Demirbas, 2010). A number of questions related to biorefinery diagnosis can be addressed using SWOT analysis. Such an investigation of the opportunities and strengths, weaknesses and threats of biorefineries as developed by IEA within the Task 42 is illustrated in Table 14.6 (de Jong et al., 2009). The concept of eco-efficiencyddefined as “creating more value with less impact”dhas been developed by The World Business Council to weigh and compare products and technologies in both aspects: environmental pressure and economic significance (WBCSD, 2000). The Organization for Economic Co-operation and Development (OECD) has defined eco-efficiency as the effectiveness with which ecological resources are used to meet human needs. Integrating the issues concerning the environmental impacts and economic value resulting from biorefinery processes allows decision makers in the business world to evaluate and compare products and technologies simultaneously, from both points of view. Organizations could be supported to establish measurable objectives of eco-efficiency and to facilitate comparisons between companies and business sectors by the standardization of definitions and decision system for calculating and reporting eco-efficiency indicators. The environmental impact ratio, defined in Figure 14.11,
234 14. BIOREFINERY SYSTEMS: AN OVERVIEW reflects how much environmental impact per environmental credit occurs in the product system (Hong Chua and Replace with Steinmüller, 2010; Kim and Dale, 2004). A scenario with a greater eco-efficiency would be more sustainable, which means that it would offer more economic value per unit of environmental impact (Fig. 14.11). Some eco-efficiency indicators were developed for different levels of biorefinery integration, following the physical flows of materials and energy (Hong Chua and Steinmüller, 2010): Level 1 addresses process integration and involves the key processes (receiving and preparation of feedstock, retreatment, conversion to bioproduct, and wastewater treatment system). Level 2 refers to agriculture integration, which means that feedstocks, including agricultural waste, are supplied in the biorefinery system at the business level that is involving low costs, while biofuels, bioelectricity and biochemicals from biorefinery are sent to the agricultural sector. Level 3 involves livestock farming integration at the business level, meaning that the organic waste from farms are supplied to the biorefinery system, while animal feed products are sent to the farm. FIGURE 14.10 The biorefinery process chain (Wagemann, 2012). Source: Reproduced with the permission of the coordinator of “Biorefineries Roadmap as part of the German Federal Government action plans for the material and energetic utilization of renewable raw materials” brochure on behalf of The Federal Government, Professor Kurt Wagemann. (For color version of this figure, the reader is referred to the online version of this book.) TABLE 14.5 Estimated costs of production in biorefinery systems may be hampered by a number of driving forces who can change their direction of action and/or importance in time (agricultural development, raw material costs, production scale, competing markets evolution, their demands and access, waste recovery Comparison of Biorefinery and Petrochemical Value Chains (Wagemann, 2012) Value Chain Biorefinery Petrochemical Raw Materials Biomasses: very complex mixture of organic compounds Mineral oil, natural gas: mixture of hydrocarbons Carbon and heteroatoms (poor in hydrogen, rich in oxygen) Carbon and hydrogen (almost no hetero atoms, poor in oxygen) Contains inorganic compounds Contains virtually no inorganic compounds Hydrous Waterless Primary Refinery Thermal and thermocatalytic (syngas) as well as biochemical (biogas) cleavage into simple molecules Distillation and thermal and thermocatalytic cleavage into simple molecules Secondary Refinery Build-up complex molecules from simple precursors (bottom-up principal) Processes Thermochemical, thermocatalytic and chemocatalytic processes Product Classes Chemicals and materials Combustibles and fuels Source: Adapted with the permission of the coordinator of “Biorefineries Roadmap as part of the German Federal Government action plans for the material and energetic utilization of renewable raw materials” brochure on behalf of The Federal Government, Professor Kurt Wagemann.
BIOREFINERY ECO-EFFICIENCY TABLE 14.6 235 SWOT* Analysis of Biorefineries Processes (de Jong et al., 2009) Strengths Weaknesses • Adds value to the sustainable use of biomass • Maximizes biomass conversion efficiencydminimizing raw material requirements • Produces a spectrum of bio-based products (food, feed, materials, and chemicals) and bioenergy (fuels, power and/or heat) feeding the full bio-based economy • Strong knowledge of infrastructure available to tackle any nontechnical and technical issues potentially hindering the deployment trajectory • Is not new, and in some market sectors (food, paper, etc.), it is common practice • Broad undefined and unclassified area • Needs involvement of stakeholders from different market sectors (agro, energy, chemical,.) over the full biomass value chain • Most promising biorefinery processes/concepts not clear • Most promising biomass value chains, including current/future market volumes/prices, not clear • Still at a stage of studying and concept development instead of real market implementation • Variability of quality and energy density of biomass Opportunities Threats • Make a significant contribution to sustainable development • Challenging national, European and global policy goalsdinternational focus on sustainable use of biomass for the production of bioenergy • Biomass availability is limited so the raw material should be used as efficiently as possibledi.e. development of multipurpose biorefineries in a framework of scarce raw materials and energy • International development of a portfolio of biorefinery concepts, including designing technical processes • Strengthening of the economic position of various market sectors (e.g. agriculture, forestry, chemical and energy) • Biorefinery is seen as hype that still has to prove its benefits in the real market • Economic change and drop in fossil fuel prices • Fast implementation of other renewable energy technologies filling market needs • No level playing field concerning bio-based products and bioenergy (assessed to a higher standard) • Global, national and regional availability and contractibility of raw materials (e.g. climate change, policies, and logistics) • High-investment capital for pilot and demonstration initiatives difficult to find, and existing industrial infrastructure is not depreciated yet • Fluctuating (long-term) governmental policies • Questioning of food/feed/fuels (land use competition) and sustainability of biomass production • Goals of end users often focused upon single product * Strengths, Weaknesses, Opportunities, and Threats Source: Reproduced with the permission of IEA Bioenergy Task Leader, Dr. Ed de Jong. and recycling alternatives, storage and production costs, distribution costs, etc.), which could be associated with the components of a complex system with various boundaries (Figure 14.12; Demirbas, 2010; Kim and Dale, 2004). Life cycle assessment (LCA) is an especially useful tool to investigate the environmental performance of l nta me n viro act En Imp Eco-efficiency = AD PO P, GW CP , H P, O AP TP, E DP, ,E P TP, Economic value added Capital investment Environmental impact ratio NPV, IRR Economic value added = ic om on is Ec alys An Market value of products Cost of raw material & fuel Environmental impact ratio = product and/or technologies. The problem to be solved in the case of biorefineries is not a simple one because these systems are characterized by some particularities that need to be considered in evaluating the processes on an LCA basis and to ensure correct results in terms of eco-efficiency (for example, sometimes it is not obvious which product should be the main output; Environmental impact Environmental credit FIGURE 14.11 Integration of economic analysis and environmental impact for eco-efficiency (ADP, abiotic resources depletion potential; GWP, global warming potential; ODP, ozone layer depletion potential; POCP, photochemical oxidation potential; HTP, human toxicity potential; ETP, ecological toxicity potential; AP, acidification potential; EP, eutrophication potential; NPV, net present value; IRR, internal rate of return). (For color version of this figure, the reader is referred to the online version of this book.)
236 14. BIOREFINERY SYSTEMS: AN OVERVIEW Surplus Local power grid •Electricity •Chemicals, enzyme •Electricity •Steam System boundary Pre-processing Cropping systems Animal waste treatment Animal operation Final processing Intermediates Final product •Sugars •Ethanol •Lipids •Biodiesel •Lignin •Nutrient Energy inputs Biorefinery •Steam •Electricity •Steam Inputs •Fertilizers •Fuel •Agrochemicals Inputs •Biopolymers •Ash •Protein Final product •Food •Other products •Chemicals Cogeneration •Ash FIGURE 14.12 Boundaries of an integrated biorefinery system (Kim and Dale, 2004). Source: Reproduced with the permission of the author of PowerPoint presentation, Prof. Bruce E. Dale. (For color version of this figure, the reader is referred to the online version of this book.) Hong Chua and Steinmüller, 2010; Laser et al., 2009). Further, the system boundaries could be different if the biorefineries are nonintegrated or integrated and this can determine the selection of system boundaries, which could also affect the eco-efficiency results, while allocation issues in particular are both important and somewhat controversial (Figure 14.12). A very common approach considers that all biomass is local since this could improve the selection of crops and cropping systems for local biorefineries, reduce opportunities for agendadriven manipulation of data and opportunities for system integration and waste utilization could be better exploited (Kim and Dale, 2004). The functional unit could be chosen as unit area of land allocated for crop biomass for a certain time period since cropping systems play an important role in the environmental performance of bio-based products, while impacts assessment could address global warming potential, nonrenewable energy, crude oil consumption, water use, acidification, eutrophication, biodegradability, less toxicity, etc. (Laser et al., 2009; Demirbas, 2010). In their study, Hong Chua and Steinmüller (2010) have identified the following main environmental influences for a biorefinery: energy consumption, material consumption, GHG emissions, acidification, and eutrophication. The eco-efficiency indicators used to account for these environmental influences are as shown in Table 14.7. Ensuring biorefinery eco-efficiency is one of the most relevant objectives of Task 42 of IEA in parallel with the projection of new perspectives in terms of competitiveness, sustainability, and safety of processing routes for biogenic raw materials to guarantee the concurrent fabrication of biofuels, commodity chemicals, new materials, heat and power. CONCLUDING REMARKS AND PERSPECTIVES Bioresource use in the forms of new and waste biomass is a great opportunity and a challenge for the future since it offers the chance of replacing fossil fuels for the production of energy carriers, materials and specialty chemicals and diminishing the market pressure in an almost carbon-neutral way. Industrial biorefineries are seen as one of the most promising directions toward a sustainable bio-based economy. Fully developed biorefineries combine biological and physicochemical processes. A weakness of biorefineries as an alternative to conventional oil refineries consists in the fact that the former is based on biofeedstock, which can require an intensive cultivation and land use. Moreover, biorefineries could compete with food requirements and needs, which would limit the land allocated to biomass for biorefineries. As a result, the future of biorefineries should consider the use of nonedible biomass and the advanced processing of biomass waste, as well as land which could not normally be used for agriculture. This type of land could be used for microalgae cultures or renewable plants. Other sources of raw material for biorefineries could be found on waste from the food industry and urban organic waste. The processing of this raw matter can be successfully and eco-efficiently carried out through the development of enzymatic systems and engineered microorganisms capable of separating useful compounds from waste. The development of these technologies should also consider the important issue of costs, since, currently, oil-based refineries offer more cost-effective solutions at the expense of environmental degradation and pollution.
TABLE 14.7 Main Eco-Efficiency Indicators for Biorefineries (Hong Chua and Steinmüller, 2010) Eco-Efficiency Indicators (EEI) Equation Terms Overall P EETEC;i ¼ PRi = Biorefinery Material Consumption (EETMC,i) EETMC;i ¼ PRi = Biorefinery GHG Emissions (EEGHG,i) EEGHG;i ¼ PRi = Biorefinery Acidification Emissions (EEAP,i) EEAP;i ¼ PRi = Biorefinery Eutrophication Emissions (EEEP,i) EEEP;i ¼ PRi = TECi ¼ PRi =ðRECI þ RECII þ RECIII þ NRECI þ NRECII þ NRECIII Þ P P P P TMCi ¼ PRi =ðRMCI þ RMCII þ RMCIII þ NRMCI þ NRMCII þ NRMCIII Þ GHGi ¼ PRi =ðGHGI þ GHGII þ GHGIII Þ APi ¼ PRi =ðAPI þ APII þ APIII Þ EPi ¼ PRi =ðEPI þ EPII þ EPIII Þ PRi, total profit from all productions sold (country currency) NRECi, total nonrenewable energy consumption of biorefinery RECi, total renewable energy consumption of biorefinery (Megajoules) RMCi, total renewable material consumption of biorefinery (kg) NRMCi , total nonrenewable material consumption of biorefinery (kg) GHGi, total greenhouse gas emissions of biorefinery (kg) APi, total acidification emissions of biorefinery (kg) EPi, total eutrophication emissions of biorefinery (kg) Energy Consumption Total Energy Consumption (EETEC;ij) EETEC;ij ¼ PRij = Nonrenewable Energy Consumption (EENRE;ij) EENRE;ij ¼ PRij = P TECij ¼ PRij =ðRECI þ RECII þ RECIII þ NRECI þ NRECII þ NRECIII Þj P NRECij ¼ PRij =ðNRECI þ NRECII þ NRECIII Þj Renewable Energy Consumption EEREC;ij ¼ EETEC ð1=EETEC  1=EENRE Þ  100% Rate (EEREC;ij) PRij, allocated profit from productions sold (country currency) NRECij, allocated nonrenewable energy consumption associated with the production of bioproduct from feedstock and biorefinery integration levels (Megajoules) RECij, allocated renewable energy consumption associated with the production of bioproduct and biorefinery integration levels (Megajoules) CONCLUDING REMARKS AND PERSPECTIVES Biorefinery Energy Consumption (EETEC,i) Material Consumption Total Material Consumption (EETMC;ij) EETMC;ij ¼ PRij = EENRM;ij ¼ PRij = P P TMCij ¼ PRij =ðRMCI þ RMCII þ RMCIII þ NRMCI þ NRMCII þ NRMCIII Þj NRMCij ¼ PRij =ðNRMCI þ NRMCII þ NRMCIII Þj PRij, allocated profit from productions sold (country currency) RMCij, allocated renewable materials consumption associated with the production (Continued) 237
238 TABLE 14.7 Main Eco-Efficiency Indicators for Biorefineries (Hong Chua and Steinmüller, 2010)dcont’d Eco-Efficiency Indicators (EEI) Equation Terms EERMC;ij ¼ EETMC ð1=EETMC  1=EENRM Þ  100% of bioproduct and biorefinery integration levels (kg) NRMCij, allocated nonrenewable materials consumption associated with the production of bioproduct and biorefinery integration levels (kg) Nonrenewable Material Consumption (EENRM;ij) Renewable Material Consumption Rate (EERMC;ij) Greenhouse Gases (GHG) EEGHG;ij ¼ PRij = P GHGij ¼ PRij =ðGHGI þ GHGII þ GHGIII Þj PRij, allocated profit from productions sold (country currency) GHGij, allocated greenhouse gas emissions associated with the production of bioproduct and biorefinery integration levels (kg) Acidification Potential (AP) Acidification Emissions (EEAP) EEAP;ij ¼ PRij = P APij ¼ PRij =ðAPI þ APII þ APIII Þj PRij, allocated profit from productions sold (country currency) APij, allocated acidification emissions associated with the production of bioproduct and biorefinery integration levels (kg SO2 equivalent) EPij ¼ PRij =ðEPI þ EPII þ EPIII Þj PRij, allocated profit from productions sold (country currency) EPij, allocated eutrophication emissions associated with the production of bioproduct and biorefinery integration levels (kg PO4 equivalent) Eutrophication Potential (EP) Eutrophication Emissions (EEEP) EEEP;ij ¼ PRij = P Notations: i refers to the level of integrations of the biorefinery; j refers to the product from the refinery. Source: Development Of Eco-Efficiency Indicators for a Biorefinery, Authors: Celia Bee Hong Chua, Horst Steinmüller (http://www.energyefficiency.at/web/artikel/eco-efficiency_indicators.html). 14. BIOREFINERY SYSTEMS: AN OVERVIEW GHG Emissions (EEGHG)
REFERENCES Acknowledgments This work was partially supported by the grant of the Romanian National Authority for Scientific Research, CNCSdUEFISCDI, project number PN-II-ID-PCE-2011-3-0559, Contract 265/2011. References Al-Zuhair, S., 2007. Production of Biodiesel: possibilities and challenges. Biofuels, Bioprod. Biorefin. 1, 57e66. Alakangas, E., Mäkinen, T., 2008. BioRefine Programme 2007e2012. Yearbook 2008, Tekes Review 293/2008. VTT Technical Centre of Finland, Tekes, Finnish Funding Agency for Technology and Innovation, Helsinki. Antonsson, S., 2008. Strategies for Improving Kraftliner Pulp Properties (Ph.D. thesis). Royal Institute of Technology, Stockholm. Asakuma, Y., Maeda, K., Kuramochi, H., Fukui, K., 2009. Theoretical study of the transesterification of triglycerides to biodiesel fuel. Fuel 88, 786e791. Batsi, D.R., Solvason, C.C., Sammons, N.E., Chambost, V., Bilhartz, D.L., Eden, M.R., El-Halwagi, M.M., Stuart, P.R., 2012. Product Portofolio Selection and Process Design for the Forest Biorefinery. In: Stuart, P.R., El-Halwagi, M.M. (Eds.), Integrated Biorefineries: Design, Analysis and Optimization. CRC Press, Boca Raton, pp. 4e36. Benninga, H., 1990. A History of Lactic Acid Making: A Chapter in the History of Biotechnology. Springer. Berntsson, T., Sanden, B., Olsson, L., Asblad, A., 2012. What is a biorefinery? In: Sanden, B. (Ed.), Systems Perspectives on Biorefineries. Chalmens University of Technology, Gäteborg, pp. 16e25. Binder, B., Raines, R.T., 2010. Fermentable sugars by chemical hydrolysis of biomass. Proc. Natl. Acad. Sci. USA 107, 4516e4521. Bio Deutschland, 2012. White biotechnology a pillar of the bioeconomy. Available from: http://www.biodeutschland.org/tl_files/content_ eng/Documents/2013/Flyer_White_Biotechnology.pdf. Biorefinery Euroview, 2008. Current situation and potential of biorefinery concept in the EEC, strategic framework and guideline for its development. FP6EEC Project. Available from: www. biorefinery.euroview.eu. Bowles, D. (Ed.), 2007. Micro and macro algae: Utility for industrial applications. CPL Press, Newbury, University of York, UK. Bridgwater, A., Meier, D., Radlein, D., 1999. An overview of fast pyrolysis of biomass. Org. Geochem. 30, 1479e1493. Brown, H., 2009. Starch Reality. Food Manufacture, UK. Available from: http://www.foodmanufacture.co.uk/Ingredients/Starchreality2. Buchholz, K., Kasche, V., Bornscheuer, U.T., 2005. Biocatalysts and Enzyme Technology. Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. Burton, R.A., Cox, P.A., 1998. Sugarbeet culture and mormon economic development in the Intermountain West. Econ. Bot. 52, 201e206. Carlson, T., Tompsett, G., Conner, W., Huber, G., 2009. Aromatic production from catalytic fast pyrolysis of biomass derived feedstocks. Top. Catal. 52, 241e252. Centi, G., Perathoner, S., 2009. From green to sustainable industrial chemistry. In: Cavani, F., Centi, S., Perathoner, S., Trifiro, F. (Eds.), Sustainable Industrial Chemistry. John Wiley and Sons, New York, pp. 11e20. Cheeptham, N., Lal, A., 2012. Starch Agar Protocol. American Society for Microbiology. Available from: http://www.microbelibrary. org/library/laboratory-test/3780-starch-agar-protocol. 239 Chen, Y., 2011. Development and application of co-culture for ethanol production by co-fermentation of glucose and xylose: a systematic review. J. Ind. Microbiol. Biotechnol. 38, 581e597. Cherubini, F., 2010. The biorefinery concept using biomass instead of oil fro producing chemicals. Energy Convers. Manage. 51, 1412e1421. Cherubini, F., Jungmeier, G., Wellisch, M., Willke, T., Skiadas, I., Van Ree, R., de Jong, E., 2009. Toward a common classification approach for biorefinery systems. Biofuels, Bioprod. Biorefin. 3, 534e546. Chisti, Y., 2007. Biodiesel from microalgae. Biotechnol. Adv. 25, 294e306. Chisti, Y., 2008. Biodiesel from microalgae beats bioethanol. Trends Biotechnol. 26, 126e131. de Jong, W., Marcotullio, G.L., 2010. Overview of biorefineries based on coproduction of furfural, existing concepts and novel developments. Int. J. Chem. React. Eng. 8. Article A69. de Jong, E., van Ree, R., 2006. Biorefineries- International Status Quo and Future Direction. Wagenigen University. Available from: www.biorefinery.nl/fileadmin/biorefinery/docs/publications/ presentations-kickoff/-2-current-status-on-biorefineries-IEA42150307.pdf. de Jong, E., van Ree, R., Kwant, I.K., 2009. Biorefineries: Adding Value to the Sustainable Utilization of Biomass. IEA Bioenergy. Available from: http://www.ieabioenergy.com/Libitem.aspx? id¼6420. de Sa, D.O.J., 2004. Applied Technology and Instrumentation for Process Control. Taylor & Francis, New York, London. de Wild, R.J., 2011. Biomass Pyrolysis for Chemicals (Ph.D. thesis). Rijksu University, Groningen, The Netherlands. Demirbas, M.F., 2009. Biorefineries for biofuel upgrading: a critical review. Appl. Energy 86, S151eS161. Demirbas, A., 2010. Biorefineries: for Biomass Upgrading Facilities. Springer Dordrecht, Heidelberg-London-New York. Devi, L., Ptasinski, K.J., Janssen, F., 2003. A review of the primary measures for tar elimination in biomass gasification processes. Biomass Bioenergy 24, 125e140. EEA, 1999. Making sustainability accountable: eco-efficiency resource productivity and innovation topic report NO11/1999. In: Proceedings of a workshop on the occasion of the Fifth Anniversary of the European Environment Agency. EEA, 28e30 October 1998, Copenhagen, Denmark. Evans, R., Milne, T., 1987. Molecular characterization of the pyrolysis of biomass. Energy Fuels 1, 123e137. FitzPatrick, M., Champagne, P., Cunningham, M.F., Whitney, R.A., 2010. A biorefinery processing perspective: treatment of lignocellulosic materials for the production of value added products. Bioresour. Technol. 101, 8915e8922. Gao, J., Andreson, D., Levie, B., 2013. Saccharification of recalcitrant biomass and integration options for lignocellulosic sugars from Catchlight Energy’s sugar process (CLE Sugar). Biotechnol. Biofuels 6, 10. http://dx.doi.org/10.1186/1754-6834-6-10. Gavrilescu, M., 2011. Sustainable Industrial Production. Ecozone Press, Iasi, Romania. Gavrilescu, M., Chisti, Y., 2005. Biotechnology e sustainable alternative for chemical industry. Biotechnol. Adv. 23, 471e499. Goransson, K., Soderlind, U., He, J., Zhang, W., 2011. Review of syngas production via biomass DFGs. Renewable Sustainable Energy Rev. 15, 482e492. Hackl, R., Harvey, S., 2010. Opportunities for Process Integrated Biorefinery Concepts in the Chemical Cluster in Stenungsund. Research Project Report, Department of Energy and Environment, Division of Heat and Power Technology. Chalmers University of Technology. Available from: http://publications.lib.chalmers.se/ records/fulltext/local_131485.pdf.
240 14. BIOREFINERY SYSTEMS: AN OVERVIEW Hajny, G.J., 1981. Biological Utilization of Wood for Production of Chemicals and Foodstuffs. United States Department of Agriculture. Available from: http://www.fpl.fs.fed.us/documnts/fplrp/ fplrp385.pdf. Harris, F.S., 1919. The Sugar-Beet in America. The MacMillan Company, New York. Available from: http://booksnow2.scholarsportal.info/ ebooks/oca4/33/sugarbeetinameri00harruoft/sugarbeetin ameri00harruoft.pdf. Hens, L., Quynh, L.X., 2008. Environmental space. In: Encyclopedia of Ecology. Academic Press, pp. 1356e1363. Hodge, D., Karim, N., Schell, D., MacMillan, J., 2008. Soluble and insoluble solid contributions to high- solid enzymatic hydrolysis of lignocellulose. Bioresour. Technol. 99, 8940e8948. Hofmann, C., 1873. A Practical Treatise on the Manufacture of Paper in All Its Branches. University Oxford. Hong Chua, C.B., Steinmüller, H., 2010. Development of Eco-efficiency Indicators for a Biorefinery. Energy Institute and Johannes Keppler University Linz. Available from: http://www.energyefficiency.at/ web/artikel/eco-efficiency_indicators.html. Huang, H.J., Ramaswamy, S., 2013. Overview of biomass conversion processes and separation and purification technologies in biorefineries. In: Ramaswamy, S., Hunag, H.J., Ramarao, B.V. (Eds.), Separation and Purification Technologies in Biorefineries. John Wiley and Sons, Ltd, pp. 1e36. Hughes, W., Larson, E., 1998. Effect of fuel moisture content on biomass-IGCC performance. J. Eng. Gas Turbines Power 120, 455e459. Huijser, S., 2009. Synthesis and Characterization of Biodegradable Polyesters (Ph.D. thesis). Eindhoven University of Technology. James, C.E., Hough, L., Khan, R., 1989. Sucrose and its derivatives. In: Herz, W., Grisebach, H., Kirby, G.W., Tamm, Ch (Eds.), Progress in the Chemistry of Organic Natural Products. Springer, Vienna, pp. 117e184. Jeffries, T., Lindblad, P., 2009. We march backward into the future. Curr. Opin. Biotechnol. 20, 255e256. Jentoft, F.C., 2003. Introduction: a historical approach to catalysis. Mod. Methods Heterog. Catal. Available from: http://www.fhiberlin.mpg.de/acnew/department/pages/teaching/pages/ teaching__winterseme ster__2003_2004/jentoft_intro_241003.pdf. Kamm, B., 2013. Introduction to biomass and biorefineries. In: Xie, H., Gathergood, N. (Eds.), The Role of Green Chemistry in Biomass Processing and Conversion. Wiley, Hoboken, New Jersey, pp. 1e26. Kamm, B., Kamm, M., Gruber, P.R., 2006. Biorefinery systems e an overview. In: Kamm, B., Gruber, R., Kamm, M. (Eds.), 2006. Biorefineries e Industrial Processes and Products. Status Quo and Future Directions, vol. 1. Wiley WCH Verlag GmbH and Co. KGaA, Weinheim, Germany, pp. 3e40. Khan, A.A., de Jong, W., Jausens, P.J., Spliethoff, H., 2009. Biomass contribution in fluidized bed boilers: potential problems and remedies. Fuel Process. Technol. 90, 21e50. Kim, S., Dale, B.E., 2004. Life cycle assessment of integrated biorefinery cropping systems: all biomass is local. In: Presentation within Agriculture as a Producer and Consumer of energy, Michigan State University. Kitano, M., Tanimoto, F., Okabayashi, M., 1975. Levulinic acid, a new chemical raw material; its chemistry and use. Chem. Econ. Eng. Rev. 7, 25e29. Knauf, M., Moniruzzaman, M., 2004. Lignocellulosic biomass processing a perspective. Int. Sugar J. 106, 147e150. Kobayashi, Y., 2010. Lipase-catalyzed polyester synthesis e a green polymer chemistry. Proc. Jpn. Acad. Ser. B 86, 338e365. Kupiainen, L., 2012. Dilute acid catalysed hydrolysis of cellulose e extension to formic acid. Acta Universitatis Ouluensis, C TecHNICA 438 (Ph.D. thesis). University of Oulu. Lapas, A., Herackeous, E., 2011. Production of Biofuels via Fischer e Tropsch Synthesis: Biomass to Liquids in Handbook of Biofuels and Technologies. Woodhead Publishing ltd, Cambridge UK, 493e529. Laser, M., Larson, E., Bale, B., Wang, M., Greene, N., Lynd, L.R., 2009. Comparative analysis of efficiency, environmental impact and process economics for mature biomass refining scenario. Biofuels, Bioprod. Biorefin. 3, 247e270. Laufenberg, G., Kunz, B., Nystroem, M., 2003. Transformation of vegetable waste into value e added products: (A) the upgrading concept, (B) practical implementations. Bioresour. Technol. 87, 167e198. Lee, R.A., Lavore, J.M., 2013. From first to third generation biofuels: challenges of producing a from a biomass of microarating complexity. Anim. Front. 3, 6e11. Lestare, D., Mudler, W., Sanders, J., 2010. Improving Jatropha curcas seed protein recovery by using counter current multistage extraction. Biochem. Eng. J. 50, 16e23. Liu, X.Y., Ding, H.B., Wang, J.Y., 2010. Food waste to bioenergy. In: Khanal, S.K., Surampalli, R.Y., Zhang, T.C., Lamsal, B.P., Tyagi, R.D., Kao, C.M. (Eds.), Bioenergy and Biofuel from Biowastes and Biomass. ASCE, Reston, Virginia, pp. 43e70. Lloyd, R.A., Harris, J.F., 1955. Wood Hydrolysis for Sugar Production. United States Department of Agriculture, Forest Products Laboratory. Available from: http://ir.library.oregonstate.edu/xmlui/ bitstream/handle/1957/2671/FPL_2029_ocr.pdf?sequence¼1. Marcus, A.I., 2005. Engineering in a Land-Grant Context: The Past, Present, and Future of an Idea. Purdue University Press. Martin, M., Grossmann, I.E., 2012. Optimal synthesis of sustainable biorefineries. In: Stuart, P.R., El-Halwagi, M.M. (Eds.), Integrated Biorefineries: Design, Analysis and Optimization. CRC Press, Boca Raton, pp. 325e347. McKillip, W.J., Collin, G., Hoke, H., Zeitsch, K.J., 2001. Furan and derivatives. In: Ulmann’s Encyclopedia of Industrial Chemistry. Wiley-VCH, Weihneim. Meher, L.C., Vidya Sagar, D., Naik, S.N., 2006. Technical aspects of biodiesel production by trans esterification a review. Renewable Sustainable Energy Rev. 10, 248e268. Melin, K., Hurne, M., 2011. Lignocellulosic biorefinery economic evaluation. Cellul. Chem. Technol. 45, 443e454. NREL, 2009. Biomass research e what is a Biorefinery? Available from: www.nrel.gov/biomass/biorefinery.html. Otulugbu, K., 2012. Production of ethanol from cellulose (sawdust). Available from: http://publications.theseus.fi/bitstream/handle/ 10024/42578/Otulugbu.pdf?sequence¼1. Paulik, C., 2011. Chemical Technology of Organic Materials. Institute of Organic Technology and Organic Materials, Johannes Keppler University of Linz. Available from: http://www.jku.at/cto/content/ e34502/e116152/e118705/OT2Bio-Refinery_1handout_ger.pdf. Peiyadarsan, S., Annamalai, K., Sweeten, J., Munkhtar, S., Holzapple, M.T., 2004. Fixed bed gasification of feedlot manure and poultry litter biomass. Trans. ASAE. 47, 1689e1696. Pennington, N.L., Baker, C.W., 1990. Sugar: User’s Guide to Sucrose. Springer, Berlin, Heidelberg, New York. Peters, F.N., 1937. Furfural as an outlet for cellulosic waste materials. Chem.Eng. News 15, 269e270. PP, 2012. Biowaste Biorefinery in Europe: Opportunities and R&D Needs, Position Paper of the Section of Environmental Biotechnology. European Federation of Biotechnology. Available from: www.efb-central.org/images/uploads/PP_biowastebiorefinery_ 26.02.2012.pdf. Ramey D., 1998. Continuous, two stage, dual path anaerobic fermentation of butanol and other organic solvents using two different strains of bacteria, US Patent 5753474. Ribeiro, N., Pinto, A., Quintella, C., da Rocha, G., Teixeira, L., Guariero, L.L.N., da Corno Rangel, M., Veloso, M.C.C.,
REFERENCES Rozende, M.J.C., da Cruz, R.S., de Oliveira, A.M., Torres, E.A., de Andrade, J.B., 2007. The role of additives for diesel and diesel blended (ethanol or biodiesel) fuels: a review. Energy Fuels 21, 2433e2445. RIRDC, 2006. Furfural chemicals and biofuels from Agriculture. Wondu Business and Technology Services, RIRDC Project No WBT-2A. Available from: https://rirdc.infoservices.com.au/ downloads/06-127.pdf. Rodsrud, G., Lersch, M., Sjode, A., 2012. History and future of world’s most advanced biorefinery in operation. Biomass Bioenergy 46, 46e59. Sanders, J., Scott, E., Mooibrock, H., 2005. Biorefinery the bridge between agricultural and chemistry In: Proceeding of the Fourteenth European Biomass Conference and Exhibition: Biomass for Energy Industry and Climate Protection, Paris, France, 17e21 October 2005. Schobert, H., 2013. Chemistry of Fossil Fuels and Biofuels. Cambridge University Press, Cambridge, UK. Sharara, M.A., Clausen, E.C., Carrier, D.J., 2012. An overview of biorefinery technology. In: Bergeron, C., Carrier, D.J., Ramswamy (Eds.), Biorefinery Co-products: Phytochemicals, Primary Metabolites and Value Added Biomass Processing. John Wiley and Sons, pp. 1e18. Sindall, R.W., 1906. Paper Technology. Charles Griffin and Company Ltd, London. Solecki, M.,Dougherty, A., Epstein, B., 2012. Advanced Biofuel Market. In: Report 2012 Meeting US Fuel Standards, San Francisco, California USA. Speight, J.G., 2011. The Biofuel Handbook. Royal Scociety of Chemistry, London, UK. Star-COLIBRI, 2011. European Biorefinery Joint Strategic Research Roadmap, Star-COLIBRI, Strategic targets for 2020-Collaboration Initiatives on Biorefineries. Available from: www.star-colibri.eu/ files/files/road-map-web.pdf. Swanson, R.M., Satrio, J.A., Brown, R.C., Platon, A., Hsu, D.D., 2010. Technico- Economic Analysis of Biofuels Production Based on Gasification. Technical Report NREL/TP-6A20e46587. National Renewable Energy Laboratory, Golden Colorado USA. van der Drift, van Doorn, J., van Verneulen, J., 2001. Ten residual biomass fuels for circulating fluidized bed gasification. Biomass Bioenergy 20, 45e56. van der Maarel, M.J.E.C., Van der Veen, B., Uitdehaag, J.C.M., Leemhuis, H., Dijkhuizen, L., 2002. Properties and applications of starch-converting enzymes of the a -amylase family. J. Biotechnol. 94, 137e155. van Ree, R., Annevelink, B., 2007. Status Report Biorefinery 2007 Report 847, Agrotechnology and Food Sciences Group, Wageningen, Netherlands Available from: http://edepot.wur.nl/42141. 241 Vink, E.T.H., Rabago, K.R., Glasner, D.A., Gruber, P.R., 2003. Applications of life cycle assessment to natureworks polylactide (PLA) production. Polym. Degrad. Stabil. 80, 403e419. Visvanathan, C., 2010. Bioenergy production from organic fraction of municipal solid waste (OFMSW) through dry anaerobic digestion. In: Khanal, S.K. (Ed.), Bioenergy and Biofuel from Biowastes and Biomass. ASCE Publications, pp. 71e86. Wagemann, K. (Ed.), 2012. Biorefineries Roadmap, as Part of the German Federal Government Action Plans for the Material and Energetic Utilisation of Renewable Raw Materials. Federal Ministry of Food, Agriculture and Consumer Protection (BMELV), Berlin. Available from: http://www.bmbf.de/pub/roadmap_ biorefineries.pdf. Watt, A., 1890. The Art of Paper-making. Crosby Lockwood and Son, London. WBCSD, 2000. Eco-efficiency: creating more value with less impact, World Business Council for Sustainable Development. Available from: http://www.wbcsd.org/web/publications/eco_efficiency_ creating_more_value.pdf. WEF, 2010. The Future of Industrial Biorefineries. World Economic Forum, Cologny, Geneva, Switzerland. Available from: http:// www3.weforum.org/docs/WEF_FutureIndustrialBiorefineries_ Report_2010.pdf. Wolfrom, M.L. (Ed.), 1970. Methods in Carbohydrate Chemistry. Academic Press. Xiu, S., Zhang, B., Shahbazi, A., 2011. Biorefinery processes for biomass conversion to liquid fuel. In: dos Santos, M.A. (Ed.), Biofuels Engineering Process Technology. Intech Open Science, pp. 167e190. Yang, X., li, T., Weston, D., Karve, A., Labbe, J.L., Gunter, L.E., Sukumar, P., Borland, A., Chen, J.-G., Wullschleger, S.D., Tschaplinski, T.J., Tuskan, G.A., 2011. Innovative biological solutions to challenges in sustainable biofuels production. In: dos Santos Bernardes, M.A. (Ed.), Biofuel Production e Recent Developments and Prospects. InTech, Croatia, pp. 375e414. Yuan, Z., Chen, B., Gani, R., 2013. Applications of process synthesis: moving from conventional processes towards biorefinery processes. Comput. Chem. Eng. 49, 217e229. Zhang, Q., Chang, J., Wang, T., Xu, Y., 2007. Review of biomass pyrolysis oil properties and upgrading research. Energy Convers. Manage. 48, 87e92. Zhang, H., Xiao, R., Huang, H., Xiao, G., 2009. Comparison of noncatalytic and catalytic fast pyrolysis of corncob in a fluidized bed reactor. Bioresour. Technol. 100, 1428e1434.
C H A P T E R 15 Catalytic Thermochemical Processes for Biomass Conversion to Biofuels and Chemicals Lin Mei Wu 1, Chun Hui Zhou 1,2,*, Dong Shen Tong 1, Wei Hua Yu 1 1 Research Group for Advanced Materials & Sustainable Catalysis (AMSC), Breeding Base of State Key Laboratory of Green Chemistry Synthesis Technology, College of Chemical Engineering and Materials Science, Zhejiang University of Technology, Hangzhou, Zhejiang, China, 2 The Institute for Agriculture and the Environment, University of Southern Queensland, Queensland, Australia *Corresponding author email: clay@zjut.edu.cn, Chun.Zhou@usq.edu.au O U T L I N E Introduction 243 Pyrolysis of Biomass Fast Pyrolysis Catalytic Pyrolysis Reactors Entrained-Fow Reactors Ablative Reactors Bubbling Fluid Bed Reactor and Circulating Fluidized Beds Rotating Cone Reactor New Systems 244 244 244 245 245 245 Gasification of Biomass 247 245 246 247 INTRODUCTION Thermochemical processing usually refers to the one in which solid reactants are heated at high temperatures for a certain period to yield the desired products. In modern times, the thermochemical processing has often been used in industry for the production of fuels, chemicals and materials. Today, the production of fuels, chemicals and materials from biomass become attractive because it has renewability, one of the advantages Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00015-2 Gasification Catalytic Gasification 247 247 Hydrothermal Liquefaction of Biomass Hydrothermal Liquefaction Feedstock Reaction Conditions Solvent Catalyst 248 248 249 249 250 250 Conclusion 251 References 251 over the fossil oil sources (Zhou et al., 2011; Wu et al., 2013). In a sense, thermochemical processing of biomass is not a new technique. Wood combustion for heating and cooking, a method that humanity have been using since prehistoric time can be regarded as a thermochemical processing of biomass. However, today’s need for thermal processing of biomass is far beyond combustion. The combination of thermal processing and catalysis is bringing about new opportunities for using biomass to produce renewable fuels, 243 Copyright Ó 2014 Elsevier B.V. All rights reserved.
244 15. CATALYTIC THERMOCHEMICAL PROCESSES FOR BIOMASS CONVERSION TO BIOFUELS AND CHEMICALS chemicals and materials (Brown, 2011). The past three decades have witnessed rapid progress in catalytically thermochemical technologies (Zhou et al., 2008; Huber et al., 2006; Fan et al., 2009). Pyrolysis, gasification and hydrothermal liquefaction are major methods frequently tested for the catalytically thermochemical conversion of biomass (Zhou et al., 2011). Other thermochemical technologies could be regarded as modification, in more or less ways, of these three methods. Relevant studies and progress have shown that these technologies are promising alternatives to process diverse biomass feedstocks to yield fine chemicals and biofuels. reaction time of only several seconds or even less (Demirbas and Arin, 2002). Among these, fast pyrolysis is of most commercial interest for production of chemicals and liquid fuels (Zhou et al., 2011). Fast pyrolysis is mainly intended to maximize the biooil yield as well as to increase the contents of the target compounds in it. To this end, there are needs to use a finely ground particle biomass feed of typically less than 3 mm, selective catalysts, a well-controlled pyrolysis temperature of around 773K, short hot vapor residence time of typically less than 2 s and rapid removal and cooling of the products (Bridgwater, 2012). Catalytic Pyrolysis PYROLYSIS OF BIOMASS Fast Pyrolysis Pyrolysis of biomass is thermal depolymerization and decomposition of biomass (TDP) in the absence of air/ oxygen. The temperature generally used is in the range of 623e973K (Goyal et al., 2008). The products, charcoal or biochar, gaseous and liquid chemicals, depend on the biomass composition, the heating rate and the temperature. According to the heating rate, pyrolysis is classified as slow pyrolysis, fast pyrolysis and flash pyrolysis. Slow pyrolysis of biomass is conducted at slow heating rates (0.1e1  C/s). In the relatively low temperatures of 573e673K, charcoal is the main product; when the temperature is are increased to >673K, the oil yield is increased (Putun et al., 2001; Ӧzbay et al., 2001; Onay and Kockar, 2004). By contrast, fast pyrolysis is conducted at higher heating rates (about 10e200  C/s) and intended to produce liquid bio-oil (Bridgwater, 2003). Flash pyrolysis is conducted at heating rates >1000  C/s within TABLE 15.1 Table 15.1 shows some results from recent studies on the pyrolysis of lignocellulosic biomass in the presence or absence of catalysts. Earlier, catalysts such as carbonates and hydroxides were mainly tested for the catalytic pyrolysis of lignocellulosic biomass. The use of alkaline compounds like NaOH, Na2CO3 and Na2SiO3 resulted in bio-oils rich in acetol and to some extent favored H2 formation. The use of Fe2(SO4)3 as a catalyst favored the formation of furfural and 4-methyl-2-methoxyphenol (Chen et al., 2008). Lu et al. revealed that SO2 4 /SnO2 were an effective catalyst to yield 5-methyl furfural (Lu et al., 2009). The selectivity varied significantly once the catalyst support was altered. For example, SO2 4 /TiO2 catalyst favored the formation of furfural; SO2 4 /ZrO2 catalyst favored the formation of furan (Lu et al., 2009). Liquid acids such as H2SO4, hydrochloric acid, phosphoric acid and solid acids such as ZSM-5, Al-MCM-41 are also used as catalysts (Table 15.1), in addition to their uses in the pretreatment of lignocellulosic biomass (Lu et al., 2011). The typical products of liquid acid catalytic Typical Pyrolysis of Lignocellulosic Biomass in the Presence or Absence of Catalysts Feedstock Catalyst Reactor Fast Pyrolysis Typical Products References Cellulose No Entrained-flow reactor 1173 K, 3 h CO, H2, CH4, hydrocarbons Lanza et al., 2009 Pine Sawdust H-ZSM-5 Conical spouted-bed reactor 673e773 K*, 50 ms, under N2 flow C4-hydrocarbons Olazar et al., 2000 Wood Ni-ZSM-5 Tubular quartz microreactor 873 K, under He flow Hydrocarbons toluene French and Czernik, 2010 Cellulose 0.1 wt% sulfuric or polyphosphoric acid Quartz boat 773 K, under He flow Levoglucosane Levoglucosenone Dobele et al., 2005 Pine Wood Sawdust NaOH Conical spouted-bed reactor 773 K, microwave heating Acetol Chen et al., 2008 Pine Wood Sawdust Fe2(SO4)3 Conical spouted-bed reactor 773 K, microwave heating Furfural Chen et al., 2008 Corncob ZnCl2 Quartz reactor 613 K Furfural Qiang et al., 2011 * Flash pyrolysis. All others are fast pyrolysis.
PYROLYSIS OF BIOMASS pyrolysis of biomass are levoglucosenone, furfural and levoglucosane (0.1 wt% sulfuric or polyphosphoric acid as catalyst) (Dobele et al., 2005; Kawamoto et al., 2007). Taking the separation between catalysts and liquid products and corrosive into consideration, the catalytic pyrolysis of lignocellulosic biomass over zeolites and many other solid catalysts have recently received much attention. Microporous zeolite and mesoporous materials such as ZSM-5, Al-MCM-41 have been used catalysts in catalytic pyrolysis of lignocellulosic biomass. In particular, hydrocarbons can be produced in considerable quantities by fast pyrolysis of biomass over these catalysts (Pattiya et al., 2008). Olazar et al. conducted the fast pyrolysis of pine sawdust catalyzed by ZSM-5 in a spouted-bed reactor using nitrogen as the carrier gas. A 12% yield of aromatic compounds was obtained (carbon) (Olazar et al., 2000). French et al. revealed that over nickel, cobalt, iron, or gallium-substituted ZSM-5 at 673e873 K and with a catalyst-to-biomass ratio of 5e10 by weight, lignocellulosic biomass was pyrolyzed to give an approximately 16 wt% yield of hydrocarbons (French and Czernik, 2010). The solid catalysts have advantages over the liquid acid catalysts. However, for solid catalysts at high temperatures, the cracking and deoxygenating activity decreased with time because of the coke formed on them (Carlson et al., 2008). Reactors Several types of reactors have been designed for fast pyrolysis of biomass, and the reactor is very crucial to fast pyrolysis of biomass. There are entrained-flow reactors (EFRs), fluid bed reactors, rotating cone reactors, and ablative reactors. Entrained-Fow Reactors In EFRs, schematically shown in Figure 15.1, biomass particles are usually fed into the reactor in a stream of hot, inert gas. Reaction is typically completed at 973e1073K within a residence time of a few seconds. Dupont et al. used the mixture of two softwoods (sylvester pine and spruce) as a model of biomass to pyrolyze in an EFR (Dupont et al., 2008). The influence of the particle size (0.4 and 1.1 mm), temperature (1073e1273 K), the presence of steam in the gas atmosphere (0 and 20 vol%) and the residence time (between 0.7 and 3.5 s for gas) on conversion and selectivity were studied. Results showed that the particle size was the most crucial parameter that influenced decomposition and more than 70 wt% of gas was produced. Ablative Reactors An ablative pyrolysis reactor is considered as a possible alternative to an EFR. The surface is heated by 245 FIGURE 15.1 Pressurized high temperature entrained flow reactor (PiTER). Source: Tremel et al., 2012; Elsevier. (For color version of this figure, the reader is referred to the online version of this book.) hot flue gas produced by combustion of pyrolysis gases or char and rotates while biomass is pressed onto the hot surface (873K). However, in general, an ablative pyrolysis reactor has difficulty in getting sufficient heat transfer from hot gases to the ablative surface and in contacting feedstock of diverse morphologies (particle shape, structure, and density) with the ablative surface. In practice, relatively few feedstocks would be suitable for ablative pyrolysis. Bubbling Fluid Bed Reactor and Circulating Fluidized Beds A fluid bed reactor is very suitable for fast pyrolysis, as the biomass is rapidly heated and there are high heat and mass transfer rates between gas, particles and catalysts and any other objects in the reactor. Vapor and solid residence time are controlled by the fluidizing gas flow rate. Bubbling fluidized beds (Figure 15.2(a)) are usually referred to simply as fluidized beds, which provide good temperature control and very efficient heat transfer to biomass particles due to high density of solids in the bed. Jung et al. pyrolyzed rice straw and bamboo sawdust in a bubbling fluidized bed equipped with a char separation system (Jung et al., 2008). They found that the maximum bio-oil yield was above 70 wt% and a higher feed rate and a smaller feed size were more favorable to the production of bio-oil.
246 FIGURE 15.2 15. CATALYTIC THERMOCHEMICAL PROCESSES FOR BIOMASS CONVERSION TO BIOFUELS AND CHEMICALS (a) Bubbling fluid bed reactor and (b) circulating fluid bed reactor. Source: Bridgwater, 2012; Elsevier. While circulating fluid bed reactors have similar features to bubbling fluid bed reactors. A main difference between them is the amount of gas used to fluidize the bed. In the circulating fluid bed reactor (Figure 15.2(b)), the gas flow is intentionally set high enough to transport particles out of the bed, which are recovered by gas cyclones and then returned to the fluidized bed. All char is burned in the secondary reactor to reheat the circulating sand or is separated as a fine powder. The circulating fluidized bed can be divided into two zones: pyrolysis zone; and reduction and cracking zone (Wu et al., 1992). In the pyrolysis zone, biomass is loaded into the bed and pyrolyzed very quickly to form char, tar, H2O, and gas (CO2, CO, CH4, CnHm and H2). In the reduction and cracking zone, pyrolysis char of contributes to secondary cracking in the vapor phase. For example, with the circulating fluidized bed as reactor, Dai et al. pyrolyzed wood powder at 773K (Dai et al., 2000). The main effects were: (1) the higher temperature and longer residence time contributed to the secondary reactions and then lead to less liquids; (2) the lower heating rate favored the carbonization and reduced the liquid production; and (3) most compounds in bio-oil were nonhydrocarbons and alkanes, aromatics, and asphalt were relatively less. Rotating Cone Reactor Rotating cone reactor is designed to achieve the intense mixing and heat transfer between biomass and heat carrier without the use of a large amount of fluidizing gas (Figure 15.3). Gas is needed for char burn-off in a secondary bubbling fluid bed combustor and sand transport recirculated to the pyrolyzer. Flash FIGURE 15.3 Rotating cone pyrolysis reactor and integrated process. Source: Bridgwater, 2012; Elsevier.
247 GASIFICATION OF BIOMASS pyrolysis of wood dust was processed in a rotating cone reactor by Wagenaar et al. the cone geometry was specified by a top angle of p/2 radians and a maximum diameter of 650 mm (Wagenaar et al., 1994). The rotating cone reactor model included the description of the particle flow behavior, the particle conversion and the gas-phase cracking of tar vapors. It appeared that the product distribution was affected by the gasphase reaction kinetics and residence time and the gas-phase residence time was determined by the available reactor volume and the feed rate of the wood particles. New Systems Recently, emerging technology is to couple a pyrolysis reactor with other catalytic reactors such as steam reformer and hydrogenation. For example, the technology of hydroprocessing is intended to convert bio-oil to petroleum-refinery compatible feedstock (Elliott et al., 2012). The combination can also be used to build a microscale pyrolysis reactor coupled to the molecular-beam mass-spectrometer (Bahng et al., 2009). It can be used as a very efficient tool for studying mechanisms of thermal and catalytic processes and to optimize process conditions for different products from a variety of feedstocks. GASIFICATION OF BIOMASS Gasification Gasification of biomass is to convert it into useful gases such as carbon monoxide, hydrogen and light hydrocarbons (Brown, 2003). Since the mid-1980s, interest has grown on the subject of catalysis for biomass gasification. The advances in this area have been driven by the need for producing tar-free gases from biomass. The avoidance of tars and the yield of hydorgen are deciding factors the economic viability of the biomass TABLE 15.2 gasification process. Major reactions in gasification are as follows (Brown, 2003). 1 carboneoxygen reactionk C þ O2 4 CO 2 Boudouard reaction C þ CO2 4 2CO carbonewater reaction C þ H2 O 4 H2 þ CO hydrogenation reaction C þ 2H2 4 CH4 wateregas shift reaction CO þ H2 O 4 H2 þ CO2 methanation CO þ 3H2 4 CH4 þ H2 O The desired product from gasification of biomass is hydrogen or syngas. Syngas can be burned directly in gas engines, be used to produce methanol, or be converted into synthetic fuels via the FischereTropsch process. Though gases are target products, gasification of biomass leaves behind solid residuals such carbon and inorganic compounds (ash). Gasification of biomass is normally performed in the presence of steam and the process depends on the occurrence of the steam-reforming reactions. Water, in the form of steam, is often added to promote additional production of hydrogen via the wateregas shift reaction. As the biomass is heated, moisture contained in the biomass is converted to steam, which can react with biomass. However, in practice, proper drying of biomass before feeding it into gasification equipment is still needed in view of energy-input. Small amounts of oxygen can also be added to the gas feed. The heat from exothermic oxidation reactions can then be used by the endothermic steam-reforming reaction. In addition, oxygen has a function to delay the catalyst deactivation by helping burn off some of the coke formed. Catalytic Gasification Table 15.2 lists some typical results from the gasification of lignocellulosic biomass in the presence of catalysts. Typical Gasification of Lignocellulosic Biomass in the Presence of Catalysts Feedstock Catalyst Gasification Typical Products References Cellulose Pt/Al2O3 623 K, 10 min, 30 MPa (N2) H2 and CH4 Usui et al., 2000 Sawdust Ru/C 773 K, 20 min, 27 MPa (N2) H2 Hao et al., 2005 Cellulose SnO2, ZnO 573 K, 1 h, 8 Mpa CO, CO2 and H2 Sinag et al., 2011 CH4 and CO2 Nieolivine 1073 K, 80 h Rich H2 gas Courson et al., 2002 Cellulose Rh/CeO2/SiO2 823e923 K, 0.1 MPa Syngas Tomishige et al., 2004 Glucose NaOH 723 K, 34 MPa H2 Onwudili and Williams, 2009 Sawdust Fe/CaO 933 K H2 Huang et al., 2012
248 15. CATALYTIC THERMOCHEMICAL PROCESSES FOR BIOMASS CONVERSION TO BIOFUELS AND CHEMICALS Biomass gasification is inevitably accompanied with tar formation. Nevertheless, tar can be effectively minimized by catalytic cracking. Naturally occurring dolomite (CaMg(CO3)2), for example, has been used as a catalyst for gasification of biomass in a fluid bed reactor to reduce the tar content by transforming it to gases (Delgado and Aznar, 1997). The mineral-based catalyst generally contains CaO, MgO, CO2 and trace minerals such as SiO2, Fe2O3 and Al2O3. The tar cracking efficiency over the dolomites depends on their chemical composition. In general, dolomites with the lowest content of CaO and MgO show the lowest tar cracking efficiency. Yu et al. gasified birch on the four types of dolomites (deposites in Zhenjiang, Nanjing, Shanxi, and Anhui, China) and a Swedish dolomite (Sala) (Yu et al., 2009). The result was that Anhui dolomite showed a low catalytic capacity to crack tar at 973 and 1073K due to its lowest content of CaO and MgO among the tested dolomites. An alternative can be naturally occurring particles of olivine, which are a mineral containing magnesium oxide, iron oxide and silica. Regarding their attrition resistance, Olivine is advantageous over dolomite (Devi et al., 2005). Alkali salts are often added to biomass by dry mixing or wet impregnation and used as catalysts for the elimination of tar and upgrading of the product gas (Li et al., 1996; Encinar et al., 1998). But it has considerable difficulty in catalyst recovery and disposal of ash. Carbonates, oxides and hydroxides of alkali metals can effectively catalyze the decomposition of tar during catalytic gasification (McKee, 1983). Earlier, for example, Mudge et al. investigated the catalytic steam gasification of wood using alkali carbonates and naturally occurring minerals (trona, borax), which were either impregnated or mixed with the biomass (Mudge and Baker, 1985). The order of activity reported was potassium > carbonate > sodium carbonate > trona > borax. The Ni-based catalysts for biomass gasification in a fluid bed reactor are typically Ni-Al based one (Garcı́a et al., 2002; Arauzo et al., 1997) and Ni/olivine one (Courson et al., 2002, 2000). Ni catalysts help to remove tars and methane and to adjust the composition of synthesis gas. Sinag et al. studied the effect of nano-sized and bulky ZnO and SnO2 at 573 K on the wateregas shift reaction in gasification of cellulose. The results showed that the wateregas shift reaction proceeded faster over ZnO catalysts than that over SnO2 catalysts. Therefore, a higher yield of hydrogen was obtained in the presence of ZnO (Sinag et al., 2011). However, catalysts often suffer from deactivation by sintering and/or coke deposition. The use of supercritical water can prevent catalyst from deactivation by means of extracting the coke precursor from the catalyst surface (Baiker, 1999). In addition, it can improve solubility of cellulosic materials and thus reduce mass-transfer limitation. It is also worth noting that, in addition to the active component in a catalyst, usually the acidity and basicity of a support is also an influential factor on product distribution and coke formation. Tasaka and coworkers disclosed that steam reforming of tar derived from cellulose gasification was efficiently catalyzed by 12 wt% Co/MgO catalyst at 873 K in a fluidized bed reactor (Tasaka et al., 2007). Supported Ru, Pt or Pd catalysts also appear promising in the catalytic gasification of lignocellulosic biomass. They were able to overcome the shortcomings of Ni-based catalysts and dolomite catalysts, although they are relatively costly. Usui et al. gasified cellulose in hot-compressed water at 623 K in the presence of a series of supported catalysts such as Zr(OH)4, (CH3COCH]C(Oe)CH3)3Fe, ferrocene, Ru3(CO)12, (CH3COCH]C(Oe)CH3)2Co, NiC2O4, NiO, Ni(OH)2, PdI2 and Cu(OH)2. After reaction for 3 h, 5 wt% Pd supported on Al2O3 showed the highest catalytic activity, leading to a 42.3 vol% yield of H2 and a 7.7 vol% yield of CH4 (Usui et al., 2000). Tomishige et al. found that the order of M/CeO2/SiO2 catalyst activity in the cedar wood gasification at 823 K was the following: Rh > Pd > Pt > Ni]Ru (Tomishige et al., 2004). For Rh/ CeO2/M-type (M]SiO2, Al2O3, and ZrO2) catalysts for cellulose gasification in a continuous-feeding fluidizedbed reactor, Asadullah et al. found that Rh/CeO2/SiO2 exhibited the best performance in terms of generating syngas or hydrogen (Asadullah et al., 2001, 2003). HYDROTHERMAL LIQUEFACTION OF BIOMASS Hydrothermal Liquefaction Hydrothermal liquefaction of biomass makes biomass react at high-temperature aqueous solutions under high vapor pressures. In the field of geochemistry and mineralogy, this method also was used for getting insights into the solubility of minerals in hot water under high pressure (Zhang et al., 2010; Wu et al., 2012; Tong et al., 2013). Hydrothermal liquefaction involves thermal depolymerization in an aqueous or organic medium. In this context, it might be called a depolymerization process using hydrous pyrolysis for decomsition of complex organic materials (for example, here biomass) into light crude oil. In this way, it is expected that under pressure and upon heat, long-chain lignocellulosic polymers decompose into short-chain petroleum-like hydrocarbons and chemicals. In this aspect, pyrolysis technologies are best suited for the conversion of dry feedstocks (<5% moisture), while hydrothermal liquefaction of biomass is ideal for processing high-moisture (i.e. wet) biomass.
HYDROTHERMAL LIQUEFACTION OF BIOMASS Direct hydrothermal liquefaction involves converting biomass to an oily liquid under elevated pressures (50e200 atm) and at low temperatures (473e673K) to keep water in either liquid or supercritical state. In a hot pressurized water for sufficient time, hydrothermal treatment of biomass breaks down the solid biopolymeric structure to liquid components and even gases. Usually, the liquid product from the hydrothermal liquefaction of lignocellulosic biomass is a complicated mixture with a wide range of compositions. It typically consists of glycoaldehyde dimers, 1,3-dihydroxyacetone dimers, anhydroglucose, soluble polyols, 5-hydroxy-methylfurfural (HMF), furfural, organic acids, phenolic compounds and even hydrocarbons. Feedstock Hydrothermal liquefaction has been applied to a wide range of biomass. One of the advantages of hydrothermal processing is the use of high-moisture biomass without the need for preliminary drying of the biomass. The feedstocks can be cellulose, hemicellulose, lignin, aquatic biomass such as duckweed, microalgae, microalgae, wastes animal manure and human sewage. In general, the presence of high cellulose and hemicelluloses content in biomass yields more bio-oil (Akhtar and Amin, 2011). For example, hardwood samples (cherry) produced more oils than softwood (cypress) due to the high lignin contents in the latter biomass (Bhaskar et al., 2008). Besides the oil yield, the oil composition is also different when different feedstocks are used. Karagöz et al. made analysis of oil compositions obtained from hydrothermal treatment of sawdust, rice husk, lignin and cellulose at 553K for 15 min (Karagöz et al., 2005b). The conclusion was that the oil from the cellulose mainly consisted of furan derivatives, whereas ligninderived oil mainly contained phenolic compounds. The compositions of oils from sawdust and rice husk contained both phenolic compounds and furans; however, phenolic compounds were dominant. But rice husk-derived oil consists of more benzenediols than sawdust-derived oil. In addition, hydrothermal liquefaction of algae biomass has also received much attention. The advantage of microalgae compared to terrestrial biomass is its much higher photosynthetic efficiency, which results in higher growth rates and improved CO2 mitigation (Brennan and Owende, 2009). Studies on the hydrothermal processing of microalgae indicated that 30e60% of the algal biomass can be converted to bio-oils (Tsukahara and Sawayama, 2005; Patil et al., 2008). With different biochemical content of pristine microalgae, the oil from the hydrothermal liquefaction of microalgae 249 is different. Biller and Ross liquefied microalgae and cyanobacteria with different biochemical contents (lipids, proteins and carbohydrates) under hydrothermal conditions at 623K, w200 bar in water (Biller and Ross, 2011). The results indicated that bio-oil formation followed the trend: lipids > proteins > carbohydrates, and proteins produced large amounts of nitrogen heterocycles, pyrroles and indoles; carbohydrates produced cyclic ketones as well as phenols while lipids were converted to fatty acids. Reaction Conditions Hydrothermal liquefaction of biomass need be accomplished with careful choices of time, temperature, pressure, catalyst and the use of reducing gases. Increasing temperature in a certain range is favorable. Temperature control is important because after reaching a maximum of the oil yield, further increase in temperature actually inhibits biomass liquefaction due to the secondary decomposition, Bourdard gas reactions and char formation (Mok and Antal, 1992; El-Rub et al., 2004; Zhong and Wei, 2004). The choice of temperature also depends on the biomass types. Rogalinski et al. carried out a kinetic study on hydrolysis of different biopolymers (Rogalinski et al., 2008). It was observed that cellulose hydrolysis rate in water at 25 MPa increased 10-fold between 513 and 583K and at 553K, a 100% of cellulose conversion was achieved within 2 min. Lignin showed a higher hydrothermal liquefaction temperature than hemicellulose and cellulose. Zhang and Wei found that the optimal temperature of wood hydrothermal liquefaction shifted to a higher value as the lignin content increased (Zhong and Wei, 2004). Pressure increases the density of solvent to facilitate solvent penetration into molecules of biomass components, which results in enhanced decomposition and extraction (Deshande et al., 1987). According to Le Chatelier’s principle, one would expect that the higher the pressure during liquefaction, the less liquid components are gasified. By maintaining pressure above the critical pressure of medium, the rate of hydrolysis and biomass dissolution can be controlled. This can be used to enhance favorable reaction pathways thermodynamically for the production of liquid fuels. However, once supercritical conditions for liquefaction are used, pressure has little or negligible influence on the yield of liquid oil or gas yield because in the supercritical region influence of pressure on the properties of water or solvent medium becomes very weak small (Kersten et al., 2006; Sangon et al., 2006). Reaction atmosphere also need to be considered. The use of reductive gases (e.g. CO and H2) generally improves oil yields with higher H/C ratios (He et al.,
250 15. CATALYTIC THERMOCHEMICAL PROCESSES FOR BIOMASS CONVERSION TO BIOFUELS AND CHEMICALS 2001). The reducing gases stabilize the products of liquefaction by inhibiting the condensation, cyclization, or repolymerization of free radicals. Hence, they help reduce char formation (Xu and Etchevery, 2008). By using H2 instead of Ar atmosphere for liquefaction, Wang et al. found that both the conversion of sawdust and the oil yield were able to be increased (Wang et al., 2007a). Besides the oil yield, the quality of gaseous product is also improved by using H2; For example, CO and C1eC4 products increased and CO2 decreased. Solvent Water is the most common medium used for hydrothermal liquefaction of biomass. The bio-oil obtained from hydrothermal liquefaction of lignocellulose in water is usually a viscous tarry lump with a high oxygen content and low heat. To make bio-oils with low viscosity and high yield, the use of organic solvents is an alternative. The tested ones include ethyl acetate (Demirbas, 2000), acetone (Liu and Zhang, 2008), methanol, ethanol, propanol, butanol, propylene glycol, ethylene glycol, diethylene glycol and so forth  (Mun and Hassan, 2004; KrZan et al., 2005). Liquefaction of biomass with proper solvents is a process that can be integrated with optimized conditions to produce fuel and valuable chemicals. Liu et al. liquefied pinewood in the presence of various solvents (water, acetone and ethanol) in the conditions of temperature range 523e723 K, starting pressure 1 MPa, reaction time 20 min (Demirbas, 2000). The results showed that the highest oil yield reached 26.5% at 473 K in ethanol and the product distribution was strongly affected by the solvent type. The major compound was 2-methoxyphenol (17.20%) for liquefaction in water, while it was 2-methoxy-4-methyl-phenol (8.23%) for liquefaction in ethanol and 4-methyl-1,2-benzenediol (9.49%) for liquefaction in acetone. Recently, it was found that co-solvents are a much more effective than the constituent monosolvents alone due to the synergistic effects of different solvents. For example, biomass conversion in TABLE 15.3 100% ethanol and 100% methanol at 573K is 43% and 42%, respectively, producing a bio-oil yield at approximately 26 and 23 wt%, while the liquefaction in the mixed 50 wt% methanol-water solution or the 50 wt% ethanol-water solution led to a conversion of biomass >95 wt% and a bio-oil yield of as high as 65 wt% at 573K (Cheng et al., 2010). The use of donor hydrogen solvents is a new option to hydrogenate the biomass fragments. These solvents not only donate hydrogen but also act as hydrogen transport vehicle and it was found that the use of tetralin solvent enhanced liquid oil yield by suppressing the formation of asphaltenes, preasphaltenes and gases compared to toluene solvent (nonhydrogen donor) (Wang et al., 2007b). For example, Wang et al. observed that in the presence of solvent the yield of oil increased to 33.1% in toluene (nonhydrogen donor) and 48.4% in tetralin (Wang et al., 2007c). Besides, tetrahydrophenanthrene, octahydrophenanthrene, hexahydropyrene, hexahydrofluorene, and tetrahydroacenaphthene are also useful solvents for hydrogenation (Akhtar and Amin, 2011). Catalyst Table 15.3 summarizes some results of catalytically hydrothermal liquefaction of lignocellulosic biomass. Hydrothermal liquefaction of biomass was significantly affected by catalyst. Lignocellulosic biomass mainly contains cellulosic polymer and lignin polymer. The former readily interacts with acid; the latter readily interacts with alkali. In the presence of alkaline catalysts, liquefaction of lignocellulosic biomass mainly leads to oil-like products (Meszaros et al., 2004; Knill and Kennedy, 2003). The conversion and yield of liquid products decreases in the following order: K2CO3 > KOH > Na2CO3 > NaOH (Karagöz et al., 2006; Akhtar et al., 2010). Typically, the equipment corrosion by caustic hydroxides is severely enhanced under subcritical and Catalytically Hydrothermal Liquefaction of Lignocellulosic Biomass Catalyst Feedstock Reaction Conditions Main Products References Ba(OH)2 or Rb2CO3 Lignin 1.5e8.6 MPa, 573 K, 1 h Phenolic compounds Tymchyshyn and Xu, 2010 Na2CO3 Woody biomass 653 K, 16e20 min, 8 MPa (H2) Heavy oil Qian et al., 2007 Ca(OH)2 Sawdust 553 K, 15 min Oil Karagöz et al., 2004a CoSO4 Cellulose 573 K, 120 s Lactic acid Kong et al., 2008 K2CO3 þ ZrO2 Waste biomass 673 K, 10 min, 22.1 MPa Oil Hammerschmidt et al., 2011 FeSO4 Jack pine powder 623 K, 40 min, 5 MPa (H2) Bio-oil Xu and Etchevery, 2008 CrCl3 Cellulose 473 K, 3 h Levulinic acid Peng et al., 2010
REFERENCES supercritical water conditions. Therefore, in this aspect, alkali and alkaline earth carbonate salts are thought to be optional catalysts. Karagöz et al. found that the alkali and alkaline salts enhanced bio-oil formation from wood hydrothermal processing and the catalytic activity of these catalysts shown a sequence of K2CO3 > KOH > Na2CO3 > NaOH > RbOH > CsCO3 > RbCO3 > CsOH based on heavy oil yield (Karagöz et al., 2004b, 2005a, 2005c). Jena et al. investigated the thermochemical liquefaction of the microalga Spirulina platensis over an alkali metal salt catalyst (Na2CO3), an alkaline earth metal salt (Ca3(PO4)2), and a transition metal oxide (NiO) and without a catalyst (Jena et al., 2012). Results showed that Na2CO3 was found to increase biocrude oil yield, resulting in 51.6% biocrude oil, which was w29.2% higher than that under noncatalytic conditions and w71% and w50% higher than those when NiO and Ca3(PO4)2 were used as catalysts, respectively. Hydrothermal processing of biomass can also be carried out over halide catalysts. Lewis acid catalysts could exhibit good catalytic properties in hydrothermal liquefaction of lignocellulosic biomass while catalytic hydrolysis is frequently conducted in the presence of Brønsted acid catalysts. Transition metal chlorides such as CrCl3, FeCl3, CuCl2 and AlCl3 (Zhang and Zhao, 2010; Li et al., 2009), including a pair of these metal chlorides (for example CuCl2 and CrCl2) (Su et al., 2009), exhibited high catalytic activity. In addition sulfates can also be used as catalysts for the catalytic liquefaction of lignocellulosic biomass. Kong et al. revealed, for example, that lactic acid can be produced from the catalytic hydrothermal liquefaction of lignocellulosic biomass in the presence of different transition metal ions like ZnSO4, NiSO4, CoSO4 or Cr2(SO4)3 (Kong et al., 2008). Recently, natural minerals are used as catalysts in the hydrothermal liquefaction of biomass. Tekin et al. reported the effects of a natural calcium borate mineral, colemanite, on the hydrothermal liquefaction of beech wood biomass (Tekin et al., 2012). The highest light bio-oil yield (11.1 wt%) and the highest heavy bio-oil yield (29.8 wt%) were obtained at 573K over colemanite catalysts. The total bio-oil yields were about 22 and 41 wt% at 573K without and with colemanite, respectively. CONCLUSION Catalytically thermochemical technologies allowed the possibilities to convert biomass into fuels and chemicals. The parameters such as temperature, pressure, feedstock, catalysts, and medium have been extensively studied. In the process of catalytic hydrothermal 251 gasification, catalysts can be naturally occurring minerals (dolomite and olivine); alkali metal catalysts; Ni, Fe, Co, and Cu-based catalysts and supported noble metal catalysts (Rh, Pd, Pt and Ru). Biomass gasification has been profiled as being CO2-neutral, having a potential to produce hydrogen and syngas. For pyrolysis, charcoal, gas and liquid are always produced simultaneously. However, by adjusting process parameters (high heating rates and very high heat transfer rates, controlled pyrolysis reaction temperature at around 773K, short hot vapor residence time, rapid removal of product char and cooling of the pyrolysis vapors), maximizing bio-oil yield could be achieved. Fast pyrolysis has now achieved a nearly commercial success and is being actively developed for producing liquid fuels. Catalytic pyrolysis of biomass could increase the content of the target compounds in the mixture products. Besides, the catalytic pyrolysis of lignocellulosic biomass over zeolites, along with integrated hydroprocesses, offer a new potential way to produce hydrocrabon fuels from biomass. Catalytically hydrothermal liquefaction of lignocellulosic biomass produces a very complex mixture of liquid products (typically consists of glycoaldehyde dimers, 1,3-dihydroxyacetone dimers, anhydroglucose, soluble polyols, 5-HMF, furfural, organic acids, phenolic compounds and even hydrocarbons). Therefore, the novel technology for separation and extraction of downstream products from hydrothermal liquefaction of lignocellulosic biomass need to be developed (Miller et al., 1999). Acknowledgments The authors wish to acknowledge the financial support from the National Natural Scientific Foundation of China (21373185), the Distinguished Young Scholar Grants from the Natural Scientific Foundation of Zhejiang Province (ZJNSF, R4100436), ZJNSF (LQ12B03004), Zhejiang “151 Talents Project”, and the projects (2010C14013 and 2009R50020-12) from Science and Technology Department of Zhejiang Provincial Government and the financial support by the open fund from breeding base of state key laboratory of green chemistry and synthesis technology. References Akhtar, J., Amin, N.A.S., 2011. A review on process conditions for optimum bio-oil yield in hydrothermal liquefaction of biomass. Renewable Sustainable Energy Rev. 15, 471e481. Akhtar, J., Kuang, S.K., Amin, N.S., 2010. Liquefaction of empty palm fruit bunch (EPFB) in alkaline hot compressed water. Renewable Energy 35, 1220e1227. Arauzo, J., Radlein, D., Piskorz, J., Scott, D.S., 1997. Catalytic pyrogasification of biomass. Evaluation of modified nickel catalysts. Ind. Eng. Chem. Res. 36, 67e75. Asadullah, M., Tomishige, K., Fujimoto, K., 2001. A novel catalytic process for cellulose gasification to synthesis gas. Catal. Commun. 2, 63e68.
252 15. CATALYTIC THERMOCHEMICAL PROCESSES FOR BIOMASS CONVERSION TO BIOFUELS AND CHEMICALS Asadullah, M., Tomishige, K., Fujimoto, K., 2003. Catalyst performance of Rh/CeO2/SiO2 in the pyrogasification of biomass. Energy Fuels 17, 842e849. Bahng, M.K., Mukarakate, C., Robichaud, D.J., Nimlos, M.R., 2009. Current technologies for analysis of biomass thermochemical processing: a review. Anal. Chim. Acta 651, 117e138. Baiker, A., 1999. Supercritical fluids in heterogeneous catalysis. Chem. Rev. 99, 453e473. Bhaskar, T., Sera, A., Muto, A., Sakata, Y., 2008. Hydrothermal upgrading of wood biomass: influence of the addition of K2CO3 and cellulose/lignin ratio. Fuel 87, 2236e2242. Biller, P., Ross, A.B., 2011. Potential yields and properties of oil from the hydrothermal liquefaction of microalgae with different biochemical content. Bioresour. Technol. 102, 215e225. Brennan, L., Owende, P., 2009. Biofuels from microalgae e a review of technologies for production, processing, and extractions of biofuels and co-products. Renewable Sustainable Energy Rev. 14, 557e577. Bridgwater, A.V., 2003. Renewable fuels and chemicals by thermal processing of biomass. Chem. Eng. J. 91, 87e102. Bridgwater, A.V., 2012. Review of fast pyrolysis of biomass and product upgrading. Biomass Bioenergy 38, 68e94. Brown, R.C., 2003. Biorenewable Resources: Engineering New Products from Agriculture. Blackwell Publishing, Ames, IA. Brown, R.C., 2011. Introduction to thermochemical processing of biomass into fuels, chemicals, and power. In: Brown, R.C. (Ed.), Thermochemical Processing of Biomass: Conversion into Fuels, Chemicals and Power. John Wiley & Sons, Ltd., England, pp. 1e10. Carlson, T.R., Vispute, T.P., Huber, G.W., 2008. Green gasoline by catalytic fast pyrolysis of solid biomass derived compounds. Chem. Sus. Chem. 1, 397e400. Chen, M.Q., Wang, J., Zhang, M.X., Chen, M.G., Zhu, X.F., Min, F.F., Tan, Z.C., 2008. Catalytic effects of eight inorganic additives on pyrolysis of pine wood sawdust by microwave heating. J. Anal. Appl. Pyrolysis 82, 145e150. Cheng, S., D’cruz, I., Wang, M., Leitch, M., Xu, C., 2010. Highly efficient liquefaction of woody biomass in hot-compressed alcoholwater co-solvents. Energy Fuels 24, 4659e4667. Courson, C., Makaga, E., Petit, C., Kiennemann, A., 2000. Development of Ni catalysts for gas production from biomass gasification. Reactivity in steam- and dry-reforming. Catal. Today 63, 427e437.  Courson, C., Udron, L., Swierczy nski, D., Petit, C., Kiennemann, A., 2002. Hydrogen production from biomass gasification on nickel catalysts: tests for dry reforming of methane. Catal. Today 76, 75e86. Dai, X.W., Wu, C.Z., Li, H.B., Chen, Y., 2000. The fast pyrolysis of biomass in CFB reactor. Energy Fuels 14, 552e557. Delgado, J., Aznar, M.P., 1997. Biomass gasification with steam in fluidized bed: effectiveness of CaO, MgO, and CaOMgO for hot raw gas cleaning. Ind. Eng. Chem. Res. 36, 1535e1543. Demirbas, A., Arin, G., 2002. An overview of biomass pyrolysis. Energy Sources 24, 471e482. Demirbas, A., 2000. Effect of lignin content on aqueous liquefaction products of biomass. Energy Convers. Manage. 41, 1601e1607. Deshande, G.V., Holder, G.D., Shah, Y.T., 1987. Effect of solvent density on coal liquefaction under supercritical conditions. Available from: <http://web.anl.gov/PCS/acsfuel/preprint%20archive/Files/ 30_3_CHICAGO_09-85_0112.pdf> (accessed 14.01.13.). Devi, L., Ptasinski, K.J., Janssen, F.J.J.G., van Paasen, S.V.B., Bergman, P.C.A., Kiel, J.H.A., 2005. Catalytic decomposition of biomass tars: use of dolomite and untreated olivine. Renewable Energy 30, 565e587. Dobele, G., Rossinskaja, G., Diazbite, T., Telysheva, G., Meier, D., Faix, O., 2005. Application of catalysts for obtaining 1,6-anhydrosaccharides from cellulose and wood by fast pyrolysis. J. Anal. Appl. Pyrolysis 74, 401e405. Dupont, C., Commandre, J.M., Gauthier, P., Boissonnet, G., Salvador, S., Schweich, D., 2008. Biomass pyrolysis experiments in an analytical entrained flow reactor between 1073 K and 1273 K. Fuel 87, 1155e1164. Elliott, D.C., Hart, T.R., Neuenschwander, G.G., Rotness, L.J., Olarte, M.V., Zacher, A.H., Solantausta, Y., 2012. Catalytic hydroprocessing of fast pyrolysis bio-oil from pine sawdust. Energy Fuel 26, 3891e3896. El-Rub, A.Z., Bramer, E.A., Brem, G., 2004. Review of catalysts for tar elimination in biomass gasification processes. Ind. Eng. Chem. Res. 43, 6911e6919. Encinar, J.M., Beltran, F.J., Ramiro, A., González, J.F., 1998. Pyrolysis/ gasification of agricultural residues by carbon dioxide in the presence of different additives: influence of variables. Fuel Process. Technol. 55, 219e233. Fan, Y.X., Zhou, C.H., Zhu, X.H., 2009. Selective catalysis of lactic acid to produce commodity chemicals. Catal. Rev. 51, 293e324. French, R., Czernik, S., 2010. Catalytic pyrolysis of biomass for biofuels production. Fuel Process. Technol. 91, 25e32. Garcı́a, L., Benedicto, A., Romeo, E., Salvador, M.L., Arauzo, J., Bilbao, R., 2002. Hydrogen production by steam gasification of biomass using NieAl coprecipitated catalysts promoted with magnesium. Energy Fuel 16, 1222e1230. Goyal, H.B., Seal, D., Saxena, R.C., 2008. Bio-fuels from thermochemical conversion of renewable resources: a review. Renewable Sustainable Energy Rev. 12, 504e517. Hammerschmidt, A., Boukis, N., Hauer, E., Galla, U., Dinjus, E., Hitzmann, B., Larsen, T., Nygaard, S.D., 2011. Catalytic conversion of waste biomass by hydrothermal treatment. Fuel 9, 555e562. Hao, X.H., Guo, L.J., Zhang, X.M., Guan, Y., 2005. Hydrogen production from catalytic gasification of cellulose in supercritical water. Chem. Eng. J. 110, 57e65. He, J.B., Zhang, Y., Yin, Y., Funk, T.L., Riskowski, G.L., 2001. Effects of feedstock pH, initial CO addition, and total solids content on the thermochemical conversion process of swine manure. Trans. ASAE 44, 697e701. Huang, B.S., Chen, H.Y., Chuang, K.H., Yang, R.X., Yang, R.X., Wey, M.Y., 2012. Hydrogen production by biomass gasification in a fluidized-bed reactor promoted by an Fe/CaO catalyst. Int. J. Hydrogen Energy 37, 6511e6518. Huber, G.W., Iborra, S., Corma, A., 2006. Synthesis of transportation fuels from biomass: chemistry, catalysts, and engineering. Chem. Rev. 106, 4044e4098. Jena, U., Das, K.C., Kastner, J.R., 2012. Comparison of the effects of Na2CO3, Ca3(PO4)2, and NiO catalysts on the thermochemical liquefaction of microalga Spirulina platensis. Appl. Energy 98, 368e375. Jung, S.H., Kang, B.S., Kim, J.S., 2008. Production of bio-oil from rice straw and bamboo sawdust under various reaction conditions in a fast pyrolysis plant equipped with a fluidized bed and a char separation system. J. Anal. Appl. Pyrolysis 82, 240e247. Karagöz, S., Bhaskar, T., Muto, A., Sakata, Y., Uddin, M.A., 2004a. Low-temperature hydrothermal treatment of biomass: effect of reaction parameters on products and boiling point distributions. Energy Fuels 18, 234e241. Karagöz, S., Bhaskar, T., Muto, A., Sakata, Y., 2004b. Effect of Rb and Cs carbonates for production of phenols from liquefaction of wood biomass. Fuel 83, 2293e2299. Karagöz, S., Bhaskar, T., Muto, A., Sakata, Y., 2005a. Catalytic hydrothermal treatment of pine wood biomass: effect of RbOH and CsOH on product distribution. J. Chem. Technol. Biotechnol. 80, 1097e1102. Karagöz, S., Bhaskar, T., Muto, A., Sakata, Y., 2005b. Comparative studies of oil compositions produced from sawdust, rice husk, lignin and cellulose by hydrothermal treatment. Fuel 84, 875e884.
REFERENCES Karagöz, S., Bhaskar, T., Muto, A., Sakata, Y., Oshiki, T., Kishimoto, T., 2005c. Low temperature catalytic hydrothermal treatment of wood biomass: analysis of liquid products. Chem. Eng. J. 108, 127e137. Karagöz, S., Bhaskar, T., Muto, A., Sakata, Y., 2006. Hydrothermal upgrading of biomass: effect of K2CO3 concentration and biomass/ water ratio on products distribution. Bioresour. Technol. 97, 90e98. Kawamoto, H., Saito, S., Hatanaka, W., Saka, S., 2007. Catalytic pyrolysis of cellulose in sulfolane with some acidic catalysts. J. Wood Sci. 53, 127e133. Kersten, S.R.A., Potic, B., Prins, W., Swaaij, W.P.M.V., 2006. Gasification of model compounds and wood in hot compressed water. Ind. Eng. Chem. Res. 45, 4169e4177. Knill, C.J., Kennedy, J.F., 2003. Degradation of cellulose under alkaline conditions. Carbohydr. Polym. 51, 281e300. Kong, L.Z., Li, G.M., Wang, H., He, W.Z., Ling, F., 2008. Hydrothermal catalytic conversion of biomass for lactic acid production. J. Chem. Technol. Biotechnol. 83, 383e388.  KrZan, A., Kunaver, M., Tisler, V., 2005. Wood liquefaction using dibasic organic acids and glycols. Acta Chim. Slov. 52, 253e258. Lanza, R., Nogare, D.D., Canu, P., 2009. Gas phase chemistry in cellulose fast pyrolysis. Ind. Eng. Chem. Res. 48, 1391e1399. Li, T.C., Yan, Y.J., Ren, Z., 1996. Study on gasification of peat and its kinetic behavior. Fuel Sci. Technol. 14, 879. Li, C.Z., Zhang, Z.H., Zhao, Z.B., 2009. Direct conversion of glucose and cellulose to 5-hydroxymethylfurfural in ionic liquid under microwave irradiation. Tetrahedron Lett. 50, 5403e5405. Liu, Z.G., Zhang, F.S., 2008. Effect of various solvents on the liquefaction of biomass to produce fuels and chemical feedstocks. Energy Convers. Manage. 49, 3498e3504. Lu, Q., Xiong, W.M., Li, W.Z., Guo, Q.X., Zhu, X.F., 2009. Catalytic pyrolysis of cellulose with sulfated metal oxides: a promising method for obtaining high yield of light furan compounds. Bioresour. Technol. 100, 4871e4876. Lu, Q.A., Dong, C.Q., Zhang, X.M., Tian, H.Y., Yang, Y.P., Zhu, X.F., 2011. Selective fast pyrolysis of biomass impregnated with ZnCl2 to produce furfural: analytical Py-GC/MS study. J. Anal. Appl. Pyrolysis 90, 204e212. McKee, D.W., 1983. Mechanisms of the alkali metal catalysed gasification of carbon. Fuel 62, 170e175. Meszaros, E., Jakab, E., Varhegyi, G., Szepesvary, P., Marosvolgyi, B., 2004. Comparative study of the thermal behavior of wood and bark of young shoots obtained from an energy plantation. J. Anal. Appl. Pyrolysis 72, 317e328. Miller, J.E., Evans, L., Littlewolf, A., Trudell, D.E., 1999. Batch microreactor studies of lignin and lignin model compound depolymerization by bases in alcohol solvents. Fuel 78, 1363e1366. Mok, W.S.L., Antal, M.J., 1992. Uncatalyzed solvolysis of whole biomass hemicellulose by hot compressed liquid water. Ind. Eng. Chem. Res. 31, 1157e1161. Mudge, L.K., Baker, E.G., Mitchell, D.H., Brown, M.D., 1985. Catalytic steam gasification of biomass for methanol and methane production. J. Sol. Energy Eng. 107, 88e92. Mun, S.P., Hassan, E.B.M., 2004. Liquefaction of lignocellulosic biomass with mixtures of ethanol and small amounts of phenol in the presence of methanesulfonic acid catalyst. J. Ind. Eng. Chem. 10, 722e727. Olazar, M., Aguado, R., Bilbao, J., Barona, A., 2000. Pyrolysis of sawdust in a conical spouted-bed reactor with a HZSM-5 catalyst. AIChE J. 46, 1025e1033. Onay, O., Kockar, O.M., 2004. Fixed bed pyrolysis of rapeseed (Brassica napus L.). Biomass Bioenergy 26, 289e299. Onwudili, J.A., Williams, P.T., 2009. Role of sodium hydroxide in the production of hydrogen gas from the hydrothermal gasification of biomass. Int. J. Hydrogen Energy 34, 5645e5656. 253 Patil, V., Tran, K.Q., Giselrød, H.R., 2008. Towards sustainable production of biofuels from microalgae. Int. J. Mol. Sci. 9, 1188e1195. Pattiya, A., Titiloye, J.O., Bridgwater, A.V., 2008. Fast pyrolysis of cassava rhizome in the presence of catalysts. J. Anal. Appl. Pyrolysis 81, 72e79. Peng, L., Lin, L., Zhang, J., Zhuang, J., Zhang, B., Gong, Y., 2010. Catalytic conversion of cellulose to levulinic acid by metal chlorides. Molecules 15, 5258e5272. Putun, A.E., Ozcan, A., Gercel, H.F., Putun, E., 2001. Production of biocrudes from biomass in a fixed bed tubular reactor. Fuel 80, 1371e1378. Qian, Y.J., Zuo, C.J., Tan, H., He, J.H., 2007. Structural analysis of biooils from sub-and supercritical water liquefaction of woody biomass. Energy 32, 196e202. Qiang, L., Wang, Z., Dong, C.Q., Zhang, Z.F., zhang, Y., Yang, Y.P., Zhu, X.F., 2011. Selective fast pyrolysis of biomass impregnated with ZnCl2: furfural production together with acetic acid and activated carbon as by-products. J. Anal. Appl. Pyrolysis 91, 273e279. Rogalinski, T., Liu, K., Albrecht, T., Brunner, G., 2008. Hydrolysis kinetics of biopolymers in subcritical water. J. Supercrit. Fluids 46, 335e341. Sangon, S., Ratanavaraha, S., Ngamprasertsith, S., Prasassarakich, P., 2006. Coal liquefaction using supercritical tolueneetetralinmixture in a semi-continuous reactor. Fuel Process. Technol. 87, 201e207. Sinag, A., Yumak, T., Balci, V., Kruse, A., 2011. Catalytic hydrothermal conversion of cellulose over SnO2 and ZnO nanoparticle catalysts. J. Supercrit. Fluids 56, 179e185. Su, Y., Brown, H.M., Huang, X.W., Zhou, X.D., Amonette, J.E., Zhang, Z.C., 2009. Single-step conversion of cellulose to 5-hydroxy-methylfurfural (HMF), a versatile platform chemical. Appl. Catal. A 361, 117e122. Tasaka, K., Furusawa, T., Tsutsumi, A., 2007. Biomass gasification in fluidized bed reactor with Co catalyst. Chem. Eng. Sci. 62, 5558e5563. Tekin, K., Karagöz, S., Bektas, S., 2012. Hydrothermal liquefaction of beech wood using a natural calcium borate mineral. J. Supercrit. Fluids 72, 134e139. Tomishige, K., Asadullah, M., Kunimori, K., 2004. Syngas production by biomass gasification using Rh/CeO2/SiO2 catalysts and fluidized bed reactor. Catal. Today 89, 389e403. Tong, D.S., Xi, X., Luo, X.P., Wu, L.M., Lin, C.X., Yu, W.H., Zhou, C.H., Zhong, Z.K., 2013. Catalytic hydrolysis of cellulose to reducing sugar over acid-activated montmorillonite catalysts. Appl. Clay Sci 74, 147e153. Tremel, A., Haselsteiner, T., Spliethoff, H., Kunze, C., 2012. Experimental investigation of high temperature and high pressure coal gasification. Appl. Energy 92, 279e285. Tsukahara, K., Sawayama, S., 2005. Liquid fuel production using microalgae. J. Jpn. Pet. Inst. 48, 251e259. Tymchyshyn, M., Xu, C.B., 2010. Liquefaction of bio-mass in hot compressed water for the production of phenolic compounds. Bioresour. Technol. 101, 2483e2490. Usui, Y., Minowa, T., Inoue, S., Ogi, T., 2000. Selective hydrogen production from cellulose at low temperature catalyzed by supported group 10 metal. Chem. Lett. 10, 1166e1167. Wagenaar, B.M., Prins, W., van Swaaiij, W.P.M., 1994. Pyrolysis of biomass in the rotating cone reactor: modelling and experimental justification. Chem. Eng. Sci. 49, 5109e5126. Wang, G., Li, W., Chen, H., Li, B., 2007a. The Direct Liquefaction of Sawdust in Tetralin. Energy Sources, Part A 29, 1221e1231. Wang, G., Li, W., Li, B.Q., Chen, H.K., 2007b. Direct liquefaction of sawdust under syngas. Fuel 86, 1587e1593. Wang, G., Li, W., Li, B.Q., Chen, H.K., Bai, J., 2007c. Direct liquefaction of sawdust under syngas with and without catalyst. Chem. Eng. Process. 46, 187e192. Wang, G., Li, W., Chen, H., Li, B., 2007b. The direct liquefaction of sawdust in tetralin. Energy Sources, Part A 29, 1221e1231.
254 15. CATALYTIC THERMOCHEMICAL PROCESSES FOR BIOMASS CONVERSION TO BIOFUELS AND CHEMICALS Wu, J.Z., Xu, B.Y., Lou, Z.F., Zhou, X.G., 1992. Performance analysis of a biomass circulating fluidized bed gasifier. Biomass Bioenergy 3, 105e110. Wu, L.M., Zhou, C.H., Keeling, J., Tong, D.S., Yu, W.H., 2012. Towards an understanding of the role of clay minerals in crude oil formation, migration and accumulation. Earth-sci. Rev. 115, 373e386. Wu, L.M., Zhou, C.H., Zhang, P.P., Yu, W.H., Tong, D.S., 2013. Oil-forming potential from hydrothermal treatment of cellulose in the presence of montmorillonite clay minerals. In: Proceedings of Second International Congress on Catalysis for Biorefineries. Dalian, China. Xu, C., Etchevery, T., 2008. Hydro-liquefaction of woody biomass in sub and supercritical ethanol with iron-based catalysts. Fuel 87, 335e345. Yu, Q.Z., Brage, C., Nordgreen, T., Sjöström, K., 2009. Effects of Chinese dolomites on tar cracking in gasification of birch. Fuel 88, 1922e1926. Ӧzbay, N., Pütün, A.E., Uzun, B.B., Pütün, E., 2001. Biocrude from biomass: pyrolysis of cottonseed cake. Renewable Energy 24, 615e625. Zhang, Z.H., Zhao, Z.K., 2010. Microwave-assisted conversion of lignocellulosic biomass into furans in ionic liquid. Bioresour. Technol. 101, 1111e1114. Zhang, D., Zhou, C.H., Lin, C.X., Tong, D.S., Yu, W.H., 2010. Synthesis of clay minerals. Appl. Clay Sci. 50, 1e11. Zhong, C., Wei, X., 2004. A comparative experimental study on the liquefaction of wood. Energy 29, 1731e1741. Zhou, C.H., Beltramini, J.N., Fan, Y.X., (Max) Lu, G.Q., 2008. Chemoselective catalytic conversion of glycerol as a biorenewable source to valuable commodity chemicals. Chem. Soc. Rev. 37, 527e549. Zhou, C.H., Xia, X., Lin, C.X., Tong, D.S., Beltramini, J., 2011. Catalytic conversion of lignocellulosic biomass to fine chemicals and fuels. Chem. Soc. Rev. 40, 5588e5617.
C H A P T E R 16 Applications of Heterogeneous Catalysts in the Production of Biodiesel by Esterification and Transesterification Luiz P. Ramos*, Claudiney S. Cordeiro, Maria Aparecida F. Cesar-Oliveira, Fernando Wypych, Shirley Nakagaki Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, Curitiba, Paraná, Brazil *Corresponding author email: luiz.ramos@ufpr.br O U T L I N E Introduction 255 Heteropolyacids 257 Zeolites 258 Clay Minerals Clay Minerals Improving Acidity Acid-Activated Clay Minerals in Biodiesel Production Case 1 Case 2 Case 3 260 262 263 263 263 264 INTRODUCTION It is well known that most of the products derived from the chemical industry involve a catalyst in at least one step of synthesis (Figueiredo and Ribeiro, 1987). However, traditional processes for chemical conversion have numerous inconveniences such as the generation of undesirable by-products and environmental pollution. For this reason, civil groups as well as governmental agencies are pressing the industrial sector to overcome these problems by developing alternative processes in which waste generation is minimized or even eliminated. This concept is also part of the atom Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00016-4 Layered Materials Layered Double Hydroxides Layered Hydroxide Salts Layered Carboxylates Layered Materials as Heterogeneous Catalysts in (Trans)Esterification Reactions 265 265 265 266 Polymeric Catalysts 269 Concluding Remarks 272 References 272 266 economy theory proposed by Trost (1991), in which the majority of the atoms present in chemical reagents must be incorporated into useful products. Many traditional processes based on homogeneous catalysis have been reviewed in order to minimize waste generation. In addition, many researchers have shown that heterogeneous catalysts are excellent alternatives to generate lower amounts of waste streams and also to improve the quality of coproducts, which may contribute with additional revenue for the overall production process. The biodiesel industry is another important sector that is following the same strategic pathway (Cordeiro 255 Copyright Ó 2014 Elsevier B.V. All rights reserved.
256 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL et al., 2012; Di Serio et al., 2008). Biodiesel is a biodegradable fuel derived from renewable sources that can be obtained by different routes such transesterification and esterification. The traditional transesterification process of oils and fats is based on the use of a homogeneous system, with methanol as the transesterification agent and a base catalyst, usually an alkoxide or a hydroxide (NaOH or KOH) that generates the corresponding alkoxide in situ (Van Gerpen and Knothe, 2009). Then, the synthetic mono alkyl esters can be used as biodiesel after suitable purification. The main problem of these processes is related to the required purification steps of the mono alkyl esters as well as glycerin, which must be recovered in good condition due to their high commercial value. Ideally, the biodiesel fuel must be free of residues formed in the chemical process like free and bonded glycerin, soaps and water, which are normally used in washing stages for purification. The presence of glycerin in the resulting biofuel is problematic because this polyol may undergo dehydration during combustion, producing a toxic unsaturated aldehyde named acrolein that is not only a dangerous atmospheric pollutant but also a reactive chemical that can be easily involved in condensation reactions, producing an accumulation of carbon deposits that may block filters and compromise the engine performance (Mittelbach and Tritthart, 1988). Soaps and free fatty acids (FFAs) cause degradation of engine components and free water can interfere with the biodiesel acid number and induce hydrolysis and biological contamination under nonadequate storage conditions (Ramos et al., 2003). The traditional fatty acid esterification processes in homogeneous media uses Brønsted acids such as sulfuric or hydrochloric as reaction catalysts (Ilgen et al., 2007). However, the extensive use of these catalysts may induce corrosion in reactor components and pipelines. Also, the purification of the monoesters produced in this way is also expensive and may require additional washing steps and distillation (Altiparmak et al., 2007). A traditional sequence for biodiesel production involves (1) the recovery of vegetable oil by pressing and/or solvent extraction, (2) the oil pretreatment to adjust its properties for transesterification, (3) the transesterification process, (4) the purification by several stages of water washing and (5) the recovery of reaction coproducts, especially glycerin. Each of these steps adds costs to the overall process. Thus, the introduction of operation units that are able to reduce the contamination degree of mono alkyl esters and glycerin may be an important measure to make biodiesel more competitive from an economic and environmental point of view. For these reasons, many researches had focused their efforts to substitute homogeneous catalysts for heterogeneous ones. The biodiesel produced in a heterogeneous system is easily purified and glycerin is of high purity, diminishing the investment to achieve a suitable market specification (Ramos et al., 2003). Many classes of chemical compounds have been tested as solids for heterogeneous catalysis to produce biodiesel by either esterification or transesterification processes. Among these, zeolites, ionic exchange resins, inorganic oxides, layered compounds, guanidines and metal complexes have been already used (Cordeiro et al., 2011). In order to have a truly heterogeneous catalytic process, the solid catalyst must not leach into the reaction medium and it also needs to be stable under the reaction conditions and reusable. While using solid catalysts for biodiesel preparation, whether by esterification or transesterification, the most common catalyst classifications are solid Brønsted acids, Brønsted bases or Lewis acids. The same solid catalyst, however, may present more than one of these sites and depending on the acidity or basicity of the solid, the catalytic performance can vary considerably (Sharma et al., 2011). Recently, in addition to this primary classification, a number of other factors have been considered while developing solid catalysts for esterification or transesterification reactions. The solids hydrophobicity, for once, is used to unveil the water tolerance. The knowledge of the pore and channel system is used to improve the mass transfer of the catalytic substrate, which for this kind of reaction presents a relatively high viscosity (Wilson and Lee, 2012). Metal oxides, mainly calcium (CaO), magnesium (MgO) and strontium (SrO) oxides, are among the most extensively studied solid bases for heterogeneous catalytic processes (Sharma et al., 2011). Among all alkaline and alkaline earth metal oxides, CaO is the most widely studied. Many are the reasons to explain this fact, including its low cost, low toxicity and low solubility in methanol, which is the most commonly used primary alcohol for the catalytic transesterification of oils and fats (Sharma et al., 2011; Kusdiana and Saka, 2001). CaO also has a long catalytic life, with high activity in many recycling processes under moderate reaction conditions (Lopez et al., 2007). Besides these advantages, CaO can be obtained from various and sometimes unusual natural sources. Naturally occurring minerals such as calcite (CaCO3) and several calcium salts (Lopez-Granados et al., 2010) as well as mollusk shells and egg shells (Cho and Seo, 2010) can be used as a source of CaO by calcination. The impregnation of different alkaline salts in zeolites followed by appropriate thermal treatment can produce basic zeolites and the resulting solids have shown good activity as heterogeneous catalysts for transesterification. Studies have shown that the basicity of the resulted zeolite can be related to the electropositive nature of the exchanged alkaline cation (Philippou et al., 2000).
HETEROPOLYACIDS The infrequent use of acid catalysis in transesterification reactions, in comparison to the base catalysis, is in part justified by the lower catalytic activities of the acid compounds. On the other hand, acid catalysts are less sensitive to several contaminants such as water and FFAs, which in many cases can deactivate the base catalyst or drive the catalytic reaction to other products (Van Gerpen and Knothe, 2009). Notwithstanding this apparent disadvantage of the acid catalyst, solid catalysts with Brønsted or Lewis acid properties have been recently investigated in heterogeneous processes. These solids are promising solid catalysts for the replacement of strong inorganic acids that, although effective in both esterification and transesterification homogeneous catalytic systems, have serious adverse factors such as corrosion of the reaction vessels. Furthermore, the use of strong inorganic acids leads to medium- and long-term environmental problems (Helwani et al., 2009a). Thus, the possibility of using solids with acid properties, rather than highly corrosive liquids, therefore replacing homogeneous processes by heterogeneous ones, may be advantageous since higher catalytic efficiencies may be obtained in more sustainable conversion processes. These are likely to outweigh the higher costs that are often associated with the rational synthesis and use of suitable solid acids. Furthermore, the research of acid catalysts has also been driven by the possible use of waste cooking oil and other cheap and widely available raw materials for biodiesel production. For such materials, the catalyst must be suitable for acting in the presence of high water and acid content, properties that are often found in low cost feedstocks. In general, solids with high acid properties usually meet these prerequisites (Oliveira et al., 2010; Zhao et al., 2012). The present work presents a discussion about the most important classes of inorganic solids and polymeric materials that have been applied in the synthesis of (m)ethyl monoesters through esterification or transesterification. However, biological systems such as immobilized lipases are not treated in this book chapter. Luckily, highly qualified reviews have been published recently on this specific subject (Di Serio et al., 2008; Fjerbaek et al., 2009; Tana et al., 2010). HETEROPOLYACIDS The heteropolyacid (HPA) solids, related to the class of polyoxometalates, are often remembered when there is a need for catalysts that are tolerant to the large amounts of water. As already discussed above, such conditions are usually found in the catalytic conversion of low-cost raw materials to liquid biofuels such as biodiesel. 257 Besides their inherent superacidity (pKHþ > 12) (Mizuno and Misono, 1998), which already ensures the achievement of relatively high yields, these compounds can be devised in such a way to produce a pore architecture and a chemical composition that meets the structure and size of the molecules that are involved in both esterification and transesterification reactions. Polyoxometalates, frequently named as POMs, are anionic metal-oxo clusters whose chemical properties can be modulated by the presence of one or more different transition metal ions and the cation used to generate the salt form. The presence of two different metal atoms per polyoxometalate molecule generates compounds with mixed metals, vanadium and molybdenum being the most commonly used. Furthermore, the presence of other atoms in the structure, besides the metal and the oxygen atoms, leads to heteropolyoxometalate compounds with the general formula (XnþMo12O40)(8n), where X can often be as W (V), Si (IV), Ge (IV) and Ti (IV). These anions can be arranged in typical structures such as Keggin and Dawson structures. The protonated form of heteropolyoxometalate anions is referred to as heteropolyacids, which may be defined as a condensed structure of different types of oxyacids. In water, it is expected that all HPA protons are dissociated. The strength of these acids in acetonitrile was estimated. For instance, the acidity of H3PW12O40 in acetonitrile is greater than that observed for the p-toluenesulfonic acid and H2SO4, two acids usually used as catalysts for (trans)esterification in homogeneous catalytic systems (Drago et al., 1997). HPAs are generally soluble in water and other polar media. Therefore, they are usually unsuitable for biodiesel production by heterogeneous catalysis. However, these anionic compounds are water insoluble when presented as salts with large cations such as Csþ. Because of this characteristic, there is a great interest in the application of this family of solid compounds in heterogeneous process, acting as acid, redox and bifunctional catalysts (Li et al., 2007). The (C16TA)H4TiPW11O40 solid, resulting from the combination of a surfactant (C16TA, cetyltrimethylammonium) with an HPA, was recently reported as a water-tolerant solid for the heterogeneous catalytic esterification of palmitic acid (Zhao et al., 2012). The observed high conversion (94.7 wt%) and high efficiency (91.8 wt% yield) were attributed to the presence of Brønsted and Lewis acid sites, its amphiphilic property and high water tolerance. The authors claimed that substrate molecules concentrate around the catalyst through hydrophobic interactions with its lipophilic tail while methanol molecules are absorbed by HPA through hydrogen bonding. The hydrophobic surroundings also promote the separation and/or repulsion of water
258 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL molecules from the surface of the catalyst. Also, the solid catalyst showed a good recyclability and its heterogeneous character was proved through several cycles of recovery and reuse. In order to heterogenize the HPAs and improve their recovery and reuse, their impregnation on zirconia was also investigated (Oliveira et al., 2010). The H3PW12O40 was immobilized on zirconia at different ratios and calcined at 200  C for 4 h. No decomposition of the Keggin anion structure was observed under these conditions. The resulting solids were used in the esterification of oleic acid with ethanol at a 20 wt% loading, 100  C and 4 h with an ethanol:acid molar ratio of 6:1, conditions under which an 88 wt% conversion of oleic acid to ethyl oleate was obtained. A small leaching (8 wt%) of the catalyst was observed at the end of the reaction, therefore affecting the reaction kinetics. The recyclability study indicated that, after being recovered, washed and thermally treated, the solid presented conversion values as high as 70 wt%, that is, 80% of the original value of 88 wt%. These examples and many others have shown that HPA and POWs represent promising catalysts for both esterification and transesterification (Giri et al., 2005; Caetano et al., 2008; Wee et al., 2010; Leng et al., 2009) but their heterogenization in different supports still needs to be improved in order to keep the process totally heterogeneous. ZEOLITES Zeolites are natural or synthetic materials, classically defined as crystalline aluminosilicate compounds (Cundy and Cox, 2003). Zeolites can be prepared by different synthetic routes with different Si/Al molar ratios, crystal structures, and proton exchange levels. These modifications favor the rationalization of catalytic properties such as pore size, hydrophobicity, strength and distribution of acid sites. When designed in a positive way, all these properties can be interesting and useful for applications in heterogeneous catalysis (Corma et al., 1989). The catalytic activity of zeolytes can derive from the properties of the cation that is present in its chemical composition. Moreover, the exchange of these cations by protons may generate different degrees of zeolite acidity, which is also an interesting property for various catalytic processes (Csicsery, 1984). In fact, acid zeolites are used in many industrial catalytic applications, mainly in the petrochemical industry. Another interesting property is the organized and uniform pore distribution and the existence of a cavity system of regular molecular dimensions ranging from <1 nm to over 10 nm, depending on the solid material. This feature may bestow rather important catalytic properties to the resulting material such as selectivity. In general, zeolites and other porous materials of similar composition and textural properties, named zeolite-like materials and zeotypes, have been prepared and used as catalysts in various chemical processes. These solid catalysts present a strong scientific appeal for green chemistry applications since they are considered environmentally benign when one takes into account their chemical composition. There are several reports regarding the use of zeolitebased catalysts for various chemical reactions. Such uses have been recently reviewed (Martinez and Corma, 2011; Rinaldi and Schuth, 2009). This versatility of uses is not only justified for their great variety in chemical composition but also because of the uniformity of their pore structure. Due to the presence of pores and channels, catalysts based on zeolites can present size selectivity that is rarely seen in other solids. This selectivity can be observed for reactants, products and transition state intermediates that are expected to control a given catalytic reaction (Csicsery, 1984). However, for the same reason, these solids not always perform well in catalytic processes that are dependent on one of their main chemical properties (e.g. acidity). This occurs mainly for processes in which the reactants have dimensions exceeding the catalytic channels and pores provided by zeolitic solid catalyst. Therefore, the structure of a zeolite catalyst must be idealized in order to have not only the appropriate chemical property but also the textural properties that would offer an array of pores and channels that are adequate for the diffusion of the chemical reactants. The strategy to meet these two goals is a challenge for the catalytic application of these solids. The preparation of zeotype materials with mesopores (2e50 nm) appears to be the solution to avoid the mass transfer limitations of zeolites in many catalytic processes. In this sense, many efforts have been made in the scientific community to prepare zeolite-like mesoporous materials that are able to address this goal (Tao et al., 2006). For applications in the esterification of fatty acids or in transesterification of oils and fats, in which large molecules are directly involved in the production of biofuels, it is expected that, apart from their high acidity, the surface of the zeolitic and/or zeotype solids should be hydrophobic enough to promote the adsorption of the substrate on the catalyst surface. In this regard, the adsorption of polar molecules may cause deactivation of catalytic sites, such as in the case of water in esterification reactions (Helwani et al., 2009a). For example, faujasite is a highly hydrophilic zeolite that presents high levels of water adsorption. Hence, this material is barely adequate for esterification because water may
ZEOLITES not only inactivate the catalytic sites on the solid surface but also compromise the reaction yields by interfering with the reaction mechanism (Nijhuis et al., 2002). The MCM-41 molecular sieves have been used as an alternative to zeolite microporous materials. Since its discovery in the early 1990s, these molecular sieves have been used as catalyst in different chemical reactions (Beck et al., 1992; Climent et al., 1999), including in the preparation of biofuels (Twaiq et al., 2003). Similar to the zeolites, these mesoporous compounds also have a regular and ordered distribution of pores (mesopores) across the solid catalyst, allowing their use for the conversion of larger molecules (Carmo et al., 2009). Moreover, the incorporation of metals in the structure of mesoporous solids may lead to acidic materials with special characteristics such as a higher hydrothermal stability. For example, the incorporation of aluminum ions in zeolitic materials can lead to a decrease in the Si/Al ratio and a subsequent increase in the amount of the solid acid sites, since it is well known that the lower the framework Si/Al ratio of the zeolite, the lower the strength of its acid sites, regardless of its higher density (Ma et al., 1996; Corma et al., 1989). Furthermore, the catalyst hydrophobicity can also be changed by modifying the Si/Al ratio. This leads to an alteration in the ability of the solid to adsorb nonpolar molecules in the catalytic reactions such as esterification and transesterification, as well as in desorption of polar molecules such as water (Luque et al., 2007). In general, high Si/Al ratios (or low aluminum contents) leads to high solid hydrophobicity. Thus, since the Si/Al ratio modifies the acidity and hydrophobicity of the catalyst, its influence on the catalytic properties is subtle, mainly in esterification reactions. The presence of water is an important factor in the conversion outcome of esterification reactions. The equilibrium constant for ester formation is very low (3.38 for the reaction of acetic acid with ethanol in nonpolar solvents) (Corma et al., 1989). So, to obtain high ester yields, the reaction must be displaced toward the products, for example, by the continuous removal of water from the system. Furthermore, the reaction can be shifted toward the products when working with a large excess of reagents. In order to segregate the water from the reaction environment, it is necessary to work with high reaction rates and this can be achieved with homogeneous acid catalysts such as sulfuric acid. However, for solid catalysts such as the zeolitic materials and zeotypes, the proper balance between strength and density of the acid sites, suitable for a good catalytic performance and water segregation, is often difficult to achieve. The rate of reactions catalyzed by zeolite catalysts and other solid materials is usually very low compared to that of sulfuric acid (Ma et al., 1996). 259 Ajaikumar and Pandurangan (2007) prepared AlMCM-41 materials with different Si/Al ratios (29, 52, 74 and 110) and used these solids in the esterification of acetic and propionic acids with various alcohols (1-hexanol, 2-ethyl-1-isoamyl alcohol and cyclohexanol). With a small addition of aluminum, which translates into a high Si/Al ratio of 110, these authors observed a higher hydrophobicity and a higher catalyst hydrothermal stability of the material concerning the amount of water formed during esterification. Furthermore, the use of more hydrophobic solid materials prevented the subsequent hydrolysis of the ester formed. On the other hand, solid catalysts with lower Si/Al ratio promoted lower levels of alcohol dehydration, which can also be favored at high temperature. As a result, the selectivity of the catalytic reaction is improved toward the ester production as the accumulation of possible by-products (etherified and dehydrated compounds) is decreased. Hence, the hydrophobicity achieved at higher Si/Al ratios was an important factor for the best catalytic performance (catalytic efficiency), whereas the use of low aluminum contents led to more selective catalytic systems. Corma et al. (1989) reported that strong BrønstedeLowry acid sites are required for the catalytic esterification of acetic acid since they are able to protonate the acetic acid carbonyl group. Working with protonated faujazite zeolite after dealuminization, these authors observed that some dealuminized HY zeolites with Si/Al ratio less than or equal to 15 had better catalytic performance. The strong acid sites present in that solid, which correspond to those aluminum vacant sites or sites occupied by one aluminum atom and the respective nearest neighbors, are more active for the catalytic esterification of fatty acids. Zeolites with high Si/Al ratio presented a more hydrophobic surface and this hydrophobicity became the predominant factor for the equilibrium shifting toward the production of alkyl esters. Also, the higher the aluminum content of the zeolite, the higher the observed adsorption effect. Carmo et al. (2009) also prepared solids based on Al-MCM-41 to investigate the relationship between the high availability of acid sites, introduced by increasing the aluminum content in the mesoporous solid, and the degree of esterification of palmitic acid with methanol, ethanol and isopropanol. However, these authors restricted their work to solids with low Si/Al ratios (8, 16 and 32) whose hydrophobicity was much smaller than the solids described in the previous work. This and most of the data already available in the literature propose the utilization of solid catalysts for esterification reactions. In general, these studies have been motivated by the technological challenge of developing a suitable catalytic system to convert vegetable oils or animal fats of high acid number in biodiesel. Hence, by the catalytic
260 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL esterification of their fatty acid content, these materials would be neutralized and become suitable for transesterification in alkaline media. The high aluminum content solid catalysts (Si/Al molar ratio of 8) prepared by Carmo et al. (2009) showed relatively modest palmitic acid conversion values. The highest value achieved in this study was 79 wt% of methyl ester. The authors did not report the effect of hydrophobicity on reaction conversion but only the increased effect of aluminum incorporation in the catalytic activity of the resulting solids. Ma et al. (1996) used different zeolitic solids (zeolite ZSM-5 and three HY zeolites with Si/Al molar ratios of 30, 5.1 and 9.3) to evaluate the relationship between the solid hydrophobicity and its aluminum content with the observed catalytic efficiency in the preparation of ethyl, n-butyl, isopentyl and benzyl acetates; ethyl and n-butyl benzoates and dioctyl phthalates. For all the solid catalysts used in this study, a high selectivity for the expected ester was observed without any formation of ether derivatives. The increase in selectivity with decreasing aluminum content was also reported by Corma et al. (1989). Insoluble inorganic salts and other inorganic solids based on transition metals can also be used as acid catalysts for transesterification. The application of sodium molybdate (Na2MoO4) and sodium tungstate (Na2WO4) has been recently reported as efficient catalysts for biodiesel production under relatively mild experimental conditions (Nakagaki et al., 2008; Santos et al., 2011). In these studies, soybean oil (0.7 mg/g KOH of acid number), degummed soybean oil containing 180 ppm of phosphatides (1.0 mg/g of KOH), and waste cooking oil (1.5 mg/g KOH) were transesterified with methanol (methanolysis). At 65e80  C using a 54:1 methanol:oil molar ratio and 5 wt% catalyst for 3e5 h, both catalysts reached conversions higher than 92 wt% regardless of the feedstock used for methanolysis. The catalytic activity of these compounds was attributed to the presence of molybdenum(VI) or tungsten(VI) strong Lewis acid sites that are probably able to polarize the methanol OeH bond leading to intermediate species that possibly have high nucleophilic character. The heterogeneous nature of the above-mentioned catalysts was investigated through their reuse in several consecutive reaction cycles. Both compounds could be reused for at least three catalytic cycles. However, part of the solid catalysts was lost during the recycling processes due to their reduced particle size and noticeable adherence to the walls of the reaction vessel. To circumvent these problems, both molybdenum (Bail et al., 2013) and tungsten (Santos et al., 2011) compounds were heterogenized in silica obtained by the solegel process and used in esterification of stearic and oleic acids. Improvements were observed in the catalysts’ recovery and reuse and a good catalytic activity was obtained in the first and subsequent recycling stages. Similarly, zirconia impregnated with tungsten oxide (ZrO2/WO2) was also investigated as an acid catalyst for both esterification and transesterification reactions with methanol (Lopez et al., 2007). CLAY MINERALS Clay minerals are composed of hydrous layered silicates that are part of the phyllosilicates family. The phyllosilicates family is broad and is roughly separated by layer types, groups, subgroups and species (Brindley and Brown, 1980). Two basic units are important to build the clay minerals, the first is silicon atoms coordinated tetrahedrically to oxygen atoms (SiO4) and divalent or trivalent metals coordinated octahedrically to hydroxyls (Mþ2/Mþ3(OH)6). The silicon tetrahedral face can share the three corners with other silicon tetrahedral to build a hexagonal two-dimensional pattern, the tetrahedral sheet. A similar procedure can be adopted by the octahedral, where basically two differences can be obtained when Mþ2 or Mþ3 atoms occupy the center of the octahedral. When Mþ2 is used and the octahedral share the edges, all octahedral sites are occupied and a two-dimensional unit is formed, the so-called octahedral sheet. This unit resembles the structure of brucite (Mg(OH)2) and the sheet is called trioctahedral. When Mþ3 is used and the octahedral share the edges, only 2/3 of the octahedral sites are occupied and the resulting two-dimensional unit resembles the structure of gibbsite Al(OH)3; in this case, the sheet is named dioctahedral. In the ideal condition, the apical oxygen of the tetrahedral sheets can be linked to one octahedral, building the clay minerals of the 1:1 layer type. The unshared hydroxyls of the octahedra lie at the center of the tetrahedra at the same “z” level of the shared apical oxygen. Under ideal conditions, two clay minerals of the 1:1 layer type can be obtained. When Mþ2 occupies the center of the octahedral, the structure of chrysotile (Mg3(OH)4 Si2O5) is obtained and the replacement of Mþ2 by Mþ3 yields the structure of kaolinite (Al2(OH)4Si2O5). As both sides of the octahedral have hydroxyls to share, one octahedral sheet can also be combined with two tetrahedral sheets, originating the 2:1 layer-type clay minerals. Again, when Mþ2 occupies the center of the octahedral, the structure of talc is obtained (Mg3(OH)2Si4O10) while its replacement by Mþ3 results in the structure of pyrophyllite (Al2(OH)4Si4O10). Figure 16.1 shows the lateral (A) and top (B) views of the above-cited compounds. In nature, the phyllosilicates are obtained through weathering, which is the phenomenon related to the
CLAY MINERALS 261 FIGURE 16.1 Lateral view (A) and top view (B) of the layered structures of chrysotile (a), kaolinite (b), talc (c) and pyrophyllite (d). (For color version of this figure, the reader is referred to the online version of this book.) disintegration and chemical alteration of rocks and minerals at the Earth’s surface in direct contact with the atmosphere, water and organism. Through this process, many different isomorphic substitutions occur either in the tetrahedral (Si by Al or Feþ3) or in the octahedral sheets (Al or Mg by Feþ2/þ3, Li, Ti, V, Cr, Mn, Co, Ni, Cu and Zn), mainly in clay minerals of the 2:1 type. The isomorphic substitution generates an excess of negative charge into the layers, which needs to be compensated with the intercalation of hydrated cations between the layers. Hence, this substitution generates the cationic exchange capacity and the plastic properties of these clay minerals, particularly when they are dispersed in water. Using the example of talc and pyrophyllite, these minerals can give origin to trioctahedral saponite ððMþ xy $nH2 OÞðM g3y ðA l; F eÞy ÞðS i4x A lx O1 0 ðO HÞ2 ÞÞ, where Mgþ2 is substituted by Al and Fe and Si, by Al. After this substitution, the excess of negative charges in the clay layers are compensated by the intercalation of hydrated cations ðMþ xy $nH2 OÞ. Another example of
262 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL trioctahedral mineral derived from pyrophyllite is the clay mineral hectorite ððMþ y $nH2 OÞðMg3y Liy ÞðSi4 O10 ðOHÞ2 ÞÞ. In the case of dioctahedral talc, the derived clay minerals are montmorillonite ððMþ y $nH2 OÞðAl2y Mgy Þ ðSi4 O10 ðOHÞ2 ÞÞ, beidellite ððMþ y $nH2 OÞAl2 ðSi4x Alx Þ þ3 O10 ðOHÞ2 Þ and nontronite ððMþ x $nH2 OÞFe ðSi4x Alx Þ O10 ðOHÞ2 Þ. Clay Minerals Improving Acidity Most of the clay minerals have low Brønsted and Lewis acidity. The Brønsted acidity and surface area can be slightly improved during drying when interstitial and intercalated cation hydration water molecules are removed. Another way to improve the acidity is to exchange the intercalated species with highly polarizing species such M3þ cations, where the hydrolysis of the solvated water molecules release protons. The Lewis acidity is normally associated with exposed Al3þ or Fe3þ at the broken crystalline edges, which can be increased by heating the clay material to temperatures above 300  C or by acid treatment (Moronta, 2004). As the thermal treatment can lead to the collapse of the clay mineral structure, the acid treatment is the most effective way to improve both Lewis and Brønsted acidity. This procedure was already described in the 1960s (Ryland et al., 1960), when acid-modified smectites provided high gasoline yields when used as a petroleum-cracking catalyst. Acid activation of clay minerals using mineral acids is not new and this procedure causes disaggregation of clay particles, elimination of impurities and improvement of their surface area, porosity and catalytic properties. Acid-activated clays are broadly spread in different industrial processes, being especially used as bleaching agent and catalysts. Although this process was found to be dependent on several factors such as the type of mineral, its crystallinity, morphology and particle size, the effects of the selective acid leaching of a 2:1 clay mineral, which is a first-order process, can be schematically seen as shown in Figure 16.2. The treatment of clay minerals with mineral acids at room temperature tends to replace the intercalated cations by hydrated protons and/or by the leached cations, improving Brønsted acidity while the structure is mostly preserved (Figure 16.2(b)). By contrast, the thermal acid activation has several effects on the structure of the clay minerals, which depend on the temperature, time of treatment and concentration and strength of the acid used. Under mild temperatures, times and acid concentration, the first effect is usually the removal of acid-soluble impurities and partial leaching of the octahedral coordinated metals from the octahedral sheet FIGURE 16.2 Steps of the acid activation. (a) Raw 2:1 clay mineral; (b) original intercalated cations are replaced by hydrated protons; (c) octahedral structural metals are leached out of the structure; (d) hydrated silica obtained by the collapse of the clay structure. (For color version of this figure, the reader is referred to the online version of this book.)
CLAY MINERALS 263 Konwar et al., 2008; Nascimento et al., 2011; Neji et al., 2011; Zatta et al., 2012, 2013; Rezende et al., 2012; Olowokere et al., 2012), as well as in the transesterification of oils and fats (Bokade and Yadav, 2009). Some specific cases will be evaluated in sequence. FIGURE 16.3 Possible mechanism for the formation of Brønsted and Lewis acid sites after treatment of a 2:1 clay mineral constituted basically of Al in the octahedral sites and Si in the tetrahedral sites. (For color version of this figure, the reader is referred to the online version of this book.) (Figure 16.2(c)). During this process, not only the Lewis acidity is generated but also the anions of the acid can be incorporated in the structure. As clays having Mg or Fe in their structure are more easily leached than octahedra occupied by Al, the acid activation must be optimized to extract the best characteristic of each clay mineral. Under extreme conditions, regardless of the type of clay mineral, all the octahedral metals are removed to produce inactive hydrated fibrous silica as reported previously (Figure 16.2(d)) (Wypych et al., 2005). The effect of an effective acid activation is the broadening of the basal X-ray diffraction peak due to the damage of the layers and, finally, to the total collapse of the structure. The physical effects of this activation process are the improvement of the surface area and pore volumes up to a specific point from which these properties are decreased. The pore radii are also constantly reduced during the treatment. For each acid and activation conditions, a new optimization needs to be reported since each clay mineral has a characteristic that depends not only on the mineralogical classification but also on the mine from which it was extracted. As an example for montmorillonites (Wilson and Clark, 2001; Zatta et al., 2012), Figure 16.3 shows a possible mechanism for the formation of Brønsted and Lewis acid sites after mineral acid treatment of a 2:1 clay mineral constituted basically of Al in the octahedral sites and Si in the tetrahedral sites. Acid-Activated Clay Minerals in Biodiesel Production Just a small number of papers have been devoted so far to the use of acid-activated clay minerals in the catalytic esterification of fatty acids (Vijayakumar et al., 2005; Case 1 (Zatta et al., 2012) Standard Texas Montmorillonite STx-1 with the chemical formula ((Ca0.27Na0.04K0.01)[Al2.41Fe(III)0.09 Mg0.71Ti0.03]Si8.00O20(OH)4), supplied by the Clay Mineral Society repository, was activated using phosphoric, nitric and sulfuric acids under different conditions of temperature, time and acid concentrations and the resulting materials were characterized by X-ray diffraction (XRD), nitrogen adsorption isotherms and Fourier transform infrared spectroscopy. Also, the presence of Lewis and Brønsted acid sites in the structure of the catalyst was characterized by pyridine adsorption. Afterward, the materials were evaluated as catalysts in the methyl esterification of lauric acid. Blank reactions carried out in the absence of any added catalyst presented conversions of 32.64, 69.79 and 79.23% for alcohol:lauric acid molar ratios of 60:1, 12:1 and 6:1, respectively. In the presence of the untreated clay and using molar ratios of 12:1 and 6:1 with 12 wt% of catalyst, conversions of 70.92 and 82.30% were obtained, respectively. For some key samples obtained by the acid activation, conversions up to 93.08% of lauric acid to methyl laurate were obtained, much higher than those observed for the thermal conversion (TC) or using raw montmorillonite. Relative good correlations were observed between the catalytic activity and the development of acid sites and structural and textural properties of the acid-leached materials. Case 2 (Zatta et al., 2013) The same sample of montmorillonite STx-1 described above was submitted to acid activation using aqueous solutions of phosphoric acid. The acid treatment was carried out under vigorous stirring at 100  C in a flatbottomed flask connected to a reflux condenser and a heating mantle. The mineral clay and the acid solution were mixed in a 1:4 ratio (mass per volume) using acid concentrations of 0.5, 1, 2 and 4 mol/l. After the acid activation process, the samples were repeatedly washed with distilled water until pH close to 7, dried at 110  C for 24 h and then heated in an oven at 250  C for 2 h. To check the influence of other acids in the activation of montmorillonite STx-1, this clay mineral was subjected to activation with hydrochloric (37% proof), nitric (65% proof) and sulfuric (98% proof) acids. The resulting acid-activated clay materials were characterized and subsequently used in the catalytic conversion of lauric, oleic and stearic acids, as well as of a complex mixture of fatty acids (tall oil) to their corresponding fatty acid methyl esters (FAMEs).
264 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL 100 K10 95 K10 PA PA K10 100 PA K10 PA 95 85 80 75 70 STX STX TC TC TC STX TC STX 65 Conversion (%) Conversion (%) 90 85 80 60 5 0 90 10 0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1 Experiment The results obtained for the best phosphoric acidactivated sample (PA) were compared to those of the TC and from a standard commercial Lewis acid catalyst (K10) and raw montmorillonite (STX). In all experiments, conversion of all samples were carried out for 2 h at a methanol:fatty acid molar ratio of 12:1 and 160  C with 12 wt% of the catalyst in relation to the oil mass. In general, the PAs and the standard Lewis acid catalyst (K10, SigmaeAldrich), which is produced by HCl activation of mineral clays at boiling temperatures, had similar catalytic activities. However, in some cases, the catalytic performance of PA was even better than that of K10 (Figure 16.4). These data were a strong indication that PA and K10 have similar chemical and physical characteristics, even though part of the layered structure is still retained in PA after acid activation. Tests of reuse of the best PA were performed (Figure 16.5) and no significant losses of activity were observed during the first four consecutive reaction cycles (see dotted horizontal line in Figure 16.5). This is a very important observation since, from any practical use in industrial processes, the catalysts must last for long time before deactivation. 4 FIGURE 16.5 Reuse experiments of the best phosphoric acid activated sample. The reactions were carried out at 160  C for 2 h with a methanol:fatty acid molar ratio of 12:1 and a catalyst loading of 12 wt% in relation to the oil mass. an oven at 110  C for 24 h and finally ground until passage through a 60 mesh sieve. The surface area of the acid-activated clays was improved from 71 m2/g to 238 m2/g for the VLn sample and from 123 m2/g to 170 m2/g for the BBn sample. The catalytic activity of these compounds was evaluated in the esterification of different organic acids, using different acylating agents and reflux conditions. In the case of methyl and hexyl esterification of lauric acid (Figure 16.6), the sulfuric acid activation of both clays greatly improved their catalytic activity and this was also valid for other acids and acylating agents. As reported in Case 2, acid activation led to catalytic activities higher than the standard Lewis acid catalyst (K10). 100 80 BBa Vla BBa Vla K10 60 K10 40 Vln BBn 20 Case 3 (Rezende et al., 2012) Two different clay minerals (namely BBn and VLn), mined in Boa Vista District of Paraı́ba (Brazil), were acid-activated using a 10 wt% suspension of clay in 4 mol/l sulfuric acid at 90  C for 2 h to produce the BBa and VLa solids, respectively. The solids were filtered under reduced pressure and washed with distilled water until the washing water had the same pH of the original material. The material was dried in 3 Use cycle Ester (%) FIGURE 16.4 Comparison among the catalytic activities of the best sample obtained after phosphoric acid (PA) activation with those of the thermal conversion (TC), raw montmorillonite (STX) and a standard commercial Lewis acid catalyst (K10). Experiments 1e4 were carried out with stearic acid, 5e8 with lauric acid, 9e12 with oleic acid and 13e16 with tall oil fatty acids. 2 BBn 0 1 Vln 2 3 4 5 6 7 8 9 10 Experiment FIGURE 16.6 Comparison between the catalytic activity of two Brazilian clay minerals (BBn and VLn) and their respective acidactivated counterparts (BBA and VLa) in the esterification of lauric acid. A standard commercial Lewis acid catalyst (K10) was also used for comparison. Experiments 1e5 correspond to the use of methanol, whereas hexanol was used in experiments 6e10.
LAYERED MATERIALS In conclusion, clay minerals are cheap inorganic materials that are readily available worldwide, environmentally friendly and suitable for the development of reusable acid catalysts for the esterification of fatty acids and the transesterification of oils and fats. The natural acidity can be improved by thermal treatment and selective acid activation. Depending on the clay minerals’ origin and genesis, different chemical compositions are possible and different acid treatments are needed to optimize the acidic properties. Normally, the catalysts can be used in several consecutive reaction cycles and, after deactivation, the residual solids can be easily disposed of or even incorporated in native clays for the production of ceramic materials, bricks and roofs, as well as in the production of porcelains. LAYERED MATERIALS Apart from clay materials, different types of layered compounds have been tested as heterogeneous catalysts in processes traditionally mediated by homogeneous catalysts, which are in some cases expensive and highly toxic. Reactions such as Michael addition, cyanoethylations of alcohols, aldolic condensations and condensation of nitro compounds with aldehydes and ketones, and ring openings can be used as examples (Centi and Perathoner, 2008). Layered materials have also been used as solid catalysts for biodiesel production through esterification and transesterification. Most applications involving this class of compounds refer to the use of layered double oxides (LDOs), which are derived from layered double hydroxides (LDHs) by controlled calcination. LDHs, layered hydroxide salts (LHSs) and layered carboxylates are less commonly used for this purpose. This section presents a brief review of the structure of these layered materials in addition to the description of their use and performance as catalysts. Layered Double Hydroxides LDHs are compounds whose individual layers are of brucite-like (Mg(OH)2) structure. In brucite, the layers are electrically neutral with magnesium cations located in the center of an octahedron with six hydroxyl groups in the vertices. The isomorphic substitution of magnesium by trivalent cations forms positively charged layers, which are stabilized by the presence of anions in the interlayer space (Bravo-Suárez et al., 2004). LDHs are represented m 3þ by the formula ½M2þ 1x Mx ðOHÞ2 Ax=m $nH2 O, where 2þ 3þ M and M are divalent and trivalent cations and Am represents an anion with an m-charge and x usually has a value between 0.25 and 0.33 (Crepaldi and Valim, 1998). 265 In this work, the chemical composition of a specific LDH will be condensed as M2þ/M3þ-A. Thus, an LDH whose layers contain Mg and Al and the counterion between the layers is nitrate will be written as Mg/ Al-NO3. One peculiar characteristic of LDHs is the memory effect. Calcination of Mg/Al or Zn/Al LDHs forms mixed and nanostructured mixed metal oxides described as LDO, which are able to reassemble the LDH structure if added back to an aqueous solution containing a salt whose anions will be intercalated between the layers in order to stabilize the LDH structure (Carlino, 1997). These anions can be different from the anions found in the original LDH. This kind of materials can substitute basic homogeneous catalysts like ammonia, ammonium salts or amines and offer an option as nonpollutant solid catalysts that can be easily separated from the reaction system and recovered. Their catalytic activity is related to the large surface area of LDHs, its solid base character and layered morphology. For instance, Zn/Al LDH containing nitrate, sulfate or orthophosphate anions have catalytic activity in esterification reactions (Hu and Li, 2004), while Mg/Al LDH intercalated with t-butoxide is active in transesterification (Choudary et al., 2000). Layered Hydroxide Salts Some layered hydroxides can also undergo isomorphic substitution of hydroxyl groups by other oxo-ions or by water molecules. In the last case, additional anions will be required to neutralize the excess of positive charge in the layers, keeping the cations unaltered, i.e. only divalent cations are present in the layers. The resulting compounds are called layered hydroxide salts. According to this description, LHSs can be classified based on the structure of copper hydroxide nitratedCu2(OH)3NO3dand zinc hydroxide nitratedZn5(OH)8(NO3)2$2H2O. The general formula for an LHS is M2þ ðOHÞ2x ðAn Þx=n $mH2 O, where M ¼ Mg, Ni, Zn, Cu, Co and A ¼ NO3 ; SO2 4 e Cl (Arizaga et al., 2007). The layers in the copper hydroxide nitrate structure are formed by octahedrons whose center is occupied by Cu2þ cations and these are coordinated to hydroxyl groups and nitrate ions that have substituted ¼ of the hydroxyl sites. This example is the easiest description of an LHS. The structure of zinc hydroxide nitrate has two main characteristics. The first is that ¼ of the Zn2þ cations in octahedral coordination with hydroxyl groups migrate out of the layers, leaving empty octahedrons and forming tetrahedrons up and down the empty octahedral sites. Then, each layer is formed by zinc cations in octahedral coordination with hydroxyl groups and in tetrahedral coordination, whose base is formed by three
266 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL hydroxyl groups shared with the main octahedral layer and its apex is occupied by water molecules. The resulting layers have residual positive charge with com2þ tetr position ½Znoct 3 Zn2 ðOHÞ8 ðH2 OÞ2  , where oct and tetr indicate octahedral and tetrahedral sites (Stählin and Oswald, 1971). The residual positive charge in the layer is neutralized with nitrate ions in the interlayer space in a perpendicular position to the layers plane. Normally, nitrate ions do not coordinate directly to the cations in the layers; however, part of the nitrate ions can be grafted to the layers by controlling the pH of the synthesis (Arizaga et al., 2008). The stacking of layers is stable because of the numerous hydrogen bonds that are formed between OH groups in the layer, nitrate ions and the interlayer water molecules. Two adjacent layers are shifted by a factor of b/2 along the (001) plane and are stacked along the basal axis. Zinc laurate melts at approximately 137  C, when nearly 30% of the all-trans conformation is lost. As a result, the structure is partially disassembled but the metals still remain coordinated to oxygen atoms, either bridging two different carboxylate groups or in a monodentade form (Taylor et al., 2006). The application of metal carboxylates in the production of biodiesel became more usual since Cordeiro et al. (2008) demonstrated that the LHS zinc hydroxynitrate (ZHN, Zn5(OH)8(NO3)2$2H2O) is rapidly transformed into zinc carboxylates when used in the catalytic conversion of lauric acid to methyl laurate. According to these authors, zinc carboxylates, which are produced in situ during a reaction course, is the actual catalytic species in the reaction system. Recently, the transformation of LDHs into metal carboxylates was also demonstrated during the catalytic esterification of fatty acids (Cordeiro et al., 2012). Layered Carboxylates Layered Materials as Heterogeneous Catalysts in (Trans)Esterification Reactions Layered carboxylates are also known as metal fatty acid soaps. These carboxylates are used in several applications such as fuel additives, cosmetic products, components of lubricants greases, catalysts, among many others. Carboxylates can be obtained by melting a selected fatty acid or fatty acid mixture with the oxide that contains the metal of interest. Temperatures above 200  C and vigorous agitation are usually required for synthesis. However, the precipitation method is more commonly used. In this method, fatty acids are neutralized with sodium hydroxide in ethanol under stoichiometric conditions. Afterward, an ion exchange is promoted by adding a salt solution of the metal of interest to the sodium carboxylate solution obtained previously, also under stoichiometric conditions. As a result, the desired metal carboxylate is precipitated and subsequently recovered by filtration (Barman and Vasudevan, 2006). Zinc is among the metals most commonly used for the synthesis of carboxylates. However, other cations, mainly di- and trivalent, have also been used such as manganese, nickel, copper and lantanum (Lisboa et al., 2012). In the layered zinc(II) carboxylates, the central zinc metal is tetrahedrally coordinated to bridging bidentate carboxylate groups forming a bilayer of the hydrocarbonic chain bridges in a syn-anti arrangement (Lacouture et al. 2000; Barman and Vasudevan, 2006). Layered carboxylates are isostructural materials and this can be observed through the repeating basal planes of their X-ray diffraction pattern below 20 of 2q (Taylor and Ellis, 2007; Lisboa et al., 2012). In long-chain metal carboxylates such as zinc laurate, methylenic groups of the hydrocarbon chains are organized in a zigzag conformation, which is also referred to as all-trans. The use of LDH as catalysts for transesterification reactions is less common if compared to the use of LDOs derived from LDH by calcination. However, the t-butoxide intercalated Mg/Al LDH (LDH/t-BU) was shown to be catalytically active for production of b-ketoesters by transesterification with primary, secondary and tertiary alcohols (Choudary et al., 2000). Serio et al. (2006) synthesized Mg/Al LDHs by coprecipitation at pH 10. After washing and drying, the LDHs were calcined at 500  C for 14 h to produce the corresponding oxides. Besides, two samples of oxides identified as MgO1 and MgO2 were obtained by calcination of Mg(OH)2 and (MgCO3)4 Mg(OH)2 at 400  C. All these oxides were tested as catalysts for soybean oil methanolysis. Reactions carried out with 10 wt% catalyst at 100  C yielded about 80% of products using the LDO solids and less than 20% with both MgO1 and MgO2. The higher activity of LDO, with respect to other catalysts, was justified by the presence of a higher concentration of very strong base sites and large pores that favored the reaction by rendering the active sites more accessible to the bulky triglyceride molecules. In another study (Serio et al., 2007), an LDO obtained in the same way was used in the methanolysis of soybean oil with and without the addition of 10 wt% of its weight in oleic acid. The reaction was carried out at 180  C for 1 h with 5 wt% catalyst using commercial MgO as a reference material. The methanolysis of neutral soybean oil was catalyzed with LDO and MgO and the yields were 92% and 75%, whereas the corresponding values for the acidified soybean oil were 80.3% and 76.6%, respectively. Unlike the direct use of LDOs, Xi and Davis (2008) rehydrated the LDO and tested the resulting material
LAYERED MATERIALS as catalyst for transesterification. The experiments started with the coprecipitation of an Mg/Al LDH with an Mg:Al molar ratio of four. The material was calcined at 500  C under nitrogen atmosphere to form the LDO and then rehydrated with vapor under nitrogen atmosphere. The crystallinity of the resulting rehydrated LDH was lower than that of the initial LDH. The absence of CO2 in the rehydration process avoided formation of carbonate ions. Hence, the counterion in the LDH structure was the hydroxyl ion. For this reason, the hydrated LDH had more Brønsted sites than a typical LDH. This material was subsequently used in the methanolysis of tributyrine and the yield of monoesters was around 80% when the reaction conditions involved 136.5 g of methanol, 43.0 g of tributyrine and 0.25 g of catalyst at 60  C for 400 min. Zeng et al. (2008) synthesized various LDHs with different Mg:Al molar ratios by coprecipitation and ripened them at 65  C. The solid LDHs were washed and dried at 90  C to be subsequently calcined in a muffle at 673e1073  C for 7 h, with the resulting oxides being tested in the transesterification of refined colza oil. The catalytic activity was correlated with the temperature and time of calcination as well as with the Mg:Al molar ratio. The best yield (90.5%) was obtained from the oxide with Mg:Al molar ratio of three that was calcined at 500  C for 12 h. In this case, the transesterification was carried out with 1.5% of catalyst in relation to the oil mass, a methanol:oil molar ratio of 6:1 and stirring at 300 rpm for 4 h at 65  C. In addition, the reuse assays showed that the catalytic activity was kept for six cycles with a slight decrease in ester yield after each cycle. Mg/Al LDOs were also tested by Xie et al. (2006) in the transesterification of soybean oil with methanol. The precursor was synthesized by coprecipitation at pH 7. The material was calcined for 8 h at different temperatures and the obtained LDO was tested in the transesterification of soybean oil with a methanol:oil molar ratio of 15:1, 7.5% of catalyst and heating under reflux. The Mg:Al molar ratio of three yielded 67% of ester, which was the best result if compared to other molar ratios of 2.0, 2.5, 3.5 and 4.0. The calcination temperature also influenced the catalytic activity. Actually, the calcination temperature affected the basic strength of the oxides as determined by the Hammett method. When the calcination temperature was increased from 300  C to 500  C, the methyl ester yield reached a maximum of 66%. The highest yield was attributed to the achievement of the highest basicity after calcination. According to XRD, this oxide corresponded to the MgO periclase phase. Temperatures above 500  C transformed the crystalline phase into spinel with less basicity and also less catalytic activity. Calcination below 500  C led Al3þ to replace Mg2þ sites and the basicity Al bonded 267 to O2 is lower than that of Mg bonded to O2. For the LDH with an Mg:Al molar ratio of three, calcination at 500  C led to the optimal basicity for catalytic applications in the methanolysis of soybean oil. Cantrell et al. (2005) reported the use of layered materials for the catalytic transesterification of glycerin tributyrate. For this purpose, a series of ½Mgð1xÞ Alx ðOHÞ2 xþ ðCO3 Þ2 x=2 compounds with the x value ranging from 0.25 to 0.55 were calcined at 500  C for 3 h under wet N2 flux (95% humidity). Also, pure Al2O3 and samples of magnesium-impregnated calcined hydrotalcite were used as reference materials and no catalytic activity was detected in any of these compounds. On the other hand, the LDOs improved their catalytic efficiency with an increase in their magnesium content, achieving a maximum ester yield of 74.8% with 25% of magnesium in the LDO structure. The reactions were always performed at the same experimental conditions (60  C for 3 h), in which pure MgO yielded only 11% of esters. In another study, heterogeneous catalytic processes were developed for the alcoholysis of triglycerides using LDOs that were impregnated with alkaline metals (Trakarnpruk and Porntangjitlikit, 2008). Mg/Al-NO3 LDHs were synthesized by coprecipitation and calcined at 450  C for 35 h. The resulting oxide was added to a potassium acetate solution in order to impregnate the oxide with potassium ions. The material was recovered from the solution, dried at 100  C for 12 h and calcined again at 500  C for 2 h. The potassium content of the resulting powder was 1.5%. FAMEs with a 96.9% ester content and methyl ester yields of 86.6% were obtained with these solids in reactions carried out at 100  C for 6 h, using 7% of catalyst and a methanol:oil molar ratio of 30:1. Liu et al. (2007) carried out the catalytic conversion of chicken fat to methyl esters using oxides that were derived from the Mg6Al2(CO3)(OH)16$4H2O hydrotalcite by calcination at different temperatures (400e800  C) for 8 h. As a result, the effect of the calcination temperature on the catalytic performance of the oxide was confirmed, as already described by Xie et al. (2006). High yields of 94 wt% were obtained when the LDH was calcined at 550  C and the reaction was carried out at 120  C for 6 h with a catalyst loading of 0.04 mg/l. The catalyst activity decreased slightly in the first recycling stage but dropped to only 60% of the original value after the fourth consecutive reaction cycle. However, the original activity could be totally recovered by calcination of the spent catalyst in air. Antunes et al. (2008) catalyzed the methanolysis of soybean oil with Mg/Al and Zn/Al oxides that were obtained by calcination of the corresponding LDH at 450  C for 12 h. Transesterification was performed for
268 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL 7 h at 70, 100 and 130  C with a methanol:oil molar ratio of 55:1. The highest activity was detected at 130  C and the yield at this temperature was 80% with MgO, 70% with Mg/Al LDO, 63% with Zn/Al LDH, 30% with ZnO, and 11% with Al2O3. Ilgen et al. (2007) used LDOs derived from Mg6Al2(OH)16CO3$4H2O for the catalytic conversion of canola oil into methyl esters. The LDH was prepared by coprecipitation of magnesium and aluminum carbonate salts at pH 10 and ripened for 18 h. After separation, the solid compound was dried at 80  C and then calcined at 500  C for 16 h. The LDO with particle diameter of 150e177 mm gave a 63% ester yield when the reaction was carried out at 60  C with methanol:canola oil molar ratio of 6:1. Higher molar ratios of 9:1, 12:1 and 16:1 decreased the ester yields to values lower than 60%, and when LDOs with other particle sizes (125, 125e150 and 150e177 mm) were used, the best performance was obtained in the range of 125e150 mm. In the same report, the use of n-hexane as a cosolvent was shown to be detrimental to methanolysis. Also, methanol resulted in better ester yields than ethanol. Barakos et al. (2008) calcined Mg/Al-CO3 at 350  C for 6 h and tested it for methanolysis of cotton oil. Samples with 95% esters were obtained for reactions carried out at 180  C, using methanol:oil molar ratios of 6:1 wt% and 1 wt% of catalyst at 2200 kPa. Albuquerque et al. (2008) prepared calcium and magnesium mixed layered hydroxides by coprecipitation from which LDO catalysts were generated. The LDH was calcined at 800  C and the resulting oxides were tested in the catalytic methanolysis of sunflower oil at 60  C. Higher yields of 92.4% were obtained in methanolysis after 3 h using a methanol:oil molar ratio of 12:1 and a 2.5 wt.% of the solid catalyst with a 3.8 Mg:Ca ratio. Macedo et al. (2006) prepared (Al2O3)4(SnO) and (Al2O3)4(ZnO) LDOs from the corresponding Sn/Al and Zn/Al LDHs and both present catalytic activity in the alcoholysis of soybean oil, even when branched alcohols were used. Yields higher than 80% were obtained with methanol after 4 h at 60  C and the recycling tests indicated that these materials did not lose their catalytic activity. Shumaker et al. (2008) used LDO catalysts to convert soybean oil in methyl esters. The LDH precursors (Mg/Al, Fe/Al and Li/Al) were obtained by coprecipitation and subsequently calcined at 450  C for 2 h. The best catalytic performance was obtained with the oxide derived from [LiAl2(OH)6](CO3)0,5.nH2O, reaching a conversion of 83.1% in 2 h with a methanol:oil molar ratio of 15:1. Under the same conditions, the LDO derived from the Mg/Al precursor yielded only 13.6% of products. The same catalysts were also tested in the methanolysis of glyceryl tributyrate. The reactions were carried out at 65  C under reflux for 1 h with 20 mmol of glyceryl tributyrate, 600 mmol of methanol and 0.1 g of catalyst. The Li/Al LDO gave yields higher than 98% while the LDOs with Mg/Al and Mg/Fe yielded only 32% and 23.9%, respectively. These results were close to the 37.1% yield that was achieved with MgO under the same conditions. These authors also observed the influence of the calcination temperature on the catalytic performance and concluded that the optimal temperature to obtain the best synthetic LDOs is between 450  C and 500  C. Ngamcharussrivichai et al. (2007) used CaO.ZnO mixed oxides as heterogeneous catalysts for the methanolysis of palm kernel oil. A layered hydroxide formed by a mixture of the divalent cations (Ca2þ and Zn2þ) was coprecipitated in alkaline media. The mixed hydroxide was then subjected to calcination between 600  C and 900  C for 2e6 h. Ester yields higher than 94% were obtained with this catalyst after 1 h at 60  C using a methanol:oil molar ratio of 30:1 and a catalyst loading of 10 wt%. Also, the mixed oxide was shown to have a Ca:Zn molar ratio of 0.25. LDHs containing Zn/Al and Mg/Al with different counterions (nitrate, chlorite and carbonate) and M2þ/M3þ ratios were synthesized by Cordeiro et al. (2012) and used as catalysts in the esterification of fatty acids with methanol. The best conversion of 97 wt% was obtained with Zn5AlCl for a reaction that was carried at 140  C with a methanol:lauric acid molar ratio of 6:1 and 2 wt% of the solid catalyst. However, all the LDHs tested were converted in situ to layered carboxylates, which preserved their catalytic activity even after several consecutive cycles of reuse. LDH compounds containing Mg2þ, Ni2þ and Al3þ were synthesized by Wang and Jehng (2011) and calcined at 500  C for 10 h to produce heterogeneous LDO catalysts for biodiesel production. The best condition for synthesis involved the use of a methanol:soybean oil molar ratio of 21, 0.3 wt% of catalyst, 105  C and 1200 rpm for 4 h, when an 87% conversion of soybean oil to methyl esters was obtained. The observed catalytic efficiency was related to the basicity and Mg content of the Mg/Al/Ni catalysts. Corma et al. (2005) also applied LDOs in the transesterification of monoesters with glycerol. LDHs were initially calcined at 450  C under nitrogen flux for 8 h to produce LDOs that were immediately rehydrated in N2 atmosphere to avoid the presence of CO2. MgO was also synthesized from magnesium oxalate by calcination at 500  C for 6 h and used as a control. The LDOs containing Li/Al had a performance better than those containing Mg/Al or MgO due to formation of stronger Lewis basic sites, since Liþ ions, which are more electropositive than magnesium, increase the charge density of the oxygen. Based on this, alumina was impregnated with KF and the resulting material revealed basicity
269 POLYMERIC CATALYSTS even higher than that of the Li/Al LDO. The catalytic conversion of glycerol to glycerin oleate with KF/ Al2O3 was 98% with monoester selectivity of 69%. Cordeiro et al. (2008) showed that the LHS zinc hydroxide nitrate [ZHN, Zn5(OH)8(NO3)2$2H2O] can be used as a heterogeneous catalyst for the esterification of fatty acids and for the transesterification of vegetable oils. For the transesterification reaction carried out at 150  C for 2 h with 5 wt% of ZHN and a methanol:palm oil molar ratio of 48:1, the resulting ester layer contained 95.7 wt% of methyl esters and the purity of the glycerin layer was as high as 93 wt%. Also, when esterification was carried at 140  C for 2 h with a methanol:lauric acid molar ratio of 4:1, the final ester layer contained 97.4 wt% of methyl laurate. In addition, these authors were able to demonstrate that ZHN turned into zinc laurated(C12H23O2)2Zndduring the reaction course and this new layered material was held responsible for the observed catalytic activity. LDHs containing Zn/Al and Mg/Al with different counterions and M2þ/M3þ ratios were used as catalysts in the esterification of fatty acids with methanol. All LDHs were synthesized by coprecipitation and high conversion rates were obtained depending on the reaction condition. For instance, a 97 wt% conversion of lauric acid to methyl laurate was obtained using a methanol:fatty acid molar ratio of 6:1 and 2 wt% of catalyst at 140  C for 2 h. However, all LDHs were also converted in situ into layered carboxylates and this new material was responsible for the observed catalytic activity, which was preserved even after several consecutive cycles of reuse (Cordeiro et al., 2012). Reinoso et al. (2012) used zinc carboxylates (acetate, laurate, palmitate, stearate and oleate) as catalysts for the methanolysis of soybean oil. Methyl ester conversions as high as 98 wt% were obtained for yields in the range of 84 wt% when the reaction was carried out for 2 h with a 30:1 methanol:oil molar ratio and a catalyst loading of 3 wt% in relation to the oil mass. Jacobson et al. (2008) developed a solid catalyst by immobilizing zinc stearate in silica using the solegel method. The resulting solids contained 6 wt% of zinc and a total available surface area of 35 m2/g. These solids were shown to be catalytically active in the methanolysis of used frying oil with high acid number (w15%). High yields of 98 wt% were obtained at 200  C for 10 h using a methanol:oil molar ratio of 18:1 and 3 wt% of catalyst in relation to the mass of the starting material. Lisboa et al. (2012) described the synthesis and characterization of layered copper(II), manganese(II), lanthanum(III) and nickel(II) laurates as well as their catalytic activity in the methyl and ethyl esterification of lauric acid. Conversions between 80 wt% and 90 wt% were observed for all catalysts when methanol was used for esterification, whereas only manganese laurate gave a reasonable catalytic activity of about 75 wt% with the use of ethanol. In general, the best results were obtained at temperatures around 140  C. Also, the structure of copper(II) and lanthanum(III) laurates was shown to be preserved after three consecutive reaction cycles. POLYMERIC CATALYSTS This section describes the application of functionalized polymers as catalysts for esterification and transesterification reactions to produce biodiesel. Polymeric catalysts consist of functionalized polymeric matrixes or polymeric matrixes that can be used as solid supports for a variety of catalysts, constituting catalytic systems (Coutinho et al., 2004a,b; Guerreiro et al., 2010; Lee and Saka, 2010; Zieba et al., 2010). These materials have long been studied as heterogeneous catalysts in systems that traditionally employ acid or basic homogeneous catalysts. Biodiesel production can be carried out in the presence of different types of catalysts. In the specific case of polymeric catalysts belonging to the class of functionalized polymers, their use in biodiesel synthesis has focused on acid catalysts, such as in the case of ion exchange resins (Ma and Hanna, 1999; Guerreiro et al., 2006; Knothe et al., 2006; Soldi et al., 2009; Rezende et al., 2008; Helwani et al., 2009b; Lee and Saka, 2010). Acid-catalyzed triacylglycerol transesterification is not commercially applied as often as catalysis in basic medium because acid catalysis in homogeneous medium is around 4000 times slower than the basecatalyzed reaction. However, acid catalysts can perform esterification and transesterification simultaneously, producing biodiesel directly from oils with high acid number. These oils are not suitable for biodiesel production via alkaline catalysis, since the FFAs promptly react with the base, generating soaps that make the separation between the ester and glycerin difficult during the washing step (Bondioli et al., 1995; Schuchardt et al., 1998; Ma and Hanna, 1999; Vicente et al., 2004; Lotero et al., 2005; Meher et al., 2006; Rezende et al., 2008; Lee and Saka, 2010). Several acid catalysts can be used in alcoholysis, especially sulfonic and sulfuric acids (Hayyan et al., 2011). Although these catalysts afford high yields of mono alkyl esters, they require high temperatures and long reaction times to achieve a satisfactory conversion rate. Another disadvantage is that residual acid catalysts can contaminate the fuel and attack the metal parts of the engine, corroding it. To avoid this situation, acid catalysts must be completely eliminated from the final product, which demands many purification steps (Canakci and Gerpen, 1999).
270 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL Some types of organic polymers and ion exchange resins can be used as polymeric catalysts, behaving as heterogeneous catalysts for esterification and transesterification reactions. Heterogeneous catalysts such as these reduce the number of biodiesel purification steps, facilitates catalyst reuse, and decreases production costs (Schuchardt et al., 1998; Choudary et al., 2000; Fukuda et al., 2001; Harmer and Sun, 2001; Ramos et al., 2003; Abreu et al., 2004; Suppes et al., 2004; Chouhan and Sarma, 2011). The pioneering studies of Merrifield on the solidphase synthesis of polypeptides and subsequent works have shown that polystyrene (PS) is a suitable polymeric support for catalysts and reagents (Merrifield, 1963; Fréchet, 1981). Styrene and divinylbenzene (DVB) are among the monomers that are most often employed to prepare solid polymeric matrixes. Polystyreneco-divinylbenzene (PS-DVB), an insoluble copolymer, results from styrene polymerization in the presence of varied amounts of DVB. The characteristics of this copolymer depend on the quantity of DVB present in the material (Kapura and Gates, 1973; Xia et al., 2012). In many cases, the success of a heterogeneous catalyst relies on the features of the polymeric material. Bergbreiter (2002) proposed that some physicochemical properties should be considered when choosing the catalyst support, including the catalytic activity, surface area, porosity, and thermal and mechanical stability of the material in the conditions of the catalyzed reaction (Bergbreiter, 2002; Chouhan and Sarma, 2011). In the field of polymer chemistry, the term resin is indistinctly employed to describe polymers with and without cross-links (Akelah and Sherrington, 1981; Sharma, 1995). Alternatively, in heterogeneous catalysis, the term resin refers to species consisting of long polymeric chains interconnected via cross-links, the so-called polymeric matrix. Polymeric matrixes are tridimensional, insoluble, and porous; their ability to exchange ions arises from the introduction of adequate functional groups into the polymeric support (Kunin et al., 1962; Akelah and Sherrington, 1981; Fréchet, 1981; Harmer and Sun, 2001; Hart et al., 2002). Polymeric matrix functionalization can be achieved in two ways: (1) monomers containing the desired functional groups (or precursors of this functional group) can be directly polymerized; (2) the polymeric support can be prepared first, and the functional group is introduced by chemical modification of the polymeric support (Molinari et al., 1979; Kucera and Jancar, 1998; Harmer and Sun, 2001). Coutinho and Rezende (2001) and Coutinho et al. (2004a) showed that supported species can be prepared by chemically modifying the copolymer base (polymeric support). These authors reported the sulfonation of a support consisting of PS reticulated with DVB (Figure 16.7). The aromatic rings on the insoluble PS-DVB copolymer react with concentrated sulfuric acid in the presence of 1,2-dichloroethane; the latter compound expands the polymeric matrix and allows sulfonation of the internal surface as well. Most of the functional groups introduced into polymeric matrices concentrate inside the resin beads (Coutinho and Rezende, 2001). Cationic resins can be used as an option for catalytic reactions involving mineral or sulfonic acids. In the presence of water, the cationic groups on the polymer display different acidity constants, as in the case of compounds with low molecular mass. The catalytic performance of an ion exchange resin is associated with the concentration of functional groups and the physicochemical properties of the support. Therefore, compared with homogeneous catalysts, different factors affect the activity of resins. The use of ion exchange resins as catalysts has many advantages: (1) despite being equivalent to strong SO3H CH2 – CH CH2 – CH CH2 CH2 – CH CH2 – CH CH2 + H 2SO4 C H2 C H CH2 – CH CH 2 SO3H ClCH2CH2Cl 94ºC/3,5 h C H2 C H CH2 – CH CH2 SO3H FIGURE 16.7 Sulfonation of a polymeric matrix consisting of polystyrene and divinylbenzene.
271 POLYMERIC CATALYSTS FIGURE 16.8 mineral acids, resins are less oxidizing and corrosive, since most of the catalytic sites are located inside the beadsdtherefore, they do not pose any hazards to the operator and are easy to store; (2) resins behave as selective catalysts and enable reaction control; (3) catalysts with high purity are recovered at the end of the reaction by simple filtration; (4) resins do not require neutralization before being separated from the reaction medium, a step that usually reduces product yield; (5) resins eliminate the steps and equipment necessary to separate the catalyst and purify the product, simplifying continuous or batch procedures based on ion exchange resins; and (6) if the resins undergo deactivation due to contamination or prolonged use, they can be reactivated via a simple procedure that does not release hazardous gases into the atmosphere (Saha and Sharma, 1996; Coutinho and Rezende, 2001; Harmer and Sun, 2001; Marquardt and Eifler-Lima, 2001; Mitsutani, 2002; Coutinho et al., 2003, 2004a,b; Kiss et al., 2006). The main drawback of ion exchange resins is that their maximum operation temperature is relatively low. Literature suggests that they should be used below 125  C to ensure long catalyst lifetime (John and Israelstam, 1960; Akelah and Sherrington, 1981; Giménez et al., 1987; Coutinho and Rezende, 2001; Rezende et al., 2008). Aromatic compounds are easy to functionalize, especially if they contain acid groups like sulfonic acids. The sulfonation of organic compounds containing benzene rings, including polymers, has been extensively reported (Ma and Hanna, 1999; Coutinho and Rezende, 2001; Coutinho et al., 2003, 2004a,b, 2006; Rezende et al., 2008; Soldi et al., 2009). Figure 16.8 represents the sulfonation PS (Soldi et al., 2009). Sulfonation significantly modifies the physical properties of linear PS, especially the polarity. Hence, sulfonated PS should remain insoluble during biodiesel production. Soldi et al. (2009) studied methods to sulfonate linear PS and applied the resulting sulfonated material as heterogeneous polymeric catalyst to produce soybean methyl esters. Raw materials with different moisture degrees and the effect of different variables on the conversion rate have been investigated; biodiesel production from soybean oil and beef tallow led to significantly improved yields. Recently, much interest has been taken in utilizing low-cost plant oil and fat containing a large amount of PS sulfonation with acetyl sulfate. FFAs. However, oils with high FFA content are difficult to transesterify using the commercially available alkaline catalyst (Zhang et al., 2003; Tesser et al., 2005; Marchetti and Errazu, 2008; Sharma et al., 2008; Liu et al., 2009; Tesser et al., 2010; Chouhan and Sarma, 2011; Shahid and Jamal, 2011; Borges and Dı́az, 2012). Canakci and Van Gerpen (1999, 2001) found the transesterification would not occur if the FFA content in the oil was beyond 3%. According to the research paper by Kouzu et al. (2011), the promising approach is to esterify FFA into FAMEs with the help of the solid acid catalyst, and there were some research papers studying utilization of several types of heterogeneous catalysts including sulfonated cation exchange resin (Russnueldt and Hoelderich, 2009; Tesser et al., 2010; Kouzu et al., 2011). With respect to utilization of the sulfonated resin for the preesterification of FFA, some researchers focused their attention on the macroreticular type but the use of two types of resins (macroreticular and gelular types) were also studied by other authors (Ramadhas et al., 2005; Soldi et al., 2009; Lam et al., 2010; Melero et al., 2010; Kouzu et al., 2011; Semwal et al., 2011; Li et al., 2012; Xia et al., 2012; Zhang et al., 2012a,b). Feng et al. (2011) investigated the continuous esterification of FFAs from acidified oil with methanol by cation exchange resin in a fixed bed reactor to prepare biodiesel and the operational stability of continuous esterification by resin in the fixed bed reactor was also conducted (McNeff et al., 2008; Shibasaki-Kitagawa et al., 2010; Feng et al., 2011; Cheng et al., 2012). According to Feng et al. (2010), from the viewpoint of cost savings, the use of cation exchange resins in heterogeneous catalytic processes may be advantageous over enzymes and supercritical methanol (Feng et al., 2010). These resins are composed of copolymers of DVB and styrene containing sulfonic acid groups attached to benzene rings and these are the active sites for esterification and transesterification (Marchetti and Errazu, 2008; Rezende et al., 2008; Russnueldt and Hoelderich, 2009; Feng et al., 2010; Kouzu et al., 2011). However, other sulfonated polymeric backbones such as AmberlystÒ , DowexÒ and NafionÒ , a perfluorinated ion exchange resin, all of them having a very strong Brønsted acid character, have also been used in these type of reactions (Özbay et al., 2008; Talukder et al., 2009; Feng et al., 2010; Park et al., 2010; Galia et al., 2011; Yin et al., 2012; Zhang et al., 2012a). In general, cation exchange resins are
272 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL preferable for esterification (Giménez et al., 1987; Chen et al., 1999; Coutinho et al., 2004b; Coutinho et al., 2006; Grob and Hasse, 2006), while anionic resins may be applied for transesterification of oils and fats (Shibasaki-Kitakawa et al., 2007; Ren et al., 2012). CONCLUDING REMARKS Truly heterogeneous catalytic processes are attractive for many practical applications due to their recyclability, structural stability, high selectivity and good catalytic performance. However, all these properties are hardly achievable in a single catalytic system. In most cases, leachable catalytic species migrate to the reaction environment, causing a partial contamination of the final product as well as a loss in catalytic activity when the solids are applied in several consecutive reaction cycles. Moreover, in many situations found in the literature, the heterogeneity of catalytic systems is not approached with proper analytical methods, resulting in wrong conclusions and/or classification of the proposed solid catalyst. These usually arise from poor data on catalyst recovery and reuse, poor characterization of the catalyst structure and high leaching levels of catalytic species. Also, in many cases, no attempt is made to fully characterize these properties and solids are classified as heterogeneous catalysts just because they are partially filterable after reaction completion. One of such flaws was nicely demonstrated by Silva et al. (2013) using bismuth-containing mixed oxides. Apart from these observations, the lack of suitable reaction controls such as in the case of TC in esterification reactions reveal unrealistic catalytic performance in reactions that are known to be autocatalytic under appropriate experimental conditions. Nevertheless, a number of rather attractive heterogeneous catalytic systems have been discovered so far for biodiesel applications, probably due to the wide scope of catalytic properties that are influential in both esterification and transesterification. However, many of these will never be able to reach industrial applications because their benchmarking was never strong enough to support further investments at large scale. References Abreu, F.R., Lima, D.G., Hamú, E.H., Wolf, C., Suarez, P.A.Z., 2004. Utilization of metal complexes as catalysts in the transesterification of Brazilian vegetable oils with different alcohols. J. Mol. Catal. A: Chem. 209, 29e33. Ajaikumar, S., Pandurangan, A., 2007. Esterification of alkyl acids with alkanols over MCM-41 molecular sieves: influence of hydrophobic surface on condensation reaction. J. Mol. Catal. A: Chem. 266, 1e10. Akelah, A., Sherrington, D., 1981. Application of functionalized polymers in organic synthesis. Chem. Rev. 81, 557e587. Albuquerque, M.C.G., Santamarı́a-González, J., Mérida-Robles, J.M., Moreno-Tost, R., Rodrı́guez-Castellón, E., Jiménez-López, A., Azevedo, D.C.S., Cavalcante Jr., C.L., Maireles-Torres, P., 2008. MgM (M ¼ Al and Ca) oxides as basic catalysts in transesterification processes. Appl. Catal. A: Gen. 347, 162e168. Altiparmak, D., Keskin, A., Koca, A., Gürü, M., 2007. Alternative fuel properties of tall oil fatty acid methyl esterediesel fuel blends. Bioresour. Technol. 98, 241e246. Antunes, W.M., Veloso, C.O., Henriques, C.A., 2008. Transesterification of soybean oil with methanol catalyzed by basic solids. Catal. Today 133e135, 548e554. Arizaga, G.G.C., Satyanarayana, K.G., Wypych, F., 2007. Layered hydroxide salts: synthesis, properties and potential applications. Solid State Ionics 178, 1143e1162. Arizaga, G.G.C., Mangrich, A.S., Gardolinski, J.E.F.C., Wypych, F., 2008. Chemical modification of zinc hydroxide nitrate and Zn/ Al-layered double hydroxide with dicarboxylic acids. J. Colloid Interface Sci. 320, 68e176. Bail, A., Santos, V.C., Freitas, M.R., Ramos, L.P., Schreiner, W.H., Ricci, G.P., Ciuffi, K.J., Nakagaki, S., 2013. Investigation of a molybdenum-containing silica catalyst synthesized by the sol-gel process in heterogeneous catalytic esterification reactions using methanol and ethanol. Appl. Catal. B: Environ. 130e131, 314e324. Barakos, N., Pasias, S., Papayannakos, N., 2008. Transesterification of triglycerides in high and low quality oil feeds over an HT2 hydrotalcite catalyst. Bioresour. Technol. 99, 5037e5042. Barman, S., Vasudevan, S., 2006. Melting of saturated fatty acid soaps. J. Phys. Chem. B 110, 22407e22414. Beck, J.S., Vartuli, J.C., Roth, W.J., Leonowicz, M.E., Kresge, C.T., Schmitt, K.D., Chu, C.T.W., Olson, D.H., Sheppard, E.W., 1992. A new family of mesoporous molecular sieves prepared with liquid crystal templates. J. Am. Chem. Soc. 114, 10834e10843. Bergbreiter, D.E., 2002. Using soluble polymers to recover catalysts and ligands. Chem. Rev. 102, 3345e3384. Bokade, V.V., Yadav, G.D., 2009. Transesterification of edible and nonedible vegetable oils with alcohols over heteropolyacids supported on acid-treated clay. Ind. Eng. Chem. Res. 48, 9408e9415. Bondioli, P., Gasparolia, A., Fedeli, E., Veronese, S., Sala, M., 1995. Storage stability of biodiesel. J. Am. Oil Chem. Soc. 72, 669e702. Borges, M.E., Dı́az, L., 2012. Recent developments on heterogeneous catalysts for biodiesel production by oil esterification and transesterification reactions: a review. Renewable Sustainable Energy Rev. 16, 2839e2849. Bravo-Suárez, J.J., Páez-Mozo, E.A., Oyama, S.T., 2004. Review of the synthesis of layered double hydroxides: a thermodynamic approach. Quim. Nova 27, 601e614. Brindley, G.W., Brown, G., 1980. Crystal structures of clay minerals and their X-ray identification. Min. Soc. Lond. 2e115. Caetano, C.S., Fonseca, I.M., Ramos, A.M., Vital, J., Castanheiro, J.E., 2008. Esterification of free fatty acids with methanol using heteropolyacids immobilized on silica. Catal. Commun. 9, 1996e1999. Canakci, M., Gerpen, J.V., 1999. Biodiesel production via acid catalysis. Trans. Am. Soc. Agric. Eng. 42, 1203e1210. Canakci, M., Gerpen, J.V., 2001. Biodiesel production from oils and fats with high free fatty acids. Trans. Am. Soc. Agric. Eng. 44, 1429e1436. Cantrell, D.G., Gillie, L.J., Lee, A.F., Wilson, K., 2005. Structurereactivity correlations in MgAl hydrotalcite catalysts for biodiesel synthesis. Appl. Catal. A: Chem. 287, 183e190. Carlino, S., 1997. The intercalation of carboxylic acids into layered double hydroxides: a critical evaluation and review of the different methods. Solid State Ionics 98, 73e84.
REFERENCES Carmo Jr., A.C., Souza, L.K.C., Costa, C.E.F., Longo, E., Zamian, J.R., Rocha Filho, G.N., 2009. Production of biodiesel by esterification of palmitic acid over mesoporous aluminosilicate Al-MCM-41. Fuel 88, 461e468. Centi, G., Perathoner, S., 2008. Catalysis by layered materials: a review. Microporous Mesoporous Mater. 107, 3e15. Chen, X., Xu, Z., Okuhara, T., 1999. Liquid-phase esterification of acrylic acid with 1-butanol catalyzed by solid acid catalysis. Appl. Catal. A: Gen. 180, 261e269. Cheng, Y., Feng, Y., Ren, Y., Liu, X., Gao, A., He, B., Yan, F., Li, J., 2012. Comprehensive kinetic studies of acidic oil continuous esterification by cation-exchange resin in fixed bed reactors. Bioresour. Technol. 113, 65e72. Cho, Y.B., Seo, G., 2010. High activity of acid-treated quail eggshell catalyst in the transesterification of palm oil with methanol. Bioresour. Technol. 101, 8515e8519. Choudary, B.M., Kantam, M.L., Reddy, Ch.V., Aranganathan, S., Santhi, P.L., Figueras, F., 2000. Mg-Al-O-t-Bu hydrotalcite: a new and efficient heterogeneous catalyst for transesterification. J. Mol. Catal. A: Chem. 159, 411e416. Chouhan, A.P.S., Sarma, A.K., 2011. Modern heterogeneous catalysts for biodiesel production: a comprehensive review. Renewable Sustainable Energy Rev. 15, 4378e4399. Climent, M.J., Corma, A., Iborra, S., Miguel, S., Primo, J., Rey, F., 1999. Mesoporous materials as catalysts for the production of chemicals: synthesis of alkyl glucosides on MCM-41. J. Catal 183, 76e82. Cordeiro, C.S., Arizaga, G.G.C., Ramos, L.P., Wypych, F., 2008. A new zinc hydroxide nitrate heterogeneous catalyst for the esterification of free fatty acids and the transesterification of vegetable oils. Catal. Commun. 9, 2140e2143. Cordeiro, C.S., Silva, F.R., Wypych, F., Ramos, L.P., 2011. Catalisadores Heterogêneos para a Produção de Monoésteres Graxos (biodiesel). Quim. Nova 34, 477e486. Cordeiro, C.S., Silva, F.R., Marangoni, R., Wypych, F., Ramos, L.P., 2012. LDHs instability in esterification reactions and their conversion to catalytically active layered carboxylates. Catal. Lett. 142, 763e770. Corma, A., Garcia, H., Iborra, S., Primo, J., 1989. Modified faujasite zeolites as catalysts in organic reactions: esterification of carboxylic acids in the presence of HY zeolites. J. Catal. 120, 78e87. Corma, A., Hamid, S.B.A., Iborra, S., Velty, A., 2005. Lewis and Brönsted basic active sites on solid catalysts and their role in the synthesis of monoglycerides. J. Catal. 234, 340e347. Coutinho, F.M.B., Rezende, S.M., 2001. Catalisadores sulfônicos imobilizados em polı́meros: Sı́ntese, caracterização e avaliação. Polı́meros 11, 222e233. Coutinho, F.M.B., Aponte, M.L., Barbosa, C.C.R., Costa, V.G., Lachter, E.R., Tabak, D., 2003. Resinas sulfônicas: Sı́ntese, caracterização e avaliação em reações de alquilação. Polı́meros 13, 141e146. Coutinho, F.M.B., Cunha, L., Gomes, A.S., 2004a. Suportes poliméricos para catalisadores sulfônicos: Sı́ntese e caracterização. Polı́meros 14, 31e37. Coutinho, F.M.B., Souza, R.R., Gomes, A.S., 2004b. Synthesis, characterization and evaluation of sulfonic resins as catalysts. Eur. Polym. J. 40, 1525e1532. Coutinho, F.M.B., Rezende, S.M., Soares, B.G., 2006. Characterization of sulfonated poly(styrene-divinylbenzene) and poly(divinylbenzene) and its application as catalysts in esterification reaction. J. Appl. Polym. Sci. 102, 3616e3627. Crepaldi, E.L., Valim, J.B., 1998. Hidróxidos duplos lamelares: sı́ntese, estrutura, propriedades e aplicações. Quim. Nova 21, 300e311. Csicsery, S.M., 1984. Shape-selective catalysis in zeolites. Zeolites 4, 202e213. 273 Cundy, C.S., Cox, P.A., 2003. The hydrothermal synthesis of zeolites: history and development from the earliest days to the present time. Chem. Rev. 103, 663e701. Di Serio, M., Tesser, R., Pengmei, L., Santacesaria, E., 2008. Heterogeneous catalysts for biodiesel production. Energy Fuels 22, 207e217. Drago, R.S., Dias, J.A., Maier, T.O., 1997. An acidity scale for Brønsted acids including H3PW12O40. J. Am. Chem. Soc. 119, 7702e7710. Feng, Y., He, B., Cao, Y., Li, J., Liu, M., Yan, F., Liang, X., 2010. Biodiesel production using cation-exchange resin as heterogeneous catalyst. Bioresour. Technol. 101, 1518e1521. Feng, Y., Zhang, A., Li, J., He, B., 2011. A continuous process for biodiesel production in a fixed bed reactor packed with cationexchange resin as heterogeneous catalyst. Bioresour. Technol. 102, 3607e3609. Figueiredo, J.L., Ribeiro, F.R., 1987. Catálise Heterogénea, first ed. Fundação Calouste Gulbenkian, Portugal. p. 1. Fjerbaek, L., Christensen, K.V., Norddahl, B., 2009. A review of the current state of biodiesel production using enzymatic transesterification. Biotechnol. Bioeng. 102, 1298e1315. Fréchet, J.M.J., 1981. Synthesis and applications of organic polymers as supports and protecting groups. Tetrahedron 37, 663e683. Fukuda, H., Kondo, A., Noda, A., 2001. Biodiesel fuel production by transesterification of oil. J. Biosci. Bioeng. 92, 405e416. Galia, A., Scialdone, O., Tortorici, E., 2011. Transesterification of rapeseed oil over acid resins promoted by supercritical carbon dioxide. J. Supercrit. Fluids 56, 186e193. Giménez, J., Costa, J., Cervera, S., 1987. Vapor-phase esterification of aceticacid with ethanol catalyzed by a macroporous sulfonated styrenedivinilbenzene (20%) resin. Ind. Eng. Chem. Res. 26, 198e2002. Giri, B.Y., Rao, K.N., Prabhavathi Devi, B.L.A., Lingaiah, N., Suryanarayana, I., Prasad, R.B.N., Sai Prasad, P.S., 2005. Esterification of palmitic acid on the ammonium salt of 12-tungstophosphoric acid: the influence of partial proton exchange on the activity of the catalyst. Catal. Commun. 6, 788e792. Grob, S., Hasse, H., 2006. Reaction kinetics of the homogeneously catalyzed esterification of 1-butanol with acetic acid in a wide range of initial compositions. Ind. Eng. Chem. Res. 45, 1869e1874. Guerreiro, L., Castanheiro, J.E., Fonseca, I.M., Martin-Aranda, R.M., Vital, J., 2006. Transesterification of soybean oil over sulfonic acid functionalized polymer membranes. Catal. Today 118, 166e171. Guerreiro, L., Pereira, P.M., Fonseca, I.M., Martin-Aranda, R.M., Ramos, A.M., Dias, J.M.L., Oliveira, R., Vital, J., 2010. PVA embedded hydrotalcite membranes as basic catalysts for biodiesel synthesis by soybean oil methanolysis. Catal. Today 156, 191e197. Harmer, M.A., Sun, Q., 2001. Solid acid catalysis using ion-exchange resins. Appl. Catal. A: Gen. 221, 45e62. Hart, M., Fuller, G., Brown, D.R., Dale, J.A., Plant, S., 2002. Sulfonated poly(styrene-co-divinilbenzene) ion-exchange resins: acidities and catalystic activities in aqueous reactions. J. Mol. Catal. A: Chem. 182e183, 439e445. Hayyan, A., Mjalli, F.S., Hashim, M.A., Hayyan, M., AlNashef, I.M., Al-Zahrani, S.M., Al-Saadi, M.A., 2011. Ethanesulfonic acid-based esterification of industrial acidic crude palm oil for biodiesel production. Bioresour. Technol. 102, 9564e9570. Helwani, Z., Othman, M.R., Aziz, N., Kim, J., Fernando, W.J.N., 2009a. Solid heterogeneous catalysts for transesterification of triglycerides with methanol: a review. Appl. Catal. A: Gen. 363, 1e10. Helwani, Z., Othman, M.R., Aziz, N., Fernando, W.J.N., Kim, J., 2009b. Technologies for production of biodiesel focusing on green catalytic techniques: a review. Fuel Process. Technol. 90, 1502e1514. Hu, C., Li, D., 2004. Polyoxometalate complexes of layered double hydroxides. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay Surfaces: Fundamentals and Applications, Interface Science and Technology, first ed., vol. 1. Elsevier Academic Press, pp. 389e390.
274 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL Ilgen, O., Dinçer, I., Yildiz, M., Alptekin, E., Boz, N., Canakci, M., Akin, A.N., 2007. Investigation of biodiesel production from canola oil using Mg/Al hydrotalcite catalysts. Turk. J. Chem. 31, 509e514. Jacobson, K., Gopinath, R., Meher, L.C., Dalai, A.K., 2008. Solid acid catalyzed biodiesel production from waste cooking oil. Appl. Catal. B: Environ. 85, 86e91. John, E.V.O., Israelstam, S.S., 1960. Use of cation exchange resin in organic reaction. J. Org. Chem. 26, 240. Kapura, J.M., Gates, B.C., 1973. Sulfonated polymers as alkylation catalysts. Ind. Eng. Chem. Prod. Res. Dev. 12, 62e67. Kusdiana, D., Saka, S., 2001. Fuel 80, 693e698. Kiss, A.A., Dimin, A.C., Rothenberg, G., 2006. Solid acid catalysts for biodiesel productiondtowards sustainable energy. Adv. Synth. Catal. 348, 75e81. Knothe, G., Krahl, J., Gerpen, J.V., Ramos, L.P. (Eds.), 2006. Manual de biodiesel. Edgard Blücher, São Paulo, pp. 1e352. Konwar, D., Gogoi, P.K., Gogoi, P., Borah, G., Baruah, R., Hazarika, N., Borgohain, R., 2008. Esterification of carboxylic acids by acid activated kaolinite clay. Indian J. Chem. Technol. 15, 75e78. Kouzu, M., Nakagaito, A., Hidaka, J., 2011. Pre-esterification of FFA in plant oil transesterified into biodiesel with the help of solid acid catalysis of sulfonated cation-exchange resin. Appl. Catal. A: Gen. 405, 36e44. Kucera, F., Jancar, J., 1998. Homogenous and heterogeneous sulfonation of polymers: a review. Polym. Eng. Sci. 38, 1e12. Kunin, R., Meitzner, E., Bortinick, N., 1962. Macroreticular ion exchange resins. J. Am. Chem. Soc. 84, 305e306. Lacouture, F., Peultier, J., Francois, M., Steinmetz, J., 2000. Anhydrous polymeric zinc(II) octanoate. Cryst. Struct. Commun. 56, 556e557. Lam, M.K., Lee, K.T., Mohamed, A.R., 2010. Homogeneous, heterogeneous and enzymatic catalysis for transesterification of high free fatty acid oil (waste cooking oil) to biodiesel: a review. Biotechnol. Adv. 28, 500e518. Lee, J.S., Saka, S., 2010. Biodiesel production by heterogeneous catalysts and supercritical technologies. Bioresour. Technol. 101, 7191e7200. Leng, Y., Wang, J., Zhu, D.R., Wu, Y.J., Zhao, P.P., 2009. Sulfonated organic heteropolyacid salts: recyclable green solid catalysts for esterifications. J. Mol. Catal. A: Chem. 313, 1e6. Li, G., Ding, Y., Wang, J., Wang, X., Suoa, Jishuan, 2007. New progress of Keggin and WellseDawson type polyoxometalates catalyze acid and oxidative reactions. J. Mol. Cat. A: Chem. 262, 67e76. Li, J., Fu, Y.-J., Qu, X.-J., Wang, W., Luo, M., Zhao, C.-J., Zu, Y.-G., 2012. Biodiesel production from yellow horn (Xanthoceras sorbifolia Bunge) seed oil using ion exchange resin as heterogeneous catalyst. Bioresour. Technol. 108, 112e118. Lisboa, F.S., Gardolinski, J.E.F.C., Cordeiro, C.S., Wypych, F., 2012. Layered metal laurates as active catalysts in the Methyl/Ethyl esterification reactions of lauric acid. J. Braz. Chem. Soc. 23, 46e56. Liu, Y., Lotero, E., Goodwin Jr., J.G., MO, X., 2007. Transesterification of poultry fat with methanol using MgeAl hydrotalcite derived catalysts. Appl. Catal. A: Gen. 331, 138e148. Liu, Y., Wang, L., Yan, Y., 2009. Biodiesel synthesis combining preesterification with alkali catalyzed process from rapeseed oil deodorizer distillate. Fuel Process. Technol. 90, 857e862. Lopez, D.E., Suwannakarn, K., Bruce, D.A., Goodwin Jr, J.G., 2007. Esterification and transesterification on tungstated zirconia: effect of calcination temperature. J. Catal. 247, 43e50. Lopez-Granados, M., Alba-Rubio, A.C., Vila, F., Martı́n, A.D., Marisacl, R., 2010. Surface chemical promotion of Ca oxide catalyst in biodiesel production reaction by the addition of monoglycerides, diglycerides and glycerol. J. Catal. 276, 229e236. Lotero, E., Liu, Y., Lopez, D.E., Suwannakarn, K., Bruce, D.A., Goodwin, J.G., 2005. Synthesis of biodiesel via acid catalysis. Ind. Eng. Chem. Res. 44, 5353e5363. Luque, R., Campelo, J.M., Luna, D., Marinas, J.M., Romero, A.A., 2007. Catalytic performance of Al-MCM-41 materials in the N-alkylation of aniline. J. Mol. Catal. A: Chem. 269, 190e196. Ma, F., Hanna, M.A., 1999. Biodiesel production: a review. Bioresour. Technol. 70, 1e15. Ma, Y., Wang, Q.L., Yan, H., Ji, X., Qiu, Q., 1996. Zeolite-catalyzed esterification I. Synthesis of acetates, benzoates and phthalates. Appl. Catal. A: Gen. 139, 51e57. Macedo, C.C.S., Abreu, F.R., Tavares, A.P., Alves, M.B., Zara, L.F., Rubim, J.C., Suarez, P.A.Z., 2006. New heterogeneous metal-oxides based catalyst for vegetable oil trans-esterification. J. Braz. Chem. Soc. 17, 1291e1296. Marchetti, J.M., Errazu, A.F., 2008. Comparison of different heterogeneous catalysts and different alcohols for the esterification reaction of oleic acid. Fuel 87, 3477e3480. Marquardt, M., Eifler-Lima, V.L., 2001. A sı́ntese em fase sólida e seus suportes poliméricos mais empregados. Quim. Nova 24, 846e855. Martinez, C., Corma, A., 2011. Inorganic molecular sieves: preparation, modification and industrial application in catalytic processes. Coord. Chem. Rev. 255, 1558e1580. McNeff, V.C., McNeff, C.L., Yan, B., Nowlan, D.T., Rasmussen, M., Gyberg, A.E., Krohn, B.J., Fedie, R.L., Hoye, T.R., 2008. A continuous catalytic system for biodiesel production. Appl. Catal. A: Gen. 343, 39e48. Meher, L.C., Vidya, D., Naik, S.N., 2006. Technical aspects of biodiesel production by transesterificationda review. Renewable Sustainable Energy Rev. 10, 248e268. Melero, J.A., Bautista, L.F., Morales, G., Iglesias, J., SánchezVázquez, R., 2010. Biodiesel production from crude palm oil using sulfonic acid-modified mesostructured catalysts. Chem. Eng. J. 161, 323e331. Merrifield, R.B., 1963. Solid phase synthesis. I. The synthesis of a tetrapeptide. J. Am. Chem. Soc. 85, 2149e2154. Mitsutani, A., 2002. Future possibilities of recently commercialized acid/base-catalyzed chemical processes. Catal. Today 73, 57e63. Mittelbach, Tritthart, M.P., 1988. Diesel fuel derived from vegetable oils, III. Emission tests using methyl ester of used frying oil. J. Am. Oil Chem. Soc. 65, 1185e1187. Mizuno, N., Misono, M., 1998. Heterogeneous catalysis. Chem. Rev. 98, 199e217. Molinari, H., Montanari, F., Quici, S., Tundo, P., 1979. Polymersupported phase-transfer catalysts. High catalytic activity of ammonium and phosphonium quaternary salts bonded to a polystyrene matrix. J. Am. Chem. Soc. 1001, 3920e3927. Moronta, A., 2004. In: Wypych, F., Satyanarayana, K.G. (Eds.), Clay SurfacesdFundamentals and Applications. Elsevier, pp. 321e344. Nakagaki, S., Bail, A., Santos, V.C., Souza, V.H.R., Vrubel, H., Nunes, F.S., Ramos, L.P., 2008. Use of anhydrous sodium molybdate as an efficient heterogeneous catalyst for soybean oil methanolysis. Appl. Catal. A: Gen. 351, 267e274. Nascimento, L.A.S., Tito, L.M.Z., Angelica, R.S., Costa, C.E.F., Zamian, J.R., Rocha, G.N., 2011. Esterification of oleic acid over solid acid catalysts prepared from Amazon flint kaolin. Appl. Catal. B: Environ. 101, 495e503. Neji, S.B., Trabelsi, M., Frikha, M.H., 2011. Esterification of fatty acid over Tunisian acid activated clay: kinetic study. J. Oleo Sci. 60, 293e299. Ngamcharussrivichai, C., Wiwatnimit, W., Wangnoi, S., 2007. Modified dolomites as catalysts for palm kernel oil transesterification. J. Mol. Cat. A: Chem. 276, 24e33. Nijhuis, T.A., Beers, A.E.W., Kapteijn, F., Moulijn, J.A., 2002. Water removal by reactive stripping for a solid-acid catalyzed esterification in a monolithic reactor. Chem. Eng. Sci. 57, 1627e1632.
REFERENCES Oliveira, C.F., Dezaneti, L.M., Garcia, F.A.C., de Macedo, J.L., Dias, J.A., Dias, S.C.L., Alvim, K.S.P., 2010. Esterification of oleic acid with ethanol by 12-tungstophosphoric acid supported on zirconia. Appl. Catal. A: Gen. 372, 153e161. Olowokere, J.A., Okafor, J.O., Mbah, G.O., 2012. Comparative studies on the catalytic esterification of butanol with ethanoic acid by some nigerian montmorillonite clays. Int. J. Environ. Sci. Manage. Eng. Res. 1, 147e159. Özbay, N., Oktar, N., Tapan, A.N., 2008. Esterification of free fatty acids in waste cooking oils (WCO): role of ion-exchange resins. Fuel 87, 1789e1798. Park, J.Y., Kim, D.K., Lee, J.S., 2010. Esterification of free fatty acids using water-tolerable Amberlyst as a heterogeneous catalyst. Bioresour. Technol. 101, S62eS65. Philippou, A., Anderson, M.W., 2000. Aldol-Type Reactions over Basic Microporous Titanosilicate ETS-10 Type Catalysts. J. Catal 189, 395e400. Ramadhas, A.S., Jayaraj, S., Muraleedharan, C., 2005. Biodiesel production from high FFA rubber seed oil. Fuel 84, 335e340. Ramos, L.P., Kucek, K.T., Domingos, A.K., Wilhelm, H.M., 2003. Biodiesel: Um projeto de sustentabilidade econômica e sócio-ambiental para o Brasil. Biotecnologia Cienc. Desenvolv. 31, 28e37. Reinoso, D.M., Damiani, D.E., Tonetto, G.M., 2012. Zinc carboxylic salts used as catalyst in the biodiesel synthesis by esterification and transesterification: study of the stability in the reaction medium. Appl. Catal. A: Gen. 449, 88e95. Ren, Y., He, B., Yan, F., Wang, H., Cheng, Y., Lin, L., Feng, Y., Li, J., 2012. Continuous biodiesel production in a fixed bed reactor packed with anion-exchange resin as heterogeneous catalyst. Bioresour. Technol. 113, 19e22. Rezende, S.M., Reis, M.C., Reid, M.G., Silva Jr., P.L., Coutinho, F.M.B., Gil, R.A.S.S., Lachter, E.R., 2008. Transesterification of vegetable oils promoted by poly(styrene-divinylbenzene) and poly(divinylbenzene). Appl. Catal. A: Gen. 349, 198e203. Rezende, M.J.C., Pereira, M.S.C., Santos, G.F.N., Aroeira, G.O.P., Albuquerque, T.C., Suarez, P.A.Z., Pinto, A.C., 2012. Preparation, characterisation and evaluation of Brazilian clay-based catalysts for use in esterification reactions. J. Braz. Chem. Soc. 23, 1209e1215. Rinaldi, R., Schuth, F., 2009. Design of solid catalysts for the conversion of biomass. Energy Environ. Sci. 2, 610e626. Russnueldt, B.M.E., Hoelderich, W.F., 2009. New sulfonic acid ionexchange resins for the preesterification of different oils and fats with high content of free fatty acids. Appl. Catal. A: Gen. 362, 47e57. Ryland, L.B., Tamale, M.W., Wilson, J.N., 1960. In: Emmett, P.H. (Ed.), 1960. Catalysis, vol. 7, Reinhold, New York, pp. 1e91. Saha, B., Sharma, M.M., 1996. Esterification of formic acid, acrylic acid and methacrylic acid with cyclohexene in bath and distillation column reactors: ion-exchange resins as catalysts. React. Funct. Polym. 28, 263e278. Santos, V.C., Bail, A., Okada, H.O., Ramos, L.P., Ciuffi, K.J., Lima, O.J., Nakagaki, S., 2011. Methanolysis of soybean oil using tungstencontaining, heterogeneous catalysts. Energy Fuels 25, 2794e2802. Schuchardt, U., Sercheli, R., Vargas, R.M., 1998. Transesterification of vegetable oils: a review. J. Braz. Chem. Soc. 9, 199e210. Semwal, S., Arora, A.K., Badoni, R.P., Tuli, D.K., 2011. Biodiesel production using heterogeneous catalysts. Bioresour. Technol. 102, 2151e2161. Serio, M., Ledda, M., Cozzolino, M., Minutillo, G., Tesser, R., Santacesaria, E., 2006. Transesterification of soybean oil to biodiesel by using heterogeneous basic catalysts. Ind. Eng. Chem. Res. 45, 3009e3014. Serio, M., Cozzolino, M., Giordano, M., Tesser, R., Patrono, P., Santacesaria, E., 2007. From homogeneous to heterogeneous catalysts in biodiesel production. Ind. Eng. Chem. Res. 46, 6379e6384. 275 Shahid, E.M., Jamal, Y., 2011. Production of biodiesel: a technical review. Renewable Sustainable Energy Rev. 15, 4732e4745. Sharma, M.M., 1995. Some novel aspects of cationic ion-exchange resins as catalysts. React. Funct. Polym. 26, 3e23. Sharma, Y.C., Singh, B., Upadhyay, S.N., 2008. Advancements in development and characterization of biodiesel: a review. Fuel 87, 2355e2373. Sharma, Y.C., Singh, B., Korstad, J., 2011. Latest developments on application of heterogeneous basic catalysts for an efficient and ecofriendly synthesis of biodiesel: a review. Fuel 90, 1309e1324. Shibasaki-Kitagawa, N., Tsuji, T., Chida, K., Kubo, M., Yonemoto, T., 2010. Simple continuous production process of biodiesel fuel from oil with high content of free fatty acid using ion-exchange resin catalysts. Energy Fuels 24, 3634e3638. Shibasaki-Kitakawa, N., Honda, H., Kuribayashi, H., Toda, T., Fukumura, T., Yonemoto, T., 2007. Biodiesel production using anionic ion-exchange resin as heterogeneous catalyst. Bioresour. Technol. 98, 416e421. Shumaker, J.L., Crofcheck, C., Tackett, S.A., Santillan-jimenez, E., Morgan, T., Ji, Y., Crocker, M., Toops, T., 2008. Biodiesel synthesis using calcined layered double hydroxide catalysts. Appl. Catal. A: Gen. 82, 120e130. Silva, F.R., Silveira, M.H.L., Cordeiro, C.S., Nakagaki, S., Wypych, F., Ramos, L.P., 2013. Esterification of Fatty Acids Using a BismuthContaining Solid Acid Catalyst. Energy Fuels 27, 2218e2225. Soldi, R.A., Oliveira, A.R.S., Ramos, L.P., César-Oliveira, M.A.F., 2009. Soybean oil and beef tallow alcoholysis by acid heterogeneous catalysis. Appl. Catal. A: Gen. 361, 42e48. Stählin, W., Oswald, H.R., 1971. The infrared spectrum and thermal analysis of zinc hydroxide nitrate. J. Solid State Chem. 3, 252e255. Suppes, G.J., Dasarik, M.A., Doskocil, E.J., Mankidy, P.J., Goff, M., 2004. Transesterification of soybean oil with zeolite and metal catalysts. Appl. Catal. A: Gen. 257, 213e223. Talukder, M.M.R., Wu, J.C., Lau, S.K., Cui, L.C., Shimin, G., Lim, A., 2009. Comparison of Novozym 435 and Amberlyst 15 as heterogeneous catalyst for production of biodiesel from palm fatty acid distillate. Energy Fuels 23, 1e4. Tana, T., Lua, J., Niea, K., Denga, L., Wanga, F., 2010. Biodiesel production with immobilized lipase: a review. Biotechnol. Adv. 28, 628e634. Tao, Y., Kanoh, H., Abrams, L., Kaneko, K., 2006. Mesopore-modified zeolites: preparation, characterization, and applications. Chem. Rev. 106, 896e910. Taylor, R.A., Ellis, H.A., 2007. Room temperature molecular and lattice structures of a homologous series of anhydrous zinc(II) n-alkanoate. Spectrochim. Acta A. 68, 99e107. Taylor, R.A., Ellis, H.A., Maragh, P.T., White, N.A.S., 2006. The room temperature structures of anhydrous zinc(II) hexanoate and pentadecanoate. J. Mol. Struct. 787, 113e120. Tesser, R., DiSerio, M., Guida, M., Nastai, M., Santacesaria, E., 2005. Kinetics of oleic acid esterification with methanol in the presence of triglycerides. Ind. Eng. Chem. Res. 44, 7978e7982. Tesser, R., Casale, L., Verde, D., Di Serio, M., Santacesaria, E., 2010. Kinetics and modeling of fatty acids esterification on acid exchange resins. Chem. Eng. J. 157, 539e550. Trakarnpruk, W., Porntangjitlikit, S., 2008. Palm oil biodiesel synthesized with potassium loaded calcined hydrotalcite and effect of biodiesel blend on elastomer properties. Renewable Energy 33, 1558e1563. Trost, B.M., 1991. The atom economy: a search for synthetic efficiency. Science 254, 1471e1477. Twaiq, F., Zabidi, N.A.M., Mohamed, A.R., Bhatia, S., 2003. Catalytic conversion of palm oil over mesoporous aluminosilicate MCM-41 for the production of liquid hydrocarbon fuels. Fuel Process. Technol. 84, 105e120.
276 16. APPLICATIONS OF HETEROGENEOUS CATALYSTS IN THE PRODUCTION OF BIODIESEL Van Gerpen, J., Knothe, G., 2009. Basis of the transesterification reaction. In: Knothe, G., Krahl, J., Van Gerpen, J. (Eds.), The Biodiesel Handbook, second ed. AOCS Press, Urbana, pp. 31e46. Vicente, G., Martinez, M., Aracil, J., 2004. Integrated biodiesel production: a comparison of different homogeneous catalysts systems. Bioresour. Technol. 92, 297e305. Vijayakumar, B., Reddy, C.R., Iyengar, P., Nagendrappa, G., Prakasha, B.S.J., 2005. Synthesis of p-tolyl stearate catalyzed by acid activated Indian bentonite. Indian J. Chem. Technol. 12, 316e321. Wang, Y., Jehng, J., 2011. Hydrotalcite-like compounds containing transition metals as solid base catalysts for transesterification. Chem. Eng. J. 175, 548e554. Wee, L.H., Bajpe, S.R., Janssens, N., Hermans, I., Houthoofd, K., Kirschhock, C.E.A., Martens, J.A., 2010. Convenient synthesis of Cu3(BTC)2 encapsulated Keggin heteropolyacid nanomaterial for application in catalysis. Chem. Commun. 46, 8186e8188. Wilson, K., Clark, J.H., 2001. Solid acids and their use as environmentally friendly catalysts in organic synthesis. Pure Appl. Chem. 72, 1313e1319. Wilson, K., Lee, A.F., 2012. Rational design of heterogeneous catalysts for biodiesel synthesis. Catal. Sci. Technol. 2, 884e897. Wypych, F., Adad, L.B., Mattoso, N., Marangon, A.A., Schreiner, W.H., 2005. Synthesis and characterization of disordered layered silica obtained by selective leaching of octahedral sheets from chrysotile and phlogopite structures. J. Colloid Interf. Sci. 283, 107e112. Xi, Y., Davis, R.J., 2008. Influence of water on the activity and stability of activated Mg/Al hydrotalcites for the transesterification of tributyrin with methanol. J. Catal. 254, 190e197. Xia, P., Liu, F., Wang, C., Zuo, S., Qi, C., 2012. Efficient mesoporous polymer based solid acid with superior catalytic activities towards transesterification to biodiesel. Catal. Commun. 26, 140e143. Xie, W., Peng, H., Chen, L., 2006. Calcined MgeAl hydrotalcites as solid base catalysts for methanolysis of soybean oil. J. Mol. Catal. A: Chem. 246, 24e32. Yin, P., Chen, L., Wang, Z., Qu, R., Liu, X., Xu, Q., Ren, S., 2012. Biodiesel production from esterification of oleic acid over aminophosphonic acid resin D418. Fuel 102, 499e505. Zatta, L., Ramos, L.P., Wypych, F., 2012. Acid activated montmorillonite as catalysts in methyl esterification reactions of lauric acid. J. Oleo Sci. 61, 497e504. Zatta, L., Ramos, L.P., Wypych, F., 2013. Acid-activated montmorillonites as heterogeneous catalysts for the esterification of lauric acid with methanol. Appl. Clay Sci 80e81, 236e244. Zeng, H., Feng, Z., Deng, X., Li, Y., 2008. Activation of MgeAl hydrotalcite catalysts for transesterification of rape oil. Fuel 87, 3071e3076. Zhang, Y., Dube, M.A., Mclean, D.D., Kates, M., 2003. Biodiesel production from waste cooking oil: 1. Process design and technological assessment. Bioresour. Technol. 89, 1e16. Zhang, H., Ding, J., Qiu, Y., Zhao, Z., 2012a. Kinetics of esterification of acidified oil with different alcohols by a cation ion-exchange resin/ polyethersulfone hybrid catalytic membrane. Bioresour. Technol. 112, 28e33. Zhang, H., Ding, J., Zhao, Z., 2012b. Microwave assisted esterification of acidified oil from waste cooking oil by CERP/PES catalytic membrane for biodiesel production. Bioresour. Technol. 123, 72e77. Zhao, J., Guan, H., Shi, W., Cheng, M., Wang, X., Li, S., 2012. A BrønstedeLewis-surfactant-combined heteropolyacid as an environmental benign catalyst for esterification reaction. Catal. Commun. 20, 103e106. Zieba, A., Drelinkiewicz, A., Konyushenko, E.N., Stejskal, J., 2010. Activity and stability of polyaniline-sulfate-based solid acid catalysts for the transesterification of triglycerides and esterification of fatty acids with methanol. Appl. Catal. A: Gen. 383, 169e181.
C H A P T E R 17 Lignocellulose-Based Chemical Products Ed de Jong 1,*, Richard J.A. Gosselink 2 1 Avantium Chemicals, Amsterdam, The Netherlands, 2Food and Biobased Research, Wageningen UR, Wageningen, The Netherlands *Corresponding author email: ed.dejong@avantium.com O U T L I N E Introduction 278 Occurrence and Composition of Lignocellulosic Biomass Storage Carbohydrates Structural Carbohydrates 278 280 280 Cellulose 280 Hemicelluloses Glucuronoxylans Glucomannan Xyloglucans Galactoglucomannans Arabinoglucuronoxylans Arabinogalactan Arabinoxylan b-(1/3, 1/4)-Glucans Complex Heteroxylans Conclusions on Carbohydrate Feedstocks 280 280 282 282 282 282 282 283 283 283 283 Lignin 283 Pretreatment Technologies Steam Explosion Liquid Hot Water Wet Oxidation Dilute and Concentrated Acid Pretreatment Alkaline (Lime) Pretreatment Process 286 286 288 288 289 289 Pretreatment Technologies Still at a Laboratory/ Conceptual Stage Ammonia Fiber Explosion/Ammonia Recycle Percolation) Ionic Liquids Sub/Supercritical Treatments Summary of Lignocellulosic Biomass Pretreatments Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00017-6 290 290 291 291 291 Lignocellulosic BiorefineriesdClassification 292 C6 and C6/C5 Sugar Platform Fermentation Products Chemical Transformation Products 295 295 296 Lignin Platform 296 Importance of Furans and Aromatics as Building Blocks for Chemicals and Fuels 297 Carbohydrate Dehydration 298 Introduction 298 Furfural Production and Applications 298 5-Hydroxymethylfurfural Formation from Hexose Feedstock 301 Relevance of 5-Hydroxymethylfurfural as a Platform 304 Chemical Conversion of Technical Lignins into Monoaromatic Chemicals Base-catalyzed Depolymerization Acid-catalyzed Depolymerization Pyrolysis Oxidative Depolymerization Reductive Hydrodeoxygenation Solvolysis Sub- and Supercritical Water Supercritical Solvents Ionic Liquids Future Perspectives of Lignin Aromatics 305 305 305 305 306 306 307 307 308 308 308 Conclusions and Further Perspectives 309 References 309 277 Copyright Ó 2014 Elsevier B.V. All rights reserved.
278 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS INTRODUCTION Around the world significant steps are being taken to move from today’s fossil-based economy to a more sustainable economy based on biomass. A key factor in the realization of a successful biobased economy will be the development of biorefinery systems allowing highly efficient and cost-effective processing of biological feedstocks to a range of biobased products, and successful integration into existing infrastructure. The recent climb in oil prices and consumer demand for environmentally friendly products have now opened new windows of opportunity for biobased chemicals and polymers. Industry is increasingly viewing chemical and polymer production from renewable resources as an attractive area for investment. Within the biobased economy and the operation of a biorefinery there are significant opportunities for the development of biobased building blocks (chemicals and polymers) and materials (fiber products, starch derivatives, coatings, resins, etc.). In many cases this happens in conjunction with the production of bioenergy or biofuels. The production of biobased products could generate US$ 10e15 billion of revenue for the global chemical industry. The economic production of biofuels is often a challenge. The coproduction of chemicals, materials, food and feed can generate the necessary added value. The world is more and more confronted with the reduction of fossil oil reserves, strong fluctuations of fossil fuel prices and the increase in CO2 emissions and the ensuing problem of the greenhouse gas effect. Recent development on the production of shale gas at various places in the world might change this picture on the short term, but the disadvantages associated with fossil resources stay in place. These environmental, social and economic challenges have created the need for sustainable alternatives to fossil fuels and chemicals (Brown, 2003; Kamm et al., 2006). The use of plant biomass as starting material is one of the alternatives to reduce the dependency on fossil oil for transportation fuels and is the main alternative to replace petrochemicals. The biomass can be transformed into energy, transportation fuels, various chemical compounds and materials such as natural fibers by biochemical, chemical, physical and thermal processes (Brown, 2003; Huber et al., 2006; Gallezot, 2012; Climent et al., 2011a,b; Lichtenthaler and Peters, 2004). The fermentation and the chemical conversion of carbohydrates into value-added compounds has received increasing interest in the last decade, and in a biorefinery different advantages may be taken from both processes (Kamm et al., 2006; Gallezot, 2012; Climent et al., 2011a; Lichtenthaler and Peters, 2004; Spiridon and Popa, 2008; Lin and Huber, 2009; Stöcker, 2008; Dhepe and Fukuoka, 2008). However, the potential competition with food and feed applications and the consequent rise in feedstock prices is an important aspect to take into consideration. Therefore the use of lignocellulosic feedstocks (often referred to as secondgeneration feedstocks) is strongly advocated. In addition to carbohydrates also substantial amounts of lignin is produced when using lignocellulosic feedstocks. In this chapter the composition of lignocellulosic biomass is discussed followed by an overview of the most important pretreatment and fractionation technologies. Especially the effect of the different technologies on the subsequent fermentative/chemocatalytic conversions is addressed. The importance is illustrated by an overview of the most important commercial as well as anticipated chemical building blocks from carbohydrates and lignin with a special emphasis on the production of furan-based building blocks from carbohydrates and aromatic building blocks originating from lignin. OCCURRENCE AND COMPOSITION OF LIGNOCELLULOSIC BIOMASS Lignocellulosic biomass is the most abundant organic compound on Earth and represents the major portion of the world’s annual production of renewable biomass. The global biomass production is about 150 billion tons annually (Balat and Ayar, 2005). Carbohydrates are by far the most omnipresent component of lignocellulosic biomass and are therefore often the preferred feedstock for the biobased economy. In fermentative processes there is sometimes more room for feedstock flexibility (proteins, triglycerides/fatty acids) but in the case of catalytic conversions such as the transformation of biomass into furan molecules you are restricted to carbohydrates. Sources of carbohydrates include conventional forestry, wood processing by-products (e.g. wood chips, sawdust, bark, pulp and paper industrial residue as black liquor), agricultural crops and surpluses (e.g. corn stover, wheat and rice straw), and so-called energy crops (e.g. switchgrass, Miscanthus, willow) grown on degraded soils and aquatic biomass (algae, seaweeds). In this chapter we will focus on lignocellulosic biomass. Typical carbohydrate compositions are shown in Table 17.1. The majority of lignocellulosic biomass consists of carbohydrates (60e80%); the other main component is lignin (20e25%); proteins are mainly found in fresh (green) plant material. Amounts of triglycerides, extractives and inorganic materials are very much species as well as harvest time dependent. The bulk of the carbohydrates present in biomass are composed of poly/oligosaccharides, such as
TABLE 17.1 Carbohydrate Composition of the Main Biomass Types C6-Sugars* Origin Species Hardwoods (Average) Mixed (stem) Softwoods (Average) Mixed (stem) Grasses Sugarcane bagasse Agricultural Residues Corn cobs 75 39 Wheat straw 57 Rice Husks 49 C5-Sugars* Glu Man Gal Rha Fuc Uro Xyl Ara 38e50% cellulose 67e75% carbohydrate 43 0.4 0.9 0.5 0.1 0.2 16 40e50% cellulose 67e75% carbohydrate 44 4.9 7.8 0.4 0.3 32e34 0.5 1.6 Lignin Reference 1.3 18e25 Fengel and Wegener 1984, Ebringerova et al., 2005 8.9 5.9 27e33 Fengel and Wegener 1984, Ebringerova et al., 2005 20e23 2 19e24 Girio et al., 2010, Han, J.S. 1998, Martin et al., 2008 N.D.{ 30 3.3 N.D. Nabarlatz et al., 2007 32 N.D.{ 20 2.8 2.6x 16e21 Nabarlatz et al., 2007, Han, J.S. 1998 30 N.D.{ 17 2 1.1x 21 Nabarlatz et al., 2007, Han, J.S. 1998 * Glu, Glucose; Man, Mannose; Gal, Galactose; Rha, Rhamnose; Fuc, Fucose; UrA, Uronic acids; Xyl, Xylose; Ara, Arabinose; Others, e.g. ash. x Ace, Acetyl groups. { N.D., Not determined; d.m., dry mass. Others* 1.4 4x OCCURRENCE AND COMPOSITION OF LIGNOCELLULOSIC BIOMASS Carbohydrate Content (% d.m.){ 279
280 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS hemicelluloses, cellulose, starch, and inulin. Sucrose is an omnipresent disaccharide consisting of a glucose and fructose moiety, whereas monosaccharides such as glucose and fructose are present in far lesser amounts. In particular, lignocellulosic plant matter is available in large quantities and is relatively cheap while aquatic biomass is given great potential for the future. Storage Carbohydrates Often the biological energy storage systems are also based on carbohydrates like sucrose (saccharose), starch and inulin, which will not be discussed here. In this overview we will only focus on the structural carbohydrates of terrestrial biomass. Structural Carbohydrates Cellulose and hemicellulose can be found in the cell wall of all terrestrial plant cells. In terrestrial biomass the combined cellulose and hemicellulose fraction represents almost always more than 50% of the total biomass based on dry weight. Cellulose is a linear polymer composed of b-D-glucopyranose (glucose) units forming microfibrils that give strength and resistance to the cell wall. The hemicellulose consists of a wide variety of polysaccharides (composed of pentoses, hexoses, hexuronic acids), which are interspersed with the microfibrils of cellulose, conferring consistency and flexibility to the structure of the cell wall (Spiridon and Popa, 2008). CELLULOSE Cellulose is the basic structural component of plant cell walls and comprises about a third of all vegetable materials. Cellulose is a complex polysaccharide, consisting of 3000 or more b-(1 / 4) linked D-glucose units (Table 17.2). It is present in wood in quantities between 40% and 50% on dry matter basis (Table 17.1). It is the most abundant of all naturally occurring organic compounds, comprising over 50% of all the carbon in vegetation. Cellulose is a straight-chain polymer where no coiling or branching occurs, and the molecule adopts an extended and rather stiff rodlike conformation. Cellulose consists of crystalline parts together with some amorphous regions. The chains can stack together to form larger microfibrils, which make cellulose highly insoluble in water. Cellulose microfibrils may also associate with water and matrix polysaccharides, such as the (1 / 3, 1 / 4)-b-D-glucans, heteroxylans (arabinoxylans (AXs)) and glucomannans (GMs) (Sinha et al., 2011; Fengel and Wegener, 1984). HEMICELLULOSES Hemicelluloses are the world’s second most abundant renewable polymers after cellulose in lignocellulosic materials. Hemicelluloses are a heterogeneous class of polymers representing, in general, 15e35% of plant biomass and which may contain pentoses (b-D-xylose and a-L-arabinose), hexoses (b-D-mannose, b-D-glucose, and a-D-galactose) and/or uronic acids a-D-4-O-methylgalacturonic and (a-D-glucuronic, a-D-galacturonic acids). Other sugars such as a-L-rhamnose and a-L-fucose may also be present in small amounts and the hydroxyl groups of sugars can be partially substituted with acetyl groups (Ebringerova et al., 2005; Peng et al., 2012; Girio et al., 2010). Composition and amounts strongly depend on plant source, plant tissue and geographical location. Hemicelluloses are usually bonded to other cell wall components such as cellulose, cell wall proteins, lignin, and phenolic compounds by covalent and hydrogen bonds, and by ionic and hydrophobic interactions. The most relevant hemicelluloses are the xylans and the GMs, with xylans being the most abundant. Xylans are the main hemicellulose components of secondary cell walls constituting about 20e30% of the biomass of hardwoods (angiosperms) and herbaceous plants. In some tissues of grasses and cereals xylans can account up to 50% (Ebringerova et al., 2005). Xylans are usually available in large amounts as by-products of forest, agriculture, agroindustries, wood and pulp and paper industries. Mannan-type hemicelluloses such as GMs and galactoglucomannans (GGMs) are the major hemicellulosic components of the secondary wall of softwoods (gymnosperms), whereas in hardwoods they occur in minor mounts. Depending on their biological origin, different hemicellulose structures can be found (Table 17.2). Upon hydrolysis, the hemicelluloses are converted into the corresponding monosaccharides (Table 17.1). The major hemicelluloses are discussed below. Glucuronoxylans Hemicelluloses in various hardwood species differ from each other both quantitatively and qualitatively. The main hemicelluloses of hardwood are glucuronoxylans (O-acetyl-4-O-methylglucurono-b-(1,4)-D-xylan; GXs), which can also contain small amounts of GMs. In hardwoods, GXs represent 15e30% of their dry mass and consist of a linear backbone of b-(1,4)-D-xylopyranosyl units. Some xylose units are acetylated at C2 and C3 and 1 in 10 molecules has an uronic acid group (4-O-methylglucuronic acid) attached by a-(1,2) linkages (Table 17.2). The percentage of acetyl groups ranges between 8% and 17% of total xylan (about 3.5e7 seven acetyl residues per 10 xylose units). The xylosidic bonds
TABLE 17.2 Main Types of Di-, Oligo- and Polysaccharides Present in Biomass based on Ebringerova et al., 2005, Girio et al., 2010 and Peng et al., 2012 Units Linkage DPx b-(1 / 4) 100e>10,000 b-D-Galp a-L-Araf b-L-Arap b-(1 / 6) a-(1 / 3) b-(1 / 3) 100e600 b-D-Glcp b-D-Xylp b-D-Xylp b-D-Galp a-L-Araf a-L-Fucp Acetyl b-(1 a-(1 b-(1 a-(1 a-(1 10e25 b-D-Manp b-D-Glcp b-D-Galp Acetyl a-(1 / 6) 40e100 GM 2e5 b-D-Manp b-D-Glcp b-(1 / 4) 40e70 Hardwoods GX 15e30 b-D-Xylp 4-O-Me-a-D-GlcpA Acetyl a-(1 / 2) 100e200 Arabinoglucuronoxylan Grasses and cereals, softwoods AGX 5e10 b-D-Xylp 4-O-Me-a-D-GlcpAb-L-Araf a-(1 / 2) a-(1 / 3) 50e185 Arabinoxylans Cereals AX 0.15e30 b-D-Xylp a-L-Araf-Feruloy a-(1 / 2) a-(1 / 3) Glucuronoarabinoxylans Grasses and cereals GAX 15e30 b-D-Xylp a-L-Araf 4-O-Me-a-D-GlcpA Acetyl a-(1 / 2) a-(1 / 3) Saccharide Type Biological Origin Cellulose All terrestrial plants Arabinogalactan Softwoods Xyloglucan Abbreviation Backbone 40e50% b-D-Glcp AG 1e3; 35** b-D-Galp Hardwoods, softwood, grasses XG 2e25 Galactoglucomannan Softwoods GGM Glucomannan Hardwoods and softwoods Glucuronoxylan Side Chains$$ / / / / / 4) 3) 2) 2) 2) HEMICELLULOSES Amount* $$ * %, Dry biomass. x Degree of polymerization. ** (Up to) in the heartwood of larches. $$ p ¼ sugar in pyranose configuration, f ¼ sugar in furanose configuration. 281
282 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS between the xylose units are easily hydrolyzed by acids, but the linkages between the uronic acid groups and xylose are very resistant. Acetyl groups are easily cleaved by alkali, and the acetate formed during kraft (alkaline) pulping of wood mainly originates from these groups. Besides these main structural units, GXs may also contain small amounts of L-rhamnose and galacturonic acid. The latter increases the polymer resistance to alkaline agents. The average degree of polymerization (DP) of GXs is in the range of 100e200 (Peng et al., 2012; Girio et al., 2010). Glucomannan In addition to xylan, hardwoods contain 2e5% of a GM, which is composed of b-glucopyranose and b-mannopyranose units linked by b-(1 / 4) bonds (Table 17.2). However, the mannose/glucose monomer ratio may vary depending on the original source of GM. The ratio of glucose to mannose varies between 1:2 and 1:1. Galactose is not present in hardwood mannan. The mannosic bonds between the mannose units are more rapidly hydrolyzed by acid than the corresponding glycosidic bonds, and GM is easily depolymerized under acidic conditions. There may be certain short side branches at the C3 position of the mannoses and acetyl groups randomly present at the C6 position of a sugar unit. The acetyl groups frequently range from 1 per 9 to 1 per 20 sugar units (Peng et al., 2012). by a-D-galactopyranosyl units attached to glucose and mannose by a-(1,6) bonds. Acetyl groups content of GGM is around 6%, corresponding, on average, to one acetyl group per three to four hexose units (Girio et al., 2010) (Table 17.2). Some GGMs are water soluble, presenting in that case higher galactose content than the insoluble GGMs. There are two main types of acetylgalactoglucomannans in softwoods, one being galactose-poor (5e8% of dry wood) and the other galactose-rich (10e15% of dry wood). The ratio of galactose:glucose:mannose is approximately 0.1:1:3 and 1:1:3, for the two woods, respectively (Peng et al., 2012). GGMs have an approximate DP between 100 and 150, which is equivalent to a molecular weight around 16,000e24,000 Da. GGMs are easily depolymerized by acids, especially the bonds between galactose and the main chain. The acetyl groups are much more easily cleaved by alkali and acid (Peng et al., 2012). GMs occur in minor amounts in the secondary wall of hardwoods (<5% of the dry wood mass) (Girio et al., 2010). As GGMs, they have a linear backbone of b-D-glucopyranosyl (Glcp) and b-D-mannopyranosyl (Manp) units but the ratio Glcp:Manp is lower. In GGMs and GMs the extent of galactosylation governs their association tendency to the cellulose microfibrils and hence their extractability from the cell wall matrix (Ebringerova et al., 2005). Arabinoglucuronoxylans Besides xylan and GM, xyloglucans (XGs) are also present in the primary cell walls of some higher plants (mainly in hardwoods, and less in softwoods) (Peng et al., 2012). They can also appear in small amounts (2e5%) in grasses. XGs consist of b-1,4-linked D-glucose (cellulosic) backbone with 75% of these residues substituted at O-6 with D-xylose. L-Arabinose and D-galactose residues can be attached to the xylose residues forming di- or triglycosyl side chains. Also L-fucose has been detected attached to galactose residues. In addition, XGs can contain O-linked acetyl groups. XGs interact with cellulose microfibrils by the formation of hydrogen bonds, thus contributing to the structural integrity of the cellulose network (Girio et al., 2010). Arabinoglucuronoxylans (AGXs) (arabino-4-Ometylglucuronoxylans) are the major components of nonwoody materials (e.g. agricultural crops) and a minor component of softwoods (5e10% of dry mass). They consist of a linear b-(1,4)-D-xylopyranose backbone containing 4-O-methyl-D-glucuronic acid and a-L-arabinofuranosyl linked by a-(1,2) and a-(1,3) glycosidic bonds (Table 17.2) (Girio et al., 2010). The xylopyranose backbone might be slightly acetylated (Peng). The typical ratio arabinose:glucuronic acid:xylose is 1:2:8. Conversely to hardwoods xylan, AGXs might be less acetylated, but may contain low amounts of galacturonic acid and rhamnose. The average DP of AGXs ranges between 50 and 185 (26). In addition, because of their furanosidic structure, the arabinose side chains are easily hydrolyzed by acids (Peng et al., 2012). Galactoglucomannans Arabinogalactan The major hemicelluloses in softwoods (gymnosperms) are acetylated GGMs, accounting up to 20e25% of their dry mass (Girio et al., 2010). GGMs consist of a linear backbone of b-D-glucopyranosyl and b-D-mannopyranosyl units, linked by b-(1,4) glycosidic bonds, partially acetylated at C2 or C3 and substituted The heartwood of larches contains exceptionally large amounts of water-soluble arabinogalactan (AG), which is only a minor constituent in other softwood species (Peng et al., 2012). Its concentration and quality are not affected by seasonal variability. AGs are highly branched polysaccharides with molecular weights Xyloglucans
283 LIGNIN ranging from 10,000 to 120,000 Da. All larch AGs isolated from the Larix sp. are of the b-(3,6)-D-galactan type and consist of galactose and arabinose in a 6 to 1 ratio. Larch AG has a galactan backbone that features b-(1 / 3) linkages and galactose b-(1 / 6) and arabinose b-(1 / 6 and 1 / 3) side chains (Peng et al., 2012) (Table 17.2). The highly branched structure is responsible for the low viscosity and high solubility in water of this polysaccharide (Peng et al., 2012). It has the ability to bind fat, retain liquid, and dispersing properties and AG also possesses a high biological activity. Larch AG is currently used in a variety of food, beverage, nutraceutical, and medicine applications (Peng et al., 2012). b-(1 / 4)-linked segments also occur. Cellulose is also b-D-glucan, which is linked by (1 / 4)-glycosidic bonds, and thus cellulose has high stiffness (crystallinity) and is insoluble in most solvents. Contrary to cellulose the b-(1 / 3) linkages existing in 1314Gs make glucans flexible and soluble (Peng et al., 2012). Complex Heteroxylans Complex heteroxylans are present in cereals, seeds, gum exudates and mucilages and they are structurally more complex. In this case the b-(1,4)-D-xylopyranose backbone is decorated with single uronic acid and arabinosyl residues and also various mono- and oligoglycosyl side chains (Girio et al., 2010). Arabinoxylan AXs are the main hemicelluloses of the grasses (Gramineae). AXs have been generally present in a variety of tissues of the main cereals: wheat, rye, barley, oat, rice, corn, and sorghum, as well as other plants (Peng et al., 2012). AXs are generally present in the starchy endosperm (flour) and outer layers (bran) of cereal grain. They are similar to hardwood xylan, but the amount of L-arabinose is higher. In AX, the linear b-(1 / 4)-DXylp backbone is substituted by a-L-Araf units in the positions 2-O and/or 3-O (Table 17.2). In addition, the AXs are also substituted by a-D-glucopyranosyl uronic unit or its 4-O-methyl derivative in the position 2-O, as can be found in wheat straw, bagasse and bamboo. O-acetyl substituents may also occur (Peng et al., 2012). According to the amount of glucuronic acid and arabinose, the types of AXs are classified as AGX and glucuronoarabinoxylan (GAX), respectively (Ebringerova et al., 2005). AGXs are the dominant hemicelluloses in the cell walls of grasses and cereals, such as sisal, corncobs and straw. Compared to AGXs, the GAXs have an AX backbone, which contains about 10 times fewer uronic acid side chains than arabinose, and also contains xylan that is double substituted by uronic acid and arabinose units. Ferulic acid and p-coumaric acid can occur esterified to the C-5 of arabinosyl units of GAXs (Peng et al., 2012). The physical and/or covalent interaction with other cell wall constituents restricts xylan extractability (Girio et al., 2010). b-(1 / 3, 1 / 4)-Glucans b-(1 / 3, 1 / 4)-glucans (1314Gs) consist of a linear chain of b-D-glucopyranosyl units linked by (1 / 3) and (1 / 4) bonds (Table 17.1). 1314Gs are present in Poaceae (grasses and cereals) as well as in equisetum, liveworts and Charophytes. The mixed-linkage glucans are dominated by cellotriosyl and cellotetrasyl units linked by b-(1 / 3) linkages, but longer Conclusions on Carbohydrate Feedstocks Storage carbohydrates are uniform in composition and relatively easily to isolate and purify. Therefore many fermentative and catalytic processes have identified these feedstocks as their initial feedstock of choice. Because of costs and societal debates (food versus fuel and indirect land use debates) many researchers both from industry and academia are investigating the use of lignocellulose as feedstock. Pure cellulose has the same advantages as starch in that it is only build up from glucose and relatively easy to hydrolyze (although much more difficult than starch) when pure (and amorphous). However, to make use of lignocellulose economically also the hemicellulose needs to be used. This overview clearly shows that due to the heterogeneity of the monosaccharides incorporated and large diversity in linkages and side groups, both enzymatic hydrolysis system as well as a catalytic/fermentative conversion system needs to be quite robust to make optimal use of the cellulose and hemicellulose fractions. LIGNIN In addition to carbohydrates the major component of lignocelullulosic biomass is lignin. Lignins are major structural components of higher plants, and confer to woody biomass its mechanical structure and resistance to environmental stress and microbial decay. Lignin, the name of which is derived from the Latin word for “wood”, accounts for 15e30 wt% of woody biomass and it is also available from agricultural residues such as straw, grass and bagasse. Lignins are built in plants starting from three basic monolignols via oxidative phenolic coupling reactions to generate the three-dimensional lignin polymer (Ralph et al., 2007). The heterogeneity of lignin polymers exists in molecular composition and linkage types between
284 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS the phenylpropane monomers, p-hydroxyphenyl- (H), guaiacyl- (G), and syringyl- (S) units. These are derived from the monolignols sinapyl-, coniferyl-, and coumaryl alcohol, respectively (Table 17.1). Lignin composition will be different not only between species, but also between different tissues of an individual plant. In softwood lignin coniferyl alcohol is the predominant building unit (over 95% guaiacyl structural elements), while in hardwoods (and dicotyl fiber crops) the ratio coniferyl/synapyl shows considerable variation. In lignins of cereal straws and grasses the presence of coumaryl alcohol leading to p-hydroxyphenylpropane structures is typical. The lignin content ranges (Table 17.1) and chemical structures of the three primary building blocks in lignocellulosic biomass are given in Table 17.3 and the occurrence and type of different interunit linkages in Table 17.4. For the production of aromatic chemicals from biorefinery lignin selection of suitable resources can be made on the occurrence of building blocks and interunit linkages next to the choice of pretreatment and isolation procedure. The presence of one or two methoxyl groups at the ring may for the production of some chemicals (e.g. guaiacol, syringol) be a requirement. For the conversion of lignin into benzene, toluene, xylene (BTX) or phenol the presence of coumaryl units (H-units) may be an advantage due to the lack of side groups next to the aromatic phenolic group (Table 17.3). The complex structure of (isolated) lignins needs suitable characterization methods and an ongoing effort for improvement of these methods have been performed during the last decades. However, results of these analytical procedures are not always consistent. TABLE 17.3 Methods for lignin characterization can be found in the literature (Gosselink et al., 2004; Baumberger et al., 2007; Tejado et al., 2007; Monteil-Rivera et al., 2013) and via the International Lignin Institute (www. ili-lignin.com). Two-dimensional nuclear magnetic resonance and pyrolysis gas chromatography mass spectrometry are now established analytical techniques for detailed structural lignin analysis (Table 17.4). From a chemical perspective, lignins are highly complex polyphenolic biopolymers with aromatic units in different configurations. Lignins are traditionally produced in the pulp and paper mills by extracting lignin upon liberation of cellulosic fibers used for paper making. Most kraft lignins are burnt within paper mills to generate heat and power, thus providing energy autonomy and lowered operating costs. The majority of lignosulfonates are used as additives in the building sector, where they provide plasticity and flowability to concrete. Lignosulfonates are also used as binders in animal feed, in road building, oil well drilling and as dispersants and coatings in pesticides used for agriculture applications. Sulfur-free lignins derived from soda pulping of annual plants such as grass and wheat straw are produced commercially and used among others in wood adhesives and in animal feed. More recently, biorefinery lignins are produced in so-called biorefinery or fractionation processes, for example for the manufacturing of cellulosic bioethanol. This side stream is for the short term used primarily as energy source, but for the medium to longer term utilization of these lignins for the production of biofuels, aromatic chemicals and materials are expected. So far limited industrial use of technical lignins is seen mainly due to the easy use as Lignin Content and Chemical Structures of Lignocellulosic Biomass Lignin (wt%) Phenylpropane Units (%) Structure Coumaryl (H) Coniferyl (G) Sinapyl (S) Softwood 27e33 e 90e95 5e10 Hardwood 18e25 e 50 50 Grasses 17e24 5 75 25 Source: Azadi et al., 2013.
285 LIGNIN TABLE 17.4 Frequencies of Different Interunit Linkage Types in Native Softwood and Hardwood Lignin per 100 C9 units Name Structure Softwood Hardwood a-O-4 40e50 50e60 b-5 10e12 3 5-5 13 3 4-O-5 3 3 b-b 3 3 Bonds to 1-position 1e3 3 Source: Henriksson et al., 2010.
286 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS energy source, the impurities in technical lignin sources, tendency to form condensed structures, inferior performance compared to synthetic compounds, unique reactivity, lack of availability of high-purity lignins and a large variety of different types of lignins (Vishtal and Kraslawski, 2011). Additionally, traditional (heterogeneous) catalysts work inferior to biorefinery lignin and need to be redesigned (Zakzeski et al., 2010). Upon depletion of fossil resources, the production of aromatic chemicals from these resources will come under stress. As lignin is by far the most abundant aromatic renewable resource on earth, lignin is the only resource that could fulfill the quantities needed for the substitution of the main aromatic compounds used in industry (Holladay et al., 2007). These are phenol, BTX, and terephthalic acid (Van Haveren et al., 2008). The annual global production of these largest aromatic chemicals is estimated at 103 million metric tons total, with benzene at 44, toluene at 22, and xylenes at 37 million metric tons (Nexant Chem. Systems, 2012). With an estimated global biomass production of about 150 billion tons annually about 30 billion tons of lignin is generated each year globally (Balat and Ayar, 2005). This amount of lignin is far exceeding the need for the conversion to aromatic chemicals even at low conversion degrees of about 10%. Currently there is a strong desire from major brand owners (e.g. Coca Cola, Pepsi, Heinz) to “green” their product portfolio by using biobased polymer building blocks. Lignin could play in the future an important role as a biobased feedstock. However, there are quite some challenges to overcome for the development of an economically viable process for the production of aromatic chemicals from lignin (Gosselink, 2011). PRETREATMENT TECHNOLOGIES There are a number of key features for the effective pretreatment of lignocellulosic biomass. The pretreatment process should have a low capital and operational cost. It should be effective on a wide range and loading of lignocellulosic feedstocks and should result in the recovery of most of the lignocellulosic components in a usable form in separate fractions. The need for preparation/ handling or preconditioning steps prior to pretreatment such as size reduction should be minimized. It should produce no or limited amounts of sugar and lignin degradation products that inhibit the growth of fermentative microorganisms or the action of hydrolytic enzymes, and it should have a low energy demand or be performed in a manner that energy invested could be used for other purposes such as secondary heating (Agbor et al., 2011). The ideal pretreatment process produces a disrupted, hydrated substrate that is easily hydrolyzed and optimized to accommodate the requirements of subsequent conversion steps, e.g. the formation of sugar degradation products and fermentation inhibitors is avoided, inorganic materials is minimized and/or optimized separation of the main constituents lignin, cellulose and hemicellulose is achieved. Pretreatment technologies are always a combination of physical/physicochemical and chemical steps. Physical pretreatment involves the size reduction by cutting, milling or grinding. Smaller particle sizes result in improved hydrolysis/solvation because of increased surface/volume ratio of the substrate resulting in improved mass transfer rates. Barakat and coworkers reviewed the dry fractionation of lignocellulosic biomass. They concluded that particle sizes must be reduced to 0.5e2 mm in order to decrease heat and mass transfer limitations and to reach a well-accepted level of digestibility. However, currently mechanical size reduction steps are not cost-effective because of too high energy demands of dry grinding operations; therefore, innovative grinding and milling processes or combinations of mechanical size reduction with others pretreatments are still required (Barakat et al., 2013). Table 17.5 highlights the advantages and disadvantages of the different pretreatment technologies that will be discussed in more detail below. The importance of highsolids loadings in biomass pretreatment has recently been reviewed (Modenbach and Nokes, 2012). Steam Explosion Steam explosion pretreatment is one of the most commonly used pretreatment technologies, as it uses a combination of physical and chemical techniques in order to open up and partly break down the structure of lignocellulosic biomass. The steam-explosion pretreatment is a (autocatalytic) hydrothermal process, which subjects the biomass to high pressures and temperatures for a short duration of time after which the system is rapidly depressurized, causing a disrupting of the three-dimensional structure of the biomass. The disruption causes a partial solubilization of the hemicellulose and lignin fraction of the biomass subsequently increasing the accessibility of the cellulose to the hydrolytic enzymes. Particle size is a major contributing factor on the effectiveness of the process, and it has been seen that relatively large particle sizes have been able to yield maximum sugar concentrations. This is a promising finding, as decreasing the particle sizes of the material requires further mechanical processing of the raw material driving up the production costs (Brodeur et al., 2011). Temperatures ranging from 190 to 270  C have been used with residence times of 1e10 min, respectively. The addition of acidic catalysts has been explored in minor amounts in order to improve hemicellulose hydrolysis during the pretreatment and cellulose
287 PRETREATMENT TECHNOLOGIES TABLE 17.5 Advantages and Disadvantages of Different Pretreatment Methods of Lignocellulosic Biomass Pretreatment Method Advantages Disadvantages Processes Pursued at Commercial or Demonstration Scale Mechanical 1. Reduce cellulose crystallinity 1. High power consumption Steam Explosion 1. Cost-effective 2. Lignin transformation and hemicellulose solubilization 3. High yield of glucose and hemicellulose in two-step process 1. Partial hemicellulose degradation 2. Acid catalyst needed to make process efficient with high lignin content material 3. Toxic compound generation 4. Incomplete destruction of lignin-carbohydrate complexes Liquid Hot Water 1. Separation of nearly pure hemicellulose from rest of feedstock 2. No need for catalyst 3. Hydrolysis of hemicellulose 1. High energy/water input 2. Solid mass left over will need to be dealt with (cellulose/lignin) 3. Long residence times Wet Oxidation 1. 2. 3. 4. Concentrated Acid 1. High glucose yield 2. Solubilizes hemicellulose 3. Ambient temperatures 1. High costs of acids and need for recovery and recyclability 2. High costs of corrosion-resistant equipment 3. Formation of inhibitors 4. Hazardous and toxic process Dilute Acid 1. Solubilizes hemicellulose 2. Tends to solubilize some lignin 1. Formation of inhibitors/degradation products 2. Risk of corrosion Alkali 1. 2. 3. 4. Organosolv 1. Efficient separation of lignin, cellulose and hemicellulose fractions 2. Low molecular mass, reactive lignins Removal of lignin Solubilizes hemicellulose Cellulose decrystallization Exothermic process Efficient removal of lignin Low inhibitor formation Increase accessible surface area Proven technology 1. Cost of oxygen 2. Equipment requirements (temperature, pressure) 1. High cost of alkaline catalyst 2. Alteration of lignin structure 3. Long residence times 1. Need for very efficient solvent recycling 2. Higher capex costs because of higher pressures and safety concerns Processes Pursued at a Laboratory or Conceptual Scale AFEX 1. High effectiveness for herbaceous material and low lignin content biomass 2. Cellulose becomes more accessible 3. Causes inactivity between lignin and enzymes 4. Low formation of inhibitors 5. Removes majority of lignin (ARP) 6. High cellulose content after pretreatment (ARP) 7. Herbaceous materials are most affected (ARP) 1. High energy costs and liquid loading (ARP) 2. Not efficient with lignin-rich feedstocks Ionic Liquids 1. Lignin and hemicellulose hydrolysis 2. Ability to dissolve high loadings of different biomass types 3. Mild processing conditions (low temperatures) 1. High solvent costs, therefore need for almost complete solvent recovery and recycle 2. Difficult separation of solvent and products 3. Buildup of inorganics in the ionic liquids 4. Chemical modifications of the ionic liquid Supercritical Fluids 1. Low degradation of sugars 2. Cost-effective 3. Increases cellulose accessible area 1. High pressure requirements 2. Lignin and hemicelluloses unaffected Microbial 1. Low energy requirements 2. Degrades lignin and hemicellulose 1. Long time required 2. Some of the glucose is digested 3. No commercial value from impure lignin and hemicellulose fractions AFEX, ammonia fiber explosion; ARP, ammonia recycle percolation. Source: Based on Agbor et al., 2011; Brodeur et al., 2011; Menon and Rao, 2012; Kumar et al., 2009; da Costa Sousa et al., 2009, Pedersen and Meyer, 2010; Limayem and Ricke, 2012; Garlock et al., 2011.
288 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS digestibility further on in the process. In addition, a reduction of inhibitory compounds formed is seen. The addition of acidic catalysts causes the hydrolysis of acetyl groups into acetic acid. The physical pretreatment is realized during the rapid decompression of the system. This causes a rapid expansion by vaporization of the saturated water within the lignocellulosic biomass; this results in the breakage of the molecular linkages, and leads to a lignocellulosic matrix very susceptible to enzymatic hydrolysis. The steam-explosion pretreatment process has been a proven technique for the pretreatment of different biomass feedstocks. It is able to generate complete sugar recovery while utilizing a low capital investment and low environmental impacts concerning the chemicals and conditions being implemented and has a higher potential for optimization and efficiency (Brodeur et al., 2011). The steamexplosion pretreatment has been demonstrated using a wide range of biomass sources including poplar chips, olive tree residues, wheat straw and corn stover. However, some disadvantages can be seen when using this process. Dilute acids are needed when using softwoods or even when increased yields are warranted for lower acetylated feedstocks. However, the addition of acids comes at a cost because it results in elevated equipment requirements and the higher formation of degradation products such as furfural and 5-hydroxymethyl furfural (HMF), which is sometimes detrimental for subsequent fermentations. In addition salts are formed because the liquors need to be neutralized. These salts need subsequently to be separated from the system and disposed. In the past the focus of many pretreatment technologies was the optimization of the cellulose recovery and subsequent conversion. However, cellulose content of biomass is seldom above 45% making the economics challenging especially because of rising feedstock and energy costs. Therefore, research and developments are now moving into the direction of complete utilization of the entire lignocellulosic biomass. Liquid Hot Water Liquid hot water (LHW) pretreatment uses water at elevated temperatures (160  Ce240  C) and high pressures to keep water in its liquid form in order to promote disintegration and separation of the lignocellulosic matrix (Ruiz et al., 2013). Time ranges from a few minutes up to an hour with temperatures between 160e240  C. The LHW process uses many of the same features as steam explosion, primarily autohydrolysis, although without the rapid decompression of the matrix. The LHW process utilizes flow through reactors of varying configurations or batch techniques with the latter being the primary emphasis at the laboratory scale (Brodeur et al., 2011). A primary goal of this process is to completely solubilize hemicellulose and separate it from the rest of the solid material while reducing the formation of inhibitors. The generation of reactive cellulose fibers for the production of glucose as well as disruption of the entire lignocellulosic matrix is achieved through the cell penetration of the biomass by the water. Both hemicellulose and part of the lignin are solubilized by the LHW acting as an acid as well as a base. On average between 40% and 60% of the total biomass dissolved in the process (Ruiz et al., 2013). There are two main product streams that are formed at the outlet of this process: the solubilized hemicellulose-rich slurry and the cellulose-rich solid fraction. The solubilized product consists primarily of (oligo)saccharides derived from nearly completely solubilized hemicellulose and lignin (35e60% of total starting material) and a minor amount of cellulose (4e15%). Temperature plays an important role in this pretreatment as the quantity of inhibitor (e.g. furfural, acetic acid, HMF, and formic acid) formation is mainly correlated with an increase in temperature. Johnson et al. evaluated inhibitors and concluded that furfural and HMF are not that toxic and that most likely other unidentified compounds are the main reason for the observed toxicity of some of the pretreatment methods (Johnson et al., 2013). The solid fraction consists mainly of more accessible cellulose because of swelling and disruption of the matrix but still needs further treatment (enzymatically/chemocatalytic) to convert it to solubilized products. The primary objective of this pretreatment is, therefore, to reduce the solubilization of cellulose as much as possible while simultaneously maximizing hemicellulose and lignin solubilization. Advantageous aspects of this pretreatment process are the relative low costs because no solvents or additives such as acid catalysts are required; furthermore, expensive reactor systems are not necessary due to the low corrosive nature of this pretreatment technique and the chemicals that are involved. As is the case with many of the pretreatment methods, the severity of the LHW process largely depends on the type of lignocellulosic material that is being used. Wet Oxidation Wet oxidation is a pretreatment technology using water and air or oxygen to fractionate biomass at temperatures above 120  C. A clear advantage of wet oxidation, in particular in combination with alkali, is the relatively mild temperature and the limited formation of fermentation inhibitors (e.g. furan aldehydes and phenolaldehydes) (Klinke et al., 2002). Wet oxidation facilitates the separation of cellulose after the majority of hemicelluloses and lignin has been solubilized. The amount of lignin removed after pretreatment ranges from 50%
PRETREATMENT TECHNOLOGIES to 70% depending on the type of biomass pretreated and the conditions used. The solid material after wet oxidation displayed a higher enzymatic convertability than the remaining solid material after steam explosion (Martin et al., 2008). Wet oxidation is effective in pretreating a variety of biomass such as wheat straw, corn stover, sugarcane bagasse, cassava, peanuts, rye, canola, faba beans, and reed (Brodeur et al., 2011; Martin et al., 2008). Wet oxidation can be combined with other pretreatment methods to further increase the yield of sugars after enzymatic hydrolysis. Combining wet oxidation with alkaline pretreatment has been shown to reduce the formation of by-products, thereby decreasing inhibition. In combination with steam explosion, in a process called wet explosion, the biomass undergoes not only the chemical reaction described above but also physical rupture. The advantages to combining wet oxidation with steam explosion includes the ability to process larger particle sizes and to operate at higher substrate loadings, up to 50% substrate (Brodeur et al., 2011). Dilute and Concentrated Acid Pretreatment Acid pretreatment involves the use of concentrated and diluted acids to break the rigid structure of the lignocellulosic material. The most commonly used acids are sulfuric (H2SO4) and hydrochloric (HCl). Dilute sulfuric acid has traditionally been used to manufacture furfural (van Putten et al., 2013a,b) by hydrolyzing the hemicellulose of mainly corncobs and bagasse into simple sugars. The pentose part (e.g. xylose) is subsequently converted into furfural. Dilute sulfuric acid has also been used commercially to pretreat a wide variety of biomass types to subsequently convert both the C6 and C5 part. Feedstocks evaluated include switchgrass, corn stover, spruce, and poplar (Brodeur et al., 2011 and references therein). Other acids have also been studied, such as phosphoric acid (H3PO4), nitric acid (HNO3) and organic acids (Brodeur et al., 2011; Kootstra et al., 2009). Due to its ability to remove hemicellulose, acid pretreatments have also been integrated in other processes in fractionating the components of lignocellulosic biomass such as the production of dissolving cellulose. Acid pretreatment (removal of hemicellulose) followed by alkali pretreatment (removal of lignin) results in relatively pure cellulose. This chemical pretreatment usually consists of the addition of concentrated or diluted acids (usually between 0.2% and 2.5% w/w) to the biomass, followed by constant mixing at temperatures between 130  C and 210  C. Depending on the conditions of the pretreatment, the hydrolysis of the sugars could take from a few minutes to hours (Brodeur et al., 2011). A key advantage of acid pretreatment is that a subsequent enzymatic hydrolysis step is sometimes not required, as the acid itself hydrolyzes the biomass to yield 289 fermentable sugars. This is especially true with concentrated acid treatments. Hemicellulose and lignin are partly solubilized with relatively minor degradation [75], and the hemicellulose is converted to monomeric and oligomeric sugars with acid pretreatment. A potential drawback is the production of fermentation inhibitors like furfural and HMF, which reduce the effectiveness of the further processes. Therefore, extensive washing and/or a detoxification step is sometimes required to remove these inhibitors before a fermentation step. Most acids have a strong corrosive nature asking for special reactor requirements (material for the reactor) in order to withstand the required experimental conditions and corrosiveness of the acids. The optimum conditions for the acid pretreatment depend highly on the targeted sugars and the purpose of the pretreatment. Up to now most times subsequent conversions were based on fermentative processes. These processes require low amounts of inhibitors (furfural, HMF, organic acids) but are relatively tolerant to inorganic components. But also for fermentative processes this is not clear cut. It was found that the optimal conditions for obtaining the maximum sugar yield depends on whether the goal is to maximize the yield after the pretreatment or after the enzymatic hydrolysis of the pretreated solids or if the goal is to obtain maximum yield after both steps (Lloyd and Wyman, 2005). Delmas (2008) studied the use of formic acid/ acetic acid pretreatment at 105  C and atmospheric pressure to fractionate wheat straw into high purity fractions of organosolv cellulose, lignin and a sugar syrup. The raw straw pulp was separated from the dissolved lignin and hemicelluloses. The pulp can be bleached with hydrogen peroxide and the commercial value of the raw pulp is close to that of eucalyptus chemical pulp. A high-purity lignin, with linear structures, was recovered from the organic medium. Alkaline (Lime) Pretreatment Process The kraft and soda processes used in chemical pulping processes are the predominant processes to produce low lignin fibers suitable for papermaking. Alkaline pretreatments have also widely been researched as a pretreatment step for biorefineries, although in general at more benign conditions compared to traditional pulping processes. Alkaline pretreatment with alkali such as NaOH, KOH, Ca(OH)2, hydrazine and anhydrous ammonia cause swelling of biomass, which increases the internal surface area of the biomass, and decreases both the DP and cellulose crystallinity. Alkaline pretreatment disrupts the lignin structure and breaks the linkage between lignin and the other carbohydrate fractions in lignocellulosic biomass, thus making the carbohydrates in the heteromatrix more
290 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS accessible. The reactivity of remaining polysaccharides increases as the lignin is removed. Acetyl and other uronic acid substitutions on hemicellulose that lessen the accessibility of enzymes to cellulose surface are also removed by alkali pretreatments (Mosier et al., 2005). However, most of the alkali is consumed. Alkali pretreatment is most effective with low lignin content biomass like agricultural residues but becomes less effective as lignin content of the biomass increases (Agbor et al., 2011). Sodium hydroxide has been extensively studied for many years; another alkali that has been used for the pretreatment of biomass is lime. Especially leftover lignocellulosic feedstocks have been shown to benefit from this method of pretreatment and include corn stover, switchgrass, bagasse, wheat, and rice straw. The conditions for alkaline pretreatment are usually less severe than other pretreatments. It can be performed at ambient conditions, but longer pretreatment times are required than at higher temperatures. The alkaline process involves soaking the biomass in alkaline solutions and mixing it at a target temperature for a certain amount of time. A neutralizing step to remove lignin and inhibitors (salts, phenolic acids, furfural, and aldehydes) is required before enzymatic hydrolysis. The advantage of lime pretreatment is that the cost of lime required to pretreat a given quantity of biomass is lowest among alkaline treatments (Brodeur et al., 2011). A number of studies have combined alkaline pretreatment with other pretreatment methods, such as wet oxidation, peroxide treatment, steam explosion, ammonia fiber explosion (AFEX), and ammonia recycled percolation (Brodeur et al., 2011; Yamashita et al., 2010). In summary, the mechanisms of alkaline pretreatment to increase cellulose accessibility is attributed to the separation of structural linkages between lignin and carbohydrates, disruption of the lignin structure and removal of acetyl group and uronic acid substitutions, with associated swelling of cellulose, decrease in DP and crystallinity, thus increasing the accessible surface area of cellulose to enzymes (Zhao et al., 2012). PRETREATMENT TECHNOLOGIES STILL AT A LABORATORY/ CONCEPTUAL STAGE Ammonia Fiber Explosion/Ammonia Recycle Percolation) The ammonia fiber/freeze explosion (AFEX) process is a physicochemical process in which the biomass is subjected to liquid anhydrous ammonia under high pressures and moderate temperatures and then is rapidly depressurized. The AFEX resembles very much the steam-explosion pretreatment technology. However, compared to the steam-explosion process the temperatures (60e100 C) are much more moderate, meaning less energy input and overall energy costs associated with the AFEX process. Major variables in the process are the operation temperature, ammonia concentration and reaction time. The temperature will influence the degree of disruption to the biomass structure, as it will affect the rapidness of the ammonia vaporization within the reactor during depressurization. Typical ammonia loading for many feedstocks are around 1 kg ammonia per kilogram dry biomass. The residence time can be altered from minutes to half an hour duration depending on the degree of saturation needed for the selected biomass (Chundawat et al., 2007). The biomass is saturated for a period of time with the ammonia in a pressurized reactor before being released to atmospheric temperature resulting in a rapid expansion of the ammonia gas causing swelling of the biomass feedstock. This creates hydrolysis of the hemicellulose fraction, a disruption in the lignincarbohydrate linkages, ammonolysis of glucuronic cross-linked bonds, and partial decrystallization of the cellulose structure, all leading to a higher accessible surface area for enzymatic attack (Chundawat et al., 2007). An important prerequisite to make the process economic is a very efficient recovery of the ammonia gas. Under typical AFEX conditions this pretreatment does not remove lignin or any other substances from the biomass; however, the lignin-carbohydrate complexes are cleaved, and the lignin is deposited on the surfaces of the material possibly causing blockage of cellulases to cellulose (Kumar et al., 2009; da Costa Sousa et al., 2009). An overview of the advantages and disadvantages is listed in Table 17.5. Ammonia recycle percolation (ARP) has often been paired with the AFEX pretreatment process, but it can have some different characteristics. In the ARP process, aqueous ammonia of concentration between 5 and 15% (wt%) is sent through a packed bed reactor containing the biomass feedstock at moderately high temperatures (140e210  C) and longer reaction times compared to the AFEX process, increasing the energy costs (Brodeur). The advantage of the ARP process over AFEX is its ability to remove the majority of the lignin (75e85%) as well as solubilize more than half of the hemicellulose (50e60%) while keeping the cellulose in its polymeric form. This results in short-chained cellulosic material containing a high amount of glucan with a high degree (>86%) of enzymatic digestibility and a limited amount of inhibitors. Up to now mostly herbaceous biomass has been treated with this process. Many of the primary concerns with the AFEX process (high energy costs and liquid loadings, along with many disadvantages associated with the AFEX process)
PRETREATMENT TECHNOLOGIES STILL AT A LABORATORY/CONCEPTUAL STAGE need to be addressed before an economical process can be envisioned (Brodeur et al., 2011). Ionic Liquids Room temperature ionic liquids (RTILs) were used for the development of new technologies in chemical and biological transformations, separations, and more recently biomass pretreatment. RTILs consist of an organic cation and an organic or inorganic anion. This tremendous variation allows solvent properties to be tailored to specific applications such as biocatalysis, particularly as nonaqueous alternatives to organic solvents. More recently, RTILs have been used as alternatives for lignocellulosic pretreatment (Mora-Pale et al., 2011). Birch wood was pretreated with N-methylmorpholine-N-oxide (NMMO or NMO) followed by enzymatic hydrolysis and fermentation to ethanol or digestion to biogas. The pretreatments were carried out with NMMO at 130  C for 3 h, and the effects of drying after the pretreatment were investigated (Goshadrou et al., 2013). Another interesting process is the use of concentrated phosphoric acid (CPA) in the pretreatment of lignocellulosic biomass (Zhao et al., 2012). After reprecipitation from CPA cellulose becomes completely amorphous and contains little lignin and hemicellulose. Further research is needed to evaluate and improve the economics of usage of ionic liquids (ILs), NMMO and CPA for pretreatment of lignocellulosic biomass. Also the integration with subsequent chemocatalytic and enzymatic/fermentative processes such as simultaneous saccharification and fermentation needs further research. Especially, the ability of microorganisms to ferment sugars in the presence of these solvents also needs to be tested to carry out a continuous process. ILs are still very expensive and need to be synthesized at a much lower cost and on a much larger scale. Other points of concern are the buildup of inorganics in the ILs introduced with the lignocellulosic biomass (especially a concern with nonwoody lignocellulosic biomass such as straw and bagasse) and chemical modifications of the ILs. So it is rather questionable if the great potential assigned to ILs can be fulfilled for bulk applications such as biomass pretreatment taking into account the aforementioned limitations. Lignocellulosic biomass pretreatment in RTIL’s is an alternative showing promise, with comparable or superior yields of fermentable sugars, than conventional pretreatments. The high number of RTILs that can be synthesized allows the design of solvents with specific physicochemical properties that play a critical role interacting with lignocellulosic biomass subcomponents. Today, these interaction mechanisms are better understood. However, future challenges rely on the ability to 291 make this process economically feasible. This might be achieved by optimizing large-scale pretreatment conditions, performing post-pretreatment steps in RTILs, reusing RTILs, recycling the RTILs with reduced energy consumption and enhancing process efficiency, and producing high-value biobased products and chemicals in addition to ethanol. Moreover, the potential high value of lignin suggests that it might instead be used in the large-scale diversified manufacture of high-value chemicals, traditionally obtained from petroleum (Mora-Pale et al., 2011). Sub/Supercritical Treatments Supercritical fluids (SCFs; conditions where the solvent is both above the critical temperature and critical pressure of the chemical) show unique properties that are different from those of either gases or liquids under standard conditions. SCFs have liquidlike densities and gaslike transport properties of diffusivity and viscosity. So, SCFs have the ability to penetrate the crystalline structure of lignocellulosic biomass overcoming the mass transfer limitations encountered with other pretreatments. Another important advantage is the fact that SCFs have tunable properties such as partition coefficients and solubility. Small changes in temperature or pressure close to critical point can result in up to 100-fold changes in solubility, which can simplify separation. Supercritical carbon dioxide (CO2) with a critical temperature (Tc) of 31  C and a critical pressure (Pc) of 7.4 MPa, as well as supercritical water has been used for biomass pretreatment. REAC fuels and Renmatix are examples of companies employing this kind of technology (Table 17.6). Other technologies such as gamma rays, ozonolysis, biological pretreatment (mainly with fungi) are still in an earlier phase and currently face challenges in scaling up and commercialization (Agbor et al., 2011; Alvira et al., 2010). Summary of Lignocellulosic Biomass Pretreatments Recently technoeconomic comparisons of some of the different pretreatment technologies have been done using identical feedstocks, and analytical methods to generate comparable data (Wyman et al., 2005, 2011; Eggeman and Elander, 2005). The results indicated that no clear winning pretreatment technology could be identified and that further optimization potential is available in the pretreatment methods. It is also clear that the optimal pretreatment technology is very much substrate dependent further hampering the surfacing of a predominant technology (Table 17.7). The effect of pH on solubilization of the different lignocellulosic
292 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS TABLE 17.6 Effect of Various Pretreatment Methods on the Chemical Composition and Chemical/Physical Structure of Lignocellulosic Biomass Increases Accessible Surface Area Pretreatment Sugar Yieldc Mechanical L Steam Explosion H H Liquid Hot Water H H Wet Oxidation H or L Dilute Acid Concentrated Acid H H Removes Hemicellulose Inhibitor Formation Removes Lignin Alters Lignin Structure Nil ND Reuse of Chemicals No H H L - H H L No Nil H Lime (Alkaline) Organosolv* Decrystallizes Cellulose No H H H H H H H L Yes H ND L L H H Yes H H H L Yes H AFEX/ARP H H L L H H Yes Ionic liquids* (NMMO and Ionic Liquids) H H L L H or L* L Yes Supercritical Fluid H H H L * Depends on the chemical nature of the solvent. H, high effect; L, low effect; ND, not determined. Source: Adapted from Mosier et al., 2005; Brodeur et al., 2011; Menon and Rao, 2012. components was nicely illustrated by Garlock et al. (2011) as depicted in Figure 17.1. Table 17.6 summarizes the effect of various pretreatment methods on the chemical composition and chemical/physical structure of lignocellulosic biomass. It can be concluded that at the moment there is no clearly winning technology also because each subsequent conversion process (e.g. fermentative, chemocatalytic) has its own set of requirements. Therefore, a wide range of technologies are currently in the progress of being scaled-up. In Table 17.7 an overview of currently worldwide developed demonstration and pilot plant facilities is presented for production of bioethanol and other chemicals. LIGNOCELLULOSIC BIOREFINERIESdCLASSIFICATION Biorefineries can be classified on the basis of a number of their key characteristics. Major feedstocks include perennial grasses, starch crops (e.g. wheat and maize), sugar crops (e.g. beet and cane), lignocellulosic crops (e.g. managed forest, short rotation coppice, and switchgrass), lignocellulosic residues (e.g. stover and straw), oil crops (e.g. palm and oilseed rape), aquatic biomass (e.g. algae and seaweeds), and organic residues (e.g. industrial, commercial and postconsumer waste). 4 KEY BIOREFINERY CHARACT E RISTICS • • • • Feedstock utilized Biorefinery platform Process Products These feedstocks can be processed to a range of biorefinery streams termed platforms. These platforms include single carbon molecules such as biogas and syngas, five- and six-carbon carbohydrates from starch, sucrose or cellulose; a mixed five- and six-carbon carbo-hydrates stream derived from hemicelluloses, lignin, oils (plant-based or algal); organic solutions from grasses; and pyrolytic liquids. These primary platforms can be converted to a wide range of marketable products using combinations of thermal, biological and chemical processes (Table 17.8). Knowledge of a biorefinery’s feedstock, platform and product allows it to be classified in a systematic manner (Cherubini et al., 2009). The classification of biorefineries enables the comparisons of biorefinery systems, improves the understanding of global biorefinery development and allows the identification of technology gaps.
TABLE 17.7 Demonstration and Pilot Plant Facilities Developed Worldwide for Production of Bioethanol and Other Chemicals Location Products Status Raw Material Pretreatment/Technology Fate of Lignin Abengoa Bioenergia Spain, Kansas, USA 75,000 tons/a EtOH Commercial facility, start-up 2013, 320,000 tons/year Corn stover, wheat straw, switchgrass Acid-catalyzed steam explosion, enzymatic hydrolysis As coproduct, recovered after distillation Beta Renewables Italy, Brazil Variable, cellulose, C5 sugars Commercial facility, start-up 2013, 270,000 tons/year Arundo donax, straw Steam explosion/enzymatic hydrolysis (PROESAÒ ) Solid biofuel Borregard Norway Cellulose, glucose, C5 sugars, lignin Pilot plant 50 kg/h, 2011 Sugarcane bagasse, corn stover, bamboo, eucalyptus, switchgrass, straw, spruce Modified neutral/acidic sulfite cook (Bali process) Performance chemicals CIMV France Cellulose, lignin, C5 sugar stream Pilot plant, in operation since 2006 Wheat straw Concentrated organic acid solvolysis High value product, linear structure Chempolis Finland Cellulose, glucose, C-5 sugars, lignin Demo scale plant, Finland, 2009, 25,000 tons/year Rice and wheat straw, corn stover, Empty Fruit Bunches, Oil Palm Fronts, bagasse, bamboo Organosolv, (Formicobio/ Formicofib process) Clariant (Süd Chemie) Germany 1000 tons/year ethanol Pilot plant, 2012, 4500 tons/year Wheat straw, corn stover or other lignocellulosic material Thermal pretreatment/enzymatic hydrolysis (Sunliquid process) Dupont USA 750 tons/year Pilot plant, 2010 Lignocellulosic, corn stover, switchgrass AFEX/enzymatic hydrolysis Inbicon (Dong Energy) Denmark 4000 tons/a EtOH, C5-molasses solid biofuel Demo facility, start-up 2009 Wheat straw Liquid hot water(hydrothermal, autocatalyzed) Solid biofuel for power-plant, recovered after distillation Iogen Canada 70,000 tons/a EtOH Commercial facility, start-up 2011 Straw (wheat, barley, oat) Modified steam explosion, enzymatic hydrolysis For steam and electricity generation recovered after enzymatic hydrolysis Blue Sugars Corporation (KL Energy) USA 4500 tons/a EtOH Demo facility, operational since 2007, 1e2 MT/h Sugarcane bagasse, wood waste, cardboard and paper Thermomechanical For steam or electricity generation, or as wood pellet Lignol Canada Lignin, cellulose, monomeric hemicellulose stream Pilot plant facility, 1 tons/day Wood, agricultural waste Organosolv (ethanol) High value lignin POET/DSM JC USA 75,000 tons/a EtOH Commercial facility, start 2013 Corn cobs Pretreatment/enzymatic hydrolysis Biogas production Pure Lignin Environmental Technology (PLET) Canada Cellulose, proteins, lignin Pilot plant since 2008, demo plant planned (2012) Softwood (pine) Weak acid pretreatment (nitric acid/ammonium hydroxide) Water-soluble lignin for products Renmatix USA C6/C5 sugar syrups Demo scale plant (100 kg/day dry biomass) Lignocellulose Supercritical fluids (Plantrose process) Sweetwater Energy/Biogasol USA Verenium Process USA 4200 tons/a EtOH Demo facility, operational since 2009 Sugarcane bagasse, energy crops, wood products and switchgrass Mild acid hydrolysis and steam explosion Lignin-rich residue burned for steam generation recovered after distillation Virdia (HCl Cleantech) USA Sugars, lignin Demo Lignocellulose Concentrated HCl, (modified Bergius) Solid fuel Weyland AS Norway Sugars, lignin Pilot plant, 2010, 75 kg/h Lignocellulosedvarious feedstocks, mostly spruce & pine Concentrated acids Lignin as value-added product Demo facility Wet oxidation/steam explosion 293 Source: Partly based on Menon and Rao, 2012; Bacovsky et al., 2013. Solid biofuel for energy generation LIGNOCELLULOSIC BIOREFINERIESdCLASSIFICATION Company
294 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS FIGURE 17.1 Cell wall model showing the general effect of pH on solubilization of hemicellulose and lignin. (A) Untreated cell wall and (B) cell wall during pretreatment. Cellulose can also be degraded under extremely acidic conditions; however, that is not portrayed in this diagram. Source: Designed by Garlock et al., 2011 based on figures from Mosier et al., 2005 and Pedersen and Meyer, 2010. (For color version of this figure, the reader is referred to the online version of this book.) TABLE 17.8 Biomass-Derived Chemical Building Blocks Cn Chemical Company Potential 1 Formic acid Maine BioProducts Pipeline Methane Many companies Growth Ethylene Braskem, DOW/Mitsui, Songyuan Ji’an Biochemical Growth Ethyl acetate Zeachem Pipeline Ethanol Many companies Growth Glycolic acid Metabolic Explorer (Metex) Pipeline Ethylene glycol India Glycols Ltd, Greencol Taiwan Growth Acetic acid Wacker Growth Lactic acid Purac, NatureWorks, Galactic, Henan Jindan, BBCA Growth Acrylic acid Cargill, Perstorp, OPXBio, DOW, Arkema 3-Hydroxy propionic acid Propylene 2 3 TABLE 17.8 Biomass-Derived Chemical Building Blocksdcont’d Cn Chemical Company Potential n-Propanol Braskem Pipeline Ethyl lactate Vertec BioSolvents Growth Isopropanol Genomatica, Mitsui chemicals Pipeline Propylene glycol (1,2-propanediol) ADM Growth n-Butanol Cathay Industrial Biotech, Butamax, Butalco, Cobalt/Rhodia Growth 1,4-Butanediol Genomatica/M&G, Genomatica/Mitsubishi Chemical, Genomatica/ Tate & Lyle Pipeline Iso-butanol Butamax, Gevo Growth Iso-butene Gevo/Lanxess Pipeline Pipeline Methyl methacrylate Lucite/Mitsubishi Rayon, Evonik/Arkema Pipeline Cargill Pipeline Succinic acid Growth Braskem/Toyota Tsusho, Mitsubishi Chemical, Mitsui Chemicals Pipeline BioAmber, Myriant, BASF/Purac, Reverdia (DSM/Roquette), PTT Chem/Mitsubishi CC Iso-butene Gevo/Lanxess Pipeline Epichlorohydrin Solvay, DOW Growth Furfural Many companies Growth 1,3-Propanediol DuPont/Tate & Lyle Growth Furfuryl alcohol a.o. Transfurans Chemicals Growth 4 5
295 C6 AND C6/C5 SUGAR PLATFORM TABLE 17.8 Cn 6 Biomass-Derived Chemical Building Blocksdcont’d Chemical Company Potential Itaconic acid a.o. Qingdao Kehai Biochemistry Co, Itaconix Pipeline Xylitol a.o. Danisco/Lenzing, Xylitol Canada Growth Isoprene/ Farnesene Goodyear/Genencor, GlycosBio, Amyris Pipeline Glutamic acid a.o. Global Biotech, Meihua, Fufeng, Juhua Growth Levulinic acid Maine BioProducts, Avantium, Segetis, Circa Group Pipeline Sorbitol a.o. Roquette, ADM Growth Adipic acid Verdezyne, Rennovia, BioAmber, Genomatica Pipeline Lysine a.o. Global Biotech, Evonik/RusBiotech, BBCA, Draths, Ajinomoto Growth FDCA Avantium Pipeline Isosorbide Roquette Growth Fermentation Products Benzene Phenol(s) 7 fermentation processes providing access to a variety of important chemical building blocks. Glucose can also be converted by chemical processing to useful chemical building blocks. Mixed six- and five-carbon platforms are produced from the hydrolysis of hemicelluloses. The fermentation of these carbohydrate streams can in theory produce the same products as six-carbon sugar streams; however, technical, biological and economic barriers need to be overcome before these opportunities can be exploited. Chemical manipulation of these streams can provide a range of useful molecules. Growth Glucaric acid Rivertop Renewables Pipeline Citric acid a.o. Cargill, DSM, BBCA, Ensign, TTCA, RZBC Growth Caprolactam DSM Pipeline Vanillin o.a. Borregaard Steady Toluene 8 Para-xylene Gevo, Draths*, UOP, Annellotech, Virent Pipeline N** PHA Metabolix, Meridian plastics (103), Tianjin Green Biosience Co. Growth Alkyl benzenes * Draths is recently acquired by Amyris. ** N means unspecified number bigger than 8. Source: Based on De Jong et al., 2012b. The number of chemical building blocks accessible through fermentation is considerable. Fermentation has been used extensively by the chemical industry to produce a number of products with chemical production through fermentation starting around the turn of the twentieth century. Around 8 million tons of fermentation products are currently produced annually (Bakker et al., 2010). • Fermentation-derived fine chemicals are largely manufactured from starch and sugar (wheat, corn, sugarcane, etc.) • The global market for fermentation-derived fine chemicals in 2009 was $16 billion and is forecast to increase to $22 billion by 2013 (Frost and Sullivan, 2011). • The market is broken down as follows: Chemical 2009 ($ millions) 2013 ($ millions) Amino Acids 5410 7821 Enzymes 3200 4900 Organic Acids (Lactic Acid 20%) 2651 4036 Vitamins and Related Compounds 2397 2286 Antibiotics 1800 2600 Xanthan 443 708 Total 15,901 22,351 An overview of current feedstocks, platforms and products is given in Figure 17.2. C6 AND C6/C5 SUGAR PLATFORM Six-carbon sugar platforms can be accessed from sucrose or through the hydrolysis of starch or cellulose to give glucose. Glucose serves as feedstock for (biological) Modern biotechnology is allowing industry to target new and previously abandoned fermentation products and improve the economics of products with commercial potential. Coupled with increasing fossil feedstock costs, cost reductions in the production of traditional fermentation products such as ethanol and lactic acid will allow derivative products to capture new or
296 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS FIGURE 17.2 Overview of the different feedstocks, platforms, conversion steps and products leading to a novel biorefinery classification system. Source: According to Cherubini et al., 2009. (For color version of this figure, the reader is referred to the online version of this book.) increased market shares. Improving cost structures will also allow previously abandoned products such as butanol to reenter the market. Many see the future abundant availability of carbohydrates derived from lignocellulosic biomass as the main driver. However, carbohydrate costs are increasing strongly in recent years and its use for nonfood products is under pressure even in China. Fermentation also gives the industry access to new chemical building blocks previously inaccessible due to cost constraints. The development of cost-effective fermentation processes to succinic, itaconic and glutamic acids promises the potential for novel chemical development. INNOVATIVE F ERMENTA TIO N PRODUCTS Succinic acid Itaconic acid Adipic acid 3-Hydroxypropionic acid/aldehyde Isoprene/farnesene Glutamic acid Aspartic acid Chemical Transformation Products Six- and five-carbon carbohydrates can undergo selective dehydration, hydrogenation and oxidation reactions to give useful products, such as sorbitol, furfural, glucaric acid, HMF and levulinic acid. Over 1 million tons of sorbitol are produced per year as a food ingredient, personal care ingredient (e.g. toothpaste) and for industrial use (ERRMA, 2011; Vlachos et al., 2010). PROMISING GLUCOSE CHEMICAL DERIVATIVES Sorbitol Levulinic acid Glucaric acid Hydroxymethylfurfural 2,5-Furan dicarboxylic acid p-Xylene LIGNIN PLATFORM Lignin offers a significant opportunity for enhancing the operation of a lignocellulosic biorefinery. It is an extremely abundant raw material contributing as much as 30% of the weight and 40% of the energy content of lignocellulosic biomass (Holladay et al., 2007). Lignin’s native structure suggests that it could play a central role as a new chemical feedstock, particularly in the formation of supramolecular materials and aromatic chemicals (Holladay et al., 2007; Hatakeyama and Hatakeyama, 2010). Up to now the vast majority of industrial applications have been developed for
IMPORTANCE OF FURANS AND AROMATICS AS BUILDING BLOCKS FOR CHEMICALS AND FUELS 297 Lignin Syngas products Hydrocarbons Methanol DME Ethanol Mixed Alcohols Fischer Tropsch Liquids C1-C7 gasses Benzene Toluene Xylene Cyclohexane Styrenes Biphenyls Phenols Oxidised products Phenol Substituted phenols Catechols Cresols Resorcinols Eugenol Syringols Coniferols Guaiacols Vanilin Vanilic acid DMSO Aromatic acids Aliphatic acids Syringaidyde Aldehydes Quinones Cyclohexanol β-keto adipate Macromolecules Carbon fibre fillers Polymer extenders Substituted lignins Themoset resins Composites Adhesives Binders Preservatives Pharmaceuticals Polyols FIGURE 17.3 Potential products from lignin. (For color version of this figure, the reader is referred to the online version of this book.) lignosulfonates. These sulfonates are isolated from acid sulfite pulping and are used in a wide range of lower value applications where the form but not the quality is important. The solubility of this type of lignin in water is an important requirement for many of these applications. Around 67.5% of world consumption of lignosulfonates in 2008 was for dispersant applications followed by binder and adhesive applications at 32.5%. Major end-use markets include construction, mining, animal feeds and agriculture uses. The use of lignin for chemical production has so far been limited due to contamination from salts, carbohydrates, particulates, volatiles and the molecular weight distribution of lignosulfonates. The only industrial exception is the limited production of vanillin from lignosulfonates (Evju, 1979). Besides lignosulfonates, kraft lignin is produced as commercial product at about 60 kton/year. New extraction technologies, developed in Sweden, will lead to an increase in kraft lignin production at the mill side for use as external energy source and for the production of value-added applications (Öhman et al., 2009). The production of bioethanol from lignocellulosic feedstocks could result in new forms of higher quality lignin becoming available for chemical applications. The Canadian company Lignol Energy has announced the production of cellulosic ethanol at its continuous pilot plant at Burnaby, British Columbia. The process is based on a wood pulping process using Canadian wood species but the pilot plant will test a range of feedstocks while optimizing equipment configurations, enzyme formulations and other process conditions (Lignol Energy. 2013). The Lignol Energy process proÔ duces a lignin product (HP-L lignin) upon which the company is developing new applications together with industrial partners. Also other lignin types will result from the different biomass pretreatment routes under development and unfortunately there is not one lignin macromolecule that will fit all applications. However, if suitable cost-effective and sustainable conversion technologies can be developed, a lignocellulosic biorefinery can largely benefit from the profit obtained from this side stream lignin (Gosselink, 2011). The production of more value-added chemicals from lignin (e.g. resins, composites and polymers, aromatic compounds, carbon fibers) is viewed as a medium- to longterm opportunity that depends on the quality and functionality of the lignin that can be obtained (Figure 17.3, Table 17.8). The potential of catalytic conversions of lignin (degradation products) has been recently reviewed (Zakzeksi et al., 2010). The main chemical building blocks can be organized by their carbon number, i.e. C1eCn. In the following sections, examples of biobased chemicals are discussed with respect to their current status and the companies that are pursuing the development of these new chemicals. IMPORTANCE OF FURANS AND AROMATICS AS BUILDING BLOCKS FOR CHEMICALS AND FUELS Aromatic compounds are important building blocks for many chemicals and polymers as well as components of fuel compositions. Furans, with their dienic structure, can replace aromatic compounds in several applications including polymers (e.g. Poly Ethylene Terephthalate by Poly Ethylene Furanoate), fuels (diesel) and pharmaceuticals (de Jong et al., 2012a; de Jong et al., 2013; Van Putten
298 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS et al., 2013a). In this paragraph we will discuss the formation of furans from carbohydrates and the formation of aromatic compounds from lignin as an example how all major components of lignocellulosic biomass can be valorized by chemocatalytic routes. Some of the most important chemical transformations of carbohydrates are arguably the hydrolysis and subsequent dehydration of polysaccharides into the furan platform products, furfural and HMF (Dias et al., 2010; Van Putten et al., 2013a,b). Furfural has a wide industrial application profile and is considered as one of the top 30 building blocks that can be produced from biomass (Dias et al., 2010; Van Putten et al., 2013b; Lange et al., 2012; Bozell and Petersen, 2010; Zeitsch, 2000a; Hoydonckx et al., 2007). HMF is promising as a versatile, renewable furan chemical for the production of chemicals, polymers and biofuels, similar to furfural (Van Putten et al., 2013a; Bozell and Petersen, 2010). While furfural has been produced on an industrial scale for decades (Dias et al., 2010; Van Putten et al., 2013b), the production of HMF has not yet reached industrial scale (Van Putten et al., 2013a; Bozell and Petersen, 2010). CARBOHYDRATE DEHYDRATION Introduction The formation of furans from sugars has been known since the early nineteenth century (Dias et al., 2010; Van Putten et al., 2013a,b). Furfural was discovered in 1821 by Döbereiner, by the distillation of bran with dilute sulfuric acid (Kamm et al., 2006; Van Putten et al., 2013b). The resulting compound was first named furfurol (the name comes from the Latin word furfur that means bran cereal, while finishing ol means oil). The furfural molecule has an aldehyde group and a furan ring with aromatic character, and a characteristic smell of almonds. In the presence of oxygen, a colorless solution of furfural tends to become initially yellow, then brown, and finally black. This color is due to the formation of oligomers/ polymers with conjugated double bonds formed by radical mechanisms and can be observed even at concentrations as low as 105 M (Zeitsch, 2000a). Despite the fact that furfural has an LD50 between 50 and 2330 mg/ kg for mice, rats, guinea pigs and dogs, man tolerates its presence in a wide variety of fruit juices, wine, coffee and tea (Zeitsch, 2000a; Hoydonckx et al., 2007). The highest concentrations of furfural are present in cocoa and coffee (55e255 ppm), in alcoholic beverages (1e33 ppm) and in brown bread (26 ppm) (Zeitsch, 2000a). There is no commercially attractive route for the production of furfural from petrochemical resources (Mamman et al., 2008). The synthesis of HMF from biomass was already described in 1895 by Düll (1895) and Kiermayer (1895). Due to its high potential as a platform chemical for a variety of applications, furfural and HMF were mentioned by Bozell in the “top 10 þ 4” list ofbiobasedchemicals (Bozell and Petersen, 2010), along with 2,5-furandicarboxylic acid (FDCA), which is formed by oxidation of HMF (Van Putten et al., 2013a). The formation of furans from sugars takes place through an acid-catalyzed dehydration of sugar molecules at elevated temperature. In general furfural is formed from C-5 sugars and HMF is formed from C-6 sugars. It is therefore not surprising that furans, especially HMF, can be found in essentially all carbohydrate containing heat-treated food. Furfural is known to have some toxic effects, whereas for HMF it is still unclear (Van Putten et al., 2013a). The hydrolysis of polysaccharides and subsequent dehydration into furfural and HMF may be promoted by Brönsted or Lewis acid catalysts (Dias et al., 2010; Van Putten et al., 2013a). Furfural production through traditional processes is accompanied by acidic waste stream production and high energy consumption. Marcotullio and de Jong state that modern furfural production process concepts will have to consider environmental concerns and energy requirements besides economics moreover will have to be integrated within widened biorefinery concepts (Marcotullio and de Jong, 2010). The industrial use of aqueous mineral acids as the catalysts, such as sulfuric acid for furfural production, poses serious operational (corrosion), safety and environmental problems (large amounts of toxic waste). Hence, it is seen desirable to replace conventional aqueous mineral acids by “green” nontoxic catalysts for converting sugars into furfural and HMF. The use of solid acids as catalysts may have several advantages over liquid acids, such as easier separation and reuse of the solid catalyst, longer catalyst lifetimes, toleration of a wide range of temperatures and pressures, and easier/safer catalyst handling, storage and disposal. A road map to furfural, HMF and levulinic acid has recently been presented by the group of Dumesic (Wettstein et al., 2012). Furfural Production and Applications The industrial production of furfural was driven by the need of the United States to become self-sufficient during the First World War. Between 1914 and 1918, intensive exploration for converting agricultural wastes into industrially more valuable products was initiated. In 1921, the Quaker Oats company in Iowa initiated the production of furfural from oat hulls using “left over” reactors (Zeitsch, 2000a). Over time, there was an increased industrial production of furfural and the discovery of new applications. Nowadays, the annual world production of furfural is about 300,000 tons and, although there is industrial production in several countries, the main
299 CARBOHYDRATE DEHYDRATION FIGURE 17.4 Some of the main outlets of furfural. Source: Dias et al., 2010. production units are located in China, the Dominican Republic and South Africa (Kamm et al., 2006; Zeitsch, 2000a; Hoydonckx et al., 2007; Mamman et al., 2008). Figure 17.4 gives an overview of some of the main outlets of furfural. Most of the furfural produced worldwide is converted through a hydrogenation process into furfuryl alcohol, which is primarily used as foundry resin but also increasingly applied as resin to improve wood durability and for the manufacturing of polymers and plastics (Dias et al., 2010). The aldehyde group and furan ring furnish the furfural molecule with outstanding properties for use as a selective solvent (Zeitsch, 2000a; Hoydonckx et al., 2007; Sain et al., 1982). Furfural has the ability to form a conjugated double bond complex with molecules containing double bonds, and therefore is used industrially for the extraction of aromatics from lubricating oils and diesel fuels, or unsaturated compounds from vegetable oils. Furfural is used as a fungicide and nematocide in relatively low concentrations (Zeitsch, 2000a). Additional advantages of furfural as an agrochemical are its low cost, safe and easy application, and relatively low toxicity to humans. Nakagawa and Tomishige (Nakagawa and Tomishige, 2012) have recently reviewed HO OH HO the catalyst system used to produce 1,5-pentanediol from tetrahydrofurfuryl alcohol. Other furan compounds obtained from furfural include levulinic acid (Gürbüz et al., 2012) and tetrahydrofuran. Furfural and many of its derivatives can be used for the synthesis of new polymers based on the chemistry of the furan ring (Hoydonckx et al., 2007; Sain et al., 1982; Win, 2005; Gandini and Belgacem, 1997; Moreau et al., 2004). Furfural derivatives are also excellent starting points for fuel applications (Lange et al., 2012; Gruter and de Jong, 2009; de Jong et al., 2012a,b). Commercially, the pentosans (mainly xylan) present in the hemicellulose fraction of agricultural streams such as corn cobs and sugarcane bagasse are hydrolyzed, using homogeneous acid catalysts in water, giving rise to pentose (xylose), which, by dehydration and cyclization reactions, leads to furfural with a theoretical mass yield of approximately 73% (Scheme 17.1). Nowadays also other feedstocks are considered. Huber and his group developed a new process to produce furfural from waste aqueous hemicellulose solutions from the pulp and paper and cellulosic ethanol industries using a continuous twozone biphasic reactor (Xing et al., 2011). A two-stage hybrid fractionation process was investigated to produce HO OH OH H+ + H HO O O O Pentosans H2O OH O CHO -3 H2O Pentose SCHEME 17.1 Net conversion of pentosans into furfural. O Furfural
300 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS cellulosic ethanol and furfural from corn stover. In the first stage, zinc chloride (ZnCl2) was used to selectively solubilize hemicellulose. During the second stage, the remaining solids were converted into ethanol using commercial cellulase and fermentative microorganisms. Yoo et al. found that the furfural yield from the hemicellulose hydrolysates could be up to 58% based on carbon (Yoo et al., 2012). Yemis and Mazza researched the potential of a microwave-assisted process that provided a highly efficient conversion of wheat straw, triticale straw, and flax shives: obtained furfural yields based on carbon were 48%, 46%, and 72%, respectively (Yemis and Mazza, 2011, 2012). Sahu and Dhepe also presented a solid acidcatalyzed one-pot method for the selective conversion of solid hemicellulose without its separation from other lignocellulosic components, such as cellulose and lignin resulting in 56% furfural yields in biphasic systems (Sahu and Dhepe, 2012). An interesting approach was disclosed by vom Stein and coworkers (vom Stein et al., 2011) by working with “real samples”. They prepared aqueous solutions of FeCl3eNaCl (or seawater) to evaluate the dehydration of xylose into furfural, which can be extracted in situ into 2-methyltetrahydrofuran (2-MTHF) as second phase. Furfural was also successfully obtained when aqueous nonpurified xylose effluents directly from lignocellulose fractionation are tested (vom Stein et al., 2011). Also Marcotullio and De Jong observed good results with FeCl3 (Marcotullio and De Jong, 2010). The hydrolysis of pentosans into pentoses in the presence of H2SO4 is faster than the dehydration of the pentose monomers into furfural (Zeitsch, 2000a; Hoydonckx et al., 2007). Hence, kinetic studies are generally focused on the rate-limiting process, i.e. the dehydration of pentoses. Xylose and arabinose are monomers found in pentosans, which can be converted into furfural, and some studies have shown that the dehydration of arabinose is slower than that of xylose (Zeitsch, 2000a; Kootstra et al., 2009). The concentration of xylose in the various raw materials is almost always much higher than that of arabinose. Considering these factors, it seems reasonable to investigate the kinetics of the dehydration process using xylose as substrate (Zeitsch, 2000a; Sain et al., 1982; Win, 2005; Gandini and Belgacem, 1997; Moreau et al., 2004, 1998; Antal et al., 1991; Root et al., 1959). In the dehydration and cyclization of xylose into furfural, three molecules of water are released per molecule of furfural produced. Huber and coworkers developed a kinetic model for the dehydration of xylose to furfural in a biphasic batch reactor with microwave heating (62). There are four key steps in their kinetic model: (1) xylose dehydration to form furfural, (2) furfural reaction to form degradation products, (3) furfural reaction with xylose to form degradation products, and (4) mass transfer of furfural from the aqueous phase into the organic phase (methyl isobutyl ketone (MIBK)). It was estimated that furfural yields in a biphasic system can reach 85%, whereas at these same conditions in a monophase system furfural yields of only 30% are obtained (Weingarten et al., 2010). Also a kinetic model for the homogeneous conversion of D-xylose in high-temperature water was developed (Kim et al., 2011). Experimental testing evaluated the effects of operating conditions on xylose conversion and furfural selectivity, with furfural yields of up to 60% observed. Also the kinetics of formic acid-catalyzed xylose dehydration into furfural and furfural decomposition was investigated using batch experiments within a temperature range of 130e200  C (Lamminpää et al., 2012). The study showed that the modeling must account for other reactions from xylose besides dehydration into furfural. Moreover, the reactions between xylose intermediate and furfural play only a minor role and that furfural decomposition reactions must take the uncatalyzed reaction in water as solvent into account (Lamminpää et al., 2012). By-products formed in the xylose reaction may also derive from the fragmentation of xylose, such as glyceraldehyde, glycolaldehyde, formic acid, lactic acid, acetol (Antal et al., 1991; Ahmad et al., 1995). As furfural is formed it can be transformed into higher molecular weight products by (1) condensation reactions between furfural and intermediates of conversion of xylose to furfural (and not directly with xylose) and (2) furfural polymerization (Zeitsch, 2000a). Aldol condensation between two molecules of furfural does not occur due to the absence of a carbon atom in Ha position in relation to the carbonyl group (Chheda and Dumesic, 2007). The side reactions (1) and (2) lead to oligomers and polymers with (1) are considered to be more relevant than (2), although published characterization studies of the by-products formed are scarce (Zeitsch, 2000a). The extent of these side reactions can be minimized by reducing the residence time of furfural in the reaction mixture and by increasing the reaction temperature (Zeitsch, 2000a,b; Root et al., 1959; Zeitsch, 2000b). If furfural is kept in the gas phase during the aqueous phase reaction it will not react with intermediates, which are “nonvolatile”. Agirrezabal-Telleria et al. (AgirrezabalTelleria et al., 2011) developed new approaches for the production of furfural from xylose. They propose to combine relatively cheap heterogeneous catalysts (Amberlyst 70) with simultaneous furfural stripping using nitrogen under semibatch conditions. Nitrogen, compared to steam, does not dilute the vapor phase stream when condensed. This system allowed stripping 65% of the furfural converted from xylose and almost 100% of selectivity in the condensate. Moreover, high initial xylose loadings led to the formation of two waterefurfural phases, which could further reduce purification costs. Constant liquidevapor equilibrium during stripping could be maintained for different xylose loadings. The modeling of the experimental data was carried out in order to obtain a liquidevapor mass transfer coefficient. This value could be used for future studies under steady-state continuous conditions
CARBOHYDRATE DEHYDRATION in similar reaction systems (Agirrezabal-Telleria, 2011). Formic acid, a by-product of furfural process (Root et al., 1959), can be an effective catalyst for dehydration of xylose into furfural. There is a growing interest in the use of formic acid as catalyst because it has low corrosiveness and can be easily separated and reused. Using response surface methodology the optimal process parameters (xylose concentration 40 g/l, formic concentration 10 g/l, and a reaction temperature 180  C) were determined to obtain high furfural yield and selectivity. Under these conditions, a maximum furfural yield of 74% and selectivity of 78% were achieved (Yang et al., 2012). Extraction using supercritical CO2 (scCO2) also enhances furfural yields (Kim et al., 2011; Sako et al., 1991, 1992). The above mechanistic considerations for the homogeneous conversion of xylose into furfural using H2SO4 as catalyst may also be considered for solid acid catalysts. Nevertheless, differences in product selectivity between homogeneous and heterogeneous catalytic processes are expected due to effects such as shape/size selectivity, competitive adsorption (related to hydrophilic/hydrophobic properties), and strength of the acid sites. Industrially, furfural is directly produced from the lignocellulosic biomass in the presence of mineral acids, mainly sulfuric acid, under batch or continuous mode operation (Table 17.9). Attempts to improve furfural yields have been made by process innovation, although the use of mineral acids remains a drawback (Zeitsch, 2000a, 69. 70). The cost and inefficiency of separating these homogeneous catalysts from the products makes their recovery impractical, resulting in large volumes of acid waste, which must be neutralized and disposed off. Other drawbacks include corrosion and safety problems. The production of furfural is therefore one of many industrial processes where the reduction or replacement of the “toxic liquid” acid catalysts by alternative “green” catalysts is of high priority. Recently Marcotullio and De Jong (Marcotullio and de Jong, 2010, 2011) shed new light on some particular aspects of the chemistry of D-xylose reaction to furfural. Their aim was to clarify the reaction mechanism leading to furfural TABLE 17.9 Industrial Processes of Furfural Production Industrial Process Catalyst Reaction Type Temperature ( C) Quaker Oats H2SO4 Batch 153 Chinese H2SO4 Batch 160 Agrifurane H2SO4 Batch 177e161 Quaker Oats H2SO4 Continuous 184 Escher Wyss H2SO4 Continuous 170 Rosenlew Acids formed from the raw material Continuous 180 301 and to define new green catalytic pathways for its production. Specifically, their objective was to reduce the use of mineral acids by the introduction of alternative catalysts, e.g. halides, in dilute acidic solutions at temperatures between 170 and 200  C (Schädel et al., 2010). Results indicate that the Cl- ions promote the formation of the 1,2-enediol from the acyclic form of xylose, and thus the subsequent acid-catalyzed dehydration to furfural. For this reason the presence of Cl- ions led to significant improvements for H2SO4 catalyzed reactions. The addition of NaCl to a 50 mM HCl aqueous solution gave 90% selectivity to furfural. Follow-up experimental results by the same group show the halides to influence at least two distinct steps in the reaction leading from D-xylose to furfural under acidic conditions, via different mechanisms. The nucleophilicity of the halides appears to be critical for the dehydration, but not for the initial enolization reaction. By combining different halides synergic effects become evident resulting in very high selectivities and furfural yields (Marcotullio and de Jong, 2011). Also Rong et al. (2012) found that the addition of inorganic salts (e.g. NaCl, FeCl3) promoted the yield of furfural from xylose. Another approach to reduce the inorganic waste streams is to perform the reaction at high temperatures. It was shown that the reaction pathway for the xylose decomposition in hightemperature liquid water can be changed by manipulating the temperature and pressure without any catalyst with a maximum furfural yield of 50% (Jing and Lu, 2007). Many attempts have been made to develop heterogeneous catalytic processes for furfural production that offer environmental and economic benefits, but to the best of our knowledge none has been commercialized (Van Putten et al., 2013b). 5-Hydroxymethylfurfural Formation from Hexose Feedstock HMF stands out among the platform chemicals for a number of reasons: It has retained all six-carbon atoms that were present in the hexoses and high selectivities have been reported for its preparation, in particular from fructose, which compares favorably with other platform chemicals, such as levulinic acid or bioethanol. HMF is formed through the acid-catalyzed dehydration of a hexose, as described in Scheme 17.2. Initially the synthesis of HMF from hexoses was performed in aqueous systems, catalyzed by homogeneous acids. SCHEME 17.2 The acid-catalyzed dehydration of hexose into HMF.
302 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS SCHEME 17.3 The dehydration of glucose and fructose through acyclic intermediates. A number of mechanistic pathways have been proposed for this reaction, which can generally be divided into two groups. The first group is based on a pathway through acyclic intermediates and the second group is based on a pathway through cyclic intermediates. Although there are differences between the various acyclic pathways proposed for the aqueous dehydration of hexoses, they generally propose the formation of the 1,2-enediol intermediate in the Lobry De BruijnAlberda Van Ekenstein transformation (Speck Jr, 1958) between fructose and glucose as the key intermediate (Anet, 1964; Feather and Harris, 1973; Kuster, 1990; Newt, 1951). This intermediate is proposed to dehydrate to a 3-deoxyglucosone, followed by further dehydration and ring closure to form HMF. A schematic representation is provided in Scheme 17.3. The proposed aqueous hexose dehydration pathways through cyclic intermediates generally assume dehydration to start at the C2 hydroxyl position of fructose (Scheme 17.4), leading to the formation of a tertiary carbocation (Van Putten et al., 2013a; Feather and Harris, 1973; Newt, 1951). This is then followed by consecutive dehydrations at C3 and C4 to form HMF. It is clear that in this proposed mechanism, glucose dehydration requires glucose to first isomerize to fructose before it can dehydrate to HMF. Under the acidic reaction conditions, however, this is unfavorable as the isomerization is base catalyzed. The HMF yields and selectivities from the dehydration of fructose, a ketose, are generally much higher than those obtained from the dehydration of glucose, which is an aldose (Van Putten et al., 2013a). The HMF SCHEME 17.4 The dehydration of fructose through cyclic intermediates.
CARBOHYDRATE DEHYDRATION yields for homogeneous acid-catalyzed fructose dehydration in water are limited to around 60% at full conversion, whereas for glucose this is only around 10% at full conversion. Fructose is known to be significantly less stable than glucose, which shows in the required reaction conditions for dehydration. Fructose dehydrates to HMF at temperatures around 100  C in the presence of acid, whereas glucose requires much more severe conditions of at least 140  C in the presence of catalyst to form only small amounts of HMF (less than 10% yield). Quite large variations are seen in the reaction conditions applied by different groups. In some cases relatively high catalyst concentrations in the order of 0.1e1 M mineral acid are applied in fructose dehydration at relatively low temperatures between 100 and 150  C with reaction times in the order of minutes. Others applied lower acid concentrations, but at either longer reaction times or higher temperatures (Van Putten et al., 2013a). Also a significant amount of work has been done with heterogeneous acid catalysts, like ion exchange resins and zeolites, showing comparable selectivities and yields to the homogeneous catalysts (Van Putten et al., 2013a). The HMF yield is limited by its inherent instability under aqueous acidic conditions. In the presence of acid HMF reacts with water (so-called HMF hydration reaction) to form levulinic acid and formic acid, as described in Scheme 17.5 (Kuster, 1990). Other undesirable side reactions are the formation of polymeric material, often referred to as humins (Kuster, 1990), and retroaldol reactions of sugars (Aida et al., 2007). In order to minimize side reactions and HMF hydration, biphasic systems have been researched in which the HMF is extracted to the organic phase (RománLeshkov et al., 2006; Cope, 1959; Kuster and van der Steen, 1977; Kuster and Laurens, 1977; Moreau et al., 1996). The major extraction solvents used are methylisobutylketone, 1-butanol and 2-butanol. The in situ extraction has improved HMF yields from fructose dehydration in some cases to around 70% at full conversion. Due to the high solubility of HMF in water relatively large amounts of solvent are needed, generally at least two equivalents, in order to extract sufficient amounts of HMF (Van Putten et al., 2013a). In the early 1980s a number of researchers started performing HMF synthesis in organic solvents (Nakamura and Morikawa, 1980; Szmant and Chundury, 1981; Brown et al., 1982). The biggest initial challenge here is that, except from high-boiling coordinating solvents like SCHEME 17.5 The acid-catalyzed hydration of HMF to levulinic acid and formic acid. 303 dimethyl sulfoxide (DMSO), N,N-dimethylformamide (DMF) and N-methylpyrrolidinone, most organic solvents do not dissolve sugars very well. The focus was mainly on solvents like DMSO and DMF, showing significant improvements in yield and selectivity (Nakamura and Morikawa, 1980; Szmant and Chundury, 1981; Brown et al., 1982; Musau and Munavu, 1987). In DMSO reaction temperatures of 100e120  C are generally applied and the solvent shows catalytic activity as yields over 90% have been reported in the absence of catalyst (Brown et al., 1982; Musau and Munavu, 1987). An important issue here is the known decomposition of DMSO at temperatures over 100  C. Since 2003 ILs have been extensively researched as solvents for HMF synthesis by many research groups; however, 20 years before that HMF synthesis in pyridinium salts was already performed by Fayet and Gelas, resulting in 70% yield starting from fructose (Fayet and Gelas, 1983). Certain ILs are known to dissolve sugars in high concentrations. The vast majority of this research has been done in imidazolium-based ILs. As is the case for the coordinating organic solvents, the HMF yields for fructose dehydration to HMF in ILs, in which the IL is often also the catalyst, are generally high (70e90%) and levulinic acid formation is in most cases not mentioned (Van Putten et al., 2013a; Zakrzewska et al., 2010). In the work on ILs some conflicting results have been published with the same or comparable ILs. As was already mentioned, HMF synthesis from glucose is much more challenging than from fructose. In 2007 Zhang and coworkers published a breakthrough in glucose dehydration to HMF by using CrCl2 as a catalyst in an imidazolium type IL (Zhao et al., 2007). They achieved an HMF yield of around 70%, which was essentially equal to the yield obtained from fructose in the same system. It is believed that CrCl2 behaves as an isomerization catalyst that forms fructose, which can be dehydrated readily to HMF. Earlier research on HMF synthesis focused mainly on fructose and polymers thereof as substrates. Recent years have seen an enormous increase in interest in the development of biobased platform chemicals as a replacement for fossil-oil based feedstock. For this reason it is preferable to use cheap feedstocks that do not compete with food. Many parties have placed their focus on cellulose, a for humans nondigestible polymer of glucose, as a feedstock. Cellulose is present in large amounts in plant waste material. Application in HMF synthesis will require both hydrolysis and dehydration of the cellulose, either in one reactor or in two separate steps. Recent years have shown a dramatic increase in research on HMF synthesis from cellulose. The main focus has been in line with the work on glucose, applying bifunctional catalyst systems, especially chromium salts in combination with a Brønsted acid.
304 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS Especially in ILs the yields approach those obtained with glucose. The substrate concentration is mostly significantly lower, due to the much lower solubility of cellulose. Also the reaction times are typically much longer for cellulose compared to glucose, likely due to the required hydrolysis prior to dehydration to HMF (Van Putten et al., 2013a). Although sugar dehydration to furans is a hot topic in academia, a lot of research has yet to be done in upscaling these processes to pilot plant and ultimately industrial scale. Especially for hexose dehydration to HMF this holds true. Only two pilot-scale processes are known for the production of HMF: a process from Süddeutsche Zucker-Aktiengesellschaft and a process from Roquette Frères. The first process concerns HMF production in around 5 kg scale from fructose and inulin, a polymer of mainly fructose, catalyzed by oxalic acid at around 140  C in water in which the purification of HMF is done by chromatographic separation (Rapp, 1987). The second process concerns fructose dehydration in a water-MIBK (1:9 v/v) biphasic system in the presence of cationic resins at temperatures between 70 and 95  C (Fleche et al., 1982). In both processes the fructose concentration in water was 20e25 wt% and the HMF yields are in the range of 40e50%. The workup procedures for HMF mentioned in these patents appear unfavorable for a large-scale plant as large-scale chromatographic separation is expensive and a very high solvent to water ratio requires a lot of energy for evaporation of the solvent from the product. In order to produce HMF or a derivative thereof in a cost-effective way, some challenges must be overcome. HMF is unstable under the reaction conditions in the presence of water, leading to the formation of levulinic acid, formic acid and polymeric materials. For this reason contact with water should be minimized. This can be achieved by performing the reaction using other solvents or by continuously extracting the HMF from the aqueous phase. The distribution of HMF over water and extraction solvents is generally not highly favorable toward the solvent, demanding large excess of extraction solvents and therefore energy-consuming workup (Román-Leshkov and Dumesic, 2009). Performing HMF synthesis in other solvents than water is an appealing option. Here the choice has to be made between solvents that have lower boiling points, but exhibit low sugar solubility, and solvents that dissolve high concentrations of sugar, like DMSO and ILs, although from which product separation is difficult due to the high affinity of HMF for these solvents. Two processes focus on the production of derivatives of HMF in order to produce the furanic product more effectively. Mascal and coworkers have focused their efforts on the production of 5-chloromethylfurfural in a biphasic system of concentrated hydrochloric acid and 1,2-dichloroethane (Mascal and Nikitin, 2008). Avantium Chemicals opened their pilot plant in December of 2011 on alcohol-based production of HMF ethers, which will be used for the production of furan-based polymers (Gruter and Dautzenberg, 2007). Relevance of 5-Hydroxymethylfurfural as a Platform Chemical HMF is a very important building block for a wide range of applications. In this paragraph applications in the areas of polymers, fine chemicals, and fuels are summarized. When HMF is produced at high efficiency follow-up products will become an attractive option to replace petrochemical analogs. An interesting molecule that can be derived from HMF is FDCA. It can be obtained via the oxidation of HMF; several oxidation methods have been described in literature (Van Putten et al., 2013a). FDCA was identified by the US Department of Energy (Bozell and Petersen, 2010) to be a key bioderived platform chemical, which in itself is the building block for polyesters, polyamides and plasticizers but FDCA can also serve as starting point for several other interesting molecules, including succinic acid, FDCA dichloride, and FDCA dimethyl ester. In addition to FDCA, other platform chemicals can be produced as well. 5-Hydroxymethylfuroic acid, 2,5-diformyl furan, the 2,5-diamino-methylfuran, and 2,5-bishydroxymethylfuran are most versatile intermediate chemicals of high industrial potential because they are six-carbon monomers that could replace, for example, adipic acid, alkyldiols, or hexamethylenediamine in the production of polymers (Van Putten et al., 2013a). 2,5-Furandicarboxaldehyde and 2,5-hydroxymethylfuroic acid can be considered intermediates to FDCA in the oxidation of HMF. De Vries, Heeres and coworkers (Buntara et al., 2011) have shown an interesting route to convert HMF into caprolactam, the monomer for nylon-6. In addition to applications in the polymer field HMF can also be used in many fine chemicals applications. In view of the rigid furan structure and the two substituents that can be easily modified, HMF has been used in quite a number of pharmaceutical studies (Van Putten et al., 2013a). HMF-derived 5-amino-levulinic acid (Binder et al., 2010) and its derivatives are herbicides. A synthesis route was published by Descotes in collaboration with Südzucker (Schinzer et al., 2004). The Maillard reaction between reducing carbohydrates and amino acids is undoubtedly one of the most important reactions in the flavor and fragrance world, leading to the development of the unique aroma and taste as well as the typical browning, which contribute to the sensory quality of thermally processed foods, such as cooked or roasted meat, roasted coffee or cocoa.
CONVERSION OF TECHNICAL LIGNINS INTO MONOAROMATIC CHEMICALS Although numerous studies have addressed the structures and sensory attributes of the volatile odor-active compounds, the information available on nonvolatile, sensory-active components generated during thermal food processing is scarce but HMF derivatives play an essential role (Van Putten et al., 2013a). HMF has also been linked to natural products, sugar derivatives (e.g. glucosylated HMF) and spiroketals (Van Putten et al., 2013a). HMF can also be a precursor of fuel components. HMF is a solid at room temperature with very poor fuel blend properties; therefore, HMF cannot be used and has not been considered as a fuel or a fuel additive. The Small Medium-sized Enterprise (SME) company Avantium is developing chemical, catalytic routes to produce furan derivatives “furanics” for a range of biofuel applications (de Jong et al., 2012a,b). Avantium targets biofuels with advantageous qualities, both over existing biofuels such as bioethanol and biodiesel as well as over traditional transportation fuels. Another major goal is minimizing the H2 demand for their production. These C5-derived furanic monoethers and C6-derived furanic diethers have a relatively high energy density, and good chemical and physical characteristics, no difference in the engine operation was observed and strongly decreased smoke and particulates emissions. The use of furans, such as HMF and furfural, as precursors of liquid hydrocarbon fuels is also an option for the production of linear alkanes in the molecular weight range appropriate for diesel or jet fuel. The group of Dumesic has researched and evaluated the different strategies possible for upgrading HMF to liquid fuels (531 Alonso et al., 2010). HMF can be transformed by hydrogenolysis to 2,5-dimethyl furan. To form larger hydrocarbons, HMF and other furfural products can be upgraded by aldol condensation with ketones, such as acetone, over a basic catalyst (NaOH) already at room temperatures (West et al., 2008). Also several levulinic acid derivatives have been proposed for fuel applications, for instance ethyl levulinate, g-valerolactone, and MTHF (Geilen et al., 2010). The conversion of HMF to fuels has recently been reviewed (Mäki-Arvela et al., 2012). CONVERSION OF TECHNICAL LIGNINS INTO MONOAROMATIC CHEMICALS The conversion of technical lignin into these monoaromatic chemicals is assumed to be a long-term application (Holladay et al., 2007). Increased worldwide research activities can be observed in this area where predominantly thermochemical approaches are under study to convert lignin model compounds and depolymerize technical lignins into the desired aromatic compounds. In general, lignin depolymerization can not only be performed in aqueous and organic phases, but 305 also in dry form. Complex mixtures are the result in which the individual mass yields barely exceeds few percent. Mostly, CeOeC bonds are cleaved, while the CeC linkages in the lignin structures are very resistant to cleavage. The use of catalysts seems to be a necessity and these activities have been recently reviewed (Zakzeski et al., 2010; Gallezot, 2012; Azadi et al., 2013) showing the following main routes for technical lignin depolymerization in (mono)aromatic chemicals. Base-catalyzed Depolymerization Most work related to base-catalyzed depolymerization (BCD) originates from the pulp and paper industry where these alkaline processes are used to depolymerize and liberate lignin from the lignocellulosic matrix as described in the previous sections. Besides extensive cleavage of the b-O-4 linkages under BCD conditions the methoxyl contents in lignin decrease with the severity of alkaline conditions. However, repolymerization of liberated lignin fragments to condensation products may occur. Alcell organosolv lignin depolymerization in alkali (0e4%) yielded 7e30% liquid products. The maximum concentration of identified phenols was 4.4%, mostly syringol (2.4%) and a limited amount of guaiacol when less severe conditions were applied. Catechol was found at higher pH and temperatures (Thring, 1994). More recently, Yuan et al. (2010) studied the base-catalyzed degradation of kraft lignin in watereethanol at 220e300  C, with phenol as the capping agent into oligomers with a negligible char and gas production. Under the conditions applied lignin could not be degraded completely into lignin monomers. Base-catalyzed lignin depolymerization with the addition of boric acid greatly facilitates the depolymerization of lignin in water, increase product selectivity and boric acid acts as a capping agent to suppress addition and condensation reactions (Roberts et al., 2011). Acid-catalyzed Depolymerization Depolymerization of Alcell lignin using Lewis acid catalysts NiCl2 or FeCl3 yielded gas, solid and liquid products including the formation of ether-soluble monomers under different reaction conditions. Both catalysts favor condensation reactions leading to insoluble residues. The low yields of organic monomers were dominated by phenolics over ketones and aldehydes (Hepditch and Thring, 2000). Pyrolysis Pyrolysis of isolated lignins gives a different product distribution than pyrolysis of wood of other lignocellulosic materials. Lignin pyrolysis occurs in a wider
306 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS temperature range (e.g. 160e900  C) compared to polysaccharides (e.g. 220e400  C) (Yang et al., 2007). Furthermore, the amount of char from isolated lignins is significantly higher compared to whole biomass pyrolysis. Solid acid catalysts such as H-Zeolite Socony Mobil-5 can effectively shift the products toward more deoxygenated compounds. Different isolated lignins pyrolyzed at temperature ranges of 500e800  C yielded bio-oil, gas and char of 16e70%, 3e39%, and 17e81%, respectively (Azadi et al., 2013). Several researchers showed that inorganic alkaline catalysts such as NaOH can facilitate depolymerization of lignin by pyrolysis and influence the product composition (Amen-Chen et al., 2001). Recently, an international study of fast pyrolysis of lignin was undertaken with contribution from 14 laboratories. Based on the results it was concluded that an impure lignin containing up to 50% carbohydrates behaves like whole biomass, while a purified lignin was difficult to process in the fast pyrolysis reactors and produced a much lower amount of a more enriched aromatic bio-oil. It was concluded that for highly pure lignin feedstocks new reactor designs will be required other than the typical fluidized bed fast pyrolysis systems (Nowakowski et al., 2010). Upgrading of lignin pyrolysis oil by catalytic hydrodeoxygenation (HDO) is often used as described by de Wild et al. (2009). More stable oil due to partial removal of oxygen is an important upgrading property. Bu et al. (2012) made a review on the catalytic HDO upgrading of lignin-derived phenols from biomass pyrolysis. This study shows that further investigation of HDO is needed to improve catalysts and optimize operation conditions, further understanding of kinetics of complex bio-oils, and availability of sustainable and cost-effective hydrogen sources. Further HDO treatments are discussed in the next session. Anellotech (2010) has developed a technology platform using catalytic pyrolysis for the claimed inexpensive production of chemicals and transportation fuels from nonfood biomass. Vispute et al. (2010) claim that all chemical conversions can be performed in one reactor, using an inexpensive catalyst. Target green chemicals are BTX. Oxidative Depolymerization In general oxidative depolymerization of lignin is carried out to produce aromatics with an increase in oxygen-containing groups, mostly aldehydes. The production of vanillin (3-methoxy-4-hydroxybenzaldehyde) by oxidative depolymerization of lignin, mainly from black liquor of sulfite pulping is the most well-known process. This commercial process is typically performed at 160e175  C under alkaline conditions using a copper catalyst by Borregard in Norway. Especially softwood lignin is yielding relatively higher amounts of vanillin as compared to hardwood lignin where syringaldehyde may prevail (Evju, 1979). Other researchers used hydrogen peroxide for oxidative depolymerization. Kraft lignin was treated at 90  C by a biomimetic system, using hemin as a catalyst and hydrogen peroxide as an oxidizing agent, which mimics the catalytic mechanism of lignin peroxidase. Relatively high yields of vanillin 19%, vanillic acid 9%, 2-methoxyphenol 2% and 4-hydroxybenzaldehyde 2% were obtained (Suparno et al., 2005). Xiang and Lee (2000) found that alkaline peroxide treatment of lignin at 80e160  C yields mainly low molecular weight organic acids (up to 50%) with only traces of aromatics, which are rapidly degraded by hydrogen peroxide. Sales et al. (2004, 2007) studied the alkaline oxidation of sugarcane soda lignin with a continuous fluid bed with a palladium chloride PdCl3.3H2O/g-Al2O3 catalyst at 100e250  C and 2e10 bar partial oxygen pressure. Total aldehyde yield on lignin was 12%. Zakzeski et al. (2010) reported other predominantly catalytic lignin oxidation processes yielding aromatic aldehydes and acids, which do not exceed 10% on lignin basis. However, lignin model compounds show in some catalytic processes good conversions, which are promising to further develop catalytic strategies for lignin depolymerization in a biorefinery concept. Voitl and Rudolf von Rohr (2010) studied a process for producing vanillin and methyl vanillate from kraft lignin by acidic oxidation in aqueous methanol with H3PMo12O40 as a homogeneous catalyst in the presence of 10 bar oxygen. A stable yield of 3.5 wt% vanillin and 3.5 wt% methyl vanillate can be obtained together with 60 wt% of oligomeric products in the extract. The monomers can be effectively separated using organic solvent nanofiltration (Werhan et al., 2012). Reductive Hydrodeoxygenation HDO is a promising upgrading technology to remove the oxygen from biomass-derived streams, for example obtained after pyrolysis. Strong emphasis is put on finding selective catalysts to minimize the use of hydrogen while maintaining the aromatic functionality of lignin. HDO of lignin model compounds can be efficiently performed over a copper chromite catalyst (Deutsch and Shanks, 2012). The hydroxymethyl group of benzyl alcohol is highly reactive to HDO. Demethoxylation of anisole is the primary reaction pathway in contrast to demethylation and transalkylation. The latter are more prevalent for conventional hydrotreating catalysts. The hydroxyl group of phenol strongly activated the aromatic ring toward cyclohexanol and cyclohexane.
CONVERSION OF TECHNICAL LIGNINS INTO MONOAROMATIC CHEMICALS When applied directly to isolated technical lignin a wide range of chemical reactions occur at 380e430  C including cleavage of interunit linkages, deoxygenation, ring hydrogenation, and removal of alkyl and methoxyl moieties. A complex bio-oil is the result, but the oxygen content of this hydropyrolysis oil is lower compared to pyrolysis oil and therefore this HDO bio-oil is chemically more stable. The hydrogen pressure, typically 50e150 bar, strongly influences the oil yield. Ideal catalysts should have high activity for hydrogenolysis and/or cracking of CeOeC and CeC linkages; low activity for ring hydrogenation; meaningful selectivity toward a certain aromatic compound or class of compounds to allow effective product isolation; high resistance against coke formation and easy regeneration; high sulfur resistance for processing sulfur-containing lignins. Bifunctional catalysts comprise an active hydrogenation metal (e.g. NiMo-Cr2O3, Pd, Co-Mo) and an acidic support such as zeolites to selectively open some CeC bonds. By using catalysts the yield of HDO bio-oil has been improved from 15% up to 81% (Azadi et al., 2013). For development of viable catalytic HDO bio-oil upgrading technologies to produce transportation fuel include (1) improved catalysts, (2) alternative hydrogen source, (3) detailed kinetics study and (4) optimizing the HDO reactions conditions suitable for existing refinery infrastructure (Bu et al., 2012). Solvolysis Alternatively, instead of the use of metal catalysts and hydrogen for hydrogenation, solvolytic depolymerization reactions were performed in the presence of hydrogen donors such as tetralin or anthracene derivatives (Dorrestijn et al., 1999). However, the high costs of these solvents that are consumed during the process prevent practical implementation. A solution to this problem could be the use of formic acid or 2-propanol as hydrogen donors (Kleinert and Barth, 2008; Kleinert et al., 2009). In the presence of relatively large amounts of formic acid and a low chain alcohol the resulting phenolic oil contains substantial amounts of aliphatic hydrocarbons, indicating that extensive hydrogenation of the resulting depolymerization products occurs (Gellerstedt et al., 2008). Another advantage of this process is the negligible formation of char. Xu et al. (2012) used this approach to depolymerize lignin with a combination of formic acid and a Pt/C catalyst in ethanol to further promote the production of lower molar mass fractions. After 4 h all lignin has been completely solubilized. The highest H/C and lowest O/C molar ratios were obtained with prolonged reaction times. Lignin depolymerization in aqueous ethanol leads to a reduced formation of char, which might be attributed to the solubility power of ethanol and the hydrogen 307 donation capability of ethanol to stabilize generated free lignin radicals (Ye et al., 2012). Zakzeski et al. 2012 used ethanol/water mixtures that greatly enhanced the solubility of different technical lignins (e.g. kraft, organosolv and sugarcane bagasse lignin) and consequently led to higher yields of monoaromatics in one-pot lignin liquid phase reforming (LPR) reactions. During solubilization extensive cleavage of various ether linkages in the macromolecule occurred. The Pt/Al2O3-catalyzed LPR reactions yielded up to 17% of monomeric guaiacol-type products for kraft lignin in the presence of H2SO4. Depending on the lignin source and the used cocatalyst, different product distributions and light gases such as hydrogen and methane were formed. Char formation was not observed in any of the reactions. HDO reduction of solubilized lignin using transition metal catalysts led to the formation of alkyl-substituted guaiacol-type molecules with isolated yields of up to 6% for Pt/Al2O3. Toledano et al. 2012 used a microwave-assisted bifunctional catalytic process using tetralin or formic acid as in situ hydrogen donating solvents lead to over 30% bio-oil yield mostly enriched in monomeric and dimeric phenolic compounds. However, the amount of biochar and residual lignin still needs to be reduced. Organosolv and kraft lignin were depolymerized using a silica-alumina catalyst in a water/1-butanol mixture to a yield of 85e88 C-mol%. In a second step the lignin-derived slurry was cracked over a ZrO2e Al2O3eFeOx catalyst in water/1-butanol Total recovered phenols is 6.6e8.6% and the conversion of methoxy phenol reached 92e94% to phenol and cresol (Yoshikawa et al., 2013). Sub- and Supercritical Water Depolymerization of lignin in sub- and supercritical water (pc > 22.1 MPa; Tc > 374  C) lead to extensive lower molar mass fragments, dealkylation and demethoxylation, but a part of these fragments tend to cross-link in larger fragments. The economic viability of this process is severely controlled by the extent to which the heat is recovered from the effluents. The yield of monomers is positively correlated with base concentration added with maximum yield of one-third of the initial lignin. Low molecular weight fraction yields increased with longer reaction times in supercritical water without catalysts at 350e400  C and 25e40 MPa. The water-soluble fraction consists of catechol (28%), phenol (7.5%), and cresols (11%), suggesting the cleavage of both ether and carbonecarbon (Wahyudiono et al., 2008). Addition of phenolics (e.g. phenol and p-cresol Okuda et al., 2004a,b, 2008; Fang et al., 2008) gives a complete depolymerization of lignin into dimers without char formation. Phenol and p-cresol depressed
308 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS cross-linking reactions due to entrapment of reactive fragments, like formaldehyde, and capping of active sites like Ca in the lignin structure. Supercritical Solvents Lignin depolymerization in supercritical solvents next to water includes ethanol, methanol, CO2, and CO2/acetone/water. The supercritical properties of these fluids are displayed in Table 17.10. The choice for using CO2 as solvent is obvious as CO2 is cheap, environmentally friendly and generally recognized as safe by the US Food and Drug Administration. scCO2 has other advantages because of its high diffusivity combined with its easily tunable solvent strength. To use CO2 under supercritical conditions, the temperature needed is low (>31  C) and the pressure needed relatively low (>7.4 MPa) in comparison to other supercritical solvents (Table 17.10). Additionally, CO2 is a gas at room temperature and pressure, which leads to a solvent-free product after pressure expansion. A drawback of scCO2 is its low polarity, which is comparable to hexane, but this problem can be overcome by using cosolvents to change the polarity of the SCF (Herrero et al., 2010). Furthermore, SCF processing based on CO2 enables the easy recycling of CO2, which is advantageous for the development of a sustainable process. Research performed on supercritical processing of lignin to produce aromatic compounds has been summarized hereafter. Depolymerization of lignin model compounds and organosolv lignin have been studied in supercritical alcohols like methanol and ethanol in a temperature range of >239  C and a pressure of >8.1 MPa. By using bases such as KOH and NaOH a high depolymerization conversion was obtained. The dominant depolymerization route is the solvolysis of ether linkages in the lignin structure while the carbonecarbon linkages are mostly stable (Miller et al., 1999; Minami et al., 2003). Yuan et al. (2010) used BCD at mild temperatures (220e300  C) of kraft lignin in watereethanol into oligomers with a negligible char and gas production. TABLE 17.10 Supercritical Fluid Parameters Solvent Carbon Dioxide Critical Temperature Tc ( C) Critical Pressure Pc (MPa) 31 7.4 Water 374 22.1 Acetone 235 4.7 Methanol 239 8.1 Ethanol 241 6.2 1-Butanol 287 4.9 Source: Reid et al., 1987. However, under the conditions applied lignin could not be completely degraded into monomers. Oxidation of lignin and lignin model compounds with peroxide was studied under scCO2 conditions in the absence of alkali. The 5-5 biphenols were shown to be degraded and in this process mostly the formation of carboxylic acids from kraft lignin was observed (Argyropoulos et al., 2006). Gosselink et al. (2012) found that hardwood and wheat straw organosolv lignins were depolymerized in supercritical carbon dioxide/acetone/water fluid at 300  C and 100 bar into 10e12% monomeric aromatic compounds. Small amounts of formic acid were introduced as in situ hydrogen donor. Furthermore, lignin is converted into a phenolic oil consisting of both monomeric and oligomeric aromatic compounds. Interestingly, maximum individual yields of 3.6% for syringol and 2.0% for syringic acid based on lignin were obtained. Depolymerized phenolic products and char were separated during this process by pressure expansion. As during this process competition occurs between lignin depolymerization and recondensation of fragments a substantial amount of char is formed. Ionic Liquids Recent work has demonstrated that ILs are excellent solvents for processing woody biomass and technical lignin. Seeking to exploit ILs as media for depolymerization of lignin, lignin model compounds were treated using Brønsted acid catalysts in 1-ethyl-3-methylimidazolium triflate at temperatures below 200  C. A 11.6% molar yield of the dealkylation product 2-methoxyphenol from the model compound 2-methoxy-4-(2-propenyl)phenol and cleaved 2-phenylethyl phenyl ether, a model for lignin ethers, was obtained. However, depolymerization of organosolv lignin to monomers failed (Binder et al., 2009). The oxidative depolymerization of lignin in 1-ethyl-3-methylimidazolium trifluoromethanesulfonate with Mn(NO3)2 catalyst yielded 11.5 wt% of pure 2,6-dimethoxy-1,4-benzoquinone (Stark et al., 2010). Hossain and Aldous (2012) reviewed the achieved results for depolymerization of lignin model compounds in ILs, but for technical lignin samples mixed results have been obtained. It should be emphasized that conversion of lignin in ILs is still at its infancy, but there is certain potential to make use of these solvents in the valorization of lignin into aromatic chemicals. Future Perspectives of Lignin Aromatics Although the research activities show that in 2013 there is a great interest in using lignin as a renewable resource for the production of aromatic chemicals, it
REFERENCES is also clear that commercial utilization will take substantial time. So far the literature results show that relatively low conversion yields to about 10 wt% based on dry lignin and resulting complex mixtures hinder the commercial utilization of these processes. The Netherlands can play an important role in the lignin aromatics valorization technologies as technology provider with the strong presence and strategic location of academia, chemical industries and other stakeholders in the value chain. In the Port of Rotterdam in the Netherlands about 5 million tons of aromatic building blocks are currently produced and distributed to the chemical industry in the Netherlands, Germany, Belgium and other countries. These aromatic bulk chemicals used and produced consist of so-called aromatic monomers like BTX, styrene and phenol. In the Netherlands in 2010 the Wageningen UR Lignin Platform was established, which plays an important role in this lignin valorization value chain development (http://www.wageningenur.nl/Lignin-Platform. htm). This is a joint research program with academia and industry dedicated to develop the entire lignin bioaromatics value chain. Besides this initiative other networks in Canada and Scandinavia work on lignin valorization topics. Considering these increased research activities on lignin conversion and valorization technologies it can be concluded that the race to produce bioaromatics from renewable feedstocks is wide open. Next to lignin as aromatic feedstock conversion of carbohydrates to aromatic chemicals is also under investigation (Dodds and Humpheys, 2013). It should also be emphasized that many of the above-discussed technologies are at a very early stage, which makes it at present unclear if and which of those routes can become cost competitive as well as sustainable. CONCLUSIONS AND FURTHER PERSPECTIVES The use of lignocellulosic feedstocks as an important source for chemicals and fuels is gaining momentum. This chapter has indicated that there are many variables to take into consideration. We have learned that lignocellulosic biomass consists of three major groups: the softwoods, hardwoods and grasses and that there is also great heterogeneity within each group. There are multiple pretreatment routes developed that are currently scaled up to pilot, demonstration and commercial scales. The optimal pretreatment technology needs to be selected based on the available feedstock and the desired product. At the moment there are no indications that one pretreatment method will be the optimal route for all feedstocks and products. Many 309 routes toward chemical building blocks based on monomeric carbohydrates are ready for scaling up; lignin conversion into monomeric building blocks needs substantial additional R&D before economical processes are within reach. References Agbor, V.B., Cicek, N., Sparling, R., Berlin, A., Levin, D.B., 2011. Biomass pretreatment: fundamentals toward application. Biotechnol. Adv. 29, 675e685. Agirrezabal-Telleria, I., Larreategui, A., Requies, J., Güemez, M.B., Arias, P.L., 2011. Furfural production from xylose using sulfonic ion-exchange resins (Amberlyst) and simultaneous stripping with nitrogen. Bioresour. Technol. 102, 7478e7485. Ahmad, T., Kenne, L., Olsson, K., Theander, O., 1995. The formation of 2-furaldehyde and formic acid from pentoses in slightly acidic deuterium oxide studied by 1H-NMR spectroscopy. Carbohydr. Res. 276, 309e320. Aida, T.M., Tajima, K., Watanabe, M., Saito, Y., Kuroda, K., Nonaka, T., Hattori, H., Smith Jr, R.L., Arai, K., 2007. Dehydration of d-glucose in high temperature water at pressures up to 80 MPa. J. Supercrit. Fluids 42, 110e119. Alonso, M.D., Bond, J.Q., Dumesic, J.A., 2010. Catalytic conversion of biomass to biofuels. Green Chem. 12, 1493e1513. Alvira, P., Tomás-Pejó, E., Ballesteros, M., Negro, M.J., 2010. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis: a review. Bioresour. Technol. 101, 4851e4861. Amen-Chen, C., Pakdel, H., Roy, C., 2001. Production of monomeric phenols by thermochemical conversion of biomass: a review. Bioresour. Technol. 79, 277e299. Anellotech Inc, 2010. Low Cost Petrochemicals from Non-food Biomass. accessed December 2012. www.anellotech.com. Anet, E.F.L.J., 1964. 3-deoxyglycosuloses (3-deoxyglycosones) and the degradation of carbohydrates. Adv. Carbohydr. Chem. 19, 181e218. Antal, M.J., Leesomboon Jr, T., Mok, W.S., Richards, G.N., 1991. Mechanism of formation of 2-furaldehyde from D-xylose. Carbohydr. Res. 217, 71e86. Argyropoulos, D.S., Gaspar, A.R., Lucia, L.A., Rojas, O.J., 2006. Supercritical CO2 oxidation of lignin. Production of high value added Products. Chim. l’Ind. 1 (88), 74e79. Azadi, P., Inderwildi, O.R., Farnood, R., King, D.A., 2013. Liquid fuels, hydrogen and chemicals from lignin: a critical review. Renewable Sustainable Energy Rev. 21, 506e523. Bacovsky, D., Ludwiczek, N., Ognissanto, M., Wörgetter, M., 2013. Status of Advanced Biofuels Demonstration Facilities in 2012. a report to IEA Bioenergy task 39. http://demoplants.bioenergy 2020.eu/files/Demoplants_Report_Final.pdf. Bakker, R.R., den Uil, H., van Ree, R., et al., 2010. Financieel-Economische Aspecten van Biobrandstofproductie e Desktopstudie Naar de Invloed van Co-Productie van Bio-Based Producten op de Financiele Haalbaarheid van Biobrandstoffen. rapport 1175. WUR Food and Biobased Research, Wageningen, Nederland. Oktober 2010 (in Dutch). http://edepot.wur.nl/202366. Balat, M., Ayar, G., 2005. Biomass energy in the world, use of biomass and potential trends. Energy Sources 27, 931e940. Barakat, A., de Vries, H., Rouau, X., 2013. Dry fractionation process as an important step in current and future lignocellulose biorefineries: a review. Bioresour. Technol. 134, 362e373. Baumberger, S., Abaecherli, A., Fasching, M., Gellerstedt, G., Gosselink, R., Hortling, B., Li, J., Saake, B., De Jong, E., 2007. Molar mass determination of lignins by size-exclusion chromatography: towards standardisation of the method. Holzforschung 61, 459e468.
310 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS Binder, J.G., Gray, M.J., White, J.F., Zhang, Z.C., 2009. Reactions of lignin model compounds in ionic liquids. Biomass Bioenergy 33, 1122e1130. Binder, J.B., Blank, J.J., Cefali, A.V., Raines, R.T., 2010. Synthesis of furfural from xylose and xylan. ChemSusChem 3, 1268e1272. Bozell, J.J., Petersen, G.R., 2010. Technology development for the production of biobased products from biorefinery carbohydratesd the US Department of Energy’s “Top 10” revisited. Green Chem. 12, 539e554. Brodeur, G., Yau, E., Badal, K., Collier, J., Ramachandran, K.B., Ramakrishnan, S., 2011. Chemical and physicochemical pretreatment of lignocellulosic biomass: a review. Enzym. Res. 2011. http:// dx.doi.org/10.4061/2011/787532. Article ID 787532, 17 pages. Brown, R.C., 2003. Biorenewable Resources e Engineering New Products from Agriculture, first ed. Blackwell Publishing, USA. Brown, D.W., Floyd, A.J., Kinsman, R.G., Roshan-Ali, Y., 1982. Dehydration reactions of fructose in non-aqueous media. J. Chem. Technol. Biotechnol. 32, 920e924. Bu, Q., Lei, H., Zacher, A.H., Wanga, L., Ren, R., Liang, J., Wei, Y., Liu, Y., Tang, J., Zhang, Q., Ruan, R., 2012. A review of catalytic hydrodeoxygenation of lignin-derived phenols from biomass pyrolysis. Bioresour. Technol. 124, 470e477. Buntara, T., Noel, S., Phua, P.H., Melian-Cabrera, I., de Vries, J.G., Heeres, H.J., 2011. Caprolactam from renewable resources: catalytic conversion of 5-hydroxymethylfurfural into caprolactone. Angew. Chem. Int. Ed. 50, 7083e7087. Cherubini, F., Jungmeier, G., Wellisch, M., Willke, T., Skiadas, I., van Ree, R., de Jong, E., 2009. Toward a common classification approach for biorefinery systems. Biofuels Bioprod. Biorefin. 3, 534e546. Chheda, J.N., Dumesic, J.A., 2007. An overview of dehydration, aldolcondensation and hydrogenation processes for production of liquid alkanes from biomass-derived carbohydrates. Catal. Today 123, 59e70. Chundawat, S.P.S., Venkatesh, B., Dale, B.E., 2007. Effect of particle size based separation of milled corn stover on AFEX pretreatment and enzymatic digestibility. Biotechnol. Bioeng. 96, 219e231. Climent, M.J., Corma, A., Iborra, 2011a. Heterogeneous catalysts for the one-pot synthesis of chemicals and fine chemicals. Chem. Rev. 111, 1072e1133. Climent, M.J., Corma, A., Iborra, 2011b. Converting carbohydrates to bulk chemicals and fine chemicals over heterogeneous catalysts. Green Chem. 13, 520e540. Cope, A.C., 1959. Production and recovery of furans. US2917520. da Costa Sousa, L., Chundawat, S.P.S., Balan, V., Dale, B.E., 2009. ‘Cradle-to-grave’ assessment of existing lignocellulose pretreatment technologies. Curr. Opin. Biotechnol. 20, 339e347. de Jong, E., Vijlbrief, T., Hijkoop, R., Gruter, G.-J.M., van der Waal, J.C., 2012a. Promising results with YXY Diesel components in an ESC testcycle using a PACCAR Diesel engine. Biomass Bioenergy 36, 151e159. de Jong, E., Higson, A., Walsh, P., Wellisch, M., 2012b. Product developments in the bio-based chemicals arena. Biofuels, Bioprod. Biorefin. 6, 606e624. de Jong, E., Dam, M.A., Sipos, L., Gruter, G.-J.M., 2013. Furandicarboxylic acid (FDCA), a versatile building block for a very interesting class of polyesters. In: Smith, P.B., Gross, R. (Eds.), ACS Symposium Series “Biobased Monomers, Polymers and Materials”, pp. 1e13. http://dx.doi.org/10.1021/bk-2012-1105.ch001. de Wild, P., Van der Laan, R., Kloekhorst, A., Heeres, E., 2009. Lignin valorisation for chemicals and (transportation) fuels via (catalytic) pyrolysis and hydrodeoxygenation. Environ. Prog. Sustainable Energy 28 (3), 461e469. Delmas, M., 2008. Vegetal refining and agrichemistry. Chem. Eng. Technol. 31 (5), 792e797. Deutsch, K.L., Shanks, B.H., 2012. Hydrodeoxygenation of lignin model compounds over a copper chromite catalyst. Appl. Catal., A 447-448, 144e150. Dhepe, P.L., Fukuoka, A., 2008. Cellulose conversion under heterogeneous catalysis. ChemSusChem 1, 969e975. Dias, A.S., Lima, S., Pillinger, M., Valente, A.A., 2010. Furfural and furfural-based industrial chemicals. In: Pignataro, B. (Ed.), Ideas in Chemistry and Molecular Sciences, Advances in Synthetic Chemistry, vol. 8. Wiley-VCH, Weinheim, pp. 167e186. Dodds, D., Humpheys, B., 2013. Production of aromatic chemicals from biobased feedstock. In: Imhof, P., van der Waal, J.C. (Eds.), Catalytic Process Development for Renewable Materials. WileyVCH Verlag GmbH & Co, pp. 183e238. Dorrestijn, E., Kranenburg, M., Poinsot, D., Mulder, P., 1999. Lignin depolymerization in hydrogen-donor solvents. Holzforschung 53 (6), 611e616. Düll, G., 1895. Über die einwirkung van oxalsäure auf inulin. Chem. Ztg. 19, 216e220. Ebringerova, A., Hromadkova, Z., Heinze, T., 2005. Hemicellulose. Adv. Polym. Sci. 186, 1e67. Eggeman, T., Elander, R.T., 2005. Process and economic analysis of pretreatment technologies. Bioresour. Technol. 96, 2019e2025. European Renewable Resources and Materials Association (ERRMA), 2011. EU-Public/Private Innovation Partnership “Building the Bioeconomy by 2020”. www.errma.com. Evju, H., 1979. Process for the preparation of 3-methoxy4-hydroxybenzaldehyde. US Patent 4151207. Fang, Z., Sato, T., Smith Jr, R.L., Inomata, H., Arai, K., Kozinski, J.A., 2008. Reaction chemistry and phase behavior of lignin in high-temperature and supercritical water. Bioresour. Technol. 99, 3424e3430. Fayet, C., Gelas, J., 1983. Nouvelle méthode de préparation du 5-hydroxyméthyl-2-furaldéhyde par action de sels d’ammonium ou d’immonium sur les mono-, oligo- et poly-saccharides. Accès direct aux 5-halogénométhyl-2-furaldéhydes. Carbohydr. Res. 122, 59e68. Feather, M.S., Harris, J.F., 1973. Dehydration reactions of carbohydrates. Adv. Carbohydr. Chem. 28, 161e224. Fengel, D., Wegener, G., 1984. Wood: Chemistry, Ultrastructure, Reactions. Kessel Verlag, Munich, Germany, ISBN 3-935638-39-6. pp. 613. Fleche, G., Gaset, A., Gorrichon, J.-P., Truchot, E., Sicard, P., 1982. Procedé de fabrication du 5-hydroxymethylfurfural. US4339387. Frost, Sullivan, 2011. Advances in Fermentation Technologies - An Industry Overview. http://www.technicalinsights.frost.com. Gallezot, P., 2012. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 41, 1538e1558. Gandini, A., Belgacem, M.N., 1997. Furans in polymer chemistry. Prog. Polym. Sci. 22, 1203e1379. Garlock, R.J., Balan, V., Dale, B.E., Pallapolu, V.R., et al., 2011. Comparative material balances around pretreatment technologies for the conversion of switchgrass to soluble sugars. Bioresour. Technol. 102, 11063e11071. Geilen, F.M.A., Engendahl, B., Harwardt, A., Marquardt, W., Klankermayer, J., Leitner, W., 2010. Selective and flexible transformation of biomass-derived platform chemicals by a multifunctional catalytic system. Angew. Chem. 122, 5642e5646. Gellerstedt, G., Li, J., Eide, I., Kleinert, M., Barth, T., 2008. Chemical structures present in biofuel obtained from lignin. Energy Fuels 22, 4240e4244. Girio, F.M., Fonseca, C., Carvalheiro, F., Duarte, L.C., Marques, S., Bogel-Łukasik, R., 2010. Hemicelluloses for fuel ethanol: a review. Bioresour. Technol. 101, 4775e4800. Goshadrou, A., Karimi, K., Taherzadeh, M.J., 2013. Ethanol and biogas production from birch by NMMO pretreatment. Biomass Bioenergy 49, 95e101.
REFERENCES Gosselink, R.J.A., 2011. Lignin as a Renewable Aromatic Resource for the Chemical Industry, Dissertation. Wageningen University, The Netherlands, ISBN 978-94-6173-100-5. Gosselink, R.J.A., Abacherli, A., Semke, H., Malherbe, R., Kauper, P., Nadif, A., van Dam, J.E.G., 2004. Analytical protocols for characterisation of sulphur-free lignin. Ind. Crops Prod. 19, 271e281. Gosselink, R.J.A., Teunissen, W., van Dam, J.E.G., de Jong, E., Gellerstedt, G., Scott, E.L., Sanders, J.P.M., 2012. Lignin depolymerisation in supercritical carbon dioxide/acetone/water fluid for the production of aromatic chemicals. Bioresour. Technol. 106, 173e177. Gruter, G.-J.M., Dautzenberg, F., 2007. Method for the Synthesis of 5-Alkoxymethylfurfural Ethers and Their Use. Furanix Technologies B. V., The Netherlands. WO2007104514. Gruter, G.-J., de Jong, E., 2009. Furanics: novel fuel options from carbohydrates. Biofuels Technol. 1, 11e17. Gürbüz, E.I., Wettstein, S.G., Dumesic, J.A., 2012. Conversion of hemicellulose to furfural and levulinic acid using biphasic reactors with alkylphenol solvents. ChemSusChem 5, 383e387. Han, J.S., 1998. Properties of nonwood fibers. In: Proceedings of the Korean Society of Wood Science and Technology Annual Meeting. The Korean Society of Wood Science and Technology, Seoul, Korea, pp. 3e12. Hatakeyama, H., Hatakeyama, T., 2010. Lignin structure, properties and applications. Adv. Polym. Sci. 232, 1e63. Henriksson, G., Li, G., Zhang, L., Lindström, M.E., 2010. Lignin utilization. In: Crocker, M. (Ed.), Thermochemical Conversion of Biomass to Liquid Fuels and Chemicals. Royal Society of Chemistry, pp. 223e262. Hepditch, M.M., Thring, R.W., 2000. Degradation of solvolysis lignin using Lewis acid catalysts. Can. J. Chem. Eng. 78 (1), 226e231. Herrero, M., Mendiola, J.A., Cifuentes, A, Ibañeza, E., 2010. Supercritical fluid extraction: Recent advances and applications. J. Chrom. A 1217, 2495e2511. Holladay, J.E., Bozell, J.J., White, J.F., Johnson, D., 2007. Top ValueAdded Chemicals from Biomass Volume II - Results of Screening for Potential Candidates from Biorefinery Lignin. http://www. ntis.gov/ordering.htm. Hossain, M.M., Aldous, L., 2012. Ionic liquids for lignin processing: dissolution, isolation, and conversion. Aust. J. Chem. 65, 1465e1477. Hoydonckx, H.E., van Rhijn, W.M., van Rhijn, W., de Vos, D.E., Jacobs, P.A., 2007. Furfural and derivatives. Ullmann’s Encycl. Ind. Chem., 1e29. Huber, G.W., Iborra, S., Corma, A., 2006. Synthesis of transportation fuels from biomass: chemistry, catalysts, and engineering. Chem. Rev. 106, 4044e4098. Jing, Q., Lu, X.-Y., 2007. Kinetics of non-catalyzed decomposition of D-xylose in high temperature liquid water. Chin. J. Chem. Eng. 15, 666e669. Jönsson, L.J., Alriksson, B., Nilvebrant, N.O., 2013. Bioconversion of lignocellulose: inhibitors and detoxification. Biotechnol. Biofuels 6, 16e26. Kamm, B., Gruber, P.R., Kamm, M. (Eds.), 2006. Biorefineries e Industrial Processes and Products, Status Quo and Future Directions, vols 1e2. Whiley-VCH, Weinheim. Kiermayer, J., 1895. Über ein furfurolderivat aus lävulose. Chem. Ztg. 19, 1003e1006. Kim, S.B., Lee, M.R., Park, E.D., Lee, S.M., Lee, H.K., Park, K.H., Park, M.J., 2011. Kinetic study of the dehydration of D-xylose in high temperature water. React. Kinet. Mech. Catal. 103, 2267e2277. Kleinert, M., Barth, T., 2008. Phenols from lignin. Chem. Eng. Technol. 31, 736e745. Kleinert, M., Gasson, J.R., Barth, T., 2009. Optimizing solvolysis conditions for integrated depolymerisation and hydrodeoxygenation of lignin to produce liquid biofuel. J. Anal. Pyrolysis 85, 108e117. 311 Klinke, H.B., Ahring, B.K., Schmidt, A.S., Thomsen, A.B., 2002. Characterization of degradation products from alkaline wet oxidation of wheat straw. Bioresour. Technol. 82, 15e26. Kootstra, A.M.J., Mosier, N.S., Scott, E.L., Beeftink, H.H., Sanders, J.P.M., 2009. Differential effects of mineral and organic acids on the kinetics of arabinose degradation under lignocellulose pretreatment conditions. Biochem. Eng. J. 43, 92e97. Kumar, P., Barrett, D.M., Delwiche, M.J., Stroeve, P., 2009. Methods for pretreatment of lignocellulosic biomass for efficient hydrolysis and biofuel production. Ind. Eng. Chem. Res. 48, 3713e3729. Kuster, B.F.M., 1990. 5-hydroxymethylfurfural. A review focusing on its manufacture. Starch/Stärke 43, 314e321. Kuster, B.F.M., Laurens, J., 1977. Preparation of 5-hydroxymethylfurfural. Part II: dehydration of fructose in a tube reactor using polyethyleneglycol as solvent. Starch/Stärke 29, 172e176. Kuster, B.F.M., van der Steen, H.J.C., 1977. Preparation of 5-hydroxymethylfurfural. Part I: dehydration of fructose in a continuous stirred tank reactor. Starch/Stärke 29, 99e103. Lamminpää, K., Ahola, J., Tanskanen, J., 2012. Kinetics of xylose dehydration into furfural in formic acid. Ind. Eng. Chem. Res. 51, 6297e6303. Lange, J.-P., van der Heide, E., van Buijtenen, J., Price, R., 2012. Furfuralda promising platform for lignocellulosic biofuels. ChemSusChem 5, 150e166. Lichtenthaler, F.W., Peters, S., 2004. Carbohydrates as green raw materials for the chemical industry. C. R. Chim. 7, 65e90. Lignol Energy, 2013. Lignol Energy Corporation Announces First Commercial Supply Agreement, for HP-L Lignin in the Thermoplastics Sector. [Online] 1 February 2013. [Cited: 18 April 2013]. http://www.lignol.ca/news/News-2013/News%20Release% 2031Jan3013%20final.pdf. Limayem, A., Ricke, S.C., 2012. Lignocellulosic biomass for bioethanol production: current perspectives, potential issues and future prospects. Prog. Energy Combust. Sci. 38, 449e467. Lin, Y.-C., Huber, G.W., 2009. The critical role of heterogeneous catalysis in lignocellulosic biomass conversion. Energy Environ. Sci. 2, 68e80. Lloyd, T.A., Wyman, C.E., 2005. Combined sugar yields for dilute sulfuric acid pretreatment of corn stover followed by enzymatic hydrolysis of the remaining solids. Bioresour. Technol. 96, 1967e1977. Mäki-Arvela, P., Salminen, E., Riittonen, T., Virtanen, P., Kumar, N., Mikkola, J.-P., 2012. The challenge of efficient synthesis of biofuels from lignocellulose for future renewable transportation fuels. Int. J. Chem. Eng. 2012 (674761). Mamman, A.S., Lee, J.-M., Kim, Y.-C., Hwang, I.T., Park, N.-J., Hwang, Y.K., Chang, J.-S., Hwang, J.-S., 2008. Furfural: hemicellulose/xylose derived biochemical. Biofuels, Bioprod. Biorefin. 2, 438e454. Marcotullio, G., de Jong, W., 2010. Chloride ions enhance furfural formation from D-xylose in dilute aqueous acidic solutions. Green Chem. 12, 1739e1746. Marcotullio, G., de Jong, W., 2011. Furfural formation from D-xylose: the use of different halides in dilute aqueous acidic solutions allows for exceptionally high yields. Carbohydr. Res. 346, 1291e1293. Martin, C., Marcet, M., Thomsen, A.B., 2008. Comparison between wet oxidation and steam explosion as pretreatment methods for enzymatic hydrolysis of sugarcane bagasse. Bioresources 3 (3), 670e683. Mascal, M., Nikitin, E.B., 2008. Direct, high-yield conversion of cellulose into biofuel. Angew. Chem. 120, 1e4. Menon, V., Rao, M., 2012. Trends in bioconversion of lignocellulose: biofuels, platform chemicals & biorefinery concept. Prog. Energy Combust. Sci. 38, 522e550.
312 17. LIGNOCELLULOSE-BASED CHEMICAL PRODUCTS Miller, J.E., Evans, L., Littlewolf, A., Trudell, D.E., 1999. Batch microreactor studies of lignin and lignin model compound depolymerization by bases in alcohol solvents. Fuel 78, 1363e1366. Minami, E., Kawamoto, H., Saka, S., 2003. Reaction behavior of lignin in supercritical methanol as studied with lignin model compounds. J. Wood Sci. 49, 158e165. Modenbach, A.A., Nokes, S.E., 2012. The use of high-solids loadings in biomass pretreatment - a review. Biotechnol. Bioeng. 109, 1430e1442. Monteil-Rivera, F., Phuong, M., Ye, M., Halasz, A., Hawari, J., 2013. Isolation and characterization of herbaceous lignins for applications in biomaterials. Ind. Crops Prod. 41, 356e364. Mora-Pale, M., Meli, L., Doherty, T.V., Linhardt, R.J., Dordick, J.S., 2011. Room temperature ionic liquids as emerging solvents for the pretreatment of lignocellulosic biomass. Biotechnol. Bioeng. 108 (6), 1229e1245. Moreau, C., Durand, R., Razigade, S., Duhamet, J., Faugeras, P., Rivalier, P., Ros, P., Avignon, G., 1996. Dehydration of fructose to 5-hydroxymethylfurfural over H-mordenites. Appl. Catal., A 145, 211e224. Moreau, C., Durand, R., Peyron, D., Duhamet, J., Rivalier, P., 1998. Selective preparation of furfural from xylose over microporous solid acid catalysts. Ind. Crops Prod. 7, 95e99. Moreau, C., Belgacem, M.N., Gandini, A., 2004. Recent catalytic advances in the chemistry of substituted furans from carbohydrates and in the ensuing polymers. Top. Catal. 27, 11e30. Mosier, N., Wyman, C.E., Dale, B.E., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., 2005. Features of promising technologies for pretreatment of lignocellulosic biomass. Bioresour. Technol. 96, 673e686. Musau, R.M., Munavu, R.M., 1987. The preparation of 5-hydroxymethyl-2-furaldehyde (HMF) from D-fructose in the presence of DMSO. Biomass 13, 67e75. Nabarlatz, D., Ebringerová, A., Montané, D., 2007. Autohydrolysis of agricultural by-products for the production of xylooligosaccharides. Carbohydr. Polym. 69, 20e28. Nakagawa, Y., Tomishige, K., 2012. Production of 1,5-pentanediol from biomass via furfural and tetrahydrofurfuryl alcohol. Catal. Today 195, 136e143. Nakamura, Y., Morikawa, S., 1980. The dehydration of D-fructose to 5-hydroxymethyl-2-furaldehyde. Bull. Chem. Soc. Jpn. 53, 3705e3706. Newt, F.H., 1951. The formation of furan compounds from hexoses. Adv. Carbohydr. Chem. 6, 83e106. Nexant ChemSystems, 2012. Global Chemicals Demand Estimate. White Plains, New York. Nowakowski, D.J., Bridgwater, A.V., Elliott, D.C., Meier, D., de Wild, P., 2010. Lignin fast pyrolysis: results from an international collaboration. J. Anal. Appl. Pyrolysis 88 (1), 53e72. Öhman, F., Theliander, H., Tomani, P., Axegard, P., 2009. A method for separating lignin from black liquor, a lignin product, and use of a lignin product for the production of fuels or materials. WO104995. Okuda, K., Man, X., Umetsu, M., Takami, S., Adschiri, T., 2004a. Efficient conversion of lignin into single chemical species by solvothermal reaction in water-p-cresol solvent. J. Phys.: Condens. Matter 16, 1325e1330. Okuda, K., Umetsu, M., Takami, S., Adschiri, T., 2004b. Disassembly of lignin and chemical recovery - rapid depolymerization of lignin without char formation in water - phenol mixtures. Fuel Process. Technol. 85, 803e813. Okuda, K., Ohara, S., Umetsu, M., Takami, S., Adschiri, T., 2008. Disassembly of lignin and chemical recovery in supercritical water and p-cresol mixture: studies on lignin model compounds. Bioresour. Technol. 99 (6), 1846e1852. Pedersen, P., Meyer, A.S., 2010. Lignocellulose pretreatment severity e relating pH to biomatrix opening. New Biotechnol 27, 739e750. Peng, F., Peng, P., Xu, F., Sun, R.-C., 2012. Fractional purification and bioconversion of hemicelluloses. Biotechnol. Adv. 30, 879e903. Ralph, J., Brunow, G., Boerjan, W., 2007. Lignins. John Wiley & Sons, Ltd. Rapp, K.M., 1987. Process for preparing pure 5-hydroxymethylfurfuraldehyde. Süddeutsche Zucker-Aktiengesellschaft. US 4740605A1. Reid, R.C., Prausnitz, J.M., Poling, B.E., 1987. The Properties of Gases and Liquids. McGraw-Hill, New York. Roberts, V., Stein, V., Reiner, T., Lemonidou, A., Li, X., Lercher, J., 2011. Towards quantitative catalytic lignin depolymerization. Chem. Eur. J. 17, 5939e5948. Román-Leshkov, Y., Dumesic, J.A., 2009. Solvent effects on fructose dehydration to 5-hydroxymethylfurfural in biphasic systems saturated with inorganic salts. Top. Catal. 52, 297e303. Román-Leshkov, Y., Chheda, J.N., Dumesic, J.A., 2006. Phase modifiers promote efficient production of hydroxymethylfurfural from fructose. Science 312, 1933e1937. Rong, C., Ding, X., Zhu, Y., Li, Y., Wang, L., Qu, Y., Ma, X., Wang, Z., 2012. Production of furfural from xylose at atmospheric pressure by dilute sulfuric acid and inorganic salts. Carbohydr. Res. 350, 77e80. Root, D.F., Saeman, J.F., Harris, J.F., Neill, W.K., 1959. Chemical conversion of wood residues, part II: kintetics of the acid-catalyzed conversion of xylose to furfural. For. Prod. J. 9, 158e165. Ruiz, H.A., Rodriguez-Jasso, R.M., Fernandes, B.D., Vicente, A.A., Teixeira, J.A., 2013. Hydrothermal processing, as an alternative for upgrading agriculture residues and marine biomass according to the biorefinery concept: a review. Renewable Sustainable Energy Rev. 21, 35e51. Sahu, R., Dhepe, P.L., 2012. A one-pot method for the selective conversion of hemicellulose from crop waste into C5 sugars and furfural by using solid acid catalysts. ChemSusChem 5, 751e761. Sain, B., Chaudhuri, A., Borgohain, J.N., Baruah, B.P., Ghose, J.L., 1982. Furfural and furfural-based industrial chemicals. J. Sci. Ind. Res. 41, 431e438. Sako, T., Sugeta, T., Nakazawa, N., Okubo, T., Sato, M., Taguchi, T., Hiaki, T., 1991. Phase equilibrium study of extraction and concentration of furfural produced in a reactor using supercritical carbon dioxide extraction. J. Chem. Eng. Jpn. 24, 449e455. Sako, T., Sugeta, T., Nakazawa, N., Okubo, T., Sato, M., Taguchi, T., Hiaki, T., 1992. Kinetic study of furfural formation accompanying supercritical carbon dioxide extraction. J. Chem. Eng. Jpn. 25, 372e377. Sales, F.G., Abreu, C.A.M., Pereira, J.A.F.R., 2004. Catalytic wet-air oxidation of lignin in a three-phase reactor with aromatic aldehyde production. Braz. J. Chem. Eng. 21 (2), 211e218. Sales, F.G., Maranhão, L.C.A., Filho, N.M.L., Abreu, C.A.M., 2007. Experimental evaluation and continuous catalytic process for fine aldehyde production from lignin. Chem. Eng. Sci. 62, 5386e5391. Schädel, C., Blöchl, A., Richter, A., Hoch, G., 2010. Quantification and monosaccharide composition of hemicelluloses from different plant functional types. Plant Physiol. Biochem. 48, 1e8. Schinzer, D., Bourguet, E., Ducki, S., 2004. Synthesis of furanoepothilone D. Chem. Eur. J. 10, 3217e3224. Sinha, A.K., Kumar, V., Makkar, H.P.S., De Boeck, G., Becker, K., 2011. Non-starch polysaccharides and their role in fish nutrition e a review. Food Chem. 127, 1409e1426. Speck Jr, J.C., 1958. The Lobry de Bruyn- Alberda van Ekenstein transformation. Adv. Carbohydr. Chem. 13, 63e103. Spiridon, I., Popa, V.I., 2008. Hemicelluloses: major sources, properties and applications. In: Belgacem, N.M., Gandini, A. (Eds.), Monomers, Polymers and Composites from Renewable Resources. Elsevier, Amsterdam, pp. 289e304.
REFERENCES Stark, K., Taccardi, N., Bosmann, A., Wasserscheid, P., 2010. Oxidative depolymerisation of lignin in ionic liquids. ChemSusChem. 3, 719e723. Stöcker, M., 2008. Biofuels and biomass-to-liquid fuels in the biorefinery: catalytic conversion of lignocellulosic biomass using porous materials. Angew. Chem. Int. Ed. 47, 9200e9211. Suparno, O., Dovington, A.D., Evans, C.S., 2005. Kraft lignin degradation products for tanning and dyeing of leather. J. Chem. Technol. Biotechnol. 80, 44e49. Szmant, H.H., Chundury, D.D., 1981. The preparation of 5-hydroxymethylfurfuraldehyde from high fructose corn syrup and other carbohydrates. J. Chem. Tech. Biotechnol. 31, 135e145. Tejado, A., Pena, C., Labidi, J., Echeverria, J.M., Mondragon, I., 2007. Physico-chemical characterization of lignins from different sources for use in phenol-formaldehyde resin synthesis. Bioresour. Technol. 98, 1655e1663. Thring, R.W., 1994. Alkaline degradation of ALCELLÒ lignin. Biomass Bioenergy 7, 125e130. Toledano, A., Serrano, L., Labidi, J., 2012. Organosolv lignin depolymerization with different base catalysts. J. Chem. Technol. Biotechnol. 87, 1593e1599. Van Haveren, J., Scott, E.L., Sanders, J.P.M., 2008. Review: bulk chemicals from biomass. Biofuels Bioprod. Biorefin. 2, 41e57. Van Putten, R.-J., van der Waal, J.C., de Jong, E., Rasrendra, C.B., Heeres, H.J., de Vries, J.G., 2013a. Hydroxymethylfurfural, a versatile platform chemical made from renewable resources. Chem. Rev. 113, 1499e1597. Van Putten, R.-J., Dias, A.S., de Jong, E., 2013b. Furan based building blocks from carbohydrates. In: Imhof, P., van der Waal, J.C. (Eds.), Catalytic Process Development for Renewable Materials. WileyVCH Verlag GmbH & Co., pp. 81e118. Vishtal, A., Kraslawski, A., 2011. Challenges in industrial applications of technical lignins. Bioresources 6, 1e22. Vispute, T.P., Zhang, H., Sanna, A., Xiao, R., Huber, G.W., 2010. Renewable chemical commodity feedstocks from integrated catalytic processing of pyrolysis oils. Science 330, 1222e1227. Vlachos, D.G., Chen, J.G., Gorte, R.J., Huber, G.W., Tsapatsis, M., 2010. Catalysis center for energy innovation for biomass processing: research strategies and goals. Catal. Lett. 140, 77e84. Voitl, T., Rudolf von Rohr, P., 2010. Demonstration of a process for the conversion of kraft lignin into vanillin and methyl vanillate by acidic oxidation in aqueous methanol. Ind. Eng. Chem. Res. 49, 520e525. vom Stein, T., Grande, P.M., Leitner, W., Domı́nguez de Marı́a, P., 2011. Iron-catalyzed furfural production in biobased biphasic systems: from pure sugars to direct use of crude xylose effluents as feedstock. ChemSusChem 4, 1592e1594. Wahyudiono, Sasaki, M., Goto, M., 2008. Recovery of phenolic compounds through the decomposition of lignin in near and supercritical water. Chem. Eng. Process.: Process Intensif. 47 (9e10), 1609e1619. Weingarten, R., Cho, J., Conner Jr, W.C., Huber, G.W., 2010. Kinetics of furfural production by dehydration of xylose in a biphasic reactor with microwave heating. Green Chem. 12, 1423e1429. Werhan, H., Farshori, A., Rudolf von Rohr, P., 2012. Separation of lignin oxidation products by organic solvent nanofiltration. J. Membr. Sci. 423e424, 404e412. West, R.M., Liu, Z.Y., Peter, M., Gartner, C.A., Dumesic, J.A., 2008. Carbonecarbon bond formation for biomass-derived furfurals and ketones by aldol condensation in a biphasic system. J. Mol. Catal. A: Chem. 296, 18e27. Wettstein, S.G., Alonso, D.M., Gürbüz, E.I., Dumesic, J.A., 2012. A roadmap for conversion of lignocellulosic biomass to chemicals and fuels. Curr. Opin. Chem. Eng. 1, 218e224. Win, D.T., 2005. Furfural e gold from garbage. Aust. J. Technol. 8, 185e190. 313 Wyman, C.E., Dale, B.E., Elander, R.T., Holtzapple, M., Ladisch, M.R., Lee, Y.Y., 2005. Coordinated development of leading biomass pretreatment technologies. Bioresour. Technol. 96, 1959e1966. Wyman, C.E., Balan, V., Dale, B.E., Elander, R.T., et al., 2011. Comparative data on effects of leading pretreatments and enzyme loadings and formulations on sugar yields from different switchgrass sources. Bioresour. Technol. 102, 11052e11062. Xiang, Q., Lee, Y.Y., 2000. Oxidative cracking of precipitated hardwood lignin by hydrogen peroxide. Appl. Biochem. Biotechnol. 84e86, 153e162. Xing, R., Qi, W., Huber, G.W., 2011. Production of furfural and carboxylic acids from waste aqueous hemicellulose solutions from the pulp and paper and cellulosic ethanol industries. Energy Environ. Sci. 4, 2193e2205. Xu, W., Miller, S.J., Agrawal, P.K., Jones, C.W., 2012. Depolymerization and hydrodeoxygenation of switchgrass lignin with formic acid. ChemSusChem 5 (4), 667e675. Yamashita, Y., Shono, M., Sasaki, C., Nakamura, Y., 2010. Alkaline peroxide pretreatment for efficient enzymatic saccharification of bamboo. Carbohydr. Pol 79, 914e920. Yang, H., Yan, R., Chen, H., Lee, D., Zheng, C., 2007. Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel 86, 1781e1788. Yang, W., Li, P., Bo, D., Chang, H., 2012. The optimization of formic acid hydrolysis of xylose in furfural production. Carbohydr. Res. 357, 53e61. Ye, Y., Zhang, Y., Fan, J., Chang, J., 2012. Novel method for the production of phenolics by combining lignin extraction with lignin depolymerisation in aqueous ethanol. Ind. Eng. Chem. Res. 51, 103e110. Yemis, O., Mazza, G., 2011. Acid-catalyzed conversion of xylose, xylan and straw into furfural by microwave-assisted reaction. Bioresour. Technol. 102, 7371e7378. Yemis, O., Mazza, G., 2012. Optimization of furfural and 5-hydroxymethylfurfural production from wheat straw by a microwave-assisted process. Bioresour. Technol. 109, 215e223. Yoo, C.G., Kuo, M., Kim, T.H., 2012. Ethanol and furfural production from corn stover using a hybrid fractionation process with zinc chloride and simultaneous saccharification and fermentation (SSF). Process Biochem. 47, 319e326. Yoshikawa, T., Yagia, T., Shinoharaa, S., Fukunagab, T., Nakasakaa, Y., Tagoa, T., Masudaa, T., 2013. Production of phenols from lignin via depolymerization and catalytic cracking. Fuel Proc. Technol. 108, 69e75. Yuan, Z., Cheng, S., Leitch, M., Xu, C., 2010. Hydrolytic degradation of alkaline lignin in hot-compressed water and ethanol. Bioresour. Technol. 101, 9308e9313. Zakrzewska, M.E., Bogel-Łukasik, E., Bogel-Łukasik, R., 2010. Ionic liquid-mediated formation of 5-hydroxymethylfurfurals. A promising biomass-derived building block. Chem. Rev. 111, 397e417. Zakzeski, J., Bruijnincx, P.C., Jongerius, A.L., Weckhuysen, B.M., 2010. The catalytic valorization of lignin for the production of renewable chemicals. Chem. Rev. 110 (6), 3552e3599. Zakzeski, J., Jongerius, A.L., Bruijnincx, P.C.A., Weckhuysen, B.M., 2012. Catalytic lignin valorization process for the production of aromatic chemicals and hydrogen. ChemSusChem 5, 1602e1609. Zeitsch, K.J., 2000a. The chemistry and technology of furfural and its many by-products. Sugar Series 13. Elsevier, The Netherlands. Zeitsch, K.J., 2000b. Furfural production needs chemical innovation. Chem. Innovation 30, 29e32. Zhao, H., Holladay, J.E., Brown, H., Zhang, Z.C., 2007. Metal chlorides in ionic liquid solvents convert sugars to 5-hydroxymethylfurfural. Science 316, 1597e1600. Zhao, X., Zhang, L., Liu, D., 2012. Biomass recalcitrance. Part II: fundamentals of different pre-treatments to increase the enzymatic digestibility of lignocellulose. Biofuels, Bioprod. Biorefin. 6, 561e579.
C H A P T E R 18 Industrial Lignins: Analysis, Properties, and Applications Alex Berlin 1,*, Mikhail Balakshin 2 1 Novozymes, Protein Chemistry Department, Davis, CA, USA, 2Renmatix, R&D Department, King of Prussia, PA, USA *Corresponding author email: axbl@novozymes.com O U T L I N E The Potential of Technical Lignins as a Renewable 315 Raw Material Feedstock Technical Lignins: Production, Properties, and Analysis Comparison of Analytical Methods for Characterization of Technical Lignins Reproducibility of 31P NMR Analytical Techniques 13 C NMR Analysis of Technical Lignins Advanced NMR Methods Molecular Weight Distribution 318 323 323 327 330 330 THE POTENTIAL OF TECHNICAL LIGNINS AS A RENEWABLE RAW MATERIAL FEEDSTOCK Lignins in their native form are the most abundant renewable aromatic polymers on earth (Kirk and Farrell, 1987). Consequently, lignins present great potential as a source of energy due to their high fuel content (26e28 MJ/ton dry lignin) rivaling the fuel content of some coals (Lora, 2006; Tomani, et al., 2011). Lignins can be combusted to produce “green” electricity, power, fuel, steam, or syngas; all these are forms of energy which are being or will be used in the future to operate industrial plants where lignins are generated as byproducts. The lignin by-products are called “technical lignins” or “industrial lignins” and they differ dramatically in properties from the native lignins found in plants. Examples of the use of technical lignins as a source of energy to run industrial plants are the pulp Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00018-8 StructureeProperties Correlations in Lignin 331 Technical Lignins: Traditional and Emerging Applications Traditional Lignin Applications Emerging Lignin Applications 332 332 332 Conclusions 333 References 333 mills deployed worldwide and the emerging lignocellulose biorefineries. Energy production is, compared to all other technical lignins applications, the one with the lowest market value, estimated at approximately 10 US$ cents/kg as coal replacement (Holladay et al., 2007). However, energy generation is the lignin application with the highest demand by volume and currently the one with the lowest technical risk. Almost every major pulp chemical mill today utilizes lignin as a source of energy. The latter is today’s common industrial practice which will likely be mirrored by future cellulosic biomass biorefineries which will use lignin as the main energy source in combination with other fuels such as raw biomass. Technical lignins are available in large volumes, primarily in kraft mill spent liquors (“black liquors”), and, to a less extent, in the spent liquors of the few remaining sulfite mills (“brown liquors”). According to our conservative estimate, ca. 6e7% of the spent liquor 315 Copyright Ó 2014 Elsevier B.V. All rights reserved.
316 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS produced at a kraft pulp mill could be used for lignin extraction without significantly affecting the plant energy balance. This represents a potential average lignin production capacity per plant in the order of 30e75 tons of lignin per kraft pulp plant per day (Domtar, 2013) assuming an average annual pulp production capacity of ca. 0.5 million tons odw pulp (Table 18.1). On the contrary, in sulfite pulp mills, the majority of the produced spent liquor can be used for TABLE 18.1 Chemical Pulp Production Capacity of Typical Pulp and Paper Mills Corporation* Plant Location Pulp Production Capacity (1000 odw tons per year) Domtar (USA/ Canada) Ashdown (AR, USA) 747 Dryden (Ontario, Canada) 328 Espanola (Ontario, Canada) 325 Hawesville (KY, USA) 430 Kamloops (BC, Canada) 380 Marlboro (SC, USA) 278 Plymouth (NC, USA) 438 Windsor (QC, Canada) 447 Fray Bentos (Uruguay) 1100 Kymi (Finland) 570 Pietarsaari (Finland) 790 Kaukas (Finland) 740 Husum (Sweden) 700 Joutseno (Finland) 670 Kaskinen (Finland) 300 Kemi (Finland) 590 Rauma (Finland) 630 Äänekoski (Finland) 520 Licantén (Chile) 140 Constitución (Chile) 355 Ránquil (Chile) 1027 Arauco (Chile) 790 San José de la Mariquina (Chile) 550 Misiones Province (Argentina) 350 Average production capacity (1000s odw tons pulp/year) 550 UPM (Finland) Metsä-Botnia (Finland) Arauco (Chile) *The information contained in this table was obtained from the respective corporate websites. lignosulfonate production given the fact that not many sulfite pulping players burn lignosulfonates for energy generation. In 2004, it was reported that 2% of the lignin available in the pulp and paper industry was commercially used comprising about 1,000,000 tons/year lignosulfonates from sulfite pulping brown liquors and <100,000 tons/year from kraft spent liquors (Gosselink et al., 2004). Assuming the forecasted annual growth rate of 2.5% (IHS-Chemical, 2012), the current lignosulfonates production should be ca. 1,200,000 tons/year. The global annual production of chemical pulps is estimated at 150 million tons/year (Vappula, 2011) with an average odw chemical pulp production per plant of ca. 0.5 million tons (Table 18.1) and a total volume of dissolved lignin in pulp making of ca. 70 million tons globally (Lora, 2010). Assuming a lignin annual production capacity per chemical pulp mill of ca. 27,000 metric tons kraft lignin per year (Domtar, 2013), it can then be concluded that the estimated annual global potential for kraft lignin production capacity from pulp mills is ca. 6e9 million tons depending on the feedstock species and the plant design which is in good agreement with the estimates reported by other authors (Glasser, 2010; Lora, 2010). When considering the potential replacement of the main petrochemicals, namely, ethylene, propylene, butadiene and benzene, toluene, and xylene (BTX) isomers, which are produced at a rate of about 300 million tons/year with a total value of over $400 billion (Lucintel, 2013), by lignin it is relevant to consider the current global potential production capacity of technical lignins. It becomes clear from the analysis performed above that in the best case scenario purified technical lignins produced globally at chemical pulp mills could potentially replace a maximum of ca. 2% of the global volume of main petrochemicals. However, the emerging lignocellulose biorefinery industry for production of biofuels and chemicals might completely change this picture. For instance, the US Department of Energy estimates that 1.3 billion tons of biomass is available in the United States alone for biorefining into transportation fuels and chemicals (Perlack et al., 2005). This amount of biomass could make available additionally 225 million tons of lignin which could be utilized for power, transportation fuels, products and various combinations of the above (Holladay et al., 2007). Assuming that 20% of this biorefinery lignin (45 million tons) will be converted into BTX and linear hydrocarbons, the result could be ca. 10% replacement of these petrochemicals by lignins produced at American biomass biorefineries alone. In other words, if we consider, in addition to the American biorefineries, a scenario where these biomass biorefining technologies will be deployed globally, we could witness, in the future, a hypothetical situation where a large fraction of petroleum-derived BTX, and perhaps of other petrochemicals too, could be replaced by lignin.
THE POTENTIAL OF TECHNICAL LIGNINS AS A RENEWABLE RAW MATERIAL FEEDSTOCK Another abundant source of technical lignins, which is often ignored, is the acid-hydrolysis (AH) lignin (“hydrolysis lignin”) which has been produced at Eastern European wood and agricultural wastes AH plants since the mid-1930s with yields in the range of 350e400 kg lignin/ton odw softwood. The annual production of such hydrolysis lignin in the former Soviet Union reached 1.5 million tons by the end of the 1980s. However, only 30e40% of the hydrolysis lignin was really utilized, whereas the rest was disposed in giant landfills nearby the wood hydrolysis plants, creating as a result serious environmental problems caused primarily by autoignition of these deposits. For example, the current lignin waste stocks in the Irkutsk region (Siberia), where only four plants are located, exceed 20 million tons (Rabinovich, 2010) equivalent to ca. 20 times today’s global commercial technical lignin market. The current annual production of AH lignin in Belarus alone is in the order of 100,000 tons (Podterob et al., 2004). The main application of the hydrolysis lignin is the production of pellets for energy generation. However, highly specialized applications, such as pharmaceutical enterosorbents, have been successfully developed and commercialized on the basis of purified hydrolysis lignin. An example of these commercial sorbents is the enterosorbent “Polyphepan” (Podterob et al., 2004). In addition to the low-value energy lignin application, a wide diversity of high-value industrial applications have been envisioned or industrially realized or demonstrated including uses as novel materials, polymeric, oligomeric, and monomeric feedstock. Some of these opportunities, such as the use of lignin or its derivatives in animal feed additives, agriculture, construction, textile, oil drilling, binders, dispersants, and composites, are today commercial realities but many others such as the production of carbon fiber precursors, the broad incorporation of lignin in synthetic polymeric blends, or the production of BTX remain longer term opportunities with great value and market potential. Both low- and high-value lignin applications are often seen as efficient vehicles to increase the productivity, reduce fossil fuel consumption, and increase the profitability of the industrial plants where lignin is produced as a by-product. For instance, the LignoboostÔ process (Tomani, 2010), a recently commercialized process by Metso Corporation (Helsinki, Finland) for lignin production from alkaline black liquors, significantly improves the profitability of the pulp and paper mill by debottlenecking the wood pulp production as a result of increasing the recovery capacity of pulping chemicals and valorizing the lignin stream. Commercial-scale lignin production based on the LignoboostÔ process has begun in February 2013 by Domtar Corp. at the Plymouth Mill (NC, USA) with a targeted rate of 75 tons/ 317 day (w27,000 tons/year), destined for a wide range of industrial applications as a bio-based alternative to the use of petroleum and other fossil fuels (Domtar, 2013). As it was mentioned earlier, in the case of emerging industries, such as the cellulosic ethanol industry, the smart utilization of residual lignin could dramatically boost the profitability of the cellulosic biofuel plants if converted into value-added chemicals such as BTX, other monomeric and oligomeric phenolic compounds, and suitable for material applications macromolecules such as carbon fiber precursors, polymeric blends, adhesives, dispersants, and others. While energy and monomeric applications for technical lignins and their derivatives often target direct replacements of fossil fuels and petrochemicals, the development of novel lignin-derived oligomeric and macromolecular entities has the potential of generating better alternatives or synergy with petrochemical feedstocks. Recent examples of the latter have been reported in the literature which illustrates this concept. For instance, recently Berlin (2011) showed that the replacement of methylene diphenyl diisocyanate (MDI) in engineered wood diisocyanate adhesives by organosolv (OS) lignin derivatives can lead to substantial improvements of the adhesive binding properties (increased modulus of rupture and modulus of elasticity) when applying the lignineMDI adhesives in engineered wood composite construction materials such as Oriented Strand Boards (OSB) while still meeting the industry standard requirements for these adhesives. A similar observation was documented when phenol in phenoleformaldehyde resins was replaced by OS lignin derivatives which resulted in a significant increase of the resin normalized bond strength (Berlin, 2012b). These two examples are important because they illustrate a fact often overseen which is the evidence that lignin derivatives can technically outperform petrochemicals when used in conjunction with the latter in certain chemical formulations. This observation hints at the possibility of not needing to completely depolymerize lignin, a longstanding unresolved challenging technical problem, into the equivalent petrochemical monomers in order to achieve similar or better performance of the lignin-derived chemicals in formulated products. On the contrary, further research efforts could be directed toward valorization strategies of technical lignins with preserve natural backbone structures to produce viable novel polymeric precursors alternative to petrochemicals. The recent resurgence of interest in lignin as a renewable raw material feedstock is evidenced by the growing number of patent applications containing the word “lignin” which have been filed between 2003 and 2012 via the World Intellectual Property Organization (WIPO; Figure 18.1). It is interesting to note the fact
318 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS FIGURE 18.1 Number of WIPO patent applications containing the word “lignin” found in May 2013 by using the WIPO search tool Patentscope for the period 2003e2012. that 25% of all these patent applications were filed by major chemical, pharmaceutical, and energy companies such as BASF, Bayer, Ciba-Geigy, Monsanto, Sumitomo, and Shell with the German chemical giants Bayer (10% lignin filings) and BASF (8% lignin filings) leading the group. The patents found in May 2013 by using the WIPO search tool Patentscope for the period 2003e 2012 (25,974 documents) represent ca. 50% of all the patents registered in the WIPO which contain the word “lignin” on the front page (52,895 documents). Today’s global lignin market is dominated by the Norwegian company Borregaard Lignotech (Norway) followed by Tembec (North AmericaeFrance) and MeadWestvaco (USA). There are a number of smaller players such as Domsjö (Finland-India), Granit SA (Switzerland), and CIMV (France), among others. TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS A detailed understanding of the structure of technical lignins is critically important in order to direct the efforts toward their valorization (Glasser, 2000). Not surprisingly, there is in the literature significant evidence suggesting that the performance of purified technical lignins can be linked to their chemical structure (Gosselink et al., 2010; Berlin, 2011, 2012a,b; Chung and Washburn, 2012). However, it is recognized that there is a fundamental lack of knowledge in the understanding of technical lignins as a polymer and their conversion to materials, so targeted modifications via refining, chemical modifications, or fractionation can be pursued to maximize their performance in formulated products (Baker and Rials, 2013). Hence, the importance of the lignin analytical methods employed to study the structure of these lignins will be discussed in detail below. Native lignin is an irregular heterogeneous polymer. The same applies to technical lignins with the particularity that the lignin heterogeneity is typically increased by the biomass processing. It is widely believed that the lignin structure is tridimensional; however, there is no solid evidence supporting this hypothesis. Some scientists question the latter claim (Ralph et al., 2004). Lignin is optically inactive. The repeated (monomeric) unit in lignin is the phenylpropane unit (or so-called the “C9-unit”) of the p-hydroxyphenyl (H), guaiacyl (G) and syringyl (S) types (Figure 18.2). Coniferous lignins are predominantly of G-type. Hardwood lignins contain both G- and S-units. The H-unit content in woody lignin is usually low; however, they can significantly contribute to the structure of nonwoody lignins, for instance, from annual fibers. In addition, annual fiber lignins contain significant amounts of cinnamic and ferrulic acid derivatives attached to lignin predominantly via ester linkages with the gamma hydroxyl of C9-units (Adler, 1977; Sakakibara, 1991; Ralph et al., 2004). The lignin C9-units contain different functional groups. The most common ones are aromatic methoxyl and phenolic hydroxyl, primary and secondary aliphatic hydroxyl, small amounts of carbonyl groups (of the aldehyde and ketone types) and carboxyl groups. The monomeric C9 lignin units are linked to form a polymer by CeOeC and CeC linkages. The most abundant lignin interunit linkage is the b-O-4 type of linkage (structures 1e4, and 7; Figure 18.2) comprising about 50% of the interunit linkages in lignin (ca. 45% in softwoods and up to 60e65% in hardwoods). Other common lignin interunit linkages are the resinol (b-b) (structure 6; Figure 18.2), phenylcoumaran (b-5) (structure 5; Figure 18.2), 5-50 (structure 12; Figure 18.2), and
TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS 319 FIGURE 18.2 Structural moieties of native lignins (1e20), technical lignins and major ligninecarbohydrate bonds (AeC). 4-O-5 (structure 11; Figure 18.2) moieties. The number of these structures varies in different lignins, but rarely exceeds 10% of the total lignin moieties. The number of other lignin moieties is usually below 5% (Adler, 1977; Sakakibara, 1991; Balakshin et al., 2008). The degree of lignin condensation (DC) is an important lignin characteristic as it is often correlated (negatively) with lignin reactivity. The definition of condensed lignin moieties found in the literature is not always clear. Most commonly, condensed lignin structures are lignin moieties linked to other lignin units via 2, 5 or 6 positions of the aromatic ring (in H-units also C-3 position). The most common condensed structures are 5-50 , b-5, and 4-O-50 structures. Since the C-5 position of the syringyl aromatic ring is occupied by a methoxyl group and therefore it cannot be involved in condensation, hardwood lignins are less condensed than softwood lignins. According to recent findings, almost all lignin in softwood and softwood pulps is linked to polysaccharides, mainly via hemicelluloses (Lawoko et al., 2005). The main types of ligninecarbohydrate (LC) linkages in
320 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS wood are phenyl glycoside bonds (A), esters (B) and benzyl ethers (C) (Figure 18.2; Helm, 2000; Koshijima and Watanabe, 2003; Balakshin et al., 2007; Balakshin et al., 2011; Balakshin et al., 2014). The occurrence of stable LC bonds in native lignins is one of the main reasons preventing selective separation of the wood components in biorefining processes. Technical lignins are obtained as a result of lignocellulosic biomass processing. Technical lignins differ dramatically from the corresponding native ones as a result of a combination of multiple reactions including catalyzed biomass hydrolysis, condensation of lignin fragments, elimination of native lignin functional groups, formation of new functional groups, and others. They are appreciably more heterogeneous (in terms of chemical structure and molecular mass) than the native lignins. Technical lignins have a high variety of structural moieties present in rather small amounts (Balakshin et al., 2003; Liitiä et al., 2003). Technical lignins can be classified from different points of view (Table 18.2). From a practical point of view, there are technical lignins originated from pulp and paper industrial processes which are considered mostly waste products without controllable chemical properties. These are kraft and soda lignins (kraft and soda pulping, correspondingly) and lignosulfonates (sulfite pulping). On the other hand, there is a big group of technical lignins from various emerging biomass biorefining processes such as different variations of AH, steam explosion (SE), and OS pretreatment, in particular. In terms of the process chemistry, and, correspondingly, the lignin chemical structure, lignins can be derived from acid- or alkali-based processes. The former includes most of the emerging biomass biorefinery pretreatments, such as AH, SE (except AFEX) and most of OS processes as well as lignosulfonates. Alkaline processes are kraft and soda pulping, AFEX pretreatment, and some OS processes. In addition to the process nature, the feedstock source has naturally an important impact on the structure of technical lignins. TABLE 18.2 Another consideration which can be used to classify technical lignins, especially in view of their application, is the presence or absence of sulfur in their structure. Therefore, kraft lignin, and, especially lignosulfonates, are sulfur-containing lignins whereas soda, OS, AH and SE lignin are sulfur-free or low-sulfur-containing lignins. In terms of the chemical structure, native lignins undergo significant degradation/modification during biomass processing. Lignin degradation occurs predominantly via cleavage of b-O-4 linkages (although the mechanisms are different for different processes), which results in an increase of phenolic hydroxyls and a decrease in lignin molecular mass. The lignin degradation also leads to a decrease in aliphatic hydroxyls, oxygenated aliphatic moieties and the formation of carboxyl groups and saturated aliphatic structures. In contrast to lignin degradation, some reversed reactions, such as lignin repolymerization/condensation, take place to some degree resulting in increase of lignin molecular mass and decrease of its reactivity. These changes are common for most of the technical lignins although the degree of transformation varies significantly depending on the process conditions (temperature, time, pH, and others) and feedstock origin. Each process provides the lignin with specific chemical characteristics. First, the reaction mechanism is quite different in acidic and basic media. The cleavage of b-O4 linkages under alkaline conditions occurs via a quinone methide intermediate which results in the formation of coniferyl alcohol-type moieties as a primary reaction product (Figure 18.3). They are not accumulated in the lignin; however, they undergo further secondary rearrangement reactions forming various (aryl-) aliphatic acids. b-5 and b-1 type of linkages of the native lignin cannot be cleaved during the process but are transformed into stilbene-type structures (structure 30; Figure 18.2). The latter are stable and are accumulated in alkaline lignins. In addition, a significant amount of vinyl ether structures (structure 29; Figure 18.2) forms Classification of Technical Lignins Lignin Type Scale Chemistry Sulphur content Purity Kraft Industrial Alkaline Moderate Moderate Soda Industrial Alkaline Free Moderate-low Lignosulfonate Industrial Acid High Low Organosolv Pilot/demo Acid Free High Hydrolysis Industrial/pilot Acid Low-free Moderate-low SE Demo/pilot Acid Low-free Moderate-low AFEX Pilot Alkaline Free Moderate-low SE, steam-exploded lignin; AFEX, ammonia fiber expansion lignin.
TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS FIGURE 18.3 321 The main lignin reactions in alkaline pulping. during soda pulping and accumulates in lignin, in contrast to kraft lignin. Another relevant structural difference between soda and kraft lignins, resulting from differences in the reaction mechanism, is the presence of aryl-glycerol type structures (structure 20; Figure 18.2) in the former. On the other hand, lignin undergoes demethylation reactions which result in formation of o-quinone structures during kraft pulping (but not in the case of soda pulping). In addition, kraft lignins contain small amounts of organically bound sulfur, likely in the form of thiol compounds (Marton, 1971; Gierer, 1980; Gellerstedt, 1996; Balakshin et al., 2003). Kraft and soda lignins show significantly higher degree of condensation than the corresponding native lignins. However, this is the result of accumulation of condensed moieties of original native lignin rather than the result of extensive condensation reactions during pulping (Balakshin et al., 2003). Kraft and soda lignins contain small amounts of carbohydrate and ash impurities. The amounts of these contaminants are dependent on feedstock origin and are significantly higher in annual fiber lignins than in woody lignins. The lignin chemistry originated from the emerging acid-based biomass biorefinery processes is very diverse (Glasser et al., 1983). The acid-based biomass biorefining can be catalyzed by addition of mineral or organic acids (from catalytic amounts to the use of organic acids as the reaction media) or without acid addition (autohydrolysis) when organic acids are generated due to cleavage of acetyl groups of lignocellulosics as well as due to the formation of acidic reaction products. Technical lignins derived from biomass biorefinery processes have been much less investigated than kraft lignins. Moreover, a high diversity of lignins is expected in the future given the large number of technical biomass pretreatment processes under either R&D or industrial deployment and the high variety of potential raw materials (softwoods, hardwoods, annual fibers, agricultural residues, etc.) as compared to the relative uniformity of pulping processes. The main pathway of lignin degradation under acidic conditions is the acidic hydrolysis of b-O-4 linkages (Figure 18.4). The major products of this reaction are the so-called Hibbert ketones (Wallis, 1971). The accumulation of Hibbert ketones in lignin results in relatively high content, as it compares to alkaline lignins, of carbonyl groups and the corresponding saturated aliphatic structures (Berlin et al., 2006). Although degradation of lignin under acidic condition occurs via vinyl ether intermediates, they do not accumulate in the lignin since vinyl ether structures are very unstable in acid media. Significant amounts of olefinic moieties were observed in lignin obtained under acidic conditions, but their nature is different from the olefinic structures of kraft and soda lignins, their exact structure is still not well understood (Berlin et al., 2006). Moreover, lignin condensation reactions under acidic conditions are more significant than those occurring in alkaline processes. Acidic lignin condensation occurs predominantly via 2 and 6 positions of the aromatic ring, in contrast to alkaline condensation which occurs predominantly at the C-5 position of the aromatic ring (Glasser et al., 1983). The DC is dependent on the acidity of the reaction media (pH and solvent used) and the process
322 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS FIGURE 18.4 Major lignin degradation reactions under acidolysis conditions (in acidic aqueous organic solvent). The cleavage of b-O-4 structures results in the formation of free phenolic moieties and Hibbert ketones. severity (temperature and time). An extreme example of highly modified technical lignins is the industrial AH lignin produced in Russia or Belarus which is obtained at 170e190  C, 2e3 h with 1% H2SO4. AH lignin is insoluble in polar organic solvents and alkaline solutions due to the strong condensation/polymerization occurred during the AH process. Hydrolysis lignin has high content of phenolic hydroxyl groups and olefinic structures. In addition, it contains 10e30% residual carbohydrates and up to 20% lipophilic extractives (Chudakov, 1983). In contrast, a significant fraction of AH lignin obtained at very high reaction temperature (>220  C) but short reaction time (<1 min) was soluble in 1 N NaOH (70e90% of AH lignin) and dioxane (50%); the amounts of carbohydrates in these soluble lignins were significantly lower, 2e4% (Glasser et al., 1983; Lora and Glasser, 2002). SE lignin is also quite degraded in terms of cleavage of b-O-4 linkages, but apparently much less condensed than AH lignins (Glasser et al., 1983; Robert et al., 1988; Li et al., 2009). OS methods of lignin production are very diverse, where different organic solvents and reaction pH can be varied during the process. Each of these processes produces lignins that are very different in their physicochemical and biochemical properties. Furthermore, the physical conditions (e.g. temperature, time, and pressure) and chemical conditions (e.g. pH and concentration of solvents) under which these processes are conducted can drastically affect the molecular weight (Mw), chemical structure, and functional groups distribution in the generated lignin derivatives (Abdelkafi et al., 2011; Balakshin et al., 2013a, b). The most investigated OS delignification technology is the Alcell process deployed at industrial scale in Eastern Canada during the 1980s. The Alcell process can be carried out in aqueous ethanol liquor at moderate acidity (no exogenous acid is added; the acidic pH results from formation of organic acids during the process). Alcell lignin is practically sulfur-free and has significantly lower amounts of carbohydrate and ash impurities compared to kraft lignins. Lignosulfonates are a special class of technical lignins and they constitute the bulk of the commercial lignins for materials and chemical applications. Lignosulfonates are primarily isolated from sulfite spent liquors. However, sulfonated or sulfomethylated kraft lignins are often used in similar to lignosulfonates applications but they have different chemical properties, in particular the sulfonic acid groups in lignosulfonates are located in the side alkyl chains, whereas in sulfonated kraft lignins they are found in the aromatic ring. Sulfur-containing lignins, in form of lignosulfonates or sulfomethylated kraft lignins, are commercialized by Borregaard (Sarpsborg, Norway), MeadWestvaco (Richmond, VA, USA), Tembec (Montreal, Quebec, Canada) and other smaller players for a wide variety of applications including dispersants for wettable powders, binders for granules and seed coatings, additives, etc. Although sulfur-containing lignins are generated during acid sulfite pulping or are produced by sulfonation of kraft lignin, the chemistry of the process and correspondingly the lignin structure is quite specific. The main reaction in sulfite pulping is sulfonation of lignin side chain, predominantly at the a-position of the propane chain as well as at the conjugated a-hydroxyl group. In addition, carbonyl groups also undergo sulfonation although the
TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS mechanism is different from that for hydroxyl groups. Introduction of highly polar sulfonate groups into the lignin structure strongly increases its solubility in aqueous solutions. Most of the lignosulfonates contain about 1 sulfonate group per 2 monomeric units. Although strong degradation of lignin is not needed to transfer sulfonated lignin from solid phase into solution, it still takes place and the degree of degradation is dependent on the reaction conditions. Therefore, the molecular mass of lignosulfonates is very high and varies strongly. The average Mw of lignosulfonates has been reported in the range of 10,000e40,000 Da and a fraction with Mw up to 100,000 Da has been isolated. An increased number of phenolic hydroxyl groups can be observed in lignosulfonates and it is strongly dependent on reaction conditions (Glennie, 1971). Comparison of Analytical Methods for Characterization of Technical Lignins There is a rather good correlation between different analytical methods used for the analysis of native lignin preparations, both in intralab (Evtuguin et al., 2001) and in interlab studies (Sakakibara, 1991; Capanema et al., 2004; Zhang and Gellerstedt, 2007; Balakshin et al., 2008). An exception is the 31P NMR analytical methodology of native lignins which yielded ca. 30% lower numbers for aliphatic hydroxyls (Pu et al., 2011) compared to other methods (Sakakibara, 1991; Balakshin et al., 2008). A rather good correlation between various methods for technical lignins analysis has been also reported (Faix et al., 1994; Cateto et al., 2008). However, a comprehensive review of published analytical data leads to a much less optimistic conclusion. Table 18.3 shows that significant variability in the structure of the same technical lignins can be observed when these lignins are analyzed by independent methods, in contrast to what is seen with the analysis of native lignin preparations. This deviation might be caused by the interference of specific lignin moieties generated during the technical process on the results of each specific analytical method which, as a rule, was developed and validated for the analysis of native lignin preparations. Various wet chemistry techniques for the analysis of lignin functional groups have been comprehensively reviewed previously (Lin and Dence, 1992; Zakis, 1994). Therefore, we will focus our discussion on major NMR spectroscopic techniques for the analysis of technical lignins, 31P and 13C NMR, as well as advanced NMR methods, which have received less attention. As 31P NMR spectroscopy of derivatized lignins becomes one of the most common techniques for lignin analysis, it is important to critically evaluate the method in a comprehensive manner. There are two main 323 modifications of the 31P NMR lignin analysis. Originally, 2-chloro-1,3,2-dioxaphospholane (31P-I method) was suggested (Archipov et al., 1991) as the derivatizing agent in this method. Later, 2-chloro-4,4,5,5-tetramethylwas reported 1,3,2-dioxaphospholane (31P-II) (Kostukevich et al., 1993; Granata and Argyropoulos, 1995) to provide better signal separation and it is currently used as the major 31P method for lignin analysis. Although good agreement has been reported between the results obtained with these two derivatization reagents (Granata and Argyropoulos, 1995), data reported later did not confirm this observation (Tables 18.3 and 18.4). The results obtained with the 31P-II method tend to underestimate results compared to the data generated by all other analytical methods. The 31P-I method tends to report significantly higher amount of aliphatic and total hydroxyl groups if compared to the data obtained by the 31 P-II methodology (Table 18.3), in contrast to the conclusions drawn in the original validation work (Granata and Argyropoulos, 1995). The significantly lower numbers reported by the 31 P-II NMR analysis, as it is compared to other analytical methods, in the structural analysis of lignins might be explained by the incomplete lignin derivatization with the phosphorylating reagent PR-II possibly due to steric hindrance of the bulky reagent containing four t-butyl groups. The use of PR-I yields apparently more quantitative results. However, the signal resolution in 31P-I is not high enough as it can clearly be seen in the publication by Akim et al. (2001). In this publication, the signals of primary hydroxyls and 5-substituted phenolic hydroxyls are heavily overlapped. The main conclusion derived from this observation is that even when the resolution of a resonance signal is formally acceptable (Argyropoulos, 1994), one cannot conclude that the resolved signals are reporting the correct values. Reproducibility of 31P NMR Analytical Techniques Although, a very high reproducibility has been reported in a specific intralaboratory study (Granata and Argyropoulos, 1995), the reproducibility of quantitative 31 P NMR spectroscopy reported in independent interlaboratory studies is much lower (Table 18.4). Moreover, even in studies conducted at the same laboratory, one can observe significant divergences between the results reported earlier (Granata and Argyropoulos, 1995) and more recently (Xia et al., 2001) for the same lignin samples (see SE aspen and poplar lignins in Table 18.3). As expected, the worst reproducibility of 31P NMR analytical methods has been observed for different types of 5-substituted phenolic hydroxyls (S-units and 5-condensed G units) in a hardwood technical lignin (Table 18.4) due to a very poor signal resolution.
324 In mmol/g per 100 C9 (or 100 Ar) Lignin Method References Aliphatic Phenolic COOH Total OH£ Alcell 31 Average (Gosselink et al., 2010; Wörmeyer et al., 2011; Vanderlaan and Thring, 1997; Cateto et al., 2010; Granata and Argyropoulos, 1995; Balakshin and Capanema, unpublished data; Saad et al., 2012) 1.51 3.80 0.32 5.31 2.54 31 (Argyropoulos, 1994) 2.68 3.91 0.34 6.59 1.46 13 (Balakshin and Capanema, unpublished data) 1.78 4.00 1.22* 5.78 13 (Cateto et al., 2010) 1.58U 3.88 5.46 P-II P-I C C Wet chem (Milne et al., 1992) 3.00 3.39 U 3.18 Alcell lab aspen Wet chem (Glasser et al., 1983) 1.59 Indulin AT (MWV) 31 Average (Gosselink et al., 2010; Cateto et al., 2010; Granata and Argyropoulos, 1995; Balakshin and Capanema, unpublished data; Beauchet et al., 2012) 2.34 3.66 31 (Argyropoulos, 1994) 3.04 3.15 13 (Balakshin and Capanema, unpublished data) 2.82 3.70 P-II P-I C 6.39 Phenolic/ Aliphatic 0.88 Total OH£ Conversion Factor 96 17.9 48 70 118 17.9 2.25 32 72 104 17.9 2.46 29U 71 100 18.3 61 115 17.9 59 109 1.13 27 Phenolic 68 4.77 0.42 Aliphatic 54 50 U U 5.99 1.57 42 66 108 18.1 6.19 1.04 55 57 112 18.1 6.52 1.13 51 67 118 18.1 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS TABLE 18.3 Content of Major Lignin Functional Groups Determined by Different Analytical Methods
13 C (Cateto et al., 2010) Wet chem Wet chem (Milne et al., 1992) (Glasser et al., 1983) 3.01 3.06 U 3.73 U 3.88 6.89 3.72 6.78 3.25 1.27 1.22 56 U 55 U 71 127 18.4 67 122 18.1 6.98 0.87 73 59 132 SW kraft Wet chem (Mörck et al., 1986) 2.2 4.3 0.80 6.5 1.95 40 79 119 18.3 Soda Sarkanda (Granit) 31 (Gosselink et al., 2010) 1.59 2.41 0.99 4.00 1.52 31 47 78 19.5 31 (Cateto et al., 2010) 1.89 2.74 0.62 4.63 1.45 37 53 90 19.5 13 (Cateto et al., 2010) 3.11 2.23 61 44 104 19.5 31 (Granata and Argyropoulos, 1995) 3.47 P-II P-II C SE aspen P-II 31 P-II P-I 13 x C-IS SE poplar 2.44 (Xia et al., 2001) 3.01 1.75 (Argyropoulos, 1994) 4.20 2.18 (Xia et al., 2001) U 2.93 U (Milne et al., 1992) 4.46 2.44 31 (Granata and Argyropoulos, 1995) 2.72 2.92 P-II P-II 31 P-I 13 x C-IS Wet chem Wet chem U (Xia et al., 2001) 2.25 2.29 (Argyropoulos, 1994) 2.97 2.46 (Xia et al., 2001) (Milne et al., 1992) (Glasser et al., 1983) U 2.12 3.08 U 3.25 0.21 1.72 Wet chem 31 0.31 2.16 3.08 2.53 5.91 0.70 67 47 114 19.3 4.76 0.58 58 34 92 19.3 6.37 0.52 81 42 123 19.3 46 126 27.1 4.65 0.41 0.00 0.59 U 80 U 6.89 0.55 86 47 133 19.3 5.64 1.08 53 57 110 19.5 4.54 1.02 44 45 89 19.5 5.44 0.83 58 48 106 19.5 51 104 24.3 60 120 19.5 48 111 4.28 6.16 5.78 1.02 1.00 0.78 *COOH þ COOR. x Quantitative 13C NMR using internal standard. U Calculated by difference: Aliphatic OH = Total OH  Phenolic OH £ COOH groups are not included Regular font, as reported; italic font, recalculated using conversion factor; “wet chem”, wet chemistry method; MWV, MeadWestvaco; SW, softwood. U 53 U 60 U 63 TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS 31 U 5.34 325
326 TABLE 18.4 Reproducibility of 31 P NMR (PR-II) for Some Reference Technical Lignins Phenolic References Aliphatic G-Condensed S Total 5-Substituted G H Total Phenolic COOH Total Hydroxyl** Phenolic/ Aliphatic Alcell 1.08 0.76 1.05 1.81 0.70 0.20 2.71 0.30 3.79 2.51 (Wörmeyer, Ingram et al., 2011) 1.28 0.24 1.45 1.69 0.64 0.11 2.44 0.26 3.72 1.91 (Vanderlaan and Thring, 1997) 1.5 5.4 2.60 (Cateto, Barreiro et al., 2010) 1.10 (Granata and Argyropoulos, 1995)* 1.84 (Saad, Radovic-Hrapovic et al., 2012) 2.2 (Balakshin, and Capanema, unpublished data) 1.57 0.77 1.55 2.32 0.99 Average 1.51 0.74 1.68 2.27 0.91 RSD, % 26.9 52.2 37.1 22.5 (Gosselink, van Dam et al., 2010) 2.08 1.3 (Beauchet, Monteil-Rivera et al., 2012) 2.25 (Granata and Argyropoulos, 1995)* 2.54 (Cateto, Barreiro et al., 2010) 2.34 1.58 (Balakshin, and Capanema, unpublished data) 2.45 Average 2.34 RSD, % 7.8 3.9 1.18 1.63 2.81 0.80 0.13 3.74 0.23 4.84 3.40 2.74 2.74 1.45 0.39 4.59 0.33 6.43 2.48 5.73 0.42 7.93 2.60 0.21 3.52 0.37 5.09 2.24 0.21 3.80 0.32 5.31 2.54 29.3 22.1 27.9 17.9 35.5 53.0 Indulin AT 1.3 1.62 0.23 3.15 0.44 5.23 1.51 1.38 1.94 0.23 3.55 0.36 5.80 1.58 3.59 0.43 6.13 1.41 1.93 1.91 1.96 0.26 4.13 0.39 6.47 1.76 1.22 1.65 1.95 0.25 3.85 0.48 6.30 1.57 1.37 1.56 1.88 0.24 3.66 0.42 5.99 1.57 17.8 7.8 6.2 10.0 11.0 8.2 8.2 13.8 0.33(?) *Recalculated from the original report (Granata and Argyropoulos, 1995) using a conversion factor reported earlier (Argyropoulos, 1994), see Table 18.3. ** COOH groups are not included; Bold font corresponds to statistically calculated values Average and Relative standard Deviation (RSD,%) 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS (Gosselink, van Dam et al., 2010)
TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS Therefore, in these cases, we consider erroneous reporting separately S-units and 5-condensed G-units, it is more reasonable to report them as “5-substituted phenolics”. The reproducibility of 31P NMR is overall better for major signals such as aliphatic hydroxyls (AliphOH), phenolic hydroxyls (PhOH), and total OH, especially for Indulin AT lignin. Surprisingly, it is still not of very high quality the results reported for Alcell lignin (Table 18.4). In part, but not completely, this observation might be explained by inconsistencies in the lignin itself. In addition, 31P NMR reports much lower carboxyl numbers than wet chemistry methods and 13C NMR methods (Table 18.3). In fact, 13C NMR methods usually report the sum of carboxyl and ester structures in general. For instance, the significantly higher numbers reported by 13 C NMR for Alcell lignin might be explained by the significant contribution of ester structures (predominantly ethyl esters) in this lignin. However, it is quite unreasonable to expect significant amounts of esters in kraft lignins isolated from high alkaline solutions. Therefore, it might be concluded that 31P NMR also underestimates the amounts of carboxyl groups in lignins. In summary, it appears that 31P NMR-I does not provide sufficient resolution even between major signals, PhOH and AliphOH, while 31P NMR-II yields significantly lower values. Moreover, separation of S-units and G-condensed (at C-5) structures is very ambiguous in 31P NMR-II analysis of technical lignins, consequently so is the evaluation of S/G ratio and the degree of condensation. The amount of 5-substituted (S þ G-condensed) and 5-unsubstitued phenolic hydroxyl should be reported instead. Comparative data for the analysis of various technical lignins by the 31P-II NMR method are summarized in Table 18.5. Due to the high degree of variability in the structure of lignins discussed above, it is difficult to make any general conclusions on the existing structural differences among the lignins originated from various technical processes (Tables 18.3e18.5). In addition to the variability linked to the feedstock origin and process variables, the numbers for different structural moieties reported can vary due to the particularities of the used analytical methodologies (Table 18.3). There is overall a lack of comprehensive comparison of technical lignins. In addition, the existing comparative studies are limited to a few lignin functionalities only, such as amounts of phenolic, aliphatic hydroxyl groups, or methoxyl groups. 13 C NMR Analysis of Technical Lignins The 13C NMR analysis of lignin has been considered as a very informative, but not very affordable method due to the very long experimental time originally required for quantitative lignin analysis (Chen and Robert, 1988). Recently, we have optimized the experimental time 327 required for 13C NMR analysis and reduced it from 70 to 15e20 h (Capanema et al., 2004, 2005a). Thus, a large amount of valuable structural information (20e30 results on structural moieties per analyzed lignin sample) can be obtained in a reasonably short experimental time which permits considering the 13C NMR method as the most productive one in lignin analysis. Furthermore, we demonstrated that the use of a CryoProbe NMR (Bruker BioSpin MRI GmbH, Germany) allows for 1 h total quantitative 13C NMR experimental time (Balakshin, Berlin et al., 2013). Therefore, 13C NMR cannot be considered a time consuming lignin analytical technique anymore. In addition, the CryoProbe yields much better signal resolution, both for 13C and Heteronuclear Single Quantum Coherence (HSQC) NMR methods. However, currently the use of CryoProbe does require tedious and professional optimization of acquisition and processing parameters to adjust the baseline, specifically for 13C NMR spectra of lignin samples. Therefore, we cannot recommend yet the use of this method on a routine basis. Further development in the use of CryoProbe technology and its use for lignin analytical chemistry is expected to mitigate this limitation. A careful optimization of the acquisition parameters for lignin analysis using a traditional probe yields 13C NMR spectra with a good and reproducible baseline, which can be easily and reliably corrected during spectra processing. This careful adjustment of the acquisition and processing parameters has enabled the recording of reproducible results even in different NMR spectrometers with a relative error of ca. 2e3% for the major lignin peaks. Unfortunately, it is not feasible for now to evaluate interlab variability of the 13 C NMR method as it has been used much less often than the 31P NMR method and data for the same lignin preparations are very limited. It is also very important to consider some issues when calculating the amount of various lignin moieties using the original 13C NMR spectra as it has been discussed earlier (Capanema et al., 2005b). However, our team has been acquiring over the past few years significant information on different technical lignins, which is hereby summarized in Table 18.6. Since the data were produced and interpreted by the same analytical methodology, their comparison is more accurate than the comparisons based on data obtained from various literature reports. The analysis of technical lignins (Table 18.6) clearly showed dramatic changes in lignin structure resulting from the delignification process. In addition to between-process variations, certain within-process lignin structural changes could be documented. One of the most important factors in these variations is the feedstock origin. Significant differences in the structure of technical lignins from different tree species were obvious and they were significantly larger than the differences in the
328 TABLE 18.5 Analysis of Different Technical Lignins by 31P-II NMR Method (mmol/g Lignin) Phenolic References Aliphatic Total 5-Substituted G-Non-condensed H Total Phenolic COOH Phenolic/ Aliphatic Alcell Average (Gosselink et al., 2010; Wörmeyer et al., 2011; Vanderlaan and Thring, 1997; Cateto et al., 2010; Granata and Argyropoulos, 1995; Balakshin and Capanema, unpublished data; Saad et al., 2012) 1.51 2.27 0.91 0.21 3.80 0.32 2.54 Pine organosolv (Pu et al., 2011; Sannigrahi et al., 2010) 7.3 0.60 1.4 0.4 2.7 0.3 0.37 Straw organosolv (Wörmeyer et al., 2011) 4.69 0.34 0.57 0.36 1.27 0.12 0.27 Miscanthus organosolv (Pu et al., 2011; 15) 1.26e3.11 1.58e0.91 0.49e0.65 2.12e3.93 0.16e0.28 Miscanthus organosolv (Pu et al., 2011; 16) 1.19 1.33 0.61 3.07 0.22 2.58 Indulin AT Average (Gosselink et al., 2010; Cateto et al., 2010; Granata and Argyropoulos, 1995; Balakshin and Capanema, unpublished data; Beauchet et al., 2012) 2.34 1.56 1.88 0.24 3.66 0.42 1.57 Curan 100 (Gosselink et al., 2010) 1.78 1.55 1.84 0 3.39 0.43 1.90 Sarkanda soda granit Average (Gosselink et al., 2010; Cateto et al., 2010) 1.74 1.34 0.77 0.47 2.58 0.81 1.48 Straw soda (Wörmeyer et al., 2011) 3.18 0.32 0.66 0.16 1.14 1.07 0.36 Hardwood soda (Gosselink et al., 2010) 1.34 1.62 0.51 0.34 2.47 1.06 1.84 Aspen steam explosion (Granata and Argyropoulos, 1995)* 3.47 1.76 0.67 2.44 0.31 0.70 (Xia et al., 2001) 3.01 (Granata and Argyropoulos, 1995)* 2.72 (Xia et al., 2001) 2.25 Pine acid hydrolysis (Pu et al., 2011; Sannigrahi et al., 2008) 3.42 0.34 1.82 Switchgrass acid hydrolysis (Pu et al., 2011; 17) 2.83 0.35 0.57 Poplar steam explosion 1.75 2.00 0.92 *Recalculated from the original report (Granata and Argyropoulos, 1995) using a conversion factor reported earlier (Argyropoulos, 1994), see Table 18.3. 2.92 0.58 0.41 1.08 2.29 1.02 0.06 2.22 0.65 0.33 1.25 0.33 0.44 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS Lignin
329 TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS TABLE 18.6 13 C NMR Analysis of Different Native and Technical Lignins (per 100 Ar) Moieties Birch MWL Spruce MWL Birch Kraft E. glob. kraft E. grandis kraft Pine kraft Indulin AT Alcell Douglas fir OS Total CO 12 21 9 14 12 11 12 35 20 Non-conjugated CO 3 5 4 5 4 4 5 16 7 Conjugated CO 9 14 5 9 8 7 7 19 13 Total COOR 4 5 20 18 16 21 16 19 6 Aliphatic COOR 3 4 18 16 13 20 15 15 5 Conjugated COOR 1 1 2 2 3 1 1 4 1 Total OH 150 138 107 128 119 108 118 104 118 Aliphatic 129 107 27 51 39 34 51 32 37 Primary 73 68 23 29 24 23 32 18 27 Secondary 56 39 3 22 15 11 19 14 10 Phenolic 20 31 80 77 80 74 67 72 81 S/G 3.02 n.a. 1.7 2.5 1.4 n.a. n.a. 1.3 n.a. ArH 209 172 191 186 218 235 209 DC 16 38 65 37 55 82 65 33 73 b-O-4 66 45 2 12 5 3 7 8 <5 b-b 11 4 3 2 3 5 4 3 1 b-5 2 9 2 1 2 3 4 3 3 OCH3 177 95 141 141 125 81 80 117 85 Oxygen aliphatic 260 86 110 79 72 94 94 86 Saturated aliphatic 15 . . 68 96 62 -OEt n.a. n.a. n.a. n.a. 13 10 Sugars 4 1 1 <1 Alk-ether 68 61 52 43 55 48 38 32 17 >35 Alk-O-Alk n.a. 42 n.a. n.a. These numbers could be recalculated on a mmol/g basis using the approximate mass of a C9-unit (ca. 180) for Organoslv and kraft Lignins (see Table 18.3). (Source: Capanema, Balakshin et al., 2005b; Berlin et al., 2006; Balakshin et al., 2008; Balakshin, Capanema unpublished data.) structure observed for native lignins in these tree species. For instance, it was shown that various hardwood lignins degraded differently during kraft pulping resulting in variations of hydroxyl and carboxyl groups, b-O4, b-b, and b-5 linkages as well as in S/G ratio and degree of condensation (Capanema et al., 2005b; Balakshin et al., 2008). In fact, species-originated variations are similar or even larger than the variations in major lignin functionalities caused by different delignification technologies, such as kraft and OS processes. The only significant differences observed between kraft and ethanol OS lignins (as analyzed by 1-D NMR) are the incorporation of ethoxyl groups and the significantly higher amounts of carbonyl groups in the latter. Most of the wet chemistry and 31P NMR methods originally yield results in mmol/g (or mass %) units. The 13C NMR method reports results in number of functional groups per aromatic ring (Ar). A conversion factor based on the C9-formulae is typically used to correlate these values (mmol/g and units/Ar), but the ratio is not obvious. The C9-formulae might not be accurate (even for high-purity lignins; contaminations would also contribute to this NMR signal) as the lignin side chain is degraded, to a certain extent, during biomass processing. The 13C NMR lignin analysis with Internal Standard (IS) allows for both types of data presentation. A very good correlation between 13 C NMR with IS and 31P NMR data for the hydroxyl
330 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS group content has been reported (Xia et al., 2001). However, in that publication, the authors did not specify if a correction for lignin acetylation was applied to the 13C NMR data or not. The proportion between the values expressed “per 100 Ar” and those in mmol/g (for the same lignin) allow us to calculate the weight of an average C9-unit (or Ar). The numbers obtained are 271 and 243 for Aspen SE and Poplar SE lignin, correspondingly (Table 18.3). These values are much higher than those calculated based on the C9-formulae (195 and 193, correspondingly; Table 18.3) and indicate that the amounts of OH groups in the 13C-IS NMR experiments have been calculated based on acetylated lignin. Therefore, recalculation as per the original (non-derivatized) lignin would give numbers of ca. 25e40% higher for the 13C NMR (with IS) vs. 31P NMR-II. It should be mentioned that a very good correlation between 13C-NMR-IS data (for a non-acetylated lignin) and the methoxyl group wet chemistry analysis, one of the most reliable analytical methods in lignin chemistry, has been reported (Xia et al., 2001). This indicates that 13C NMR data should be considered as more realistic and that the 31P-II NMR method produced significantly underestimated numbers (probably due to incomplete derivatization), in agreement with the earlier discussed results. In summary, 13C NMR with IS is probably the best analytical approach to obtain the most comprehensive and reliable lignin structural information expressed, both in mol% (units/Ar) and in mmol/g. Unfortunately, very little has been reported so far on the methodology development and its validation of the 13C NMR lignin analysis with IS vs other analytical techniques. Overall, the lignin scientific community believes, based on a few publications, that there is a good correlation between the different methods used for the analysis of the technical lignin chemical structures. However, a comprehensive review of the existing database (especially of independent publications) clearly shows that this is not the case. In fact, the deviation between reported data using even the same analytical method (such as 31P NMR) for the same lignin preparation is often very significant. Moreover, the deviation between different analytical methods is in the range of the differences observed between different lignin types. This conclusion indicates that significant efforts still should be made to address these deviations and to standardize the analytical methodology for technical lignins analysis. Meanwhile, we should remember the general principle that it is naturally more accurate to compare structural data obtained with the same analytical method in the same lab. The use of “reference” lignin samples (well-investigated lignins, such as Alcell and Indulin AT) would be also very beneficial to ensure at least a reliable relative comparison. Advanced NMR Methods Routine analytical methods, even comprehensive C NMR analysis, have failed to reveal the key lignin structures originated by different biomass processes, for example, differences between Alcell OS and kraft Indulin lignins (Tables 18.3 and 18.6). Therefore, this methodology is not sufficient to distinguish typical structural features of different lignins neither in qualitative nor in quantitative analytical mode. This indicates the necessity of advanced analytical methods to describe the key characteristics in the structure of technical lignins. Two-dimensional NMR methods, specifically the HSQC technique, allow to distinguish specific structural characteristics of various technical lignins (Capanema et al., 2001; Balakshin et al., 2003; Liitiä et al., 2003). The most advanced structural characterization has been achieved so far for kraft lignins. The HSQC analysis of OS lignins showed significant amounts of specific structures (Balakshin et al., 2000; Capanema et al., 2001; Berlin et al., 2006), however, their exact signal assignment was not possible due to limited NMR data for specific model compounds. Further studies are required in order to perform the proper assignments in this type of lignins. The first attempts to quantify specific lignin functionalities in different technical lignins have already been undertaken (Capanema et al., 2008). The 2D NMR approach pursued by the authors of the latter article focused on the quantification of lignin moieties, which were not possible to quantify with 13C NMR alone (and other 1D NMR techniques), especially of those structures formed during pulping. The study provided important quantitative information on various structural lignin units, such as condensed lignin moieties, products of bO-4 bond cleavage, vinyl, and alkyl-aryl structures, saturated aliphatic moieties and others, as well as lignin-carbohydrate linkages (Table 18.7). Another useful advanced NMR analytical method is DEPT 13C NMR that allows for quantification of specific lignin functionalities overlapped in routine 13C NMR spectra (Gellerstedt and Robert, 1987). These advanced NMR methods showed the possibility of expanding our understanding of the structure of technical lignins. However, more comprehensive studies and cross-validation of the advanced methodologies with independent methods are needed before these methods can be routinely used. 13 Molecular Weight Distribution The accurate determination of lignin Mw is an important aspect of lignin characterization. The lignin Mw is
TECHNICAL LIGNINS: PRODUCTION, PROPERTIES, AND ANALYSIS TABLE 18.7 Amounts of Different Structural Units per 100 Ar (per 100 Monomeric Units) Quantified by Combination of 13C and HSQC NMR Techniques Units PSDL PKRL EKRL EKDL 1-G 0.7 5.2 5.8 1.7 1-S na na 21.8 2.2 C 1.1 2.8 2.3 0.8 5 1.3 1.8 0.8 0.7 6 2.7 4.0 7.5 4.6 1.5 6.4 1.2 18 17 2.9 4.2 1.3 10 1.4 2.4 28 6.9 3.8 20 3.5 0.6 30 4.1 25.2 3.4 13.1 29 2.0 nd nd tr 21e26 4.7 2.1 nd 2.4 13 e 3.1 nd 0.2 16 2.8 2.2 nd 0.3 S/G na na 3.0 2.1 7.2 1.5 tr PSDL, pine soda dissolved lignin; PKRL, pine kraft residual lignin; EKRL, Eucalyptus globulus kraft residual lignin; EKDL, Eucalyptus grandis kraft dissolved lignin. See Figure 18.2 for lignin structural units numbering reference. Source: Capanema et al., 2008. typically analyzed by a combination of size-exclusion HPLC and a detection method (refractive index, evaporative light scattering, or others) using calibrants which mimic the typical Mws and chemistry of lignins. However, lignin Mw determination still remains a challenge. For instance, a recent joined study by seven major European lignin research groups reported a reasonably small (<15%) intralab variability in the determination of lignin Mw. However, the interlab differences were very large, up to 49 times(!) difference even under standardized conditions (Baumberger et al., 2007). StructureeProperties Correlations in Lignin In spite of the large number of studies dedicated to the utilization of technical lignins, the correlation between lignin chemical structure and its properties and functions has not been established yet for most industrial applications. Lignin functional properties are most often correlated with physical properties, such as glass transition point (Tg), solubility and, sometimes, with molecular mass distribution (Glasser, 2000) as well as with such compositional features as the ash, carbohydrate, sulfur, and carbon content. The effect of the chemical structure of lignins on lignin performance in specific 331 applications is usually anticipated based on fundamental knowledge rather than on experimentally established correlations. For example, the behavior of lignin in polyurethane production is correlated with the amount of hydroxyl groups predominantly. The utilization of lignin in phenoleformaldehyde (PF) resins requires the unsubstituted 5-position of the aromatic ring (o-position to the phenolic hydroxyl group) and therefore higher proportion of G-units is desirable, in contrast to S-units, which cannot participate in this reaction. The comparison of lignin reactivity is typically limited to the comparison of lignins originated from different technical processes and different feedstock origins (Tejado, 2007; Mansouri and Salvado, 2006; Evtuguin et al. 1998; Rials and Glasser, 1986). An attempt to correlate lignin structure with its performance includes, for instance, the observation that the presence of ethyl groups in Alcell lignin act as an internal plasticizer and improve the lignin performance in poly(ethylene oxide) blends (Kubo and Kadla, 2004). Another attempt was the correlation of the quantity of aliphatic, phenolic hydroxyl groups and methoxyl groups as well as the lignin Mw with antioxidant lignin properties (Pan et al., 2006). However, the correlations between the amount of phenolic hydroxyls, Mn and the Radical Scavenging Index observed in this study were rather poor or inexistent (R2 ¼ 0.53 or lower) implying that the dependence is more complex and it requires comprehensive lignin structural elucidation. In summary, it would appear that the main reason for the lack of clear structureefunctional performance correlations is the high heterogeneity and variability of technical lignins and the absence of widely accepted and understood, quantitative, fast, and simple analytical techniques (Glasser, 2000). In the past few years, leveraging all the recent advances made in the development of new lignin analytical techniques, a very comprehensive work on correlation between process parameters and the structure and properties of the produced technical lignins was undertaken at the R&D Department of Lignol Innovations, Ltd (Vancouver, Canada) (Balakshin, Berlin et al., 2013). Three categories of feedstock (softwoods, hardwoods, and annual fibers) including various biomass species sourced from different continents were processed under at least 30 different combinations of process conditions (time, temperature, ethanol concentration, pH, and L:S ratio). The extracted lignins were analyzed using advanced rapid and comprehensive 13C high-resolution NMR spectroscopy coupled with a CryoProbe technology for lignin structural characterization, as described above, along with traditional methods for lignin composition analysis, Mw, thermal properties (Tg), antioxidant activity, and other properties generating results for over 50 different characteristics for each lignin sample. This effort generated a very comprehensive
332 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS database including several thousands of data points covering a wide diversity of OS lignins (Balakshin and Berlin, 2010). This unique database has been playing a very important role in developing Lignol’s lignin applications helping in the selection of best lignin candidates for specific customer needs. Moreover, very accurate models correlating process parameters and lignin characteristics have been also developed on the basis of these studies (Balakshin, Berlin et al., 2013). TECHNICAL LIGNINS: TRADITIONAL AND EMERGING APPLICATIONS According to the Global Industry Classification Standard (GICSÒ) developed by Morgan Stanley Capital International and Standard & Poor’s, the global industries can be divided into 10 sectors: energy, materials, industrials, consumer discretionary, consumer staples, health care, financials, information technology, telecommunication services, and utilities. After a review of each of the 10 GICSÒ industrial sectors, it should be noted that lignin finds applications in almost every sector of the industry with the exception of financials, telecommunication services, and information technology. The latter illustrates the versatility of lignin chemistry and its potential. Overall, lignin applications can be divided into traditional and emerging. The traditional applications of lignosulfonates and sulfonated kraft lignins are primarily lower value such as dust control (ca. 11% market), concrete admixtures (ca. 50% market), and oil well drilling muds (ca. 4% market) while most of the emerging applications target higher value applications, in which the chemical versatility of lignin could be fully leveraged, or very large volume applications such as the production of BTX and other petrochemicals. Traditional Lignin Applications The traditional lignin applications include, primarily, lignin uses where lignin plays a replacement role for a relatively low-value chemical or material. As it was mentioned earlier, the largest current use of industrial lignin is fuel. However, there are a number of matured industrial lignin applications, constituting the bulk of the higher value lignin commercial chemical market, and this includes, primarily, lignosulfonates or sulfonated kraft lignins in large-medium-size markets such as additives for concrete admixtures, dust control, feed and food additives, dispersants, resin and binder compositions, and oil well drilling. Examples of smaller markets for technical lignins include carbon black, emulsifiers, water treatment, cleaning chemicals, leather tanning, battery expanders, and rubber additives (Higson, 2011). There are many comprehensive reviews in the literature which cover traditional lignin applications (ACS, 1999; Holladay et al., 2007); therefore, we will not discuss them here in detail. Emerging Lignin Applications The growth of the lignin business in this century depends, for the most part, on the ability of chemists and materials scientists to develop novel original applications and processes for lignin valorization. From 1990 to 2010, a number of novel lignin applications have been proposed and described in the scientific and patent literature. Some of these novel applications have been piloted or demonstrated at larger than laboratory scale. For instance, the production of lignin-based carbon fiber (LCF) (Baker and Rials, 2013) is one of the brightest examples of successful lignin upgrading technology which have been scaled up to pilot scale. Another example of nontraditional lignin applications which is being commercialized with a market potential comparable to carbon fiber is the use of upgraded technical lignins in polymers, in particular, in thermoplastic, thermoset, and composite applications. Not far in the future, we will witness the rise of processes aiming at depolymerization of lignins to produce valuable oxygenated aromatic compounds, and possibly olefins too, in replacement of petrochemicals. This section briefly summarizes these three relevant examples of emerging nontraditional lignin applications which in our opinion present the largest potential, in terms of volume and value, for commercialization of technical lignins in the future. Among all the emerging lignin applications, the manufacturing of LCFs is perhaps the one with the largest market potential in terms of value. This dream finds its beginning in the early 1960s, in Japan, when a group led by Dr. Sugio Otani from the Department of Chemistry, Faculty of Technology, Gunma University developed a technology to turn lignin into carbonized fibers (Otani et al., 1969). The Kajima Corporation, a giant Japanese construction company, working together with Dr. Otani, had developed, on the basis of this early invention, a ligninefiber-based reinforced cement with superior strength properties. Since the early work of Dr. Otani, the quest to develop commercially viable LCF applications has been focused on achieving lignin and lignin blends with properties enabling high rate of fiber spinning (>2000 m/s), yielding scalable and fast conversion technologies, as well as high yield. However, regardless of all the efforts made so far by multiple research groups worldwide, in particular, by the Carbon Fiber Composites Consortium (Oak Ridge National Laboratory), the development of high-quality structural LCF has proved to be very challenging. The main challenge remains achieving the required LCF engineering properties imposed by existing carbon fiber spinning
333 REFERENCES technologies. In particular, the best LCF prototypes have shown a relative low fiber tensile strength (w1 GPa) and low fiber elastic modulus (<100 GPa) while most structural applications (automotive, sport goods, wind turbines, and aerospace applications) require tensile strengths well above 1 GPa and elastic moduli over 100 GPa. Attempts to overcome these limitations have been made where lignin is blended with synthetic polymers, such as acrylonitrile, and copolymerized to yield hybrid LCF with acceptable tensile strengths (Maradura et al., 2012). In addition to the LCF mechanical limitations, the relatively high cost can be pointed out as another limiting factor deterring LCF commercialization. Owing to these hurdles, the focus of LCF R&D efforts has been recently shifting from structural applications toward functional uses where LCF seem better suited. Examples of functional uses are hightemperature insulating materials, CO2 and other gases capture sorbents, controlled adsorption and release of macromolecules, and capacitors. On the cost reduction side, besides the improvement of technical lignins as a raw material, so their upgrading costs can be minimized, there is the potential to realize cost reductions by way of alternative fiber spinning methods that are not bound by the stringent technical requirements needed for the traditional melt-spinning of carbon fiber precursors (Baker and Rials, 2013). In addition to cost reduction and attempts to improve LCF mechanical properties, the development of novel technologies to further upgrade lignin will be required to meet the demanding industrial standards. Valorization of lignin can be achieved also via its incorporation in polymeric materials, in particular, in composite materials containing thermosets, such as, phenoleformaldehyde, ureaeformaldehyde, and epoxy resins (Zhao et al., 2001; Mankar et al., 2012; Wang et al., 2012; Yin and Di, 2012), and, thermoplastics, such as lignin blends for extrusion applications with polyesters (Li and Sarkanen, 2002), polyamides (Nitz et al., 2001), polyacrylonitrile (Seydibeyo glu, 2012), and polyethylenes (Casenave et al., 2009). Technical lignins find also applications in structural materials such as polyurethanes which are recognized as one of the most versatile classes of polymeric materials and they show good compatibility with lignin given the presence of both aromatic and aliphatic hydroxyl groups within the lignin structure (Cateto et al., 2010; Faria et al., 2012). The cornerstone of a viable wide incorporation of technical lignins in synthetic polymer blends is their compatibility with the chemical matrices. The compatibility between lignin and synthetic polymers, often more hydrophobic than lignin, can be achieved, for instance, via chemical modification of lignin through esterification (Li and Sarkanen, 2002) or during lignin production by modifying the reaction conditions. Finally, the depolymerization of lignins to produce valuable oxygenated aromatic compounds is an application which, if successful, will consume most of the technical lignin supply in the future. However, this goal faces, perhaps, one of the most challenging lignin technological barriers (Holladay et al., 2007). Multiple approaches have been attempted to achieve the evasive lignin depolymerization target. The thermochemical methods including pyrolysis (thermolysis), gasification, hydrogenolysis, chemical oxidation, and hydrolysis under supercritical conditions are the major methods studied with regard to lignin depolymerization (Pandey and Kim, 2011). The biochemical route, due to its relative low capital required for deployment and low-energy operation, has been extensively researched as a way of selectively depolymerizing lignin but its high cost and relatively long reaction time continues to be a barrier for commercialization (Chen et al., 2012a,b). CONCLUSIONS The low value of technical lignins as a fuel has been evidenced in this review. However, new technological developments for smart utilization of technical lignins can lead to much higher value market opportunities, possibly with lignin-based chemicals superior to petrochemicals as it was illustrated in the case of lignin-based hybrid resins. Further technological advancements in the upgrading and optimization of technical lignins for materials applications is currently constrained by our ability to better understand their chemical structure and reactivity as well as by our capacity to efficiently purify and adapt them to the needs of the demanding chemical industry. Advancements in developing comprehensive and validated lignin analytical methods applied to purposeful lignin applications will help with mitigating the existing technological hurdles. Regrettably, the efforts made toward developing lignin as a chemical feedstock remain very modest compared, for instance, to those made toward the utilization of other biomass components such as cellulose and hemicellulose. The refining of complex raw chemical streams into building blocks is a matured technology best exemplified by the petroleum and gas industries. Therefore, the development of lignin-refining technologies into valuable chemicals should also be possible in the near future if adequate resources are directed toward this goal. References Abdelkafi, F., Ammar, H., et al., 2011. Structural analysis of alfa grass (Stipa tenacissima L.) lignin obtained by acetic acid/formic acid delignification. Biomacromolecules 12 (11), 3895e3902. ACS, 1999. Lignin: Historical, Biological, and Materials Perspectives. American Chemical Society, Washington, DC.
334 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS Adler, E., 1977. Lignin chemistry: past, present and future. Wood Sci. Technol. 11, 169e218. Akim, L., Argyropoulos, D., et al., 2001. Quantitative 31P NMR spectroscopy of lignins from transgenic poplars. Holzforschung 55 (4), 386e390. Archipov, Y., Argyropoulos, D., et al., 1991. 31P NMR spectroscopy in wood chemistry. I. Model compounds. J. Wood Chem. Technol. 11 (2), 137e157. Argyropoulos, D., 1994. Quantitative phosphorus-31 NMR analysis of six soluble lignins. J. Wood Chem. Technol. 14, 65e82. Baker, D., Rials, T., 2013. Recent advances in low-cost carbon fiber manufacture from lignin. J. Appl. Polym. Sci. http://dx.doi.org/ 10.1002/APP.39273. Balakshin, M., Berlin, A., 2010. Lignol’s Biorefinery Technology - Unlocking the Biomass Value. Bio-Based Chemicals East: Strategic Science and the Business of Breakthrough Technologies, Boston, MA, USA. Balakshin, M., Capanema E. unpublished data Balakshin, M., Capanema, E., et al., 2000. The Use of 2D NMR Spectroscopy on Structural Analysis of Residual and Technical Lignin. Proceedings of the Sixth European Workshop on Lignocellulosics and Pulp, Bordeaux, France. Balakshin, M., Capanema, E., et al., 2003. Elucidation of the structures of residual and dissolved pine kraft lignins using an HMQC NMR technique. J. Agric. Food Chem. 51 (21), 6116e6127. Balakshin, M., Capanema, E., et al., 2007. A fraction of MWL with high concentration of lignin-carbohydrate linkages: isolation and analysis with 2D NMR spectroscopic techniques. Holzforschung 61, 1e7. Balakshin, M., Capanema, E., et al., 2008. Recent advances in the isolation and analysis of lignins and ligninecarbohydrate complexes. In: Hu, T. (Ed.), Characterization of Lignocellulosic Materials. Blackwell, Oxford, UK, pp. 148e170. Balakshin, M., Capanema, E.A., et al., 2011. Quantification of lignincarbohydrate linkages with high-resolution NMR spectroscopy. Planta 233, 1097e1110. Balakshin, M., Berlin, A., et al., 2013. Derivatives of Native Lignin from Softwood Feedstocks. USPTO. USA. Lignol Innovations, Ltd. US8431635 B2. Balakshin, M., Berlin, A., et al., 2013. Processes for Recovery of Derivatives of Native Lignin. USPTO. USA. Lignol Innovations, Ltd. US8378020 B1. Balakshin, M., Yu, M., Capanema, E.A., et al., 2014. Isolation and analysis of Lignin-Carbohydrate Complexes (LCC) preparations with traditional and advanced methods. A review. In: Studies in Natural Products Chemistry. Elsevier, Amsterdam, in press. Baumberger, S., Abaecherli, A., et al., 2007. Molar mass determination of lignins by size-exclusion chromatography: towards standardisation of the method. Holzforschung 61 (4), 459e468. Beauchet, R., Monteil-Rivera, F., et al., 2012. Conversion of lignin to aromatic-based chemical (L-chems) and biofuels (L-fuels). Bioresour. Technol. 121, 328e334. Berlin, A., Balakshin, M., et al., 2006. Inhibition of cellulase, xylanase and beta-glucosidase activities by softwood lignin preparations. J. Biotechnol. 125 (2), 198e209. Berlin, A., 2011. Binder Compositions Comprising Lignin Derivatives. W.I.P.O., WO2011097719. Berlin, A., 2012a. Carbon Fibre Compositions Comprising Lignin Derivatives. E. P. Office. Europe. Lignol Innovations, Ltd. Berlin, A., 2012b. Resin Compositions Comprising Lignin Derivatives. European Patent Office, EP2435457. Capanema, E., Balakshin, M., et al., 2001. Structural analysis of residual and technical lignins by 1H-13C correlation 2D NMRspectroscopy. Holzforschung 55 (3), 302e308. Capanema, E., Balakshin, M., et al., 2004. A comprehensive approach for quantitative lignin characterization by NMR spectroscopy. J. Agric. Food Chem. 52 (7), 1850e1860. Capanema, E., Balakshin, M., et al., 2005a. Quantitative characterization of a hardwood milled wood lignin by nuclear magnetic resonance spectroscopy. J. Agric. Food Chem. 53 (25), 9639e9649. Capanema, E., Balakshin, M., et al., 2005b. Isolation and Characterization of Residual Lignins from Hardwood Pulps: Method Improvements. Proceedings of the Thirteenth International Symposium on Wood, Fibre, and Pulping Chemistry (ISWFPC), Auckland, New Zealand. Capanema, E.A., Balakshin, M.Yu., et al., 2008. Quantitative analysis of technical lignins by a combination of 1H-13C HMQC and 13C NMR methods. In: Proceedings of International Conference on Pulping, Papermaking and Biotechnology. Nov. 4-6. Nanjing, China. 647e651. Casenave, S., Aı̈t-Kadi, A., et al., 2009. Mechanical behaviour of highly filled lignin-polyethylene compistes made by catalytic grafting. Can. J. Chem. Eng. 74 (2), 308e315. Cateto, C., Barreiro, M., et al., 2008. Lignins as macromonomers for polyurethane synthesis: a comparative study on hydroxyl group determination. J. Appl. Polym. Sci. 109 (5), 3008e3017. Cateto, C., Barreiro, M., et al., 2010. Polyurethane as a Viable Route to Valorise Lignin. XXI Encontro Nacional da TECNICELPA/VI CIADICYP 2010, Lisbon, Portugal. Chen, C.-L., Robert, D., 1988. Characterization of Lignin by 1H and 13C NMR Spectroscopy, Vol. 61B. Academic Press Inc, New York pp. 137e174. Chen, Q., Marshall, M.N., et al., 2012. Effects of laccase on lignin depolymerization and enzymatic hydrolysis of ensiled corn stover. Bioresour. Technol. 117 (0), 186e192. Chen, Y., Sarkanen, S., et al., 2012. Lignin-degrading enzyme activities. Methods Mol. Biol. 908, 251e268. Chudakov, M., 1983. Industrial Utilization of Lignin. Lesnaya Promishlennost, Moscow, Russia. Chung, H., Washburn, N., 2012. Improved Lignin Polyurethane Properties with Lewis Acid Treatment. ACS Applied Materials and Interfaces. Domtar, March 12, 2013. Domtar inaugurates commercial lignin production. Biomass Mag. Retrieved June 19, 2013, from biomassmagazine.com. Evtuguin, D.V., Andreolety, J.P., et al., 1998. Polyurethanes based on oxygen-organosolv lignin. Eur. Polym. J. 34, 1163e1169. Evtuguin, D., Neto, C., et al., 2001. Comprehensive study on the chemical structure of dioxane lignin from plantation Eucalyptus globulus wood. J. Agric. Food Chem. 49, 4252e4261. Faix, O., Argyropoulos, D., et al., 1994. Determination of hydroxyl groups in lignins. Evaluation of 1H-, 13C-, 31P-, FTIR and wet chemistry methods. Holzforschung 48, 387e394. Faria, F., Evtuguin, D., et al., 2012. Lignin-based polyurethane doped with carbon nanotubes for sensor applications. Polym. Int. 61 (5), 788e794. Gellerstedt, G., Robert, D., 1987. Quantitative 13C NMR analysis of kraft lignins. Acta Chem. Scand., Ser. B 41, 541e546. Gellerstedt, G. (Ed.), 1996. Chemical Structure of Pulp Components. Pulp Bleaching: Principles and Practice. Tappi Press, Peachtree Corners, GA. Gierer, J., 1980. Chemical aspects of kraft pulping. Wood Sci. Technol. 14, 241e266. Glasser, W., Barnett, C., et al., 1983. The chemistry of several novel bioconversion lignins. J. Agric. Food Chem. 31 (5), 921e930. Glasser, W., 2000. Classification of lignin according to chemical and molecular structure. In: Glasser, W., Northey, R., Schultz, T. (Eds.), Lignin: Historical, Biological, and Material Perspectives, pp. 216e238. Washington, DC. Glasser, W., 2010. New Perspectives on Lignin Conversion (International Lignin Biochemicals Conference).
REFERENCES Glennie, D., 1971. Reactions in Sulfite Pulping. Wiley e Interscience, New York. Gosselink, R., de Jong, E., et al., 2004. Co-ordination network for lignin - standardisation, production, and applications adapted to market requirements (EUROLIGNIN). Ind. Crops Prod. 20, 121e129. Gosselink, R., van Dam, J., et al., 2010. Fractionation, analysis, and PCA modeling of properties of four technical lignins for prediction of their application potential in binders. Holzforschung 64 (2), 193e200. Granata, A., Argyropoulos, D., 1995. 2-chloro-4,4,5,5-tetramethyl1,3,2-dioxaphospholane, a reagent for the accurate determination of the uncondensed and condensed phenolic moieties in lignins. J. Agric. Food Chem. 43 (6), 1538e1544. Helm, R. (Ed.), 2000. Lignin-Polysaccharide Interaction in Woody Plants. Lignin: Historical, Biological, and Material Perspectives. ACS Symposium Series, Washington, DC. Higson, A., 2011. Lignin Factsheet. Holladay, J.E., White, J.F., et al., 2007. Top Value-added Chemicals from Biomass. In: Results of Screening for Potential Candidates from Biorefinery Lignin, Vol. 2. Pacific Northwest National Laboratory (PNNL) and the National Renewable Energy Laboratory (NREL). IHS-Chemical, 2012. Lignosulfonates. Kirk, T.K., Farrell, R.L., 1987. Enzymatic "Combustion": the microbial degradation of lignin. Annu. Rev. Microbiol. 41, 465e501. Koshijima, T., Watanabe, T., 2003. Association between Lignin and Carbohydrates in Wood and Other Plant Tissues. Springer-Verlag Berlin Heidelberg GmbH, New York. Kostukevich, N., Filipov, V., et al., 1993. Determination of the Hydroxyl Containing Functional Groups of the Oxygen-acetic Lignins by 31P NMR Spectroscopy. Proceedings of the Seventh International Symposium on Wood and Pulping Chemistry, Beijing, China. Kubo, S., Kadla, J., 2004. Poly(ethylene oxide)/organosolv lignin blends: relationship between thermal properties, chemical structure, and blend behavior. Macromolecules 37, 6904e6911. Lawoko, M., Henriksson, G., et al., 2005. Structural differences between the lignincarbohydrate complexes present in wood and in chemical pulps. Biomacromolecules 6 (6), 3467e3473. Li, Y., Sarkanen, S., 2002. Alkylated kraft lignin-based thermoplastic blends with aliphatic polyesters. Macromolecules 35 (26), 9707e9715. Li, J., Gellerstedt, G., et al., 2009. Steam explosion lignins; their extraction, structure and potential as feedstock for biodiesel and chemicals. Bioresour. Technol. 100 (9), 2556e2561. Liitiä, T., Maunu, S., et al., 2003. Analysis of technical lignins by twoand three-dimensional NMR spectroscopy. J. Agric. Food Chem. 51 (8), 2136e2143. Lin, S., Dence, C., 1992. Methods of Lignin Chemistry. Springer-Verlag, Heidelberg/Berlin/New York. Lora, J., Glasser, W., 2002. Recent industrial applications of lignin: a sustainable alternative to nonrenewable materials. J. Polym. Environ. 10 (1e2), 39e48. Lora, J., 2006. Biorefinery Non-wood Lignins: Potential Commercial Impact, 3e6. PAPTAC Ninety-Second. Annual Meeting. Lora, J., 2010. Utilization Opportunities for Biorefinery Lignins: an Industrial Perspective (International Lignin Biochemicals Conference, Toronto, Canada). Lucintel, 2013. Global Basic Petrochemicals Industry 2013e2018: Trend, Profit, and Forecast Analysis. Mankar, S., Chaudhari, A., et al., 2012. Lignin in phenol-formaldehyde adhesives. Int. J. Knowl. Eng. 3 (1), 116e118. Mansouri, N.-E., Salvado, J., 2006. Structural characterization of technical lignins for the production of adhesive: Application of lignosulfonates, kraft, soda-anthraquinone, organosolv and ethanol process lignins. Industrial Crops and Products. 24, 8e16. 335 Maradura, S., Kimb, C., et al., 2012. Preparation of carbon fibers from a lignin copolymer with polyacrylonitrile. Synth. Mater. 162, 453e459. Marton, J. (Ed.), 1971. Reaction in Alkaline Pulping. Lignins. Occurrence, Formation, Structure and Reactions. Wiley e Interscience, New York. Milne, T., Chum, H., et al., 1992. Standardized analytical methods. Biomass Bioenergy 2 (1e6), 341e366. Mörck, R., Yoshida, H., et al., 1986. Fractionation of kraft lignin by successive extraction with organic solvents. I. Functional groups, 13 C-NMR-spectra and molecular weight distribution. Holzforschung (Suppl. 51e60). Nitz, H., Semke, H., et al., 2001. Influence of lignin type on the mechanical properties of lignin based compounds. Macromol. Mater. Eng. 286 (12), 737e743. Otani, S., Fukuoka, K., et al., 1969. Method for Producing Carbonized Lignin Fiber. U. S. P. Office. USA. 3,461,082. Pan, X., Kadla, J., et al., 2006. Organosolv ethanol lignin from hybrid poplar as a radical scavenger: relationship between lignin structure, extraction conditions, and antioxidant activity. J. Agric. Food Chem. 54 (16), 5806e5813. Pandey, M., Kim, C., 2011. Lignin depolymerization and conversion: a review of thermochemical methods. Chem. Eng. Technol. 34 (1), 29e41. Perlack, R., Wright, L., et al., 2005. Biomass as Feedstock for a Bioenergy and Bioproducts Industry: The Technical Feasibility of a Billion-Ton Annual Supply. Podterob, A., Bogdanovich, Y., et al., 2004. Detoxicating properties of polyphepan evaluated in model experiments. Pharm. Chem. J. 38 (8), 49e54. Pu, Y., Cao, S., et al., 2011. Application of quantitative 31P NMR in biomass lignin and biofuel precursors characterization. Energy Environ. Sci. 4, 3154e3166. Rabinovich, M., 2010. Wood hydrolysis industry in the Soviet Union and Russia: a mini-review. Cellul. Chem. Technol. 44 (4e6), 173e186. Ralph, J., Lundquist, K., et al., 2004. Lignins: natural polymers from oxidative coupling of 4-hydroxyphenylpropanoids. Phytochem. Rev. 3, 29e60. Rials, T.G., Glasser, W.G., 1986. Engineering plastics from lignin. XIII Effect of lignin structure on polyurethane network formation. Holzforschung. 40, 353e360. Robert, D., Bardet, M., et al., 1988. Structural changes in aspen lignin during steam explosion treatment. Cellul. Chem. Technol. 22, 221e230. Saad, R., Radovic-Hrapovic, Z., et al., 2012. Sorption of 2,4-dinitroanisole (DNAN) on lignin. J. Environ. Sci. 24 (5), 808e813. Sakakibara, A. (Ed.), 1991. Chemistry of Lignin. Wood and Cellulose Chemistry. Marcel Dekker Inc, New York. Sannigrahi, P., Ragauskas, A., et al., 2008. Effects of two-stage dilute acid pretreatment on the structure and composition of lignin and cellulose in loblolly pine. BioEnergy Res. 1, 205e214. Sannigrahi, P., Ragauskas, A., et al., 2010. Lignin structural modifications resulting from ethanol organosolv treatment of loblolly pine. Energy Fuels 24 (1), 683e689. Seydibeyo glu, M., 2012. A novel partially biobased PAN-lignin blend as a potential carbon fiber precursor. J. Biomed. Biotechnol. Tejado, A., Pena, C., et al., 2007. Phisico-chemical characterization of lignins from different sources for use in phenol-formaldehyde resin synthesis. Bioresource Technology. 98, 1655e1663. Tomani, P., Axegard, P., et al., 2011. LignoBoost Kraft Lignin - A New Renewable Fuel and a Valuable Fuel Additive (International Bionergy & Bioproducts Conference. Atlanta, GA, USA). Tomani, P., 2010. The Lignoboost process. Cellul. Chem. Technol. 44 (1e3), 53e58.
336 18. INDUSTRIAL LIGNINS: ANALYSIS, PROPERTIES, AND APPLICATIONS Vanderlaan, M., Thring, R., 1997. Polyurethanes from Alcell lignin fractions obtained by sequential solvent extraction. Biomass Bioenergy 14, 525e531. Vappula, H., 2011. Pulp Market Review - Energy and Pulp Business Group. U. T. B. Company, Helsinki. Wallis, A., 1971. Solvolysis by acids and bases. In: Sarkanen, K., Ludvig, C. (Eds.), Lignins. Occurrence, Formation, Structure and Reactions. Wiley e Interscience, New York. Wang, Z., Xue, J., et al., 2012. Synthesis of wood lignin-ureaformadehyde resin adhesive. Adv. Mater. Res. 560e561, 242e246. Wörmeyer, K., Ingram, B., et al., 2011. Comparison of different pretreatment methods for lignocellulisic materials. Part II: influence of pretreatment on the properties of rye straw lignin. Bioresour. Technol. 102, 4157e4164. Xia, Z., Akim, L., et al., 2001. Quantitative 13C NMR of lignins with internal standards. J. Agric. Food Chem. 49, 3573e3578. Yin, Q., Di, M., 2012. Preparation and mechanical properties of lignin/epoxy resin composites. Adv. Mater. Res. 482-484, 1959e1962. Zakis, G., 1994. Functional Analysis of Lignins and Their Derivatives. Tappi Press, Atlanta, GA. Zhang, L., Gellerstedt, G., 2007. Quantitative 2D HSQC NMR determination of polymer structures by selecting suitable internal standard references. Magn. Reson. Chem. 45 (1), 37e45. Zhao, B., Chen, G., et al., 2001. Synthesis of lignin-based expoxy resin and its characterization. J. Mater. Sci. Lett. 20 (9), 859e862.
C H A P T E R 19 Amino-Based Products from Biomass and Microbial Amino Acid Production K. Madhavan Nampoothiri*, Vipin Gopinath a, M. Anusree a, Nishant Gopalan, Kiran S. Dhar Biotechnology Division, National Institute for Interdisciplinary Science and Technology (NIIST), CSIR, Trivandrum, Kerala, India, aEqual contributors *Corresponding author email: madhavan85@hotmail.com O U T L I N E Amino Acids Glutamic Acid Lysine Methionine Threonine Arginine Aromatic Amino Acids 337 338 339 340 340 340 341 Aspartame 341 Poly(Amino Acid)s Cyanophycin/Cyanophycin Granule Polypeptide Production of Cyanophycin Biodegradability of Cyanophycin Applications for Cyanophycin 341 342 343 343 343 AMINO ACIDS The amino acid industry has shown an exponential growth since its infancy in the 1950s. It has grown from extracting flavor enhancers from seaweeds, to fermenting high-purity, optically active forms in hundreds and thousands tons. The isolation of a bacterial strain producing glutamic acid and an efficient screening method to identify the highest producer by the Japanese researchers of the Kyowa Hakko Kogyo Co. was the key event in the amino acid fermentation industry. Until Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00019-X Poly-g-Glutamic Acid Production of PGA Biodegradability of PGA ε-Poly-L-Lysine Production of ε-Poly-L-Lysine Degradation of Polylysine Applications of Polylysine 343 344 344 345 345 345 345 Polyamines Putrescine Cadaverine 345 346 348 Conclusion and Perspectives 349 References 349 then, there was no suitable commercial process for the mass production of amino acids. Later on, it received further boost when the workers of the same organization reported a homoserine auxotrophic lysine producer. This discovery led to the development of a commercially viable fermentation process for lysine fermentation with a conversion efficiency of 26% from glucose. Bioprocess engineering and strain improvement methods have contributed to the massive growth of the industry. The essential amino acids hold a major place in the global amino acid market, as these cannot be 337 Copyright Ó 2014 Elsevier B.V. All rights reserved.
338 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION synthesized in the organisms and have to be supplied externally. The annual demand for feed-grade amino acids globally is about 2.43 million tons with an estimated value of US $6  109. The global amino acid market is estimated to hit US $12.8  109 by the end of 2017 (Chapman, 2012). There has been a substantial increase in the demand for amino acids in the last 30 years with a steady growth rate in the market. It is estimated that in that last 10 years the market demand for amino acids has doubled with glutamic acid and lysine on the top of the chart. Corynebacterium glutamicum is generally used for amino acid production, with an estimated annual production of 2,160,000 tons of L-glutamate and 1,480,000 tons of L-lysine (Zahoor et al., 2012). The commonly used methods for amino acid production are extraction, chemical synthesis, enzymatic conversion and fermentation. Selection of the best method depends on the cost of the raw materials used, overall production cost, purification methods adopted, marketability and demand. Cost of production and environmental impact can further be reduced by using sugars from agricultural, industrial or municipal wastes than pure and refined sugars. Microbial amino acid process for biorefining applications will lead to a cleaner environment and lesser production costs. Glutamic Acid Glutamic acid is a nonessential amino acid finding its major application in the flavor industry. It was first isolated in its pure form from wheat gluten by Ritthausen (1866) and Dr Kikunae Ikeda (1908) found that monosodium glutamate was responsible for the flavor of brown kelp used in Japanese food preparations. The discovery was soon patented, and Ajinomoto began the commercial production of monosodium glutamate from acid hydrolysate of wheat gluten and defatted soybean in 1909. At the end of the 1950s, the team led by Kinoshita, of Kyowa Hakko Kogyo, isolated the glutamate-excreting soil bacterium, C. glutamicum. This led to the beginning of the fermentative production of amino acids in large quantities, which in turn made revolutionary changes in the amino acid production and the flavor industry. The bacterium was able to accumulate large amounts of glutamate in the cell, and was excreted out of the cell when triggered either by change in temperature, addition of surfactants, antibiotics or biotin deficiency (Shimizu and Hirasawa, 2006). These triggering mechanisms altered cell wall permeability to help excrete the amino acid. Nampoothiri et al. (2002) showed that expression of genes of lipid synthesis and altered lipid composition modulates L-glutamate efflux of C. glutamicum. In the same year as Kinoshita, Donald J Kita and Jackson Heights reported glutamic acid production from Cephalosporium species (Kita and Jackson, 1957). Generally, the preferred sources of glutamic acid production were refined sugars like glucose, fructose and sucrose. In a short period, diverse substrates were explored in place of refined sugars for glutamic acid production, including starch hydrolysates, molasses hydrocarbons and methanol. Enzymatic hydrol, a waste product after enzymatic hydrolysis of grains and starch in industries was fermented by Brevibacterium divaricatum with a yield of 0.19 g glutamic acid per gram of enzymatic hydrolysate (McCutchan et al., 1962). Fermentation with wheat bran and rice bran extracts gave maximum of 50% w/w glutamic acid by Bacillus ammoniagenes (Hong et al., 1974). Glutamic acid production has already been demonstrated with palm waste hydrolysate, cassava starch hydrolysate, date waste hydrolysate and rice hydrolysate by different bacterial strains (Das et al., 1995; Nampoothiri and Pandey, 1999; Tavakkoli et al., 2009). Different fermentation methods also were developed so as to utilize agricultural wastes. Sugarcane bagasse was used as inert substrate for glutamate production by solid state fermentation (Nampoothiri and Pandey, 1996) and similarly sugarcane molasses-enriched medium was also used with carrageenan-immobilized C. glutamicum cells. Advances in the biorefinery system research basically demands value addition in the overall process. Cell surface expression of alpha amylase was reported in C. glutamicum for glutamate production, which could find use in whole crop biorefinery. The heterologous expression of pentose sugar utilizing genes in C. glutamicum helped to establish a way into the lignocellulosic feedstock biorefinery for amino acid production. Here, the heterologous expression of araBAD operon from Escherichia coli resulted in the utilization of arabinose for production of L-glutamate. Coutilization of xylose and arabinose along with glucose has also been demonstrated in rice straw and wheat bran hydrolysates for the amino acid production (Gopinath et al., 2011). Later this strain was further improved for accelerated growth and production of lysine, glutamate, ornithine and putrescine, by overexpressing xyl B of C. glutamicum along with other genes involved in pentose sugar utilization (Meiswinkel et al., 2012). A detailed review on pentose sugar utilization by C. glutamicum for production of various value-added commodities was made by Gopinath et al. (2012). The biodiesel industry generates waste glycerol streams, and the engineering of glycerol utilization pathway in C. glutamicum points toward an efficient utilization of glycerol for glutamate production (Rittmann et al., 2008). Efforts were also made for the direct utilization of cellulosic materials. The heterologous expression of Corynebacterium thermocellum endoglucanase in C. glutamicum resulted in the production of 178 mg/l glutamate from 15 g/l barley b-glucan, with the synergistic action of external b-glucosidase from Aspergillus oryzae (Tsuchidate et al., 2011).
AMINO ACIDS Lysine Lysine was first isolated from casein by Drechsel (1889). However, microbial conversion of a-amino adipic acid and diaminopimelic acid to lysine was reported much later (Haulaham and Mitchell, 1948; Davis, 1952). Microbial fermentation was initiated, when the Kyowa Hakko Kogyo group reported a homoserine auxotroph of C. glutamicum that produced increased amounts of lysine. Classical mutagenesis, strain improvements and metabolic pathway engineering led to increased production of lysine (Wittmann and Becker, 2007). Genes coding for enzymes in the amino acid biosynthetic pathway were overexpressed or disrupted, either alone or in combinations for amino acid overproduction. For example, the identification and cloning of 339 lysine exporter gene lysE helped to increase lysine excretion (Vrljic et al., 1996). Point mutations in the pyc, lysC and hom genes (Figure 19.1) were introduced into the wild-type C. glutamicum chromosome to get the engineered lysine producer DM1729 (Georgi et al., 2005). Pyruvate carboxylase and phosphoenol pyruvate carboxykinase are two anaplerotic enzymes in Corynebacterium for growth on carbohydrates. The results from various studies indicate that overexpression of pyc gene encoding pyruvate carboxylase redirects the carbon flux toward lysine production and overexpression of pck encoding phosphoenol pyruvate carboxykinase is counteractive to amino acid production (Peters Wendisch et al., 1998, 2001; Riedel et al., 2001). Substrate range for lysine production is very vast including dextrose, sucrose and fructose as the refined FIGURE 19.1 Genes involved in lysine biosynthetic pathways in Corynebacterium glutamicum. pyc, pyruvate carboxylase; lysC, aspartate kinase; dapA, dihydrodipicolinate synthase; dapD, succinyl transferase; dapC, aminotransferase; dapE, desuccinylase; dapF, epimerase; ddh, diaminopimelate dehydrogenase and hom, homoserine dehydrogenase. (For color version of this figure, the reader is referred to the online version of this book.)
340 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION sugars. Organic acids such as acetic acid, propionic acid, benzoic acid, formic acid, malic acid, citric acid and fumaric acid; alcohols such as ethanol, propanol, inositol and glycerol and hydrocarbons, oils and fats such as soybean oil, sunflower oil, groundnut oil and coconut oil as well as fatty acids such as palmitic acid, stearic acid and linoleic acid were also used. These substances may be used individually or as mixtures (Anastassiadis, 2007). An array of nondefined sugar substrates like cane and beet molasses, blackstrap molasses and starch hydrolysates were used for industrial fermentation (Ikeda, 2003). Inability of lysine-producing microbes was circumvented by engineering them for direct starch utilization. The expression of amylase genes on the cell surface of C. glutamicum enabled simultaneous saccharification and fermentation of raw corn starch, potato starch and sweet potato starch, rather than using starch hydrolysates and refined sugars (Tateno et al., 2007; Berens et al., 2001). In the direction of enabling the efficient utilization and conversion of hemicellulosic biomass-derived sugars, C. glutamicum has been engineered to utilize arabinose and xylose. In the fermentation section of biorefinery, downstream processes with least steps help in better energy efficiency, low product cost and waste management exemplified by spray dried lysine with 78% purity called “Biolys”. Under the greenbiorefineries, the brown juice produced from the green crop drying units is acidified with lactic acid fermentation by Lactobacillus species and is stored and transported for lysine fermentation (Thomsen et al., 2004). Biodiesel industry generated glycerol can also be utilized for lysine production. Methionine Methionine is a limiting amino acid in the monogastric’s feed. Its isolation was first reported from casein (Mueller, 1923) and since then methionine was commercially produced either by chemical synthesis or enzymatic methods. Unlike other amino acids, methionine has an advantage that it can be supplied to animal feed as a racemate or a racemic mixture as the mammals are able to convert it to utilizable form with a methionine racemase enzyme. Even so, the microbial production has added advantages over the racemate that it helps optimal nutrient utilization. With the discovery of fermentative production of amino acids by Kinoshita, attempts were being made to commercialize methionine production by submerged fermentation. The initial attempts on microbial fermentation were done in the 1970s (Kase and Nakayama, 1974). As amino acid synthesis is an energy expensive process and is feedback inhibited, the wild-type bacteria were not reported to overproduce methionine. Hence, overproduction was achieved either by classical mutagenesis or deregulation of the biosynthetic pathway. Species of Corynebacterium and Brevibacterium have much simpler regulatory mechanisms than E. coli for methionine biosynthesis and are the preferred microbes for overproduction. Corynebacterium is also able to switch between the transsulfuration pathway and the direct sulfhydrylation for methionine production (Hwang et al., 2002). Even though attempts are still being made to tailor the methionine producers to utilize raw sugars from complex sources, glucose and maltose are most extensively used. Substrates for fermentation varied over coconut water, banana, cassava, molasses, sugarcane juice, etc. (Pham et al., 1992). Methanol and n-alkanes were also used (Morinaga et al., 1982; Ghosh and Banerjee, 1986). Threonine Threonine is the third limiting amino acid after lysine and methionine in the animal feeds and was discovered by W.C. Rose (Rose, 1931). Presently, the major share of commercially produced threonine employs fermentative production. The microbial strains employed in the production process were genetically manipulated for threonine overproduction including Serratia marcescens (Komatsubara et al., 1978), C. glutamicum and E. coli (Dong et al., 2011). Escherichia coli dominates as the threonine producer due to their dynamic growth pattern and better substrate utilization, but cannot be used in the synthesis of pharmaceutical grade amino acid owing to their endotoxin production. The generally recognized as safe status, the highly defined genome database and scope for further genetic manipulations make C. glutamicum dominate the pharmaceutical grade threonine production. Generally the strains employed in methionine fermentation utilize directly available monosaccharide sugars. The broadening of substrate utilization range to polysaccharides and biomass-derived sugars including pentoses is highly desired, as this helps in cost reduction and flexibility of the overall industrial production process. Most importantly it will help utilize the renewable sugar sources that will otherwise go underutilized. But this will require manipulations in the sugar uptake systems of the microorganisms under concern, as both the industrial strainsdE. coli and C. glutamicumdare unable to directly utilize polysaccharides. Escherichia coli strains are reported to have pentose utilization systems, but this capability has to be incorporated in the strains with amino acid production ability and resistance to biomass pretreatment-derived inhibitors. Thus, the extended substrate utilization spectrum will be a step toward production of threonine from biorefinery. Arginine Isolation of L-arginine was first reported from lupin seedlings. The analog-resistant mutants of E. coli,
341 POLY(AMINO ACID)S Saccharomyces cerevisiae and Bacillus subtilis excreted arginine in production medium, but were lower than C. glutamicum (Utagawa, 2004). An arginine overproducer with an argR (arginine repressor) deletion and expression of feedback-insensitive N-acetylglutamate kinase was modified with E. coli araBAD. The recombinant strain was able to utilize 77% of arabinose from the production medium to accumulate arginine, when glucose and arabinose were used as the sole carbon source. Arginine production can be further increased by the overexpression of argininosuccinate synthase gene. An ornithine overproducer C. glutamicum argFR double mutant was manipulated to utilize arabinose as the carbon source (Schneider et al., 2011). It was able to utilize 98% arabinose (244 mM) from the production medium. The ability of arabinose utilization points to the possibilities of amino acid production from lignocellulosic hydrolysates. The ability of simultaneous utilization of mixed sugars (Sasaki et al., 2008) and ability to withstand lignocellulosic pretreatment-derived inhibitors were demonstrated earlier in C. glutamicum. The integration of amino acid production excellence with these characters will boost the chances of amino acid production from the lignocellulosic feedstock biorefinery. Aromatic Amino Acids Aromatic amino acids include tryptophan, phenylalanine, tyrosine and histidine and all of them were isolated in the 1880s. Tyrosine was isolated by Liebig in 1846 and phenylalanine from lupins by Schulze in 1881. While histidine was reported by Kossel and Hedin in 1896, tryptophan was first isolated from casein by Frederick Hopkins in 1901. Even though the aromatic amino acids are produced by microbial fermentation, high production levels are not reached. Commercial production of these amino acids also includes extraction and enzymatic conversion. In 2006, Kyowa Hakko Kogyo claimed development of the world’s first fermentation-based method for the commercial production of L-tyrosine. Biodiesel industry-generated crude glycerol can be biorefined by phenylalanine-producing E. coli cells (Khamduang et al., 2009). The use of glycerol resulted in phenylalanine yields up to 0.58 g/g, which is twice as compared to production with sucrose. Tryptophan and histidine were produced from mixed sugars pentoses and hexoses by genetically modified E. coli (Savrasova et al., 2004). Microbial amino acid production process for biorefining application will be technically feasible, only if the nutrient requirements are met in invariably same quantity and quality. These raw material production costs must be lower than starch hydrolysates or refined sugars and the coproducts have to use the existing fermentation machinery and infrastructure for economic feasibility. Most of the microbial amino acid fermentations occurs at a temperature range of 30e40  C and at near neutral pH. The fermentation medium has to be neutralized and proper cooling systems for temperature maintenance, agitation and aeration has to be in place for the amino acid production to match with the expected values. Alternatives are development or utilization of strains that are pH and temperature tolerant and overproduce amino acids or choosing coproducts and organisms having the same substrate utilization spectrum and physical requirements. The use of microbial amino acid fermentation for biorefining resulted in improved ground water quality, lower ammonia and nitrate excretion from poultry and livestock. This is due to the substitution of optimal quantities of the limiting amino acids in place of soybean meal. The sugar solutions from the lignocellulose feedstock biorefinery will have a fair concentration of inhibitors like furfural and hydroxymethyl furfurals and syringaldehyde, which is toxic to the microorganisms. The inhibitor tolerance of amino acid producers will also be a deciding factor in the biorefining concept. Fortunately, the amino acid producer C. glutamicum has shown tolerance toward inhibitors at growtharrested conditions and high cell densities. ASPARTAME Aspartame is a methyl ester of dipeptides consisting of aspartic acid and phenylalanine. It was accidentally discovered in 1965 by the chemist James M. Schlatter while working on an antiulcer drug (Walters, 1991). Aspartame is 160e220 times sweeter than sucrose and is used as artificial sweetener in foods and beverages. Aspartame is produced by coupling microbial fermentation and synthesis. Phenylalanine and aspartic acid are produced by microbial fermentation and phenylalanine is reacted with methanol to form the methyl ester. Aspartic acid is also treated in such a way as to protect active sites by benzyl rings. Then the modified amino acids are mixed in a reaction tank at appropriate temperatures to get aspartame intermediates (Figure 19.2). It is further treated with acetic acid, purified, crystallized and powdered to produce aspartame. Methods for direct enzymatic synthesis and chemical synthesis are also reported. POLY(AMINO ACID)S In a microbial (or higher) system, proteins are synthesized with the help of an information template in the form of an mRNA molecule and the translatory machinery of ribosomes and amino acylated tRNA; however, poly(amino acid)s are synthesized enzymatically without the requirement of an information template.
342 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION FIGURE 19.2 Aspartame synthesis. (For color version of this figure, the reader is referred to the online version of this book.) While proteins owe their functionality to the tertiary structure that they attain, the poly(amino acid)s owe it normally to their physical properties, usually the number of repeating units more dominant in a pool of poly (amino acid)s. Another striking difference between proteins and poly(amino acid)s is the fact that one type of a protein will contain an exact number of amino acids, while poly(amino acid)s display wide polydispersity, i.e. in a single organism, the size of the same poly (amino acid) will vary. The fact that these polymers are biocompatible with human physiology and for applications other than pharmacology, the polymer is biodegradable, makes it an attractive alternative to the widely used polymers obtained through the petrochemical route, which are often nonbiodegradable, and end up accumulating in the environment. These poly(amino acid)s have been shown to be useful in multitude of applications, like controlled drug release, preparation of bioplastics, use as antimicrobial additive in food, as well as superabsorbers (replacement of polyacrylate gels) (Obst and Steinbüchel, 2004). They may also be used as a viable source of dipeptide neutraceuticals (Sallam and Steinbüchel, 2010). There are three identified poly(amino acid)s so far that have been reported to be produced from microbial source; they are cyanophycin, poly-g-glutamate (PGA) and ε-poly lysine (Figure 19.3). In poly(glutamic acid) the amide linkages are formed on the a-amino group to then g-carboxyl group in the polymer backbone, whereas in poly(lysine), the a-carboxyl group is linked to the ε-amino group of lysine. In the case of cyanophycin, almost equivocal amounts of arginine and aspartic acid are arranged as a polyaspartate backbone, with arginine moieties linked to the b-carboxyl group of almost every aspartic acid residue. Cyanophycin/Cyanophycin Granule Polypeptide Cyanophycin is the ideal nitrogen storing molecule, because every repeating unit has about five atoms of nitrogen, and it is insoluble at physiological conditions inside the cell protoplasm. Due to its insolubility, it does not cause detrimental shifts in the osmolarity of the cell, hence helps in cell survival (Oppermann-Sanio and Steinbüchel, 2002). Cyanobacteria normally produce cyanophycin when the organism senses a decrease in sulfur, phosphate and significantly by the decrease in nitrogen concentration in the surrounding milieu (Lin et al., 2012). Apart from cyanobacteria, cyanophycin granule polypeptide (CGP) has been found in some strains of Synechococcus sp. (Hai et al., 1999). FIGURE 19.3 Amino compounds from microbial sources. (For color version of this figure, the reader is referred to the online version of this book.)
POLY(AMINO ACID)S Production of Cyanophycin Cyanophycin is enzymatically produced by the action of cyanophicin synthases on small primers of cyanophycin. Because of the slow growth of cyanobacteria in photobioreactors, large-scale production of CGP is hindered by the lower cell densities and thus used to get only low yield of CGP with respect to the cell dry mass (CDM). To solve this problem cyanophycin synthase gene or CphA gene has been identified and cloned into a wide range of organisms, from E. coli to eukaryotic microbes like S. cerevisiae (Steinle et al., 2008), commercially important strains like Ralstonia eutropha, C. glutamicum, Pseudomonas spp., etc. (Aboulmagd et al., 2001) and even higher plants like potato and tobacco (Neumann et al., 2005). Recombinant CGP produced were shorter in length (21e35 repeating units), along with a small percentage of arginine replaced by incorporated lysine. Till some time, the highest CGP production was attained by using Acinetobacter calcoaceticus, with a value of 48% CDM, with the addition of exogenous arginine, along with the addition of other carbon and nitrogen sources (Elbahoul et al., 2005). The economic viability of recombinant strains is always a problem for largeescale usage, due to the sheer amount of antibiotics that have to be added to the medium. However, the hunt for a commercially compatible strain for economical production of CGP led to a method for obtaining high cell densities and high yield of CGP with Ralstonia eutropha. The strain is poly hydroxy butyrate (PHB) negative (the wild type produces PHB) and was devoid of the 2-keto-3-deoxyphosphogluconate aldolase (eda) gene (Lin et al., 2012). The plasmid carrying the cphA gene was constructed along with the eda gene; for the microbe to survive, the plasmid had to be retained, and in other words, a selective pressure for plasmid retention was achieved, without the use of antibiotic resistance genes, thus making the process economically viable. In general, the production was optimized with a basic mineral medium, Mineral Salts Medium (MSM), with sufficient supplements of fructose, NH3, K2SO4, MgSO4.7H2O, Fe(III) NH4-citrate, CaCl2.2H2O, and trace elements. A 30 l pilot study gave promising yields of water-insoluble CGP and water-soluble CGP, contributing to 47.5% and 5.8% (w/w) of CDM, and a cell density as high as 57 g/l CDM was obtained (Elbahloul et al., 2005). CGP is normally purified by acid extraction, which involves the solubilization of CGP in acidic solutions of pH 1, followed by washes with distilled water, which renders it insoluble again. Alternative production strategies include the use of molecular farming approach, and expression of the cphA gene in certain specific tissues of selected plants. Potato and tobacco have been successfully transformed with the cphA gene; however, production of CGP within 343 the plant tissues would lead to slow growth and fleshy leaves. Use of the stable cyanophycin synthase for direct synthesis of cyanophycin has been suggested by certain authors as an alternative to the intricately controlled fermentative production of cyanophycin (Hai et al., 2002). Biodegradability of Cyanophycin Biodegradation of cyanophycin is observed in all the organisms that naturally produce the polymer, as it serves as a reserve carbon and nitrogen pool. Cyanophycin can be depolymerized by intracellular cyanophycinase, which also has been isolated from Synechocystis sp. strain PCC6308 (Hai et al., 2002). The cyanophycinase gene or the cphE gene has been found to be located downstream of the cphA gene. The cyanophycinase enzyme does not cleave the polymer into arginine and aspartate. Recent studies have shown that cyanophycin is degraded by most of the gut bacteria through the anaerobic route within a time period ranging from 1 day to 7 days (Sallam and Steinbüchel, 2009). The above studies have opened the doors to the use of cyanophycin directly as nutrient supplement. Applications for Cyanophycin Cyanophycin can be hydrolyzed to its constituent amino acids, aspartic acid, and arginine. These amino acids may be utilized directly in food and pharmaceutical applications. Cyanophycin can be stripped off of arginine through chemical modifications, so as to produce polyaspartate. Polyaspartate is a polyanionic polymer that can be utilized for production of biodegradable surfactants, and can be utilized for applications pertaining to polyacrylate (Schwamborn M, 1998). It has been discovered recently that cyanophycin can be degraded by the gut bacteria obtained from a diverse group of organisms ranging from mammals, birds, and fishes. This opens routes for the use of cyanophycin directly as a nutritional substrate, instead of constituent dipeptides or amino acids, which would require additional investments of time and money (Sallam A and Steinbüchel A 2009). Even though a considerable amount of research has been carried out for the viable production of cyanophycin, the complete potential for the various bulk chemicals that may be obtained from cyanophycin is not attained yet. Poly-g-Glutamic Acid PGA was first observed by Ivanovics et al. produced by Bacillus anthracis strain during 1937. The traditional Japanese food natto (fermented product made from soyabean) consists of PGA and a fructan, and is produced through fermentation using the strain Bacillus natto. PGA produced can be essentially of three types,
344 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION a polymer purely made of either D-glutamic acid, L-glutamic acid, or D and L-glutamic acid. Besides Bacillus, other microbial species from Archaea (Natrialba aegyptiaca, Natronococcus occultus), eubacteria as well as eukaryotes (hydra) were found to produce PGA (Kocianova et al., 2005; Hezayen et al., 2001). PGA is one of the most utilized poly(amino acid)s, with multitude of applications ranging from agriculture to cosmetics, food industry and even pharmaceuticals (Buescher and Margaritis, 2007). Production of PGA PGA can be produced only through the microbial route, unlike alpha poly glutamic acid, which can be chemically synthesized. Among PGA producers, a broad classification was established in the form of glutamic acid-dependent PGA producers and glutamic acidindependent PGA producers, depending on the mandatory requirement or nonrequirement for glutamic acid as a major component in the nutrient medium. Examples of glutamic acid-dependent producers of PGA include B. subtilis IFO 3335, B. subtilis NX-2, and Bacillus licheniformis ATCC 9945A and examples of glutamic acid-independent PGA producers include B. subtilis TAM-4 and B. licheniformis SAB-26. Initial studies for the production of PGA was carried out using the strain B. licheniformis ATCC 9945A. The production medium formulated for PGA production contains a relatively high concentration (20e1230 mM) of Mnþ2 and the more it was, the amount of D-glutamic acid present as well (Leonard et al., 1958a,b). Bacillus subtilis IFO 3335 was originally isolated from natto, a traditional fermented food in Japan, which has mucilage containing PGA and a levan. PGA productivity of this strain was higher than that of B. licheniformis ATCC 9945A. Bacillus subtilis IFO 3335 could produce PGA at levels of 9.6 g/l, with an optimized medium with major components including 30 g/l of L-glutamic acid, 20 g/l of citrate, and 5 g/l ammonium sulfate. It was found that L-glutamic acid merely stimulated the production of PGA, which could initiate PGA productions at lower concentrations (0.1 g/l), without the addition of 30 g/l of L-glutamic acid (Kunioka and Goto, 1994; Kunioka, 1995). Cheng et al. (1989) isolated B. licheniformis A35 while looking for amino acid producer under denitrifying conditions. A strain with extremely high production rates of PGA was isolated from fermented bean curd (Shi et al., 2006). The strain was found to be B. subtilis ZJU7. In an optimized culture medium containing 60 g/l sucrose, 60 g/l tryptone and 80 g/l L-glutamic acid and after cultivated at 37  C for 24 h, the yield of g-PGA reached 54.4 g/l. In an interesting study, a coculture of C. glutamicum S9114 along with B. subtilis ZJU7 was attempted to reduce the input costs of addition of L-glutamic acid into the medium, and it yielded 32.8 g/l of PGA after 24 h of fermentation (Shi et al., 2007). Various studies regarding the fermentative production, downstream processing and characterization of PGA have been reported in the literature. The review by Bajaj and Singal (2011) provides updated information on fermentative production of PGA by various bacterial strains and effect of fermentation conditions and media component on production of PGA in submerged as well as solid state fermentation. PGA can be extracted from the fermentation medium by two different methods. A crude method involves centrifuging cells followed by precipitation of PGA by methanol or ethanol (chilled) (Goto and Kunioka, 1992). Subsequent purification steps, involving gel permeation chromatography followed by reprecipitation have to be used to get PGA in a pure form. Another method exploiting the specific interaction of Cuþ2 ions with PGA, gives relatively purer PGA (Troy, 1973). Further, the purity of the polymer is checked through peptide hydrolysis, followed by thin layer chromatography, to assess the constituent amino acids. Large amounts of PGA are produced microbially by the Japanese company Meiji Seika Kaisha Ltd employing B. subtilis strain F-021. Biodegradability of PGA PGA like the other poly(amino acid)s is a degradable polymer. The polymer can withstand temperatures up to 60  C, beyond which the amide bonds start getting hydrolyzed. PGA is resistant to proteases that cleave alpha peptide bonds. Two types of enzymes are involved in the degradation of PGA, endo-g-glutamyl peptidase and exo-g-glutamyl peptidase. Exo-g-glutamyl peptidase consists of two subunits and is a key enzyme in glutathione metabolism (Ogawa et al., 1991; Xu and Strauch, 1996). This enzyme catalyzes the formation of g-glutamic acid di- and tripeptides in vitro. Endo-gglutamyl peptidase is secreted into the medium by g-PGA-producing B. subtilis and B. licheniformis. It subsequently cleaves high molecular weight g-PGA into fragments as small as 105 Da (Goto and Kunioka, 1992). Attempts to isolate microbes that can utilize PGA as the sole source of carbon and nitrogen source were also successful (Obst and Steinbüchel, 2004). APPLICATIONS OF PGA PGA has been a keen interest of research as far as applications are concerned and hence a plethora of applications for this biopolymer has been developed. PGA has been used in the food industry as an additive to flour to increase the moisture-retaining capacity of the dough, as well as to improve the texture and shelf life of bread. The calcium salt of g-PGA can be added to health food in order to increase the Ca2þ concentration, thus contributing to the prevention of osteoporosis (Ashiuchi et al., 2004). Addition of PGA improved the solubility
345 POLYAMINES and hence the availability of vitamins and also caused sustained release of these vitamins, which led to increased absorption of these vitamins. PGA salts are known to be used as antifreeze agents in food. The antifreeze action of the salt increases with the decreasing size of the salt of the polymer (Shih et al., 2003). PGA has been suggested for water treatment, as PGA complexes with a lot of metal ions, like Caþ2, Feþ3, Alþ3, etc. (Kunioka, 2004). Esters of PGA have been used to test the ability of PGA to form bioplastics with required properties (Kubota et al., 1995). Hydrogels that can be used for applications such as controlled drug release, biosensors, diagnostics, and bioseparators can be produced by using g-PGA and poly(ethylene glycol)methacrylate (Yang et al., 2002). PGA has been used as adjuvants in vaccines, and also as a delivery agent for hydrophobic drugs, increasing their bioavailability. PGA has also been used as a medical adhesive for surgical wounds (sutureless wound closure). PGA hydrogels can be used as three-dimensional scaffolds for tissue engineering (Matsusaki et al., 2005). ε-Poly-L-Lysine The polymer was discovered by (Shima and Sakai, 1977). The polymer has been found to be heat stable, and can even withstand autoclaving for 20 min. These properties led to the use of polylysine as a food preservative on a commercial scale in Japan (Yoshida and Nagasawa, 2003). Production of the compound is influenced substantially by the pH of the medium. Polylysine has not known to form any secondary or tertiary structure, and its microbicidal activity is attributed to its polycationic nature. Production of ε-Poly-L-Lysine The strain Streptomyces albulus 346 spp. lysinopolymerus was the first organism to be isolated as a polylysine producer, following which many improvements were carried out on the same for industrial production of polylysine. Later more producers from the genus Streptomyces, Kitasatospora, and an ergot fungus epichole were found to produce polylysine. Polylysine is now industrially produced by aerobic fermentation, using a mutant derived from S. albulus 346, isolated from soil. A maximum amount of polylysine, 0.5 g/l was reported under optimized conditions set at pH 6.0 (Shima and Sakai, 1981). A mutant strain resistant to S-(2-aminoethyl)-L-cysteine, an analogue for lysine and glycine were derived and gave higher productivity values of up to 20 mg/l of polylysine, after 120 h, with glucose as the carbon source and ammonium sulfate as the nitrogen source (Hiraki et al., 1998). Streptomyces albulus 410, the strain that has been exploited for commercial production of polylysine displayed the two-stage polylysine production (pH 6.0 and pH 3.0e5.0). Accurate control of the process and pH led to a maximal production of about 48.3 g/l. It was found that the best pH for increasing the cellular mass was pH 6.0, while pH below 4.2 was favorable for high levels of polylysine production (Shi et al., 2007). Every poly(amino acid) produced is poly dispersed (has variable molecular weight), which in turn makes it difficult to obtain the product of the required specification. This problem was recently tackled to an extent, when new isolates belonging to the genus Streptomyces was obtained, which could produce nearly monodispersed polylysine. Degradation of Polylysine There is not much report on the biodegradation of polylysine. The polylysine-resistant strain Chryseobacterium sp. OJ7 also was postulated to have a polylysinedegrading enzyme, with an exopeptidase activity (Obst and Steinbüchel, 2004). Polylysine was shown to be susceptible and was degraded by commercially available enzymes proteases A, P and peptidase R from A. oryzae, Aspergillus melleus and Rhizopus oryzae (Kito et al., 2002). Applications of Polylysine The application of polylysine as a food preservative has been established. Other uses of polylysine include the production of an emulsifying agent through its conjugation with dextran. Polylysine can be used to coat biochips and surfaces of cell culture flasks, so as to provide a biocompatible adherent surface. One of the most industrially relevant applications of polylysine is its use as a drug delivery agent and an aid in cell transformation due to its polycationic nature. POLYAMINES Alkaline organic compounds with an aliphatic, saturated carbon backbone having at least two primary amino groups, and a varying number of secondary amino groups are referred to as polyamines (Schneider and Wendisch, 2011). The polyamines were first discovered by Antonie van Leeuwenhoek (1678) when he isolated some “three-sided” crystals (sperminephosphate crystals) from human semen. The charge on the polyamines is distributed along the entire length of the carbon chain, making them unique and distinct from the point charges of the cellular bivalent cations. Their positive charge enables polyamines to interact electrostatically withpolyanionicmacromolecules within the cell. Due to this they can modulate diverse cellular processes such as transcription and translation (Wallace et al., 2003), biosynthesis of siderophores (Brickman and
346 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION Armstrong, 1996), take part in acid resistance (Foster, 2004), protect from oxygen toxicity (Jung et al., 2003), etc. They have a role in signaling for cellular differentiation (Sturgill and Rather, 2004) and are essential for plaque biofilm formation (Patel et al., 2006). They are also found as a part of gram-negative bacterial outer membranes (Takatsuka and Kamio, 2004). Transgenic activation of polyamine catabolism profoundly disturbs polyamine homeostasis in most tissues, creates a complex phenotype affecting skin, female fertility, fat depots, pancreatic integrity and regenerative growth (Janne et al., 2004). In the nucleosome, polyamine depletion results in partial unwinding of DNA and unmasking of sequences previously buried in the particle. These sequences are potential binding sites for factors regulating transcription (Morgan et al., 1987). This, together with the fact that polyamines favor the formation of triplex DNA at neutral pH, may provide a mechanism whereby polyamines regulate the transcription of growth regulatory genes such as c-myc (Hampel et al., 1991; Celano et al., 1992). Since polyamines play a wide range of activities in a living cell their relative intracellular concentrations may vary from species to species, and they can reach up to the millimolar range (Miyamoto et al., 1993). The most common polyamines in bacteria and Archaea are putrescine (a diamine also named as 1,4-diaminobutane) and cadaverine (diamine also named 1,5-diaminopentane) (Figure 19.3). In addition to the above-mentioned polyamines, the pathways for the biosynthesis of 1,3-diaminopropane, norspermidine, homospermidine, and thermine are known in some bacteria and Archaea (Tabor and Tabor, 1985). The polyamine family also contains a number of uncommon longer or branched-chain polyamines, which were found in extremophiles and which seem to play an essential role for growth under such extreme conditions (Oshima, 2007). Polyamines are found in all living species, except two orders of Archaea, Methanobacteriales and Halobacteriales (Hamana and Matsuzaki, 1992). Polyamines are used in a wide variety of commercial applications due to their unique combination of reactivity, basicity, and surface activity. With a few exceptions, they are used predominantly as intermediates in the production of functional products (e.g. polyamides/epoxy curing, fungicide, anthelmintics/pharmaceuticals, petroleum production, oil and fuel additives, paper resins, chelating agents, fabric softeners/surfactants, bleach activator, asphalt chemicals) (Kroschwitz and Seidel, 2004). The main commercial interest in biogenic polyamines is their use in the polymer industry. Today, the only example of an industrial polyamide containing a biogenic diamine, which can also be synthesized by bacteria, is nylon-4, 6. This polyamide is produced from putrescine and adipic acid (hexanedioic acid). Putrescine Putrescine apparently has a specific role in skin physiology and neuroprotection (Janne et al., 2005). Fermentative production of putrescine can be achieved by manipulating arginine decarboxylase (ADC) pathway or ornithine decarboxylase (ODC) pathway in E. coli (Figure 19.4(a)) and C. glutamicum (Figure 19.4(b)) and of which, the ODC pathway is preferable as it comprises only a single reaction compared to two or three reactions of the ADC pathway. To increase L-ornithine formation, its conversion to L-arginine may be blocked; however, this results in unfavorable auxotrophy for L-arginine. Thus, the maintenance of prototrophy with concomitant high L-ornithine supply is a focus in strain construction. The pathway for biosynthesis of L-arginine and L-ornithine, the substrates of the initial decarboxylase reactions in the ADC and ODC pathway, respectively, are similar in E. coli and C. glutamicum. There is some difference in ornithine synthesis between them; C. glutamicum has a cyclic pathway while E. coli has a linear pathway (Glansdorff and Xu, 2007). The cyclic pathway is economical in terms of metabolic cost for ornithine synthesis when compared to linear pathway, because in linear pathway there is a concomitant hydrolysis of acetyl-CoA to acetic acid. The L-ornithine was then converted to citrulline by ornithine carbamoyl phosphate transferase ArgF (EC 2.1.3.3). The synthesis of all enzymes in the pathway is subject to repression by L-arginine, which is mediated by the repressor ArgR in E. coli and C. glutamicum (Glansdorff and Xu, 2007). In order to use a microorganism in industrial fermentations or biotransformations the organism should possess high tolerance to the desired product. Concentrations of up to 66 g/l putrescine reduced the growth rate of C. glutamicum by 34% and that of E. coli by 78% (Schneider and Wendisch, 2010). In order to overproduce putrescine in E. coli, several attempts have been done so far. The ADC pathway is completed by agmatinase SpeB, which hydrolyzes agmatine to putrescine and urea. While urea cannot be reused by E. coli, putrescine can be utilized by E. coli as a sole carbon source. The overexpression of ODC genes speC (b2965) and of speF (b0693) in the wildtype genetic background led to comparable results as 0.72 or 0.87 g/l of putrescine accumulated in batch cultures (Eppelmann, 2006). The simultaneous overexpression of speF and speAB, the ADC encoding gene speA as well as speB coding for the agmatinase of E. coli (b2938, b2937) increased putrescine accumulation up to 1.03 g/l (Eppelmann et al., 2006). In order to increase the putrescine production a base strain was constructed, by inactivating the putrescine degradation and utilization pathways, and the engineered E. coli strain was able to produce 1.68 g/l of putrescine with a yield of 0.168 g/
POLYAMINES 347 FIGURE 19.4 (a) Engineered putrescine and cadaverine production pathways used in E. coli. GdH, glutamic acid dehydrogenase (EC1.4.1.4); ArgA, amino acid N-acetyltransferase (EC 2.3.1.1); ArgB, acetylglutamic acid kinase (EC 2.7.2.8); ArgC, N-acetylglutamylphosphate reductase (EC 1.2.1.38); ArgD, acetylornithine aminotransferase (EC 2.6.1.11); ArgE, acetylornithinase (EC 3.5.1.16); ArgF, ornithine carbamoyltransferase (EC 2.1.3.3); ArgG, argininosuccinic acid synthetase (EC 6.3.4.5); ArgH, argininosuccinic acid lyase (EC 4.3.2.1); Pepck, phosphoenolpyruvic acid carboxykinase (EC 4.1.1.32); Ppc, phosphoenolpyruvic acid carboxylase (EC 4.1.1.31); Pyc, pyruvic acid carboxylase (EC 6.4.1.1); AspC, aspartic acid aminotransferase (EC 2.6.1.1); LysC, aspartokinase (EC 2.7.2.4); Asd, aspartic acid semialdehyde dehydrogenase (EC 1.2.1.11); MetL, ThrA bifunctional aspartokinase/ homoserine dehydrogenase (EC 2.7.2.4/1.1.1.3); DapA, dihydrodipicolinic acid synthase (EC 4.2.1.52); DapB, dihydrodipicolinic acid reductase (EC 1.3.1.26); DdH, meso-diaminopimelic acid dehydrogenase (EC 1.4.1.16); LysA, diaminopimelic acid decarboxylase (EC 4.1.1.20); ODC, ornithine decarboxylase (EC 4.1.1.17); ADC, arginine decarboxylase (EC 3.5.3.1). (b) Engineered putrescine and cadaverine production pathways used in C. glutamicum. GdH, glutamic acid dehydrogenase (EC1.4.1.4); ArgJ, bifunctional ornithine acetyltransferase/N-acetylglutamic acid synthase (EC 2.3.1.35/2.3.1.1); ArgB, acetylglutamic acid kinase (EC 2.7.2.8); ArgC, Nacetylglutamylphosphate reductase (EC 1.2.1.38); ArgD, acetylornithine aminotransferase (EC 2.6.1.11); ArgE, acetylornithinase (EC 3.5.1.16); ArgF, ornithine carbamoyltransferase (EC 2.1.3.3); ArgG, argininosuccinic acid synthetase (EC 6.3.4.5); ArgH, argininosuccinic acid lyase (EC 4.3.2.1); Pepck, phosphoenolpyruvic acid carboxykinase (EC 4.1.1.32); Ppc, phosphoenolpyruvic acid carboxylase (EC 4.1.1.31); Pyc, pyruvic acid carboxylase (EC 6.4.1.1); AspC, aspartic acid aminotransferase (EC 2.6.1.1); LysC, aspartokinase (EC 2.7.2.4); Asd, aspartic acid semialdehyde dehydrogenase (EC 1.2.1.11); MetL, ThrA, bifunctional aspartokinase/homoserine dehydrogenase (EC 2.7.2.4/1.1.1.3); DapA, dihydrodipicolinic acid synthase (EC 4.2.1.52); DapB, dihydrodipicolinic acid reductase (EC 1.3.1.26); DdH, meso-diaminopimelic acid dehydrogenase (EC 1.4.1.16); LysA, diaminopimelic acid decarboxylase (EC 4.1.1.20); ODC, ornithine decarboxylase (EC 4.1.1.17); ADC, arginine decarboxylase (EC 3.5.3.1).
348 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION g glucose. A further optimization by 25% was achieved by promoter exchange of genes encoding the enzymes converting L-glutamic acid into L-ornithine, as well as the exchange of speFepotE promoter (potE encodes the ornithineeputrescine antiporter) (Qian et al., 2009). In contrast to E. coli, C. glutamicum is unable to degrade and utilize putrescine as a carbon source. The expression of genes of the ADC and ODC pathway from E. coli in the wild-type background of C. glutamicum only led to minor amounts of putrescine. The deletion of argR and argF led to accumulation of L-ornithine but rendered the resulting strain arginine auxotrophic. When speC and speF from E. coli were expressed in the argReargF deletion strain of C. glutamicum, production of 5 g/l putrescine resulted, which was about 50 times higher than strains endowed with the ADC pathway. To avoid costly supplementation with L-arginine and the strong feedback inhibition of the key enzyme N-acetylglutamate kinase (ArgB) by L-arginine, a plasmid addiction system for low-level argF expression was developed. This strain resulted in putrescine yields on glucose from less than 0.001 up to 0.26 g/g, the highest yield in bacteria reported to date and was named as PUT21. In fed-batch cultivation with C. glutamicum PUT21, a putrescine titer of 19 g/l at a volumetric productivity of 0.55 g/l h and a yield TABLE 19.1 of 0.16 g/g glucose was achieved (Schneider et al., 2012). Moreover, while plasmid segregation of the initial strain required antibiotic selection, plasmid segregation in C. glutamicum PUT21 was fully stable for more than 60 generations without antibiotic selection even in the presence of L-arginine. Cadaverine Cadaverine can be overproduced by introduction of an overproduced lysine decarboxylase. The corresponding substrate, L-lysine, is synthesized in E. coli and C. glutamicum by similar pathways covering 10 enzymatic steps initiating from the tricarboxylic acid cycle intermediate oxaloacetate. The three initial steps in this pathway lead to aspartic acid semialdehyde, which is the branch point for biosynthesis of the amino acids, L-methionine, L-threonine, L-isoleucine and L-lysine (Figure 19.4). However, there were substantial differences in the enzyme systems possessed by E. coli and C. glutamicum. When it is LysC from C. glutamicum that is additionally feedback inhibited by L-threonine, it was ThrA from E. coli that is subject to feedback inhibition by L-threonine (Park and Lee, 2010). The tolerance of E. coli for cadaverine seems to be lower compared to putrescine. The biomass formed in the presence of 51 g/l Characteristics of Microbial Putrescine and Cadaverine Production Polyamine Substrate Organism Cultivation Method C [g/l] Y(P/S) [g/g] References Putrescine Glucose E. coli Fermentor (fed-batch) 5.1 nd Eppelmann et al. (2006) Putrescine Glucose E. coli Fermentor (fed-batch) 24.2 nd Qian et al. (2009) Putrescine Glucose C. glutamicum Shake flask 6 0.12 Schneider and Wendisch (2010) Putrescine Glucose C. glutamicum Fermentor (fed-batch) 19 0.16 Schneider et al. (2012) Cadaverine Lysine E. coli Fermentor (fed-batch) 69 e Nishi et al. (2007) Cadaverine Glucose E. coli Fermentor (fed-batch) 9.6 0.12 Qian et al. (2011) Cadaverine Glucose C. glutamicum Fermentor (fed-batch) 2.6 0.05 Mimitsuka et al. (2007) Cadaverine Glucose C. glutamicum Shake flask 3.4 nd Verseck et al. (2008) Cadaverine Glucose C. glutamicum Fermentor (fed-batch) 5.0 0.09 Tateno et al. (2007) Cadaverine Starch C. glutamicum Fermentor (fed-batch) 2.4 0.05 Tateno et al. (2007) Cadaverine Glucose C. glutamicum Shake flask 1.7 0.17 Kind et al. (2010b) Cadaverine Glucose C. glutamicum Shake flask 1.1 0.11 Kind et al. (2010b) Cadaverine Glucose C. glutamicum Shake flask 1.3 0.13 Kind et al. (2010a) Cadaverine Glucose C. glutamicum Fermentor (fed-batch) nd Völkert et al. (2010) Cadaverine Xylose C. glutamicum Shake flask 1.4 0.11 Buschke et al. (2011) Cadaverine Hemicellulose hydrolysate C. glutamicum Shake flask 2 nd Buschke et al. (2011) nd - Not determined Source: Schneider, J. and Wendisch, V.F. (2011); with modification. 72
REFERENCES cadaverine was reduced by 30% in comparison to the same molar concentration of putrescine (Qian et al., 2011, 2009). Corynebacterium glutamicum was tested for growth on solid medium and grew even at concentrations of up to 31 g/l cadaverine (Mimitsuka et al., 2007). Escherichia coli strains overexpressing the lysine decarboxylase gene cadA (b4131) in the wild-type genetic background led to accumulation of 0.8 g/l cadaverine by growing cells. To avoid side reactions of enzymes active with putrescine toward cadaverine, a number of genes were deleted: the spermidine synthase gene speE, the spermidine acetyltransferase gene speG, the putrescine importer gene puuP, the putrescine aminotransferase gene puuA and ygjG, which encodes the initial enzyme of the second putrescine degradation pathway and is known to be active in vitro with cadaverine. The resulting strain was able to accumulate 1.2 g/l cadaverine. Production of cadaverine was increased by 10% as a consequence of enhancing the flux of L-aspartic acid toward L-lysine by overexpression of dapA via promoter exchange. In fed-batch cultivation, this strain produced 9.6 g/l cadaverine (Qian et al., 2011). Cadaverine production in C. glutamicum was also achieved by insertional inactivation of homoserine dehydrogenase gene, hom (cg1337, Figure 19.4, B-1) with cadA from E. coli. The resultant strain secretes 2.6 g/l cadaverine in the supernatant. The expression of cadA was driven by the strong kanamycin resistance gene promoter. But the strain was auxotrophic for L-methionine, L-threonine, and L-isoleucine (Mimitsuka et al., 2007). A different approach with biosynthetic lysine decarboxylase (LdcC) from E. coli led to 30% more cadaverine production than overexpression of cadA (Kind et al., 2010b). Later the C. glutamicum DAP-3c cadaverine-producing strain’s substrate spectrum was broadened for hemicellulose utilization by introducing xylA and xylB genes from E. coli (Buschke et al., 2011). Through various studies reasonable titers and productivities were achieved for putrescine and cadaverine (Table 19.1). CONCLUSION AND PERSPECTIVES This chapter outlined the microbial production of amino acids, poly(amino acid)s and polyamines known so far. Biotechnological production of amino acids today serves a market with strong prospects of growth. In the foreground are the fermentation processes, which are now widely established in the production of proteinogenic amino acids; this can be extended to the production of other amino products like poly(amino acid)s and polyamines. The potential that will be leveraged in the future by modern methods and new findings in system biology will further stimulate and strengthen 349 microbial production of amino products. Modern methods such as directed evolution will allow development of customized, highly selective, and stable enzymes and whole cell biocatalysts, as well as efficient and ecologically sustainable production of the required products. It became a need to assess the feasibility of implementing, in addition to the established chemical processes, a biorefinery concept based on renewable raw materials. The poly(amino acid)s production does not have the luxury of background knowledge regarding the metabolic process leading to their synthesis when compared to the amino acids and polyamines. Even then, poly(amino acid)s was produced in recently good titers by using newly isolated strains and their genetically manipulated versions. However, the genetic engineering strategies were yet to attain maximal potential in polyamine and poly(amino acid)s producing strains. Acknowledgments The authors are thankful to various funding agencies such as DBT, New Delhi, DST New Delhi, and BMBF, Germany for different grants to work on microbial production of amino acids. References Aboulmagd, E., Voss, I., Oppermann-Sanio, F.B., Steinbüchel, A., 2001. Heterologous expression of cyanophycin synthetase and cyanophycin synthesis in the industrial relevant bacteria Corynebacterium glutamicum and Ralstonia eutropha and in Pseudomonas putida. Biomacromolecules 2, 1338e1342. Anastassiadis, S., 2007. L-Lysine fermentation. Recent Pat Biotechnol. 1, 11e24. Ashiuchi, M., Shimanouchi, K., Nakamura, H., Kamei, T., Soda, K., Park, C., Sung, M.H., Misono, H., 2004. Enzymatic synthesis of highmolecular mass poly-gamma-glutamate and regulation of its stereochemistry. Appl. Environ. Microbiol. 70, 4249e4255. Bajaj, I., Singhal, R., 2011. Poly (glutamic acid)dan emerging biopolymer of commercial interest. Bioresour. Technol. 102, 5551e5561. Berens, S., Kalinowski, J., Puehler, A., 2001. Corynebacterium glutamicum expressing heterologous amylase. European Patent Application EP1156113. Brickman, T.J., Armstrong, S.K., 1996. The ornithine decarboxylase gene odc is required for alcaligin siderophore biosynthesis in Bordetella spp.: putrescine is a precursor of alcaligin. J. Bacteriol. 178, 54e60. Buescher, J.M., Margaritis, A., 2007. Microbial biosynthesis of polyglutamic acid biopolymer and applications in the biopharmaceutical, biomedical and food industries. Crit. Rev. Biotechnol. 27, 1e19. Buschke, N., Schröder, H., Wittmann, C., 2011. Metabolic engineering of Corynebacterium glutamicum for production of 1,5-diaminopentane from hemicellulose. Biotechnol. J. 6, 306e317. Celano, P., Berchold, C.M., Kizer, D.L., Weeraratna, A., Nelkin, B.D., Baylin, S.B., Casero Jr, R.A., 1992. Characterisation of an endogenous RNA transcript with homology to the antisense strand of the human c-myc gene. J. Biol. Chem. 267, 15092e15096. Chapman, P., 3 July, 2012. Amino Acids Market to Hit $12.8 Billion by 2017. Available from: http://www.companiesandmarkets.com/
350 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION News/Chemicals/Amino-acids-market-to-hit-12-8- billion-by-2017/ NI3838. Cheng, C., Asada, Y., Aida, T., 1989. Production of g-polyglutamic by Bacillus licheniformis A35 under denitrifying conditions. Agric. Biol. Chem. 53, 2369e2375. Das, K., Anis, M., Azemi, B.M., Ismail, N., 1995. Fermentation and recovery of glutamic acid from palm waste hydrolysate by ionexchange resin column. J. Biotechnol. Bioeng. 48, 551e555. Davis, B., 29 Mar, 1952. Biosynthetic interrelations of lysine, diaminopimelic acid, and threonine in mutants of E. coli. Nature 169 (4300), 534e536. Dong, X.Y., Quinn, P.J., Wang, X.Y., 2011. Metabolic engineering of Escherichia coli and Corynebacterium glutamicum for the production of L-threonine. Biotechnol. Adv. 29, 11e23. Drechsel, E., 1889. Zur Kenntniss der Spaltungsprodukte des Caseı̈ns. J. Prakt. Chem. 39, 425e429. Elbahloul., Y., Krehenbrink, M., Reichelt, R., Steinbüchel, A., 2005. Physiological conditions conducive to high cyanophycin content in biomass of Acinetobacter calcoaceticus strain ADP1. Appl. Environ. Microbiol. 71, 858e866. Eppelmann, K., Nossin, P.M.M., Raeven, L.J.R.M., Kremer, S.M., Wubbolts, M.G., 2006. Biochemical Synthesis of 1,4-butanediamine. WO2006005603. Foster, J.W., 2004. Escherichia coli acid resistance: tales of an amateur acidophile. Nat. Rev. Microbiol. 2, 898e907. Georgi, T., Rittmann, D., Wendisch, V.F., 2005. Lysine and glutamate production by Corynebacterium glutamicum on glucose, fructose and sucrose: roles of malic enzyme and fructose-1,6-bisphosphatase. Metab. Eng. 7, 291e301. Ghosh, B.B., Banerjee, A.K., 1986. Production of methionine and glutamic acid from n-alkanes by Serratia marcescens var. kiliensis. Folia Microbiol. 31, 106e112. Glansdorff, N., Xu, Y., 2007. Amino acid biosynthesisdpathways, regulation and metabolic engineering. In: Wendisch, V.F. (Ed.), Microbial Arginine Biosynthesis: Pathway, Regulation and Industrial Production, Microbiology Monographs, vol. 5. Springer Berlin Heidelberg, Berlin, pp. 219e257. Gopinath, V., Meiswinkel, T.M., Wendisch, V.F., Nampoothiri, K.M., 2011. Amino acid production from rice straw and wheat bran hydrolysates by recombinant pentose-utilizing Corynebacterium glutamicum. Appl. Microbiol. Biotechnol. 92, 985e996. Gopinath, V., Anusree, M., Dhar, K.S., Nampoothiri, K.M., 2012. Corynebacterium glutamicum as a potent biocatalyst for the bioconversion of pentose sugars to value-added products. Appl. Microbiol. Biotechnol. 93, 95e106. Goto, A., Kunioka, M., 1992. Biosynthesis and hydrolysis of poly(gglutamic acid) from Bacillus subtilis IFO3335. Biosci. Biotechnol. Biochem. 56, 1031e1035. Hai, T., Oppermann-Sanio, F.B., Steinbüchel, A., 1999. Purification and characterization of cyanophycin and cyanophycin synthetase from the thermophilic Synechococcus sp. MA19. FEMS Microbiol. Lett. 181, 229e236. Hai, T., Oppermann-Sanio, F.B., Steinbüchel, A., 2002. Molecular characterization of a thermostable cyanophycin synthetase from the thermophilic cyanobacterium Synechococcus sp. strain MA19 and in vitro synthesis of cyanophycin and related polyamides. Appl. Environ. Microbiol. 68, 93e101. Hamana, K., Matsuzaki, S., 1992. Polyamines as a chemotaxonomic marker in bacterial systematics. Crit. Rev. Microbiol. 18, 261e283. Hampel, K.J., Crosson, P., Lee, J.S., 1991. Polyamines favour triplex DNA formation at neutral pH. Biochemistry 30, 4455e4459. Hezayen, F.F., Rehm, B.H., Tindall, B.J., Steinbüchel, A., 2001. Transfer of Natrialba asiatica B1T to Natrialba taiwanensis sp. nov. and description of Natrialba aegyptiaca sp. nov., a novel extremely halophilic, aerobic, non-pigmented member of the Archaea from Egypt that produces extracellular poly(glutamic acid). Int. J. Syst. Evol. Microbiol. 51, 1133e1142. Hiraki, J., Hatakeyama, M., Morita, H., Izumi, Y., 1998. Improved epolyL-lysine production of an S-(2-aminoethyl)-L-cysteine resistant mutant of Streptomyces albulus. Seibutsu Kogaku Kaishi 76, 487e493. Hong, W., Soon, H., Yung, Chil, Seung Hee, C., 1974. Kor. Jour. Microbiol. 12, 115e130. Houlahan, M.B., Mitchell, H.K., 1948. Evidence for an interrelation in the metabolism of lysine, arginine and pyrimidines in Neurospora. Proc. Natl. Acad. Sci. USA 34, 465e470. Hwang, B.J., Yeom, H.J., Kim, Y., Lee, H.S., 2002. Corynebacterium glutamicum utilizes both transsulfuration and direct sulfhydrylation pathways for methionine biosynthesis. J. Bacteriol. 184, 1277e1286. Ikeda, K., 1908. Japanese patent 4805. Ikeda, M., 2003. Amino acid production processes. Adv. Biochem. Eng. Biotechnol. 79, 1e35. Janne, J., Alhonen, L., Pietila, M., Keinanen, T.A., 2004. Genetic approaches to the cellular functions of polyamines in mammals. Eur. J. Biochem. 271, 877e894. Janne, J., Alhonen, L., Pietila, M., Keinanen, T.A., Pietila, M., Uimari, A., Pirinen, E., Hyvonen, M.T., Jarvinen, A., 2005. Animal disease models generated by genetic engineering of polyamine metabolism. J. Cell. Mol. Med. 9, 865e882. Jung, I.L., Oh, T.J., Kim, G., 2003. Abnormal growth of polyamine deficient Escherichia coli mutant is partially caused by oxidative stress-induced damage. Arch. Biochem. Biophys. 418, 125e132. Kase, H., Nakayama, K., 1974. Production of O-acetyl-L-homoserine by methionine analog-resistant mutants and regulation of homoserine-O-transacetylase in Corynebacterium glutamicum. Agr. Biol. Chem. 38, 2021e2030. Khamduang, M., Packdibamrung, K., Chutmanop, J., Chisti, Y., Srinophakun, P., 2009. Production of L-phenylalanine from glycerol by a recombinant Escherichia coli. J. Ind. Microbiol. Biotechnol. 36, 1267e1274. Kind, S., Jeong, W.K., Schröder, H., Zelder, O., Wittmann, C., 2010a. Identification and elimination of the competing N-acetyldiaminopentane pathway for improved production of diaminopentane by Corynebacterium glutamicum. Appl. Environ. Microbiol. 76, 5175e5180. Kind, S., Jeong, W.K., Schröder, H., Wittmann, C., 2010. Systemswide metabolic pathway engineering in Corynebacterium glutamicum for bio-based production of diaminopentane. Metab. Eng. 12, 341e351. Kita, D.A., Jackson Heights, N.Y., 1957. Production of glutamic acid by cephalasporum. United States Patent 2789939. Kito, M., Onji, Y., Yoshida, T., Nagasawa, T., 2002. Occurrence of epsilon-poly-L-lysine-degrading enzyme in epsilon-poly-L-lysinetolerant Sphingobacterium multivorum OJ10: purification and characterization. FEMS Microbiol. Lett. 207, 147e151. Kocianova, S., Vuong, C., Yao, Y., Voyich, J.M., Fischer, E.R., DeLeo, F.R., Otto, M., 2005. Key role of poly-gamma-DL-glutamic acid in immune evasion and virulence of Staphylococcus epidermidis. J. Clin. Invest. 115, 688e694. Komatsubara, S., Kisumi, M., Murata, K., Chibata, I., 1978. Threonine production by regulatory mutants of Serratia marcescens. Appl. Environ. Microbiol. 35, 834e840. Kroschwitz, J.I., Seidel, A., 2004. KirkeOthmer Encyclopedia of Chemical Technology, fifth ed. Wiley-Interscience, Hoboken. Kubota, H., Nambu, Y., Endo, T., 1995. Convenient esterification of poly(-glutamic acid) produced by microorganism with alkyl halides and their thermal properties. J. Polym. Sci. A: Polym. Chem. 33, 85e88. Kunioka, M., 1995. Biosynthesis of poly(g-glutamic acid) from L-glutamine, citric acid, and ammonium sulfate in Bacillus subtilis IFO3335. Appl. Microbiol. Biotechnol. 44, 501e506.
REFERENCES Kunioka, M., 2004. Biodegradable water absorbent synthesized from bacterial poly(amino acid)s. Macromol. Biosci. 4, 324e329. Kunioka, M., Goto, A., 1994. Biosynthesis of poly(g-glutamic acid) from L-glutamic acid, citric acid, and ammonium sulfate in Bacillus subtilis IFO3335. Appl. Microbiol. Biotechnol. 40, 867e872. Leonard, C.G., Houseright, R.D., Thorne, C.B., 1958a. Effects of some metallic ions on glutamyl polypeptides, by Bacillus subtilis. J. Bacteriol. 76, 499e503. Leonard, C.G., Houseright, R.D., Thorne, C.B., 1958b. Effects of some metallic ions on the optical specificity of glutamyl synthase and glutamyl transferase of Bacillus licheniformis. Biochem. Biophys. Acta 62, 432e434. Lin, K., Elbahloul, Y., Steinbüchel, A., 2012. Physiological conditions conducive to high cell density and high cyanophycin content in Ralstonia eutropha strain H16 possessing a KDPG aldolase genedependent addiction system. Appl. Microbiol. Biotechnol. 93, 1885e1894. Matsusaki, M., Serizawa, T., Kishida, A., Akashi, M., 2005. Novel functional biodegradable polymer. III. The construction of poly(gamma-glutamic acid)-sulfonate hydrogel with fibroblast growth factor-2 activity. J. Biomed. Mater. Res. A 73, 485e491. McCutchan, et. al., 1962. Glutamic acid fermentation. US patent 3061521. Meiswinkel, M.T., Gopinath, V., Lindner, S.N., Nampoothiri, K.M., Wendisch, V.F., 2012. Accelerated pentose utilization by Corynebacterium glutamicum for accelerated production of lysine, glutamate, ornithine and putrescine. Microbial. Biotechnol.. http:// dx.doi.org/10.1111/1751-7915.12001. Mimitsuka, T., Sawai, H., Hatsu, M., Yamada, K., 2007. Metabolic engineering of Corynebacterium glutamicum for cadaverine fermentation. Biosci. Biotechnol. Biochem. 71, 2130e2135. Miyamoto, S., Kashiwagi, K., Ito, S., Watanabe, S., Igarashi, K., 1993. Estimation of polyamine distribution and polyamine stimulation of protein synthesis in Escherichia coli. Arch. Biochem. Biophys. 300, 63e68. Morgan, J.E., Blankenship, J.W., Matthews, H.R., 1987. Polyamines and acetylpolyamines increase the stability and alter the conformation of nucleosome core particles. Biochemistry 26, 3643e3649. Morinaga, Y., Tani, Y., Yamada, H., 1982. L-Methionine production by ethionine resistant mutant of facultative methylotroph Pseudomonas FM 518. Agric. Biol. Chem. 46, 473e480. Mueller, J.H., 1923. A new sulfur-containing amino-acid isolated from the hydrolytic products of protein. J. Biol. Chem. 56, 157e169. Nampoothiri, K.M., Pandey, A., 1996. Solid state fermentation for Lglutamic acid production using Brevibacteriuı́m sp. Biotech. Lett. 18, 199e204. Nampoothiri, K.M., Pandey, A., 1999. Fermentation and recovery of Lglutamic acid from cassava starch hydrolysate by ion-exchange resin column. Rev. Microbiol. 30, 258e264 (online). Nampoothiri, K.M., Hoischen, C., Bathe, B., Mockel, B., Pfefferle, W., Krumbach, K., Sahm, H., Eggeling, L., 2002. Expression of genes of lipid synthesis and altered lipid composition modulate L-glutamate efflux of C. glutamicum. Appl. Microbiol. Biotechnol. 58, 89e96. Neumann, K., Stephan, D.P., Ziegler, K., Hühns, M., Broer, I., Lockau, W., Pistorius, E.K., 2005. Production of cyanophycin, a suitable source for the biodegradable polymerpolyaspartate, in transgenic plants. Plant. Biotechnol. J. 3, 249e258. Nishi, K., Endo, S., Mori, Y., Totsuka, K., Hirao, Y., 2007. Method for producing cadaverine dicarboxylate. US 7189543 B2. Obst, M., Steinbüchel, A., 2004. Microbial degradation of poly(amino acid)s. Biomacromolecules 5 (4), 1166e1176. Ogawa, Y., Hosoyama, H., Hamano, M., Motai, H., 1991. Purification and properties of gamma-glutamyltranspeptidase from Bacillus subtilis (natto). Agric. Biol. Chem. 55, 2971e2977. 351 Oppermann-Sanio, F.B., Steinbüchel, A., 2002. Occurrence, functions and biosynthesis of polyamides in microorganisms and biotechnological production. Naturwissenschaften 89, 11e22. Oshima, T., 2007. Unique polyamines produced by an extreme thermophile, Thermus thermophilus. Amino Acids 33, 367e372. Park, J.H., Lee, S.Y., 2010. Metabolic pathways and fermentative production of L-aspartate family amino acids. Biotechnol. J. 5, 560e577. Patel, C.N., Wortham, B.W., Lines, J.L., Fetherston, J.D., Perry, R.D., Oliveira, M.A., 2006. Polyamines are essential for the formation of plague biofilm. J. Bacteriol. 188, 2355e2363. Peters-Wendisch, P.G., Kreutzer, C., Kalinowski, J., Pátek, M., Sahm, H., Eikmanns, B.J., 1998. Pyruvate carboxylase from Corynebacterium glutamicum: characterization, expression and inactivation of the pyc gene. Microbiology 144, 915e927. Peters-Wendisch, P.G., Schiel, B., Wendisch, V.F., Katsoulidis, E., Möckel, B., Sahm, H., Eikmanns, B.J., 2001. Pyruvate carboxylase is a major bottleneck for glutamate and lysine production by Corynebacterium glutamicum. J. Mol. Microbiol. Biotechnol. 3, 295e300. Pham, C.B., Galvez, F.C.F., Padolina, W.G., 1992. Methionine production by batch fermentation from various carbohydrates. ASEAN Food J. 7, 34e37. Qian, Z., Xia, X., Lee, S.Y., 2009. Metabolic engineering of Escherichia coli for the production of putrescine: a four carbon diamine. Biotechnol. Bioeng. 104, 651e662. Qian, Z., Xia, X., Lee, S.Y., 2011. Metabolic engineering of Escherichia coli for the production of cadaverine: a five carbon diamine. Biotechnol. Bioeng. 108, 93e103. Riedel, C., Rittmann, D., Dangel, P., Möckel, B., Petersen, S., Sahm, H., Eikmanns, B.J., 2001. Characterization of the phosphoenolpyruvate carboxykinase gene from Corynebacterium glutamicum and significance of the enzyme for growth and amino acid production. J. Mol. Microbiol. Biotechnol. 3, 573e583. Ritthausen, H., 1866. Über die Glutaminsaure [About glutamic acid]. J. Prakt. Chem. 99, 454e462. Rittman, D., Lindner, S.N., Wendisch, V.F., 2008. Engineering of a glycerol utilization pathway for amino acid production by Corynebacterium glutamicum. Appl. Environ. Microbiol. 74, 6216e6222. Rose, W.C., 1931. Feeding Experiments with Mixtures of Highly Purified Amino Acids. I. The Inadequacy of Diets Containing Nineteen Amino Acids. J. Biol. Chem. 94, 155e165. Sallam, A., Steinbüchel, A., 2009. Cyanophycin-degrading bacteria in digestive tracts of mammals, birds and fish and consequences for possible applications of cyanophycin and its dipeptides in nutrition and therapy. J. Appl. Microbiol. 107, 474e484. Sallam, A., Steinbüchel, A., 2010. Dipeptides in nutrition and therapy: cyanophycin-derived dipeptides as natural alternatives and their biotechnological production. Appl. Microbiol. Biotechnol. 87, 815e828. Sasaki, M., Jojima, T., Inui, M., Yukawa, H., 2008. Simultaneous utilization of D-cellobiose, D-glucose, and D-xylose by recombinant Corynebacterium glutamicum under oxygen-deprived conditions. Appl. Microbiol. Biotechnol. 81, 691e699. Savrasova, E.A., Sycheva, E.V., Michurina, T.A., Kozlov, Y.I., 2004. Process for producing l-amino acids by fermentation of a mixture of glucose and pentoses. US Patent 20040229321A1. Schneider, J., Wendisch, V.F., 2010. Putrescine production by engineered Corynebacterium glutamicum. Appl. Microbiol. Biotechnol. 88, 859e868. Schneider, J., Wendisch, V.F., 2011. Biotechnological production of polyamines by bacteria: recent achievements and future perspectives. Appl. Microbiol. Biotechnol. 91, 17e30. Schneider, J., Niermann, K., Volker, V.F., 2011. Production of the amino acids L-glutamate, L-lysine, L-ornithine and L-arginine from
352 19. AMINO-BASED PRODUCTS FROM BIOMASS AND MICROBIAL AMINO ACID PRODUCTION arabinose by recombinant Corynebacterium glutamicum. J. Biotechnol. 154, 191e198. Schneider, J., Eberhardt, D., Wendisch, V.F., 2012. Improving putrescine production by Corynebacterium glutamicum by fine-tuning ornithine transcarbamoylase activity using a plasmid addiction system. Appl. Microbiol. Biotechnol. 95, 169e178. Schwamborn, M., 1998. Chemical synthesis of polyaspartates: a biodegradable alternative to currently used polycarboxylate homoand copolymers. Polym. Degrad. Stabil. 59, 39e45. Shi, F., Xu, Z.N., Cen, P.L., 2006. Efficient production of poly-gglutamic acid by Bacillus subtilis ZJU-7. Appl. Biochem. Biotechnol. 133, 271e282. Shi, F., Xu, Z.N., Cen, P.L., 2007. Microbial production of natural poly amino acid. Sci. China Ser. B: Chem. 50, 291e303. Shih, I.L., Van, Y.T., Sau, Y.Y., 2003. Antifreeze activities of poly(gamma-glutamic acid) produced by Bacillus licheniformis. Biotechnol. Lett. 25, 1709e1712. Shima, S., Sakai, H., 1977. Polylysine produced by streptomyces. Agric. Biol. Chem. 41, 1807e1809. Shima, S., Sakai, H., 1981. Poly-L-lysine produced by Streptomyces. Part II. Taxonomy and fermentation studies. Agric. Biol. Chem. 45, 2497e2502. Shimizu, H., Hirasawa, T., 2006. Production of glutamate and glutamate-related amino acids: molecular mechanism analysis and metabolic engineering. In: Wendisch, V.F. (Ed.), Microbiology Monographs, Amino Acid BiosynthesisdPathways, Regulation and Metabolic Engineering, vol. 5. Springer Berlin Heidelberg, Berlin, pp. 1e38. Steinle, A., Oppermann-Sanio, F.B., Reichelt, R., Steinbüchel, A., 2008. Synthesis and accumulation of cyanophycin in transgenic strains of Saccharomyces cerevisiae. Appl. Environ. Microbiol. 74, 3410e3418. Sturgill, G., Rather, P.N., 2004. Evidence that putrescine acts as an extracellular signal required for swarming in Proteus mirabilis. Mol. Microbiol. 51, 437e446. Tabor, C.W., Tabor, H., 1985. Polyamines in microorganisms. Microbiol. Rev. 49, 81e99. Takatsuka, Y., Kamio, Y., 2004. Molecular dissection of the Selenomonas ruminantium cell envelope and lysine decarboxylase involved in the biosynthesis of a polyamine covalently linked to the cell wall peptidoglycan layer. Biosci. Biotechnol. Biochem. 68, 1e19. Tateno., T., Fukuda, H., Kondo, A., 2007. Production of L-lysine from starch by Corynebacterium glutamicum displaying alpha-amylase on its cell surface. Appl. Microbiol. Biotechnol. 74, 1213e1220. Tavakkoli, M., Hamidi-Esfahani, Z., Azizi, M.H., 2009. Optimization of Corynebacterium bacterium glutamic acid production by response surface methodology. Food Bioprocess Technol. 5, 92e99. Thomsen, M.H., Bech, D., Kiel, P., 2004. Manufacturing of stabilised brown juice for L-lysine from university lab scale over pilot scale to industrial production. Chem. Biochem. Eng. Q. 18, 37e46. Troy, F.A., 1973. Chemistry and biosynthesis of the poly(g-D-glutamyl) capsule in Bacillus licheniformis. I. Properties of the membranemediated biosynthetic reaction. J. Biol. Chem. 248, 305e315. Tsuchidate, T., Tateno, T., Okai, N., Tsutomu, T., Ogino, C., Kondo, A., 2011. Glutamate production from b-glucan using endoglucanasesecreting Corynebacterium glutamicum. Appl. Microbiol. Biotechnol. 90, 895e901. Utagawa, T., 2004. Production of arginine by fermentation. J. Nutr. 134, 2854Se2857S. van Leeuwenhoek, A., 1678. Observationes D. Anthonii Leeuwenhoek, de Natis e semine genitali Animalculis. Phil. Trans. R. Soc. Lond. 12, 1040e1043. Verseck, S., Häger, H., Karau, A., Eggeling, L., Sahm, H., 2008. A process for the fermentative production of cadaverine. DE 102007005072 A1. Völkert, M., Zelder, O., Ernst, B., Jeong, W.K., 2010. Method for fermentatively producing 1,5-diaminopentane. US 20100292429 A1. Vrljic, M., Sahm, H., Eggeling, L., 1996. A new type of transporter with a new type of cellular function: L-lysine export from Corynebacterium glutamicum. Mol. Microbiol. 22, 815e826. Wallace, H.M., Fraser, A.V., Hughes, A., 2003. A perspective of polyamine metabolism. Biochem. J. 376, 1e14. Walters, E.D., 1991. The rational discovery of sweeteners. In: Walters, D.E., Orthoefer, F.T., DuBois, G.E. (Eds.), Sweeteners. Discovery, Molecular Design, and Chemoreception. American Chemical Society, Washington DC, pp. 1e11. Wittmann, C., Becker, J., 2007. The L-lysine story: from metabolic pathways to industrial production. In: Wendisch, V.F. (Ed.), Microbiology Monographs, Amino Acid BiosynthesisdPathways, Regulation and Metabolic Engineering, vol. 5. Springer Berlin Heidelberg, Berlin, pp. 39e70. Xu, K., Strauch, M.A., 1996. Identification, sequence, and expression of the gene encoding gamma-glutamyltranspeptidase in Bacillus subtilis. J. Bacteriol. 178, 4319e4322. Yang, G., Chen, J., Qu, Y.B., Lun, S.Y., 2002. Effects of metal ions on gamma-poly (glutamic acid) synthesis by Bacillus licheniformis. Sheng Wu Gong Cheng Xue Bao 17, 706e709. Yoshida, T., Nagasawa, T., 2003. Epsilon-poly-L-lysine: microbial production, biodegradation and application potential. Appl. Microbiol. Biotechnol. 62, 21e26. Zahoor, A., Linder, S.N., Wendisch, V.F., 2012. Metabolic engineering of Corynebacterium glutamicum aimed at alternative carbon sources and new products. Computational Struct. Biotechnol. J. 3 (4). e201210004. http://dx.doi.org/10.5936/csbj.201210004.
C H A P T E R 20 Production of Phytochemicals, Dyes and Pigments as Coproducts in Bioenergy Processes Hanshu Ding*, Feng Xu Department of Protein Chemistry, Novozymes Inc., Davis, California, USA *Corresponding author email: hdin@novozymes.com; fxu@novozymes.com O U T L I N E Industrial Phytochemicals Overview Colorants (Pigment, Dye, and Ink) Dietary, Nutraceutical, Food or Feed Additives Bioactive or Pharmaceutical Phytochemicals Phytochemicals for Personal Care or Other Uses 353 353 355 356 356 358 Production of Industrial Phytochemicals Extraction and Isolation from Specific Plants Coproduction from Processing (Biorefinery) of Staple Crops Production from Cultured Plant Cells Production from Microbial Fermentation Production from Algae via Aquaculture 358 358 Coproduction of Phytochemicals in Bioenergy Processes 358 360 360 361 References 361 361 362 362 363 363 363 361 INDUSTRIAL PHYTOCHEMICALS Overview Phytochemicals may be defined as chemicals derived or derivable from plants. Phytochemical sources may include not only unprocessed trees, crops, grains, fruits, nuts, vegetables and legumes, but also processed plant-derived materials such as starch, sugars and oil. Phytochemicals of commercial interest have demonstrated or suspected utilities for dietary, bioactive, therapeutic and industrial technical uses. Based on chemical structure, major groups of phytochemicals include (Figure 20.1) the following: Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00020-6 Coproduction from Starch- or Sugar-Based Bioenergy Processes Coproduction from Plant Oil-Based Bioenergy Processes Coproduction from Lignocellulose (Biomass)-Based Bioenergy Processes Coproduction from Bio-Oil, Syn-Gas, or Algal Bioenergy Processes Colocation of Fermentative Phytochemicals Production with Bioenergy Processes Utilization of Phytochemical Production By-Products for Bioenergy 1. Carotenoids (carotenes or xanthophylls, e.g. a- or b-carotene, b-cryptoxanthin, lycopene, lutein, and zeaxanthin) and homologs (e.g. crocin) 2. Flavonoids (anthocyanins, flavanols, flavanones, flavonols, flavones, and isoflavones), and condensed tannin and xanthones 3. Other phenolic/quinonics (e.g. tocopherols, curcumin, resveratrol, carminic acids, alizarin, purpurin, lignans (dimeric phenyl propanoid), tannic acid, thymol, and capsaicin) 4. Alkaloids (e.g. caffeine, nicotine, quinine, vinblastine, and opiates) and other N-contained compounds (e.g. chlorophylls, flavins, betalain, indole-3-carbinol, galanthamine, and indigo) 353 Copyright Ó 2014 Elsevier B.V. All rights reserved.
354 20. PRODUCTION OF PHYTOCHEMICALS, DYES AND PIGMENTS AS COPRODUCTS IN BIOENERGY PROCESSES O H3C H3C H 3C N + CH3 N O N Carotenoids: β-Carotene N CH3 H3C O CH3 CH3 Flavonoids: Anthocyanin CH3 N N Alkaloids: Caffeine Alkaloids: Nicotine H3C CH3 H3C CH3 H 3C H3C H3C CH3 O CH3 H3C H3C CH3 H3C H H OH H3C CH3 H CH3 H CH3 H3C HO Phenolic/quinonics: Tocopherols Phytosterols: Sitosterol O CH3 H3C SH HO O HN H O O O H 2N HN Protease inhibitors: Papain-inhibitor E64 Lipids: Lecithin O OH O O O HO H O HO O H O OH O O OH O O HO O H HO O H O OH O O HO H O OH O O HO O H HO O H O OH O O HO H O OH O O HO HO O O H H O OH O O OH HO O H HO O HO O O O H HO O OH CH3O OCH3 O OH O HO O OH O O O OH HO O OCH3 HO HO OCH3 O HO O OH O H Carbohydrates: Cellulose O CH3O OCH3 O OH O OCH3 OH O HO O OCH3 O O H HO O OH OH O OH OCH3 CH3O OH HO HO O H OCH3 OH HO OH O OH OH OH O OH OH HO HO H O N H H2N H3C S-compounds: γ-glutamylcysteines NH O H3C CH3 O O O N+ H3C O N H O P O CH3 O HO O O HO CH3 OCH3 OH OH OCH3 Carbohydrates: Lignin FIGURE 20.1 Major groups of phytochemicals. Source: Drawings from ChemSpider and Sigmaaldrich.com are used in this and Figures 20.2e20.4. (For color version of this figure, the reader is referred to the online version of this book.) 5. S-contained compounds (e.g. g-glutamylcysteines, (allyl)cysteine sulfoxides, and isothiocyanates) 6. Phytosterols (e.g. sitosterol, stigmasterol, campesterol or 4-desmethyl sterols), saponin, digoxin, and other terpenoids (e.g. artemisinin, paclitaxel, and camphor) 7. Polymeric carbohydrates (e.g. cellulose, hemicellulose, b-glucan, pectin, gum, inulin, and resistance starch), oligosaccharides (e.g. oligofructose), and lignin 8. Lipids and volatiles such as lecithin, essential oils, and menthol 9. Proteases such as bromelain and papain, as well as protease inhibitors It is reported that about 8000 phenolics (including w4000 flavonoids), w20,000 terpenoids, w10,000 alkaloids, w700 carotenoids, and w250 phytosterols are known, and many have been shown with various
355 INDUSTRIAL PHYTOCHEMICALS Colorants (Pigment, Dye, and Ink) functions (Watkins and Chaudhry, 2013). Comprehensive studies have been carried out on phytochemicals from whole grains, fruits and vegetables (Liu, 2007; Piironen et al., 2000). Phytochemicals may have different physical, chemical and biological properties, thus suitable for different industrial usages as colorants (pigment, dye, and ink), dietary food/feed additives or nutraceuticals, bioactive/ pharmaceutical ingredients, personal care (cosmetic, perfume) agents, or other useful materials. For instance, carotenoids, polyphenols, flavonoids and tocopherols may be used as antioxidant or antiinflammatory agents; alkaloids may be used as analgesic, antispasmodic or mental disorder-relieving agents; and carotenoids may be used as coloring agents. Phytochemicals are of great interest for industrial, technical, household, health care or other uses, due to their renewability, performance, safety, environmentfriendliness, and diversity in structure and activity. The use of phytochemicals started at the dawn of humanity, has contributed to the civilization, and is reemerging along with the advancement of bioenergy, biobased chemicals, and biorefinery. Many phytochemicals are chromophoric, reflecting lights that cover the visible wavelength range. Ubiquitous phytocolorants include chlorophyll (green) and carotenoids (yellow-red) from leaves and stems of plants, while more specific colorants may exist in flowers, fruits or other parts of plants (Figure 20.2). Colorants could also be produced from algae, bacteria, or fungi including the saprophytes (Gupta et al., 2011; Rymbai et al., 2011; Matthews and Wurtzel, 2007; Mortensen, 2006; Dufossé, 2006; Mapari et al., 2005; Adrio and Demain, 2003; Sengupta, 2003). Colorants are used mostly as dyestuff, food/feed additives or cosmetic agents. Traditional plant-extracted/ derived dyestuffs include saffron from saffron crocus plant, madder (red) from madder plants (Rubia), and indigo from Indigofera plants (at present chemical synthesis from fossil feedstocks provides most indigo dyes, although microbial route has been explored). Commonly used food or feed colorants derived from plants include extracts or isolates from specifically grown plants, such as bixin and norbixin (annatto), betalains (including betanin), curcumin (turmeric), crocin CH2 CH3 - N H3C Chlorophyll CH3 Mg2+ N N Betanin O O H3C CH H3C O CH3 H3C O O O OH O H 3C H N O HO N H3C H3C OH N+ HO OH HO O OH H3C O OH O Curcumin HO Purpurin Indigo O OH O N H O O H 3C OH O O O H N OH OH OH H 3C H CH3 CH3 H3C CH3 O CH3 CH3 H3C HO CH3 CH3 CH3 Canthaxanthin H3C CH3 H3C CH3 OH O OH H3C H3C CH3 H3C Astaxanthin HO CH3 OH HO H3C CH3 CH2 Crocin OH H3C CH3 OH O HO CH3 H3C H 3C H3C H 3C OH O OH O CH3 Lutein OH OH O O O O O H3C H3C CH3 O CH3 HO O O OH OH OH FIGURE 20.2 Representative phytocolorants. (For color version of this figure, the reader is referred to the online version of this book.)
356 20. PRODUCTION OF PHYTOCHEMICALS, DYES AND PIGMENTS AS COPRODUCTS IN BIOENERGY PROCESSES FIGURE 20.3 Plant-derived dietary, food or feed additives, and nutraceuticals. (For color version of this figure, the reader is referred to the online version of this book.) (saffron), and carotenoids (including b-carotene, lutein, canthaxanthin and astaxanthin). The colorants also include extracts or isolates from agricultural residues, such as anthocyanins and carotenoids. The colorants may be produced microbially, as exemplified by the carotenoids such as astaxanthin, b-carotene, lutein, and riboflavin (Chattopadhyay et al., 2008). Some plant-derived colorants, such as lutein and b-carotene, are used as cosmetic agents. In addition to plants, algae also produce colorants of industrial interest. For instance, phycobiliproteins have uses as natural dyes, cosmetic agents or food colorants (in addition to health applications) (Spolaore et al., 2006). Phytocolorants may also be used for thermoplastic (van den Oever et al., 2004). Dietary, Nutraceutical, Food or Feed Additives A wide range of phytochemicals have long been used for dietary or neutraceutical purposes (Rao, 2012; Wang and Weller, 2006). Several vitamin homologs or precursors, such as b-carotene (for vitamin A), tocopherol (for vitamin E), and ascorbic acid (vitamin C), are widely produced from plants. Numerous nutraceutical phytochemicals, such as anthocyanin and flavonoids, are known for antioxidant or other bioactivities. Commonly used food or feed additives include lutein, canthaxanthin and b-carotene as dietary or coloring agent, astaxanthin for aquaculture such as salmon farming, essential oils, menthol, camphor, caffeine, tannin, capsaicin, wood flavor or liquid smoke (water-diluted bio-oil; Venderbosch and Prins, 2010; Di Blasi et al., 2010) for flavor or aroma, anthocyanins as antimicrobial agents (Chattopadhyay et al., 2008), papain and bromelain for meat processing, lecithin for emulsification, and dietary fibers. Figure 20.3 shows representative dietary phytochemicals. Bioactive or Pharmaceutical Phytochemicals Phytochemicals with a wide diversity in structure and bioactivity have long been sources for pharmaceutical agents (“phytomedicines”, Pandey et al., 2011). One important bioactive group is carotenoids.Astaxanthinhas uses as anticarcinogens, antioxidant, antiinflammation (da Fonseca et al., 2011), cholesterol effector, pain reliever (Skjanes et al., 2012) or immune system booster (Abad and Turon, 2012). Canthaxanthin also suits for antioxidant or antiinflammation uses (Skjanes et al., 2012). b-Carotene may serve to prevent erythema or arthritis (Skjanes et al., 2012). Another major bioactive phytochemical group is isoflavones, especially those from soybean. For instance, daidzein, genistein and glycitein are antioxidative or estrogenic, potentially beneficial in cancer, heart disease or obesity prevention. Phytosterols or triterpenes are phytochemicals beneficial for cholesterol, cancer or immune system-related considerations. Saponins (saponenols glycosides) are also thought to be beneficial for certain cancer, heart, liver illness treatments (Wu and Kang, 2011; Guclu-Ustundag and Mazza, 2007; Zhao and Moghadasian, 2010). Organic sulfur compounds, such as allylsulfides, also have
INDUSTRIAL PHYTOCHEMICALS anticancer potential (Cerella et al., 2011). Polyaminehydroxycinnamic amide conjugates are antioxidative and antimelanogenic (Choi et al., 2007). Lignans are also antioxidants; so are phytic and cinnamic acid ester glycosides (Wu and Kang, 2011; Guclu-Ustundag and Mazza, 2007). Menthol and essential oils are used as topical analgesic or antiitching agents, decongestants or oral hygiene ingredients. Capsaicin also has topical uses for relieving pain, itch or inflammation. Highly effective, pharmacologically well-studied medicines originated from plants include quinine, ephedrine, artemisinin, paclitaxel (Taxol) and vinblastine, galanthamine and digoxin, as well as opiates like morphine and cocaine. Plant-derived precursors for medicines include 10-deacetylbaccatin (for paclitaxel), ()-shikimic FIGURE 20.4 this book.) 357 acid (for oseltamivir phosphate or Tamiflu), diosgenin (for various steroid hormones), salicylic acid or salicin (for acetylsalicylic acid or aspirin) (Pandey et al., 2011). There are also phytochemicals that can be used for agriculture or forestry protection (‘agrochemicals’, Huter, 2011; Dayan et al., 2009). For example,leptospermonefromCallistemoncitrinusplant is used as herbicide (Salim et al., 2008); lemongrass oil as pesticide or herbicide; essential oils (e.g.D-limonene), pyrethrum, nicotine and rotenone as insecticide; thymol and pyrolyzed tobacco bio-oil as biocide (O’Brien et al., 2009; Dayan et al., 2009); and corn gluten meal and essential oils for weed control. The structures of many bioactive phytochemicals are shown in Figure 20.4. Bioactive or pharmaceutical phytochemicals. (For color version of this figure, the reader is referred to the online version of
358 20. PRODUCTION OF PHYTOCHEMICALS, DYES AND PIGMENTS AS COPRODUCTS IN BIOENERGY PROCESSES Phytochemicals for Personal Care or Other Uses Various phytochemicals are used for personal care because of their performance and renewability. Betaine (trimethylglycine ammonium salt) has significant potential for hair care (Kripp, 2006), lutein and b-carotene as colorant, and menthol and citronella oil for insect repelling. There are other industrial uses of phytochemicals. Lecithin is useful for antifoaming, dispersion, stabilization, or wetting. Tannin is used for leather processing (tannery), wood products (e.g. particle board) adhesion (Frihart, 2010), or anticorrosion. Lignin may be used for making resins. PRODUCTION OF INDUSTRIAL PHYTOCHEMICALS Extraction and Isolation from Specific Plants Conventional productions involve various mostly physical but sometimes chemical methods to isolate and enrich phytochemicals from selected wild or purposely farmed plants. Representative methods consist of solid-liquor extraction (including steam distillation), liquideliquid extraction, or membrane separation, whose choices are based on effectiveness (low cost) and efficiency (recovery, especially for labile or low-abundance phytochemicals). In principle, phytochemical productions may involve mechanical grinding of feedstocks, single or multiple steps of extraction, and enrichment or purification of final products. The extraction (leaching) parts may be simple binary systems or assisted by enhanced energy inputs (ultrasound, microwave, high pressure, sub- or supercritical condition) (Huang and Ramaswamy, 2012). Proper selection of solvent, adsorbent, and other conditions is critical. Chemical transformation is also applied to convert phytochemical precursors to final products, as exemplified by sulfuric acid treatment of madder to yield alizarin or purpurin. Plant-derived colorants are mostly produced by the methods listed above: betalains (including betanin) extracted from red beet (Beta vulgaris); bixin/norbixin (annatto) from the tree Bixa orellana (Chattopadhyay et al., 2008); gossypol from cotton seed; lutein from marigold (Tagetes erecta); capsanthin/capsorubin from paprika; capsorubin from Capsicum annuum; crocin from saffron (Crocus sativus) flower; anthocyanins from grape skin, apple or cranberry; acylated anthocyanins from black carrot; curcumin (turmeric) from Curcuma longa; carminic acid from Dactylopius coccus; alizarin or purpurin from madder (Rubia) plants; chlorophyll from spinach; and indigo from Indigofera or Isatistinctoria (woad) plants. For plant-derived S-compounds, glutamylcysteine or (allyl)cysteine sulfoxide are prepared from Allium species (including garlic); betaine from sugar beet (Kripp, 2006); tannin from tea, quebracho, chestnut or barks; caffeine from coffee and tea plant; nicotine from tobacco; camphor from camphor laurel; bromelain from pineapple; papain from papaya; essential oils from a variety of fruits, seeds, leaves, woods, barks and roots; and menthol from mint. Plant-derived drugs or precursors have been prepared by combinations of the methods listed above: quinine from cinchona tree, artemisinin from sweet wormwood (Artemisia annua), paclitaxel from Pacific yew (Taxus brevifolia)’s endophytic fungi, 10-deacetylbaccatin from a few yews, ()-shikimic acid from shikimi tree, diosgenin from Dioscorea plants, cytisine from Cytisus laburnum, vinblastine from Madagascar periwinkle (Catharanthus roseus), salicylic acid from willow bark, salicin from meadow sweet (Filipendula ulmaria), galanthamine from Caucasian snowdrop (Galanthus caucasicus), digoxin from foxglove (Digitalis lanata), and ephedrine from Ephedra sinica (Simard et al., 2012; Braz-Filho, 1999). Plant-derived phytochemicals active as plant protection and other bioactive agents are also produced from specific plants. For instance, pyrethrum is prepared from Chrysanthemum cinerariifolium and Chrysanthemum coccineum, rotenone from jicama vine, thymol from thyme (O’Brien et al., 2009; Dayan et al., 2009), and flavonoid glycosides or polymethoxylated flavones from citrus peels (juice-extracted residues) (Manthey, 2012). Coproduction from Processing (Biorefinery) of Staple Crops In addition to specific plants, staple crops also provide phytochemicals at large scale, as coproducts of well-developed, comprehensive food production processes, which may be considered as the first generation of biorefineries. Soybean processing involves multitier steps yielding multiple streams. The major soybean products include oil, feedstuff, and fermented soy food. Minor products include full-fat soy flours, soy concentrate, soy protein isolates, and lecithin. Phytochemicals that can be produced as coproducts from soybean processing include carotenoids, isoflavones and saponin; protease inhibitors from protein fractions; as well as lecithin, phytosterols, and tocopherols from oil. Soybean processing in general consists of preparatory steps (cleaning, drying, mechanic disruption or grinding, or conditioning) and oil-extraction steps (mechanical pressing or solvent extraction, refining, bleaching, and hydrogenation). Coproducts are prepared by extractive distillation, adsorption, membrane filtration, and super- or subcritical fluid extraction (Kannan et al., 2012; Zijlstra et al., 2012) (Figure 20.5).
359 PRODUCTION OF INDUSTRIAL PHYTOCHEMICALS Soybean Crushing, grinding Extraction Wet meal Miscella Toaster Drying Desolventizing Soybean meal Crude soybean oil Carotenoids Isoflavones Saponin Protease inhibitor Refining Soy lecithin Lecithin Phytosterols Tocopherols Refined soybean oil FIGURE 20.5 Schematic soybean processing, with potential phytochemical coproduction. One of the two major corn processings is wet milling. Wet milling yields major products ranging from starch, starch-fermented ethanol (first-generation bioethanol), Wet or Dried Distillers Grains (residues from ethanol fermentation) or Dried Distillers Grains and Solubles or Stillage (WDDG or DDG, DDGS, which are widely used as feed), and steep liquor. Phytochemicals that might be generated as coproducts include flavonoids, phytosterols, carotenoids, polyamine-hydroxycinnamic acid amide conjugates (Rausch, 2012; Moreau et al., 2009) from steep liquor, steeped corn, oil-extracted residues, stillage, or unfermented residues, although their production has not been widely integrated in current corn wet milling factories. Corn wet milling process include mechanic disruption, liquid extraction (steeping, acidic, basic or SO2 impregnation), screening, oil pressing, evaporation, centrifugation, fermentation and distillation. Corn dry milling is another major corn processing, mainly geared for bioethanol production. The process does not have steeping and germ-processing (oil extraction) as the wet milling does, resulting in separation and enrichment of most corn phytochemicals in unfermented residues and stillage (Rausch, 2012; Rausch and Belyea, 2006) (Figure 20.6). Wet milling Dry milling Corn Corn Steeping Corn steep liquor First milling germ separation Germ for oil and to CGF Second milling fiber separation Fiber to corn gluten feed (CGF) Starch-protein separation Grinding Water Cooking Saccharification Fermentation Corn gluten meal (CGM) Enzymes Yeast CO2 Downstream processing Ethanol DDGs Unfermented residues Starch washing Ethanol Water Starch Polyolefins polyurethane DDG, DDGS Corn steep liquor Flavonoids Phytosterols Carotenoids Polyamine-hydroxycinnamic acid amide conjugates FIGURE 20.6 Schematic corn wet milling and dry milling processes, with potential phytochemical coproduction.
360 20. PRODUCTION OF PHYTOCHEMICALS, DYES AND PIGMENTS AS COPRODUCTS IN BIOENERGY PROCESSES and germ as primary products, but also gluten, fiber, bran oil, or other phytochemicals as secondary products. The processing mainly comprises different levels of milling and fractionation (air classification, sieving, etc.), sometimes also with extraction (Kraus, 2006). Soybean, rapeseed, sun flower seed, peanut, olive, coconut, etc Handling, conditioning Hexane Hexane extraction Hexane recovery (Steam distillation) Water, acid, base, or enzyme Methanol Meal processing Meal Carotenoids Isoflavones Saponin Protease inhibitor Oil degumming Lecithin Oil deodorizing Phytosterols Tocopherols Transesterification Catalyst Phase separation Crude biodiesel Acid Free fatty acid separation Glycerin/water/methan ol separation Free fatty acid Crude glycerin FIGURE 20.7 Schematic vegetable oil and biodiesel production processes, with potential phytochemical coproduction. Vegetable oil production involves multitier, multiphase steps (Figure 20.7). Major plant sources for vegetable oils include palm, soybean, rapeseed, sunflower seed, peanut, cotton seed, coconut, and olive, and (to less extent) corn, hazelnut, grape seed, sesame, flax seed, safflower, rice bran, etc. Besides oil and cake (or meal, oil-extracted residues), other coproducts come from mechanically or chemically separated substances prior to oil extraction as well as refining by-products. Degumming and deodorizing of crude vegetable oils result in the production of lecithin and tocopherols or phytosterols, respectively. General processes of vegetable oil production include feedstock disruption by mechanical, chemical or enzymatic means, mechanical pressing, phase separation, solvent extraction, and refining (degumming, neutralizing, bleaching and deodorizing) (Panpipat et al., 2012; Febrianto and Yang, 2011; Muth et al., 1998; Dunford, 2012). Postharvest processing of wheat, rice, oat, or other cereals generates not only flour or milled rice, bran, Production from Cultured Plant Cells High-value phytochemicals may be produced from cultured plant cells, which can be far more expensive and demand far more complex, highly specialized technologies (to promote and sustain the growth, propagation, vitality and productivity of the cells) in comparison with other methods. Such approaches are generally developed for phytochemicals of pharmaceutical uses, as exemplified by the production of paclitaxel or other taxoids from cultured Taxus plants cells. Anthocyanin production from Ajugareptans, Aralia, Euphorbia milli, Fragaria, Oxalis, Perilla, Vitis, grapes or carrot have also been explored (Chattopadhyay et al., 2008; Tripathi and Tripathi, 2003). Production from Microbial Fermentation Microbial fermentation may be a viable phytochemical-producing technology, alternative to those directly targeting plants. A fermentative process is independent of plant harvesting cycle, suits for process engineering and control, and could be more economical than plant cell culturing. Specifically selected wild-type or genetically engineered microbes are fed (in addition to N sources, growth-stimulators, and other medium ingredients) with inexpensive fermentable sugars such as isolated glucose, corn steep liquor, whey or other coproducts from various plants, crops or dairy processings (Gupta et al., 2011; Dufosse, 2006; Mapari et al., 2005; Adrio and Demain, 2003). After fermentation, various steps such as cell disruption and solvent extraction may be applied to obtain, enrich or purify phytochemical products. Full or semicommercial processes for phytochemical production by microbial fermentation have been developed, as exemplified by the production of riboflavin from fungi (Eremothecium ashbyii, Ashbya gossypi), yeast (Candida guilliermondii, Debaryomyces subglobosus), or bacteria (Clostridium acetobutylicum) (Chattopadhyay et al., 2008), as well as other vitamins (Shimizu, 2008). Carotenoids may also be produced by microbial fermentation, as exemplified by the production of b-carotene from B. trispora or Phycomyces blakesleeanus; lycopene from Fusarium sporotrichioides or bacterium Erwinia uredovora; zeaxanthin from a Flavobacterium sp.; astaxanthin from Xanthophyllomyces dendrorhous, Rhodotorula glutinis, Rhodotorula gracilis, Rhodotorula rubra or Rhodotorula graminis; canthaxanthin from bacterium
COPRODUCTION OF PHYTOCHEMICALS IN BIOENERGY PROCESSES Bradyrhizobium sp.; and isorenieratene from bacterium Brevibacterium aurartiacum, Streptomyces mediolani, or Mycobacterium aurum. Production of certain therapeutic phytochemicals in microbial fermentation has been reported as well (Demain and Adrio, 2008). Production from Algae via Aquaculture As known producers of many compounds identical or homologous to plant-derived phytochemicals of industrial interest, algae have been explored for phytochemical production. Currently, the majority of commercial b-carotene is produced from Dunaliella salina and Dunaliella bardawil. Astaxanthin may be produced from Haematococcus lacustris; canthaxanthin from H. lacustris, Coelastrella striolata or Chlorella zofingiensis; and lutein from Muriellopsis sp., Scenedesmus almeriensis or Chlamydomonas zofingiensis (Skjanes et al., 2012; Chattopadhyay et al., 2008). Algae may also produce vitamins and bioactive or dietary amino acids, proteins (e.g. phycobiliproteins from Spirulina (Arthospira) platensis), lipids or fatty acids, or phycocolloids (agar, carrageenan and alginate) (Brennan et al., 2012; Becker, 2004). Those algae may be grown and harvested either outdoor (aquaculture) or indoor inside factory tanks, as selected wild types or genetically engineered strains. COPRODUCTION OF PHYTOCHEMICALS IN BIOENERGY PROCESSES Productions of biofuels (bioenergy processes) use biobased (sustainable) feedstocks, and convert energylatent plant or algal molecules (formed by photosynthesis) to molecules more suited or amenable as fuels (see chapters 10, 11, 15, 18e20 of this book). The key intermediate, platform molecules of bioenergy processes, such as glucose for starchy or cellulosic bioethanol production, may be used to produce nonfuel biochemicals, such as organic acids, polyols, polymers or plastics (Clark et al., 2012; Dapsens et al., 2012; Koutinas et al., 2007, 2006). As the economics and scale of biofuel and biochemical productions continue to grow, it becomes more important to enhance the values of various coand by-products, so that the biofuel or biochemical production can be upgraded or expanded to a more efficient biorefinery capable of maximally utilizing biobased feedstocks. Such expansion is desirable considering the colocation and energy/material source sharing, as well as the integrability of the existing phytochemical production technologies (mentioned in Section (Production of Industrial Phytochemicals)), with the bioenergy processes. Many approaches might be taken to recover phytochemicals during bioenergy or biochemical 361 production processes. For instance, an upstream fractionation (e.g. dry milling or air classification) might be added to allow further processing of crudely separated feedstock components. Coproduction from Starch- or Sugar-Based Bioenergy Processes Starch- or sugar(cane)-based bioethanol processes are fully commercialized, and are the major bioethanol providers at present. The main feedstocks are corn and sugarcane, and to a less extent, potato, cassava and sugar beet. Starch is converted by amylolytic enzymes to fermentable sugars (mainly glucose), sucrose is squeezed out from cane or beet, and sugars are fermented by yeast to ethanol. For corn ethanol processes, the main coproducts are DDGS for feed, as well as steep liquor, gluten meal, corn oil, and fiber from wet milling (Zhang et al., 2012, Figure 20.6). Main by-products are corn stover and cob for corn ethanol processes, and bagasse for sugarcane ethanol process. These byproducts are currently being developed as feedstocks for lignocellulosic ethanol. Many phytochemicals of industrial interest might be obtained or derived from the co- or by-products of starch- or sugar-based bioenergy processes, as exemplified in Section (Coproduction from Processing (Biorefinery) of Staple Crops). Obtaining betaine from sugar beet has been shown (Kripp, 2006). It has also been reported that polyolefins (e.g. polypropylene and polyethylene), polymerized polyurethane or other biomaterials may be made from DDGS (Diebel et al., 2012, Tatara et al., 2007; Cheesbrough et al., 2008). Coproduction from Plant Oil-Based Bioenergy Processes Plant oil-based biodiesel processes are commercialized (Figure 20.7). Biodiesel productions use mostly extracted plant or vegetable oils, isolated animal fats (from meat or dairy productions), or to less extent recycled cooking oils, and transesterify the triglycerides with short-chain primary alcohols (e.g. methanol or (bio)ethanol) to make fatty acid esters (diesels). The oil extraction from plants (solideliquid extraction) may allow coextraction of numerous phytochemicals as coproducts. The main coproduct from the transesterification is glycerol, which can be used directly as solvent, or as feedstock for microbial fermentation production of other biochemicals, such as propanediol, dihydroxyacetone, succinic acid, polyglycerols or polyhydroxyalkanoate, or for hydrothermal-chemical conversion to H2 (Khanna et al., 2012; Kosmider et al., 2011; Kannan et al., 2012). Grown on glycerol,
362 20. PRODUCTION OF PHYTOCHEMICALS, DYES AND PIGMENTS AS COPRODUCTS IN BIOENERGY PROCESSES astaxanthin production by bacterium Phaffia rhodozyma (X. dendrorhous) or Sporobolomyces ruberrimus (Valduga et al., 2009), b-carotene by B. trispora (Mantzouridou et al., 2008), prodigiosin or other carotenoids by R. glutinis (da Silva et al., 2009), various carotenoids by Rhodosporidium paludigenum (Yimyoo et al., 2011), as well as b-carotene by microalgae Clamidomonas acidophila (Langner et al., 2009), astaxanthin by Schizochytrium sp., and various carotenoids by Thraustochytrium have been demonstrated. Coproduction from Lignocellulose (Biomass)Based Bioenergy Processes Lignocellulosic (or second-generation) bioenergy is being intensively developed, due to its use of renewable but nonfood or feed feedstocks; valorization of agricultural, forestry, first-generation bioenergy, or municipal by-products or waste; and potential to significantly replace fossil feedstocks for energy or chemical industries. From lignocellulosic biomass materials, cellulose and hemicellulose are converted by (hemi)cellulolytic enzymes or chemical means to fermentable sugars (mainly glucose and xylose), which are then fermented by yeast or bacteria to ethanol or other chemicals. The processes could run alone, fed by selected biomass feedstocks, or along with the starch/sugar bioethanol or biodiesel processes, fed by the lignocellulosic by-products from firstegeneration bioenergy processes. The main by-product from lignocellulosic bioenergy processes (Figure 20.8) is lignin or lignaceous residue, whose valorizations are the focus of rigorous research efforts and may include uses for the production of phenolics (e.g. vanillin, vinyl guaiacol, ferulic acid), lignans, carbon fiber, or as additives for paper and pulp industry, roadbed construction, or soil augmentation (Ceylan et al., 2012; Chapters 22, 23 of this book). Other byproducts from lignocellulosic bioenergy processes include stillage and pretreatment liquor, which may be inhibitory to fermentation or enzymatic hydrolysis but rich in phenolics and oligosaccharides (Klinke et al., 2002; Persson et al., 2002). Mechanically separated biomass components (upstream to pretreatment, hydrolysis, and fermentation) may also serve as sources for Biomass feedstock Enzymes Biomass pretreatment (Hemi)cellulose hydrolysis Pretreatment liquor Phenolics Oligosaccharides phytochemicals, such as tree barks for tannin production. Woods, especially those not suited for conventional forestry products, are attractive feedstocks for lignocellulosic bioenergy. Prior to enzymatic or chemical conversion to fermentable sugars, woody materials might be subjected to treatments (such as the leaching processes widely used for dedicated phytochemical production, as mentioned in Section (Extraction and Isolation from Specific Plants) to yield extractives comprising phenolics (phenols, flavonoids, and anthocyanins), terpenoids (essential oils), nitrogen-containing phytochemicals (alkaloids) or organic acids (citric, oxalic, acetic, malic, benzoic, etc.) (Huang and Ramaswamy, 2012; Turley et al., 2006). In addition to agricultural and forestry by-products (e.g. corn stover, wheat straw, and wood residues), switchgrass and other dedicated “energy crops” may serve as viable feedstocks for not only bioenergy but also phytochemical coproducts. For instance, valued phytochemicals like antioxidants and flavonoids might be extracted from switchgrass prior to the pretreatment of the bioenergy process (Huang and Ramaswamy, 2012; Uppugundla et al., 2009; Wang and Weller, 2006). Coproduction from Bio-Oil, Syn-Gas, or Algal Bioenergy Processes Bio-oil and syn-gas are fuels chemically converted from lignocellulosic feedstocks. Aggressively pursued and developed as viable bioenergy, bio-oil and syn-gas productions rely on thermal-chemical conversion (or pyrolysis) of biomass or other lignocellulosic feedstocks to combustible oily substances (comprising numerous oxygenated hydrocarbons) or H2eCO gas mix, respectively. Depending on process conditions, up to hundreds of compounds may be present in bio-oils, with numerous chemicals of interest (other than combustibility) among them (Abou-Zaid and Scott, 2012; Venderbosch and Prins, 2010; Briens et al., 2008; Demirbas, 2009). These may include polyphenols, proanthocyanidins, tannins, flavonoids or organic acids. Some of the (phenolic) compounds may possess biocidal activity, making them useful as pesticide, bactericide, antitermite Ethanol Glucose, xylose fermentation Phenolics Lignans Carbon fiber Distillation: ethanol recovery Lignin residue FIGURE 20.8 A schematic lignocellulosic bioenergy process, with potential phytochemical coproduction.
REFERENCES agent, or wood preservatives (Di Blasi et al., 2010). For instance, bio-oil made from tobacco can have antimicrobes or insect activity (Hossain et al., 2013). Bio-oil made from lignin may provide substances that replace formaldehyde-phenol resins in particle board (Venderbosch and Prins, 2010). Algae have attracted intensive research and development efforts for bioenergy production, because algal processes might directly be driven by photosynthesis (thus fixing CO2) or yield hydrocarbons (for drop-in refining or use as fuel). The main by-product from algal bioenergy production is the post-hydrocarbon-harvest algal mass, which might be used as feed or fertilizer. Some algae species can produce phenolics, terpenoids, carotenoids, alkaloids or sterols at significant levels (Huang and Ramaswamy, 2012; Brennan et al., 2012). Phytochemical productions from algae (as mentioned in Section (Production from Algae via Aquaculture)) might be combined with hydrocarbon production, to further valorize algal bioenergy processes. Colocation of Fermentative Phytochemicals Production with Bioenergy Processes Microbial fermentation production of valuable phytochemicals may be colocated (onsite, integral use of materials or energy streams) with starch- or lignocellulose-based bioenergy processes, to benefit from locally produced, inexpensive fermentable sugars (intermediates or by-products of bioenergy processes) (Thomsen et al., 2006). In principle, all microbial fermentations for biochemicals production may be run on sugars converted from starch, sugarcane or lignocellulose in various bioenergy processes, to yield chemicals like surfactants, polyols or organic acids (Choi et al., 2007; Aalford and Morel, 2006; Mapari et al., 2005). For instance, X. dendrorhous can be grown on cellulasesdigested pine and produce carotenoids (Chattopadhyay et al., 2008), and Serratia marcescens can be grown on processed cassava waste to produce prodigiosin (Casullo de Araújo et al., 2010). Utilization of Phytochemical Production By-Products for Bioenergy Production of valuable phytochemicals (such as those for therapeutic, cosmetic, dietary or agricultural uses) from either wild-type or transgenic plants generates lignocellulosic by-products. Such materials might serve as feedstocks for (onsite) bioenergy production. This might add value, reduce waste, and enhance raw material or energy use efficiency. For instance, the woody residues from vinblastine or vincristine production in C. roseus (Braz-Filho, 1999) or other natural products 363 production (Simard et al., 2012), or the citrus peel residues from furanocoumarins, flavonoid glycosides, polymethoxylated flavones, triterpenoids, limonoids or peel oil extraction (Manthey, 2012), have potential as bioenergy feedstocks. Phytochemical production has focused mainly on therapeutic, dietary or cosmetic agents from specific fruits, flowers, nuts, vegetables, or other plant sources. In comparison, less attention has been paid on phytochemicals coproduction in current or future bioenergy processes. Integral coproductions of biofuels, biochemicals, phytochemicals and other valuable materials are imperative for highly efficient and viable bioenergy and biorefinery processes (Huang and Ramaswamy, 2012). In some cases, relatively simple combinations or colocations of existing bioenergy and phytochemicals processes may suffice to coproduce biofuels, biochemicals and phytochemicals. In other cases, new production technology or process engineering may need to be developed. To maximize the economy of raw materials and energy utilization and minimize the carbon footprint, bioenergy processes will evolve into more comprehensive biorefineries in which the coproduction of industrial phytochemicals plays an important role. References Aalford, S.N., Morel du Boil, P.G., 2006. A survey of value addition in the sugar industry. Proc. S. Afr. Sug. 80, 39e61. Abad, S., Turon, X., 2012. Valorization of biodiesel derived glycerol as a carbon source to obtain added-value metabolites: focus on polyunsaturated fatty acids. Biotechnol. Adv. 30, 733e741. Abou-Zaid, M., Scott, I.M., 2012. Pyrolysis bio-oils from temperate forests: fuels, phytochemicals and bioproducts. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefinery Co-products. Phytochemicals, Primary Metabolites and Value-Added Biomass Processing. John Wiley, Chichester, UK, pp. 311e325. Adrio, J.L., Demain, A.L., 2003. Fungal biotechnology. Ind. Microbiol. 6, 191e199. Becker, W., 2004. Microalgae in human and animal nutrition. In: Richmond, A. (Ed.), Handbook of Microalgae Culture: Biotechnology and Applied Phycology. Wiley-Blackwell, Oxford, pp. 312e351. Braz-Filho, R., 1999. Brazilian phytochemical diversity: bioorganic compounds produced by secondary metabolism as a source of new scientific development, varied industrial applications and to enhance human health and the quality of life. Pure Appl. Chem. 71, 1663e1672. Brennan, L., Mostaert, A., Murphy, C., Owende, P., 2012. Phytochemicals from algae. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefinery Co-products. Phytochemicals, Primary Metabolites and Value-added Biomass Processing. John Wiley, Chichester, UK, pp. 199e240. Briens, C., Piskorz, J., Berruti, F., 2008. Biomass valorization for fuel and chemicals production - a review. Int. J. Chem. React. Eng. 6, 1e49. Casullo de Araújo, H.W., Fukishima, K., Campos Takaki, G.M., 2010. Prodigiosin production by Serratia marcescens USC 1549 using renewable-resources as a low cost substrate. Molecules 15, 6931e6940.
364 20. PRODUCTION OF PHYTOCHEMICALS, DYES AND PIGMENTS AS COPRODUCTS IN BIOENERGY PROCESSES Cerella, C., Kelkel, M., Viry, E., Dicato, M., Jacob, C., Diederich, M., 2011. Naturally occurring organic sulfur compounds: an example of a multitasking class of phytochemicals. In: Rasooli, I. (Ed.), Phytochemicals - Bioactivities and Impact on Health. InTech, Rijeka, Croatia, pp. 3e42. Ceylan, H., Kim, S., Gopalakrishnan, K., 2012. Sustainable utilization of bio-fuel co-product in roadbed stabilization. In: Gopalakrishnan, K., van Leeuwen, J.H., Brown, R.C. (Eds.), Sustainable Bioenergy and Bioproducts: Value Added Engineering Applications (Green Energy and Technology). Springer-Verlag, London, pp. 117e130. Chattopadhyay, P., Chatterjee, S., Sen, S.K., 2008. Biotechnological potential of natural food grade biocolorants. Afr. J. Biotechnol. 7, 2972e2985. Cheesbrough, V., Rosentrater, K.A., Visser, J., 2008. Properties of distillers grains composites: a preliminary investigation. J. Polym. Environ. 16, 40e50. Choi, S.W., Lee, S.K., Kim, E.O., Oh, J.H., Yoon, K.S., Parris, N., Hicks, K.B., Moreau, R.A., 2007. Antioxidant and antimelanogenic activities of polyamine conjugates from corn bran and related hydroxycinnamic acids. J. Agric. Food Chem. 55, 3920e3925. Clark, J.H., Luque, R., Matharu, A.S., 2012. Green chemistry, biofuels, and biorefinery. Annu. Rev. Chem. Biomol. Eng. 3, 183e207. da Fonseca, R.A.S., da Silva Rafael, R., Kalil, S.J., Burkert, C.A.V., Burkert, J.F.M., 2011. Different cells disruption methods for astaxanthin recovery by Phaffia rodozyma. Afr. J. Biotechnol. 10, 1165e1171. da Silva, G.P., Mack, M., Contiero, J., 2009. A promising and abundant carbon source for industrial microbiology. Biotechnol. Adv. 27, 30e39. Dapsens, P.Y., Mondelli, C., Pérez-Ramı́rez, J., 2012. Biobased chemicals from conception toward industrial reality: lessons learned and to be learned. ACS Catal. 2, 1487e1499. Dayan, F.E., Cantrell, C.L., Duke, S.O., 2009. Natural products in crop protection. Bioorg. Med. Chem. 17, 4022e4034. Demain, A.L., Adrio, J.L., 2008. Strain improvement for production of pharmaceuticals and other microbial metabolites by fermentation. Prog. Drug Res. 65, 251e289. Demirbas, A., 2009. Biorefineries: current activities and future developments. Energ. Convers. Manage. 50, 2782e2801. Di Blasi, C., Branca, C., Galgano, A., 2010. Biomass screening for the production of furfural via thermal decomposition. Ind. Eng. Chem. Res. 49, 2658e2671. Diebel, W., Reddy, M.M., Misra, M., Mohanty, A.K., 2012. Material property characterization of co-products from biofuel industries: potential uses in value-added biocomposites. Biomass Bioenergy 37, 88e96. Dufossé, L., 2006. Microbial production of food grade pigments. Food Technol. Biotechnol. 44, 313e321. Dunford, N.T., 2012. Co-products from cereal and oilseed biorefinery systems. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefinery Co-Products. Phytochemicals, Primary Metabolites and Value-Added Biomass Processing. John Wiley, Chichester, UK, pp. 93e115. Febrianto, N.A., Yang, T.A., 2011. Producing high quality edible oil by using eco-friendly technology: a review. Adv. J. Food Sci. Technol. 3, 317e326. Frihart, C., 2010. Biobased adhesives and non-conventional bonding. In: Rowell, R.M., Caldeira, F., Rowell, J.K. (Eds.), Sustainable Development in the Forest Products Industry. Universidade Fernando Pessoa, Porto, Portugal, pp. 98e113. Guclu-Ustundag, O., Mazza, G., 2007. Saponins: properties, applications and processing. Crit. Rev. Food Sci. Nutr. 47, 231e258. Gupta, C., Garg, A.P., Prakash, D., Goyal, S., Gupta, S., 2011. Microbes as potential source of biocolours. Pharmacol Online. 2, 1309e1318. Hossain, M.M., Scott, I.M., McGarvey, B.D., Conn, K., Ferrante, L., Berruti, F., Briens, C., 2013. Toxicity of lignin, cellulose and hemicellulose-pyrolyzed bio-oil combinations: estimating pesticide resources. J. Anal. Appl. Pyrol. 99, 211e216. Huang, H.J., Ramaswamy, S., 2012. Separation and purification of phytochemicals as co-products in biorefineries. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefinery Co-Products. Phytochemicals, Primary Metabolites and Value-Added Biomass Processing. John Wiley, Chichester, UK, pp. 37e53. Huter, O.F., 2011. Use of natural products in the crop protection industry. Phytochem. Rev. 10, 185e194. Kannan, A., Rayaprolu, S., Hettiarachchy, N., 2012. Bioactive soy co-products. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefinery Co-Products. Phytochemicals, Primary Metabolites and Value-Added Biomass Processing. John Wiley, Chichester, UK, pp. 117e131. Khanna, S., Goyal, A., Moholkar, V.S., 2012. Microbial conversion of glycerol: present status and future prospects. Crit. Rev. Biotechnol. 32, 235e262. Klinke, H.B., Ahring, B.K., Schmidt, A.S., Thomsen, A.B., 2002. Characterization of degradation products from alkaline wet oxidation of wheat straw. Bioresour. Technol. 82, 15e26. Kosmider, A., Leja, K., Czaczyk, K., 2011. Improved utilization of crude glycerol by-product from biodiesel production. In: Montero, G. (Ed.), Biodiesel - Quality, Emissions and By-products. InTech, Rejika, Croatia, pp. 341e364. Koutinas, A.A., Wang, R., Campbell, G.M., Webb, C., 2006. A whole crop biorefinery system: a closed system for the manufacture of nonfood products from cereals. In: Kamm, B., Gruber, P.R., Kamm, M. (Eds.), Biorefineries - Industrial Processes and Products: Status Quo and Future Directions, vol. 1. Wiley-VCH, Weinheim, pp. 165e191. Koutinas, A.A., Wang, R.H., Webb, C., 2007. The biochemurgist bioconversion of agricultural raw materials for chemical production. Biofuels., Bioprod. Biorefin. 1, 24e38. Kraus, G.A., 2006. Phytochemicals, dyes, and pigments in the biorefinery context. In: by Kamm, B., Gruber, P.R., Kamm, M. (Eds.), Biorefineries - Industrial Processes and Products: Status Quo and Future Directions, vol. 2. Wiley-VCH, Weinheim, pp. 315e323. Kripp, T.C., 2006. Biobased consumer products for cosmetics. In: Kamm, B., Gruber, P.R., Kamm, M. (Eds.), Biorefineries - Industrial Processes and Products: Status Quo and Future Directions, vol. 2. Wiley-VCH, Weinheim, pp. 409e442. Langner, U., Jakob, T., Stehfest, K., Wilhelm, C., 2009. An energy balance from absorbed photons to new biomass for Chlamydomonas reinhardtii and Chlamydomonas acidophila under neutral and extremely acidic growth conditions. Plant Cell Environ. 32, 250e258. Liu, R.H., 2007. Whole grain phytochemicals and health. J. Cereal Sci. 46, 207e219. Manthey, J.A., 2012. Potential value-added co-products from citrus fruit processing. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefineries, Biorefinery Co-products. Phytochemicals, Primary Metabolites and Value-added Biomass Processing. John Wiley, Chichester, UK, pp. 153e178. Mantzouridou, F., Naziri, E., Tsimidou, M.Z., 2008. Industrial glycerol as a supplementary carbon source in the production of betacarotene by Blakeslea trispora. J. Agric. Food Chem. 56, 2668e2675. Mapari, S.A.S., Nielsen, K.F., Larsen, T.O., Frisvad, J.C., Meyer, A.S., Thrane, U., 2005. Exploring fungal biodiversity for the production of water-soluble pigments as potential natural food colorants. Curr. Opin. Biotechnol. 16, 231e238. Matthews, P.D., Wurtzel, E.T., 2007. Biotechnology of food colorant production. In: Socaciu, C. (Ed.), Food Colorants: Chemical and Functional Properties. CRC Press, Boca Raton, FL, USA, pp. 347e398.
REFERENCES Moreau, R.A., Lampi, A.M., Hicks, K.B., 2009. Fatty acid, phytosterol, and polyamine conjugate profiles of edible oils extracted from corn germ, corn fiber, and corn kernels. J. Am. Oil Chem. Soc. 86, 1209e1214. Mortensen, A., 2006. Carotenoids and other pigments as natural colorants. Pure Appl. Chem. 78, 1477e1491. Muth, M.K., Depro, B.M., Domanico, J.L., 1998. Vegetable Oil Production: Industry Profile. US Environmental Protection Agency document A.98.41. http://www.epa.gov/ttnecas1/regdata/IPs/ Vegetable%20Oil_IP.pdf. O’Brien, K.P., Franjevic, S., Jones, J., 2009. Green Chemistry and Sustainable Agriculture: The Role of Biopesticides. http:// advancinggreenchemistry.org/wp-content/uploads/Green-Chemand-Sus.-Ag.-the-Role-of-Biopesticides.pdf. Pandey, M., Debnath, M., Gupta, S., Chikara, S.K., 2011. Phytomedicine: an ancient approach turning into future potential source of therapeutics. J. Pharmacogn. Phytother. 3, 27e37. Panpipat, W., Xu, X., Guo, Z., 2012. Towards a commercially potential process: enzymatic recovery of phytosterols from plant oil deodoriser distillates mixture. Process Biochem. 47, 1256e1262. Persson, P., Andersson, J., Gorton, L., Larsson, S., Nilvebrant, N.O., Jönsson, L.J., 2002. Effect of different forms of alkali treatment on specific fermentation inhibitors and on the fermentability of lignocellulose hydrolysates for production of fuel ethanol. J. Agric. Food Chem. 50, 5318e5325. Piironen, V., Lindsay, D.G., Miettinen, T.A., Toivo, J., Lampi, A.M., 2000. Plant sterols: biosynthesis, biological function and their importance to human nutrition. J. Sci. Food Agric. 80, 939e966. Rao, V., 2012. In: Phytochemicals as Nutraceuticals e Global Approaches to Their Role in Nutrition and Health. (ed.). InTech, Rijeka, Croatia. Rausch, K., 2012. Phytochemicals from corn: a processing perspective. In: Bergeron, C., Carrier, D.J., Ramaswamy, S. (Eds.), Biorefineries, Biorefinery Co-products. Phytochemicals, Primary Metabolites and Value-added Biomass Processing. John Wiley, Chichester, UK, pp. 55e92. Rausch, K.D., Belyea, R.L., 2006. The future of co-products from corn processing. Appl. Biochem. Biotechnol., 47e86. Rymbai, H., Sharma, R.R., Srivastav, M., 2011. Biocolorants and its implications in health and food industry - a review. Int. J. Pharm. Tech. Res. 3, 2228e2244. Salim, A.A., Chin, Y.W., Kinghorn, A.D., 2008. Drug discovery from plants. In: Ramawat, K.G., Mérillon, J.M. (Eds.), Bioactive Molecules and Medicinal Plants. Springer, Berlin, pp. 1e24. Sengupta, S., 2003. Natural, “green” Dyes for the Textile Industry. The Massachusetts Toxics Use Reduction Institute University of Massachusetts Lowell Technical Report No. 57. http://www.google. com/url?sa¼t&rct¼j&q¼natural%20green%20dyes%20for%20the %20textile%20industry&source¼web&cd¼1&ved¼0CC8QFjAA &url¼http%3A%2F%2Fwww.turi.org%2Fcontent%2Fdownload% 2F3242%2F29575%2Ffile%2FReport%252057%2520-%2520 sengupta.pdf&;ei¼tuovUdOTGZCZhQfu74CYCg&usg¼AFQj CNGU8AeXikvG83-fwL6jvTrs8t2qXw" (accessed 28.02.13.). Shimizu, S., 2008. Vitamins and Related Compounds: Microbial Production. In Biotechnology Set Second edition volume 10eSpecial Processes IVeSpecial Substances. Wiley-VCH Verlag, Weinheim. pp. 318e340. Simard, F., Pichette, A., Legault, J., 2012. New bioactive natural products from Canadian boreal forest. In: Bergeron, C., Carrier, D.J., 365 Ramaswamy (Eds.), Biorefineries, Biorefinery Co-products. Phytochemicals, Primary Metabolites and Value-added Biomass Processing. John Wiley, Chichester, UK, pp. 241e258. Skjånes, K., Rebours, C., Lindblad, P., 2012. Potential for green microalgae to produce hydrogen, pharmaceuticals and other high value products in a combined process. Crit. Rev. Biotechnol. http://dx.doi.org/10.3109/07388551.2012.681625. Spolaore, P., Joannis-Cassan, C., Duran, E., Isambert, A., 2006. Commercial applications of microalgae. J. Biosci. Bioeng. 101, 87e96. Tatara, R.A., Suraparaju, S., Rosentrater, K.A., 2007. Compression molding of phenolic resin and corn-based DDGS blends. J. Polym. Environ. 15, 89e95. Thomsen, M.H., Andersen, M., Kiel, P., 2006. Plant juice in the biorefinery - use of plant juice as fermentation medium. In: Kamm, B., Gruber, P.R., Kamm, M. (Eds.), Biorefineries - Industrial Processes and Products: Status Quo and Future Directions, vol. 1. WileyVCH, Weinheim, pp. 295e374. Tripathi, L., Tripathi, J.N., 2003. Role of biotechnology in medicinal plants. Trop. J. Pharm. Res. 2, 243e253. Turley, D.B., Chaudhry, Q., Watkins, R.W., Clark, J.H., Deswarte, F.E.I., 2006. Chemical products from temperate forest tree species developing strategies for exploitation. Ind. Crop Prod. 24, 238e243. Uppugundla, N., Engelberth, A., Ravindranath, S.V., Clausen, E.C., Lay, J.O., Gidden, J., Carrier, D.J., 2009. Switchgrass water extracts: extraction, separation and biological activity of rutin and quercitrin. J. Agric. Food Chem. 57, 7763e7770. Valduga, E., Tatsch, P.O., Tiggemann, L., Zeni, J., Colet, R., Cansian, J.M., Treichel, H., Luccio, M., 2009. Evaluation of the conditions of carotenoids production in a synthetic medium by Sporidiobolus salmonicolor (CBS 2636) in a bioreactor. Int. J. Food Sci. Technol. 44, 2445e2451. van den Oever, M.J.A., Boeriu, C.G., Blaauw, R., van Haveren, J., 2004. Colorants based on renewable resources and food-grade colorants for application in thermoplastics. J. Appl. Polym. Sci. 92, 2961e2969. Venderbosch, R.H., Prins, W., 2010. Fast pyrolysis technology development. Biofuel. Bioprod. Biorefin. 4, 178e208. Wang, L., Weller, C.L., 2006. Recent advances in extraction of nutraceuticals from plants. Trends Food Sci. Technol. 17, 300e312. Watkins, R., Chaudhry, Q., 2013. Useful Chemicals from the Main Commercial Tree Species in the UK. The Food and Environment Research Agency of UK, online database (accessed 28.02.13.). https://secure.fera.defra.gov.uk/treechemicals/. Wu, X., Kang, J., 2011. Phytochemicals in soy and their health effects. In: Rasooli, I. (Ed.), Phytochemicals - Bioactivities and Impact on Health. InTech, Rijeka, Croatia, pp. 43e76. Yimyoo, T., Yongmanitchai, W., Limtong, S., 2011. Carotenoid production by Rhodosporidium paludigenum DMKU3-LPK4 using glycerol as the carbon source. Kasetsart J. Nat. Sci. 45, 90e100. Zhang, X., Beltranena, E., Christensen, C., Yu, P., 2012. Use of a dry fractionation process to manipulate the chemical profile and nutrient supply of a coproduct from bioethanol processing. J. Agric. Food Chem. 60, 6846e6854. Zhao, Z., Moghadasian, M.H., 2010. Bioavailability of hydroxycinnamates: a brief review of in vitro and in vivo studies. Phytochem. Rev. 9, 133e145. Zijlstra, R.T., van Kessel, A.G., Drew, M.D., 2012. Ingredient Fractionation: The Value of Value-added Processing for Animal Nutrition (accessed 28.02.13.). http://www.prairieswine.com/ pdf/1713.pdf.
C H A P T E R 21 Recent Developments on Cyanobacteria and Green Algae for Biohydrogen Photoproduction and Its Importance in CO2 Reduction Y. Allahverdiyeva*, E.M. Aro, S.N. Kosourov Department of Biochemistry, University of Turku, Turku, Finland *Corresponding author email: allahve@utu.fi O U T L I N E Introduction 367 Mechanisms of Hydrogen Photoproduction Oxygenic Photosynthesis Nitrogenases Alternative Nitrogenases Hydrogenases Uptake Hydrogenases Bidirectional Hydrogenases Green Algal [FeeFe]-Hydrogenases 368 368 369 369 370 370 370 371 Hydrogen Photoproduction by Cyanobacteria Strategies to Improve H2 Production in Cyanobacteria Genetic Modifications Introducing Foreign Enzymes and Semiartificial Systems 372 373 373 374 374 375 Hydrogen Photoproduction by Green Algae Light-Dependent Hydrogen Production Pathways Role of H2 Photoproduction in Green Algae Long-Term H2 Production by Green Algae Hydrogen Photoproduction by Nutrientdeprived Green Algae Strategies to Improve H2 Photoproduction in Green Algae 375 375 377 377 References 382 378 381 373 INTRODUCTION It is well known that the fossil energy resources are limited. Despite the fact that millions of years of photosynthesis were required to ensure the fossil fuel formation and accumulation, the current consumption of fossil fuels occurs at a rapid rate. Such utilization of fossil fuels creates extreme damage to the environment, increasing the CO2 level in atmosphere and leading to global warming and pollution on the Earth. Future scenarios predict an increase in CO2 partial pressure in the atmosphere from the current levels of approximately 380, to about 750 and up to 1000 matm until the end of this century (Raupach et al., 2007). Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00021-8 Elimination of Competing Electron Transfer Routes Biodiversity Immobilization There is an urgent need to switch to alternative, environmentally friendly and renewable energy sources. An efficient strategy for production of bioenergy would employ photosynthetic microorganisms, which are collectively, a significant player in the global carbon cycle. Cyanobacteria and green algae have inherited mechanisms for production of hydrogen, which possesses all properties of a clean and efficient energy carrier. Although the natural production of hydrogen by these microorganisms is negligible at the current state, there is a huge potential for engineering and synthetic biology advances of cyanobacteria and green algae toward commercially profitable production of hydrogen and other biofuels. 367 Copyright Ó 2014 Elsevier B.V. All rights reserved.
368 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION MECHANISMS OF HYDROGEN PHOTOPRODUCTION The electrons extracted from water on the lumenal side of PSII are transferred via the PSII reaction center, plastoquinone (PQ), the Cyt b6f complex and plastocyanin to PSI, which after excitation directs electrons to ferredoxin (Fd), ferredoxin-NADPþ reductase (FNR) and, finally, to generate reduced nicotinamide adenine dinucleotide phosphate (NADPH). This process is known as the linear electron transport (LET). Concomitantly with electron transfer reactions, protons are transferred inside of the thylakoid lumen creating a proton gradient across the thylakoid membrane, which in turn drives adenosine triphosphate (ATP) production via the ATP synthase complex. Sometimes, the electrons are recycled from NADPH or Fd to PQ in the process known as the cyclic electron transport, whereby DpH is generated without production of NADPH. NADPH produced by LET is further used by carbon metabolism, and many other metabolic pathways. The excess of reduced carbon is stored in cells as carbohydrates or lipids. An unique feature of photosynthetic microorganisms is that under specific conditions, most of them are able to redirect the flow of electrons originated from water splitting to the enzymes that mediate H2 production (Figure 21.1). Biophotolysis of water by microalgae has been under investigation for over 70 years. H2 production by the anaerobically adapted and CO2-depleted suspension of Scenedesmus obliquus in light was reported for the first time by Gaffron and Rubin (1942). Three decades later, it was revealed that filamentous cyanobacteria, Anabaena cylindrica is also able to evolve H2 and O2 simultaneously under Ar atmosphere (Benemann and Weare, 1974). Despite intensive research on the structure and function of photosynthetic protein complexes, we are still lacking a fundamental understanding of the molecular factors regulating the entire electron transfer chain from water to H2 in oxygenic photosynthetic organisms. In this chapter, we mainly focus on H2 production by oxygenic photosynthetic microorganisms via the Oxygenic Photosynthesis Cyanobacteria and green algae are photosynthetic microorganisms widespread in nature that survive even in extreme climatic conditions. They are able to harness solar energy and convert it into chemical energy by simultaneous splitting of water to molecular oxygen and protons with following fixation of CO2, according to the general equation of photosynthesis: 6CO2 þ 6H2 O / 6ðCH2 OÞ þ 6O2 (21.1) The photosynthetic electron transfer reactions are usually divided into two stagesdthe “light reactions”, where light energy is converted into the chemical energy of strong reductants and the “dark reactions”, where CO2 is reduced into organic compounds by using chemical energy obtained from the light reactions. The simplified scheme of photosynthetic light reactions is presented in Figure 21.1. Photosynthetic light reactions involve electron flow through three major protein complexes: photosystem II (PSII), Cytochrome b6f (Cyt b6 f), and photosystem I (PSI) embedded into the thylakoid membrane. The light reactions start with capture of photons by the pigment molecules in the antenna complexes and subsequent transfer of light energy to PSI and PSII reaction centers, where primary charge separation occurs and photosynthetic electron transport reactions are initiated. The two reaction centers, PSII and PSI, function simultaneously, but in series. PSII is the only known biocatalyst that can oxidize water, which is energetically a poor electron donor. The oxidationereduction midpoint potential of water is þ0.82 Vat pH 7. In PSII the photolysis of water is driven by the oxidized reaction center, P680þ (the midpoint potential of P680/P680þ is þ1.2 V at pH 7). FIGURE 21.1 Simplified schematic CO2 H2 2H+ H2 metabolism NAD(P)H 2H Carbohydrates 2O2 2H2O O2 RubisCO ? Carbon metabolism Flv1, Flv3 (Mehler) NADPH NADP+ ? NAD(P) ADP + P ATP FNR Fd QA QB ? NDH-2 Membrane Cyt b PQH Pheo PQ FB FA PQ FX A1 Cyt f PQH FeS P680 A0 P700 PC 2H O2 + 4H+ 2H2O PSII PC H+ NDH-1 Cyt b6 f PSI H+ ATP synt model of the electron transport routes in unicellular cyanobacteria. For details, see the text. Black arrows show the routes of photosynthetic electron transport. Dashed arrow shows that under certain conditions the electrons from the photosynthetic electron transport chain can be redirected to H2 metabolism allowing H2 photoproduction to occur. Dashed lines marked with ‘?’ show possible electron transfer routes via the NDH-2 complex and photorespiratory pathway. (For color version of this figure, the reader is referred to the online version of this book.)
MECHANISMS OF HYDROGEN PHOTOPRODUCTION light-dependent direct and indirect biophotolysis pathways. During direct biophotolysis H2 is derived from the electrons originated from water splitting at PSII, whereas for indirect biophotolysis electrons are mainly supplied by degradation of intracellular carbon compounds produced in photosynthetic carbon reduction reactions. Nitrogenases Many cyanobacteria are able to fix atmospheric N2 into ammonia (NH3) and produce H2 as a by-product. The reaction of nitrogen fixation is catalyzed by nitrogenase, a complex metalloenzyme and results in the formation of 1 mol of H2 per 1 mol of fixed N2 (Phelps and Wilson, 1942): þ  N2 þ 8H þ 8e þ 16ATP / 2NH3 þ 16ðADP þ Pi Þ þ H2 ; (21.2) where Pi is inorganic phosphate. Nitrogenases have relatively low turnover numbers. N2 fixation is an energy-expensive process that requires two ATP molecules per electron transfer. It has, however, an advantage of catalyzing an irreversible reaction and not being inhibited by H2 accumulation. The best studied type of nitrogenase is conventional molybdenum (Mo)-nitrogenase, which is encoded by the structural genes nifHDK1. Very little is known about the regulation of nif genes. Nitrogenase consists of two components: dinitrogenase, or the MoFe protein composed of NifD and NifK subunits, and dinitrogenase reductase, or the Fe protein, consisting of two subunits of NifH. The substrate-binding and reducing active site is located in the MoFe protein. The Fe protein containing a [4Fee4S] cluster and a Mg-ATP binding site acts as electron donor to a MoFe protein. This cluster accepts electrons from Fd or flavodoxin. The fixation of N2 is always accompanied by H2 evolution (Hadfield and Bulen, 1969). The reason for production of H2 as a by-product is not yet clear. It could be a result of unavoidable leakage of reducing potential, or formation of H2 could be a prerequisite for binding of N2 to the active site (Burgess and Lowe, 1996). Besides reducing N2, nitrogenase can reduce a number of other substrates with triple bonds. Importantly, in the absence of N2 as a substrate, nitrogenase exclusively catalyzes ATP-dependent reduction of Hþ to H2 (Benemann and Weare, 1974; Pickett, 1996). 8Hþ þ 8e þ 16ATP / 16ðADP þ Pi Þ þ 4H2 (21.3) Indeed, in terms of H2 production by N2 fixation in heterocystous cyanobacteria the N2 is a much more potent inhibitor than O2 (Yeager et al., 2011). This is logical, 369 due to the fact that heterocysts can protect enzymes from external O2. Since the replacement of N2 with argon (Ar) gas is an expensive approach for optimization of H2 production, an alternative method to genetically modify the catalytic site of the nitrogenase enzyme has been chosen as more appropriate. Recently, sitedirected mutations have been introduced to several amino acid residues coordinating the MoeFe active site of the nif1-enzyme in attempts to direct the electron flow selectively to H2 production in atmospheric N2 condition (Masukawa et al., 2010). Importantly, several mutant strains demonstrated nearly similar rate of H2 production under N2 and Ar atmosphere. Moreover, these strains accumulated significantly high levels of H2 under atmospheric N2 as compared to the reference strains. Alternative Nitrogenases In addition to the conventional Mo-nitrogenase, nif1, N2-fixing microorganisms possess also alternative nitrogenases: second type of Mo-nitrogenase, nif 2, vanadium (V)-nitrogenase, vnf, and Fe-nitrogenase, anf (Bothe et al., 2010). Presence of the Fe-nitrogenase in cyanobacteria has not yet been documented. Mo-nitrogenase 2, encoded by nifHDK2, is expressed in both vegetative cells and heterocysts of Anabaena variabilis under N2-fixing and anaerobic conditions (Schrautemeier et al., 1995; Thiel et al., 1995). Unicellular Chroococcidiopsis, inhabiting in a gypsum rock, where the shards provide a microaerobic, low light environment, also possesses the alternative nif 2 system. Based on the phylogenetic analysis of nif H sequences, it has been suggested that nif 2 is characteristic of unicellular or filamentous nonheterocystous cyanobacteria fixing N2 only under microaerobic conditions (Boison et al., 2004). Weyman and coworkers reported the amino acid substitution in nifD2 as a first step toward the development of nitrogenase mutants in A. variabilis, which produces large amounts of H2 in N2 containing atmosphere (Weyman et al., 2010). V-nitrogenase has a V-Fe cofactor in the active site. It is encoded by vnfHDGK genes and is expressed in heterocysts only under Mo-deficient conditions, in the presence of V (Kentemich et al., 1988; Thiel, 1993; Thiel et al., 1995). Biochemical and spectroscopic investigations of purified proteins isolated from Azotobacter vinelandii have revealed a mechanistic difference between the MoeFe and VeFe catalytic site and in H2 evolution mechanisms (Lee et al., 2009). N2 þ 12Hþ þ 12e þ 24ATP / 2NH3 þ 24ADP þ 24Pi þ 3H2 (21.4) As can be seen from the Eqns (21.2) and (21.4), the distribution of electrons and protons are different for Mo- and V-nitrogenases. V-nitrogenase can produce three
370 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION times more H2 per mole of N2 reduced compared to Mo-nitrogenase (Eady, 1996). For this reason, the production of H2 by vnf system is likely to be more efficient and it is, therefore, worth searching for organisms possessing alternative nitrogenases. For a long time the presence of alternative nitrogenases was confirmed only for A. variabilis and Anabaena azollae (Ni et al., 1990; Thiel, 1993). Recent screening of 14 different cyanobacterial strains has revealed 8 strains with nif2, and 4 strains with vnf nitrogenases (Masukawa et al., 2009), suggesting that alternative hydrogenases are not unique. Hydrogenases There are two classes of hydrogenases that commonly present in phototrophic organisms: the [FeeFe]hydrogenase and the [NieFe]-hydrogenase. The [FeeFe]hydrogenase is found in green algae and some bacteria, and it is the most active H2-forming enzyme. It demonstrates about 100 times higher activity than the [NieFe]hydrogenase. However, it is irreversibly inactivated when exposed to O2 (see Section Green Algal [FeeFe]-Hydrogenases in this chapter). Cyanobacteria possess two types of [NieFe]-hydrogenases (Houchins, 1984), which are more tolerant to O2 and only temporarily inactivated upon exposure to O2. Uptake Hydrogenases In cyanobacteria, the uptake hydrogenases (encoded by hupSL genes) catalyze the consumption of H2 produced by the nitrogenase. Thus, the net H2 evolution by N2-fixing cyanobacteria is barely observed under natural conditions. Uptake hydrogenase has been found in all N2-fixing cyanobacteria studied so far. Nevertheless, a few N2-fixing Synechococcus strains lacking an uptake hydrogenase have been reported (Ludwig et al., 2006; Steunou et al., 2008). It is believed that the uptake hydrogenase transfers electrons from H2 back to the photosynthetic and respiratory electron transport chains, and thus partially regains the energy used for N2 fixation. The cellular/subcellular localization of the uptake hydrogenase is controversial and seems to be species specific. The data obtained from the N2-fixing filamentous nonheterocystous cyanobaterium, Lyngbya majuscule, revealed higher specific labeling associated with the thylakoid membranes, suggesting that the cyanobacterial uptake hydrogenase is a membrane-bound protein (Seabra et al., 2009). However, it lacks a membrane-spanning region. Therefore, the presence of the third subunit, which would anchor the uptake hydrogenase to the membrane and link electron transfer from the enzyme to the respiratory or photosynthetic chains, has been suggested (Tamagnini et al., 2007). In some heterocystous cyanobacteria, such as Anabaena PCC 7120, the uptake hydrogenase enzyme was detected only in heterocysts, while in other cyanobacteria, such as Nostoc punctiforme, it is localized in both vegetative and heterocyst cells, corresponding most probably to inactive and active pools of the enzyme (Camsund et al., 2011; Seabra et al., 2009). Since the uptake hydrogenase is an obstacle for H2 production, mutations disrupting the structural hupSL genes have been constructed to improve the H2 production in N2-fixing cyanobacteria (Happe et al., 2000; Lindberg et al., 2002; Masukawa et al., 2002; Schutz et al., 2004; Yoshino et al., 2007; Khetkorn et al., 2012). These mutants produced about four- to sevenfold more H2 than the control strain. In addition, inactivation of uptake hydrogenase had no major effect on cell growth and heterocyst differentiation. Quantitative shotgun proteomics and physiological approaches on the uptake hydrogenase mutant of N. punctiforme demonstrated that the mutant strain undergoes metabolic and structural alterations to compensate for the amount of electrons lost as a release of H2 (Ekman et al., 2011). Construction of mutant strains combining several improvements is likely to be a better approach toward sustainable H2 production. To this end, the single- and the double-mutant strains lacking the homocitrate synthase genes, nifV1 and nifV2, were constructed using the DhupL strain of Anabaena PCC 7120 as the parental strain (Masukawa et al., 2007). The catalytic MoeFe center binds homocitrate, which is necessary for N2 fixation, but in the absence of homocitrate gene MoeFe center binds citrate: in a Klebsiella mutant this was shown to demonstrate low N2 fixation but high H2 production activity in a N2 atmosphere (Mayer et al., 2002). In line with this result, the DhupLDnifV1 cells also demonstrated high H2 production rate and heterocyst frequency compared to the parental DhupL in N2 atmosphere (Masukawa et al., 2007). Bidirectional Hydrogenases The bidirectional hydrogenases can either produce or consume H2 according to the cellular redox environment. The enzyme functions in dark fermentation and under specific conditions in photoproduction of H2. In cyanobacteria the bidirectional hydrogenase consists of two structural moieties: the hydrogenase (encoded by hoxYH) and the diaphorase unit (encoded by hoxUFE) capable of oxidation of NAD(P)H. Over the last several years, significant progress has been achieved in the identification of the transcription factors, such as LexA and AbrB-like proteins, which are members of the complex signal cascade that directs the expression of the bidirectional hydrogenase genes (Oliveira and Lindblad, 2009). On the basis of high sequence similarity, it has been hypothesized that the diaphorase subunit of the bidirectional hydrogenase also serves as the three missing
MECHANISMS OF HYDROGEN PHOTOPRODUCTION activity subunits of cyanobacterial respiratory NDH-1 complex (Appel and Schulz, 1996). However, more recent results have not supported this hypothesis since the mutants lacking the diaphorase subunits do not show malfunction of the respiratory activity (Boison et al., 1999). Bidirectional hydrogenase has been found in all non-N2-fixing and some N2-fixing cyanobacteria. Thus, many filamentous N2-fixing cyanobacteria contain both the bidirectional and the uptake hydrogenase. However, a few species have only the uptake hydrogenase. In cyanobacteria, the bidirectional hydrogenase is constitutively expressed under both aerobic and anaerobic conditions but is active only in the dark, anoxic conditions or during the transition from dark to light (Cournac et al., 2004; Schutz et al., 2004). The biological function of bidirectional hydrogenase in filamentous cyanobacteria is not well understood (Tamagnini et al., 2007). Mutational studies with hox-defective mutants suggested that the bidirectional hydrogenase in N2-fixing cyanobacteria does not support N2 fixation (Masukawa et al., 2002). In non-N2-fixing cyanobacteria the bidirectional hydrogenase is the main H2-producing enzyme and it is thought to interact with photosynthetic pathways (Ludwig et al., 2006). However, H2 production catalyzed by bidirectional hydrogenases is only transient (less than 30 s in light) since it is quickly inhibited by increasing photosynthetic O2 evolution (Cournac et al., 2004). In line with this, no transient H2 evolution was detected in different Hox deletion mutants studied by Aubert-Jousset et al. (2011). It is hypothesized that the bidirectional hydrogenase functions as a safety electron sink thereby removing excess reducing equivalents during the dark, anaerobic to light transition in unicellular Synechocystis cells (Appel et al., 2000; McIntosh et al., 2011). This hypothesis is interesting due to the natural environment of cyanobacteria being highly dynamic, with rapid fluctuations in light intensity. Such fluctuations might strongly unbalance the function of the photosynthetic complexes, resulting in production of reactive oxygen species and destroying photosynthetic apparatus. Cyanobacteria have unique flavodiiron proteins, Flv1 and Flv3, functioning as a strong electron sink at the end of light reactions by directing excess electrons to O2 without production of reactive oxygen species, thus maintaining the redox balance of the electron transport chain (Helman et al., 2003; Allahverdiyeva et al., 2011, 2012). A bidirectional hydrogenase possibly takes over the role of a strong electron sink upon dark to light transitions during anaerobiosis, the condition created in cyanobacterial mats and blooms, and where the Flv1 and Flv3 pathway is not functional (Gutthann et al., 2007). In Synechocystis, both NADPH and NADH can act as electron donors for the bidirectional hydrogenase. Recent studies showed that NADH is a preferential substrate of the diaphorase moiety, whereas NADPH is an 371 efficient activator of the bidirectional hydrogenase (Aubert-Jousset et al., 2011). These results are in line with the observed dynamics of H2 production during darkelight transition. In the dark anaerobic conditions, H2 is produced by oxidation of NADH, the major product of glycolysis assimilation. Sudden exposure to light produces NADPH by photosynthetic electron transfer chain, which functions as an activator of the hydrogenase and begins consumption of H2. Although there is no strong evidence for direct electron donation from reduced Fd (E0 ¼ 0.42 V), which is a stronger reductant than NADH (E0 ¼ 0.315 V), to the bidirectional hydrogenase, such an electron transfer cannot be completely excluded (McNeely et al., 2011). Direct linkage of Fd to the bidirectional enzyme in mutants lacking the diaphorase domain could be employed to improve cyanobacterial H2 production. Accounting for the high affinity of the bidirectional hydrogenase to H2, it has been suggested that the enzyme can function in utilization of H2 under physiological conditions. However, it should be kept in mind that the bidirectional hydrogenase reversibly evolves H2 under dark, anaerobic conditions as a result of fermentation of photosynthetically stored carbon intermediates in cyanobacteria. Recent electrochemical investigations of the bidirectional enzyme from Synechocystis PCC 6803 have revealed unexpected properties. The rate of H2 production at low pH and low H2 pressure was shown to be about 1.4 times faster than the rate of H2 consumption at high pH and high H2 pressure (McIntosh et al., 2011). Green Algal [FeeFe]-Hydrogenases Hydrogen photoproduction in green algae is catalyzed by [FeeFe]-hydrogenases. Earlier reports have suggested the existence of [NieFe]-hydrogenase in S. obliquus (Zinn et al., 1994), but presently S. obliquus is considered as having only the [FeeFe]-hydrogenase (Wunschiers et al., 2001; Florin et al., 2001). Some green algal species do not show hydrogenase activity at all (Brand et al., 1989; Boichenko and Hoffmann, 1994). Currently, the presence of genes encoding [FeeFe]-hydrogenases has been proved in the following species: Shlamydomonas reinhardtii (Happe and Kaminski, 2002; Forestier et al., 2003), Chlorella fusca (Winkler et al., 2002), Shlamydomonas noctigama (Skjanes et al., 2010), Volvox carteri (Prochnik et al., 2010), Tetraselmis subcordiformis (Yan et al., 2011) and Chlorella variabilis (Meuser et al., 2011). Algal [FeeFe]hydrogenases in vivo interact with Fd, a terminal acceptor of photosynthetic electron transport chain (Chang et al., 2007). In contrast to [NieFe]-hydrogenase enzymes, [FeeFe]-hydrogenases have significantly higher turnover rate (6000e9000/s) and usually catalyze H2 production instead of H2 uptake (Frey, 2002). However, the possible role of these enzymes in H2 uptake under high H2 partial
372 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION pressure has also been suggested (Kosourov et al., 2012). Unfortunately, the [FeeFe]-hydrogenases are extremely sensitive to O2 that irreversibly inactivates purified enzymes within seconds (Ghirardi et al., 1997). The most studied green alga, C. reinhardtii has two monomeric [FeeFe]-hydrogenases: HydA1 and HydA2 with a molecular mass of around 48 kD (Happe and Kaminski, 2002; Forestier et al., 2003). Both proteins are nuclear encoded and contain putative transit sequences that target them to the chloroplast. The hydA1 gene shows 74% similarity to hydA2 and encodes protein that is 68% identical to HydA2. Two homologous hydrogenases are typically observed in almost all green algae showing hydrogenase activity (Winkler et al., 2004). Nevertheless, some species have three [FeeFe]hydrogenase enzymes (Skjanes et al., 2010). The physiological basis for the presence of two and more hydrogenases in green algae has not been determined. In C. reinhardtii cells, HydA1 most probably participates in the light-dependent H2 production pathway (Happe and Naber, 1993). Examination of relative enzyme activities by gene-silencing techniques indicate that HydA1 catalyzes the majority of the hydrogenase activity, but the role of HydA2 in algal H2 production has not been clearly resolved (Godman et al., 2010). Recently, Meuser and et al. (2012) using the single hydA1, hydA2 and double hydA1/hydA2 knockout mutants showed that HydA2 also participates in H2 photoproduction. However, according to the authors, its contribution in the light-dependent process does not exceed 25%. The next important step in this direction should be investigation of the role of these enzymes in the H2 uptake, including the mechanisms of photoreduction (Kosourov et al., 2012). HYDROGEN PHOTOPRODUCTION BY CYANOBACTERIA Cyanobacteria have different life forms: some species are unicellular, others form colonies and filaments, or live in symbiosis with eukaryotic organisms. Accordingly, the protection of O2 sensitive enzymes from photosynhetically evolved oxygen has evolved through several different strategies. In the absence of combined nitrogen, many filamentous N2-fixing cyanobacteria physically separate oxygenic photosynthesis and N2-fixing enzymes by differentiating specialized heterocyst cells, which are regularly spaced among vegetative cells. Mature heterocysts are unique cells providing a microaerobic environment suitable for the enzymes involved in N2 fixation. The microaerobic environment inside of heterocysts is maintained by an elevated rate of respiration, lack of active PSII complexes resulting in an absence of photosynthetic O2 evolution, and a thick cell wall (Wolk et al., 1994). These two cell types, the vegetative cells and heterocysts, depend on each other. During diazotrophic growth the vegetative cells perform photosynthetic CO2 fixation and provide the heterocysts with organic carbon intermediates, like sucrose (Lopez-Igual et al., 2010), whereas heterocysts provide vegetative cells with fixed nitrogen required for cell growth (Figure 21.2). Since H2 production in heterocysts depends on carbohydrates produced in vegetative cells, this process of H2 production has been classified as indirect water biophotolysis. Heterocyst differentiation is tightly regulated by NtcA, a global transcription factor of carbon and nitrogen metabolism (Zhao et al., 2010). HetR is another essential protein specifically involved in the initial steps of heterocyst development. The patterned differentiation of heterocysts is controlled by the ratio of activator, HetR, and suppressor molecules, peptides derived from PetS and HetN (Muro-Pastor and Hess, 2012; Risser and Callahan, 2009). For the heterocystous strain, Anabaena PCC 7120 the frequency of heterocysts is approximately 10% under optimal laboratory growth conditions. Such a low frequency of heterocysts in the filament might result in modest yields of net H2 production. Thus, one possible strategy to improve H2 production is to increase the number of heterocysts in the filaments. However, although the overexpression of HetR resulted in an overall enhancement of heterocyst frequency up to 29% in FIGURE 21.2 Simplified schematic view of spatial separation of oxygenic photosynthesis in the vegetative cells and N2 fixation/H2 production in heterocysts. (For color version of this figure, the reader is referred to the online version of this book.)
HYDROGEN PHOTOPRODUCTION BY CYANOBACTERIA Anabaena PCC 7120 mutant, no increase in the nitrogenase activity of the filaments took place (Buikema and Haselkorn, 2001). It is possible that the relative decrease in the number of vegetative cells makes them incapable of producing enough reducing power to be transferred to heterocysts for enhanced H2 production. Unicellular and filamentous nonheterocystous N2-fixing cyanobacteria apply mostly temporal separation mechanism, by performing photosynthesis during the daytime and N2 fixation at night (Compaore and Stal, 2010). The energy generated by photosynthesis is stored in glycogen granules, which are later subjected to oxidative breakdown. Trichodesmium are unique cyanobacteria, because these filamentous nonheterocystous cyanobacteria are able to fix N2 simultaneously with oxygenic photosynthesis during the photoperiod (Berman-Frank et al., 2001). Nitrogenase is localized in subsets of cells in each trichome, which also contain photosynthetic complexes. During hours of high N2 fixation the cells can turn photosynthetic activity down within 10 min, which is observed as unequally distributed inactive zones in whole filaments. Importantly, the PSII activity was shown to be essential for N2 fixation in Trichodesmium (Berman-Frank et al., 2001). According to the authors, Trichodesmium utilizes photosynthetic electron transport to support N2 fixation and concomitantly enhances the Mehler reaction, which efficiently eliminates the evolved O2. Recently published complete genome sequence of Trichodesmium erythraeum (http://www.ncbi.nlm.nih.gov) shows that the strain indeed possesses the genes encoding the Flv1 and Flv3 proteins, which are involved in the “Mehlerlike” reaction in cyanobacteria. Thus, the nitrogenase enzyme in this organism is protected from O2 by a combined and modulated temporal and spatial segregation of N2 fixation and oxygenic photosynthesis within individual cells (Berman-Frank et al., 2001). The N2-fixing, unicellular cyanobacteria Cyanothece has recently attracted lots of research interest as a highly efficient H2 producer under natural aerobic conditions. Cyanothece sp. ATCC 51142 is the best hydrogen producer among the known wild-type cyanobacterial strains (Bandyopadhyay et al., 2010). Also, the ability to grow phototrophically, mixotrophically, and heterotrophically makes this strain an attractive organism for biotechnology. Cyanothece demonstrates temporal separation of oxygenic photosynthesis and N2 fixation by performing photosynthesis in daytime and N2 fixation at night. Moreover, these alternating processes are regulated by an intrinsic circadian rhythm. The genome sequence reveals the presence of the bidirectional [NieFe]-hydrogenase, uptake hydrogenase, and the conventional MoeFe nitrogenase. Diazotrophically grown Cyanothece cells entrained in 12-h light/12-h dark cycles exhibit a light-induced H2 production 373 (specific rate >150e300 mmol H2 mg/Chl h) under aerobic conditions during “subject dark” (Bandyopadhyay et al., 2010). Interestingly, the robust circadian rhythm of Cyanothece allows cells to fix N2 and produce H2 at reasonably high rates even when grown under continuous light (Min and Sherman, 2010). In the presence of combined nitrogen, Cyanothece produces H2 at very low rates, 2e10 mmol H2 mg/Chl h. This H2 production is catalyzed by the bidirectional hydrogenase and is dependent on PSII activity. In diazotrophically grown cultures, the production of H2 is driven by the nitrogenase enzyme and the activity of the enzyme is linked to PSI and respiratory electron flow (Min and Sherman, 2010). Moreover, the rates of H2 production in Cyanothece 51142 could be greatly enhanced when cells were grown in the presence of additional carbon sources, as observed in cultures supplemented with high concentrations of CO2 or glycerol (Bandyopadhyay et al., 2010). Photoproduction of H2 can be significantly enhanced by increasing reductant availability via dark anaerobic preincubation. This indicates the tight coupling of H2 photoproduction to the dark, anaerobic metabolism (Skizim et al., 2012). Recently, it was reported that Cyanothece can coproduce H2 and O2 over 100 h under continuous illumination and uninterrupted photosynthetic electron transport (Melnicki et al., 2012). Of course, Cyanothece has a very flexible metabolism and the existence of intracellular O2 gradient within the cells cannot be excluded. Despite many interesting papers describing the H2 production in Cyanothece, the molecular mechanisms behind the regulation of the nitrogenase and protection against oxygenic photosynthesis are still under debate. Strategies to Improve H2 Production in Cyanobacteria Genetic Modifications Most important genetic modifications applied to cyanobacteria in order to improve their H2 production capacity have already been discussed above. Therefore, just a few recent and important findings, will be mentioned below. Introducing Foreign Enzymes and Semiartificial Systems Most attempts of heterologous expression of genetically modified hydrogenase in cyanobacteria have not been successful due to the complexity of transcriptional regulation and maturation of the hydrogenases. A successful heterologous expression of Fd-dependent hydrogenase from Clostridium in the Synechococcus PCC7942 was demonstrated and resulted in about threefold higher H2 production activity compared to the wild type (Asada et al., 2000). More recent studies related to
374 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION the identification and characterization of the genes involved in maturation and regulation of the [FeeFe]and [NieFe]-hydrogenases and nitrogenases (Rubio and Ludden, 2005) as well as the development of heterologously expressed enzyme systems have opened new opportunities. In an attempt to engineer an organism, which can produce H2 even under aerobic conditions, the [FeeFe]-hydrogenase operon from Shewanella oneidensis MR-1 containing all maturation genes was successfully expressed in the heterocysts of Anabaena PCC 7120 (Gartner et al., 2012). Active [FeeFe]-hydrogenase was detected in aerobically grown Anabaena PCC 7120 under diazotropic growth. Despite significantly higher turnover number of the [FeeFe]-hydrogenase, in situ H2 production rate was only about 20% of that by nitrogenases (Masukawa et al., 2010). One of the reasons for such low activity might be that Fd in Anabaena is not an effective electron donor to the [FeeFe]-hydrogenase. Some hydrogenases, like the [NieFe]-hydrogenase of Ralstonia eutropha are able to perform H2 cycling in the presence of ambient O2. Employing this interesting property of the Ralstonia hydrogenase, a [NieFe]hydrogenase (Hox) from Ralstonia has been fused with the extrinsic PsaE subunit, which is located at the acceptor site of PS I, by genetic engineering. The resulted Hox/PsaE fusion exhibited in vitro self-assembly with a cyanobacterial PSI lacking the PsaE subunit and lightdriven H2 was evolved (0.58 mmol H2 mg/Chl h, Ihara et al., 2006a). In another study, Cytochrome c3 (Cyt c3) from Desulfovibrio vulgaris was chemically cross-linked to PsaE protein and the Cyt c3/PsaE complex was rebound to a PsaE-free PSI complex and introduced to a solution containing the D. vulgaris [NieFe]hydrogenase enzyme (Ihara et al., 2006b). When illuminated, light-induced H2 generation was observed at a maximum rate of 0.30 mmol H2 mg/Chl h in the presence of Fd and FNR, physiological acceptors of PSI. These results suggest that the Cyt c3/PSI complex may produce H2 in vivo. However, the observed in vitro H2 production rate was quite low, most probably due to poor electron transfer coupling between PSI and the hydrogenase in solution. A further coupling approach has been applied where the FB cluster of PSI and catalytic nanoparticle surface have been covalently bound via a “molecular wire”. Upon illumination, this semiartificial system generated up to 70 mol H2 PSI/mol min (Grimme et al., 2008, 2009). In this innovative system, covalent binding between FB (an electron cofactor of PSI) and a “molecular wire” (catalytic nanoparticle surface) enables high rates of H2 evolution. Finally, the direct coupling of photosynthesis and H2 production has also been performed on a gold surface (Krassen et al., 2009). For stable connection of the [NieFe]-hydrogenase to PSI, the extrinsic PsaE subunit was fused to the electron-transferring subunit of the membrane-bound [NieFe]-hydrogenase by genetic engineering. The resulting Hox/PsaE protein was purified and incubated with isolated PSI from Synechocystis sp. PCC 6803 lacking the PsaE subunit (PSIDPsaE). PSIDPsaE and HoxPsaE were assembled on a gold surface and electrons provided by the gold electrode were transferred to PSI with the aid of the soluble electron carrier N-methylphenazonium methyl sulfate. Upon light illumination the hydrogenase-PSI hybrid system demonstrated H2 production a rate of 4500 mol H2/min mol (Krassen et al., 2009). Elimination of Competing Electron Transfer Routes For surviving under different environmental conditions, cyanobacteria have developed a number of different alternative electron transfer pathways, which at the same time decrease the photosynthetic and H2 production efficiency. In order to improve H2 photoproduction from cyanobacteria, the competing alternative electron transport pathways should be diminished. This will guide future genetic and metabolic engineering efforts to modulate the major energetic pathways to avoid “wasteful” electron flow and to channel major electron flux to H2 production. Recent studies have confirmed that the alternative electron-transport routes in cyanobacteria are also strong competitors for H2 production. Under Ci-deprivation conditions the Flv1/Flv3 proteins in cooperation with photorespiratory pathway might flux up to 60% of electrons to O2, functioning as a powerful electron sink for the electrons originated from watersplitting PSII (Allahverdiyeva et al., 2011). In line with this, recent studies on deletion mutants of the respiratory electron transport complexes, terminal oxidases and the NdhB subunit of the NDH-1 complex have revealed increased hydrogenase activity and the production of H2 (Cournac et al., 2004; Gutthann et al., 2007). Biodiversity Cyanobacteria are very diverse organisms. For many decades only a limited number of cyanobacterial strains have been used as model laboratory organisms for H2 research and by the end of 2000 only a few attempts to screen large cyanobacterial culture collections for H2 production were recorded (Berchtold and Bachofen, 1979; Lambert and Smith, 1977). Recently, more emphasis has been given to the biodiversity of cyanobacteria in order to identify promising H2 producers with flexible metabolic pathways (Yeager et al., 2011; Yoshino et al., 2007; Allahverdiyeva et al., 2010). Screening of 400 cyanobacteria strains from the University of Helsinki Culture Collection revealed that about 50% of these strains produced easily detectable amounts of H2. Ten of them produced similar or up to
HYDROGEN PHOTOPRODUCTION BY GREEN ALGAE four times as much of H2 as the uptake hydrogenase mutants of Anabaena PCC 7120 (Masukawa et al., 2002) and N. punctiforme ATCC 29133 (Lindberg et al., 2002), specifically engineered to produce higher amounts of H2 (Allahverdiyeva et al., 2010). All 10 best H2 producers were N2-fixing, heterocystous filamentous strains. Notably, the changes in environmental parameters had differential effects on H2 production, depending on the strain. Therefore, it is necessary to test multiple environmental conditions when screening for superior H2producing strains (Yeager et al., 2011). Optimization of culture conditions and genetic modification of new strains would enhance further H2 production. Immobilization In general, cyanobacterial H2 production is difficult to sustain for long time periods in liquid cultures in photobioreactors. Utilization of suspension cultures in a two-stage system for H2 production is an even more complicated and energy-consuming process due to the centrifugation or sedimentation steps required for cell harvesting, media changes, and dilutions of cell density during the switch between the different phases. Suspension cultures require intensive mixing, which in turn causes damage to the fragile cyanobacteria filaments. This system is hard to scale up. Use of immobilized cyanobacterial cells in specially designed laboratoryscale photobioreactors would be a good solution to the above-mentioned problems of liquid cultures. Immobilization of biomolecules and whole cells on various substrates and into different gels, such as solid surfaces like porous glass, supported films, (nano) porous materials, (nano)fibers, foams, inorganic and organic hydrogels, latex, nanotubes, and nanoparticles, has been studied extensively (see for review Meunier et al., 2011). Application of immobilization usually improves the stability of the enzymes, and increases light-utilization efficiency. Immobilization of algal cells on a solid phase made of glass has been used for extended H2 production (Laurinavichene et al., 2006). A green alga, C. reinhardtii entrapped in thin alginate films demonstrated extended H2 photoproduction due to increased light-utilization efficiency and better tolerance against O2 (Kosourov and Seibert, 2009). Several attempts also have been made to immobilize cyanobacterial cells in order to improve H2 production. These include Plectonema boryanum within alginate beads (Sarkar et al., 1992), Oscillatoria in agar matrix (Phlips and Matsui, 1986) Phormidium valderianum together with Halobacterium halobium and E. coli within polyvinyl alcohol-alginate beads (Bagai and Madamwar, 1998). Recent immobilization of the Calothrix 336/3 strain in thin alginate film resulted in extended production of H2 even after 40 days of immobilization (Leino et al., 2012). Immobilization has also been found to have a 375 positive effect also on viability of cells. Twelve weeks after initial immobilization, entrapped cells from recovered films produced H2 nearly as efficiently as the fresh cells in newly made films. Moreover, the immobilized cells of Calothrix 336/3, Anabaena PCC 7120 and DhupL mutant of Anabaena were viable for over 10 months in the initial nutrient medium without addition of CO2. The demonstrated long-term viability of entrapped cells is a very important issue for economical use of cyanobacterial in H2 production systems. HYDROGEN PHOTOPRODUCTION BY GREEN ALGAE Light-Dependent Hydrogen Production Pathways Like some cyanobacteria, many species of eukaryotic green algae are capable of direct water biophotolysis (Boichenko and Hoffmann, 1994). In green algae, water biophotolysis typically proceeds in two steps: H2 O þ 2Fdox / 2Hþ þ 1=2O2 þ 2Fdred 2Hþ þ 2Fdred / H2 þ 2Fdox (21.5) (21.6) The first reaction is common to all oxygenic phototrophs, and was explained above (Section Oxygenic Photosynthesis). The second step, however, occurs only under anaerobic or microaerobic conditions, since it is extremely oxygen sensitive (Ghirardi et al., 1997). The reaction is catalyzed by the [FeeFe]-hydrogenase enzyme that accepts electrons from photosynthetically reduced Fd and reduces protons of water to molecular hydrogen (Happe and Naber, 1993; Happe and Kaminski, 2002; Foriestier et al., 2003). Although six different Fds are known in green algae, it is most likely that only PetF serves as the physiological electron donor to [FeeFe]-hydrogenase and, thus, links the photosynthetic electron-transport chain to the hydrogenase-driven reaction in vivo (Happe and Naber, 1993; Winkler et al., 2009; Long et al., 2008). In photosynthetically active algal cells, the direct water biophotolysis process usually occurs when cultures are exposed to the light after a period of dark, anaerobic adaptation (Gaffron and Rubin, 1942; Greenbaum, 1982; Appel and Schulz, 1998). It proceeds at very high initial rates (up to 300 mmol H2 mg/Chl h) that are comparable to the rates of O2 evolution in green algae under optimal light conditions (Boichenko and Hoffmann, 1994; Boichenko et al., 2004). It should be noted, however, that such high rates of H2 photoproduction in healthy algal cells occur only for a very short period of time and that the duration of the process
376 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION depends directly on the light intensity. For example, in dark-adapted Chlorella vulgaris cultures the H2 photoproduction rate reaches the maximum in 2.5 s after exposing cells to about 0.03 W/m2 light and the kinetics is linear for at least 1 min (Boichenko et al., 1983). Under 2 W/m2 illumination, cells show the maximum rate after 0.6 s, which starts declining soon after 1 s. The process can also be extended in the presence of 3-(30 ,40 dichlorophenyl)-1,1-dimethylurea (DCMU), a specific inhibitor of electron transport from PSII to the PQ-pool (Happe and Naber, 1993; Florin et al., 2001). In this case, however, H2 photoproduction is driven through the photofermentation pathway (see below) and thus depends on the level of stored carbohydrates and proteins. Nevertheless, the extension of the H2 photoproduction period in DCMU-treated cultures indicates that inhibition of H2 evolution in dark-adapted and DCMU-untreated algae in light is mainly due to a fast accumulation of O2 inside algal chloroplasts. This inhibits hydrogenase-driven reaction and switches the physiological state of cells from anaerobic to aerobic. Indeed, using the Clark-type O2 and H2 sensors, Boichenko and et al. (1983) showed that the decrease in the rate of H2 photoproduction in algae is always followed by accumulation of O2 in the culture. The sensitivity to O2 occurs at four levels: (1) gene transcription, (2) [FeeFe]-hydrogenase maturation, (3) activity of the hydrogenase catalytic site, and (4) competition for photosynthetic reductants (Ghirardi, 2006). Although H2 photoproduction in green algae is a short-term phenomenon, its theoretical sunlight to hydrogen conversion efficiency (STHE) is higher than in N2-fixing cyanobacteria. In addition, algae split water and, contrasting many anoxygenic photosynthetic bacteria, they do not require any organic substrates for production of H2 gas. The maximum efficiency for the direct water biophotolysis process has been estimated at around 10% (Bolton, 1996; Akkerman et al., 2002). Since this value is comparable to the power conversion efficiency of present commercial silicon solar cell modules, which are rated at around 10e11% with watersplitting process considered (Blankenship et al., 2011), the high efficiency of algal H2 photoproduction raises the possibility of industrial application of the process in the future. At the current state, however, the direct water biophotolysis systems are not cost-effective and will require significant biochemical and engineering improvements to achieve commercial viability. Taking into account a high sensitivity of algal hydrogenases to O2, much attention in the past was focused on the development of the efficient methods of keeping cultures anaerobic throughout the H2 production period. These methods included the addition of O2 scavengers (Healey, 1970; Randt and Senger, 1985), the use of added reductants (Randt and Senger, 1985), sparging cultures with inert gases (e.g. argon or helium) (Greenbaum, 1982; Greenbaum et al., 2001) and the addition of PSII inhibitors (Gfeller and Gibbs, 1984; Fouchard et al., 2005). Although some of these methods indeed prolong H2 production in algae, none of them resulted in bulk production of H2 gas. The expense of such approach also limits their application on larger scales. Many early research efforts also concentrated on screening for naturally better H2 producers. As a result, a few dozen algal species were tested for their ability to photoproduce H2 gas after the period of dark anaerobic adaptation. Many of them, but not all, were capable of direct water biophotolysis only for a very short period of time (Ben-Amotz et al., 1975; Boichenko and Hoffmann, 1994). According to Eqns (21.5) and (21.6), direct water biophotolysis gives a maximum theoretical H2 to O2 (mol:mol) ratio of 2 to 1. Since both PSII and PSI are involved in the process, the minimum number of quanta required for generating one H2 molecule is equal to four. In many short-term experiments performed with darkadapted algae, this value was above 4, but in some cases only 2e2.5 quanta were required (Greenbaum, 1988; Boichenko et al., 1989, 2004). The latter value indicates that H2 photoproduction in green algae may occur through a mechanism independent of water oxidation. The existence of this pathway was also confirmed by the experiments with inhibitors of the photosynthetic electron transport chain. It was found that DCMU, a PSII inhibitor (see above), does not completely block H2 photoproduction in algal cells, while 2,5-dibromo3-methyl-6-iso-propyl-p-benzoquinone, which blocks PQ oxidation by the Cyt b6 f complex, inhibits the process almost completely (Ben-Amotz et al., 1975; Kosourov et al., 2003; Antal et al., 2009). In contrast to water biophotolysis, this pathway depends on the metabolic oxidation of organic substrates that are coupled to PSI and the [FeeFe]-hydrogenases through the PQ-pool (Stuart and Gaffron, 1972; Gfeller and Gibbs, 1984; Gibbs et al., 1986). According to Gibbs and coworkers, starch, acetate and proteins could be the main substrates for photofermentation in algae, which results in production of CO2 and H2 gases. For example, the full degradation of 1 mol (in glucose equivalents) starch will give up to 6 mol CO2 and up to 12 mol H2: C6 H12 O6 þ 6H2 O / 6CO2 þ 12H2 (21.7) In C. reinhardtii and other green algae, degradation of substrates through photofermentation is not complete. The major by-products are formate, acetate, ethanol, and, in some rare cases, lactate and glycerol (Gfeller and Gibbs, 1984; Kreuzberg, 1984; Zhang et al., 2002; Kosourov et al., 2003). According to a number of studies, the final composition varies significantly on the strain and the environmental condition. Since accumulation of
HYDROGEN PHOTOPRODUCTION BY GREEN ALGAE organic substrates, mainly starch, also depends on PSII activity, the photofermentation pathway in green algae proceeds in two stages and, thus, can be considered as an indirect water biophotolysis. Starch accumulated through photosynthetic activity later undergoes degradation through glycolysis that yields pyruvate, ATP and reductants. In green algae, oxidation of reductants may occur via Nda2, a monomeric class-II type NAD(P) dehydrogenase, which feeds electrons into the photosynthetic electron transport chain at the point of the PQ-pool (Jans et al., 2008; Desplats et al., 2008). From the PQ-pool, electrons follow to [FeeFe]-hydrogenase through PSI and Fd. Since both direct and indirect water biophotolysis pathways are linked to [FeeFe]hydrogenase via PSI, it becomes clear that in algae they operate simultaneously most of the time. Nevertheless, their contribution into the overall H2 photoproduction process varies depending on the organism (Meuser et al., 2009) and the physiological state of algal cells (Laurinavichene et al., 2004). In the indirect water biophotolysis process, O2 evolution and H2 production stages can be separated from each other either temporally or spatially, thus circumventing the apparently inherent O2 sensitivity of H2 photoproduction. Despite lower efficiency (as compared to the direct water splitting), separation of the process into two distinct stages gives a significant advantage to the biotechnological applications (Benemann, 1994, 1996). According to the early Benemann’s concept, during the first aerobic stage algal cultures accumulate biomass enriched in carbohydrates as a result of photosynthetic CO2 fixation. During the second, anaerobic stage carbohydrates and other materials stored in biomass are processed to H2 gas. The two stages are separated either physically in two different photobioreactors or temporally through additional dark adaptation and fermentation periods. In this approach, O2-evolving activity inside the cells is totally inactivated during the H2 production stage without application of any inhibitors, and algae evolve H2 gas through the photofermentation pathway linked to PSI. The second stage can also be driven in the dark by fermentative bacteria. Thus, the discovery of the indirect water biophotolysis pathway was the first step toward development of the protocol for long-term H2 photoproduction in green algae. Role of H2 Photoproduction in Green Algae Despite extensive research, the role of H2 photoproduction in physiology of green algae still remains unclear. Some genera of green algae do not show hydrogenase activity and so do not produce H2 even under dark anaerobic conditions (Brand et al., 1989; Winkler et al., 2002; Boichenko et al., 2004). Some strains demonstrate a very low rate of H2 production in light, even if they have a 377 high hydrogenase activity (Boichenko and Hoffmann, 1994; Guan et al., 2004; Skjanes et al., 2008). In algae showing active H2 photoproduction, hydrogenase has been proposed as a regulatory valve preventing overreduction of photosynthetic apparatus during transition from dark anaerobic to light aerobic conditions (Appel and Schulz, 1998). Under anaerobic conditions, CO2 fixation in algal cells is repressed due to inactivation of Rubisco, the key enzyme of the CalvineBenson cycle. Reactivation of Rubisco upon illumination is a very slow process, which takes up to 2 min (Campbell and Ogren, 1990). During this time, hydrogenase may act as an alternative sink of electrons for the electron transport chain, preventing photodamage of the photosynthetic apparatus. Although very attractive, this hypothesis has never been proved experimentally. Moreover, experiments performed by Tsygankov and coauthors have demonstrated results contradicting the hypothesis, as cell viability between the mutant lacking the hydrogenase enzyme and the parental strain was almost the same after the period favorable to H2 photoproduction (Tsygankov, 2012). Long-Term H2 Production by Green Algae As mentioned above, all metabolic pathways leading to H2 production in green algae are extremely sensitive to O2 due to fast and irreversible inhibition of [FeeFe]hydrogenase enzyme(s) and competition from different respiratory processes for the reductants. Significant efforts to surmount the O2 sensitivity issue have been made, but still the algal strain with the O2-tolerant H2 photoproduction has not been generated (Ghirardi, 2006; Ghirardi and Mohanty, 2010). Therefore, in photosynthetically active algal cells under saturating light conditions the most efficient H2 photoproduction process (via direct biophotolysis) lasts for a few seconds. For the industrial system, however, H2 photoproduction should be extended from the scale of seconds to at least days. When optimized, such process may yield H2 at a cost of around $3/kg H2 for the upper bound performance (w9% STHE) and slightly above $8/kg H2 for the near term performance (w1.5e2% STHE) (Blake et al., 2008; James et al., 2009). Unfortunately, at the current state sustained H2 photoproduction in photosynthetically active green algae is only possible at the expense of efficiency. For example, long-term H2 production is usually observed in cultures under very low light intensities or even in the dark when O2 evolution proceeds very slowly or does not proceed at all (Kondratieva and Gogotov, 1983; Aparicio et al., 1985). Interestingly, Batyrova et al. (2012) recently observed a stable but negligible rate of H2 production in very dense cultures placed under normal light conditions. The algal cells demonstrated a high hydrogenase activity, but
378 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION efficient H2 photoproduction was not observed. Most probably, under these conditions H2 evolution was driven by the cells in the inner part of the photobioreactor, while H2 photoproduction in the illuminated algae was limited by the coevolved O2. One of the possibilities of driving the process under normal photosynthetic conditions is to sparge cultures continuously with inert gas, such as argon or nitrogen, that removes rapidly the PSII-evolved O2 gas (Greenbaum, 1982; Greenbaum et al., 2001). Before the sulfur-deprivation protocol was developed, this was the only way for the long-term H2 generation in algal cultures. Using a confined bioreactor, Greenbaum et al. (2001) showed several cycles of simultaneous H2 and O2 photoproduction during 1 h intervals after extensive purging the cultures in the dark for 2 h by N2. The experiment lasted for over 1400 h (58 days) and required periodic additions of CO2 gas into the photobioreactor for restoration of photosynthetic activity and replenishment of carbohydrates in algal cells. The average stoichiometric ratio of H2 to O2 was 2.8, indicating that reducing equivalents for H2 were derived from endogenous reductants, most likely starch, as well as water. Therefore, H2 photoproduction in this case was partly driven by the cells via indirect biophotolysis pathway. Nevertheless, even under extensive purging of the cultures with N2 and good mixing conditions in the photobioreactor, the H2 evolution rate was limited due to O2 buildup in the liquid phase (Greenbaum et al., 2001). These experiments were later repeated under sulfurdeprived conditions (but with some modifications) and demonstrated a simultaneous improvement in the rate of H2 photoproduction upon declining the O2-evolving activity in algal cultures (Ghirardi et al., 2000). The prolonged H2 evolution by green algae can also be induced by a full or partial inhibition of the watersplitting activity of PSII in cells. The full inhibition can be achieved by applying DCMU (Gfeller and Gibbs, 1984; Fouchard et al., 2005). In contrast to the N2-sparging approach, H2 photoproduction in DCMU-treated cultures depends totally on the amount of carbohydrates or other substrates stored by the cells during the growth period and, thus, the process is limited to only one cycle. The partial inhibition of the PSII activity occurs in sulfur-deprived algae (Melis et al., 2000) and in certain mutants with manipulated expression levels of the D1 reaction center protein (Surzycki et al., 2007) or in the mutant cells affected in the PSII water-splitting complex (Makarova et al., 2005, 2007). For the establishment of an anaerobic environment and H2 production in algal cultures, the partial inhibition of the PSII activity in cells should achieve the point when the O2 produced by the PSII centers is consumed sufficiently by respiration. However, H2 photoproduction under these conditions usually proceeds at lower rates as compared to the initial rates in dark-adapted algae exposed to the light. In contrast to the full inhibition approach, several cycles of H2 production are possible (Ghirardi et al., 2000), and thus the process can be driven continuously (Laurinavichene et al., 2006, 2008; Kim et al., 2010). Hydrogen Photoproduction by Nutrient-deprived Green Algae One of the most remarkable events in investigating H2 metabolism in green algae was the discovery of sustained H2 photoproduction in C. reinhardtii cultures under sulfur-deprived conditions (Melis et al., 2000; Ghirardi et al., 2000). In this approach, the long-term H2 photoproduction is possible due to a metabolic switch occurring in sulfur-deprived algal cells, which separate temporarily the O2-evolving, aerobic (Eqn (21.5)) and H2-producing, anaerobic (Eqn (21.6)) stages in the same culture. Sulfur deprivation causes the partial and reversible inhibition of PSII-dependent water-splitting activity in algae. As demonstrated by Wykoff et al. (1998), C. reinhardtii cells lose gradually up to 75% of the initial PSII activity within the first 24 h of sulfur starvation. The reduction of H2O-splitting activity was also shown under deprivation of other nutrients, such as nitrogen, phosphorus, Fe and Mn (Wykoff et al., 1998; Ghirardi et al., 2000; Philipps et al., 2012) but usually with a significant delay, as compared to sulfur starvation. The repression of the linear electron flow from the PSII centers under nutrient starvation is a common phenomenon not only for green algae but also for cyanobacteria (Sauer et al., 2001) and high plants (Dietz and Heilos, 1990; Ferreira and Teixeira, 1992), and is a good example of how photosynthetic organisms adjust the rate of photosynthesis to the stress conditions. Continuous nutrient starvation reduces the capacity for de novo protein biosynthesis and CO2 fixation, and, as a result, decreases the demand of the cells in the photosynthetic reductants (Grossman, 2000). Under these conditions, the repression of the O2-evolving activity and linear electron flow protects the photosynthetic apparatus from overreduction, generation of reactive oxygen species and photoinhibition. Numerous experiments showed that the inhibition of O2-evolving activity in nutrient-deprived cells is mostly caused by the loss of PSII centers (Kolber et al., 1988; Wykoff et al., 1998). In the absence of basic nutrients such as nitrogen, phosphorus or sulfur, the cells cannot efficiently resynthesize D1 protein, the key component of the PSII complex, and the PSII repair cycle is blocked (Melis and Chen, 2005). Nutrient deprivation, however, has little effect on cellular respiration, especially in the first few days (Melis et al., 2000). As a result, the rate of photosynthetic
HYDROGEN PHOTOPRODUCTION BY GREEN ALGAE O2 evolution falls below the rate of respiratory O2 uptake and algal cultures, if sealed in photobioreactors with a little headspace volume, become anaerobic in the light (Melis et al., 2000). In sulfur-deprived cultures, this usually happens within the first 24 h. The establishment of anaerobiosis in the sealed photobioreactor induces the expression of [FeeFe]-hydrogenase enzymes in algal cells (Happe and Kaminski, 2002; Forestier et al., 2003). [Fe-Fe]-hydrogenase accepts electrons from the photosynthetic electron-transport chain and algae start producing H2 in the light. If not optimized, H2 photoproduction lasts for several days (Melis et al., 2000; Ghirardi et al., 2000). Under continuous flow of the medium containing sulfur in a micromolar range, algae produce H2 gas for several months, although at substantially low rates (Fedorov et al., 2005; Laurinavichene et al., 2006). The most interesting results obtained from sulfur-deprivation experiments are summarized in Table 21.1. As shown in the table, the rates and the yields of H2 photoproduction in algal cultures vary depending on the experimental conditions. In C. reinhardtii wildtype strains the rate usually does not exceed 13 mmol mg/Chl h, while some genetically modified strains are able to produce H2 with rates up to 27 mmol mg/Chl h. In green algae, sulfur deprivation demonstrates the strongest inhibitory effect on PSII (Wykoff et al., 1998; Ghirardi et al., 2000) most probably due to the lowest intracellular sulfur reserves. The later studies showed that the same principle works for phosphorusdeprived (Batyrova et al., 2012) and nitrogen-deprived (Philipps et al., 2012) microalgae. Phosphorus-depleted cultures start producing H2 gas only after the initial growth period on the phosphorus-free medium. Growing algae utilize an intracellular pool of reserved phosphorus. When they reach the point of phosphorus starvation, PSII in algal cells is inactivated in a manner similar to sulfur-starved algae. Despite a considerable delay in the establishment of anaerobic conditions, phosphorus-deprived algae produce only slightly less H2 gas than sulfur-deprived cultures under the same experimental conditions, but they also accumulate less starch reserves during the growth stage (Batyrova et al., 2012). Nitrogen-deprived algae behave in a similar way. They produce H2, but with a significant delay (Philipps et al., 2012). In contrast to phosphorusdeprived cells, the delay in nitrogen-deprived cultures seems to be caused by slower inactivation of PSII centers. These algae also accumulate significantly more starch reserves than sulfur-deprived algae, but degrade them slower. As a result, they produce considerably less H2 overall. Inability to efficiently channel electrons from carbohydrate oxidation toward the hydrogenase enzyme likely causes the degradation of the Cyt b6 f complex upon nitrogen starvation and lowers amounts of PetF. Nevertheless, nitrogen-deprived cultures may 379 have a higher potential for the light-independent H2 production pathway (Philipps et al., 2012). The vast majority of experiments completed on H2 production by nutrient-deprived microalgae have been undertaken so far with C. reinhardtii cultures. However, other species of green algae also produce H2 gas under this condition (Winkler et al., 2002; Skjanes et al., 2008; Meuser et al., 2009). Successful H2 production has been demonstrated by sulfur-depriving C. noctigama and Chlamydomonas euryale (Skjanes et al., 2008). Sulfurdeprived cultures of S. obliquus, Platymonas subcordiformis, Scenedesmus vacuolatus, Chlamydomonas vectensis, Chlamydomonas pyrenoidosa, Desmodesmus subspicatus, Pseudokirchneriella subcapitata, Chlamydomonas moewusii and Lobochlamys culleus generate only minor amounts of H2 gas (Winkler et al., 2002; Guan et al., 2004; Skjanes et al., 2008; Meuser et al., 2009). Some other tested species, such as Dunaliella salina and C. vulgaris demonstrate no detectible hydrogenase activities and do not produce H2 under sulfur-deprived conditions (Cao et al., 2001; Winkler et al., 2002). H2 photoproduction in nutrient-deprived algae depends both on the residual PSII activity remaining in cells after inactivation (Antal et al., 2003; Kosourov et al., 2003) and on the catabolism of starch accumulated during the first 18e24 h of sulfur deprivation (Fouchard et al., 2005; Ghirardi et al., 2000; Kosourov et al., 2003; Tsygankov et al., 2002; Zhang et al., 2002). The contribution of these two pathways in H2 photoproduction varies depending on the stage of sulfur deprivation (Laurinavichene et al., 2004) and, most probably, on the strain used in the experiment (Chochois et al., 2009). In the wild-type C. reinhardtii CC-124 strain, starch degradation may donate up to 20% electrons to hydrogenase enzymes in the middle of the H2 production stage (Kosourov et al., 2003). Besides contribution to H2 photoproduction, the degradation of starch and other stored organic substrates fuels the respiratory consumption of O2 produced by the residual PSII activity and therefore is responsible for maintaining culture anaerobiosis and for protecting hydrogenase enzymes from O2 inactivation (Fouchard et al., 2005; Kosourov et al., 2007). The importance of efficient respiration for H2 photoproduction was further proved by inhibitory analysis (Antal et al., 2009) and in the respiratorydeficient mutants (Table 21.1). Under photoheterotrophic conditions (when acetate is the only substrate), accumulation of starch in algae in the beginning of sulfur deprivation is tightly linked to consumption of acetate from the medium. The respiration of acetate provides the cells with a substrate for CO2 fixation. It also helps with the establishment of anaerobiosis in the photobioreactor (Kosourov et al., 2007). The use of acetate in the growth medium, however, increases the expense associated with maintenance of
380 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION TABLE 21.1 The Rates and Yields of H2 Photoproduction by the Sulfur-Deprived, Wild-Type C. reinhardtii Strains and Some Mutants under Different Experimental Conditions Maximum Specific Rate of H2 Production, mmol mg/Chl h Total Yield of H2 Gas, mmol/l References e 4.7 Melis et al., 2000 12.5 23.1 Kosourov et al., 2012 e 2.3 Tsygankov et al., 2006 e 1.6 Lecler et al., 2011 1. Photoheterotrophic, 150 mmol/ m2 s PAR from two sides, 28  C, initial pH at 7.3, synchronized culture 5.9 6.6 Kosourov et al., 2002 2. The same as above, but initial pH at 7.7 and unsynchronized culture 9.4 7.7 Kosourov et al., 2003 3. The same as above, but 140 mmol/ m2 s PAR from two sides and improved culture mixing 9.8 6.9 Giannelli et al., 2009 e w22 Kruse et al., 2005 Hemschemeier et al., 2008 Strain Experimental Condition WT, 137C mtþ 1. Photoheterotrophic, 25  C, 200 mmol/m2 s PAR from two sides 2. The same as above, but 70 mmol/ m2 s PAR from one side and low H2 partial pressure, initial pH 7.3 3. Photoautotrophic, 28  C, 110 mmol/m2 s PAR during the photosynthetic stage and 20 mmol/m2 s PAR during the hydrogen production stage (from two sides), pH was stabilized at 7.4 during the first stage 4. Photoheterotrophic, 25  C, 500 mmol/m2 s PAR WT, CC-124 mt- Stm6 (Affected in the State Transition) Photoheterotrophic, 100 mmol/m2 s PAR, 25  C CC-2803 (RubiscoDeficient Strain) 1. Photoheterotrophic, 100 mmol/ m2 s PAR 3.8 e 2. The same as above, but sulfurreplete 5.4 e CC-4169 (Antennae Mutant, Affected in tla1) Photoheterotrophic, 285 mmol/m2 s PAR, 25  C immobilized in alginate films 3.8 e Kosourov et al., 2011 Respiratory-Deficient Mutants 1. Photoheterotrophic, 500 mmol/ m2 s PAR, 25  C, mutant defective in mitochondrial complex I (NADH:ubiquinone oxidoreductase) e 1.3 Lecler et al., 2011 2. The same as above, mutant defective in mitochondrial complex III (ubiquinol cytochrome c oxidoreductase) e 0.3 3. The same as above, mutant defective in both I and III complexes e 0.07 1. Photoheterotrophic, 28  C, improved culture mixing, 70 mmol/m2 s PAR from two sides 19 21 Torsillo et al., 2009 2. The same as above, but 140 mmol/ m2 s PAR from two sides 27.5 23.6 Scoma et al., 2012 L159I-N230Y (Substitution in the D1 PSII Protein) WT, wild type; PAR, photosynthetic active radiation.
HYDROGEN PHOTOPRODUCTION BY GREEN ALGAE the system and should therefore be avoided. Recently, Tsygankov et al. (2006) showed that H2 photoproduction in green algae is also possible under autotrophic conditions, when cultures are supplied with CO2 gas instead of acetate. In this experiment, authors used the microprocessor-controlled bioreactor system for a controllable addition of CO2 gas. The unique aspect of this system is that cells are provided with appropriate amounts of CO2, in accordance with the demands of the culture. Under these conditions, algae accumulate enough starch that can later be used for the establishment of anaerobiosis in the culture and for the removal of O2 during the H2 production stage (Kosourov et al., 2007). Using the special light regime, the authors generated almost the same amounts of H2 gas as in photoheterotrophic cultures (Tsygankov et al., 2006; Tolstygina et al., 2009). Strategies to Improve H2 Photoproduction in Green Algae Among major barriers to optimal H2-production yields in algal cultures are high sensitivity of algal [FeeFe]hydrogenases to O2 inactivation, low light saturation levels of photosynthesis, competition for reductant from alternative metabolic pathways, state transition and establishment of cyclic electron flow around PSI, and the reversible nature of the hydrogenasee driven reaction. All these barriers have been extensively studied by different research groups in the last few years. Clearly, the O2 sensitivity of [FeeFe]-hydrogenases is the major barrier preventing the application of green algal H2 photoproduction in commercial systems. Several approaches for solving the O2-sensitivity issue have been suggested: (1) identifying and implementing mutations, which narrow the channel(s) of the [FeeFe]-hydrogenase enzyme for blocking access of O2 molecules to the catalytic center (Cohen et al., 2005; Posewitz et al., 2009); (2) selecting for O2-tolerant enzymes through random mutagenesis (Nagy et al., 2007; Stapleton and Swartz, 2010); (3) introducing enzymes from other organisms, which are more stable to O2 inactivation, into algal cells. None of these approaches have yet resulted in a mutant with improved O2 tolerance. Nevertheless, Stapleton and Swartz (2010) applying the directed evolution approach identified a version of C. reinhardtii HydA1 with a fourfold increase in catalytic activity as compared to the wild-type enzyme. Low light-utilization efficiency in mass cultures is another important factor precluding the use of H2producing green algae in practical applications (Torzillo et al., 2003). In algal suspensions, light intensity decreases with the depth of the culture. The light attenuation is more pronounced in dense cultures, where 381 shading limits the productivity of inner parts of the culture. On the contrary, algae in the upper layers suffer from photoinhibition, which is more pronounced under high light intensities. The latter significantly limits application of high light intensities for improving the overall algae productivity. The problem can be addressed in part by immobilizing algae in thin layers or films. Immobilization fixes algal cells within a controllable volume and allows uniform light distribution to the cells that makes light utilization per volume basis more efficient. Indeed, immobilization of sulfurdeprived C. reinhardtii cultures on glass fiber matrices demonstrated significant improvements both in the volumetric rate of H2 photoproduction and in the duration of the process (Laurinavichene et al., 2006). This technique used the property of microalgae to form biofilm on the glass surface. The attachment of cells occurred through natural colonization that, if required, can be accelerated by activating glass fibers with 3-(2-aminoethyl-aminopropyl)-trimethoxysilane (Tsygankov et al., 1994). Later studies of immobilized algae with either a constant flow of medium containing micromolar sulfate concentrations or cycling of immobilized cells between minus and plus sulfate conditions improved the duration of H2 production up to at least 3 months (Laurinavichene at al., 2008). However, due to irregular colonization of glass fibers by the algal cells, the system showed significant physical and physiological heterogeneities in different parts of the matrix, resulting in irregular light and nutrient distributions, and decreasing the overall performance of H2 photoproduction. In order to improve the light absorption properties of immobilized microalgae, Kosourov and Seibert (2009) entrapped cells within thin alginate films. This technique produced films with uniform distribution of algal cells within the matrix that had very high cell densities (up to 2000 mg total Chl per ml of the matrix). As a result, the light conversion efficiency in alginate films at w60 mE/m2 s PAR (photosynthetic active radiation) achieved 1.5% for the period of the maximum H2production rate and was close to 1% for the whole period of nutrient deprivation. Another approach for improving light utilization efficiency in mass algal cultures is to find or generate algal mutants with a small chlorophyll antenna size. Strains with the truncated antennae allow greater transmittance of irradiance through the ultrahigh cell density culture without significant dissipation of light energy and, as a result, have a higher photosynthetic productivity in outdoor conditions. Recently, C. reinhardtii mutants with truncated chlorophyll antennae were generated and characterized (Polle et al., 2000, 2003). These mutants have shown promise in increasing the light utilization efficiency and the overall productivity in mass cultures (Polle et al., 2002, 2003), but suspensions have not
382 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION established anaerobiosis and so have failed to produce H2 gas under sulfur-deprived conditions. Despite this, nutrient-deprived mutants with truncated chlorophyll antennae produced H2 after immobilization within thin alginate films (Table 21.1). These mutants showed higher efficiency of H2 photoproduction than the parental CC-425 strain under saturating light conditions (Kosourov et al., 2011). H2 photoproduction in green algae competes with a number of different metabolic pathways for the reductant originated in photosynthesis (Hemschemeier and Happe, 2011). Here, CO2 fixation is one of the most important. The affinity of Fd to FNR is very high and on the order of 0.6 mM (Kurisu et al., 2005), while the affinity of Fd to hydrogenase enzyme is only about 10 mM (Roessler and Lien, 1984). It is clear that electrons in healthy algal cells will be preferably directed toward reduction of NADPþ and, hence, toward CO2 fixation. Sulfur-deprived algae, however, inactivate Rubisco, the key enzyme of CO2 fixation, by the time of the establishment of anaerobiosis in the photobioreactor (Zhang et al., 2002). Zhang and coauthors showed that only about 3% of this protein is present in cells during the H2 production stage. This finding suggests that the photosynthetically generated reductants in sulfurdeprived algae are preferably used for generation of H2, but not for CO2 fixation. It is important to note here, that according to Hemschemeier et al. (2008) the Rubisco-deficient C. reinhardtii, CC-2803 strain produces H2 gas even under sulfur-replete conditions (Table 21.1). H2 evolution in this strain is almost completely dependent on electron flow from PSII. This finding shows that flow of electrons in engineered green algae can be successfully redirected toward H2 photoproduction. The competition for the photosynthetically generated reductant from other metabolic pathways is less studied. There has been some evidence for the competition from nitrate reductase (Aparicio et al., 1985). However, the wild-type CC-124 and 137C strains of C. reinhardtii that are commonly used in sulfur-deprivation experiments (Table 21.1) carry the nit1 and nit2 mutations and cannot grow on nitrate. Therefore, the question about possible competition for the reductant between hydrogenase and nitrate reductase should be studied in detail using the Sager’s line of C. reinhardtii wild-type strains (Pröschold et al., 2005). A novel approach for preventing competition for the reductant from other metabolic pathways is formation of a fused complex of Fd and hydrogenase. In vitro analysis of such a complex showed that replacing the hydrogenase with the Fd/hydrogenase fusion switches the bias of electron transfer from FNR to hydrogenase and results in an increased rate of H2 photoproduction (Yacoby et al., 2011). This experiment indicates that the idea of the formation of a fused Fd/hydrogenase complex is promising, but should be checked in the C. reinhardtii mutant in vivo. Another barrier for the industrial H2 photoproduction system involves the redirection of photosynthetic electron flow from linear to cyclic and production of ATP, which results in a nonproductive pathway and decreased H2 production under anaerobic conditions. In green algae, this process also involves phosphorylation and dissociation of PSII-light-harvesting antenna and results in the so-called state 1 to state 2 transitions that lead to higher excitation of PSI over PSII. A promising approach for prolongation of H2 production in algae has recently been proposed by Kruse et al. (2005). They generated the mutants affected in state transition. These mutants are blocked in state 1 that inhibits cyclic electron flow around PSI. One of these mutants, stm6, accumulated larger starch reserves under sulfur deprivation and produced almost five times more H2 gas than the wild type (Table 21.1). H2 photoproduction in green algae is driven by the bidirectional [FeeFe]-hydrogenase enzyme that catalyzes not only the forward (H2 photoproduction) but also the reverse (H2 uptake) reaction. Under high H2 partial pressure in the photobioreactor, the rate of reverse reaction is significant (Kosourov et al., 2012). The authors showed that the decrease in H2 partial pressure improves significantly the yields and rates of H2 photoproduction in algal cultures. They also suggested the existence in sulfur-deprived algae of H2-uptaking pathways, either photoreduction or oxy-hydrogen reaction. The possibility of photoreduction in nutrientdeprived algae is questionable because of the significant degradation of the Rubisco enzyme by the time H2 photoproduction begins (Zhang et al., 2002). Therefore, it is most likely that nutrient-deprived algae utilize H2 gas through the indirect oxy-hydrogen reaction involving the chlororespiration pathway from [FeeFe]hydrogenase(s) to O2 through Fd, NADPþ/NADPH and the PQ-pool. If H2-uptaking pathway(s) does exist, H2 photoproduction in green algae can be further improved by downregulating this pathway(s). Acknowledgments This work was financially supported by the Academy of Finland Center of Excellence project (118637) and by the Kone foundation (YA, SNK). References Akkerman, I., Janssen, M., Rocha, J., Wijffels, R.H., 2002. Photobiological hydrogen production: photochemical efficiency and bioreactor design. Int. J. Hydrogen Energy 27, 1195e1208. Allahverdiyeva, Y., Leino, H., Saari, L., Fewer, D., Shunmugam, S., Sivonen, K., Aro, E.M., 2010. Screening for biohydrogen production by cyanobacteria isolated from the Baltic Sea and Finnish lakes. Int. J. Hydrogen Energy 35, 1117e1127.
REFERENCES Allahverdiyeva, Y., Ermakova, M., Eisenhut, M., Zhang, P., Richaud, P., Hagemann, M., Cournac, L., Aro, E.M., 2011. Interplay between flavodiiron proteins and photorespiration in Synechocystis sp. PCC 6803. J. Biol. Chem. 286, 24007e24014. Allahverdiyeva, Y., Mustila, H., Ermakova, M., Bersanini, L., Richaud, P., Ajlani, G., Battchikova, N., Cournac, L., Aro, E.M., 2012. Flavodiiron proteins Flv1 and Flv3 enable cyanobacterial growth and photosynthesis under fluctuating light. Proc. Natl. Acad. Sci. USA 110, 4111e4116. Antal, T.K., Krendeleva, T.E., Laurinavichene, T.V., Makarova, V.V., Ghirardi, M.L., Rubin, A.B., Tsygankov, A.A., Seibert, M., 2003. The dependence of algal H2 production on photosystem II and O2 consumption activities in sulfur-deprived Chlamydomonas reinhardtii cells. Biochim. Biophys. Acta 1607, 153e160. Antal, T.K., Volgusheva, A.A., Kukarskih, G.P., Krendeleva, T.E., Rubin, A.B., 2009. Relationships between H2 photoproduction and different electron transport pathways in sulfur-deprived Chlamydomonas reinhardtii. Int. J. Hydrogen Energy 34, 9087e9094. Aparicio, P.J., Azuara, M.P., Ballesteros, A., Fernandez, V.M., 1985. Effects of light intensity and oxidized nitrogen sources on hydrogen production by Chlamydomonas reinhardii. Plant Physiol. 78, 803e806. Appel, J., Schulz, R., 1996. Sequence analysis of an operon of a NAD(P)-reducing nickel hydrogenase from the cyanobacterium Synechocystis sp. PCC 6803 gives additional evidence for direct coupling of the enzyme to NAD(P)H-dehydrogenase (complex I). Biochim. Biophys. Acta 1298, 141e147. Appel, J., Schulz, R., 1998. Hydrogen metabolism in organisms with oxygenic photosynthesis: hydrogenases as important regulatory devices for a proper redox poising? J. Photochem. Photobiol. Biol. 47, 1e11. Appel, J., Phunpruch, S., Steinmuller, K., Schulz, R., 2000. The bidirectional hydrogenase of Synechocystis sp. PCC 6803 works as an electron valve during photosynthesis. Arch. Microbiol. 173, 333e338. Asada, Y., Koike, Y., Schnackenberg, J., Miyake, M., Uemura, I., Miyake, J., 2000. Heterologous expression of clostridial hydrogenase in the Cyanobacterium synechococcus PCC7942. Biochim. Biophys. Acta 1490, 269e278. Aubert-Jousset, E., Cano, M., Guedeney, G., Richaud, P., Cournac, L., 2011. Role of HoxE subunit in Synechocystis PCC 6803 hydrogenase. FEBS J. 278, 4035e4043. Bagai, R., Madamwar, D., 1998. Prolonged evolution of photohydrogen by intermittent supply of nitrogen using a combined system of Phormidium valderianum, Halobacterium halobium, and Escherichia coli. Int. J. Hydrogen Energy 23, 545e550. Bandyopadhyay, A., Stöckel, J., Min, H., Sherman, L.A., Pakrasi, H.B., 2010. High rates of photobiological H2 production by a cyanobacterium under aerobic conditions. Nat. Commun. 1, 139. Batyrova, K.A., Tsygankov, A.A., Kosourov, S.N., 2012. Sustained hydrogen photoproduction by phosphorus-deprived Chlamydomonas reinhardtii cultures. Int. J. Hydrogen Energy 37, 8834e8839. Ben-Amotz, A., Erbes, D.L., Riederer-Henderson, M.A., Peavey, D.G., Gibbs, M., 1975. H2 metabolism in photosynthetic organisms: I. dark H2 evolution and uptake by algae and mosses. Plant Physiol. 56, 72e77. Benemann, J.R., 1994. Feasibility analysis of photobiological hydrogen production. In: Block, D.L., Versiroglu, T.N. (Eds.), Hydrogen Energy Progress X, Proceedings of Tenth World Hydrogen Energy Conference, Cocoa Beach, Florida, pp. 931e940. Benemann, J.R., 1996. Hydrogen biotechnology: progress and prospects. Nat. Biotech. 14, 1101e1103. Benemann, J.R., Weare, N.M., 1974. Hydrogen evolution by nitrogenfixing Anabaena cylindrica cultures. Science 184, 174e175. Berchtold, M., Bachofen, R., 1979. Hydrogen formation by cyanobacteria cultures selected for nitrogen fixation. Arch. Microbiol. 123, 227e232. 383 Berman-Frank, I., Lundgren, P., Chen, Y.B., Küpper, H., Kolber, Z., Bergman, B., Falkowski, P., 2001. Segregation of nitrogen fixation and oxygenic photosynthesis in the marine cyanobacterium Trichodesmium. Science 294, 1534e1537. Blake, D.M., Amos, W.A., Ghirardi, M.L., Seibert, M., 2008. Chapter 5: materials requirements for photobiological hydrogen production. In: Jones, R.H., Thomas, G.J. (Eds.), Materials for the Hydrogen Economy. CRC Press, Boca Raton, FL, pp. 123e145. Blankenship, R.E., Tiede, D.M., Barber, J., Brudvig, G.W., Fleming, G., Ghirardi, M., Gunner, M.R., Junge, W., Kramer, D.M., Melis, A., Moore, T.A., Moser, C.C., Nocera, D.G., Nozik, A.J., Ort, D.R., Parson, W.W., Prince, R.C., Sayre, R.T., 2011. Comparing photosynthetic and photovoltaic efficiencies and recognizing the potential for improvement. Science 332, 805e809. Boichenko, V.A., Hoffmann, P., 1994. Photosynthetic hydrogen production in prokaryotes and eukaryotes e occurrence, mechanisms and functions. Photosynthetica. 30, 527e552. Boichenko, V.A., Arkhipov, V.N., Litvin, F.F., 1983. Simultaneous measurements of fluorescence induction and hydrogen evolution of Chlorella under anaerobic conditions. Biophys. (Russ). 28, 976e979. Boichenko, V.A., Satina, L.Y., Litvin, F.F., 1989. Efficiency of hydrogen photoproduction in algae and cyanobacteria. Fiziol. Rast. 36, 239e247. Boichenko, V.A., Greenbaum, E., Seibert, M., 2004. Hydrogen production by photosynthetic microorganisms. In: Archer, M.D., Barber, J. (Eds.), Photoconversion of Solar Energy: Molecular to Global Photosynthesis. Imperial College Press, London, pp. 397e452. Boison, G., Bothe, H., Hansel, A., Lindblad, P., 1999. Evidence against a common use of the diaphorase subunits by the bidirectional hydrogenase and by the respiratory complex I in cyanobacteria. FEMS Microbiol. Lett. 174, 159e165. Boison, G., Mergel, A., Jolkver, H., Bothe, H., 2004. Bacterial life and dinitrogen fixation at a gypsum rock. Appl. Environ. Microbiol. 70, 7070e7077. Bolton, J.R., 1996. Solar photoproduction of hydrogen. In: Report to the International Energy Agency, under agreement on the production and utilization of hydrogen. IEA/H2/TR-96. pp. 1e48. Bothe, H., Schmitz, O., Yates, M.G., Newton, W.E., 2010. Nitrogen fixation and hydrogen metabolism in cyanobacteria. Microbiol. Mol. Biol. Rev. 74, 529e554. Brand, J.J., Wright, J., Lien, S., 1989. Hydrogen production by eukaryotic algae. Biotechnol. Bioengy 33, 1482e1488. Buikema, W.J., Haselkorn, R., 2001. Expression of the Anabaena hetR gene from a copper-regulated promoter leads to heterocyst differentiation under repressing conditions. Proc. Natl. Acad. Sci. USA 98, 2729e2734. Burgess, B.K., Lowe, D.J., 1996. Mechanism of molybdenum nitrogenase. Chem. Rev. 96, 2983e3011. Campbell, W.J., Ogren, W.L., 1990. Electron transport through photosystem I stimulates light activation of ribulose bisphosphate carboxylase/oxygenase (rubisco) by rubisco activase. Plant Physiol. 94, 479e484. Camsund, D., Devine, E., Holmqvist, M., Yohanoun, P., Lindblad, P., Stensjö, K., 2011. A HupS-GFP fusion protein demonstrates a heterocyst-specific localization of the uptake hydrogenase in Nostoc punctiforme. FEMS Microbiol. Lett. 316, 152e159. Cao, H., Zhang, L., Melis, A., 2001. Bioenergetic and metabolic processes for the survival of sulfur-deprived Dunaliella salina (Chlorophyta). J. Appl. Phycol. 13, 25e34. Chang, C.H., King, P.W., Ghirardi, M.L., Kim, K., 2007. Atomic resolution modeling of the ferredoxin:[FeFe] hydrogenase complex from Chlamydomonas reinhardtii. Biophys. J. 93, 3034e3045. Chochois, V., Dauvillee, D., Beyly, A., Tolleter, D., Cuine, S., Timpano, H., Ball, S., Cournac, L., Peltier, G., 2009. Hydrogen
384 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION production in Chlamydomonas: PSII-dependent and independent pathways differ in their requirement on starch metabolism. Plant Physiol. 151, 613e640. Cohen, J., Kim, K., Posewitz, M., Ghirardi, M.L., Schulten, K., Seibert, M., King, P., 2005. Molecular dynamics and experimental investigation of H(2) and O(2) diffusion in [Fe]- hydrogenase. Biochem. Soc. Trans. 33, 80e82. Compaore, J., Stal, L.J., 2010. Oxygen and the light-dark cycle of nitrogenase activity in two unicellular cyanobacteria. Environ. Microbiol. 12, 54e62. Cournac, L., Guedeney, G., Peltier, G., Vignais, P.M., 2004. Sustained photoevolution of molecular hydrogen in a mutant of Synechocystis sp strain PCC 6803 deficient in the type I NADPH-dehydrogenase complex. J. Bacteriol. 186, 1737e1746. Desplats, C., Mus, F., Cuine, S., Billon, E., Cournac, L., Peltier, G., 2008. Characterization of Nda2, a plastoquinone-reducing type II NAD(P)H dehydrogenase in Chlamydomonas chloroplasts. J. Biol. Chem. 284, 4148e4157. Dietz, K.J., Heilos, L., 1990. Carbon metabolism in spinach leaves as affected by leaf age and phosphorus and sulfur nutrition. Plant Physiol. 93, 1219e1225. Eady, R.R., 1996. Structurefunction relationships of alternative nitrogenases. Chem. Rev. 96, 3013e3030. Ekman, M., Ow, S.Y., Holmqvist, M., Zhang, X., van Wagenen, J., Wright, P.C., Stensjo, K., 2011. Metabolic adaptations in a H2 producing heterocyst-forming cyanobacterium: potentials and implications for biological engineering. J. Proteome Res. 10, 1772e1784. Fedorov, A.S., Kosourov, S., Ghirardi, M.L., Seibert, M., 2005. Continuous hydrogen photoproduction by Chlamydomonas reinhardtii: using a novel two-stage, sulfate-limited chemostat system. Appl. Biochem. Biotechnol. 121-124, 403e412. Ferreira, R.M.B., Teixeira, A.R.N., 1992. Sulfur starvation in Lemna leads to degradation of ribulose-bisphosphate carboxylase without plant death. J. Biol. Chem. 267, 7253e7257. Florin, L., Tsokoglou, A., Happe, T., 2001. A novel type of iron hydrogenase in the green alga Scenedesmus obliquus is linked to the photosynthetic electron transport chain. J. Biol. Chem. 276, 6125e6132. Forestier, M., King, P., Zhang, L., Posewitz, M., Schwarzer, S., Happe, T., Ghirardi, M.L., Seibert, M., 2003. Expression of two [Fe]hydrogenases in Chlamydomonas reinhardtii under anaerobic conditions. Eur. J. Biochem. 270, 2750e2758. Fouchard, S., Hemschemeier, A., Caruana, A., Pruvost, J., Legrand, J., Happe, T., Peltier, G., Cournac, L., 2005. Autotrophic and mixotrophic hydrogen photoproduction in sulfur-deprived Chlamydomonas cells. Appl. Environ. Microbiol. 71, 6199e6205. Frey, M., 2002. Hydrogenases: hydrogen-activating enzymes. Chembiochem 3, 153e160. Gaffron, H., Rubin, J., 1942. Fermentative and photochemical production of hydrogen in algae. J. Gen. Physiol. 26, 219e240. Gfeller, R.P., Gibbs, M., 1984. Fermentative metabolism of Chlamydomonas reinhardtii, I: analysis of fermentative products from starch in dark and light. Plant Physiol. 75, 212e218. Ghirardi, M.L., 2006. Hydrogen production by photosynthetic green algae. Ind. J. Biochem. Biophys. 43, 201e210. Ghirardi, M., Mohanty, P., 2010. Oxygenic hydrogen photoproductionecurrent status of the technology. Curr. Sci. 98, 499e507. Ghirardi, M.L., Togasaki, R.K., Seibert, M., 1997. Oxygen sensitivity of algal hydrogen production. Appl. Biochem. Biotechnol. 63, 141e151. Ghirardi, M.L., Zhang, L., Lee, J.W., Flynn, T., Seibert, M., Greenbaum, E., Melis, A., 2000. Microalgae: a green source of renewable H2. Trends Biotechnol. 18, 506e511. Giannelli, L., Scoma, A., Torzillo, G., 2009. Interplay between light intensity, chlorophyll concentration and culture mixing on the hydrogen production in sulfur-deprived Chlamydomonas reinhardtii cultures grown in laboratory photobioreactors. Biotechnol. Bioengy 104, 76e90. Gibbs, M., Gfeller, R.P., Chen, C., 1986. Fermentative metabolism of Chlamydomonas reinhardtii. III: photoassimilation of acetate. Plant Physiol. 82, 160e166. Godman, J.E., Molnar, A., Baulcombe, D.C., Balk, J., 2010. RNA silencing of hydrogenase(-like) genes and investigation of their physiological roles in the green alga Chlamydomonas reinhardtii. Biochem. J. 431, 345e351. Greenbaum, E., 1982. Photosynthetic hydrogen and oxygen production: kinetic studies. Science 196, 879e880. Greenbaum, E., 1988. Energetic efficiency of hydrogen photoevolution by algal water splitting. Biophys. J. 54, 365e368. Greenbaum, E., Blankinship, S.L., Lee, J.W., Ford, R.M., 2001. Solar photobiochemistry: simultaneous photoproduction of hydrogen and oxygen in a confined bioreactor. J. Phys. Chem. B. 105, 3605e3609. Grimme, R.A., Lubner, C.E., Bryant, D.A., Golbeck, J.H., 2008. Photosystem I/molecular wire/metal nanoparticle bioconjugates for the photocatalytic production of H2. J. Am. Chem. Soc. 130, 6308e6309. Grimme, R.A., Lubner, C.E., Golbeck, J.H., 2009. Maximizing H2 production in photosystem I/dithiol molecular wire/platinum nanoparticle bioconjugates. Dalton Trans. 10106e10113. Grossman, A., 2000. Acclimation of Chlamydomonas reinhardtii to its nutrient environment. Protist 151, 201e224. Guan, Y.F., Deng, M.C., Yu, X.J., Zhang, W., 2004. Two-stage photobiological production of hydrogen by marine green alga Platymonas subcordiformis. Biochem. Eng. J. 19, 69e73. Gutthann, F., Egert, M., Marques, A., Appel, J., 2007. Inhibition of respiration and nitrate assimilation enhances photohydrogen evolution under low oxygen concentrations in Synechocystis sp. PCC 6803. Biochim. Biophys. Acta 1767, 161e169. Gärtner, K., Lechno-Yossef, S., Cornish, A.J., Wolk, P.C., Hegg, E.L., 2012. Expression of Shewanella oneidensis MR-1 [FeFe]-Hydrogenase genes in Anabaena sp. Strain PCC 7120. Appl. Environ. Microbiol. 78, 8579e8586. Hadfield, K.L., Bulen, W.A., 1969. Adenosine triphosphate requirement of nitrogenase from Azotobacter vinelandii. Biochem. 8, 5103e5108. Happe, T., Kaminski, A., 2002. Differential regulation of the Fe-hydrogenase during anaerobic adaptation in the green alga Chlamydomonas reinhardtii. Eur. J. Biochem. 269, 1022e1032. Happe, T., Naber, J.D., 1993. Isolation, characterization and N-terminal amino acid sequence of hydrogenase from the green alga Chlamydomonas reinhardtii. Eur. J. Biochem. 214, 475e481. Happe, T., Schutz, K., Bohme, H., 2000. Transcriptional and mutational analysis of the uptake hydrogenase of the filamentous cyanobacterium Anabaena variabilis ATCC 29413. J. Bacteriol. 182, 1624e1631. Healey, F.P., 1970. The mechanism of hydrogen evolution by Chlamydomonas moewusii. Plant Physiol. 45, 153e159. Helman, Y., Tchernov, D., Reinhold, L., Shibata, M., Ogawa, T., Schwarz, R., Ohad, I., Kaplan, A., 2003. Genes encoding a-type flavoproteins are essential for photoreduction of O2 in cyanobacteria. Curr. Biol. 13, 230e235. Hemschemeier, A., Fouchard, S., Cournac, L., Peltier, G., Happe, T., 2008. Hydrogen production by Chlamydomonas reinhardtii: an elaborate interplay of electron sources and sinks. Planta 227, 397e407. Hemschemeier, A., Happe, T., 2011. Alternative photosynthetic electron transport pathways during anaerobiosis in the green alga Chlamydomonas reinhardtii. Biochim. Biophys. Acta 1807, 919e926. Houchins, J.P., 1984. The physiology and biochemistry of hydrogen metabolism in cyanobacteria. Biochim. Biophys. Acta 768, 227e255.
REFERENCES Ihara, M., Nishihara, H., Yoon, K., Lenz, O., Friedrich, B., Nakamoto, H., Kojima, K., Honma, D., Kamachi, T., Okura, I., 2006a. Light-driven hydrogen production by a hybrid complex of a [NiFe]-hydrogenase and the cyanobacterial photosystem I. Photochem. Photobiol. 82, 676e682. Ihara, M., Nakamoto, H., Kamachi, T., Okura, I., Maedal, M., 2006b. Photoinduced hydrogen production by direct electron transfer from photosystem I cross-linked with cytochrome c3 to [NiFe]hydrogenase. Photochem. Photobiol. 82, 1677e1685. James, B.D., Baum, G.N., Perez, J., Baum, K.N., 2009. Technoeconomic Boundary Analysis of Biological Pathways to Hydrogen Production. Report No. SR-560-46674. NREL. p. 207. Jans, F., Mignolet, E., Houyoux, P.A., Cardol, P., Ghysels, B., Cuine, S., Cournac, L., Peltier, G., Remacle, C., Franck, F., 2008. A type II NAD(P)H dehydrogenase mediates light-independent plastoquinone reduction in the chloroplast of Chlamydomonas. Proc. Natl. Acad. Sci. USA 105, 20546e20551. Kentemich, T., Danneberg, G., Hundeshagen, B., Bothe, H., 1988. Evidence for the occurrence of the alternative, vanadium-containing nitrogenase in the cyanobacterium Anabaena variabilis. FEMS Microbiol. Lett. 51, 19e24. Khetkorn, W., Lindblad, P., Incharoensakdi, A., 2012. Inactivation of uptake hydrogenase leads to enhanced and sustained hydrogen production with high nitrogenase activity under high light exposure in the cyanobacterium Anabaena siamensis TISTR 8012. J. Biol. Eng. 6, 19. Kim, J.P., Kim, K.-R., Choi, S.P., Han, S.J., Kim, M.S., Sim, S.J., 2010. Repeated production of hydrogen by sulfate re-addition in sulfur deprived culture of Chlamydomonas reinhardtii. Int. J. Hydrogen Energy 35, 13387e13391. Kolber, Z., Zehr, Z., Falkowski, P., 1988. Effects of growth irradiance and nitrogen limitation on photosynthetic energy conversion in photosystem II. Plant Physiol. 88, 923e929. Kondratieva, E., Gogotov, I., 1983. Production of molecular hydrogen in microorganisms. In: Advances in Biochemical Engineering/ Biotechnology. Microbial Activities, vol. 23. Springer-Verlag, Berlin, Heidelberg, New York, Tokyo, pp. 139e191. Kosourov, S.N., Seibert, M., 2009. Hydrogen photoproduction by nutrient-deprived Chlamydomonas reinhardtii cells immobilized within thin alginate films under aerobic and anaerobic conditions. Biotechnol. Bioengy 102, 50e58. Kosourov, S., Tsygankov, A., Seibert, M., Ghirardi, M.L., 2002. Sustained hydrogen photoproduction by Chlamydomonas reinhardtii: effects of culture parameters. Biotechnol. Bioeng. 78, 731e740. Kosourov, S., Seibert, M., Ghirardi, M.L., 2003. Effects of extracelular pH on the metabolic pathways of sulfur-deprived, H2-producing Chlamydomonas reinhardtii cultures. Plant Cell Physiol. 44, 146e155. Kosourov, S., Patrusheva, E., Ghirardi, M.L., Seibert, M., Tsygankov, A., 2007. A comparison of hydrogen photoproduction by sulfur-deprived Chlamydomonas reinhardtii under different growth conditions. J. Biotechnol. 128, 776e787. Kosourov, S.N., Ghirardi, M.L., Seibert, M., 2011. A truncated antenna mutant of Chlamydomonas reinhardtii can produce more hydrogen than the parental strain. Int. J. Hydrogen Energy 36, 2044e2048. Kosourov, S.N., Batyrova, K.A., Petushkova, E.P., Tsygankov, A.A., Ghirardi, M.L., Seibert, M., 2012. Maximizing the hydrogen photoproduction yields in Chlamydomonas reinhardtii cultures: the effect of the H2 partial pressure. Int. J. Hydrogen Energy 37, 8850e8858. Krassen, H., Stripp, S., von Abendroth, G., Ataka, K., Happe, T., Heberle, J., 2009. Immobilization of the [FeFe]-hydrogenase CrHydA1 on a gold electrode: design of a catalytic surface for the production of molecular hydrogen. J. Biotechnol. 142, 3e9. 385 Kreuzberg, K., 1984. Starch fermentation via a formate-producing pathway in Chlamydomonas reinhardtii, Chlorogonium elongatum and Chlorella fusca. Physiol. Plant 61, 87e94. Kruse, O., Rupprecht, J., Bader, K.P., Thomas-Hall, S., Schenk, P.M., Finazzi, G., Hankamer, B., 2005. Improved photobiological H2 production in engineered green algal cells. J. Biol. Chem. 280, 34170e34177. Lambert, G.R., Smith, G.D., 1977. Hydrogen formation by marine blue-green algae. FEBS Lett. 83, 159e162. Laurinavichene, T., Tolstygina, I., Tsygankov, A., 2004. The effect of light intensity on hydrogen production by sulfur-deprived Chlamydomonas reinhardtii. J. Biotechnol. 114, 143e151. Laurinavichene, T.V., Fedorov, A.S., Ghirardi, M.L., Seibert, M., Tsygankov, A.A., 2006. Demonstration of sustained hydrogen photoproduction by immobilized, sulfur-deprived Chlamydomonas reinhardtii cells. Int. J. Hydrogen Energy 31, 659e667. Laurinavichene, T.V., Kosourov, S.N., Ghirardi, M.L., Seibert, M., Tsygankov, A.A., 2008. Prolongation of H2 photoproduction by immobilized, sulfur limited Chlamydomonas reinhardtii cultures. J. Biotechnol. 134, 275e277. Lecler, R., Godaux, D., Vigeolas, H., Hiligsmann, S., Thonart, P., Franck, F., Cardol, P., Remacle, C., 2011. Functional analysis of hydrogen photoproduction in respiratory-deficient mutants of Chlamydomonas reinhardtii. Int. J. Hydrogen Energy 36, 9562e9570. Lee, C.C., Hu, Y., Ribbe, M.W., 2009. Unique features of the nitrogenase VFe protein from Azotobacter vinelandii. Proc. Natl. Acad. Sci. USA 106, 9209e9214. Leino, H., Kosourov, S.N., Saari, L., Sivonen, K., Tsygankov, A.A., Aro, E.M., Allahverdiyeva, Y., 2012. Extended H2 photoproduction by N2-fixing cyanobacteria immobilized in thin alginate films. Int. J. Hydrogen Energy 37, 151e161. Lindberg, P., Schutz, K., Happe, T., Lindblad, P., 2002. A hydrogenproducing, hydrogenase-free mutant strain of Nostoc punctiforme ATCC 29133. Int. J. Hydrogen Energy 27, 1291e1296. Long, H., Chang, C.H., King, P.W., Ghirardi, M.L., Kim, K., 2008. Brownian dynamics and molecular dynamics study of the association between hydrogenase and ferredoxin from Chlamydomonas reinhardtii. Biophys. J. 95, 3753e3766. Lopez-Igual, R., Flores, E., Herrero, A., 2010. Inactivation of a heterocystspecific invertase indicates a principal role of sucrose catabolism in heterocysts of Anabaena sp. J. Bacteriol. 192, 5526e5533. Ludwig, M., Schulz-Friedrich, R., Appel, J., 2006. Occurrence of hydrogenases in cyanobacteria and anoxygenic photosynthetic bacteria: implications for the phylogenetic origin of cyanobacterial and algal hydrogenases. J. Mol. Evol. 63, 758e768. Makarova, V.V., Kosourov, S.N., Krendeleva, T.E., Kukarskikh, G.P., Ghirardi, M.L., Seibert, M., Rubin, A.B., 2005. Photochemical activity of photosystem II and hydrogen photoproduction in sulfur-deprived Chlamydomonas reinhardtii mutants D1-R323D and D1-R323L. Biofizika 50, 1070e1078. Makarova, V.V., Kosourov, S., Krendeleva, T.E., Semin, B.K., Kukarskikh, G.P., Rubin, A.B., Sayre, R.T., Ghirardi, M.L., Seibert, M., 2007. Photoproduction of hydrogen by sulfur-deprived Chlamydomonas reinhardtii mutants with impaired photosystem II photochemical activity. Photosynth. Research. 94, 79e89. Masukawa, H., Mochimaru, M., Sakurai, H., 2002. Disruption of the uptake hydrogenase gene, but not of the bi-directional hydrogenase gene, leads to enhanced photobiological hydrogen production by the nitrogen-fixing cyanobacterium Anabaena sp. PCC 7120. Appl. Microbiol. Biotechnol. 58, 618e624. Masukawa, H., Inoue, K., Sakurai, H., 2007. Effects of disruption of homocitrate synthase genes on Nostoc sp. strain PCC 7120 photobiological hydrogen production and nitrogenase. Appl. Environ. Microbiol. 73, 7562e7570.
386 21. RECENT DEVELOPMENTS ON CYANOBACTERIA AND GREEN ALGAE FOR BIOHYDROGEN PHOTOPRODUCTION Masukawa, H., Zhang, X., Yamazaki, E., Iwata, S., Nakamura, K., Mochimaru, M., Inoue, K., Sakurai, H., 2009. Survey of the distribution of different types of nitrogenases and hydrogenases in heterocyst-forming cyanobacteria. Mar. Biotechnol. 11, 397e409. Masukawa, H., Inoue, K., Sakurai, H., Wolk, C.P., Hausinger, R.P., 2010. Site-directed mutagenesis of the Anabaena sp. strain PCC 7120 nitrogenase active site to increase photobiological hydrogen production. Appl. Environ. Microbiol. 76, 6741e6750. Mayer, S.M., Gormal, C.A., Smith, B.E., Lawson, D.M., 2002. Crystallographic analysis of the MoFe protein of nitrogenase from a nifV mutant of Klebsiella pneumoniae identifies citrate as a ligand to the molybdenum of iron molybdenum cofactor (FeMoco). J. Biol. Chem. 277, 35263e35266. McIntosh, C.L., Germer, F., Schulz, R., Appel, J., Jones, A.K., 2011. The [NiFe]-hydrogenase of the cyanobacterium Synechocystis sp. PCC 6803 works bidirectionally with a bias to H2 production. J. Am. Chem. Soc. 133, 11308e11319. McNeely, K., Xu, Y., Ananyev, G., Bennette, N., Bryant, D.A., Dismukes, G.C., 2011. Synechococcus sp. strain PCC 7002 nifJ mutant lacking pyruvate: ferredoxin oxidoreductase. Appl. Environ. Microbiol. 77, 2435e2444. Melis, A., Chen, H.C., 2005. Chloroplast sulfate transport in green algae: genes, proteins and effects. Photosynth. Res. 86, 299e307. Melis, A., Zhang, L., Forestier, M., Ghirardi, M.L., Seibert, M., 2000. Sustained photobiological hydrogen gas production upon reversible inactivation of oxygen evolution in the green alga Chlamydomonas reinhardtii. Plant Physiol. 122, 127e136. Melnicki, M.R., Pinchuk, G.E., Hill, E.A., Kucek, L.A., Fredrickson, J.K., Konopka, A., Beliaev, A.S., 2012. Sustained H-2 production driven by photosynthetic water splitting in a unicellular cyanobacterium. mBio 3, e00197. Meunier, C.F., Yang, X.-Y., Rooke, J.C., Su, B.-L., 2011. Biofuel cells based on the immobilization of photosynthetically active bioentities. Chem. Cat. Chem. 3, 476e488. Meuser, J.E., Ananyev, G., Wittig, L.E., Kosourov, S., Ghirardi, M.L., Seibert, M., Dismukes, G.C., Posewitz, M.C., 2009. Phenotypic diversity of hydrogen production in chlorophycean algae reflects distinct anaerobic metabolisms. J. Biotechnol. 142, 21e30. Meuser, J.E., Boyd, E.S., Ananyev, G., Karns, D., Radakovits, R., Murthy, U.M.N., Ghirardi, M.L., Dismukes, G.C., Peters, J.W., Posewitz, M.C., 2011. Evolutionary significance of an algal gene encoding an [FeFe]-hydrogenase with F-domain homology and hydrogenase activity in Chlorella variabilis NC64A. Planta. 234, 829e843. Meuser, J.E., D’Adamo, S., Jinkerson, R.E., Mus, F., Yang, W., Ghirardi, M.L., Seibert, M., Grossman, A.R., Posewitz, M.C., 2012. Genetic disruption of both Chlamydomonas reinhardtii [FeFe]hydrogenases: Insight into the role of HYDA2 in H2 production. Biochem. Biophys. Res. Commun. 417, 704e709. Min, H.T., Sherman, L.A., 2010. Genetic transformation and mutagenesis via single-stranded DNA in the unicellular, diazotrophic cyanobacteria of the genus Cyanothece. Appl. Environ. Microbiol. 76, 7641e7645. Muro-Pastor, A.M., Hess, W., 2012. Heterocyst differentiation: from single mutants to global approaches. Trends Microbiol. 20, 548e557. Nagy, L., Meuser, J., Plummer, S., Seibert, M., Ghirardi, M., King, P., Ahmann, D., Posewitz, M., 2007. Application of gene-shuffling for the rapid generation of novel [FeFe]-hydrogenase libraries. Biotechnol. Lett. 29, 421e430. Ni, C.V., Yakuninin, A.F., Gogotov, I.N., 1990. Influence of molybdenum, vanadium, and tungsten on growth and nitrogenase synthesis of the free-living cyanobacterium Anabaena azollae. Microbiology 59, 395e398. Oliveira, P., Lindblad, P., 2009. Transcriptional regulation of the cyanobacterial bidirectional Hox-hydrogenase. Dalton Trans. 45, 9990e9996. Phelps, A.S., Wilson, P.W., 1942. Occurrence of hydrogenase in nitrogen-fixing organisms. Proc. Soc. Exp. Biol. Med. 47, 473e476. Philipps, G., Happe, T., Hemschemeier, A., 2012. Nitrogen deprivation results in photosynthetic hydrogen production in Chlamydomonas reinhardtii. Planta 235, 729e745. Phlips, E.J., Mitsui, A., 1986. Characterization and optimization of hydrogen production by a salt water blue-green algae Oscillatoria sp. strain Miami BG 7. II Use of immobilization for enhancement of hydrogen production. Int. J. Hydrogen Energy 11, 83e89. Pickett, C.J., 1996. The Chatt cycle and the mechanism of enzymic reduction of molecular nitrogen. J. Biol. Inorg. Chem. 1, 601e606. Polle, J.E.W., Benemann, J.R., Tanaka, A., Melis, A., 2000. Photosynthetic apparatus organization and function in wild type and a Chl b-less mutant of Chlamydomonas reinhardtii. Dependence on carbon source. Planta 211, 335e344. Polle, J.E.W., Kanakagiri, S., Jin, E., Masuda, T., Melis, A., 2002. Truncated chlorophyll antenna size of the photosystems e a practical method to improve microalgal productivity and hydrogen production in mass culture. Int. J. Hydrogen Energy 27, 1257e1264. Polle, J.E.W., Kanakagiri, S., Melis, A., 2003. tla1, a DNA insertional transformant of the green alga Chlamydomonas reinhardtii with a truncated light-harvesting chlorophyll antenna size. Planta 217, 49e59. Posewitz, M.C., Dubini, A., Meuser, J.E., Seibert, M., Ghirardi, M.L., 2009. Hydrogenases, hydrogen production, and anoxia. In: Stern, D.B. (Ed.), The Chlamydomonas Sourcebook, Second ed., vol. 2. Academic Press, Oxford, UK, pp. 217e255. Prochnik, S.E., Umen, J., Nedelcu, A.M., Hallmann, A., Miller, S.M., Nishii, I., Ferris, P., Kuo, A., Mitros, T., Fritz-Laylin, L.K., Hellsten, U., Chapman, J., Simakov, O., Rensing, S.A., Terry, A., Pangilinan, J., Kapitonov, V., Jurka, J., Salamov, A., Shapiro, H., Schmutz, J., Grimwood, J., Lindquist, E., Lucas, S., Grigoriev, I.V., Schmitt, R., Kirk, D., Rokhsar, D.S., 2010. Genomic analysis of organismal complexity in the multicellular green alga Volvox carteri. Science 329, 223e226. Pröschold, T., Harris, E.H., Coleman, A.W., 2005. Portrait of a species: Chlamydomonas reinhardtii. Genetics 170, 1601e1610. Randt, C., Senger, H., 1985. Participation of the two photosystems in light dependent hydrogen evolution in Scenedesmus obliquus. Photochem. Photobiol. 42, 553e557. Raupach, M.R., Marland, G., Ciais, P., Le Quere, C., Canadell, J.G., Klepper, G., Field, C.B., 2007. Global and regional drivers of accelerating CO2 emissions. Proc. Natl. Acad. Sci. USA. 104, 10288e10293. Risser, D.D., Callahan, S.M., 2009. Genetic and cytological evidence that heterocyst patterning is regulated by inhibitor gradients that promote activator decay. Proc. Natl. Acad. Sci. USA. 106, 19884e19888. Roessler, P.G., Lien, S., 1984. Purification of hydrogenase from Chlamydomonas reinhardtii. Plant. Physiol 75, 705e709. Rubio, L., Ludden, P., 2005. Maturation of nitrogenase: a biochemical pazzle. J. Bacteriol. 187, 405e414. Sarkar, S., Pandey, K.D., Kashyap, A.K., 1992. Hydrogen photoproduction by filamentous nonheterocystous cyanobacterium Plectonema boryanum and simultaneous release of ammonia. Int. J. Hydrogen Energy 17, 689e694. Sauer, J., Schreiber, U., Schmid, R., Volker, U., Forchhammer, K., 2001. Nitrogen starvation-induced chlorosis in Synechococcus PCC 7942. Low-level photosynthesis as a mechanism of long-term survival. Plant Physiol. 126, 233e243. Schrautemeier, B., Neveling, U., Schmitz, S., 1995. Distinct and differentially regulated Mo-dependent nitrogen-fixing systems evolved for heterocysts and vegetative cells of Anabaena variabilis
REFERENCES ATCC 29413: characterization of the fdX1/2 gene regions as part of the nif1/2 gene clusters. Mol. Microbiol. 18, 357e359. Schutz, K., Happe, T., Troshina, O., Lindblad, P., Leitao, E., Oliveira, P., Tamagnini, P., 2004. Cyanobacterial H2 production e a comparative analysis. Planta 218, 350e359. Scoma, A., Krawietz, D., Faraloni, C., Giannelli, L., Happe, T., Torzillo, G., 2012. Sustained H2 production in a Chlamydomonas reinhardtii D1 protein mutant. J. Biotechnol. 157, 613e619. Seabra, R., Santos, A., Pereira, S., Moradas-Ferreira, P., Tamagnini, P., 2009. Immunolocalization of the uptake hydrogenase in the marine cyanobacterium Lyngbya majuscula CCAP 1446/4 and two Nostoc strains. FEMS Microbiol. Lett. 292, 57e62. Skizim, N.J., Ananyev, G.M., Krishnan, A., Dismukes, G.C., 2012. Metabolic pathways for photobiological hydrogen production by nitrogenase-and hydrogenase-containing unicellular cyanobacteria Cyanothece. J. Biol. Chem. 287, 2777e2786. Skjånes, K., Knutsen, G., Kallqvist, T., Lindblad, P., 2008. H2 production from marine and freshwater species of green algae during sulfur deprivation and considerations for bioreactor design. Int. J. Hydrogen Energy 33, 511e521. Skjånes, K., Pinto, F.L., Lindblad, P., 2010. Evidence for transcription of three genes with characteristics of hydrogenases in the green alga Chlamydomonas noctigama. Int. J. Hydrogen Energy 35, 1074e1088. Stapleton, J.A., Swartz, J.R., 2010. A cell-free microtiter plate screen for improved [FeFe] hydrogenases. PLoS ONE 5, e10554. Steunou, A.S., Jensen, S.I., Brecht, E., Becraft, E.D., Bateson, M.M., Kilian, Q., Bhaya, D., Ward, D.M., Peters, J.W., Grossman, A.R., Kuhl, M., 2008. Regulation of nif gene expression and the energetics of N2 fixation over diel cycle in a hot spring microbial mat. ISME J. 2, 364e378. Stuart, T.S., Gaffron, H., 1972. The mechanism of hydrogen photoproduction by several algae. Planta. 106, 101e112. Surzycki, R., Cournac, L., Peltier, G., Rochaix, J., 2007. Potential for hydrogen production with inducible chloroplast gene expression in Chlamydomonas. Proc. Natl. Acad. Sci. USA. 104, 17548e17553. Tamagnini, P., Leita, E., Oliveira, P., Ferreira, D., Pinto, F., Harris, D.J., Heidorn, T., Lindblad, P., 2007. Cyanobacterial hydrogenases: diversity, regulation and applications. FEMS Microbiol. Rev. 31, 692e720. Thiel, T., 1993. Characterization of genes for an alternative nitrogenase in the cyanobacterium Anabaena variabilis. J. Bacteriol. 175, 6276e6286. Thiel, T., Lyons, E.M., Erker, J., Ernst, A., 1995. A second nitrogenase in vegetative cells of a heterocyst-forming cyanobacterium. Proc. Natl. Acad. Sci. USA 92, 9358e9362. Tolstygina, I.V., Antal, T.K., Kosourov, S.N., Krendeleva, T.E., Rubin, A.B., Tsygankov, A.A., 2009. Hydrogen production by photoautotrophic sulfur-deprived Chlamydomonas reinhardtii pregrown and incubated under high light. Biotechnol. Bioengy 102, 1055e1061. Torzillo, G., Pushparaj, B., Masojidek, J., Vonshak, A., 2003. Biological constraints in algal biotechnology. Biotechnol. Bioprocess. Eng. 8, 338e348. Torzillo, G., Scoma, A., Faraloni, C., Ena, A., Johanningmeier, U., 2009. Increased hydrogen photoproduction by means of a sulfurdeprived Chlamydomonas reinhardtii D1 protein mutant. Int. J. Hydrogen Energy 34, 4529e4536. Tsygankov, A.A., 2012. Hydrogen production: light-driven processes e green algae. In: Hallenbeck, P.C. (Ed.), Microbial Technologies, Advanced Biofuels Production, pp. 29e51. 387 Tsygankov, A.A., Hirata, Y., Miyake, M., Asada, Y., Miyake, J., 1994. Photobioreactor with photosynthetic bacteria immobilized on porous-glass for hydrogen photoproduction. J. Ferment. Bioeng. 77, 575e578. Tsygankov, A., Kosourov, S., Seibert, M., Ghirardi, M.L., 2002. Hydrogen photoproduction under continuous illumination by sulfur-deprived, synchronous Chlamydomonas reinhardtii cultures. Int. J. Hydrogen Energy 27, 1239e1244. Tsygankov, A.A., Kosourov, S.N., Tolstygina, I.V., Ghirardi, M.L., Seibert, M., 2006. Hydrogen production by sulfur-deprived Chlamydomonas reinhardtii under photoautotrophic conditions. Int. J. Hydrogen Energy 31, 1574e1584. Weyman, P.D., Pratte, B., Thiel, T., 2010. Hydrogen production in nitrogenase mutants in Anabaena variabilis. FEMS Microbiol. Lett. 304, 55e61. Winkler, M., Hemschemeier, A., Gotor, C., Melis, A., Happe, T., 2002. [Fe]-hydrogenases in green algae: photo-fermentation and hydrogen evolution under sulfur deprivation. Int. J. Hydrogen Energy 27, 1431e1439. Winkler, M., Maeurer, C., Hemschemeier, A., Happe, T., 2004. The isolation of green algal strains with outstanding H2eproductivity. In: Miyake, J., Igarashi, Y., Roegner, M. (Eds.), Biohydrogen III. Elsevier Science, Oxford, pp. 103e115. Winkler, M., Kuhlgert, S., Hippler, M., Happe, T., 2009. Characterization of the key step for light-driven hydrogen evolution in green algae. J. Biol. Chem. 284, 36620e36627. Wolk, C.P., Ernst, A., Elhai, J., 1994. In: Bryant, D.A. (Ed.), Heterocyst Metabolism and Development. The Molecular Biology of Cyanobacteria. Kluwer, Dordrecht, pp. 769e823. Wünschiers, R., Stangier, K., Senger, H., Schulz, R., 2001. Molecular evidence for a Fe-hydrogenase in the green alga Scenedesmus obliquus. Curr. Microbiol. 42, 353e360. Wykoff, D.D., Davies, J.P., Melis, A., Grossman, A.R., 1998. The regulation of photosynthetic electron transport during nutrient deprivation in Chlamydomonas reinhardtii. Plant Physiol. 117, 129e139. Yacoby, I., Pochekailov, S., Toporik, H., Ghirardi, M.L., King, P.W., Zhang, S., 2011. Photosynthetic electron partitioning between [FeFe]-hydrogenase and ferredoxin:NADPþ-oxidoreductase (FNR) enzymes in vitro. Proc. Natl. Acad. Sci. USA 108, 9396e9401. Yan, F., Chen, Z., Li, W., Cao, X., Xue, S., Zhang, W., 2011. Purification and characterization of a hydrogenase from the marine green alga Tetraselmis subcordiformis. Process Biochem. 46, 1212e1215. Yeager, C.M., Milliken, C.E., Bagwell, C.E., Staples, L., Berseth, P.A., Sessions, H.T., 2011. Evaluation of experimental conditions that influence hydrogen production among heterocystous cyanobacteria. Int. J. Hydrogen Energy 36, 7487e7499. Yoshino, F., Ikeda, H., Masukawa, H., Sakurai, H., 2007. High photobiological hydrogen production activity of a Nostoc sp. PCC 7422 uptake hydrogenase-deficient mutant with high nitrogenase activity. Mar. Biotechnol. 9, 101e112. Zhang, L., Happe, T., Melis, A., 2002. Biochemical and morphological characterization of sulfur-deprived and H2-producing Chlamydomonas reinhardtii (green alga). Planta. 214, 552e561. Zhao, M.X., Jiang, Y.-L., He, Y.X., Chen, Y.F., Teng, Y.B., Chen, Y., Zhang, C.C., Zhou, C.Z., 2010. Structural basis for the allosteric control of the global transcription factor NtcA by the nitrogen starvation signal 2-oxoglutarate. Proc. Natl. Acad. Sci. USA 107, 12487e12492. Zinn, T., Schnackenberg, J., Haak, D., Romer, S., Schulz, R., Senger, H., 1994. Evidence for nickel in the soluble hydrogenase from the unicellular green alga Scenedesmus obliquus. J. Biosci. 49, 33e38.
C H A P T E R 22 Engineered Cyanobacteria: Research and Application in Bioenergy Gustavo B. Leite, Patrick C. Hallenbeck* Département de Microbiologie et Immunologie, Université de Montréal, Montréal, Québec, Canada *Corresponding author email: patrick.hallenbeck@umontreal.ca O U T L I N E Introduction 389 Engineering Cyanobacteria Strains, Tools and Methods 392 392 Cyanobacteria as a Production System for Biofuels: 393 Current Status Hydrogen 393 Hydrogen Bioproduction 394 Hydrogen-Evolving Enzymes 394 Hydrogen Bioproduction 395 Ethanol 398 INTRODUCTION Paleontological and geochemical data as well as molecular analysis of the plastid genome point to a single prokaryote as the origin of several groups of organisms scattered throughout the tree of life, including the entire kingdom of Plantae (Knoll, 2008; Yoon, 2004). A cyanobacterial ancestor is believed to be the only organism ever to couple together two photosystems, harvesting electrons from water to produce energy-rich molecules such as adenosine triphosphate (ATP) and reduced nicotinamide adenine dinucleotide phosphate (NADPH) (Knoll, 2008) (Figure 22.1). These molecules provide the necessary chemical energy, protons and electrons for cellular reactions and the synthesis of other molecules, most importantly powering CO2 fixation through the Calvin-Benson-Bassham cycle. This event is thought to have happened between the mid-Archean and early Proterozoic eras (2000e3000 millions of years ago). The Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00022-X Ethylene Microbial Production of Ethylene Bioproduction of Ethylene Using efe Isoprene Butyraldehyde and Butanol Photosynthetic Production of Aliphatic Alcohols and Alkanes 398 399 399 400 401 402 Conclusion and Outlook 403 References 403 atmosphere was poor in oxygen and rich in CO2, and the oceans were rich in salts and minerals; perfect conditions for the first algal blooms. The invention of oxygenic photosynthesis conferred a great advantage to this ancient cyanobacterium, starting widespread speciation and changing the composition of the atmosphere through the oxidation of water into protons and molecular oxygen (Figure 22.1). This was probably the first universally relevant instance of primary production and established a food chain by transforming inorganic nutrients into organic molecules that could be used by heterotrophic organisms (Knoll, 2008). The role of primary producers, so important in fully establishing life on earth, is still equally important today, when cyanobacteria are thought to be responsible for 25% of all carbon dioxide fixation and together with eukaryotic microalgae sustain most of oceanic life, fixing CO2 and carrying out important steps in various biogeochemical nutrient cycles (Field et al., 1998). 389 Copyright Ó 2014 Elsevier B.V. All rights reserved.
390 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY FIGURE 22.1 Scheme of light reactions in oxygenic photosynthesis. Photosystem II oxidizes a water molecule, harvesting the electron that will be used to synthesize NADPH, and producing an electrochemical gradient through the release of protons (Hþ) that will be used by ATP synthase to drive phosphorylation of ADP. The ATP and NADPH that are produced are used by the Calvin cycle for CO2 fixation (dark reactions). (For color version of this figure, the reader is referred to the online version of this book.) Humanity is totally dependent on photosynthesis for food and fuel. As well as a source of organic carbon, mankind relies on photosynthesis as energy source, through the use of fossil fuels, ancient photosynthetic products stored and cooked under pressure for millions of years, the burning of readily available biomass, or more recently through the use of biofuel crops as a new source of liquid fuels. Sugarcane or corn ethanol and biodiesel have been produced from crops for more than 40 years, with a greatly increased role the last two decades. These first-generation biofuels are presently being produced at large scale, with worldwide production of ethanol and biodiesel of 50 billion and 9 billion liters, respectively, in 2007. Even though these seem like significant quantities, biofuels still represent a miniscule fraction of the world’s primary energy use; in 2011, 161 tons per day of renewable liquid biofuels were produced, whereas 12 million tons per day of crude oil were consumed (BP, 2012). Humans have been constantly perfecting agricultural technology since the dawn of civilization, and with the green revolution, food crop yields have shown considerable increases decade after decade, although this progress is now stagnating in many food producing areas (Ray et al., 2012). At any rate, given the enormous demand for energy and the predicted increase in the world’s population to 9 billion by 2050, it is evident that there is not enough arable land to satisfy both nutritional and energy demands through food and fuel crops. Of course, in addition to renewable energy derived through photosynthesis, other sources of sustainable energy exist: solar, wind, geothermal, hydroelectric, etc., but together these energy sources cannot supply the quantity and types of energy demanded worldwide since electricity is not suitable for all applications. Modern society is built around liquid and gaseous fuels, which are very efficient energy carriers suitable for a variety of applications, in particular mobile power. Liquid biofuels are essentially photosynthetically derived compounds, at present sustainably produced through the cultivation of energy crops, but as discussed above, this directly competes with the production of food crops. A possible and promising alternative for sustainable energy production system is intimately related to crude oil formation over the previous millions of years. Before the appearance of vascular land plants on earth, ancestral cyanobacteria were already occupying a large variety of environments and now, after a long period of evolution, cyanobacteria and the microalgae formed through endosymbiosis of cyanobacteria, can be isolated from virtually any natural water sample, from extremely fresh water to hypersaline lakes, from snow in the Arctic Circle to hot or relatively dry environments. The richness of this speciation over billions of years can be appreciated through the variety of morphological forms that are found. These organisms show themselves to be a promising system for the production of hydrocarbons and other desirable products. Cultivation can be carried out using nonarable land; seawater and wastewater have been shown to support growth, bioremediating effluents while fixing atmospheric carbon dioxide into
INTRODUCTION possible commercial products. The rather simple nutrition requirements of these organisms highlight the capability of their metabolism to produce all the molecules needed for cellular growth. Their pathways frequently contain metabolites with commercial interest that can be readily used or easily processed into a final product (Figure 22.2). 391 Although cyanobacteria and eukaryotic algae share these attributes, cyanobacteria have the additional advantage of being relatively easily manipulated genetically. Thus, using cyanobacteria, if a desired product is not naturally produced, genetic engineering techniques allow the insertion of genes or even entire pathways to make novel products, either high-value compounds or FIGURE 22.2 Scheme of the TCA cycle with the alternatives proposed for cyanobacteria, blue pathways on the bottom (Zhang et al., 2011), and for production of ethylene through the ethylene-forming enzyme isolated from P. syringae (orange pathway in the center). The lack of homologous genes for 2-oxoglutarate dehydrogenase in cyanobacteria led to the idea that they have an incomplete TCA cycle, working as two branched chains of reactions (oxidative and reductive) generating succinate from fumarate. However, a new 2-oxoglutarate decarboxylase recently described in Synechococcus sp. was the missing piece that closes the TCA cycle in cyanobacteria. Homologs to this gene were found in all cyanobacteria already sequenced with the exception of Prochlorococcus and marine Synechococcus sp. Source: Zhang et al., 2011. (For interpretation of the references to color in this figure legend, the reader is referred to the online version of this book.)
392 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY commodity chemicals such as biofuels. Of course any molecule that is produced using cyanobacteria could be produced in other microorganisms, especially fermentation workhorses such as Saccharomyces cerevisiae or Escherichia coli, but these are heterotrophs requiring carbon compounds previously fixed through photosynthesis, i.e. agriculturally produced. Thus, the cyanobacteria are uniquely positioned to carry out CO2 fixation driven by solar energy capture while at the same time being amenable of genetic engineering to produce a wide variety of liquid and gaseous biofuels. In this chapter the current achievements on research toward the production of biofuels and crude oil substitutes using cyanobacteria as a model organism are reviewed. As will be seen, although much has already been achieved in terms of engineering toward the production of biofuels, in most cases productivity is the greatest bottleneck, although some steps in downstream processing also present many challenges. Thus, at present, use of a cyanobacterial system for commercial production of biofuels at cost-effective levels still faces significant hurdles. ENGINEERING CYANOBACTERIA The major argument for using cyanobacteria or eukaryote microalgae for biofuel production is the possibility to directly couple photosynthesis with product formation. This strategy could have sustainable and economic advantages. The financial appeal is related to the production chain, with CO2 fixation directly producing the desired fuel in a single organism. Thus, the biofuel is recovered at the production site, avoiding as a consequence the processing of photosynthetically produced sugars in a second-stage microbial fermentation. This process also has great ecological and sustainability appeal since atmospheric CO2 is being recycled into fuels without using the conventional agriculture system, leaving arable land available for food crops. Nevertheless, the inherently low value and high demand characteristics of fuels present a challenge for the development of biofuel production. The volume of fuel required to fulfill the needs of the transportation sector is massive, in contrast to their low market value, which must be at least as cheap as bottled water. The achievement of this goal requires the solution of major challenges in civil and mechanical engineering, chemistry, and biology. In the biological arena, the main challenge is strain development. The ideal cyanobacterium for biofuel production would have a high quantum efficiency of photosynthesis and well-defined carbon partitioning, where the CO2 fixed would be primarily directed to “housekeeping” metabolism and the targeted product. To achieve this goal, two main venues are being followed: high-throughput bioprospecting, which seeks naturally occurring species, enzymes and pathways adaptable for cultivation and economic exploitation, or the use of genetic engineering, where a model organism is genetically modified to introduce and/or to enhance the production of a desired molecule. In this section the available tools are discussed as well as some paths toward the improvement of photosynthetic quantum efficiency. Strains, Tools and Methods Originating in an environment without available fixed carbon, cyanobacteria have evolved as versatile organisms, capable of producing a large variety of organic compounds from simple inorganic sources that can be directly used or transformed into a commercial product. When the desired molecule is not naturally produced, genes or entire pathways can be introduced through a variety of methods and product yields can be increased by driving cell metabolism toward the desired product. There are more than 3350 species of cyanobacteria already described, with hundreds available in culture collections (Guiry and Guiry). To date, 87 cyanobacterial genomes have been sequenced and deposited in public databases but only a few strains have been used in genetic manipulation studies (Heidorn et al., 2011). Many molecular tools are currently available and genetic manipulation can be pursued through conjugation, electroporation or natural transformation. These techniques are constantly being revised or optimized for each host species and sample protocols are available elsewhere (Heidorn et al., 2011). So far, no cyanophage able to perform transduction has been described, nevertheless this technique is still the object of great interest (Koksharova and Wolk, 2002). Natural transformation is an appealing feature found in some cyanobacterial strains, with two standing out as being frequently used in genetic manipulation studies, Synechocystis sp. PCC 6803 (Pasteur Culture Collection) and Synechococcus sp. PCC 7002 (Grigorieva and Shestakov, 1982). These two strains are of significant interest due to the high yield of mutants achieved through this technique, making it widely used for both pure and applied science, from plant physiology studies to metabolic engineering aiming for the commercial production of biomolecules. The high frequency of transformants with natural transformation is intimately linked with the nature of the transferred genetic material, with chromosomal DNA reaching up to 100-fold more viable transformants than when replicative plasmids are used as the source of DNA (Golden and Sherman, 1983; Shestako and Khyen, 1970). In fact, this is true specifically for replicative plasmids since most of the transformation efficiency is
CYANOBACTERIA AS A PRODUCTION SYSTEM FOR BIOFUELS: CURRENT STATUS recovered when a suicide plasmid is used (Tsinoremas et al., 1994). Thus, it would seem that the final localization of the inserted DNA plays a key role in the transformation efficiency. This is argued to be related to the postreplicative processing of chromosomal DNA together with a putative robust recombination mechanism in these species (Flores et al., 2008). Natural transformation has being reported to be associated with pilus-related genes (Yoshihara et al., 2001; Yura, 1999), a natural machinery putatively adapted to take up exogenous DNA with such high efficiency that different artificial procedures intended to increase the transformation yield fail to improve the frequency of viable mutants (Zang et al., 2007). Unfortunately, natural transformation is not widespread in the cyanobacterial phylum and many species require other techniques for the efficient introduction of exogenous DNA. Electroporation was first demonstrated in Anabaena sp. (Thiel and Poo, 1989) and today has been optimized for many strains. It has been shown to be effective despite the low yield in many cases (Koksharova and Wolk, 2002). Unlike what is observed for green algae (Kilian et al., 2011), the procedures and electric pulse settings are not very different from those used with other bacterial phyla (Heidorn et al., 2011). However, even though it can be an effective method, the ease of natural transformation and the higher yield of conjugation have left electroporation behind as a choice for mutagenesis. Conjugation is the most commonly used technique for genetic engineering in terms of the diverse species with which it can be used, and, with the filamentous N2 fixing (heterocyst forming) cyanobacteria, it is the only effective technique thus far described. With the advent of molecular biology, plasmids of cyanobacterial origin were actively sought with the intention of producing shuttle vectors allowing their transfer from E. coli to Synechococcus (Golden and Sherman, 1983). Since then, E. coli has been widely used for conjugation with many filamentous strains, such as Nostoc sp. and Anabaena sp., and single cell strains, like Synechococcus sp. and Synechocystis sp. Although incorporation of DNA into the chromosome of many strains has proved to be relatively easily achieved when using linear DNA or suicide plasmids, it has proved challenging to make cyanobacteria harbor replicative plasmids. During conjugation, the plasmid is relaxed and single-stranded DNA is driven to the recipient cell through the type four secretion system by the enzyme relaxase. Once in the recipient cell, the transferred DNA will have its antisense strand resynthesized and this newly reformed plasmid can integrate itself into the genome or autoreplicate. The vectors used in cyanobacteria must contain the replicons for both organisms, donor and recipient, a mobilization site (origin of transfer, e.g. bom, nic and oriT), a selective marker effective for both organisms, 393 and a codon optimization to avoid the broad range of restriction enzymes harbored by cyanobacteria, which has been found to be an important hurdle to successful conjugation (Elhai et al., 1997; Flores et al., 2008; Wolk et al., 1984). Extra enzymes might be needed to ensure a successful transfer, which could be encoded on secondary (aka helper) plasmids. Among these special enzymes are some endonucleases, intended to cut the cargo plasmid at the bom site and promote transfer, and methylases to protect the transferred DNA against the restriction enzymes in the recipient. Detailed procedures, strategies and strains used are amply reviewed elsewhere (Heidorn et al., 2011). CYANOBACTERIA AS A PRODUCTION SYSTEM FOR BIOFUELS: CURRENT STATUS Hydrogen Frequently cited as the fuel of the future, hydrogen production, storage and utilization are being widely investigated. As a transportation fuel it presents a series of challenges in every link of the chain, from production to storage and distribution. Although having a low volumetric energy density, hydrogen has the highest energy density per mass and the simple fact that its combustion generates almost only water and heat has seduced entire generations. “Yes, my friends, I believe that water will one day be employed as fuel, that hydrogen and oxygen which constitute it, used singly or together, will furnish an inexhaustible source of heat and light, of an intensity of which coal is not capable” (Verne). Cars that could run on water with minimal energy consumption have captured the imagination of many people and, not surprisingly, have inspired frauds like the almost magical conversion of saltwater into fuel using radiofrequency radiation, claimed by John Kanzius and broadcast live countrywide from Philadelphia, or the notorious “StanleyMeyer’s water fuel cell” to be used in an internal combustion engine, where a special device could split water giving an energy output sufficient to generate mechanical energy for the vehicle with enough leftover to power a fuel cell that would provide more hydrogen and oxygen through water splitting. Considering that the combustion of hydrogen and oxygen regenerates water, both systems obviously defy the first and second laws of thermodynamics (Ball, 2007). Despite the motivation behind these schemes, they touched upon the most limiting step in the development of the hydrogen fuel technology: production. In current industrial practice, hydrogen can be produced by pyrolysis, electrolysis or by steam reforming of hydrocarbons. The last is the dominant method, applied to fossil fuels,
394 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY usually natural gas (methane). This makes hydrogen both expensive and unsustainable. Hydrogen Bioproduction Molecular hydrogen (H2) is the lightest gas possible. When released into the atmosphere it diffuses quickly toward the troposphere, thus, at the sea level it can only be found in trace amounts. For this reason, very little naturally occurring H2 is available and therefore a sustainable production system must be found if this molecule is to be used as a fuel. Efficient biological production of hydrogen could represent a breakthrough in the development of this energy carrier and many different approaches are being followed toward this goal. Undoubtedly, among all the possible fuels that could be produced by cyanobacteria, it is hydrogen that has received the most attention. Here we discuss the biological mechanisms for hydrogen production and advances toward yield improvements in cyanobacteria. In the light reactions of photosynthesis, light is captured by photosystems I and II, acting together to transform solar energy into chemical energy, splitting water into molecular oxygen and protons (Hþ) and the reducing agent NADPH. The transmembrane proton gradient that is formed is used by ATP synthase to combine adenosine diphosphate (ADP) þ Pi into ATP (Figure 22.1). This set of reactions is rather interesting because it effectively conserves ubiquitous solar energy in energy-dense molecules using an abundant substrate, water. Ironically, cyanobacteria (and all plants) had been all along for millions of years the very sought after solution for breaking the strong bond between oxygen and hydrogen in the water molecule without using the special radiofrequency of John Kanzius or the mysterious fuel cell of Stanley Meyer. During the water-splitting process, oxygen is released in its molecular form (O2), while hydrogen, in the form of protons, is further used to produce two molecules of high-energy content: ATP and NADPH. Together, they feed energy into the Calvin-Benson-Bassham cycle, where CO2 is fixed into organic molecules, as well as into many other reactions related to cellular homeostasis or secondary metabolism. Alternatively, before it is used to generate NADPH, the high-energy electron generated by photosynthesis can be directly used for the evolution of hydrogen, a process called direct biophotolysis (Benemann and Weare, 1974). Therefore, hydrogen evolution through this route does not require CO2 fixation, and solar energy and water, together with the required enzymes, are sufficient for H2 formation (Hallenbeck and Benemann, 2002). The major problem with this process is that hydrogenases, the hydrogenevolving enzymes, are extremely sensitive to oxygen (O2) and are irreversibly inactivated by even small concentrations of this gas. Thus, hydrogen evolution is usually a short-lived process, with a burst of hydrogen evolution when transitioning from a dark cycle into light as increasing oxygenic photosynthesis quickly inactivates the hydrogenase. Some species, especially filamentous ones (e.g. Anabaena sp. and Nostoc sp.), capable of forming specialized cells called heterocysts, can be shown to produce hydrogen over prolonged periods in light, as the heterocysts provide an oxygen-free environment that protects the hydrogenase against inactivation. In indirect biophotolysis, the captured light energy is used to fix CO2 and the organic molecules that are produced are stored as reserve material. Under normal conditions, part of these carbon reserves will be oxidized over the dark period to maintain cellular homeostasis. However, under proper conditions such a culture can be induced to produce hydrogen, thus separating hydrogen evolution temporally and spatially from the oxygen evolved by oxygenic photosynthesis (Hallenbeck, 2011). Thus, hydrogenase activity is maintained and the simultaneous production of hydrogen and oxygen, an explosive mixture when concentrated in the headspace of a bioreactor, is avoided. Hydrogen-Evolving Enzymes Hydrogenases in cyanobacteria have been studied for over 35 years (Benemann and Weare, 1974; Hallenbeck and Benemann, 1978) and many variations of hydrogenases have been described in different bacterial phyla (Vignais and Billoud, 2007). These enzymes are frequently classified into three different groups: nitrogenase, the reversible hydrogenase (Hox), and the uptake hydrogenase (Hup) (Ghirardi et al., 2007). HUPdHYDROGEN UPTAKE ENZYME Hup is a [NiFe] hydrogenase that occurs associated with the thylakoid membrane (Seabra et al., 2009). This enzyme shows the least sensitivity to oxygen among the three classes. Its function is in the oxidation of H2, returning the captured electrons to cellular electron transfer reactions. To date it has been found only in N2-fixing strains and appears to have an intimate relationship with nitrogenase (Marreiros et al., 2013). Under natural conditions, nitrogenase functions to reduce atmospheric N2 to NH3, producing H2 in an unavoidable side reaction. It is thought that Hup functions to recycle the recently formed H2, which is oxidized back into protons or reacted with O2 in a respiratory oxyhydrogen reaction, protecting the nitrogenase from O2 inactivation, avoiding an excessive build up of H2 in the cell and recovering part of the ATP used in its formation (Bothe et al., 2010; Tamagnini et al., 2007). In the nitrogen-fixing cyanobacteria, transcription of the Hup-encoding genes hupSL is associated with the nitrogen depletion response and
CYANOBACTERIA AS A PRODUCTION SYSTEM FOR BIOFUELS: CURRENT STATUS 395 is under the regulation of the NtcA, the global nitrogen regulator (Weyman et al., 2008). Hup inactivation increases the production of H2 two- to threefold in most cyanobacteria (Ludwig et al., 2006; Tamagnini et al., 2007). heterocyst can maintain an internal anoxic environment since the expression of PSII is repressed. Hydrogen production therefore is supported through the use of carbon compounds delivered by the neighboring vegetative cells. NITROGENASEdA GRATUITOUS HYDROGENASE REVERSIBLE HYDROGENASE (HOX) In nature this complex enzyme carries out a critical function, breaking the three covalent bonds of molecular nitrogen (N2) providing ammonia to the cell and closing the nitrogen cycle. This process consumes a large amount of energy in the form of ATP and highenergy electrons (Eqn (22.1)), producing NH3 with the coproduction of hydrogen in an unavoidable side reaction. In addition to nitrogenase, N2-fixing cyanobacteria can have a second hydrogen-evolving enzyme, the so-called reversible hydrogenase (Hox). This enzyme is a heteropentameric complex that is formed by a hydrogenase module (HoxHY) and a diaphorase module (HoxEFU), which transfers electrons from NAD(P)H to the hydrogenase module (Bothe et al., 2010). Like Hup, Hox is a [NiFe] hydrogenase, but in this case it shows a high sensitivity to O2. Its expression is totally independent from that of nitrogenase and varies among species. In some cases it is under the control of the circadian clock, where it is shown to promote hydrogen production in the dark (Hallenbeck and Benemann, 1978; Schmitz et al., 2001). The bidirectional hydrogenase is not taxon specific, being found in many different groups of cyanobacteria, and its location and organization in the chromosome are also heterogeneous. Recent studies regarding Hox transcription factors have elucidated many aspects of its regulatory mechanisms, which are reviewed elsewhere (Oliveira and Lindblad, 2009). N2 þ 10Hþ þ 8e þ 16ATP/2NH3 þ H2 þ 16ADP (22.1) The most common nitrogenase is the Mo-Fe nitrogenase, which is characterized by a complex iron-sulfur cluster containing molybdenum. While performing nitrogen fixation, up to one-fourth of the electron flux goes toward the reduction of hydrogen. Variations of this enzyme includes the substitution of the molybdenum by vanadium or iron (V-Fe and FeeFe nitrogenases, respectively), which, although a greater proportion of electrons are allocated to hydrogen production, in fact show a lower net flux of electrons to hydrogen since their overall reaction rates are much lower than that of the Mo-Fe enzyme, limiting the application of these variants in bioproduction systems. One option that is an interesting strategy for H2 production, to increase the electron flux into H2, is cultivation in the absence of N2, since nitrogenase turnover continues, but now the electron flux goes totally toward hydrogen evolution. In addition, the growth arrest caused by the nutrient limitation is of interest as this decouples hydrogen evolution from biomass production, therefore potentially leaving more energy available for H2 production (Benemann and Weare, 1974). Even so, the expression of an oxygen-sensitive enzyme in an O2 rich milieu is counter productive. To overcome this problem, temporal separation between N2 fixation and photosynthesis can be used, where during the day the photosynthetic machinery works toward the carbon fixation, which then can be consumed to power nitrogenase and consequently proton reduction. Interestingly, the peak of hydrogen production in indirect biophotolysis occurs when the cell is reilluminated, possibly due to a burst in ATP synthesis before the oxygen formed by PSII (Figure 22.1) reaches a toxic level for the nitrogenase. Heterocyst forming species on the other hand can perform direct biophotolysis by carrying out nitrogen fixation in the differentiated cell during the day. The Hydrogen Bioproduction As discussed above, nonbiological production of hydrogen is energy intensive and often associated with the production of greenhouse gas. Biologically, hydrogen can be produced by a variety of microorganisms possessing one of several different hydrogenases. In the cyanobacteria, enzymes involved in hydrogen metabolism belong to one of the three families discussed above: Hox, Hup or nitrogenase. The uptake hydrogenase (Hup) is not useful for hydrogen evolution since it is poised to work unidirectionally, toward the recycling of H2 into Hþ. When hydrogen is produced by a heterotrophic organism, an organic carbon source (ultimately derived from photosynthesis) is used to provide protons and chemical energy to fuel hydrogen evolution. Ironically, this is also true for cyanobacteria carrying out direct or indirect biophotolysis, at least on the molecular level. As discussed above, a complete photosynthetic apparatus uses water as proton donor, releasing molecular oxygen (Figure 22.1). Thus, the high sensitivity of hydrogenase to this gas dictates that both reactions cannot occur in the same place at the same time. The solution found by Nature was the most obvious one: changing the timing (indirect biophotolysis) or the space (direct biophotolysis). In indirect biophotolysis the cell uses the chemical energy stored through the capture of sunlight, as
396 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY NADPH and ATP (Figure 22.1), to fix CO2 into organic compounds. These energy reserve molecules are then consumed in the dark to drive cellular metabolism, including nitrogen fixation by nitrogenase. The separation of these reactions occurs naturally in several cyanobacterial species by circadian control and in these strains dark hydrogen production by either nitrogenase or the bidirectional hydrogenase is frequently reported (Prabaharan et al., 2010; Troshina et al., 2002). An interesting characteristic found in many of these strains is a burst of hydrogen production when cells are reilluminated. This phenotype was characterized as a function of the bidirectional hydrogenase and hydrogen production ceases quickly as the O2 produced by photosystem II (Figure 22.1) accumulates in the cell, inactivating hydrogenase The production of H2 is thought to serve as an electron sink, helping the cell return to the proper redox state for carrying out the light reactions. In practice, indirect biophotolysis could possibly be done as a large-scale production using a two-stage cultivation system. In a first stage, the cells are cultivated in the light and biomass is formed through photosynthesis. When the desired cell concentration is achieved and the cells have stored enough fixed carbon, a dark anaerobic cultivation could follow, favoring proton reduction to hydrogen by hydrogenase. Thus, the water-splitting reaction is separated from H2 production in time and space. This system has being already demonstrated, where nitrogen limitation was also used to induce glycogen accumulation and increase hydrogen production yield in the second stage through the nitrogenase enzyme (Huesemann et al., 2009). In a similar approach with Synechococcus sp., the carbon accumulated in the first stage was converted into hydrogen in a second stage by a [NiFe] hydrogenase (McNeely et al., 2010). Even so, some continual synthesis of nitrogenase is necessary to replace oxygen-damaged nitrogenase (Murry et al., 1983). As discussed above, since heterocysts lack a complete photosynthetic apparatus, the necessary reductant is derived from fixed carbon imported from the neighboring vegetative cells through specialized interconnecting pore structures (Mariscal and Flores, 2010). The imported sugar is sucrose (Lopez-Igual et al., 2010) and it is metabolized though the oxidative pentose pathway (Summers et al., 1995) Thus, hydrogen production by heterocysts is essentially indirect biophotolysis on a microscopic scale, and since the energy captured by photosynthesis is first stored as fixed carbon, the maximal possible theoretical conversion efficiencies are reduced. However, this system has been attractive due to its inherent robustness and has been studied for almost four decades (Benemann and Weare, 1974). Very reasonable conversion efficiencies, sustained for days to weeks, were achieved in early studies using nitrogen-limited cultures. Under laboratory conditions where higher efficiencies can be expected, conversion efficiencies (total incident light energy to free energy of hydrogen produced) were shown to be 0.4% (Weissman and Benemann, 1977). Cultures incubated under natural sunlight (Figure 22.3) were able to attain an average conversion efficiency of 0.1% (Miyamoto et al., 1979a). Remarkably, even though there have been a large number of studies since, very little improvement in yields H2 PRODUCTION BY HETEROCYSTOUS CYANOBACTERIA Solar energy capture and hydrogen evolution by some filamentous cyanobacterial strains proceeds naturally in the presence of oxygen by confining the oxygen-sensitive processes to the heterocyst, a cell type that emerged shortly after the oxygenation of the earth’s atmosphere in what has been called the Oxygen Catastrophe or Great Oxidation Event 2.6 billion years ago (Kumar et al., 2010; Mariscal and Flores, 2010). In this case the evolved hydrogen is produced by nitrogenase whose expression is restricted to the heterocyst under normal aerobic conditions (Murry et al., 1984). A number of mechanisms are employed to protect nitrogenase from oxygen damage; heterocysts lack photosystem II so do not produce oxygen, gas diffusion into the heterocyst is restricted by a unique cell wall structure, and heterocysts possess a very active membrane-bound respiratory system that consumes trace amounts of entering oxygen. FIGURE 22.3 Tubular photobioreactors operating under “airlift”conditions were used to demonstrate prolonged (over 30 days) simultaneous oxygen and hydrogen evolution by nitrogen-limited cultures of the heterocystous cyanobacterium, Anabaena cylindrica. Source: Miyamoto et al 1979d. (For color version of this figure, the reader is referred to the online version of this book.)
CYANOBACTERIA AS A PRODUCTION SYSTEM FOR BIOFUELS: CURRENT STATUS has been obtained. Thus, recent reports of conversion efficiencies found z0.7% under laboratory conditions (Berberoglu, 2008; Sakurai and Masukawa, 2007; Yoon et al., 2006) and 0.03e0.1% with natural sunlight (Sakurai and Masukawa, 2007; Tsygankov et al., 2002). Similar low efficiencies have been found with thermophilic strains, which at least have the possible advantage of requiring less cooling (Miyamoto et al., 1979b,c). There should be room for improvement as theoretical efficiencies with this nitrogenase-based system have been calculated to be around 4.6% (Hallenbeck, 2011). Since observed conversion efficiencies are lower than predicted, different strategies might be employed in order to improve overall performance, which is critically important since light conversion efficiencies directly impact on the photobioreactor footprint (doubling efficiency should halve the required surface area for the same amount of fuel production). For one thing, genetic engineering could be applied to optimizing the size of the photosynthetic antenna, since part of the reduction in efficiency is thought to be due to inefficient use of light energy at high intensities where more photons are captured than can be used and the excess energy is wasted. Another point that could be addressed is the hydrogen producing catalyst. Since half of the photon requirement is needed to provide ATP to nitrogenase action, replacing it with a hydrogenase, which does not require ATP for proton reduction, should in principle have an energy sparing effect. In a recent attempt to verify this, the [FeFe] hydrogenase from Shewanella oneidensis was expressed in Anabaena sp. under the control of a heterocyst-specific promoter with the required maturation genes (Gartner et al., 2012). Although it could be shown that active hydrogenase was made under the proper conditions, the increase in hydrogen production above the levels due to the coexisting nitrogenase was disappointingly small. Of course, under these conditions the two enzymes compete for the reductant; the true test would be to do this in a strain lacking nitrogenase activity. Finally, it might in principle be a possible way to increase hydrogen production by increasing heterocyst frequency. However, heterocyst frequency might already be close to optimal since even in long-term studies the H2/O2 ratio is close to the desired stoichiometry of two, what one would expect for optimal coupling between oxygen-generating photosynthesis in the vegetative cells and hydrogen production by heterocysts. H2 PRODUCTION BY NONHETEROCYSTOUS CYANOBACTERIA Although the heterocyst/nitrogenase-based system has been the most studied, some other known cyanobacterial hydrogen-producing reactions could potentially be used for biological hydrogen production. These include the unicellular and nonheterocystous 397 filamentous cyanobacteria, which possess nitrogenase and are able to fix nitrogen in nature. Two strategies are employed to avoid oxygen inhibition. In some unicellular species, oxygen evolution and nitrogen fixation (or hydrogen production) are separated in time since photosynthesis and nitrogen fixation are under circadian control with photosynthesis taking place during the day and nitrogen fixation being maximal during the night period. The filamentous cyanobacterium Trichodesmium uses a strategy of spatial segregation where nitrogen fixation occurs in cells located in the middle of the bundle carrying out the oxygen-sensitive nitrogenase reactions and the others carrying out the normal photosynthetic reactions (Berman-Frank, 2001). The unicellular cyanobacterium Cyanothece has been the subject of a number of recent studies demonstrating prolonged hydrogen production in the light mediated by nitrogenase. In one study, considerable hydrogen production (up to 465 mmol per milligram of chlorophyll per hour) was shown, the growth conditions were very stringent and hydrogen production was only observed when the culture was submitted to nitrogen starvation, sparged with argon to remove any oxygen formed through photosynthesis, supplemented with glycerol and cultivated under low light (Bandyopadhyay et al., 2010; Min and Sherman, 2010). Glycerol, in addition to serving as a possible additional energy source to support nitrogenase activity, appears to release nitrogenase from diurnal control (Aryal et al., 2013). Another recent study found appreciable hydrogen and oxygen production with nitrogen-depleted cultures that were incubated under continuous illumination (Melnicki et al., 2012). Light saturation curves and photosynthesis inhibition studies indicate that the hydrogen is evolved indirectly from the fixed carbon produced through photosynthesis. Here again, the requirements for continuous illumination (it can hardly be energetically positive to produce hydrogen using artificial illumination) and for argon sparging raise serious hurdles to practicality. Thus, although a nice proof of principle, such a system would hardly be economically viable. Many cyanobacteria also possess Hox, a soluble reduced nicotinamide adenine dinucleotide (NADH)linked [NiFe] hydrogenase. This reversible hydrogenase is capable of hydrogen evolution, in particular when dark-adapted cells are reilluminated (Schwarz et al., 2010). As discussed above, this forms an electron valve, readjusting the poise of the photosynthetic apparatus, but activity is quickly inhibited with renewed oxygen evolution. A recent survey showed that a diversity of cyanobacteria contains this enzyme and that there is great variability in both the amounts of hydrogen made by this enzyme and the pattern of hydrogen evolution (Kothari et al., 2012). This enzyme is also responsible for evolution during dark fermentation of endogenous
398 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY reserves, principally glycogen, and hydrogen production by this pathway can be enhanced through lowering of the hydrogen partial pressure (Ananyev et al., 2012). At least in Synechocystis, hydrogen production by Hox can be increased by eliminating the master regulator AbrB2, which normally represses synthesis of Hox (Dutheil et al., 2012; Leplat et al., 2013). In a recent attempt to increase hydrogen production, heterologous expression of the [FeFe] hydrogenase from Clostridium acetobutylicum was carried out in the non-nitrogen-fixing cyanobacterium Synechococcus (Ducat et al., 2011). Active hydrogenase was formed under proper conditions, but in vivo light-driven hydrogen production from this system was significant only when the cultures were incubated under an inert atmosphere and oxygenic photosynthesis was completely inhibited. carbon, to ethanol, pyruvate decarboxylase (pdc) and alcohol dehydrogenase (adh). This alone is sufficient to produce low millimolar levels of ethanol in the medium upon prolonged (5e10 days) incubation and growth. Further increases, obviously necessary for practical production, have been achieved through a variety of means, including better transcriptional control and further metabolic engineering. Most of this development work is being done at private enterprise laboratories, but a recent published report (Gao et al., 2012) shows that impressive increases in yields can be achieved by integrating a foreign pdc and a native adh into the genome of Synechocystis and abolishing carbon flux into polyhydroxybutyrate synthesis. Ethanol The chemical industry relies on simpler molecules to build complex compounds, which are used in a variety of applications. Among these organic compounds, ethylene is the building block with the highest demand, being used to manufacture “everyday” products such as polyethylene terephthalate (PET bottles), polyester, antifreeze, and others (Ungerer et al., 2012). Ethylene (or ethene) is the second simplest unsaturated hydrocarbon, it consists of two carbons, with a double bond, and four hydrogens (H2C]CH2). It is one of the products of pyrolysis, and has been used as a fuel since the early nineteenth century as one of constituents of the gaseous fuel in gas lamps. Ethylene is the most important compound in the chemical industry in terms of market volume, it has a heat of combustion higher than that of gasoline or diesel and can be used as a transportation fuel or to produce electric energy in stationary plants. Currently, ethylene is a petroleum derivative produced through steam cracking. It reached a production of 100 million metric tones in 2005, accounting for 30% of all petrochemical commodities (McCoy et al., 2006; Saini and Sigman, 2012). The fluctuation in crude oil prices over the last few years (EIADOE, 2012), the imminent threat of peak oil (Nashawi et al., 2010), and the existence of biological pathways for its production coupled with the ease of harvesting a gas like ethylene, make this chemical a good target for the development of a sustainable biological production system. The most common occurrence of ethylene in nature is as a hormone found in vascular plants, where it is associated with many effects such as defoliation, responses to temperature stress, mechanical injury, and for promoting fruit ripening (Abeles, 1972). In addition to vascular plants, many other plants and algae have been shown to be able to produce ethylene and even if it not found in animals, this gaseous hormone has been shown to induce regulatory responses in invertebrate and mammalian cells (Perovic et al., 2001). While hydrogen production, or at least direct biophotolysis, can be driven directly by photosynthesis, all other biofuels must use the capacity of cyanobacteria to drive carbon dioxide fixation with photosynthetically derived energy, ATP and reductant. However, once fixed by the Calvin-Benson-Bassham cycle, the newly recycled carbon can be converted to useful biofuels through the introduction of novel (to cyanobacteria) metabolic pathways (Angermayr et al., 2009). The first such cyanobacterial-derived biofuel that was demonstrated was ethanol (Deng and Coleman, 1999; Dexter and Fu, 2009), and its production is the only cyanobacterialproduced biofuel under active investigation and commercial development (Algenol Biofuels: http://www. algenolbiofuels.com/). Algenol Biofuels is presently claiming production at “around $1.00 per gallon using sunlight, carbon dioxide and saltwater at production levels above 9000 gallons of ethanol per acre per year”. At an average solar insolation for Florida of 19.8 MJ/day and since ethanol has a higher heating value of 29.7 MJ/ kg, this translates to a claim of a very impressive 2.8% conversion efficiency. Now another company, Joule Unlimited (http://www.jouleunlimited.com/), has stepÒ ped into the picture, offering to sell SunFlow-E through its fuel company, Joule Fuels. Their process uses genetically modified thermophilic cyanobacterium containing Moorella alcohol dehydrogenase, and their Web site claims are even more spectacular with targets of up to 25,000 gallons per acre (7.8% conversion efficiency) and $0.60 per gallon at full-scale commercial production. Cyanobacteria can naturally produce relatively minute amounts of ethanol so at the simplest level, creating a cyanobacterium that produces higher levels of ethanol involves boosting flux through the ethanol pathway through the introduction of the key enzymes for conversion of pyruvate, generated by glycolysis of the fixed Ethylene
CYANOBACTERIA AS A PRODUCTION SYSTEM FOR BIOFUELS: CURRENT STATUS The most common biosynthetic pathway for ethylene production is the Yang cycle that occurs in plants, where it is produced from methionine in a three-step reaction, having S-adenosylmethionine (AdoMet) and 1-aminocyclopropane-1-carboxylic acid (ACC) as precursors. However, the cellular response to this hormone occurs at very low concentrations, a characteristic that, together with the fast and easy diffusion of this gas into plant tissues, makes the conversion of AdoMet to ACC, catalyzed by ACC synthase, and from ACC to ethylene (ACC oxidase) a tightly regulated process. Both enzymes are multigenic with differential regulation through distinct promoters and operators for groups of genes of the same enzyme (Nakatsuka et al., 1998). The methionine used in this pathway is recycled through the Yang cycle (Taiz and Zeiger, 2002; Wang et al., 2002). Microbial Production of Ethylene The great influence of this gaseous hormone on different plant organs made it an interesting target for pathogens. Indeed, the mold Penicillium digitatum has been known to produce ethylene since the mid-1950s (Wang et al., 1962) and its production was shown for prokaryotic plant pathogens in the early 1960s (Freebairn and Buddenhagen, 1964). Not surprisingly, the pathways found in these microorganisms are not analogous to the one in plants. So far, two distinct routes for ethylene production have been described in microbes: a 2-oxoglutarate-dependent pathway and the 2-keto4-methyl-thiobutyric acid (KMBA) pathway (Nagahama et al., 1991, 1992). The latter is the most common among microorganisms, composed of a series of chemical and enzymatic reactions, by which only trace amounts of ethylene are usually produced (Ogawa et al., 1990). The former pathway has been found to be more efficient, with 2-oxoglutarate being used as substrate in a singlestep reaction by the ethylene-forming enzyme (EFE). This pathway has been found in several different microorganisms, including P. digitatum, Chaetomium globosum, Phycomyces nitens, Fusarium oxysporum, and in different pathovars of Pseudomonas syringae, where a comparison study found the pv. phaseolicola to be the most efficient ethylene-producing strain (Weingart et al., 1999). This enzyme catalyzes simultaneously two reactions (Fukuda et al., 1992b): 2  oxoglutarate4ethylene þ 3CO2 þ H2 O (22.2) 2  oxoglutarate þ L  arginine þ O2 4succinate þCO2 þ guanidine þ ðSÞ  1  pyrroline  5 carboxylate þ H2 O (22.3) 399 These reactions are rather interesting as they keep the tricarboxylic acid (TCA) cycle closed through a shortcut, converting 2-oxoglutarate directly into succinate with the formation of ethylene “as a by-product” (Figure 22.2), and therefore substituting for the steps catalyzed by 2-oxoglutarate dehydrogenase and succinyl-CoA synthetase (Figure 22.2). The original two-step reaction between 2-oxoglutarate and succinate generates one NADH and one guanosine triphosphate, which are not produced by EFE. Thus, competition of the two pathways for substrate 2-oxoglutarate would lower the formation of NADH. Since NADH also has a role as an inhibitor for four enzymes associated with the TCA cycle, pyruvate dehydrogenase, isocitrate dehydrogenase, 2-oxoglutarate dehydrogenase, and citrate synthase (Figure 22.2), this could potentially upregulate the reactions performed by these enzymes, from the decarboxylation of pyruvate to 2-oxoglutarate. The last is a direct and indirect substrate to the EFE, directly to generate ethylene (Eqn (22.2)) and indirectly, as it is also a substrate for the synthesis of arginine, required for the simultaneous reaction of this enzyme (Eqn (22.3) and Figure 22.2). Bioproduction of Ethylene Using efe When producing any molecule of commercial interest with a microorganism, prospecting for the key gene is as important as the choice of the host to be used. The evolutionary convergence of ethylene production is highlighted by the three pathways delineated above: Yang cycle, KMBA and 2-oxoglutarate, with the last having been found to be the most efficient when overproduction is desired. The evolutionary radiation of the mobile plasmid encoding the efe gene among the different species and strains might have produced a naturally optimized gene that could be used in commercial production. A comparison between 20 P. syringae strains revealed a high amino acid sequence similarity between five pathovars, with P. syringae pv. phaseolicola PK2 being the most efficient, giving a twofold higher production of ethylene (Weingart et al., 1999). However, these variations are likely to be due to differences in regulation as the sequence of amino acids of efe of these five strains differs by only one codon. Using the efe gene encoded by an indigenous plasmid from P. syringae pv. phaseolicola PK2, ethylene production was reported in E. coli with a tenfold increase when compared to the original strain, P. syringae (Fukuda et al., 1992a; Ishihara et al., 1995), showing that the efe gene alone was sufficient for ethylene production. When a high-copy-number plasmid containing efe was transconjugated into Pseudomonas putida and P. syringae, ethylene production was increased, but surprisingly, production was 27- and 8-fold higher, respectively, than the wild type, whereas the amount of protein produced in the cloned P. syringae was 20-fold higher
400 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY (Ishihara et al., 1996), suggesting the presence of a posttranscription regulatory system. Using cellulose as substrate, Tao et al. showed the production of ethylene in Trichoderma viride through the heterologous expression of efe from P. syringae pv. glycinea. Thus, the use of agriculture wastes as substrate for ethylene production was proved to be feasible, but the recombinant filamentous fungus produced only very small amounts of ethylene (Tao et al., 2008). So far, cyanobacteria have been shown to be the best model for the bioproduction of ethylene. Of course, many barriers still have to be crossed and commercial production is far from reality at present, but the last few years have seen encouraging reports where the productivity was increased several fold without compromising cell fitness, suggesting that the true production limit might be much higher. The efe (EFE) from P. syringae was originally cloned into Synechococcus elongatus PCC 7942 (Fukuda et al., 1994; Sakai et al., 1997; Takahama et al., 2003). The first problem area, the production of only trace amounts of ethylene by the transformants (Fukuda et al., 1994), was later shown to be due to the nature of the promoter used. A systematic evaluation of different promoters showed the psbA1 promoter is more efficient for efe expression than those (lac and efe) previously used in other reports, achieving production rates up to 240 nl/ml h or 451 nl/ml h OD730 (Takahama et al., 2003). However, these recombinants showed high genetic instability. Sequencing of the heterologous gene from mutants that had ceased to produce ethylene showed punctual mutations at a defined sequence of five nucleotides, suggested to be a possible hot-spot site for spontaneous mutagenesis (Takahama et al., 2003). Nevertheless, active ethylene-producing strains showed signs of metabolic stress, evidenced by their yellow-green color. When these strains had ceased ethylene production due to spontaneous mutation of efe (genetic instability), they recovered the normal blue-green phenotype. In another strategy (Ungerer et al., 2012), Synechocystis sp. PCC 6803 was used as model organism. Toxicity to ethylene was tested, efe was codon optimized and artificially synthesized, eliminating the bases at the putative mutational hot spot by conservative substitution. As well, efe was placed under the control of the psbA1 promoter. A semicontinuous culture using a clone containing two copies of efe was sustained over a three-week period, reaching a constant production of 3100 nl/ml h, compared to the previous result of 240 nl/ml h (Takahama et al., 2003). The peak of the specific productivities was 380 nl/ml h OD730 for one efe copy and 580 nl/ mL h OD730 for two copies, respectively, and when in semicontinuous culture, the average rate was 200 nl/ ml h OD730. The additional copy of the efe gene presented some production improvement when compared with the previous work from Takahama et al., (451 nl/ ml h OD730 compared to 580 nl/ml h OD730) but the real advance for the field can be seen from the healthy state of the culture. The growth rate, the color of the culture and the growth curve were the same for wild type and the mutants containing one or two copies of efe. This shows that there is no toxicity either by the product or by the metabolic route used to produce ethylene. In addition to the zero toxicity, the release of five carbons per ethylene formed does not seem to present a burden to the cell, as shown by the growth pattern of the single and double mutant when compared with the wild type. Nevertheless, the metabolic consequences to the cell of a higher rate of ethylene production are unknown and a physiological approach would help to understand how far ethylene production can be pushed and what to target to improve the final yield. Isoprene As with ethylene, isoprene is a medium-value biochemical that is produced through steam cracking of oil. It is actually an important by-product of ethylene production and is almost entirely used for production of a synthetic substitute for natural rubber. It is also naturally produced by many plants as a heat stress response, where it was shown to increase the stability of photosynthetic membranes at high temperatures (Sharkey et al., 2001). It can represent as much as 2% of all carbon fixed by oak leaves at a temperature of 30  C (Sharkey, 1996), showing the physiological importance of this compound. The enzyme isoprene synthase (ispS) was shown to produce isoprene in plants, converting one of the products of the methylerythritol phosphate (MEP) pathway, dimethylallyl-diphosphate (DMADP), into isoprene (Silver and Fall, 1991; Silver and Fall, 1995). Prokaryotes were suggested to be able to produce isoprene after reports of the detection of this compound in the headspace of culture broth on many species (Kuzma et al., 1995), with emphasis on Bacillus subtilis. Not surprisingly, sequence analysis of bacterial genome could not identify any gene homologous to the ispS. found in plants (Withers et al., 2007). So far, functional genomics has also failed to identify the pathway for isoprene production in prokaryotes. Sequence-independent methods showed that 19,000 E. coli clones transformed with DNA fragments from B. subtilis in an environment where DMADP and IPP (isopentenyl pyrophosphate) levels were selectively toxic, showed that no single enzyme was sufficient to convert DMADP to isoprene, where the few clones that managed to survive, preferably converted it to a prenyl alcohol (Withers et al., 2007). As all isoprenoids are thought to be solely produced from DMADP and IPP (Xue and Ahring, 2011), the conversion of the metabolites involved in MEP or mevalonate pathway
CYANOBACTERIA AS A PRODUCTION SYSTEM FOR BIOFUELS: CURRENT STATUS 401 FIGURE 22.4 Pathway alternatives for n-butanol bioproduction. The alcohol n-butanol is naturally produced in different microorganisms in small quantities, where it can be synthesized either through the CoA-dependent pathway or the keto acids pathway. (For color version of this figure, the reader is referred to the online version of this book.) to isoprene in bacteria could be a phenotype derived from convergent evolution using a multistep reaction diverged from those pathways (Izumikawa et al., 2010; Withers et al., 2007; Xue and Ahring, 2011). Bioproduction of isoprene is feasible and has already been demonstrated in E. coli expressing heterologous ispS (Miller et al., 2001; Zhao et al., 2011). Of course productivity is an issue and different strategies were tried to increase isoprene production. Simultaneous expression of heterologous enzymes involved in MEP or mevalonate pathways was shown to be effective in both cases (Yang et al., 2012; Zhao et al., 2011). Julsing et al. also showed that the individual expression of the genes encoding enzymes involved in the MEP pathway did not affect isoprene production with the exception of the dxs gene, encoding the enzyme that catalyzes the first reaction of the MEP pathway, which significantly improved isoprene production (Julsing et al., 2007). Cyanobacteria produce DMAPP through the MEP pathway for secondary metabolites and, albeit with no natural production of isoprene being reported yet, transformation and expression of heterologous ispS were shown to be sufficient for production of isoprene. Lindberg et al. reported isoprene production using Synechocystis sp. PCC 6803 as a model organism harboring ispS from Pueraria montana (Kudzu) (Lindberg et al., 2010). The transgene was inserted at the psbA2 locus and mutants did not show any disturbance in growth when compared to the wild type. This was a well-achieved proof of concept, and the low productivity reported, 50 mg per gram of dry cell weight per day, can be much improved through metabolic engineering. However, the use of cyanobacteria to produce isoprene has issues different from metabolic yield: to develop a production system of a molecule with a half-life of only a couple hours in the presence of light is particularly challenging in a photosynthetic organism. To overcome this issue, the development of special photobioreactors is made in parallel to the molecular research, where the properties of isoprene as a volatile hydrophobic compound, easily separated from a culture broth and concentrating in the headspace, are exploited (Lindblad et al., 2012). The production of a gas in microorganisms is an interesting strategy because one does not need to harvest the cells, the product is concentrated in the gaseous phase of the culture. However, the cultivation techniques and the purification of this gas from a complex mixture represents an important step in the production chain and, as shown in this case, should develop together. Butyraldehyde and Butanol Butanol has many desirable properties as a fuel and thus is a suitable target for modification of
402 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY cyanobacteria. In fact, as a fuel it is superior to ethanol, being less corrosive and less volatile. Thus, it can easily be mixed with hydrocarbon-based fuels and used in the same infrastructure. A number of recent studies have shown that engineered cyanobacteria can in fact make surprisingly high levels of this compound, at rates that in fact surpass published rates for ethanol production by engineered cyanobacteria. Since cyanobacteria produce this fuel directly through photosynthetically driven CO2 fixation, it is appropriate to compare the productivity per area of this process, as presently described, with that required for other biofuels, be it growing corn to produce the necessary sugars, or growing algae to produce biodiesel. Such a comparison shows that butanol production by cyanobacteria could be much better than making fuels from corn and very comparable to making biodiesel from microalgae (Sheehan, 2009). Many cyanobacteria are capable of producing volatile compounds, including higher alcohols, but the natural production levels are miniscule (Hasegawa et al., 2012). In order to make fuel molecules at significant quantities, new pathways must be introduced as well as changes made to the native metabolic pathways. Studies on creating heterotrophic bacterial strains capable of producing butanol demonstrated that two possible routes were useful: the 2-ketoacid pathway, normally involved in amino acid biosynthesis, and the acetyl-CoA pathway, found in organisms such as Clostridium that naturally produce butanol during fermentation (Figure 22.4). The first successful attempt in this direction was to engineer S. elongates to produce isobutyraldehyde through the 2-ketoacid pathway (Atsumi et al., 2009). Isobutyraldehyde is a precursor for isobutanol and other chemicals of interest and has the advantage of being highly volatile, easing its recovery from the culture broth thus removing product inhibition. The strategy applied consisted of boosting carbon flux through the pathway from pyruvate to 2-ketoisovalerate by integration into the genome of three foreign genes, alsS, ilvC and ilvD, catalyzing these steps, as well as kivd from Lactococcus lactis, the gene encoding the ketoacid decarboxylase enzyme that converts 2-ketoisovalerate to isobutyraldehyde. Overall carbon flux was then increased by integrating an additional copy of Rubisco (rbcLS) and the resulting strain produced 6230 mg isobutyraldehyde per liter per hour , a production rate that is higher than any other fuel molecule made by cyanobacteria to date. Additionally, it was demonstrated that isobutanol could be formed if a foreign alcohol dehydrogenase (YqhD from E. coli) was introduced, but titers were lower, presumably due to product inhibition. However, the isomer that is made by the 2-ketoisovalerate pathway is isobutanol, a fuel additive, but not nearly as desirable in itself as a fuel as n-butanol, the product of the acetyl-CoA pathway or the 2-ketobutyrate pathway (Figure 22.4). Metabolic engineering was used to create an n-butanol-producing strain of S. elongatus by introducing the hbd, crt, and adhE2 genes from C. acetobutylicum, the ter gene from Treponema denticola, and the atoB gene (instead of thl) from E. coli (Lan and Liao, 2011). However, n-butanol was only produced by this strain under anaerobic conditions, either in the light when photosystem II was inhibited by DCMU, or in the dark, which gave the highest production, a meager 20.8 mg/L/h. It was suggested that anaerobic conditions were necessary since some of the enzymes introduced are oxygen sensitive, severely limiting its usefulness. On the other hand, metabolic fluxes are obviously different during dark metabolism than during photosynthesis and the difference could be in the supply of a key metabolite. In line with this, in a more recent attempt to create an n-butanol-producing strain, flux through acetyl-CoA was increased by substituting an irreversible ATP hydrolysis step leading to the formation of acetoacetyl CoA (Lan and Liao, 2011). Other improvements consisted of substituting NADPH-requiring enzymes for NADH enzymes. With these changes, it was possible to demonstrate light-dependent n-butanol production, but at 62.5 mg/L/h this is well below (by a factor of 100) the initial promising results with butyraldehyde. This system would need very significant improvement before it could be considered for practical biofuel production. Photosynthetic Production of Aliphatic Alcohols and Alkanes For many fuel purposes, alkanes are more desirable than the other biofuels already discussed. For example, jet fuel standards (Jet-A or JP-8) demand a fuel with high energy density, low viscosity, low freezing point and good physical-chemical compatibility. These criteria cannot be met with fuels such as ethanol or fatty acid methyl esters, biodiesel. Being able to directly make alkanes would have a great payoff as these biofuels are “drop-in” fuels, able to directly substitute for presently used petroleum-based fuels as they could be used with existing infrastructure and would require no engine modification, etc. Cyanobacteria, like some other bacteria, have long been recognized as being able to synthesize at least very small quantities of alkanes, which in fact can serve as a biogeochemical marker for their presence in the past (Han et al., 1968; Winters et al., 1969). This was taken advantage of in a recent demonstration of the heterotrophic production of alkanes using a modified E. coli that expressed the alkane biosynthetic pathway from a
REFERENCES cyanobacterium, consisting of an acyl-carrier protein reductase, which produces a fatty aldehyde, and an aldehyde decarbonylase (Schirmer et al., 2010). This allowed the production and secretion of a variety of C13eC17 alkanes and alkenes. Of course it would be desirable to actually do this in a cyanobacterium, and one study examined this through the heterologous expression of fatty acyl-CoA reductase in Synechocystis (Tan et al., 2011), which allowed the production of small quantities of aliphatic alcohols. The acc genes, encoding acetyl-CoA carboxylase (ACCase), which catalyses what is believed to be the rate-limiting step of fatty acid biosynthesis, were introduced into the genome in hopes of boosting alkane production, but only insignificant quantities were made. Further work is required to demonstrate significant alkane synthesis by a cyanobacterium. However, it may prove difficult to greatly boost alkane synthesis in this oxygen-evolving organism as the critical enzyme, aldehyde decarbonylase, has recently been shown to be a di-iron enzyme with an unusual mechanism that requires anaerobic conditions for full activity (Das et al., 2011). CONCLUSION AND OUTLOOK As discussed in this chapter, recent studies have shown the great promise for biofuels production by cyanobacteria. Unique among possible biofuel producers, cyanobacteria combine the attributes of being able to carry out photosynthesis-driven carbon dioxide fixation and to be easily manipulated genetically. The next few years should see advances in increasing the production rates and titers of the different demonstrated biofuels as well as perhaps the widening of the spectrum of possible biofuels. Nevertheless, for cyanobacterial systems to live up to their potential, a number of serious hurdles must be overcome. These include the development of reliable methods of stable cyanobacterial mass culture at high levels of productivity and the demonstration of costeffective harvesting strategies. Harvesting presents a real dilemma no matter what the biofuel. If the biofuel is contained within the cell, then the biomass has to be removed from the culture medium, of which it is less than 1% by weight. If the biofuel is an excreted liquid, then this will necessarily be quite dilute and require substantial concentration. If the biofuel is a gaseous product, the culture will have to be enclosed in airtight transparent material at a substantial cost given the large surface areas that would be required. Of course, the payoff to solving these problems would be enormous and this is likely to inspire future research and development in this area. 403 References Abeles, A.L., 1972. Biochemical pathway of stress-induced ethylene. Plant Physiol. 50, 496e498. Ananyev, G.M., Skizim, N.J., Dismukes, G.C., 2012. Enhancing biological hydrogen production from cyanobacteria by removal of excreted products. J. Biotechnol. 162, 97e104. Angermayr, S.A., Hellingwerf, K.J., Lindblad, P., Teixeira de Mattos, M.J., 2009. Energy biotechnology with cyanobacteria. Curr. Opin. Biotechnol. 20, 257e263. NCBI: http://www.ncbi.nlm.nih.gov. Aryal, U.K., Callister, S.J., Mishra, S., Zhang, X., Shutthanandan, J.I., Angel, T.E., Shukla, A.K., Monroe, M.E., Moore, R.J., Koppenaal, D.W., Smith, R.D., Sherman, L., 2013. Proteome analyses of strains ATCC 51142 and PCC 7822 of the diazotrophic cyanobacterium Cyanothece sp. under culture conditions resulting in enhanced H2 production. Appl. Environ. Microbiol. 79, 1070e1077. Atsumi, S., Higashide, W., Liao, J.C., 2009. Direct photosynthetic recycling of carbon dioxide to isobutyraldehyde. Nat. Biotechnol. 27, 1177e1180. Ball, P., 2007. Burning Water and Other Myths. news@nature. Bandyopadhyay, A., Stöckel, J.S.O., Min, H., Sherman, L.A., Pakrasi, H.B., 2010. High rates of photobiological H2 production by a cyanobacterium under aerobic conditions. Nat. Commun. 1, 139e147. Benemann, J.R., Weare, N.M., 1974. Hydrogen evolution by nitrogenfixing Anabaena cylindrical cultures. Science 184, 174e175. Berberoglu, H., 2008. Effect of nutrient media on photobiological hydrogen production by Anabaena variabilis ATCC 29413. Int. J. Hydrogen Energy 33, 1172e1184. Berman-Frank, I., 2001. Segregation of nitrogen fixation and oxygenic photosynthesis in the marine cyanobacterium Trichodesmium. Science 294, 1534e1537. Bothe, H., Schmitz, O., Yates, M.G., Newton, W.E., 2010. Nitrogen fixation and hydrogen metabolism in cyanobacteria. Microbiol. Mol. Biol. Rev. 74, 529e551. BP, 2012. BP Statistical Review of World Energy June 2012. BP. p. 48. http://bp.com/statisticalreview. Das, D., Eser, B.E., Han, J., Sciore, A., Marsh, E.N.G., 2011. OxygenIndependent Decarbonylation of Aldehydes by Cyanobacterial Aldehyde Decarbonylase: A New Reaction of Diiron Enzymes. Angewandte Chemie International Edition 50 (31), 7148e7152. Deng, M.D., Coleman, J.R., 1999. Ethanol synthesis by genetic engineering in cyanobacteria. Appl. Environ. Microbiol. 65, 523e528. Dexter, J., Fu, P., 2009. Metabolic engineering of cyanobacteria for ethanol production. Energy Environ. Sci. 2, 857. Ducat, D.C., Way, J.C., Silver, P.A., 2011. Engineering cyanobacteria to generate high-value products. Trends. Biotechnol. 29, 95e103. Dutheil, J., Saenkham, P., Sakr, S., Leplat, C., Ortega-Ramos, M., Bottin, H., Cournac, L., Cassier-Chauvat, C., Chauvat, F., 2012. The AbrB2 autorepressor, expressed from an atypical promoter, represses the hydrogenase operon to regulate hydrogen production in Synechocystis strain PCC6803. J. Bacteriol. 194, 5423e5433. EIADOE, 2012. Primary energy consumption by source and sector, 2011 384(null). In: EIA - US Energy Information Administration, vol. 1. http://www.eia.gov/aer. Elhai, J., Vepritskiy, A., MuroPastor, A.M., Flores, E., Wolk, C.P., 1997. Reduction of conjugal transfer efficiency by three restriction activities of Anabaena sp. strain PCC 7120. J. Bacteriol. 179, 1998e2005. Field, C.B., Behrenfeld, M.J., Randerson, J.T., Falkowski, P., 1998. Primary production of the biosphere: integrating terrestrial and oceanic components. Science 281, 237e240. Flores, E., Muro-Pastor, A.M., Meeks, J.C., 2008. Gene transfer to cyanobacteria in the laboratory and in nature. In: Herrero, A., Flores, E. (Eds.), The Cyanobacteria: Molecular Biology, Genomics and Evolution. Caister Academic Press.
404 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY Freebairn, H.T., Buddenhagen, I.W., 1964. Ethylene production by Pseudomonas solanacearum. Nature 202, 313e314. Fukuda, H., Ogawa, T., Ishihara, K., Fujii, T., Nagahama, K., Omata, T., Inoue, Y., Tanase, S., Morino, Y., 1992a. Molecular-cloning in Escherichia coli, expression, and Nucleotide-sequence of the gene for the ethylene-forming enzyme of Pseudomonas syringae pv phaseolicola-Pk2. Biochem. Biophys. Res. Commun. 188, 826e832. Fukuda, H., Ogawa, T., Tazaki, M., Nagahama, K., Fujii, T., Tanase, S., Morino, Y., 1992b. Two reactions are simultaneously catalyzed by a single enzyme: the arginine-dependent simultaneous formation of two products, ethylene and succinate, from 2-oxoglutarate by an enzyme from Pseudomonas syringae. Biochem. Biophys. Res. Commun. 188, 483e489. Fukuda, H., Sakai, M., Nagahama, K., Fujii, T., Matsuoka, M., Inoue, Y., Ogawa, T., 1994. Heterologous expression of the gene for the ethylene-forming enzyme from Pseudomonas syringae in the cyanobacterium Synechococcus. Biotechnol. Lett. 16, 1e6. Gao, Z., Zhao, H., Li, Z., Tan, X., Lu, X., 2012. Photosynthetic production of ethanol from carbon dioxide in genetically engineered cyanobacteria. Energy Environ. Sci. 5, 9857. Gartner, K., Lechno-Yossef, S., Cornish, A.J., Wolk, C.P., Hegg, E.L., 2012. Expression of Shewanella oneidensis MR-1 [FeFe]-Hydrogenase genes in Anabaena sp. Strain PCC 7120. Appl. Environ. Microbiol. 78, 8579e8586. Ghirardi, M.L., Posewitz, M.C., Maness, P.-C., Dubini, A., Yu, J., Seibert, M., 2007. Hydrogenases and hydrogen photoproduction in oxygenic photosynthetic organisms. Annu. Rev. Plant Biol. 58, 71e91. Golden, S.S., Sherman, L.A., 1983. A hybrid plasmid is a stable cloning vector for the cyanobacterium Anacystis nidulans R2. J. Bacteriol. 155, 966e972. Grigorieva, G., Shestakov, S., 1982. Transformation in the cyanobacterium Synechocystis sp. 6803. FEMS. Microbiol. Lett. 13, 367e370. Guiry, M.D., Guiry, G.M. Algaebase. National University of Ireland, Galway. http://www.algaebase.org. Hallenbeck, P.C., Benemann, J.R., 1978. Characterization and partial purification of reversible hydrogenase of Anabaena cylindrica. FEBS Lett. 94, 261e264. Hallenbeck, P.C., Benemann, J.R., 2002. Biological hydrogen production; fundamentals and limiting processes. Int. J. Hydrogen. Energy. 27, 1185e1193. Hallenbeck, P.C., 2011. Hydrogen production by cyanobacteria. In: Hallenbeck, P.C. (Ed.), Microbial Technologies in Advanced Biofuels Production. Springer US, Boston, MA, pp. 15e28. Han, J., McCarthy, E.D., Hoeven, W.V., Calvin, M., Bradley, W.H., 1968. Organic geochemical studies, II. a preliminary report on the distribution of aliphatic hydrocarbons in algae, in bacteria, and in a recent lake sediment. Proc. Natl. Acad. Sci. USA. 59, 29e33. Hasegawa, M., Nishizawa, A., Tsuji, K., Kimura, S., Harada, K.-I., 2012. Volatile organic compounds derived from 2-Keto-Acid decarboxylase in Microcystis aeruginosa. Microb. Environ. 27, 525e528. Heidorn, T., Camsund, D., Huang, H.-H., Lindberg, P., Oliveira, P., Stensjö, K., Lindblad, P., 2011. Synthetic biology in cyanobacteria: engineering and analyzing novel functions. In: Voigt, C. (Ed.), Methods in Enzymology, Synthetic Biology, Methods for Part/ Device Characterization and Chassis Engineering, Part A, vol. 497. Elsevier academic press inc, 525 B Street, suite 1900, San Diego, CA 92101-4495, USA, pp. 539e579. Huesemann, M.H., Hausmann, T.S., Carter, B.M., Gerschler, J.J., Benemann, J.R., 2009. Hydrogen generation through indirect biophotolysis in batch cultures of the nonheterocystous nitrogenfixing cyanobacterium Plectonema boryanum. Appl. Biochem. Biotechnol. 162, 208e220. Ishihara, K., Matsuoka, M., Inoue, Y., Tanase, S., Ogawa, T., Fukuda, H., 1995. Overexpression and in-vitro reconstitution of the ethylene-forming enzyme from Pseudomonas syringae. J. Ferment. Bioeng. 79, 205e211. Ishihara, K., Matsuoka, M., Ogawa, T., Fukuda, H., 1996. Ethylene production using a broad-host-range plasmid in Pseudomonas syringae and Pseudomonas putida. J. Ferment. Bioeng. 82, 509e511. Izumikawa, M., Khan, S.T., Takagi, M., Shin-ya, K., 2010. Spongederived Streptomyces producing isoprenoids via the mevalonate pathway. J. Nat. Prod. 73, 208e212. Julsing, M.K., Rijpkema, M., Woerdenbag, H.J., Quax, W.J., Kayser, O., 2007. Functional analysis of genes involved in the biosynthesis of isoprene in Bacillus subtilis. Appl. Microbiol. Biotechnol. 75, 1377e1384. Kilian, O., Benemann, C.S.E., Niyogi, K.K., Vick, B., 2011. Highefficiency homologous recombination in the oil-producing alga Nannochloropsis sp. Proc. Natl. Acad. Sci. USA. 108, 21265e21269. Knoll, A., 2008. Cyanobacteria and earth history. In: Herrero, A., Flores, E. (Eds.), The Cyanobacteria: Molecular Biology, Genomics and Evolution. Caister Academic Press. Koksharova, O.A., Wolk, C.P., 2002. Genetic tools for cyanobacteria. Appl. Microbiol. Biotechnol. 58, 123e137. Kothari, A., Potrafka, R., Garcia-Pichel, F., 2012. Diversity in hydrogen evolution from bidirectional hydrogenases in cyanobacteria from terrestrial, freshwater and marine intertidal environments. J. Biotechnol. 162, 105e114. Kumar, K., Mella-Herrera, R.A., Golden, J.W., 2010. Cyanobacterial heterocysts. Cold Spring Harbor Perspect. Biol. 2, a000315. Kuzma, J., Nemecek-Marshall, M., Pollock, W.H., Fall, R., 1995. Bacteria produce the volatile hydrocarbon isoprene. Curr. Microbiol. 30, 97e103. Lan, E.I., Liao, J.C., 2011. Metabolic engineering of cyanobacteria for 1-butanol production from carbon dioxide. Metab. Eng. 13, 353e363. Leplat, C., Champeimont, R., Saenkham, P., Cassier-Chauvat, C., JeanChristophe, A., Chauvat, F., 2013. Genome-wide transcriptome analysis of hydrogen production in the cyanobacterium Synechocystis: towards the identification of new players. Int. J. Hydrogen Energy 38, 1866e1872. Lindberg, P., Park, S., Melis, A., 2010. Engineering a platform for photosynthetic isoprene production in cyanobacteria, using Synechocystis as the model organism. Metab. Eng. 12, 70e79. Lindblad, P., Lindberg, P., Oliveira, P., Stensjö, K., Heidorn, T., 2012. Design, engineering, and construction of photosynthetic microbial cell factories for renewable solar fuel production. Ambio 41, 163e168. Lopez-Igual, R., Flores, E., Herrero, A., 2010. Inactivation of a heterocyst-specific invertase indicates a principal role of sucrose catabolism in heterocysts of Anabaena sp. J. Bacteriol. 192, 5526e5533. Ludwig, M., Schulz-Friedrich, R., Appel, J., 2006. Occurrence of hydrogenases in cyanobacteria and anoxygenic photosynthetic bacteria: Implications for the phylogenetic origin of cyanobacterial and algal hydrogenases. J. Mol. Evol. 63, 758e768. Mariscal, V., Flores, E., 2010. In: Hallenbeck, P.C. (Ed.), 2010. Multicellularity in a Heterocyst-forming Cyanobacterium: Pathways for Intercellular Communication, vol. 675. Springer-Verlag Berlin, Heidelberger Platz 3, D-14197, Berlin, Germany, pp. 123e135. Marreiros, B.C., Batista, A.P., Duarte, A.M.S., Pereira, M.M., 2013. A missing link between complex I and group 4 membrane-bound [NiFe] hydrogenases. Biochim. Biophys. 1827, 198e209. McCoy, M., Reisch, M., Tullo, A.H., Short, P.L., Tremblay J-, F., 2006. Production: growth is the norm. Chem. Eng. News Arch. 84, 59e68.
REFERENCES McNeely, K., Xu, Y., Bennette, N., Bryant, D.A., Dismukes, G.C., 2010. Redirecting reductant flux into hydrogen production via metabolic engineering of fermentative carbon metabolism in a cyanobacterium. Appl. Environ. Microbiol. 76, 5032e5038. Melnicki, M.R., Pinchuk, G.E., Hill, E.A., Kucek, L.A., Fredrickson, J.K., Konopka, A., Beliaev, A.S., 2012. Sustained H2 production driven by photosynthetic water splitting in a unicellular cyanobacterium. mBio 3, e00197ee00212. Miller, B., Oschinski, C., Zimmer, W., 2001. First isolation of an isoprene synthase gene from poplar and successful expression of the gene in Escherichia coli. Planta 213, 483e487. Min, H., Sherman, L.A., 2010. Hydrogen production by the unicellular, diazotrophic cyanobacterium Cyanothece sp. strain ATCC 51142 under conditions of continuous light. Appl. Environ. Microbiol. 76, 4293e4301. Miyamoto, K., Hallenbeck, P.C., Benemann, J.R., 1979a. Effects of nitrogen supply on hydrogen production by cultures of Anabaena cylindrica. Biotechnol. Bioeng. 21, 1855e1860. Miyamoto, K., Hallenbeck, P.C., Benemann, J.R., 1979b. Hydrogen production by the thermophilic alga Mastigocladus laminosus effects of nitrogen, temperature, and inhibition of photosynthesis. Appl. Environ. Microbiol. 38, 440e446. Miyamoto, K., Hallenbeck, P.C., Benemann, J.R., 1979c. Nitrogenfixation by thermophilic blue-green algae (cyanobacteria) temperature characteristics and potential use in biophotolysis. Appl. Environ. Microbiol. 37, 454e458. Miyamoto, K., Hallenbeck, P.C., Benemann, J.R., 1979d. Solar energy conversion by nitrogen limited cultures of Anabaena cylindrical. J. Ferment. Technol. 57, 287e293. Murry, M.A., Hallenbeck, P.C., Benemann, J.R., 1984. Immunochemical evidence that nitrogenase is restricted to the heterocysts in Anabaena cylindrica. Arch. Microbiol. 137, 194e199. Murry, M.A., Hallenbeck, P.C., Esteva, D., Benemann, J.R., 1983. Nitrogenase inactivation by oxygen and enzyme turnover in Anabaena cylindrica. Can. J. Microbiol. 29, 1286e1294. Nagahama, K., Ogawa, T., Fujii, T., Fukuda, H., 1992. Classification of ethylene-producing bacteria in terms of biosynthetic pathways to ethylene. J. Ferment. Bioeng. 73, 1e5. Nagahama, K., Ogawa, T., Fujii, T., Tazaki, M., Tanase, S., Morino, Y., Fukuda, H., 1991. Purification and properties of an ethyleneforming enzyme from Pseudomonas syringae pv. phaseolicola PK2. J. Gen. Microbiol. 137, 2281e2286. Nakatsuka, A., Murachi, S., Okunishi, H., Shiomi, S., Nakano, R., Kubo, Y., Inaba, A., 1998. Differential expression and internal feedback regulation of 1-aminocyclopropane-1-carboxylate synthase, 1-aminocyclopropane-1-carboxylate oxidase, and ethylene receptor genes in tomato fruit during development and ripening. Plant. Physiol. 118, 1295e1305. Nashawi, I.S., Malallah, A., Al-Bisharah, M., 2010. Forecasting world crude oil production using multicyclic Hubbert model. Energy Fuels 24, 1788e1800. Ogawa, T., Takahashi, M., Fujii, T., Tazaki, M., Fukuda, H., 1990. The role of NADH: Fe(III)EDTA oxidoreductase in ethylene formation from 2-Keto-4-methylthiobutyrate. J. Ferment. Bioeng. 69, 287e291. Oliveira, P., Lindblad, P., 2009. Transcriptional regulation of the cyanobacterial bidirectional Hox-hydrogenase. Dalton Trans., 9990. Perovic, S., Seack, J., Gamulin, V., Muller, W., Schroder, H.C., 2001. Modulation of intracellular calcium and proliferative activity of invertebrate and vertebrate cells by ethylene. BMC Cell Biol. 2. art. no.e7. Prabaharan, D., Kumar, D.A., Uma, L., Subramanian, G., 2010. Dark hydrogen production in nitrogen atmosphere - an approach for sustainability by marine cyanobacterium Leptolyngbya valderiana BDU 20041. Int. J. Hydrogen Energy 35, 10725e10730. 405 Ray, D.K., Ramankutty, N., Mueller, N.D., West, P.C., Foley, J.A., 2012. Recent patterns of crop yield growth and stagnation. Nat. Commun. 3, 1293. Saini, V., Sigman, M.S., 2012. Palladium-catalyzed 1,1-difunctionalization of ethylene. J. Am. Chem. Soc. 134, 11372e11375. Sakai, M., Ogawa, T., Matsuoka, M., Fukuda, H., 1997. Photosynthetic conversion of carbon dioxide to ethylene by the recombinant cyanobacterium, Synechococcus sp. PCC 7942, which harbors a gene for the ethylene-forming enzyme of Pseudomonas syringae. J. Ferment. Bioeng. 84, 434e443. Sakurai, H., Masukawa, H., 2007. Promoting R & D in photobiological hydrogen production utilizing mariculture-raised cyanobacteria. Mar. Biotechnol. 9, 128e145. Schirmer, A., Rude, M.A., Li, X., Popova, E., del Cardayre, S.B., 2010. Microbial biosynthesis of alkanes. Science 329, 559e562. Schmitz, O., Boison, G., Bothe, H., 2001. Quantitative analysis of expression of two circadian clock-controlled gene clusters coding for the bidirectional hydrogenase in the cyanobacterium Synechococcus sp. PCC7942. Mol. Microbiol. 41, 1409e1417. Schwarz, C., Poss, Z., Hoffmann, D., Appel, J., 2010. In: Hallenbeck, P.C. (Ed.), 2010. Hydrogenases and Hydrogen Metabolism in Photosynthetic Prokaryotes, vol. 675. SPRINGER-VERLAG BERLIN, Heidelberger Platz 3, D-14197, BERLIN, GERMANY, pp. 305e348. Seabra, R., Santos, A., Pereira, S., Moradas-Ferreira, P., Tamagnini, P., 2009. Immunolocalization of the uptake hydrogenase in the marine cyanobacterium Lyngbya majuscula CCAP 1446/4 and two Nostoc strains. FEMS Microbiol. Lett. 292, 57e62. Sharkey, T.D., 1996. Emission of low molecular mass hydrocarbons from plants. Trends Plant Sci. 1, 78e82. Sharkey, T.D., Chen, X.Y., Yeh, S., 2001. Isoprene increases thermotolerance of fosmidomycin-fed leaves. Plant Physiol. 125, 2001e2006. Sheehan, J., 2009. Engineering direct conversion of CO2 to biofuel. Nat. Biotechnol. 27, 1128e1129. Shestako, S., Khyen, N.T., 1970. Evidence for genetic transformation in blue-green alga Anacystis nidulans. Mol. Gen. Genet. 107, 372. Silver, G.M., Fall, R., 1991. Enzymatic synthesis of isoprene from dimethylallyl diphosphate in aspen leaf extracts. Plant Physiol. 97, 1588e1591. Silver, G.M., Fall, R., 1995. Characterization of aspen isoprene synthase, an enzyme responsible for leaf isoprene emission to the atmosphere. J. Biol. Chem. 270, 13010e13016. Summers, M.L., Wallis, J.G., Campbell, E.L., Meeks, J.C., 1995. Genetic evidence of a major role for glucose-6-phosphate-dehydrogenase in nitrogen-fixation and dark growth of the cyanobacterium Nostoc sp strain Atcc-29133. J. Bacteriol. 177, 6184e6194. Taiz, L., Zeiger, E., 2002. In: Taiz, Lincoln (Ed.), Plant Physiology, third ed. Sinauer Associates, Sunderland, MA, p. 672. Takahama, K., Matsuoka, M., Nagahama, K., Ogawa, T., 2003. Construction and analysis of a recombinant cyanobacterium expressing a chromosomally inserted gene for an ethylene-forming enzyme at the psbAI locus. J. Biosci. Bioeng. 95, 302e305. Tamagnini, P., Leitao, E., Oliveira, P., Ferreira, D., Pinto, F., Harris, D.J., Heidorn, T., Lindblad, P., 2007. Cyanobacterial hydrogenases: diversity, regulation and applications. FEMS Microbiol. Rev. 31, 692e720. Tan, X., Yao, L., Gao, Q., Wang, W., Qi, F., Lu, X., 2011. Photosynthesis driven conversion of carbon dioxide to fatty alcohols and hydrocarbons in cyanobacteria. Metab. Eng. 13, 169e176. Tao, L., Dong, H.-J., Chen, X., Chen, S.-F., Wang, T.-H., 2008. Expression of ethylene-forming enzyme (EFE) of Pseudomonas syringae pv. glycinea in Trichoderma viride. Appl. Microbiol. Biotechnol. 80, 573e578. Thiel, T., Poo, H., 1989. Transformation of a filamentous cyanobacterium by electroporation. J. Bacteriol. 171, 5743e5746.
406 22. ENGINEERED CYANOBACTERIA: RESEARCH AND APPLICATION IN BIOENERGY Troshina, O., Serebryakova, L., Sheremetieva, M., Lindblad, P., 2002. Production of H-2 by the unicellular cyanobacterium Gloeocapsa alpicola CALU 743 during fermentation. International Journal of Hydrogen Energy 27, 1283e1289. Tsinoremas, N.F., Kutach, A.K., Strayer, C.A., Golden, S.S., 1994. Efficient gene-transfer in Synechococcus sp strains Pcc-7942 and Pcc6301 by Interspecies conjugation and chromosomal recombination. J. Bacteriol. 176, 6764e6768. Tsygankov, A.A., Fedorov, A.S., Kosourov, S.N., Rao, K.K., 2002. Hydrogen production by cyanobacteria in an automated outdoor photobioreactor under aerobic conditions. Biotechnol. Bioeng. 80, 777e783. Ungerer, J., Tao, L., Davis, M., Ghirardi, M., Maness, P.-C., Yu, J., 2012. Sustained photosynthetic conversion of CO2 to ethylene in recombinant cyanobacterium Synechocystis 6803. Energy Environ. Sci. 5, 8998e9006. Verne, J. The Mysterious Island. Plain Label Books. Vignais, P.M., Billoud, B., 2007. Occurrence, classification, and biological function of hydrogenases: an overview. Chem. Rev. 107, 4206e4272. Wang, C.H., Persyn, A., Krackov, J., 1962. Role of the Krebs cycle in ethylene biosynthesis. Nature 195, 1306e1308. Wang, K., Li, H., Ecker, J.R., 2002. Ethylene biosynthesis and signaling networks. Plant Cell 14, S131eS151. Weingart, H., Völksch, B., Ullrich, M.S., 1999. Comparison of ethylene production by Pseudomonas syringae and Ralstonia solanacearum. Phytopathology 89, 360e365. Weissman, J.C., Benemann, J.R., 1977. Hydrogen production by nitrogen-starved cultures of Anabaena cylindrica. Appl. Environ. Microbiol. 33, 123e131. Weyman, P.D., Pratte, B., Thiel, T., 2008. Transcription of hupSL in Anabaena variabilis ATCC 29413 is regulated by NtcA and Not by Hydrogen. Appl. Environ. Microbiol. 74, 2103e2110. Winters, K., Parker, P.L., Vanbaale, C., 1969. Hydrocarbons of bluegreen algae - geochemical significance. Science 163, 467. Withers, S.T., Gottlieb, S.S., Lieu, B., Newman, J.D., Keasling, J.D., 2007. Identification of isopentenol biosynthetic genes from Bacillus subtilis by a screening method based on isoprenoid precursor toxicity. App. Environ. Microbiol. 73, 6277e6283. Wolk, C.P., Vonshak, A., Kehoe, P., Elhai, J., 1984. Construction of shuttle vectors capable of conjugative transfer from Escherichia coli to nitrogen-fixing filamentous cyanobacteria. Proc. Natl. Acad. Sci. USA. 81, 1561e1565. Xue, J., Ahring, B.K., 2011. Enhancing isoprene production by genetic modification of the 1-Deoxy-D-Xylulose-5-Phosphate pathway in Bacillus subtilis. Appl. Environ. Microbiol. 77, 2399e2405. Yang, J., Zhao, G., Sun, Y., Zheng, Y., Jiang, X., Liu, W., Xian, M., 2012. Bio-isoprene production using exogenous MVA pathway and isoprene synthase in Escherichia coli. Bioresour. Technol. 104, 642e647. Yoon, H.S., 2004. A molecular timeline for the origin of photosynthetic eukaryotes. Mol. Biol. Evol. 21, 809e818. Yoon, H.S., Haeshin, J., Kim, M., Junsim, S., Park, T., 2006. Evaluation of conversion efficiency of light to hydrogen energy by Anabaena variabilis. Int. J. Hydrogen Energy 31, 721e727. Yoshihara, S., Geng, X., Okamoto, S., Yura, K., Murata, T., Go, M., Ohmori, M., Ikeuchi, M., 2001. Mutational analysis of genes involved in pilus structure, motility and transformation competency in the unicellular motile cyanobacterium Synechocystis sp. PCC 6803. Plant Cell. Physiol. 42, 63e73. Yura, K., 1999. Putative mechanism of natural transformation as deduced from genome data. DNA Res. 6, 75e82. Zang, X., Liu, B., Liu, S., Arunakumara, K.K.I.U., Zhang, X., 2007. Optimum conditions for transformation of Synechocystis sp. PCC 6803. J. Microbiol. 45, 241e245. Zhang, S., Bryant, D.A., 2011. The Tricarboxylic Acid Cycle in Cyanobacteria. Science 334 (6062), 1551e1553. Zhao, Y., Yang, J., Qin, B., Li, Y., Sun, Y., Su, S., Xian, M., 2011. Biosynthesis of isoprene in Escherichia coli via methylerythritol phosphate (MEP) pathway. Appl. Microbiol. Biotechnol. 90, 1915e1922.
C H A P T E R 23 Sustainable Farming of Bioenergy Crops Adrian Muller Research Institute of Organic Agriculture FiBL, Zurich, Switzerland; Institute for Environmental Decisions, Swiss Federal Institutes of Technology (ETH), Zurich, Switzerland email: adrian.mueller@fibl.org O U T L I N E Introduction 407 Criteria for Sustainable Farming and Sustainable Food Systems Conventional Agricultural Production Sustainable Agricultural Production Sustainable Food Systems What is Sustainable Bioenergy Production? Ways of Comparisons 409 409 409 410 410 410 INTRODUCTION Is bioenergy a sustainable energy source? A positive answer to this question is a key if bioenergy shall become a significant sustainable energy source for future societies. The answer to this question depends on various aspects of the production and use of bioenergy. The most prominent topic there is the greenhouse gas (GHG) balance, as this is the key motivation to investigate bioenergy at all. For most cases, the GHG balance is positive, albeit not at a tremendously high rate and negative values are due to large emissions from direct and indirect land use change (ILUC), e.g. if palm oil plantations are established on peatland rainforest (Faist Emmenegger et al., 2012; PBL, 2010; Fargione et al., 2008). Clearly, in the use phase, GHG emissions are counted as zero due to the overall assumption of renewable biomass provision for bioenergy. However, over the whole life Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00023-1 Sustainability Criteria for Biofuel Production Biomass Use 410 412 How Much Bioenergy may be Produced Sustainably? Global Bioenergy Potential Bioenergy Potential on Farm Level 412 413 414 Conclusions 415 References 415 cycle of bioenergy, GHG emissions arise at various steps, in particular, in the agricultural production phase (Faist Emmenegger et al., 2012). The GHG balance usually plays the role of a fundamental decision criterion in favor or against bioenergy types. With a zero or negative balance, bioenergy will not contribute to and may even adversely affect climate change mitigation. However, even in this case, one could promote one argument for bioenergy, namely that it replaces nonrenewable energy sources with renewable ones. This is particularly attractive for liquid fuels as there are currently no other alternatives available than liquid fuels from biological sources. In the following, we often focus on liquid biofuels as the discussion of sustainability of bioenergy is developed furthest for those and most data are available for those bioenergy types. The findings on the sustainability of agricultural production of crops for liquid biofuel, however, apply to agricultural production of any bioenergy type and we will 407 Copyright Ó 2014 Elsevier B.V. All rights reserved.
408 23. SUSTAINABLE FARMING OF BIOENERGY CROPS also report findings for biogas production, for example, on which there is also a considerable amount of research. Besides the GHG balance, many other criteria are needed to assess the sustainability of bioenergy. They range from environmental impacts of the agricultural production process over emissions in the use phase (e.g. nitrous oxide emissions from biomass-fueled power plants) to socioeconomic aspects, such as production costs or effects on labor (Elbehri et al., 2013; Faist Emmenegger et al., 2012; Delucchi, 2010). The assessments of environmental impacts show that most bioenergy crops perform worse than the fossil fuel baseline regarding many criteria of sustainability in agricultural production. The best performance is realized for some residue or forestry-based bioenergy sources such as fuels from wood products (Faist Emmenegger et al., 2012). The negative performance of bioenergy crops is due to the fact that current conventional agricultural production has many adverse environmental effects (Matson et al., 1997; IAASTD, 2009; see e.g. Tomei and Upham, 2009 on soy-based biodiesel in Argentina for an exemplification). As important as these findings on environmental performance is, the key point in the current discussion besides the GHG balance is land competition between bioenergy crops and food production (Rathman et al., 2010; Delzeit et al., 2010; HLPE, 2013). Bioenergy crops compete for fertile land with food crops and bioenergy use of food crops such as maize directly competes with food use. This dynamics has been behind the volatile and high food prices in recent years as, e.g. for maize and grain in 2007e2008 (NYT, 2007; HLPE, 2013). There are several options available to address all these challenges. First, there are alternative ways of agricultural production that reduce the adverse impacts from farming, a key example being organic agriculture (Rossi, 2012). These alternatives usually focus on systemic aspects of the whole production systems, emphasize closed nutrient cycles, sustaining soil fertility and plant health and the role of ecosystem services, e.g. for pest and disease control. Second, how strong land competition may become depends on the concrete situation, i.e. on total bioenergy demand, relative prices for food and energy, the policy environment with incentives and regulations, etc. Very circumspect planning of any larger scale bioenergy strategy seems crucial to have a chance to avoid such adverse effects (e.g. Damen, 2010 for Thailand, wa Gathui and Ngugi, 2010 for vulnerable regions in Kenya, or Palola and Walker, 2010 on oil palms). Standards and certification are also suggested as a means to address the potential land use competition, e.g. by explicitly excluding bioenergy from fertile croplands (e.g. GEF et al., 2013). Third, new forms of bioenergy emerge (IEA, 2010). Instead of the so-called “first-generation” bioenergy that is based on use of oil, starch or sugar contents of crops to manufacture biodiesel, “second-generation” bioenergy relies on the lignocellulosic contents of the crops. This allows a much wider range of crops to be utilized for biofuel production, in particular nonfood woody crops such as switchgrass or Eucalyptus (IEA, 2010). It also allows utilization of basically the whole plant and of crop residues for biofuel production. This reduces land demand per unit energy for bioenergy. However, it is not expected that the second-generation biofuels will play a significant role in the near future as much research is still needed (IEA, 2010). An aspect that is largely missing in this discussion on environmental impacts of bioenergy production and land competition is the role of biomass quantities. Large quantities of biomass are needed for bioenergy strategies that significantly contribute to the global energy supply. On the other hand, biomass plays a key role as a fertilizer in sustainable agricultural production systems. There is thus a direct competition for biomass between exporting it from the agricultural production system for bioenergy use and recycling it as a fertilizer (Muller, 2009). Data and studies on how much biomass may be exported for bioenergy use from sustainable agricultural farming systems are scarce, but indications that it will not be much dominate, as we will discuss further down. The key topic in this discussion is thus less whether agricultural production of bioenergy crops can be done in a sustainable way, as sustainable agricultural production is well established and can be implemented for any crop production and as options to reduce land competition are around in principle, albeit challenging to implement. The question is thus rather how much bioenergy can be sustainably produced in such a context where biomass is not a waste output from agricultural production but an essential fertilizer input in sustainable agricultural production systems and where fertile land is needed for food production. In this chapter, we focus on the sustainable production of bioenergy crops with a focus on the farm level. This thus covers the farm operations, but it does not cover processing, transport, storage and use of bioenergy. Possible disposal of waste after bioenergy use is shortly addressed in the context of fertilizer use (cf. Section Bioenergy Potential on Farm Level). This chapter also covers more aggregate aspects of agricultural production such as land and water resource use and more systemic aspects related to the whole food system, such as the competition for land, water and biomass between food and bioenergy production. In this context, we also address general aspects of food security and in the conclusions also relate to the role of meat and milk production as another sector that competes for scarce land resources with crop-based food production. We restrict this analysis to agricultural bioenergy. We thus address forestry only occasionally and do not
CRITERIA FOR SUSTAINABLE FARMING AND SUSTAINABLE FOOD SYSTEMS address aquaculture such as for algae-based biomass production for energy use or biomass production in industrial contexts, such as in bioreactors (Chen et al., 2011). As biological processes are involved in all these cases, the key topics we discuss in the following are relevant also there: resource use and resource competition, environmental impacts of the production (e.g. water and air pollution, intoxication due to pesticide use, etc.), and effects on the sustainability of the food production system as a whole. On the other hand, harvesting the energy capture potential of photosynthesis in purely industrial production of biomass in bioreactors could be an option, as the production system can work well separated from the natural environment. For such production, no ecosystem inputs such as fertile soils and water bodies are used as inputs or sinks (e.g. for nutrient runoff) and thus potentially depleted. Environmental impacts can thus be kept to a minimum. Such operations do however not refer to agricultural but rather to industrial production of biomass for bioenergy use and we do not further pursue this here. Furthermore, they are still in the research and development phase and commercial use is not expected in the near future (Pragya et al., 2013). This chapter is organized as follows. In Section Criteria for Sustainable Farming and Sustainable Food Systems, we discuss the criteria of sustainable agricultural production and sustainable food systems. In Section What is Sustainable Bioenergy Production? we assess how bioenergy production may be implemented if it needs to meet these criteria. We also address sustainability criteria for bioenergy production as formulated by a range of institutions and relate them to the criteria for sustainable agricultural production from Section Criteria for Sustainable Farming and Sustainable Food Systems. Section How Much Bioenergy may be Produced Sustainably? provides some discussion on how much agricultural bioenergy may be produced in a sustainable way and Section Conclusions concludes. CRITERIA FOR SUSTAINABLE FARMING AND SUSTAINABLE FOOD SYSTEMS Conventional Agricultural Production Over the decades from 1960 to 2010, agricultural production has increased more than threefold and per capita provision of calories has increased by one-third (FAOSTAT, 2013). This has greatly contributed to feeding an ever-increasing population (from 3109 to 6.7109 billion over this period). This development is usually subsumed under the label “Green revolution” (IFAD, 2001; Evenson and Gollin, 2003). Starting in the 1960s, this development was based on a strict focus on monocropping with high yielding species and varieties, 409 irrigation and mechanization where available and increased use of mineral fertilizers, pesticides and herbicides. The successes of the green revolution are evident, but so are the downturns related to it (Matson et al., 1997; DFID, 2004). The focus on monocropping, chemical fertilizers, pesticides, irrigation and mechanization has left an increasingly negative legacy regarding adverse effects on soil fertility, i.e. increased soil degradation, salinization and depletion of water bodies, on intoxication of the environment, biodiversity loss, loss of ecosystem services, on eutrophication of water bodies and animal health (Matson et al., 1997). Current agriculture is well able to feed the world and will be able in 2050 to feed more than 9 billion people, given projected yield increases realize (Alexandratos and Bruinsma, 2012). The challenge is not the average supply of calories per capita but their distribution globally and the fact that a third is lost or wasted globally (Godfray et al., 2010). However, in the light of the adverse effects of the green revolution and climate change, such yield increases may be compromised and sustained agricultural production calls for alternative cropping practices and a fundamental shift in the agricultural production system (IAASTD, 2009; Müller et al., 2010). Sustainable Agricultural Production A new revolution in agricultural production is thus needed. On the production level, a sustainable future agricultural system needs to focus on mitigating and avoiding the adverse effects of current agricultural practices. It needs to focus on crop diversity, ecosystem services, soil protection and fertility, nutrient and water use cycling, biocontrol of pests, diseases and weeds and reduced pesticide use. A range of alternative production approaches are available (Eyhorn et al., 2003; Pretty et al., 2006; Rossi, 2012), such as agroecologybased approaches, focusing on utilization of ecological concepts (Altieri, 1995), or integrated pest management, focusing on reducing pesticide use via managing pest populations in such a way that damages remain low (Bajwa and Kogan, 2002). The role model for these alternative approaches is organic agriculture with its ban on most pesticides, focus on soil fertility, plant health and closed nutrient cycles, utilization of optimized crop rotations and crop diversity, organic fertilizers and ecosystem functions for pest and weed control (FAO, 2002; Eyhorn et al., 2003; IFOAM, 2006). Organic agriculture is the role model as it addresses all adverse effects of conventional production, adopts a systemic approach and is well established and tested for decades and embedded in a context of governance, information provision, training and extension institutions that make it the best-developed alternative production system. Organic agriculture performs better than conventional
410 23. SUSTAINABLE FARMING OF BIOENERGY CROPS agriculture with respect to most environmental indicators on a per-hectare basis (Schader et al., 2012). The biggest drawback is its generally lower yields (Seufert et al., 2012; De Ponti et al., 2012; Badgley et al., 2007). Lower yields predominantly manifest in comparison to high-yielding intensive conventional agriculture. In developing countries, in a context of currently nooptimal conventional production systems, organic yields are on par or even higher for well-managed organic farms. The lower yields can result in a less favorable per kilogram produce assessment of environmental impacts for some products in organic agriculture (Schader et al., 2012). We emphasize that we do not address socioeconomic aspects of organic production here, such as the need for information and extension services to train farmers and potential challenges of the conversion from organic to conventional agriculture. Interestingly, the key principles and practices of organic agriculture become increasingly important in conventional agriculture, mainly due to the increasing need to contribute to climate change mitigation and adaptation but also due to the increasingly important discussion on global biodiversity losses. Optimized crop rotations with deep-rooting forage legumes and use of organic fertilizers, for example, are promoted in the context of climate change mitigation and adaptation to improve soil fertility and increase soil organic carbon levels (Smith et al., 2008) and reducing nitrogen loads are key to protect biodiversity. consumption of animal products that are mainly based on grassland feed (and some by-products of food production) thus comprise an optimal option for a sustainable food system (e.g. preliminary results from the FAO-SOL-model, Schader et al., 2012). WHAT IS SUSTAINABLE BIOENERGY PRODUCTION? Ways of Comparisons When assessing the sustainability of energy crop production, the first criterion is usually its GHG performance with regard to a fossil baseline. For this comparison, the baseline fuel mix plays a crucial role, as the increasing importance of unconventional fossil fuel sources such as oil sands will increase GHG emissions from the baseline and its general environmental impacts, thus relatively improving the performance of bioenergy production (Faist Emmenegger et al., 2012). This bears the danger that biofuel options with increasing environmental impacts and less favorable GHG balances become relatively more sustainable. Here, we adopt a different focus as we are primarily interested in the sustainability of bioenergy production with reference to sustainability in agricultural production systems and in food systems in general. This takes all sustainability criteria into account and does not focus on the GHG balance. Sustainable Food Systems Agricultural production is only one aspect of food systems. Sustainability in agriculture also needs to address more encompassing concepts such as food security and land availability on a regional level. If agriculture would switch to organic, 10e20% more land would be needed due to lower yields, given diets do not change and the same amount of wastage is produced as today. However, dietary change is a key topic for sustainable food systems, as a large part of agriculture’s environmental charge stems from animal husbandry. Reducing meat, egg and dairy product consumption levels would greatly help to reduce environmental pressure from agricultural production. Focusing on feeding animals on grasslands and not on food crops such as soy and maize would reduce the need for land, as calorie production from crops is much more efficient than from animals. Furthermore, about 30% of agricultural production are lost or wasted globally (Godfray et al., 2010). Reducing this would also contribute to reduced agricultural land use. Such reduced land use would on the other hand reduce pressure to further increase yields. Organic production in combination with reduced waste and lower Sustainability Criteria for Biofuel Production There exist several approaches proposing sustainability criteria for bioenergy (e.g. Cramer et al., 2007; EU, 2009; IEA, 2010; RSB, 2011; GEF et al., 2013, where the last two are very detailed). They usually cover some goal for GHG emission reductions, e.g. a 35% reduction of aggregate emissions over some time period with respect to the baseline as suggested in EU regulations (50% from 2017 to 60% from 2018 onwards, EU, 2009). While this seems a clear criterion, its assessment is complex. The choice of different default values, soil carbon stock data and land use change definitions, for example, is behind the huge differences in GHG balances between two GHG calculation tools as assessed in Hennecke et al. (2013), one of them being the tool used by the Roundtable of Sustainable Biomaterials (RSB) whose sustainability criteria are discussed below. Other aspects are decisive as well. The choice of the time horizon over which aggregate reductions have to be achieved and the choice of the social discount rate, which influences the relative importance of current and future emissions, also greatly influence the outcome (De Gorter and Tsur, 2009).
WHAT IS SUSTAINABLE BIOENERGY PRODUCTION? The sustainability criteria proposed for biofuel production that relate to agricultural production and the food system, address land competition, biodiversity, environmental impacts on soil, water and air, and social aspects. Land competition, resp. absence thereof and the related food security are by far the most prominent criteria in the discussion (HLPE, 2013). Potential drivers for land competition are many. First, there is the fact that bioenergy crops need fertile land to achieve economically interesting yields. Biofuels on marginal lands are some option in smallholder and communityself-sufficiency contexts, but for commercial supply of biofuels in significant shares of total global energy demand, production on fertile land that potentially is used for food production is necessary (cf. Section Bioenergy Potential on Farm Level). Similarly, bioenergy crop production depends on water availability and nutrient inputs as any other agricultural production system. Thus, a second potential competition is not only on fertile lands, but also on land with sufficient water availability (Lysen and van Egmond, 2008), in particular in the context of climate change, where water scarcity will become a prevalent problem in many regions (Meehl et al., 2007). Third, relative price differences between bioenergy and food production will be and have been a key driver behind land competition as without further regulation, land will be allocated to the most profitable production. Stated differently, an increasing demand for biofuels leads to higher prices, which triggers an increasing supply for it, with corresponding land use (HLPE, 2013). It has to be emphasized that land competition between food and other uses is not new and relative profitability has always been a key driver behind this. As Nhantumbo and Salomão (2010) state, “Competition for higher-value resources existed well before the biofuels campaign was initiated. In this sense, biofuels production per se cannot be blamed for land use conflicts, as the same types of conflicts have occurred in other economic activities. But, in conjunction with other activities like mining, forestry and tourism, biofuels projects further exacerbate competition for land, water and other resources” (p. 4). The key point is thus that biofuel expansion increases the pressure on the already scarce resource of fertile land. In principle, policy measures can be used to mitigate these adverse effects. However, their implementation is often riddled with difficulties and land-use rights protection, enforcement of laws and regulations, etc. have to be carefully considered when establishing potentially promising institutions for sustainable land use. Nhantumbo and Salomão (2010) illustrate these challenges for the case of Mozambique and draw a rather pessimistic picture. The land use debate is further complicated by ILUCs (Wicke et al., 2012). Those arise, for example, if biofuel 411 expansion in one region (e.g. sugar cane in southern Brazil) leads to land use change in another region (in this case, deforestation for livestock production in northern Brazil). The rationale behind this example is the fact that expanding sugarcane in the South is at the expense of already existing pastures in this region, that then themselves relocate at the expense of other uses such as natural forests (Andrade De Sa et al., 2013). Such effects are very difficult to clearly identify and assess (Wicke et al., 2012). This is also the case in the detailed analysis in Andrade De Sa et al. (2013) who find only very weak significant statistical effects. Nevertheless, there is evidence from many descriptive studies that the potential presence of such mechanisms must not be neglected (PBL, 2010). ILUC is not only relevant for the competition between different land uses but also for the GHG balances of biofuels, as it can have considerable negative effects on those (PBL, 2010; Faist Emmenegger et al., 2012). Biodiversity criteria mainly refer to the ban of using forests or protected areas for bioenergy production (e.g. Cramer et al., 2007; EU, 2009) or to being attentive not to use invasive species as bioenergy crops (UNEP, 2010). The use of protected areas can also be seen as a particular aspect of land competition from biofuel production. As mentioned above, biofuel production competes not only with food for land but also with other uses, such as biodiversity protection and also with fiber and biomaterial production that all depend on land availability. Invasive species are seen as a potential danger, due to already existing cases but also due to general characteristics of biofuel crops that also correlate with invasiveness (e.g. fast growth or tolerance to wide range of soil and climate conditions, UNEP, 2010). Much less prominent in this discussion are the adverse effects of current agricultural production on biodiversity (mainly due to overfertilization with nitrogen and pesticide use), albeit those are a key driver behind biodiversity losses (Galloway et al. 2008). This is mentioned in Bindraban et al. (2009) and adoption of agricultural practices with low negative effects on biodiversity is a criterion in GEF (2013), but not in EU (2009) or RSB (2011). Other environmental impacts largely remain rather unspecific in the criteria suggested, although the range of adverse environmental effects of current agricultural production as described above will also realize in bioenergy cropping systems. EU (2009) for example only posits that the production has to meet the Community environmental requirements and in GEF et al. (2013), water contamination is assumed to be no issue if legal requirements are met. The size of the adverse environmental effects depends on the types of crops. Grassland or wood products usually perform better than annual crops, for example, WBGU (2009). Somewhat more
412 23. SUSTAINABLE FARMING OF BIOENERGY CROPS detailed criteria are usually given in reference to soilrelated aspects such as soil fertility and soil organic carbon contents (see e.g. Cramer et al., 2007; EU, 2009; RSB, 2011; GEF et al., 2013). Regarding soil organic carbon, some sustainability criteria explicitly exclude bioenergy cropping on peatlands and other carbon-rich soils (EU, 2009). On the other hand, some bioenergy crops are judged to be advantageous for soil carbon levels, mainly grassland and forest-based bioenergy. The effect on soil carbon is not that clear for some perennial crops and rather negative for annual crops (WBGU, 2009). Social sustainability criteria, finally, sometimes tend to be formulated on a very general level. EU (2009), for example, only requires that source countries for bioenergy have “ratified and implemented” (p. 97) a range of conventions referring to labor rights, gender aspects, etc. RSB (2011) and GEF et al. (2013), on the other hand, are quite detailed on the social aspects that cover a range of important criteria for social sustainability in agricultural production. RSB (2011) and GEF et al. (2013) also make long-term economic viability of bioenergy projects a criterion for their sustainability assessment. It is not mentioned as a criterion, but bioenergy crops can have some risk-spreading characteristics as they can increase production diversity of a farm and as their demand and price dynamics likely follows different patterns than food or fiber crop demand and prices. Some types of bioenergy crops can also be used for direct on-farm energy provision without much investment needs, such as Jatropha, for example. These energy crops thus have the potential to increase energy access and reduce workload of women considerable in case they have to collect fuel-wood from far away, which is a common situation in many poor regions in developing countries. Energy crops can also provide specific income sources for women, as many case-studies show (Karlsson and Banda, 2009). However, there is similar evidence of problematic situations from case-studies, and whether a bioenergy project is advantageous for single farmers, the community and women in particular strongly depends on the concrete design and institutional context. Further example of positive cases are given in Practical Action Consulting (2009), and some negative cases for example in Ribeiro and Matavel (2009), focusing on Jatropha in Mozambique. The choice of sustainability criteria for bioenergy thus reflects the classical topics of sustainability criteria, with a focus on environmental aspects and climate change in particular. An additional aspect is land competition, which is covered extensively in the discussion. The focus on environmental criteria is understandable as bioenergy is a climate change mitigation strategy and prime impacts of agricultural production are in the environment. However, besides GHG emissions and, partly, biodiversity, the assessment of environmental criteria remains rather weak. For a comprehensive sustainability assessment, the topical breadth and depth in analysis must be improved. Generally, bioenergy production has the same impacts as any agricultural production and sustainability in bioenergy production largely links to sustainable agricultural production. Biomass Use The assessment of proposed sustainability criteria for bioenergy production shows that the competition for biomass between bioenergy use and for fertilizing sustainable agricultural production systems is no topic. It is covered marginally in some other publications on sustainable bioenergy, e.g. in Bindraban et al. (2009) or Blonz et al. (2008), although it is of key relevance for sustainable agricultural production. Some publications address this topic as a caveat of agricultural or forestry residues use, as exporting too much of them causes soil carbon losses and soil degradation (e.g. WBGU, 2009). Only Muller (2009) takes up this topic in depth. The export of biomass from the fields for bioenergy use also exports nutrients that have to be replaced by other fertilizers, i.e. mineral fertilizers. The overuse of mineral fertilizers is however a key driver behind many environmental problems of current agricultural production. Sustainable agricultural production systems are based on closed nutrient cycles and organic fertilizers (compost from crop residues, roots and residues that remain on and in the fields, and manure from livestock operations). Those are keys for soil fertility and increased soil organic carbon levels (Lal, 2008; Gattinger et al., 2012). The nutrient export becomes particularly relevant for second-generation biofuels, where basically the whole plant can be used and no unused residues remain, resp. where cellulosic residues from any crops can be utilized (IEA, 2010). This is even suggested as a strategy to mitigate land use competition, as residues come without additional land requirements and feedstock for second-generation biofuel is claimed to often grow on marginal lands (IEA, 2010, 2011). On marginal lands in particular, high organic matter inputs are key to improve soil fertility, though. Also for bioenergy crops, yields tend to be lower and erratic on marginal lands and economic viability of bioenergy projects is often given on fertile land only (Bindraban et al., 2009). Thus, regarding the biomass competition, the most promising options to avoid land use competition seem particularly problematic. HOW MUCH BIOENERGY MAY BE PRODUCED SUSTAINABLY? As illustrated above, sustainable production of biomass for energy use is in principle possible. Thus,
HOW MUCH BIOENERGY MAY BE PRODUCED SUSTAINABLY? the key question is how much biomass may be produced sustainably on a global level. We basically discern two types of approaches to this question. First, there are various assessments of the global biomass potential for energy use. They are based on assessments of the suitability of global land areas for biomass production and corresponding yields, usually imposing the condition that food, feed and fiber demand need to be met. The second approach focuses on single farms or farming systems and estimates how much bioenergy may be produced in such, given the specific agronomic characteristics. These latter studies can also involve experimental data from case studies. Global Bioenergy Potential A range of literature assesses the global bioenergy potential employing various models and assumptions. WBGU (2009), Chum et al. (2011) and HLPE (2013) contain some recent reviews of this literature. Some very gross global comparisons are illustrative. If all harvested biomass today (including crops, forage, wood, and residues) would be used for energy use, this would cover about one-third of today’s energy supply (HLPE, 2013). This total harvested biomass corresponds to about 230 Exajoule (EJ) of primary energy per year. Chum et al. (2011) give a gross estimate for the technical bioenergy potential of 100e300 EJ/a in 2050, showing a wide range of uncertainty, though. This amount of bioenergy would very roughly cover between 10% and 60% of total primary energy supply in 2050, which ranges between 500 and 1000 EJ/a, based on 164 scenarios (Chum et al., 2011, Figure 10.3). Nevertheless, this number is illustrative as it roughly corresponds to a situation, where a biomass quantity equaling the total current biomass production was used for bioenergy in 2050. Current biomass energy use is about 50 EJ/a. Chum et al. (2011) base these estimates on a literature review that assess the physical biomass production potential based on land suitability and crop yields. These assessments also rely on considerably gains from yield increases and agricultural technology progress. While Chum et al. (2011) evaluate these numbers rather positively, HLPE (2013) considers them very problematic. WBGU (2009) in detail presents their model for the assessment of bioenergy potentials, relating to clearly stated general sustainability boundaries (biodiversity conservation, food security, climate change and acidification mitigation, and soil protection) and arrive at an estimate of 80e170 EJ/a. A key drawback in these estimates is the lack of economic and agronomic considerations in a systematic way. Both Chum et al. (2011) and HLPE (2013) emphasize the role of economic aspects, i.e. costs of such bioenergy development. WBGU (2009) also emphasize 413 that their estimate is purely technical given some sustainability boundaries and that the economic potential likely is considerably lower. But only few studies address market interactions between food and energy crops in economic equilibrium models and none of those was used to assess the bioenergy potential presented. Supply costs curves for various bioenergy crops resp. food crops for energy use are provided for illustration in Chum et al. (2011), for example, but no consequences on crop prices in interaction with demand are derived from this. Infrastructure and access to the suitable land are neither addressed explicitly. However, to realistically assess any competition for land between food and energy crop production, biomass and land markets need to be included in the model, as the relative profitability of food or bioenergy production on a certain area of land will drive production decisions and food and bioenergy supplydunless strong governmental regulations are imposed on the bioenergy and land markets. Thus, even if the biomass potential for energy uses is very large, its effect on food market prices needs to be assessed in detail to derive robust statements on land competition. The second aspect that is missing is agronomic characteristics of biomass production. The estimates reviewed in Chum et al. (2011) consider some sustainability criteria when trying to assess how much biomass may be produced without additional conversion of forests and grassland to cropland, which may be the effects on biodiversity, or via investigating whether bioenergy crops may even serve to improve degraded soils. However, agronomic aspects of the crop production system are largely neglected. Water requirements and potential problems related to that are mentioned explicitly, but are not captured explicitly in the models referred. Furthermore, any crop production needs to be fertilized if yield decreases after some years and soil degradation should be avoided. Yield assessments for biomass production do however not differentiate for nitrogen inputs. In addition, fertilizing with mineral fertilizers only is not enough, as organic fertilizers are crucial to halt soil degradation (e.g. Lal, 2008; Blanco-Canqui and Lal, 2009). This problem is mentioned in Chum et al. (2011), but it does not become effective for determining the biomass potential although it directly conflicts with the basic mechanism of bioenergy cropping, which is exporting a high amount of biomass, resp. using biomass residues formerly left on the field. Utilizing the biomass potential referred to above may thus result in considerably increased total global fertilizer use with respect to today (and in addition, crop production has to increase by 70% as well, Alexandratos and Bruinsma, 2012). This would put additional pressure on the nitrogen cycle and lead to corresponding adverse environmental effects (Erisman et al., 2010). WBGU (2009)
414 23. SUSTAINABLE FARMING OF BIOENERGY CROPS contains a review of several bioenergy crops that also contains some agronomic aspects. They conclude that only grasslands and some forestry have a sustainable potential for bioenergy provision. This is also reflected in their model, which assesses the bioenergy potential on additional grassland and forestry use, as only those meet their sustainability boundaries, besides some use of residues. Bioenergy Potential on Farm Level As we have seen in the previous section, assessments of the global bioenergy potential are based on land use and land availability consideration subject to several sustainability criteria. These assessments thus tend to disregard agronomic boundary conditions. WBGU (2009) is one exception and also explicitly includes such aspects on a very aggregate level in their model, by assuming that only 60% of residues can be used for energy production technically (and only 30% economically), given that part of the residue biomass needs to be left on the fields in order to avoid soil degradation. In contrast to such global or regional assessments, farm or farming system-based assessments are in principle able to account for such agronomic boundaries. Rossi (2012) reviews a range of sustainable farming systems as options for sustainable biomass production. He points out the role of biomass as a fertilizer and for soil fertility, but does not provide quantitative assessments of how much biomass may be exported from these systems for bioenergy use. Even more, the case studies presented in Rossi (2012) often do not address bioenergy production at all but only illustrate the advantageous performance of the respective farming system along a range of sustainability criteria. There is however other research that provides detailed quantitative analysis. Meyer and Priefer (2012) for example discuss the potential of biogas production in organic agriculture, based on case-study farms in Germany. Biogas fits neatly into organic production systems, as in organic farms, much biomass that can be used as feedstock for biogas plants is around (from grassclover leys in the crop rotations, for example) and the biogas slurry can be used as a fertilizer. Meyer and Priefer (2012) provides also some forecast on the potential for such bioenergy production in Germany, assuming that the biogas is used for electricity production and also utilizing the heat generated in the power plants. Assuming 20% of agricultural production being organic (political goals for 2020 are 20% in Germany) and equipped with biogas facilities, 7 TWh/a electricity could be provided plus 50% of this energy in heat. Assuming a total electricity demand of 535 TWh/a in Germany in 2030 BMU 2011), similar biogas production on all farms would provide 6e7% of this (35 TWh/a). Also, Anspach (2009) finds that biogas production fits well into organic production systems. Using biogas slurry as fertilizer has also some additional advantages regarding yields, environmental impacts and weed control (as seeds of weeds e.g. in manure are killed in the biogas digester). The potential of biogas production is also recognized by authors of more aggregate studies, e.g. Bindraban et al. (2009). This biogas production is assumed to work largely without bioenergy cropping and only uses residues and manure. Thus, it does not lead to competition with food production. Currently, the reality in Germany is different, though, as co-substrates are imported to a significant part in biogas digesters and part of those are specifically grown for biogas production (e.g. maize). Another body of literature focuses on energy selfsufficiency of organic farms, motivated by the unsustainable use of fossil fuels also in organic production systems (Carter et al., 2012; Christen and Dalgaard, 2013; Halberg et al., 2008; Oleskowicz-Popiel et al., 2012; Pugesgaard et al., 2013). Those studies are from Denmark and serve as further illustration for the bioenergy production in sustainable agricultural production systems. They generally find that energy self-sufficiency of organic farms is possible and that sometimes even some small energy surplus can be generated. Carter et al. (2012) are somewhat different, as they focus on a GHG life-cycle analysis and do not address nutrient recycling aspects at all. Pugesgaard et al. (2013) find that energy self-sufficiency is also possible with nitrogen self-sufficiency. The energy selfsufficiency described in these studies comes at the expense of increased land demand or lower yields, though a fact that is not emphasized in these studies but that is crucial for our more encompassing assessment of sustainable bioenergy production. Fredriksson et al. (2006) find 4e10% increased land demand for energy self-sufficiency of the farm. We emphasize that self-sufficiency means that such a farm does not produce any energy for the wider society. In Fredriksson et al. (2006), this is achieved with utilization of firstgeneration bioenergy, thus the agronomy is similar to ordinary food production and biomass exports are also similar. Halberg et al. (2008) achieve energy selfsufficiency and improved nutrient availability by using land that has been set-aside in the baseline (8.5% of total farmland) for energy production. It is not discussed which environmental effects this has. Pugesgaard et al. (2013) use 10e20% of the farm area for biogas feedstock production and report lower food yields. Either are milk yields reduced by more than 50% due to lower cattle numbers (while cash crop yields are increased by 60e120% due to improved N fertilization of cash crops), or cash crop yields are reduced by 10e30%. The scenario with 120% increased cash crop utilizes additional 20%
REFERENCES farmland of meadows and is thus not fully comparable to the baseline. Also in this case, energy production thus comes at the expense of lower yields or higher land use. A clear assessment of what this means regarding food security is however not possible, as the differences should be translated in total calorie and protein provision for human nutrition. Interesting though is the fact that part of this energy provision is possible in scenarios that go along with some dietary change only, as animal products are reduced. CONCLUSIONS Our analysis shows that bioenergy without land competition is difficult. While general land use models exhibit quite some potential for bioenergy production also under several sustainability constraints, they lack a due assessment of nutrient use, supply and demand in the agricultural production phase. On-farm studies reveal that increased land use or reduced yields cannot be avoided even for moderate bioenergy generation (e.g. to make a farm energy self-sufficient) unless only biogas is produced. We draw several conclusions from this assessment of sustainable farming of bioenergy crops. First, for a thorough assessment of the sustainability of bioenergy, systemic views have to be adopted. It is not enough to assess the GHG balance on a life-cycle basis. Bioenergy as a climate change mitigation strategy needs to be analyzed in the context of the whole food system including agricultural production. Much work has been done in this direction. Land use modeling and also sustainability criteria for bioenergy account for a wide range of aspects, such as the competition for land. However, as a second point, we want to emphasize that fertilization and nutrient cycles play a minor role in the assessment of bioenergy and its sustainable production only. This is a significant lack in analysis, as biomass plays a key role as fertilizer in sustainable agricultural production systems and as feedstock for bioenergy production. Agronomic aspects of crop fertilization and nitrogen use need to play a significant role in sustainability assessments of bioenergy. Third, we may point out biogas production as one viable option, where biomass can in principle be used for both ends at the same timedas feedstock for biogas plants and as fertilizer in the form of biogas slurry, after having passed through the biogas digester. Biogas production can be designed in such a way that it fits into agricultural production systems without additional land demand. However, as promising as it is for local energy generation, the aggregate potential remains small. In addition, it is no option for producing liquid biofuels. 415 Fourth, land competition is a key challenge, in particular for liquid biofuel production. Many models to assess the bioenergy potential globally or regionally exist, but they should be improved by adding much more detailed interaction with the energy markets. Such models need to be able to capture land use allocation based on the relative profitability of energy or food production. Most models focus on assessing physical potentials which is a key basis for this, and they mention economic constraints for developing the technical bioenergy potential, but how strong a land competition will emerge hinges on such relative profitability, resp. prices and on demand and supply elasticities, i.e. how much demand and supply changes with prices. In addition, these land use models need to incorporate agronomic aspects. Nitrogen demand of energy crops, corresponding fertilizer demand, its environmental effects and linkages between yields and nutrient inputs need to be captured in much more detail to arrive at reliable conclusions. If it comes to assessing bioenergy potentials in the context of sustainable agricultural production systems, the need to capture fertilizer and nutrient dynamics in more detail is directly linked to biomass flows that must be captured adequately between energy and fertilizer use. Fifth, some improved standard for sustainable bioenergy could help in this. We thus suggest to combine the RSB (2011) and GEF et al. (2013) standards and to enhance them with agronomic aspects related to nutrient and biomass use and recycling. References Alexandratos, N., Bruinsma, J., 2012. World Agriculture Towards 2030/2050. The 2012 Revision, ESA Working Paper No. 12-03. FAO, Rome. Altieri, M., 1995. Agroecology: The Science of Sustainable Agriculture, second ed. Westview Press, Boulder, CO. Andrade de Sá, S., Di Falco, S., Palmer, C., 2013. Dynamics of indirect land-use change: empirical evidence from Brazil. J. Environ. Econ. Manage. Anspach, V., 2009. Status quo, Prespektiven und wirtschafltiche Potenziale der Biogaserzeugung auf landwirtschaftlichen Betrieben im ökologischen Landbau (Ph.D. thesis). University of Kassel. Badgley, C., Moghtader, J., Quintero, E., Zakem, E., Chappell, M., Aviles-Vazquez, K., Samulon, A., Perfecto, I., 2007. Organic agriculture and the global food supply. Renewable Agric. Food Syst. 22, 86e108. Bajwa, W.I., Kogan, M., 2002. Compendium of IPM Definitions (CID)dWhat Is IPM and How Is It Defined in the Worldwide Literature?. IPPC Publication no. 998. Integrated Plant Protection Center (IPPC), Oregon State University, Corvallis, OR. Bindraban, P., Bulte, E., Conjin, S., Eickhout, B., Hoogwijk, M., Londo, M., 2009. Can biofuels be sustainable by 2020? An assessment for an obligatory blending target of 10% in the Netherlands, Scientific Assessment and Policy Analysis Report WAB 500102 024, Research Programme on Scientific Assessment and Policy Analysis for Climate Change (WAB).
416 23. SUSTAINABLE FARMING OF BIOENERGY CROPS Blanco-Canqui, H., Lal, R., 2009. Corn stover removal for expanded uses reduces soil fertility and structural stability. Soil Sci. Soc. Am. J. 73 (2), 418e426. Blonz, J., Vajjhala, S., Safirova, E., 2008. Growing Complexities: A Cross-sector Review of U.S. Biofuels Policies and Their Interactions. RFF Discussion Paper 08e47. Resources for the Future RFF, Washington DC. BMU, 2011. Leitstudie 2010-Langfristszenarien und Strategien für den Ausbau der erneuerbaren Energien in Deutschlandbei Berücksichtigung der Entwicklung in Europa und global. Bundesministerium für Umwelt, Naturschutz und Reaktorsicherheit. Carter, S., Hauggaard-Nielsen, H., Heiske, S., Jensen, M., Thomsen, S., Schmidt, J., Johansen, A., Ambus, P., 2012. Consequences of field N2O emissions for the environmental sustainability of plant-based biofuels produced within an organic farming system. Global Change Biol. Bioenergy 4 (4), S. 435e452. Chen, C., Yeh, K., Aisyah, R., Lee, D., Chang, J., 2011. Cultivation, photobioreactor design and harvesting of microalgae for biodiesel production: a critical review. Bioresour. Technol. 102, 71e81. Christen, B., Dalgaard, T., 2013. Buffers for biomass production in temperate European agriculture: a review and synthesis on function, ecosystem services and implementation. Biomass Bioenergy 55, S. 53e67. Chum, H., Faaij, A., Moreira, J., Berndes, G., Dhamija, P., Dong, H., Gabrielle, B., Goss Eng, A., Lucht, W., Mapako, M., Masera Cerutti, O., McIntyre, T., Minowa, T., Pingoud, K., 2011. Bioenergy. In: Edenhofer, O., Pichs-Madruga, R., Sokona, Y., Seyboth, K., Matschoss, P., Kadner, S., Zwickel, T., Eickemeier, P., Hansen, G., Schlömer, S., von Stechow, C. (Eds.), IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Cramer, J., Wissema, E., de Bruijne, M., Lammers, E., Dijk, D., Jager, H., van Bennekom, S., Breunesse, E., Horster, R., van Leenders, C., Wonink, S., Wolters, W., Kip, H., Stam, H., Faaij, A., Kwant, K., 2007. Testing Framework for Sustainable Biomass, EnergieTransitie, the Netherlands. http://www.senternovem.nl/ energietransitiegg/documentatie/downloads_rapporten_en_achter grondinformatie.asp. Damen, B., 2010. BEFS Thailand e Key Results and Policy Recommendations for Future Bioenergy Development. FAO Environment and Natural Resources Working Paper 43. De Gorter, H., Tsur, Y., 2009. Towards a Genuine Sustainability Standard for Biofuel Production. Working Paper of the Department of Applied Economics and Management. Cornell University, NY. Delucchi, M., 2010. Impacts of biofuels on climate change, water use, and land use. Ann. N. Y. Acad. Sci. 1195, 28e45. Delzeit, R., Gömann, H., Holm-Müller, K., Kreins, P., Kretschmer, B., Münch, J., Peterson, S., 2010. Analysing Bioenergy and Land Use Competition in a Coupled Modelling System: the Role of Bioenergy in Renewable Energy Policy in Germany. Kiel Working Papers No 1653. De Ponti, T., Rijk, B., van Ittersum, M., 2012. The crop yield gap between organic and conventional agriculture. Agric. Syst. 108, 1e9. DFID, 2004. Agricultural Sustainability. Working Paper Nr. 12. UK Department of International Development (DFID), London. http:// dfid-agriculture-consultation.nri.org/summaries/wp12.pdf. Elbehri, A., Segerstedt, A., Liu, P., 2013. Biofuels and the Sustainability Challenge. Food and Agriculture Organization of the United Nations FAO. Erisman, J., van grinsven, H., Leip, A., Mosier, A., Blecker, A., 2010. Nitrogen and biofuels; an overview of the current state of knowledge. Nutr. Cycling Agroecosyst. 86, 211e223. EU, 2009. Directive 2009/30/EC of the European parliament and of the council of 23 April 2009. Off. J. Eur. Union. L. 140/88, 5.6. 2009. Evenson, R.E., Gollin, D., 2003. Assessing the impact of the green revolution, 1960e2000. Science 300, 758e762. Eyhorn, F., Heeb, M., Weidmann, G., 2003. IFOAM Training Manual for Organic Agriculture in the Tropics. International Federation of Organic Agriculture Movements (IFOAM). http://www.fibl.org/ english/publications/training-manual/. Faist Emmenegger, M., Gmünder, S., Reinhard, J., Zah, R., Nemecek, T., Schnetzer, J., Bauer, C., Simons, A., Doka, G., 2012. Harmonisation and Extension of the Bioenergy Inventories and Assessment. End Report to the Swiss Federal Office of energy, 31th August 2012. FAO, 2002. In: El-Hage Scialabba, N., Hattam, C. (Eds.), Organic Agriculture, Environment and Food Security. Food and agriculture organisation of the United Nations FAO. http://www.fao.org/ docrep/005/y4137e/y4137e00.htm. FAOSTAT, 2013. Food Balance Sheets, Food and Agriculture Organization of the United Nations FAO, Rome. Fargione, J., Hill, J., Tilman, D., Polasky, S., Hawthorne, P., 7 February 2008. Land clearing and the biofuel carbon Debt. Sci. Express. http://dx.doi.org/10.1126/science.1152747. Fredriksson, H., Baky, A., Bernesson, S., Nordberg, A., Noren, O., Hansson, P.-A., 2006. Use of on-farm produced biofuels on organic farms e evaluation of energy balances and environmental loads for three possible fuels. Agric. Syst. 89, 184e203. Galloway, J., Townsend, A., Erisman, J., Bekunda, M., Cai, Z., Freney, J., Martinelli, L., Seitzinger, S., Sutton, M., 2008. Transformation of the nitrogen cycle: resent trends, questions, and potential Solutions. Science 320 (5878), 889e892. Gattinger, A., Muller, A., Häni, M., Skinner, C., Fliessbach, A., Buchmann, N., Mäder, P., Stolze, M., Smith, P., El-Hage Scialabba, N., Niggli, U., 2012. Enhanced top soil carbon stocks under organic farming. Proc. Natl. Acad. Sci. USA. 109 (44), 18226e18231. http://dx.doi.org/10.1073/pnas.1221886110. GEF/UNEP/FAO/UNIDO, 2013. Global Assessments and Guidelines for Sustainable Liquid Biofuel Production in Developing Countries. GEF, UNEP, FAO, UNIDO. http://energy-l.iisd.org/news/ experts-assess-biofuels-screening-toolkit-at-unido-biofuelsconference/. Godfray, H., Beddington, J., Crute, I., Haddad, L., Lawrence, D., Muir, J., Pretty, J., Robinson, S., Thomas, S., Toulmin, C., 2010. Food security: the challenge of feeding 9 billion people. Science 327 (6). Halberg, N., Dalgaard, R., Olesen, J.E., Dalgaard, T., 2008. Energy selfreliance, net-energy production and GHG emissions in Danish organic cash crop farms. Renewable Agric. Food Syst. 23 (1), 30e37. Hennecke, A., Faist, M., Reinhardt, J., Junquera, V., Neeft, J., Fehrenbach, H., 2013. Biofuel greenhouse gas calculations under the European renewable energy directive e a comparison of the bioGrace tool vs the tool of the roundtable on sustainable biofuels. Appl. Energy 102, 55e62. http://www.sciencedirect.com/science/ article/pii/S0306261912003066. HLPE, 2013. Biofuels and Food Security. High Level Panel of Experts on Food Security and Nutrition HLPE, Committee on World Food Security. IAASTD, 2009. Agriculture at the Crossroads, International Assessment of Agricultural Scinece and Technology for Development (IAASTD), Island Press, Washington DC. IEA, 2010. Sustainable Production of Second-generation Biofuels. International Energy Association (Information Paper). IEA Bioenergy, 2011. Bioenergy, Land Use Change and Climate Change Mitigation (Background report).
REFERENCES IFAD, 2001. Rural Poverty Report 2001: The Challenge of Ending Rural Poverty. International Fund for Agricultural Development (IFAD). http://www.ifad.org/poverty/. IFOAM, 2006. The Principles of Organic Agriculture. International Federation of Organic Agriculture Movements (IFOAM). http:// www.ifoam.org/about_ifoam/principles/index.html. Karlsson, G., Banda, K. (Eds.), 2009. Biofuels for Sustainable Rural Development and Empowerment of Women e Case Studies from Africa and Asia. Energia. Lal, R., 2008. Crop residues as soil amendments and feedstock for bioethanol production. Waste Manage. 28 (4), 747e758. Lysen, E., van Egmond, S. (Eds.), 2008. Biomass AssessmentdAssessment of Global Biomass Potentials and Their Links to Food, Water, Biodiversity, Energy Demand and Economy. Report 500102 012, Netherlands Research Programme on Scientific Assessment and Policy Analysis for Climate Change (WAB) http://www.rivm.nl/bibliotheek/rapporten/500102012.pdf. Matson, P.A., Parton, W.J., Power, A.G., Swift, M.J., 1997. Agricultural intensification and ecosystem properties. Science 277, 504e509. Meehl, G.A., Stocker, T.F., Collins, W.D., Friedlingstein, P., Gaye, A.T., Gregory, J.M., Kitoh, A., Knutti, R., Murphy, J.M., Noda, A., Raper, S.C.B., Watterson, I.G., Weaver, A.J., Zhao, Z.-C., 2007. Global climate projections. In: Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L. (Eds.), Climate Change 2007: The Physical Science Basis. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Meyer, R., Priefer, C., 2012. Ökologischer Landbau und Bioenergieerzeugung e Zielkonflikte und Lösungsansätze. Büro für Technikfolgenabschätzung beim deutschen Bundestag. Arbeitsbericht Nr. 151. Müller, C., Bondeau, A., Popp, A., Waha, K., Fader, M., 2010. Climate Change Impacts on Agricultural Yields. Background Note to the World Development Report 2010. Muller, A., 2009. Sustainable agriculture and the production of biomass for energy use. Clim. Change 94, 319e331. Nhantumbo, I., Salomão, A., 2010. Biofuels, Land Access and Rural Livelihoods in Mozambique. International Institute for Environment and Development IIED, London. NYT, February 1, 2007. Thousands in Mexico City protest rising food prices. N. Y. Times. Oleskowicz-Popiel, P., Kádár, Z., Heiske, S., Klein-Marcuschamer, D., Simmons, B., Blanch, H., Schmidt, J., 2012. Co-production of ethanol, biogas, protein fodder and natural fertilizer in organic farming e evaluation of a concept for a farm-scale biorefinery. Bioresour. Technol. 104, 440e446. Palola, E., Walker, N., 2010. Food, Fuel, or Forests? Charting a Responsible U.S. Role in Global Palm Oil Expansion. National Wildlife Federation. PBL, 2010. How to Deal with Indirect Landuse Change in the EU Renewable Energy Directive? Netherlands Environmental Assessment Agency PBL. Practical Action Consulting, January 2009. Small-Scale Bioenergy Initiatives: Brief Description and Preliminary Lessons on Livelihood Impacts from Case Studies in Asia, Latin America and Africa. Prepared for PISCES and FAO by Practical Action Consulting. 417 Pragya, N., Pandey, K., Sahoo, P., 2013. A review on harvesting, oil extraction and biofuels production technologies from microalgae. Renewable Sustainable Energy Rev. 24, 159e171. Pretty, J.N., Noble, A.D., Bossio, D., Dixon, J., Hine, R.E., Penning de Vries, F., Morison, J., 2006. Resource conserving agriculture increases yields in developing countries. Environ. Sci. Technol. 40 (4), 1114e1119. Pugesgaard, S., Olesen, J., Dalgaard, T., Jørgensen, U, 2013. Biogas in organic agriculture - effects on productivity, energy selfsufficiency and greenhouse gas emissions. Renewable Agric. Food Syst. Rathman, R., Szklo, A., Schaeffer, R., 2010. Land use competition for production of food and liquid biofuels: an analysis of the arguments in the current debate. Renewable Energy 35, 14e22. Ribeiro, D,Matavel, N., 2009. Jatropha! A socio-economic pitfall for Mozambique, Justiça Ambiental & União Nacional de Caponeses, Alliance Sud, Arbeitsgruppe Schweiz Kolumbien, Basler Appell gegen Gentechnologie, Bio Suisse, Brot für Alle, Caritas, erklärung von Bern, Fastenopfer, HEKS, Kleinbauern-Vereinigung, l Pro Natura, Reformierte Kirchen, SWISSAID, Terre des Hommes, Uniterre. Rossi, A., 2012. Good Environmental Practices in Bioenergy Feedstock Production e Making Bioenergy Work for Climate and Food Security. FAO Environment and Natural Resources Management. Working Paper Nr. 49. RSB, 2011. Consolidated RSB EU RED Principles & Criteria for Sustainable Biofuel Production. Roundtable of Sustainable Biofuels RSB. http://rsb.org/pdfs/standards/RSB-EU-RED-Standards/1105-10-RSB-STD-11-001-01-001-vers-2-0-Consolidated-RSB-EU-REDPCs.pdf. Schader, C., Stolze, M., Gattinger, A., 2012. Environmental performance of organic agriculture. In: Boye, J., Arcand, Y. (Eds.), Green Technologies in Food Production and Processing. Springer, New York. Seufert, V., Ramankutty, N., Foley, J., 2012. Comparing the yields of organic and conventional agriculture. Nature. Smith, P., Martino, D., Cai, Z., Gwary, D., Janzen, H.H., Kumar, P., McCarl, B., Ogle, S., O’Mara, F., Rice, C., Scholes, R.J., Sirotenko, O., Howden, M., McAllister, T., Pan, G., Romanenkov, V., Schneider, U., Towprayoon, S., Wattenbach, M., Smith, J.U., 2008. Greenhouse gas mitigation in agriculture. Philos. Trans. R. Soc. Tomei, J., Upham, P., 2009. Argentinean soy-based biodiesel: an introduction to production and impacts. Energy Policy 37, 3890e3898. UNEP, 2010. Gain or Pain e Biofuels and Invasive Species. Bioenergy Issue Paper Series Nr. 3. United Nations Environmental Programme UNEP. wa Gathui, T., Ngugi, W., 2010. Bioenergy and Poverty in Kenya: Attitude, Actors and Activities (Policy Innovation Systems for Clean Energy Security PISCES Working Paper). WBGU, 2009. Welt im Wandel. Zukunftsfähige Bioenergie und nachhaltige Landnutzung. Wissenschaftlicher Beirat Globale Umweltveränderung der Bundesregierung WBGU, Berlin. Wicke, B., Verweij, P., van Meij, H., van Vuuren, D., Faaij, A., 2012. Indirect land use change: review of existing models and strategies for mitigation. Biofuels 3 (1), 87e100.
C H A P T E R 24 Bioenergy Technology and Food Industry Waste Valorization for Integrated Production of Polyhydroxyalkanoates Vasiliki Kachrimanidou 1, Nikolaos Kopsahelis 1, Colin Webb 2, Apostolis A. Koutinas 1,* 1 2 Department of Food Science and Human Nutrition, Agricultural University of Athens, Athens, Greece, School of Chemical Engineering and Analytical Science, University of Manchester, Manchester, England, United Kingdom *Corresponding author email: akoutinas@aua.gr O U T L I N E Introduction 419 PHA Structure and Properties 420 PHA Production Integrated in Biorefinery Concepts 421 Valorization of Biodiesel Industry By-Products 424 Valorization of Second-Generation Bioethanol 427 Industry By-Products Valorization of By-Product Streams from 427 Food Industries INTRODUCTION The imminent depletion of fossilized raw materials and increasing environmental concerns has paved the way toward the development of a sustainable bio-based economy. Biorefinery concepts constitute a significant aspect of the future bioeconomy era where renewable raw materials, such as widely available lignocellulosic biomass in conjunction with industrial by-products and waste streams, will be utilized for the production of value-added commercial products, including biofuels, chemicals, biodegradable polymers and antioxidants among others. However, the establishment of a new industrial sector is a difficult task not only Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00024-3 PHA Production from Winery By-Products PHB Production from Confectionery and Bakery Industry Waste Streams PHB Production from Whey 428 429 430 Conclusions and Future Perspectives 430 References 430 because of the viability of new technological advances but also because of the transition from the nonrenewable to the sustainable era should occur smoothly in order to avoid job losses and economic turmoil. A smooth transition can be achieved through the integration of sustainable processing schemes in those conventional industrial plants that generate waste and by-product streams suitable for bioconversion or green chemical conversion into value-added products. Petroleum-derived plastics constitute an everyday commodity used mainly for packaging and disposable materials. According to the U.S. Environmental Protection Agency, 31 million t of waste was generated in 2010, while only 8% of the total plastic waste was 419 Copyright Ó 2014 Elsevier B.V. All rights reserved.
420 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION recovered for recycling (Anonymous, 2012a). Other conventional waste management approaches, such as incineration and discarding to landfills that are currently applied for plastic disposal, are not considered as sustainable options. Bearing in mind that plastic consumption rates are constantly increasing, it is nowadays common knowledge that the only way to deal with petroleum-derived plastics is the development of biodegradable counterparts that serve the same, or even more, market outlets and societal needs. The production of fully biodegradable polymers can be achieved either via microbial production of small molecules (e.g. organic acids and alcohols) that could be subsequently used as monomers for polymer synthesis or through direct fermentative production of tailormade biodegradable polymers. Polyhydroxyalkanoates (PHAs) are biodegradable polyesters accumulated intracellularly during fermentation that belong to the second category. The wide versatility of PHA properties coupled with their potential to substitute major petroleum-derived plastics has driven research studying their synthesis and exploring industrial implementation for more than 30 years. Despite the numerous attempts to create industrial processes for PHA production by many manufacturers in different countries, their industrial production is still limited due mainly to high production costs that should coincide with the production of tailor-made PHAs possessing the required thermophysical and mechanical properties. The high production cost is mainly dependent on the use of commercial sources of carbon and nutrients and the significant cost required for downstream separation and purification of PHAs from microbial mass. Moreover, fermentation parameters, such as carbon source to PHA conversion yield, productivity, final PHA concentration and PHA content, are critical for the development of highly efficient production processes. Consequently, the use of low- or even no-cost waste and by-products streams could provide alternative renewable feedstocks to reduce process costs. In recent years, research focuses on the evaluation of various low-cost renewable resources that may lead to reduction of PHA production costs. However, it is nowadays clear that traditional fermentation processes should be restructured and integrated in advanced biomass refining schemes in order to reduce process economics, improve environmental impact, produce many products for different markets and create synergies between different industrial sectors. This study reports recent efforts targeting PHA production from low-cost renewable resources and proposes potential biomass refining schemes that could be developed for simultaneous production of fuels, chemicals, value-added products and PHAs. Such biorefinery schemes could be either developed in new industrial plants or integrated in existing industrial plants through the valorization of waste or by-product streams. PHA STRUCTURE AND PROPERTIES PHAs are a group of polyesters produced intracellularly as carbon and energy reserve granules. Intracellular accumulation of PHAs is usually observed when one nutrient (e.g. nitrogen, phosphorus, and oxygen) is present in the fermentation broth in limiting concentration, while, in the same time, there is an available excess source of carbon. Most bacterial strains, such as Cupriavidus necator (formerly classified as Ralstonia eutropha and Alcaligenes eutrophus), accumulate PHAs as secondary products under nutrient-limiting conditions. However, there are some bacterial strains, such as recombinant Escherichia coli and Alcaligenes latus, that synthesize PHAs as a primary metabolite during microbial growth. The generic chemical formula of PHAs is presented in Figure 24.1, where R represents the alkyl side groups of varying chain lengths. The most common monomers are 3-hydroxybutyrate (3HB, m ¼ 1 and R]CH3), 3hydroxyvalerate (3HV, m ¼ 1 and R]CH2CH3) and 3-hydroxyhexanoate (3HHx, m ¼ 1 and R]CH2CH2CH3). The simplest and most widely studied member of the PHA family is called poly-(3-hydroxybutyrate) (PHB). Three of the most widely studied copolymers are poly-(3hydroxybutyrate-co-3-hydroxyvalerate) [P(3HB-co-3HV), y ¼ 1], poly-(3-hydroxybutyrate-co-3-hydroxyhexanoate) [P(3HB-co-3HHx), y ¼ 2] and poly-(3-hydroxybutyrate-co4-hydroxybutyrate) [P(3HB-co-4HB]. Recent research has also focused on the production of terpolymers, such as P(3HB-co-4HB-co-3HV) and P(3HB-co-3HV-co-3HHx) O R CH (CH2)m C O CH3 ΟH Ο CH2 CH C ΟH Generic formula of hydroxy alkanoates (HA) n PHB CH3 Ο CH (CH2)y O CH3 CH2 C Ο CH O CH2 n C x P(3HB-co-3HA) O CH3 Ο CH CH2 O C Ο n CH2 CH2 CH2 C x P(3HB-co-4HB) FIGURE 24.1 Generic chemical formula of HA monomers and common PHA copolymers.
PHA PRODUCTION INTEGRATED IN BIOREFINERY CONCEPTS (Bhubalana et al., 2010; Cavalheiro et al., 2012). PHAs can be categorized according to the number of carbon atoms into short-chain-length PHAs (monomers containing three to five carbon atoms) and medium-chain-length PHAs (monomers containing more than six carbon atoms) (Asbhy et al., 2011; Du et al., 2012). A historical overview of PHA research and industrial applications is presented in previous publications (Solaiman et al., 2006; Verlinden et al., 2007; Du et al., 2012). As previously mentioned, one of the most important advantages of PHAs is their biodegradability and biocompatibility. Under aerobic conditions, PHAs are degraded to carbon dioxide and water, while under anaerobic conditions, methane and water are the final products. Hence, these compounds can be utilized from various microorganisms living in soil and water as carbon source for their growth, without toxic effects to the environment. PHB was the first member of the PHA family that was identified after isolation from Bacillus megaterium (Lemoigne, 1926). It can be produced by many bacterial strains (especially various strains of C. necator) in high concentrations (more than 150 g/l) and intracellular content (more than 80% on a dry weight basis) from commercial carbon sources (mainly glucose as well as fructose and sucrose) and starch hydrolysates (Ryu et al., 1997; Yu et al., 2003). In addition, the physical properties of PHB are similar to polypropylene. However, the brittle and thermally unstable nature of PHB limits its commercial applications and constitutes one of the major reasons that have prevented its production in large-scale operations. The high crystallinity of PHB (55e80%), associated with the formation of large spherulites, is the main reason that causes the brittle nature of PHB. It should be stressed though that the application of appropriate processing methodologies could reduce the undesirable mechanical properties of PHB, which could be used for the production of ductile films (Barham et al., 1992). Furthermore, the molecular weight of the PHB homopolymer produced by many bacterial strains, under varying fermentation condition and utilization of different feedstocks may also result in a biopolymer with improved characteristics (Kusaka et al., 1999). P(3HB-co-3HV) was the first copolymer of the PHA family that was identified and subsequently produced on industrial scale by Imperial Chemical Industries using a R. eutropha strain. The incorporation of 3HV units in different proportions in the copolymer by this R. eutropha strain was only possible after the addition of propionic acid as a carbon source precursor that induced the metabolic synthesis of 3HV units. The production of P(3HB-co-3HV) also demonstrated that it is feasible to alter the properties of PHAs by controlling fermentation conditions. For instance, 421 the addition of increasing propionic acid concentrations during PHA accumulation results in increasing proportions of 3HV units (expressed as mol%) in the P(3HB-co-3HV) copolymer. In this way, it was demonstrated that the incorporation of 3HV units in the P(3HB-co-3HV) copolymer results in improved mechanical properties (Byrom, 1987; Choi and Lee, 1997). Since the identification and commercial production of P(3HB-co-3HV) copolymers, research has focused on the identification or modification of microbial strains capable of producing PHA copolymers without addition of carbon source precursors or the production of different PHA copolymers, with addition of carbon source precursors, containing two, three or four monomers that demonstrate desired properties (Madden et al., 2000; Loo et al., 2005; Koller et al., 2007a). For instance, the archeon Haloferax mediterranei accumulates 72.8% (w/w) of P-(3HB-co-3HV) that contains 6 mol% 3HV units directly from whey sugars, while it produces the terpolymer P(3HB-co-3HV-co-4HB) when it is supplemented with 3HV and 4HB precursors (Koller et al., 2007a). Table 24.1 presents the specific properties of various PHAs compared with major petroleum-derived plastics. Nowadays, it is widely accepted that PHA physical properties can vary from brittle PHB homopolymers with high crystallinity to flexible PHA copolymers with lower crystallinity, such as P(3HB-co-3HV) and P(3HB-co-3HHx), to elastic PHA copolymers, such as P(3HB-co-4HB) and P(3hydroxyoctanoate-co-3-hydroxydecanoate) (Wolf et al., 2005; Whitehouse et al., 2006). In the last 30 years, PHAs have been identified as potential biopolymers for a wide spectrum of end-uses including food packaging, flushable hygiene products, tissue engineering applications, adhesives, agriculture and biocomposites (Wolf et al., 2005). PHA PRODUCTION INTEGRATED IN BIOREFINERY CONCEPTS It is nowadays widely recognized that successful implementation of industrial PHA production will only be achieved through satisfaction of sustainability aspects coupled with production of biodegradable polymers with desirable properties. Sustainability aspects include cost-competitiveness, environmental benignness and production of biodegradable polymers that serve certain market and societal needs. Additional advantages will be provided through the ability to produce PHAs with adjustable properties that could be used in different end-uses by simple modification of fermentation conditions. For instance, the production of different types of PHAs that could be used for both commodity
422 TABLE 24.1 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION Thermophysical Properties of Various PHAs Polymer Co-Monomer (mol%) Tg* ( C) Tmx ( C) DHm{ (J/g) Xc** % Td(50%)xx ( C) References PHB e 4 177 84 60  5 e Shimamura et al. (1994) PHB e 4 180 e e e Akaraonye et al. (2010) 2.3 163 e e 252 Han et al. (2010) 1.2 169 49.00 33 e Lee et al. (2008) P(3HB-co-3HV) P(3HB-co-3HV) {{ 4.1 {{ 8 {{ P(3HB-co-3HV) 12 e 153 66.40 45 e Gunaratne and Shanks (2005) P(3HB-co-3HV) 12{{ 1.9 155 72.61 49 271 Garcia et al. (2013) {{ P(3HB-co-3HV) 30 2.1 111 e e 352 He et al. (2001) P(3HB-co3HHx) 3.5*** 1 140 and 151 44 e e Tsuge et al. (2004) P(3HB-co3HHx) 5*** 1 to 3 125e138 and 142e155 38e47 e e Loo et al. (2005) P(3HB-co3HHx) 11*** 1 136 60 40  5 e Shimamura et al. (1994) P(3HB-co3HHx) 17*** 2 130 39 29  5 e Shimamura et al. (1994) P(3HB-co-4HB) 12xxx 4.3 124 56 38.2 270{{{ Luo et al. (2009) Polypropylene e 10 176 e 70 e Akaraonye et al. (2010) Polystyrene e 100 240 e e e Akaraonye et al. (2010) * Glass transition temperature. x Melting temperature. { Enthalpy of fusion. ** Percentage of crystallinity. xx Degradation temperature. {{ 3-hydroxyvalerate. *** 3-hydroxyhexanoate. xxx 4-hydroxybutyrate. {{{ Td(5%). (e.g. food packaging) and specialty (e.g. scaffolds for tissue engineering applications) end-uses by simple modification of fermentation parameters could provide process flexibility. An important innovation on future PHA-based processes will be the creation of cascade processing schemes in order to increase resource efficiency (Anonymous, 2012b). Cascade processing is based on the reutilization of packaging material after its use (also called postconsumer plastics) for other commercial purposes. For instance, hydrolysis into monomers could create valueadded platform molecules for the chemical industry. In addition, bioplastics could be used as replacements for coal and heating fuel due to their high calorific value (Anonymous, 2012b). Reutilization of PHA-based packaging materials is strongly dependent on the development of suitable recycling technologies. Despite their significant advantages, industrial production of PHAs is hindered by high production cost. Previous attempts to produce PHAs in large scale had to rely on conventional fermentation technologies that cannot compete with low-cost petroleum-derived plastics. As mentioned earlier, raw material supply is one of the most important factors that should be optimized in order to reduce processing costs. For this reason, recent research focuses on the utilization of low-cost feedstock for PHA production (e.g. molasses, crude glycerol, whey, animal fats, and waste cooking oils among others) aiming to substitute for conventional and expensive carbon sources. Table 24.2 presents the results regarding PHA production from various waste and by-product streams. However, even if waste or byproduct streams are used as fermentation feedstocks, aerobic cultivation for PHA production in industrial scale operations is still an expensive unit operation. For this reason, integration of PHA production into existing industrial plants or the development of new industrial plants for PHA production should be combined
TABLE 24.2 PHA Production from Various Crude Renewable Resources, Waste and By-Product Streams By-Product or Waste Stream Type of PHA Strain Maximum Cell Weight Max PHA Concentration (g/l) PHA Content (%) Productivity (g/l h) References 38 0.84 Cavalheiro et al. (2009) Cupriavidus necator DSM 545 PHB 68.8 26.1 Bagasse hydrolysates Ralstonia eutropha PHA 11.1 6.3 56.5 e Yu and Stahl (2008) Crude glycerol and rapeseed hydrolysates Cupriavidus necator DSM 545 P(3HB-co-3HV) 19.6 10.9 55.6 0.12 Garcia et al. (2013) Wheat-derived media (shake flask cultures) Cupriavidus necator NCIMB 11599 PHB 73.2 51.1 70 0.3 Koutinas et al. (2007b) Wheat-derived media (bioreactor cultures) Wautersia eutropha NCIMB 11599 PHB 175.2 162.8 93 0.89 Xu et al. (2010) Soybean oil Ralstonia eutropha H16 PHB 126 95.8 76 0.99 Kahar et al. (2004) Ralstonia eutropha PHB 4 (DSM 541) P(3HB-co-3HHx) 138 102.1 74 1.06 Oleic acid Pseudomonas putida PGA1 PHAs-mcl 30.2 44.8 0.19 Marsudi et al. (2007) Hydrolyzed whey Haloferax mediterranei DSM 1411 PHA 11 5.5 50 0.05 Koller et al. (2007b) Pseudomonas hydrogenovora DSM 1749 10.83 1.3 12 0.03 Hydrogenophaga pseudoflava DSM 1034 6.75 2.7 40 0.05 - 13.52 Pseudomonas hydrogenovora DSM 1749 PHB 10.58 1.27 12 0.03 P(3HB-co-3HV) 12 1.44 12 0.05 Cheese whey Methylobacterium sp. ZP24 PHB 3.54 64 0.09 Nath et al. (2008) Saccharified waste potato starch Ralstonia eutropha NCIMB 11599 PHB 179 94 55 1.47 Haas et al. (2008) Extruded rice bran and extruded corn starch Haloferax mediterranei ATCC 33500 PHB 140 77.8 55.6 0.65 Huang et al. (2006) Sugarcane molasses and corn steep liquor Bacillus megaterium PHB 3.6 2.2 59.4 e Gouda et al. 2001 Sugarcane molasses and urea Bacillus megaterium BA-019 PHB 72.6 30.5 42 1.27 Kulpreecha et al. 2009 Hydrolyzed whey permeate Hydrolyzed whey permeate and valerate 5.53 Koller et al. (2008) PHA PRODUCTION INTEGRATED IN BIOREFINERY CONCEPTS Waste glycerol 423
424 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION with the production of value-added co-products. This can be achieved through fractionation of agricultural resources or by-product and waste streams from existing industrial processes. PHA production cost increases further due to downstream separation and purification of PHAs from residual microbial mass. Several methods have been reported for the recovery of PHAs based on the utilization of organic solvents such as acetone, chloroform or dichloroethane. However, these methods are unfavorable for large-scale production since solvents increase operational cost and additional equipment for solvent recovery is often needed. Alternative extraction methods have been also proposed including enzymatic lysis of residual microbial mass (Kapritchkoff et al., 2006; Verlinden et al., 2007), supercritical fluid extraction (Hejazi et al., 2003), mechanical disruption of bacterial cells coupled with chemical treatment, autolysis of bacterial cells, and chemical treatment under acidic or alkaline conditions (Yu and Chen, 2006; Verlinden et al., 2007). Several studies have also focused on the estimation of PHA production costs from different feedstocks (Choi and Lee, 1997; van Wegen et al., 1998; Posada et al., 2011). However, there are limited studies on the evaluation of integrated biorefineries focusing on the fractionation of the initial raw material combining the production of PHAs with the extraction or production of value-added co-products. In addition, future costing studies should also focus on the evaluation of the potential to integrate PHA production in existing industries. In recent years, several studies focused on the production of PHAs from low-cost renewable resources (Akaraonye et al., 2010; Koller et al., 2010; Du et al., 2012). This study focuses on the presentation of representative biorefinery concepts targeting the production of PHAs and other value-added products. In particular, PHA production could be combined with biofuel and bioenergy production. Valorization of Biodiesel Industry By-Products Both edible (e.g. rapeseed, soybean or palm) and nonedible (e.g. Jatropha) oilseeds can be used for biodiesel production. Biodiesel production is mainly achieved from soybean in the USA, rapeseed (or sunflower in lower quantities) in Europe and palm oil in South-East Asian countries. Biodiesel production could be also achieved using nonconventional resources including microbial oil produced by yeast, fungi and algae (Meng et al., 2009). The continuous growth of biodiesel production coincides with proportional production of by-products streams. The main by-product is glycerol that is generated during transesterification of triglycerides with predominantly methanol leading to the production of fatty acid methyl esters and glycerol (10%, w/w). It has been estimated that by 2021, the share of biodiesel production from vegetable oils will increase and the worldwide biodiesel production from such oils is projected to reach approximately 30  106 t (Anonymous, 2012c). This means that approximately 3  106 t of glycerol will be available for chemical and biopolymer production. Crude glycerol streams produced from biodiesel plants have purities in the range of 77e90% (w/w) (Mothes et al., 2007). The main impurities in crude glycerol are water, methanol, residual fatty acids and corresponding esters, and salts (NaCl or K2SO4) in varying proportions depending on the extent of glycerol purification (Mothes et al., 2007). Purification methods for glycerol have been proposed (Chatzifragkou and Papanikolaou, 2012) for the removal of impurities and salts after biodiesel production but they seem to be rather unprofitable, especially for small industries. Novel uses of glycerol involve both green chemical conversions and microbial bioconversions. Glycerol represents an easily assimilated carbon source for many microorganisms. Crude glycerol has been evaluated as carbon source for various microbial bioconversions, such as 1,3-propanediol, citric acid, ethanol, succinic acid, propionic acid, microbial oil and PHAs (Koutinas et al., 2007a; da Silva et al., 2009; Koutinas and Papanikolaou, 2011; Sarris et al., 2011). Biodiesel production from oilseeds leads to the production of oilseed meals as a valuable by-product stream. Oilseed meal is the protein- and carbohydraterich fraction that remains after the extraction of oil. The main conventional commercial outlet for oilseed meals is as animal feed. In the period 2012e2021, biodiesel production from edible vegetable oils will still rely mainly on rapeseed and sunflower. However, biodiesel production from palm oil is projected to increase twofold (Anonymous, 2012c). Based on recent estimates, approximately 315  106 t of oilseed meals are expected to be produced by 2021, corresponding to an increase up to 23% based on current production capacities (Anonymous, 2012c). Future biodiesel industries could be converted into novel biorefineries through valorization of crude glycerol and oilseed meal streams leading to the production of biodiesel, chemicals, food and feed ingredients and biopolymers such as PHAs. Ashby et al. (2004, 2011) evaluated the production and properties of PHAs accumulated by the bacterial strains Pseudomonas oleovorans NRRL B-14682 and Pseudomonas corrugata 388 cultivated on crude glycerol. Ashby et al. (2011) reported that the molecular weight of PHB was decreased with increasing methanol concentration in crude glycerol. Mothes et al. (2007) and Garcia et al. (2013) evaluated the effect of NaCl and K2SO4 on PHA production during fermentation with the bacterial strains Paracoccus denitrificans, C. necator JMP 134 and C. necator DSMZ 545. These salts are present in crude
PHA PRODUCTION INTEGRATED IN BIOREFINERY CONCEPTS glycerol depending on the catalyst (NaOH or KOH) employed during transesterification of triglycerides. The inhibition caused by NaCl on PHA production is more pronounced at significantly lower concentrations than K2SO4. Mothes et al. (2007) reported that bioreactor fermentations with C. necator JMP 134 cultivated on crude glycerol and inorganic chemicals as additional nutrients could lead to the production of PHB contents up to 70% (w/w). Crude glycerol has also been employed in bioreactor fermentations for the production of PHB using the bacterial strain C. necator DSM 545 leading to 50% (w/w) PHB content and 1.1 g/l h PHB productivity (Cavalheiro et al., 2009). Tanadchangsaeng and Yu (2012) stressed that the productivity (around 0.92 g/l ) of glycerol fermentation to PHB synthesis is lower than the one achieved from glucose. Crude glycerol could be also combined with other carbon sources that could be used as precursors for the production of PHA copolymers (Cavalheiro et al., 2012). The production of P(3HB-co-4HB-co-3HV) was reported when C. necator DSM 545 was cultivated on crude glycerol, propionic acid (stimulator of 4HB accumulation and 3HV precursor) and g-butyrolactone (4HB precursor). In all studies presented above, inorganic chemicals were used as nutrient supplements. Apart from fermentation efficiency of PHA production, it is also crucial to assess the properties of the polymer produced and the associated production cost. Tanadchangsaeng and Yu (2012) reported that although the thermal and physical properties of the PHB produced from glycerol is similar to the one produced from glucose, the molecular weight of the glycerolderived homopolymer is lower than the molecular weight of the PHB produced from glucose. Posada et al. (2011) presented a comparative technoeconomic evaluation of PHB production from crude glycerol using two different bacterial strains, C. necator and B. megaterium, and three different downstream separation strategies. Fermentation of C. necator resulted in the production of 81.6 g/l of which 57.1 g/l was PHB, significantly higher than B. megaterium. The most costcompetitive process involved PHB production in fedbatch fermentations with C. necator followed by PHB separation and purification with heat pretreatment, enzymatic alkaline digestion, centrifugation, washing, evaporation, and spray drying. Posada et al. (2011) reported also that glycerol purification to 98% (w/w) contributes approximately 6% of the overall PHB production cost, thereby slightly affecting the total cost. In this study, it was concluded that the PHB production cost from crude glycerol could be as low as US$2/kg depending on the downstream process utilized. PHA production from crude glycerol could be combined with the valorization of oilseed meals or residues remaining after extraction of microbial oil. For instance, 425 rapeseed meal could be utilized for the production of various value-added fractions including protein isolates, carbohydrates, hulls, phenolic compounds and glucosinolates with various applications such as animal feed, pesticidal agent, bioactive proteins, glues and adhesives, paper coatings and ingredients for cosmetics among others (Anonymous, 2011; Egues et al., 2010). Another alternative application of oilseed meals is based on the production of complex nutrient supplements for fermentation processes including PHA production. In this way, commercial inorganic chemicals will be replaced improving the sustainability of the whole biorefinery concept. Oilseed meals contain significant quantities of protein, minerals and other necessary nutrients for microbial growth. Enzymatic hydrolysis of protein to amino acids and peptides, and phytic acid to phosphorus could provide a hydrolysate suitable for PHA production. Crude enzymes could be produced via solid-state fermentation employing appropriate fungal strains and oilseed meals as substrates (Wang et al., 2010; Kachrimanidou et al. 2013). Wang et al. (2010) reported the production of a nutrient-rich hydrolysate from rapeseed meal with a free amino nitrogen content of 2016.2 mg/l and inorganic phosphorus (IP) of 304 mg/l that was subsequently used successfully as nutrient supplement combined with glucose as carbon source for the cultivation of Saccharomyces cerevisiae. Garcia et al. (2013) investigated the generation of a microbial feedstock through hydrolysis of rapeseed meal, which was combined with crude glycerol as the sole fermentation medium for PHA production. Fed-batch fermentations resulted in a production of 10.9 g/l P(3HB-co-3HV) without addition of any precursor. The properties of the biopolymer produced were also examined, leading to the conclusion that this bioprocess could be incorporated in rapeseed-based biodiesel plants contributing to the sustainability of biodiesel biorefineries. Kachrimanidou et al. (2013) reported the utilization of sunflower meal for the production of nutrient-rich hydrolysates that could be subsequently supplemented with crude glycerol for the production of 9.9 g/l P(3HB-co-3HV) with a PHA content of 50% (w/w) in shake-flask fermentations using the microbial strain C. necator DSM 545 without addition of any precursor. Figure 24.2 presents a biorefinery concept in which sunflower (or other oilseed) meal is utilized only for the production of fermentation feedstock involving production of crude enzymes via solid-state fermentation followed by hydrolysate production via enzymatic hydrolysis. Preliminary bioreactor fermentations carried out in fed-batch mode at the Agricultural University of Athens with the microbial strain C. necator DSM 7237 cultivated on sunflower meal hydrolysate and crude glycerol lead to the production of more than 20 g/l
426 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION FIGURE 24.2 Utilization of sunflower-derived Sunflower seed biodiesel industry by-products for PHA production. Mechanical pressing and hexane extraction Vegetable Oil Transesterification Sunflower meal Solid state fermentation Aspergillus oryzae Crude enzymes Biodiesel Cupriavidus necator Crude glycerol Enzymatic hydrolysis Microbial bioconversion Carbon source Nitrogen-rich source Polyhydroxyalkanoates PHB with a PHB content of approximately 70% (w/w). However, this processing scheme does not take advantage of the full potential of sunflower meal that contains value-added ingredients that could be isolated contributing in the development of a more sustainable biorefinery approach. Figure 24.3 presents a sunflower-based biorefinery where besides fermentation feedstock, sunflower meal is also used for the production of an antioxidant-rich stream and a protein isolate product. The sunflower seed is covered by the hull that could be removed before oil separation by mechanical pressing and solvent extraction in biodiesel production processes. The protein content in sunflower meals can be increased via dehulling and complete oil extraction. The composition of sunflower meal is variable and is highly dependent on cultivation conditions, sunflower variety and the industrial process used for biodiesel production. Dehulling or partial dehulling of sunflower seeds could provide a byproduct that could be used for the production of energy, hemicelluloses, organic amendment for the soils, and biomaterial (Anonymous, 2011). The sunflower meal that remains as a by-product after (partial) dehulling and complete oil extraction could be fractionated in three different fractions (i.e. a protein-rich fraction, a lignocellulose-rich fraction and a liquid fraction) by a simple sedimentation/flotation process based on the formation of an aqueous suspension (Bautista et al., 1990; Parrado et al., 1991). This separation is based on the different densities of major components in sunflower Sunflower seeds Energy generation biomaterial hemicellulose production Biodiesel Transesterification Oil Crude glycerol Partial dehulling Mechanical pressing and hexane extraction Protein isolate Carbon source Natural adhesive food/feed additive biopolymers Protein-rich fraction Antioxidants Microbial fermentation Polyhydroxyalkanoates FIGURE 24.3 Partly dehulled sunflower meal Aqueous extraction Lignocellulosic fraction Solid state fermentation Nutrient supplement Enzymatic hydrolysis Advanced sunflower-based biorefinery concept. Liquid fraction
PHA PRODUCTION INTEGRATED IN BIOREFINERY CONCEPTS meal. Subsequently, antioxidants can be removed from the protein-rich fraction, as well as from the lignocellulosic fraction. The most important of the phenolic compounds found in sunflower is chlorogenic acid. The protein isolate extracted from the protein-rich fraction, after treatment with acid and alkaline solutions, can be utilized for the production of biopolymers and edible films. Yust et al. (2003) improved protein extraction from sunflower meal through treatment with alkalase. Remaining fractions can be used as substrate in solid-state fermentation with a fungal strain of Aspergillus oryzae, an efficient producer mostly of proteolytic enzymes. The solids at the end of solid-state fermentation can be used as enzyme-rich medium for hydrolysis of macromolecules contained in remaining sunflower. The liquid fraction from sunflower meal fractionation can be used as suspension liquid in enzymatic hydrolysis, aiming at the generation of a nutrient-rich supplement. At the end of hydrolysis, remaining solids are separated from the hydrolysate by centrifugation, and can be possibly used for combustion to generate heat or as a carbohydrate-rich resource for the production of hydrolysates for other fermentations. The nutrientrich supplement has been used successfully for enhanced production of PHB. The advanced biorefinery concept results in the production of three products (antioxidants, protein isolate and PHB) from the same raw material presenting a high potential of improved process economics. Valorization of Second-Generation Bioethanol Industry By-Products Cellulose, hemicellulose and lignin are the main components found in lignocellulosic raw materials and the corresponding composition is dependent on the biomass resource. Production of sugar-rich hydrolysates from lignocellulosic biomass requires treatment with combined thermochemical treatment and enzymatic hydrolysis. Previous studies on utilization of lignocellulosic resources have focused on hydrolysis of cellulose and hemicellulose fractions to simple sugars for microbial fermentation mainly aiming to bioethanol production. Nonetheless, given the interest arising in biopolymers production, bioethanol production could be combined with PHA production. Cellulose could be utilized for the production of bioethanol, while sugars from hemicellulose could be utilized for the production of PHAs. In this way, conventional processes employed for the production of bioethanol from lignocellulosic biomass could be upgraded into advanced biorefinery concepts. Silva et al. (2004) screened 55 strains as potential PHB producers from xylose and identified Burkholderia sacchari IPT 101 and Burkholderia cepacia IPT 048 that were subsequently evaluated via cultivations on xylose and 427 bagasse hydrolysates. Intracellular PHB content reached 62% and 53% for the two strains, respectively, when grown on bagasse hydrolysates. Keenan et al. (2006) utilized detoxified hemicellulose hydrolysates from lignocellulosic resources for the production of P(3HBco-3HV) with B. cepacia through supplementation with levulinic acid (0.25e0.5%) to achieve a P(3HB-co-3HV) concentration of 2 g/l, a P(3HB-co-3HV) content of 40% (w/w) and 3HV composition of 16e52 mol%. When xylose and levulinic acid were used in microbial bioconversions with B. cepacia, the P(3HB-co-3HV) concentration and 3HV composition achieved were up to 4.2 g/l and 61 mol%, respectively. Sugarcane bagasse hydrolysates were evaluated for PHA synthesis via fermentation of R. eutropha (Yu and Stahl, 2008). The effect of inoculum concentration, dilution of hydrolysate and implementation of an adapted strain was studied regarding PHA accumulation, which reached up to 57% (w/w) polymer content. PHB was the major polymer accumulated, whereas copolymers could be also produced that presented high ductility. PHA production could be incorporated in existing bioethanol production facilities from both sugar cane in Brazil and cereals, such as wheat and corn, in other countries worldwide. Sugar cane utilization for bioethanol production generates significant quantities of bagasse, a lignocellulosic raw material that could be used for combined production of ethanol from cellulose and PHAs from hemicellulose sugars (mainly xylose). Integration of PHA production in existing cereal-based facilities used for bioethanol production could be achieved by incorporating straw utilization as raw material for combined production of bioethanol and PHAs. Such integrated biorefinery concepts could improve the sustainability of first-generation bioethanol production plants. Valorization of By-Product Streams from Food Industries The term “food waste” covers the wastes (and byproduct streams) that are generated during the whole food supply chain starting from production of the raw material followed by the processing into edible products by the food industry and the final disposal by consumers, restaurants or catering services. Valorizing the waste derived from the food industry sector would result in the creation of novel biorefineries leading to restructured and advanced industrial plants that will not only satisfy the traditional market of food production but also other markets that are nowadays dependent on petroleum to provide the necessary feedstocks. Food processing waste streams constitute renewable resources enriched in carbohydrates, protein, oils and fats, phenolic compounds and various micronutrients.
428 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION PHA Production from Winery By-Products utilization of this waste stream, since it contains lignocellulosic fractions that can be hydrolyzed and further used in microbial bioconversions. Solid-state fermentation for production of hydrolytic enzymes has also been reported using grape marc as solid support (Botella et al., 2005). Wine lees is the remaining residue after the end of the fermentation stage. It is a rich source of ethanol, tartaric acid, phenolic compounds and yeast cells. Wine lees can be used for the production of potable alcohol (wine lees mainly produced by large wineries), as nutrient supplement for fermentation (Bustos et al., 2004; Salgado et al., 2010), for the production of tartaric acid (Versari et al., 2001; Rivas et al., 2006) and as raw material for composting (Diaz et al., 2002; Nogales et al., 2005). A novel process has been developed at the Agricultural University of Athens targeting the creation of a novel biorefinery concept based on wine lees valorization (Figure 24.4). The process starts with centrifugation or filtration of wine lees in order to separate the liquid stream that can be used for ethanol production via distillation. The ethanol produced can be used as potable or fuel ethanol depending on the purity. Current processes produce potable ethanol. Ethanol could be also used as a platform chemical to supply the future sustainable chemical industry. Alternatively, ethanol could be also utilized as carbon source for microbial fermentation aiming to PHB production by the bacterial strain C. necator NCIMB 12080 (Senior et al., 1986). This, however, may not be a cost-competitive alternative when compared to the traditional potable ethanol market. The remaining liquid after ethanol extraction can be used in subsequent hydrolysis stages to increase the presence of nutrients. Wine production constitutes an important industrial sector in many countries around the world, such as the South European countries, United States, Chile and Australia. Wine making generates both solid and liquid by-products. Residues from wine production involve mainly trimming wastes, grape stalk, grape pomace or marc, wine lees and winery wastewater. These byproducts are currently supplied to ethanol distilleries (e.g. in the case of wine lees), used (if possible) as fertilizers or processed as wastes in order to reduce the environmental impact caused by their disposal to the environment. However, given the fact that environmental policies are changing, new practices should be applied aiming at valorization of winery by-product streams. Ongoing research focuses on valorization of residues from wine making. Trimming wastes are rich in cellulose, hemicellulose and lignin. Combined thermochemical treatment with enzymatic hydrolysis can be applied to convert cellulose and hemicelluloses into C5 and C6 sugars that can be assimilated by microorganisms. Delignification steps are usually required since the complex structure of lignin prevents hydrolysis of polysaccharide. Bustos et al. (2005) evaluated the use of trimming wastes and wine lees aiming at the production of lactic acid through simultaneous saccharification and fermentation carried out by Lactobacillus rhamnosus. Trimming wastes could be also used as solid support in solid-state fermentations for the production of various enzymes (Sanchez et al., 2002). Grape pomace or marc is the solid fraction remaining after the extraction and it consists of skins, pulp, seeds and stems of grapes. Research has focused on efficient Wine lees Liquid Distillation Ethanol Solids Residual solids 1 Alcohol-free nutrient rich liquid Potable alcohol Biofuel Platform chemical Treatment with HCl Residual solids2 (rich in yeast cells) Enzymatic hydrolysis of yeast cells Polyhydroxyalkanoates FIGURE 24.4 Antioxidants Nutrient-rich supplement for microbial fermentations Liquid rich in tartaric acid Tartrate salts Tartaric acid Crude enzymes produced via solid state fermentation Addition of carbon sources (e.g. crude glycerol, lignocellulosic hydrolysates) Valorization of wine lees.
PHA PRODUCTION INTEGRATED IN BIOREFINERY CONCEPTS The solid fraction that remains after centrifugation of wine lees contains phenolic compounds with antioxidant properties, tartrate salts and yeast cells. A phenolic-rich fraction can be easily isolated via solvent extraction. Tartrate salts can be subsequently separated from yeast cells via treatment with hydrochloric acid. Versari et al. (2001) extracted tartaric acid with purity up to 99% from three different winery by-product streams, including wine lees. Moreover, Nurgel and Canbas (1998) investigated the production of tartaric acid from grape pomace. Use of tartaric acid is well established in wine making in order to adjust the pH of the must prior to fermentation. Tartaric acid could be also used as food additive. After the extraction of phenolic compounds and tartrate salts, residual wine lees solids are subjected to enzymatic hydrolysis with the addition of crude enzymes produced via solid-state fermentation of a fungal strain of A. oryzae on wheat bran. The ethanol-free medium that remains after the distillation step is used as liquid in the hydrolysis stage. In this stage, yeast cells are lysed and converted into a nutrient-rich supplement similar to yeast extract. This supplement is rich in various sources of nitrogen (e.g. amino acids and peptides), phosphorus and various trace elements. This nutrient supplement can be combined with a carbon source (e.g. crude glycerol from biodiesel industries) as fermentation media for the production of PHB with C. necator. Preliminary experiments with C. necator DSM 7237 and crude glycerol as carbon source showed that PHB production is feasible using wine lees hydrolysates. However, supplementation with a low quantity of minerals is necessary showing that this nutrient supplement is deficient in some minerals. The wine lees hydrolysate could be combined with a sugar-rich hydrolysate derived from treatment of lignocellulosic streams derived during wine production. PHB Production from Confectionery and Bakery Industry Waste Streams Significant quantities of waste streams are generated annually from confectionery industries and bakeries. The waste streams from the industrial sectors mentioned above produce flour-, starch- or sugar-rich waste streams generated either during processing or as endof-date products returned from the market. Confectionery waste streams are currently used as animal feed, for composting or are discarded to landfills. However, these low-cost materials constitute renewable feedstocks that could be used for the development of novel biorefinery schemes. Anaerobic digestion from various food waste streams and biodiesel production from cooking oils are predominant alternatives that have been proposed for the utilization of various food waste streams. Current research on confectionery waste streams and waste bread is rather limited, but in recent years research has started to focus on the valorization of such waste streams. Dorado et al. (2009) utilized hydrolysates derived from wheat milling by-products as fermentation media for the production of succinic acid (50.6 g/l). Leung et al. (2012) developed a two-stage bioprocess involving solid-state fermentation and enzymatic hydrolysis of waste bread to produce a fermentation feedstock for the production of succinic acid (47.3 g/l at a conversion yield of 0.55 g SA/g bread) using the bacterial strain Actinobacillus succinogenes. A potential biorefining concept for the production of PHAs and biodiesel from confectionery industry waste streams is presented in Figure 24.5. In the case of confectionery wastes that contain high oil content, this could be removed via solvent extraction. The oil obtained from this step can be used for biodiesel production. Remaining fractions will be rich in directly assimilable sugars such as glucose, fructose, sucrose and lactose as well as starch and protein. Utilizing starch- and Biodiesel production Oil extraction (if required) Flour-and starchbased waste streams Enzymatic hydrolysis Crude enzyme production Wheat milling by-products Solid state fermentation 429 Fermentation medium Microbial fermentation Polyhydroxyalkanoates Additional wastes Aspergillus awamori FIGURE 24.5 Valorization of confectionery industry waste streams.
430 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION protein-rich waste streams as sources of carbon and nitrogen in fermentation processes demands the conversion of starch into glucose and protein into amino acids and peptides. The amylolytic and proteolytic enzymes required for the hydrolysis of these macromolecules could be produced via solid-state fermentation using the fungal strain Aspergillus awamori cultivated on wheat milling by-products. The fermented solids, rich in amylolytic and proteolytic enzymes, are subsequently combined with confectionery waste to produce hydrolysates that can be used in fermentation processes for the production of platform chemicals, microbial oil or PHB. The production of PHB or PHAs from confectionery industry wastes could be employed for the production of biodegradable packaging materials for the same industry. The proposed process is based on the results that were achieved for the production of PHB using wheat as the whole raw materials (Koutinas et al., 2007a, 2007b; Xu et al., 2010). In this biorefinery concept, wheat is fractionated into bran and gluten as valueadded co-products, while remaining fractions are used for the production of fermentation media suitable for the production of PHB via fed-batch cultures using the microbial strain Wautersia eutropha NCIMB 11599. Xu et al. (2010) developed a fermentation process for the production of PHB from wheat-derived fermentation media during fed-batch cultures in a bioreactor. The highest PHB concentration achieved was 162.8 g/l. However, wheat is regarded a food resource and should not be used for chemical production. Starch- or flourrich food wastes could be used, instead of wheat, as a renewable resource for the production of PHB. PHB Production from Whey Whey is the main by-product occurring from cheese manufacture and lactose is one of the primary components. Current whey valorization processes mainly focus on the production of whey powder, whey protein concentrate or whey protein isolate. Utilization of whey in fermentation processes has been widely investigated, given the fact that it is produced in many countries in significant quantities. Furthermore, whey valorization will also contribute to the improvement of the environmental impact of the cheese industry because whey disposal is a notorious environmental burden. Future cheese industries could incorporate integrated processing schemes for the production of whey protein and PHAs. Koller et al. (2010) reviewed various bioconversions that employed whey permeate as carbon source aiming at the production of PHAs. Different strategies were proposed concerning uses of whey permeate; direct conversion as substrate or hydrolysis of lactose to glucose and galactose were examined. Moreover, Wong and Lee (1998) presented PHB production from whey powder with recombinant E. coli in pH-stat cultures. In fed-batch cultures with additions of concentrated whey solution, the corresponding dry cell weight and PHB concentrations were 87 and 69 g/l, respectively. The PHB content reached up to 80% (w/ w). These results established that PHB fermentation process from whey could be industrially employed, increasing the sustainability and market alternatives of traditional cheese producing plants. Whey protein concentrate and isolate that could be extracted from whey by ultrafiltration and evaporation steps can be applied as food additives. Moreover, they are considered to possess therapeutic properties and for this reason, whey protein concentrate was applied for treatment of various clinical disorders. Furthermore, whey protein ingredients are added to food targeting to improve their functional or technological properties. Hence, keeping that in mind, biorefinery schemes based on whey utilization could be easily proposed. CONCLUSIONS AND FUTURE PERSPECTIVES The necessity to eliminate our dependence on fossil resources will lead to an inevitable reconstruction of the current industry in order to introduce the utilization of renewable resources and produce chemicals, fuels and materials in a sustainable manner. Implementation of biorefinery concepts into existing industrial facilities provide an alternative processing option, taking into consideration that industrial by-products and waste streams are generated in significant quantities and currently, they are inadequately utilized. Consequently, production of value-added products from waste and by-product streams will enhance sustainability and diversify market opportunities. Furthermore, production of biofuels should coincide with chemical and biodegradable polymer production to enhance their sustainability. This study showed potential industries where biofuel and food production could coincide with PHA production. This research area is currently at the inception phase and significant effort is required in order to develop the technologies that will be implemented on industrial scale. References Akaraonye, E., Keshavarz, T., Roy, I., 2010. Production of polyhydroxyalkanoates: the future green materials of choice. J. Chem. Technol. Biotechnol. 85, 732e743. Anonymous, 2011. WP2: Optimisation of Primary Processing. Report of Deliverables. Project Title: Developing advanced Biorefinery schemes for integration into existing oil production/transesterification plants. Project Number: 213637 http://www.york.ac. uk/res/sustoil/Pages/Deliverable%202-5.pdf.
REFERENCES Anonymous, 2012c. OECD-FAO Agricultural Outlook 2012e2021 (Chapter 5): Oilseeds and Oilseed products http://www.fao.org/ fileadmin/templates/est/COMM_MARKETS_MONITORING/ Oilcrops/Documents/OECD_Reports/Ch5StatAnnex.pdf. Anonymous, 2012a. Wastes-Resource Conservation- Common Wastes & Materials. U.S. Environmental Protection Agency. http://www. epa.gov/osw/conserve/materials/plastics.htm. Anonymous, 2012b. Renewable Resources for the Production of Bioplastics: Impact on Agriculture - Status and Outlook. Fact sheet. European Bioplastics http://en.european-bioplastics.org/wpcontent/uploads/2011/04/fs/Renewable_resources_eng.pdf. Ashby, R.D., Solaiman, D.K.Y., Strahan, G.D., 2011. Efficient utilization of crude glycerol as fermentation substrate in the synthesis of poly(3-hydroxybutyrate) biopolymers. J. Am. Oil Chem. Soc. 88, 949e959. Ashby, R.D., Solaiman, D.K.Y., Foglia, T.A., 2004. Bacterial poly(hydroxyalkanoate) polymer production from the biodiesel co-product stream. J. Polym. Environ. 12, 105e112. Barham, P.J., Barker, P., Organ, S.J., 1992. Physical properties of poly(hydroxybutyrate) and copolymers of hydroxybutyrate and hydroxyvalerate. FEMS Microbiol. Lett. 103, 289e298. Bautista, J., Parrado, J., Machado, A., 1990. Composition and fractionation of sunflower meal: use of the lignocellulosic fraction as substrate in solid state fermentation. Biol. Wastes 32, 225e233. Bhubalana, K., Rathia, D.N., Abeb, H., Iwatac, T., Sudesh, K., 2010. Improved synthesis of P(3HB-co-3HV-co-3HHx) terpolymers by mutant Cupriavidus necator using the PHA synthase gene of Chromobacterium sp. USM2 with high affinity towards 3HV. Polym. Degrad. Stab. 95, 1436e1442. Botella, C., de Ory, I., Webb, C., Cantero, D., Blandino, A., 2005. Hydrolytic enzyme production by Aspergillus awamori on grape pomace. Biochem. Eng. J. 26, 100e106. Bustos, G., Moldes, A.B., Cruz, J.M., Domı́nguez, J.M., 2004. Evaluation of vinification lees as a general medium for Lactobacillus strains. J. Agric. Food Chem. 52, 5233e5239. Bustos, G., Moldes, A.B., Cruz, J.M., Domı́nguez, J.M., 2005. Production of lactic acid from vine-trimming wastes and viticulture lees using a simultaneous saccharification fermentation method. J. Sci. Food Agric. 85, 466e472. Byrom, D., 1987. Polymer synthesis by microorganisms: technology and economics. Trends Biotechnol. 5, 246e250. Cavalheiro, J.M.B.T., de Almeida, M.C.M.D., Grandfils, C., da Fonseca, M.M.R., 2009. Poly(3-hydroxybutyrate) production by Cupriavidus necator using waste glycerol. Process. Biochem. 44, 509e515. Cavalheiro, J.M.B.T., Raposo, R.S., de Almeida, M.C.M.D., Cesário, M.T., Sevrin, C., Grandfils, C., da Fonseca, M.M.R., 2012. Effect of cultivation parameters on the production of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) and poly(3-hydroxybutyrate-4-hydroxybutyrate-3-hydroxyvalerate) by Cupriavidus necator using waste glycerol. Bioresour. Technol. 111, 391e397. Chatzifragkou, A., Papanikolaou, S., 2012. Effect of impurities in biodiesel-derived waste glycerol on the performance and feasibility of biotechnological processes. Appl. Microbiol. Biotechnol. 95, 13e27. Choi, J., Lee, S.Y., 1997. Process analysis and economic evaluation for poly(3-hydroxyutyrate) production by fermentation. Bioprocess Eng. 17, 335e342. da Silva, G.P., Mack, M., Contiero, J., 2009. Glycerol: a promising and abundant carbon source for industrial microbiology. Biotechnol. Adv. 27, 30e39. Diaz, M.J., Madejón, E., López, F., López, R., Cabrera, F., 2002. Optimization of the rate vinasse/grape marc for co-composting process. Process. Biochem. 37, 1143e1145. 431 Dorado, M.P., Lin, S.K.C., Koutinas, A., Du, C., Wang, R., Webb, C., 2009. Cereal-based biorefinery development: utilisation of wheat milling by-products for the production of succinic acid. J. Biotechnol. 143, 51e59. Du, C., Sabirova, J., Soetaert, W., Lin, C., 2012. Polyhydroxyalkanoates production from low-cost sustainable raw materials. Curr. Chem. Biol. 6, 14e25. Egues, I., Gonzalez Alriols, M., Herseczki, Z., Marton, G., Labidi, J., 2010. Hemicelluloses obtaining from rapeseed cake residue generated in the biodiesel production process. J. Ind. Eng. Chem. (Seoul, Repub. Korea) 16, 293e298. Garcia, I.L., Dorado, M.P., Kopsahelis, N., Alexandri, M., Papanikolaou, S., Villar, M.A., Koutinas, A.A., 2013. Evaluation of by-products from the biodiesel industry as fermentation feedstock for poly(3-hydroxybutyrate-co-3-hydroxyvalerate) production by Cupriavidus necator. Bioresour. Technol. 130, 16e22. Gouda, M.K., Swellam, A.E., Omar, S.H., 2001. Production of PHB by a Bacillus megaterium strain using sugarcane molasses and corn steep liquor as sole carbon and nitrogen sources. Microbiol. Res. 156, 201e207. Gunaratne, L.M.W.K., Shanks, R.A., 2005. Multiple melting behaviour of poly(3-hydroxybutyrate-co-hydroxyvalerate) using step-scan DSC. Eur. Polym. J. 41, 2980e2988. Haas, R., Jin, B., Zepf, F.T., 2008. Production of poly(3-hydroxybutyrate) from waste potato starch. Biosci. Biotechnol. Biochem. 72, 253e256. Han, J., Li, M., Hou, J., Wu, L., Zhou, J., Xiang, H., 2010. Comparison of four phaC genes from Haloferax mediterranei and their function in different PHBV copolymer biosyntheses in Haloarcula hispanica. Saline Syst. 6, 1e10. He, J.-D., Cheung, M.K., Yu, P.H., Chen, G.-Q., 2001. Thermal analyses of poly(3-hydroxybutyrate), poly(3-hydroxybutyrate-co3-hydroxyvalerate), and poly(3-hydroxybutyrate-co3-hydroxyhexanoate). J. Appl. Polym. Sci. 82, 90e98. Hejazi, P., Vasheghani-Farahani, E., Yamini, Y., 2003. Supercritical fluid disruption of Ralstonia eutropha for poly(beta-hydroxybutyrate) recovery. Biotechnol. Prog. 19, 1519e1523. Huang, T.Y., Duan, K.J., Huang, S.Y., Chen, C.W., 2006. Production of polyhydroxyalkanoates from inexpensive extruded rice bran and starch by Haloferax mediterranei. J. Ind. Microbiol. Biotechnol. 33, 701e706. Kachrimanidou, V., Kopsahelis, N., Chatzifragkou, A., Papanikolaou, S., Yanniotis, S., Kookos, I., Koutinas, A.A., 2013. Utilisation of by-products from sunflower-based biodiesel production processes for the production of fermentation feedstock. Waste Biomass Valor 4, 529e537. Kahar, P., Tsuge, T., Taguchi, K., Doi, Y., 2004. High yield production of polyhydroxyalkanoates from soybean oil by Ralstonia eutropha and its recombinant strain. Polym. Degrad. Stab. 83, 79e86. Kapritchkoff, F.M., Viotti, A.D., Alli, R.C.P., Zuccolo, M., Pradella, J.G.C., Maiorano, A.E., Miranda, E.A., Bonomi, A., 2006. Enzymatic recovery of polyhydroxybutyrate produced by Ralstonia eutropha. J. Biotechnol. 122, 453e462. Keenan, T.M., Nakas, J.P., Tanenbaum, S.W., 2006. Polyhydroxyalkanoate copolymers from forest biomass. J. Ind. Microbiol. Biotechnol. 33, 616e626. Koller, M., Atlic, A., Dias, M., Reiterer, A., Braunegg, G., 2010. Microbial PHA production from waste raw materials. In: Chen, G.Q. (Ed.), Plastics from Bacteria: Natural Functions and Applications, vol. 14. Springer-Verlag, Berlin, pp. 85e119. Koller, M., Hesse, P., Bona, R., Kutschera, C., Atlic, A., Braunegg, G., 2007b. Potential of various Archae- and Eubacterial strains as industrial polyhydroxyalkanoate producers from whey. Macromol. Biosci. 7, 218e226. Koller, M., Hesse, P., Bona, R., Kutschera, C., Atlic, A., Braunegg, G., 2007a. Biosynthesis of high quality polyhydroxyalkanoate co- and
432 24. BIOENERGY TECHNOLOGY AND FOOD INDUSTRY WASTE VALORIZATION FOR INTEGRATED PRODUCTION terpolyesters for potential medical application by the archaeon Haloferax mediterranei. Macromol. Symp. 253, 33e39. Koller, M., Bona, R., Chiellini, E., Grillo Fernandes, E., Horvat, P., Kutschera, C., Hesse, P., Braunegg, G., 2008. Polyhydroxyalkanoate production from whey by Pseudomonas hydrogenovora. Bioresour. Technol. 99, 4854e4863. Koutinas, A.A., Papanikolaou, S., 2011. Biodiesel production from microbial oil. In: Luque, R., Campelo, J., Clark, J.H. (Eds.), Handbook of Biofuels Production - Processes and Technologies. Woodhead Publishing Limited, pp. 177e198. Koutinas, A.A., Wang, R.H., Webb, C., 2007a. The biochemurgist bioconversion of agricultural raw materials for chemical production. Biofuels, Bioprod. Biorefin. 1, 24e38. Koutinas, A.A., Xu, Y., Wang, R., Webb, C., 2007b. Polyhydroxybutyrate production from a novel feedstock derived from a wheat-based biorefinery. Enzyme Microb. Technol. 40, 1035e1044. Kulpreecha, S., Boonruangthavorn, A., Meksiriporn, B., Thongchu, N., 2009. Inexpensive fed-batch cultivation for high poly(3hydroxybutyrate) production by a new isolate of Bacillus megaterium. J. Biosci. Bioeng. 107, 240e245. Kusaka, S., Iwata, T., Doi, Y., 1999. Properties and biodegradability of ultra-high-molecular-weight poly[(R)-3-hydroxybutyrate] produced by a recombinant Escherichia coli. Int. J. Biol. Macromol. 25, 87e94. Lee, W.H., Loo, C.Y., Nomura, C.T., Sudesh, K., 2008. Biosynthesis of polyhydroxyalkanoate copolymers from mixtures of plant oils and 3-hydroxyvalerate precursors. Bioresour. Technol. 99, 6844e6851. Lemoigne, M., 1926. Produits de déshydration et de polymerization de l’acide b-oxybutyrique. Bull. Soc. Chim. Biol. 8, 770e782. Leung, C.C.J., Cheung, A.S.Y., Zhang, A.Y.Z., Lam, K.F., Lin, C.S.K., 2012. Utilisation of waste bread for fermentative succinic acid production. Biochem. Eng. J. 65, 10e15. Loo, C.Y., Lee, W.H., Tsuge, T., Doi, Y., Sudesh, K., 2005. Biosynthesis and characterization of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) from palm oil products in a Wautersia eutropha mutant. Biotechnol. Lett. 27, 1405e1410. Luo, L., Wei, X., Chen, G.Q., 2009. Physical properties and biocompatibility of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) blended with poly(3-hydroxybutyrate-co-4-hydroxybutyrate). J. Biomater. Sci. 20, 1537e1553. Madden, L.A., Anderson, A.J., Asrar, J., Berger, P., Garrett, P., 2000. Production and characterization of poly(3-hydroxybutyrate-co3-hydroxyvalerate-co-4-hydroxybutyrate) synthesized by Ralstonia eutropha in fed-batch cultures. Polymer 41, 3499e3505. Marsudi, S., Tan, I.K.P., Gan, S.N., Ramachandran, K.B., 2007. Production of medium chain length polyhydroxyalkanoates from oleic acid using Pseudomonas putida PGA1 by fed batch culture. Makara, Teknologi 11, 1e4. Meng, X., Yang, J., Xu, X., Zhang, L., Nie, Q., Xian, M., 2009. Biodiesel production from oleaginous microorganisms. Renewable Energy 34, 1e5. Mothes, G., Schnorpfeil, C., Ackermann, J.-U., 2007. Production of PHB from crude glycerol. Eng. Life Sci. 7, 475e479. Nath, A., Dixit, M., Bandiya, A., Chavda, S., Desa, A.J., 2008. Enhanced PHB production and scale up studies using cheese whey in fed batch culture of Methylobacterium sp. ZP24. Bioresour. Technol. 99, 5749e5755. Nogales, R., Cifuentes, C., Benı́tez, E., 2005. Vermicomposting of winery wastes: a laboratory study. J. Environ. Sci. Health Part B 40, 659e673. Nurgel, C., Canbas, A., 1998. Production of tartaric acid from pomace of some anatolian grape cultivars. Am. J. Enol. Vitic. 49, 95e99. Parrado, J., Bautista, J., Machado, A., 1991. Production of soluble enzymatic protein hydrolysate from industrially defatted nondehulled sunflower meal. J. Agric. Food Chem. 39, 447e450. Posada, J.A., Naranjo, J.M., López, J.A., Higuita, J.C., Cardona, C.A., 2011. Design and analysis of poly-3-hydroxybutyrate production processes from crude glycerol. Process. Biochem. 46, 310e331. Rivas, B., Torrado, A., Moldes, A.B., Domı́nguez, J.M., 2006. Tartaric acid recovery from distilled lees and use of the residual solid as an economic nutrient for Lactobacillus. J. Agric. Food Chem. 54, 7904e7911. Ryu, H.W., Hahn, S.K., Chang, Y.K., Chang, H.N., 1997. Production of poly(3-hydroxybutyrate) by high cell density fed-batch culture of Alcaligenes eutrophus with phosphate limitation. Biotechnol. Bioeng. 55, 28e32. Salgado, J.M., Rodriguez, N., Cortés, S., Domı́nguez, J.M., 2010. Improving downstream processes to recover tartaric acid, tartrate and nutrients from vinasses and formulation of inexpensive fermentative broths for xylitol production. J. Sci. Food Agric. 90, 2168e2177. Sanchez, A., Ysunza, F., Beltran-Garcia, M., Esqueda, M., 2002. Biodegradation of viticulture wastes by Pleurotus: a source of microbial and human food and its potential use in animal feeding. J. Agric. Food Chem. 50, 2537e2542. Sarris, D., Galiotou-Panayotou, M., Koutinas, A., Komaitis, M., Papanikolaou, S., 2011. Citric acid, biomass and cellular lipid production by Yarrowia lipolytica strains cultivated on olive mill wastewater-based media. J. Chem. Technol. Biotechnol. 86, 1439e1448. Senior, P.J., Collins, S.H., Richardson, K.R., 1986. Copolymers of Poly(b-Hydroxybutyric Acid) and Poly(b -hydroxyvaleric Acid) Are Produced by Culturing Alcohol-utlising Strains of Alcaligenes eutrophus on a Carbon Source Including Primary Alcohols Having an Odd Number of Carbon Atoms Such as Propan-1-ol. European Patent Office. Publication Number 0204442 A2. Shimamura, E., Kasuya., K., Kobayashi, G., Shiotani, T., Shima, Y., Doi, Y., 1994. Physical properties and biodegradability of microbial poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). Macromolecules 27, 878e880. Silva, L.F., Taciro, M.K., Ramos, M.E.M., Carter, J.M., Pradella, J.G.C., Gomez, J.G.C., 2004. Poly-3-hydroxybutyrate (P3HB) production by bacteria from xylose, glucose and sugarcane bagasse hydrolysate. J. Ind. Microbiol. Biotechnol. 31, 245e254. Solaiman, D.K.Y., Ashby, R.D., Foglia, T.A., Marmer, W.N., 2006. Conversion of agricultural feedstock and coproducts into poly(hydroxyalkanoates). Appl. Microbiol. Biotechnol. 71, 783e789. Tanadchangsaeng, N., Yu, J., 2012. Microbial synthesis of polyhydroxybutyrate from glycerol: gluconeogenesis, molecular weight and material properties of biopolyester. Biotechnol. Bioeng. 109, 2808e2818. Tsuge, T., Saito, Y., Kikkawa, Y., Hiraishi, T., Doi, Y., 2004. Biosynthesis and compositional regulation of poly[(3-hydroxybutyrate)-co(3-hydroxyhexanoate)] in recombinant Ralstonia eutropha expressing mutated polyhydroxyalkanoate synthase genes. Macromol. Biosci. 4, 238e242. Van Wegen, R.J., Ling, Y., Middelberg, A.P.J., 1998. Industrial production of polyhydroxyalkanoates using Escherichia coli: an economic analysis. Trans. IChemE 76, 417e426. Verlinden, R.A.J., Hill, D.J., Kenward, M.A., Williams, C.D., Radecka, I., 2007. Bacterial synthesis of biodegradable polyhydroxyalkanoates. J. Appl. Microbiol. 102, 1437e1449. Versari, A., Castellari, M., Spinabelli, U., Galassi, S., 2001. Recovery of tartaric acid from industrial enological wastes. J. Chem. Technol. Biotechnol. 76, 485e488. Wang, R., Shaarani, S.M., Godoy, L.C., Melikoglu, M., Vergara, C.S., Koutinas, A., Webb, C., 2010. Bioconversion of rapeseed meal for the production of a generic microbial feedstock. Enzyme Microb. Technol. 47, 77e83. Whitehouse, R.S., Zhong, L., Daughtry, S., 2006. Compositions Comprising Low Molecular Weight Polyhydroxyalkanoates and
REFERENCES Methods Employing Same. United States Patent No. US 7,094,840 B2. Wolf, O., Crank, M., Patel, M., Marscheider-Weidemann, F., Schleich, J., Husing, B., Angerer, G., 2005. Techno-economic Feasibility of Large Scale Production of Bio-based Polymers in Europe. Technical Report EUR 22103 EN. Wong, H.H., Lee, S.Y., 1998. Poly-(3-hydroxybutyrate) production from whey by high-density cultivation of recombinant Escherichia coli. Appl. Microbiol. Biotechnol. 50, 30e33. Xu, Y., Wang, R.-H., Koutinas, A.A., Webb, C., 2010. Microbial biodegradable plastic production from a wheat-based biorefining strategy. Process. Biochem. 45, 153e163. 433 Yu, H., Shi, Y., Yin, J., Shen, Z., Yang, S., 2003. Genetic strategy for solving chemical engineering problems in biochemical engineering. J. Chem. Technol. Biotechnol. 78, 283e286. Yu, J., Stahl, H., 2008. Microbial utilization and biopolyester synthesis of bagasse hydrolysates. Bioresour. Technol. 99, 8042e8048. Yu, J., Chen, L.X.L., 2006. Cost-effective recovery and purification of polyhydroxyalkanoates by selective dissolution of cell mass. Biotechnol. Prog. 22, 547e553. Yust, M.M., Pedroche, J., Megı́as, C., Girón-Calle, J., Alaiz, M., Millán, F., Vioque, J., 2003. Improvement of protein extraction from sunflower meal by hydrolysis with alcalase. Grasas Aceites 54, 419e423.
C H A P T E R 25 Advances and Innovations in Biochar Production and Utilization for Improving Environmental Quality Charles Hyland*, Ajit K. Sarmah Department of Civil & Environmental Engineering, The University of Auckland, Auckland, New Zealand *Corresponding author email: chyl531@aucklanduni.ac.nz O U T L I N E Introduction 435 Properties of Biochar Chemical and Physical Microbiological Effects and Synergisms 436 436 437 Utilization of Biochar for Environmental Quality Carbon Sequestration Sorption of Plant Nutrients and Other Pollutants Soil Greenhouse Gas Emissions Soil-Specific Biochar Design 438 438 438 439 440 INTRODUCTION In its simplest material context, biochar is incompletely combusted organic matter that is applied to soil; however, this material has attracted considerable attention in recent years due to goals that may be achievable through its production and end use. In this chapter, advances and innovations in biochar production and intermediate uses are presented and discussed. A considerable proportion of the world’s natural soil organic carbon content comprises black carbon, a pool that is resistant to microbial degradation and was deposited from historic fires (Krull et al., 2008). Augmenting this nearly ubiquitous pool with additional recalcitrant carbon has recently been the subject of much scientific research focused on soil improvement and carbon (C) sequestration, which has garnered notable support from some of the world’s most well-known climate Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00025-5 Postpyrolysis Indirect Application of Biochar Water Filtration Soil Nutrient Reclamation Biochar As Container Growth Medium and Container 440 440 441 441 Conclusions, Knowledge Gaps, and Research Needs 443 References 444 scientists and environmental advocates, such as Al Gore (Gore, 2009), Tim Flannery (Flannery, 2009), James Hansen (Hansen et al., 2008), and James Lovelock (Lovelock, 2009). Biochar is a form of anthropogenic recalcitrant carbon produced for the purpose of application to soil that draws inspiration from the practices of precolonial Native Amazonians who transformed some of the world’s poorest soils into extremely fertile soils that remain productive into the present, centuries after their production ceased (Lehmann, 2008). This rediscovered concept represents both a stable form of C-rich soil organic matter and a means to ameliorate degraded soils into fertile soils (Lehmann et al., 2006). C sequestration is achievable through a range of thermochemical conversion processes (Table 25.1), including pyrolysis, a process that produces biochar as a byproduct. Recalcitrant carbon, including particulates emitted by fossil fuel combustion, may represent up to 435 Copyright Ó 2014 Elsevier B.V. All rights reserved.
436 25. ADVANCES AND INNOVATIONS IN BIOCHAR PRODUCTION AND UTILIZATION FOR IMPROVING ENVIRONMENTAL QUALITY TABLE 25.1 Biomass Pyrolytic Conversion Processes Process Temperature ( C) Heating Rate End Products Torrefaction 230e300 Very low Biochar Slow Pyrolysis 380e530 Low Biochar, bio-oils, gas Fast Pyrolysis 380e530 High Bio-oils, relatively little biochar Combustion 700e1400 Very high Gas Gasification >750 Very high Gas 30% of all soil C globally (Schmidt et al., 2001). Woolf et al. (2010) reported that by increasing biochar production net anthropogenic greenhouse gas (GHG) emissions could be reduced by as much as 12% through the substitution of pyrolysis oils and gases for fossil fuels. Recent reviews and studies have highlighted the potential environmental as well as agronomic benefits of biochar (Kookana et al., 2011). While many of these potential environmental benefits of biochar are largely known through glasshouse and field trials, there have been findings that also showed negative impacts of biochar utilization as soil amendment (Clough et al., 2010). Given that biochar can be produced from a variety of feedstocks, and quality of biochar is dependent on the type of feedstock and pyrolysis conditions, not all biochars are made equal. The pyrolysis bioenergy industry is currently in the early stages of commercialization (Downie and van Zwieten, 2013) and is growing slowly due to the vast heterogeneity of biochar feedstocks, production methods, and end uses (Jahirul et al., 2012). The economic viability of commercial pyrolysis is likely to increase in tandem with the price of C credits assigned to biochar (Laird et al., 2009). Due to the characteristic persistence of biochar after incorporation into soil, prerequisite testing of biochars is crucial in order to avoid production of biochars that either form toxic compounds, such as polycyclic aromatic hydrocarbons formed as a result of pyrolysis (Verheijen et al., 2010; Kloss et al., 2012), or retain toxins, such as heavy metals, that were present in a feedstock, such as municipal solid waste or sewage sludge (Sparkes and Stoutjesdijk, 2011; Kookana et al., 2011). Minimal predictive capability currently exists in relation to biochar performance in the field (Sohi et al., 2010). Field trials, preceded by initial laboratory and Glasshouse testing, are the preferred biochar efficacy evaluation methods for conventional biochar, as well as novel biochar uses that are currently being investigated. Thorough research must be conducted in order to ensure that biochars suited to address specific issues are selected from the diverse array of biochars that are currently being produced, in order to avoid potential socioeconomic and environmental damage (Barrow, 2012). Rather than applying arbitrarily chosen biochars and passively observing their effects on the environment, a forefront of biochar research is the custom design of biochar for specific targeted end uses. Biochar is commonly referred to as a soil conditioner (Lehmann et al., 2006) in order to contextually differentiate the material from direct fertilizers. Given this, it rationally follows that biochars and their deployment methods be engineered to enhance the soil and the surrounding environment in a contextually desirable manner. For example, if it is known that biochar efficacy is increased once the biochar surfaces are coated with clay minerals in a given soil after a period of time, then producing biochars that already possess this quality at the time of incorporation into soil would be a significant advancement. Evaluating intermediate uses for biochar, that is functions that biochar can perform after production and before soil incorporation, may simultaneously improve the efficacy of the biochar and reduce preexistent downstream and atmospheric pollution. An example of an intermediate use discussed in this chapter is the employment of biochar contained within permeable mesh bags for the filtration of runoff and leachate water. PROPERTIES OF BIOCHAR As mentioned earlier, two important factors in biochar production are the type of feedstock used and pyrolysis conditions as they affect the physical and chemical properties of the biochar that is produced. Depending on the origin of the feedstock (e.g. cellulose, lignin, lingocellulose, hemicellulose) the chemical and structural composition of pyrolyzed biochar can change, and thus when biochar is used as a soil amendment, its behavior, function and fate in soils could be different. For instance, Winsley (2007) showed that when woodbased feedstocks are pyrolyzed, coarse and resistant biochars were generated with nearly 80% carbon contents, because the rigid ligninolytic nature of the source material is still retained in the biochar residue. The following sections discuss the chemical, physical, and biological properties of soil amended with biochar with particular focus on how the various properties influence the soilbiochar interactions and consequently improve the soil’s biological health and crop productivity. Chemical and Physical In an effort to offset anthropogenic C emissions to the atmosphere, a number of geoengineering technologies have been proposed, which can be divided into the
437 PROPERTIES OF BIOCHAR Respiration Respiration CO2 CO2 Photosynthesis Photosynthesis 5% 25% 50% 50% 50% Pyrolysis 25% Photosynthetic carbon cycle Photosynthetic + biochar carbon cycle FIGURE 25.1 Carbon sequestration through pyrolysis of plant biomass. Source: Adapted from Lehmann, 2007. two broad categories of solar radiation management and CO2 removal, with C sequestration through biochar incorporation into soil in the latter (Vaughan and Lenton, 2011). What distinguishes biochar CO2 sequestration (Figure 25.1) from competing CO2 removal technologies, such as ocean iron fertilization and CO2 geological injection, are two factors. First, biochar C sequestration relies on photosynthesizing plants to draw CO2 from the atmosphere and is stored in the form of charcoal, whereas methods such as CO2 geological injection rely on relatively new and untested storage methods, including mechanically forcing supercritical CO2 into depleted fossil fuel reservoirs or deep sea sediments (Vaughan and Lenton, 2011). Second, charcoal storage in soils is already ubiquitous and the clearest challenges in the short term involve optimizing economic profitability, rather than projected efficacy (Spokas et al., 2012a). C sequestration through pyrolysis for biochar production is a result of the conversion of C forms present in the feedstock into recalcitrant forms present in biochar. Pyrolysis consists of a succession of changes in chemical structural composition as temperature increases. Cellulose and lignin are degraded as volatile compounds are driven off between 250  C and 350  C, followed by lateral growth and coalescence of polyaromatic graphene sheets, and culminating with carbonization, constituted by the expulsion of the majority of non-C atoms above 600  C (Verheijen et al., 2010). Additionally, as pyrolysis temperature and residence time increase, H/C and O/C ratios of biochar decrease and aromaticity, the extent to which aromatic rings are connected, increases, resulting in greater recalcitrance against degradation (Kookana et al., 2011). Recent research suggests that biochars intended for both soil improvement and C sequestration should possess recalcitrant carbon of >15%, O/C ratios of <0.4, H/C ratios of <0.6, polyaromatic hydrocarbon contents below background soil values, and a surface area of >100 m2 g (Schimmelpfennig and Glaser, 2012). As pyrolysis temperatures increase, biochar specific surface area and microporosity analogously increase. Biochar chemical and physical properties are due to both the composition of the feedstock and the extent of the alterations undergone during pyrolysis (Kookana et al., 2011). The diversity of biochar chemical and physical qualities achievable through varied pyrolysis conditions and feedstocks is reflected in the diversity of effects on soil biota that may be achievable. Biochar is sterile when produced, yet has been observed to have beneficial effects on soil microbes that play essential roles in nutrient cycling. These effects are complex; however, they are linked to the chemical and physical properties of the biochar employed, which can serve as both a source of nutrients and as a habitat for soil microbes (Lehmann and Rondon, 2006; Lehmann et al., 2011). Microbiological Effects and Synergisms Literature focused on the effects of biochar on soil biota is currently sparse relative to biochar chemical
438 25. ADVANCES AND INNOVATIONS IN BIOCHAR PRODUCTION AND UTILIZATION FOR IMPROVING ENVIRONMENTAL QUALITY and physical effects on soil (Lehmann et al., 2011). Published research fails to adequately address the diverse spectrum of biochars within a single study, typically evaluating the effects of a small number of biochars on microbial and root abundance for a small number of soil types. Depending on the feedstock from which biochar is produced and the pyrolysis conditions involved in the feedstock conversion, biotoxic substances may persist through or be generated during pyrolysis. Standardized methods, including germination rates of various crops and degree of earthworm avoidance, have recently been proposed in order to assess biochar toxicity to avoid inadvertent detrimental effects on crop production and the environment (Busch et al., 2012; Rogovska et al., 2012). The addition of biochar to compost systems and the resultant effects on both the microbial community dynamics and final compost quality have been evaluated, and the potential for synergistic effects and increased soil C stability are known (Fischer and Glaser, 2012). While the aromatic core of biochar has been observed to remain unchanged during the composting process with manure, C and plant-available nutrients are drawn into its pores and adhere to its surfaces, elevating the cation exchange capacity (CEC) and acid neutralizing capacity, and enhancing the functionalization of biochar surfaces, although the mechanisms responsible warrant additional research (Prost et al., 2013). Biochar has been observed to affect mycorrhizal symbioses, yet the mechanisms responsible are still being determined (Warnock et al., 2007) and recently, it has been proposed that soil nitrogen may act as a switch controlling the proliferation of mycorrhizae and the subsequent oxidation of fresh biochar surfaces (LeCroy et al., 2013). Compared to compost, biochar is a much more stable soil amendment and the addition of biochar to composts has been shown to dramatically increase compost stability (Bolan et al., 2012). UTILIZATION OF BIOCHAR FOR ENVIRONMENTAL QUALITY Carbon Sequestration The potential of biochar to sequester atmospheric C for centuries is certainly one of its most attractive qualities. As global anthropogenic C emissions continue to increase, C sequestration using biochar employs photosynthesis to draw C from the atmosphere, and pyrolysis to convert that photosynthetically sequestered C into forms that are mostly not biologically degradable. Fossil fuel-derived energy and most biofuels are regarded as carbon-positive, due to the positive net emissions of C from their production and use. Carbon-neutral biofuels are those that result in no net emissions of C resulting from their production and use. Pyrolysis energy production combined with biochar incorporation into soil has been described as what may be the only carbon-negative energy production system known (Mathews, 2008). Carbon-negative energy (Figure 25.2) sequesters more C from the atmosphere than is released through its production and use. If biochar is produced from waste feedstocks that would otherwise be microbially degraded and contribute to anthropogenic C emissions, then it is possible to sequester a significant portion of anthropogenic C emissions through pyrolysis (Lehmann et al., 2006). While estimates of biochar C sequestration potential vary between studies depending on methodological details, recent studies reported 1.65 GtC/y or 19% of anthropogenic carbon dioxide emissions could be offset (Lee and Day, 2013; Lee et al., 2013). When combined with projected adoptions of renewable energy systems by the year 2100, it has been estimated that through the pyrolysis of agricultural residues, silvicultural residues, organic waste from industry, and urban waste, 5.5e9.5 GtC/y is achievable (Lehmann et al., 2006). This would exceed current fossil fuel emissions and thereby represent a potential remediation, as opposed to a conservation tool. Others have recently projected the possibility of up to 15.6 GtC/y (Smith et al., 2013). While the C sequestration potential of biochar is currently appealing, it must be deployed carefully in order to minimize the risk of damaging the soil, due to the irreversibility of biochar application to soil (Sohi et al., 2008). Sorption of Plant Nutrients and Other Pollutants The potential benefits of biochar to improving soil health through nutrient addition, and consequent improvements in fertilizer use efficiency have been well recognized through glasshouse and field trials (Sohi et al., 2010; Verheijen et al., 2010). These studies have shown that biochar and biochar-amended soil help retain plant nutrients and this is one of the means by which biochar application to soils is known to improve soil quality and increase crop yields in some cases. However, it is important to recognize that if the nutrients are retained by biochar particles in soil to a degree that plants are unable to take up the nutrients, this could impact productivity (Kookana et al., 2011). Studies have shown that retention of nitrogen (N) (Hollister et al., 2013; Liu et al., 2013), as well as phosphorus (P) and potassium (K) (Schnell et al., 2012), is achievable through incorporation of biochar, demonstrating potential agronomic and wider environmental benefits. Within the context of an agroecosystem, plant nutrients are essential and often the limiting factor
439 UTILIZATION OF BIOCHAR FOR ENVIRONMENTAL QUALITY Energy use Carbon Carbon Carbon Carbon-neutral Carbon Energy use Carbon Carbon-positive Carbon Carbon Carbon Energy use Carbon-negative FIGURE 25.2 Carbon-positive, carbon-neutral, and carbon-negative bioenergy systems. Source: Adapted from Mathews, 2008. determining the extent of crop growth. However, once N or P leave an agroecosystem, either through overland flow or leaching, they could potentially pose a threat to surface waters, with P being the limiting nutrient governing eutrophication of fresh water and N being the limiting nutrient governing the eutrophication of estuarine and ocean systems (Brady and Weil, 2008). Similar to gaseous C emissions, aqueous N and P losses from agricultural soil have global effects. While the majority of biochar research focuses on short-term impacts of its application, more long-term field research focused on net C sequestration, net GHG emissions, microbial community dynamics, nutrient use efficiency, and water use efficiency is needed (Ippolito et al., 2012). Furthermore, an increased fundamental understanding of the mechanisms underlying the interactions between biochar and soil in order to optimize agricultural systems and protect the environment should be a further focus (Spokas et al., 2012b). While applied plant nutrients outside of an agricultural context represent a threat to the environment that biochar has demonstrated potential to address, it is possible to retain other pollutants (e.g. heavy metals and pesticides) using biochar, as well. When biochar is applied to cadmium (Cd)-, copper (Cu)-, and lead (Pb)-contaminated soils, these metals have been observed to become immobilized, decreasing phytotoxicity and bioavailability, and vastly improved crop production (Park et al., 2011). Other studies have shown biochar to exhibit strong sorption and degradation inhibition of pesticide residues, leading to potential concerns regarding long-term accumulation in biochar amended soils treated with pesticides (Kookana, 2010). Some biochars have also been shown to retain estrogenic steroid hormones on dairy farm soils (Sarmah et al., 2010). While the potential for soil organic and inorganic contaminants (e.g. metal remediation, pesticide accumulation, and hormone retention) remediation are valid, it is important to consider that the enormous heterogeneity of biochars with respect to their chemical qualities and resultant effects on soil pollutants remain largely uninvestigated (Kookana et al., 2011). Research on potential agronomic and environmental applications of biochar is currently in its infancy and it is through the establishment and monitoring of additional long-term field trials that its full potential could be realized (Sarmah, 2009). Soil Greenhouse Gas Emissions In addition to the potential of biochar to partially offset anthropogenic C emissions through C sequestration, biochar has been observed to inhibit the release of GHGs from soil, thereby reducing net emissions of GHGs as a side effect of C sequestration. Decreased carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O) emissions from soils have been observed following biochar applications (Spokas et al., 2010), although increased emissions of N2O have been
440 25. ADVANCES AND INNOVATIONS IN BIOCHAR PRODUCTION AND UTILIZATION FOR IMPROVING ENVIRONMENTAL QUALITY (a) (b) Feedstock drying and shredding Feedstock drying and shredding Preparation Pyrolysis of feedstock Pyrolysis of feedstock Preparation Mixing with clay, minerals Biochar employed for surface water filtration Pyrolysis Pyrolysis Surface loading Conversion Application to soil Incorporation Incorporation Soil amelioration Soil amelioration Direct application Indirect application Torrefaction End use Biochar, bio oil, and biogas use Biochar FIGURE 25.3 Biochar, bio oil, and biogas use Biochar mineral complex End use (a) Biochar enhancement using clay. Adapted from Joseph et al., 2013. (b) Direct and indirect application of biochar to soil. observed in some cases (Clough et al., 2010). N2O, a GHG with a global warming potential hundreds of times greater than CO2, emissions have been shown to be reduced by a variety of biochars, yet the mechanisms responsible vary depending on soil moisture (Saarnio et al., 2013) and N content and forms (Kammann et al., 2012; van Zwieten et al., 2010). Additionally, biochar has been demonstrated to specifically reduce earthworm-derived CO2 and N2O emissions (Augustenborg et al., 2012). While there is an abundance of shortterm field and laboratory derived data to indicate the N2O emissions reduction potential of biochar, there is a dearth of information on the long-term field studies on this topic (Ussiri and Lal, 2013). Biochar is becoming increasingly broad in its contextual definition. At least one study has been conducted to evaluate the potential of biochar to reduce landfillderived CH4 emissions. A recent study reported that when landfill cover soil was mixed with 20% biochar, nearly 200% more CH4 was adsorbed compared to control soil, and 100% biochar was found to adsorb over 10-fold more CH4 than control soil (Yaghoubi, 2011). Biochar has also been shown to effectively decrease cattle enteric CH4 emissions while increasing animal growth, by 22% and 25%, respectively, when biochar was mixed with animal feed (Leng et al., 2012). alleviate the organic matter constraints of a given soil, using a procedure similar in effect to those used by chemical fertilizer companies to determine the macroand micronutrient requirements of a given soil (Joseph et al., 2013). While some researchers aim to produce improved biochars by simply blending multiple biochars together (Novak and Busscher, 2013), truly novel biochars are being produced by pyrolyzing an organic feedstock mixed with clay and other minerals, in order to produce a biochar mineral complex (BMC), which may represent the current forefront of custom biochar design. This BMC production process (Figure 25.3(a)) enables the production and customization of biochars through blending of materials before pyrolysis, followed by subsequent blending of materials with desirable chemical qualities, which are then subjected to torrefaction treatments. This can facilitate the loading of the biochar surfaces with additional plant-available nutrients and enhanced CEC, representing a clear progression from biochar as a soil conditioner toward biochar as an organic fertilizer (Lin et al., 2012a,b). POSTPYROLYSIS INDIRECT APPLICATION OF BIOCHAR Water Filtration Soil-Specific Biochar Design Published data on biochar and its interactions with soil are increasingly detailed in relation to feedstocks and production conditions (Sohi et al., 2010). As this body of literature grows, it will enable biochar producers to more predictably custom design biochars to While considerable research has been conducted on soil remediation through adsorption of pollutants following biochar application, relatively little data have been published on the sorption of pollutants by biochars preceding biochar application to soil. Most of the few studies that have been published on this subject
441 POSTPYROLYSIS INDIRECT APPLICATION OF BIOCHAR TABLE 25.2 Selected Examples of Biochar Employed As an Adsorbent to Remove Contaminants from Aqueous Solutions Biochar Temperature ( C) Function Source Pine Needles 180 U adsorption Zhang et al., 2013a Mulch 200 Urban road runoff Cu, Zn, and Cd adsorption Xaio-Jun et al., 2012 Dairy Manure 200, 350 Cu, Zn, and Cd adsorption Xu et al., 2013 Pine Wood 400e450 - Mohan et al., 2012 - F adsorption Pine Bark 400e450 F adsorption Mohan et al., 2012 Rice Husk 500e550 C6H5OH adsorption Liu et al., 2011 Corncob 500e550 C6H5OH adsorption Liu et al., 2011 Cottonwood and Maghemite (g-Fe2O3) 600 As adsorption Zhang et al., 2013b use biochar to filter substances from surface waters that are not suitable for application to agricultural soil, such as arsenic (As), Cd, Cu, fluoride (F), phenol (C6H5OH), uranium (U), and Zn (Table 25.2). Considering that these substances are better suited to be contained in a secure landfill than be applied on productive agricultural soil, further research should be conducted to determine if these biochars could be as effective at reducing CH4 emissions from soil as demonstrated by Yaghoubi (2011). Soil Nutrient Reclamation While some substances, such as heavy metals, pesticides, and hormones, are not desirable in soil amendments, other substances, such as N and P, are regarded as pollutants in surface waters, yet are essential plant nutrients within agroecosystems. If biochar is used to filter N and P from surface waters before in is incorporated into agricultural soil (Figure 25.3(b)), the process could reclaim a portion of the nutrients that are lost from agroecosystems to surface waters, and benefit both crop yield and surface water quality. If effective designer biochars are created for the purpose of N and P retention, it is feasible that those biochars, once saturated with N and P, could supply a crop area of some size with an adequate supply of runoff-derived nutrients to replace N and P fertilizer inputs entirely, and becoming effectively nutrient-neutral. Furthermore, it is feasible that if an analogous relocation of nutrients is performed where the deposition site is an area that would otherwise not receive fertilizer, and not be a likely source of runoff, such as a well-managed silvicultural system, then a nutrient-negative system could be created (Figure 25.4). Runoff and water-induced erosion occur when precipitation exceeds the infiltration capacity of a soil, and gravity carries it downhill over the soil surface into a drainage ditch or natural waterway. Common practices employed to reduce erosion caused by runoff and retain eroded soil particles by slowing the velocity of water flow within drainage ditches are well established. These practices include lining the sides of channels with either large angular rocks or rectangular wire mesh containers filled with smaller rocks, lining the bottom of the channel with grass sod, various bioengineering techniques involving the establishment of trees along the channel edge, and fixing straw bales in order to intercept eroded soil particles (Brady and Weil, 2008). Biochar contained within either reusable synthetic or single use biodegradable mesh containers could be used to simultaneously slow and filter overland flow (Figure 25.5). This may be complementary or even preferable to the aforementioned runoff and erosion control methods due to the added benefits of nutrient-saturated biochar application to soil (Figure 25.6). In areas where runoff currently flows directly into natural waterways, an enclosed biochar overland flow filter (Figure 25.7) fitted with mesh containers of biochar may be an effective option for reducing nutrient losses to surface water, allowing those nutrients to be relocated to soil and ultimately into plant biomass. Similarly, tile drain effluent may be filtered using mesh containers of biochar. Like the biochar overland flow filter discussed above, end caps on the tile drain with drain holes only in the upper half will increase the residence time of nutrient-laden water with the biochar and may increase filtration efficiency (Figure 25.8). Biochar As Container Growth Medium and Container Biochar has been reported to have the potential to be used as an amendment in plant nurseries.
442 25. ADVANCES AND INNOVATIONS IN BIOCHAR PRODUCTION AND UTILIZATION FOR IMPROVING ENVIRONMENTAL QUALITY Nutrients Nutrients Nutrients Nutrients Nutrients Nutrient use Nutrient use Nutrients Nutrients Nutrients Nutrients Nutrient positive Nutrient use Nutrient neutral Nutrient negative FIGURE 25.4 Nutrient-positive, nutrient-neutral, and nutrient-negative agricultural and silvicultural systems. For example, when 25% biochar was mixed with 75% peat, enhanced hydraulic conductivity and water retention were observed (Dumroese et al., 2011). Additionally, this study showed that the expansion of pelletized biochar (biochar that has been compressed with a binding agent in order to increase particle size), when wetted nearly offset the shrinkage typically exhibited by peat over time. A coconut fiber and tuff growing medium was shown to induce improved resistance of tomatoes to the necrotrophic Run off and ero sion Bio ov char flow erland fil dra ina ter in ge ditc h Slope fungus Botrytis cinerea when mixed with biochar at rates as low as 0.5% w/w (Elad et al., 2011). In another study, coconut fiber and tuff growing medium was shown to increase leaf size, plant height, flower development, and crop yield in pepper plants across all application rates from 1% to 5% w/w (Graber et al., 2010). In addition to container growth media, research is currently being conducted into the effects of plant containers constructed from molded biochar (Pulver, 2013). Run off and ero sion Bio ov char flow erland fil dra ina ter in ge ditc h Slope Nutrient saturated biochar FIGURE 25.5 Biochar overland flow filter stage 1. (For color version of this figure, the reader is referred to the online version of this book.) FIGURE 25.6 Biochar overland flow filter stage 2. (For color version of this figure, the reader is referred to the online version of this book.)
443 CONCLUSIONS, KNOWLEDGE GAPS, AND RESEARCH NEEDS Pipe Runoff inlet Hinge Pipe section connector End cap and drain holes Soil / slope Water line Runoff inlet Hinge FIGURE 25.7 Enclosed biochar overland flow filter concept with and without end cap to maximize liquid residence time for use in areas where runoff currently flows directly into natural waterways. (For color version of this figure, the reader is referred to the online version of this book.) End cap without drain holes End cap with drain holes Pipe section connector Water line FIGURE 25.8 Biochar tile drain flow filter concept with end cap for maximizing liquid residence time. (For color version of this figure, the reader is referred to the online version of this book.) CONCLUSIONS, KNOWLEDGE GAPS, AND RESEARCH NEEDS It is critical to understand that biochar is not a single material, but rather an entire class of materials (Spokas et al., 2012a) with a broad spectrum of chemical, physical, and biological properties that are drawn from both the diversity of feedstocks, production methods, and postproduction intermediary uses. It is also equally important to recognize the environmentally beneficial functions that biochar can perform after production and before application to soil and that there may be
444 25. ADVANCES AND INNOVATIONS IN BIOCHAR PRODUCTION AND UTILIZATION FOR IMPROVING ENVIRONMENTAL QUALITY Feedstock production and preparation Pyrolysis Ruminant feed supplement Landfill cover Water filtration, container media, and containers Soil FIGURE 25.9 Biochar end use decision process. desirable uses for biochars that are not suited to soil amelioration (Figure 25.9). Long-term field trial data related to biochar functions and properties as they change over time are extremely limited (Verheijen et al., 2010). Glasshouse projects that may display potential field scale benefits of biochar should be conducted, and continually monitored in order to measure, rather than project, what may be achievable for agriculture and the environment using biochar. It will be essential moving forward to be able to predict the sorption longevity and saturation point of biochars for pesticides and other pollutants. It is unknown if over time biochar in soil will lose or retain its ability to deactivate herbicides (Kookana et al., 2011). Similar temporal uncertainties exist in relation to most other biochar characteristics, aside from C stability. The understanding of both the short- and long-term effects of biochar on soil microbial communities remains limited (Sohi et al., 2008), yet is of critical importance due to the important role of microbes in many nutrient cycles and pollutant degradation pathways. Biochar uses that precede its incorporation into soil remain largely uninvestigated. Research related to the potential suitability of biochar for intermediate uses before application to soil, such as surface water filtration, enteric mitigation of methane production in ruminants, container media, and landfill cover is almost nonexistent. However, biochar itself has only recently expanded to become the focus of scientific research worldwide, so perhaps research into indirect biochar uses will progress accordingly. Acknowledgments The authors would like to thank Christian Pulver at Cornell University for his comments on the chapter. References Augustenborg, C.A., Hepp, S., Kammann, C., Hagan, D., Schmidt, O., Müller, C., 2012. Biochar and earthworm effects on soil nitrous oxide and carbon dioxide emissions. J. Environ. Qual. 41, 1203e1209. Barrow, C., 2012. Biochar: potential for countering land degradation and for improving agriculture. Appl. Geogr. 34, 21e28. Bolan, N.S., Kunhikrishnan, A., Choppala, G.K., Thangarajan, R., Chung, J.W., 2012. Stabilization of carbon in composts and biochars in relation to carbon sequestration and soil fertility. Sci. Total Environ. 424, 264e270. Brady, N.C., Weil, R.R., 2008. The Nature and Properties of Soils. Prentice Hall, Upper Saddle River pp. 1e980. Busch, D., Kammann, C., Grünhage, L., Müller, C., 2012. Simple biotoxicity tests for evaluation of carbonaceous soil additives: establishment and reproducibility of four test procedures. J. Environ. Qual. 41, 1023e1032. Clough, T.J., Bertram, J.E., Ray, J.L., Condron, L.M., O’Callaghan, M., Sherlock, R.R., Wells, N.S., 2010. Unweathered wood biochar impact on nitrous oxide emissions from a bovine-urine-amended pasture soil. Soil Sci. Soc. Am. J. 74, 852e860. Downie, A., van Zwieten, L., 2013. Biochar: a coproduct to bioenergy from slow-pyrolysis technology. In: Lee, J.W. (Ed.), Advanced Biofuels and Bioproducts. Springer ScienceþBusiness Media, New York, pp. 97e117. Dumroese, R.K., Heiskanen, J., Englund, K., Tervahauta, A., 2011. Pelleted biochar: chemical and physical properties show potential use as a substrate in container nurseries. Biomass Bioenergy 35, 2018e2027. Elad, Y., Cytryn, E., Meller Harel, Y., Lew, B., Graber, E.R., 2011. The biochar effect: plant resistance to biotic stresses. Phytopathol. Mediterr. 50, 335e349. Fischer, D., Glaser, B., 2012. Synergisms between compost and biochar for sustainable soil amelioration. In: Kumar, S. (Ed.), Management of Organic Waste. InTech, New York, pp. 167e198. Flannery, T., 2009. Foreword. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science and Technology. Earthscan Publications, Oxford, pp. xxviiexxviii. Gore, A., 2009. Our Choice: A Plan to Solve the Climate Crisis. Rodale Books, Emmaus, 1e416. Graber, E.R., Harel, Y.A., Kolton, M., Cytryn, E., Silber, A., David, D.R., Tsechansky, L., Borenshtein, M., Elad, Y., 2010. Biochar impact on development and productivity of pepper and tomato grown in fertigated soilless media. Plant Soil 337, 481e496. Hansen, J., Sato, M., Kharecha, P., Beerling, D., Berner, R., MassonDelmotte, V., Pagani, M., Raymo, M., Royer, D.L., Zachos, J.C., 2008. Target atmospheric CO2: where should we aim? Open Atmos. Sci. J. 2, 217e231. Hollister, C.C., Bisogni, J.J., Lehmann, J., 2013. Ammonium, nitrate, and phosphate sorption to and solute leaching from biochars prepared from corn stover (Zea mays L.) and oak wood (Quercus spp.). J. Environ. Qual. 42, 137e144. Ippolito, J.A., Laird, D.A., Busscher, W.J., 2012. Environmental benefits of biochar. J. Environ. Qual. 41, 967e972. Jahirul, M.I., Rasul, M.G., Chowdhury, A.A., Ashwath, N., 2012. Biofuels production through biomass pyrolysis-a technological review. Energies 5, 4952e5001. Joseph, S., van Zwieten, L., Chia, C., Kimber, S., Munroe, P., Lin, Y., Marjo, C., Hook, J., Thomas, T., Nielsen, S., Donne, S., Taylor, P., 2013. Designing specific biochars to address soil constraints: a developing industry. In: Ladygina, N., Rineau, F. (Eds.), Biochar and Soil Biota. CRC Press, Boca Raton, pp. 165e201. Kammann, C., Ratering, S., Eckhard, C., Müller, C., 2012. Biochar and hydrochar effects on greenhouse gas (carbon dioxide, nitrous oxide, and methane) fluxes from soils. J. Environ. Qual. 41, 1052e1066. Kloss, S., Zehetner, F., Dellantonio, A., Hamid, R., Ottner, F., Liedtke, V., Schwanninger, M., Gerzabek, M.H., Soja, G., 2012. Characterization of slow pyrolysis biochars: effects of feedstocks
REFERENCES and pyrolysis temperature on biochar properties. J. Environ. Qual. 41, 990e1000. Kookana, R.S., 2010. The role of biochar in modifying the environmental fate, bioavailability, and efficacy of pesticides in soils: a review. Aust. J. Soil Res. 48, 627e637. Kookana, R.S., Sarmah, A.K., van Zwieten, L., Krull, E., Singh, B.P., 2011. Biochar application to soil: agronomic and environmental benefits and unintended consequences. In: Sparks, D.L. (Ed.). Advances in Agronomy, vol. 112. Elsevier, Amsterdam, pp. 103e143. Krull, E., Lehmann, J., Skjemstad, J., Baldock, J., Spouncer, L., 2008. The global extent of black carbon in soils: is it everywhere? In: Schröder, H.G. (Ed.), Grasslands: Ecology, Management and Restoration. Nova Science Publishers, Inc, New York, pp. 13e17. Laird, D.A., Brown, R.C., Amonette, J.E., Lehmann, J., 2009. Review of the pyrolysis platform for co-producing bio-oil and biochar. Biofuels, Bioprod. Biorefin. 3, 547e562. LeCroy, C., Masiello, C.A., Rudgers, J.A., Hockaday, W.C., Silberg, J.J., 2013. Nitrogen, biochar, and mycorrhizae: alteration of the symbiosis and oxidation of the char surface. Soil Biol. Biochem. 58, 248e254. Lee, J.W., Day, D.M., 2013. Smokeless biomass pyrolysis for producing biofuels and biochar as a possible arsenal to control climate change. In: Lee, J.W. (Ed.), Advanced Biofuels and Bioproducts. Springer ScienceþBusiness Media, New York, pp. 23e34. Lee, J.W., Hawkins, B., Li, X., Day, D.M., 2013. Biochar fertilizer for soil amendment and carbon sequestration. In: Lee, J.W. (Ed.), Advanced Biofuels and Bioproducts. Springer ScienceþBusiness Media, New York, pp. 57e68. Lehmann, J., 2007. Bio-energy in the black. Front. Ecol. Environ. 5, 381e387. Lehmann, J., 2008. Terra preta nova: where to from here? In: Woods, W.I., Teixeira, W.G., Lehmann, J., Steiner, C., WinklerPrins, A.M.G.A., Rebellato, L. (Eds.), Amazonian Dark Earths: Wim Sombroek’s Vision. Springer ScienceþBusiness Media, New York, pp. 473e486. Lehmann, J., Gaunt, J., Rondon, M., 2006. Bio-char sequestration in terrestrial ecosystems-a review. Mitig. Adapt. Strat. Gl. 11, 403e427. Lehmann, J., Rillig, M.C., Thies, J., Masiello, C.A., Hockaday, W.C., Crowley, D., 2011. Biochar effects on soil biota-a review. Soil Biol. Biochem. 43, 1812e1836. Lehmann, J., Rondon, M., 2006. Bio-char soil management on highly weathered soils in the humid tropics. In: Uphoff, N., Ball, A.S., Fernandes, E., Herron, H., Husson, O., Liang, M., Palm, C., Pretty, J., Sanchez, P., Sanginga, N., Thies, J. (Eds.), Biological Approaches to Sustainable Soil Systems. CRC Press, Boca Raton, pp. 517e530. Leng, R.A., Preston, T.R., Inthapanya, S., 2012. Biochar reduces enteric methane and improves growth and feed conversion in local “Yellow” cattle fed cassava root chips and fresh cassava foliage. Livest. Res. Rural Dev. 24. Article #212. Retrieved March 14, 2013, from: http://www.lrrd.org/lrrd24/12/sang24212.htm. Lin, Y., Munroe, P., Joseph, S., Henderson, R., 2012a. Migration of dissolved organic carbon in biochars and biochar-mineral complexes. Pesqui. Agropecu. Bras. 47, 677e686. Lin, Y., Munroe, P., Joesph, S., Ziolokowski, A., van Zwieten, L., Kimber, S., Rust, J., 2012b. Chemical and structural analysis of enhanced biochars: thermally treated mixtures of biochar, chicken litter, clay and minerals. Chemosphere 91, 35e40. Liu, N., Sun, Z., Wu, Z., Zhan, X., Zhang, K., Zhao, E., Han, X., 2013. Adsorption characteristics of ammonium nitrogen by biochar from diverse origins in water. Adv. Mater. Res. 664, 305e312. Liu, W., Zeng, F., Jiang, H., Zhang, X., 2011. Preparation of high adsorption capacity bio-chars from waste biomass. Bioresour. Technol. 102, 8247e8252. 445 Lovelock, J., 2009. The Vanishing Face of Gaia: A Final Warning. Basic Books, New York, 1e304. Mathews, J.A., 2008. Carbon-negative biofuels. Energy Policy 36, 940e945. Mohan, D., Sharma, R., Singh, V.K., Steele, P., Pittman, C.U., 2012. Fluoride removal from water using bio-char, a green waste, lowcost adsorbent: equilibrium uptake and sorption dynamics modeling. Ind. Eng. Chem. Res. 51, 900e914. Novak, J.M., Busscher, W.J., 2013. Selection and use of designer biochars to improve characteristics of southeastern USA coastal plain degraded soils. In: Lee, J.W. (Ed.), Advanced Biofuels and Bioproducts. Springer ScienceþBusiness Media, New York, pp. 69e96. Park, J.H., Choppala, G.K., Bolan, N.S., Chung, J.W., Chuasavathi, T., 2011. Biochar reduces the bioavailability and phytotoxicity of heavy metals. Plant Soil 348, 439e451. Prost, K., Borchard, N., Siemens, J., Kautz, T., Séquaris, J., Möller, A., Amelung, W., 2013. Biochar affected by composting with farmyard manure. J. Environ. Qual. 42, 164e172. Pulver, C., 2013. Personal communication. Rogovska, N., Laird, D., Cruse, R.M., Trabue, S., Heaton, E., 2012. Germination tests for assessing biochar quality. J. Environ. Qual. 41, 1014e1022. Saarnio, S., Heimonen, K., Kettunen, R., 2013. Biochar addition indirectly affects N2O emissions via soil moisture and plant N uptake. Soil Biol. Biochem. 58, 99e106. Sarmah, A.K., 2009. Potential risk and environmental benefits of waste derived from animal agriculture. In: Ashworth, G.S., Azevedo, P. (Eds.), Agricultural Wastes. Nova Science Publishers, Inc, New York, pp. 1e18. Sarmah, A.K., Srinivasan, P., Smernik, R.J., Manley-Harris, M., Antal Jr., M.J., Downie, A., van Zwieten, L., 2010. Retention capacity of biochar-amended New Zealand dairy farm soil for an estrogenic steroid hormone and its primary metabolite. Aust. J. Soil Res. 48, 648e658. Schmidt, M.W.I., Skjemstad, J.O., Czimczik, C.I., Glaser, B., Prentice, K.M., Gelinas, Y., and Kuhlbusch, T.A.J., 2001. Comparative analysis of black carbon in soils. Global Biogeochem. Cy. 15, 163e167. Schimmelpfennig, S., Glaser, B., 2012. One step forward toward characterization: some important material properties to distinguish biochars. J. Environ. Qual. 41, 1001e1013. Schnell, R.W., Veitor, D.M., Provin, T.L., Munster, C.L., Capareda, S., 2012. Capacity of biochar application to maintain energy crop productivity: soil chemistry, sorghum growth, and runoff water quality effects. J. Environ. Qual. 41, 1044e1051. Smith, P., Haberl, H., Popp, A., Erb, K., Lauk, C., Harper, R., Tubiello, F., de Siqueira Pinto, A., Jafari, M., Sohi, S., Masera, O., Böttcher, H., Berndes, G., Bustamante, M., Ahammad, H., Clark, H., Dong, H., Elsiddig, E. A., Mbow, C., Ravindranath, N. H., Rice, C. W., Robledo-Abad, C., Romanovskaya, A., Sperling, F., Herrero, M., House, J. I., Rose, S, in press. How much land based greenhouse gas mitigation can be achieved without compromising food security and environmental goals? Glob. Change Biol. 19, 2285-2302 Sohi, S.P., Krull, E., Lopez-Capel, E., Bol, R., 2010. A review of biochar and its use and function in soil. In: Sparks, D.L. (Ed.). Advances in Agronomy, vol. 112. Elsevier, Amsterdam, pp. 47e82. Sohi, S., Lopez-Capel, E., Krull, E., Bol, R., 2008. In: Krull, E. (Ed.), Biochar, Climate Change and Soil: a Review to Guide Future Research. CSIRO, Collingwood, pp. 1e56. Sparkes, J., Stoutjesdijk, P., 2011. Biochar: Implications for Agricultural Productivity. ABARES, Canberra, 1e48. Spokas, K.A., Baker, J.M., Reicosky, D.C., 2010. Ethylene: potential key for biochar amendment impacts. Plant Soil 333, 443e452. Spokas, K.A., Cantrell, K.B., Novak, J.M., Archer, D.W., Ippolito, J.A., Collings, H.P., Boateng, A.A., Lima, I.M., Lamb, M.C.,
446 25. ADVANCES AND INNOVATIONS IN BIOCHAR PRODUCTION AND UTILIZATION FOR IMPROVING ENVIRONMENTAL QUALITY McAloon, A.J., Lentz, R.D., Nichols, K.A., 2012a. Biochar: a synthesis of its agronomic impact beyond carbon sequestration. J. Environ. Qual. 41, 973e989. Spokas, K.A., Novak, J.M., Venterea, R.T., 2012b. Biochar’s role as an alternative N-fertilizer: ammonia capture. Plant Soil 350, 35e42. Ussiri, D., Lal, R., 2013. Nitrous oxide sources and mitigation strategies. In: Ussiri, D., Lal, R. (Eds.), Soil Emission of Nitrous Oxide and Its Mitigation. Springer ScienceþBusiness Media, New York, pp. 243e275. van Zwieten, L., Kimber, S., Morris, S., Downie, A., Berger, E., Rust, J., Scheer, C., 2010. Influence of biochars on flux of N2O and CO2 from a ferrosol. Aust. J. Soil Res. 48, 555e568. Vaughan, N.E., Lenton, T.M., 2011. A review of climate geoengineering proposals. Clim. Change 109, 745e790. Verheijen, F., Jeffery, S., Bastos, A.C., van der Velde, M., Diafas, I., 2010. Biochar Application to Soils - A Critical Scientific Review of Effects on Soil Properties, Processes and Functions. EUR 24099 EN. Office for the Official Publications of the European Communities, Luxembourg. pp. 1e149. Warnock, D.D., Lehmann, J., Kuyper, T.W., Rillig, M.C., 2007. Mycorrhizal responses to biochar in soildconcepts and mechanisms. Plant Soil 300, 9e20. Winsley, P., 2007. Biochar and bioenergy production for climate change mitigation. N. Z. Sci. Rev. 64, 5e10. Woolf, D., Amonette, J.E., Perrott, F.A., Lehmann, J., Joseph, S., 2010. Sustainable biochar to mitigate global climate change. Nat. Commun. 1, 1e9. Xiao-Jun, Z., Da-Fang, F., He, L., 2012. Adsorption removal of copper, zinc and cadmium in aqueous solutions and road runoff by carbonized mulch: heavy metal removal by carbonized mulch. In: Hang, P.S. (Ed.), Proceedings of the 2012 International Conference on Biomedical Engineering and Biotechnology. IEEE Computer Society, Washington, pp. 1180e1185. Xu, X., Cao, X., Zhao, L., Wang, H., Yu, H., Gao, B., 2013. Removal of Cu, Zn, and Cd from aqueous solutions by the dairy manurederived biochar. Environ. Sci. Pollut. Res. 20, 358e368. Yaghoubi, P., 2011. Development of Biochar-Amended Landfill Cover for Landfill Gas Mitigation (Thesis). University of Illinois at Chicago, USA. Zhang, M., Gao, B., Varnoosfaderani, S., Hebard, A., Yao, Y., Inyang, M., 2013a. Preparation and characterization of a novel magnetic biochar for arsenic removal. Bioresour. Technol. 130, 457e462. Zhang, Z., Cao, X., Liang, P., Liu, Y., 2013b. Adsorption of uranium from aqueous solution using biochar produced by hydrothermal carbonization. J. Radioanal. Nucl. Chem. 295, 1201e1208.
C H A P T E R 26 Biochar Processing for Sustainable Development in Current and Future Bioenergy Research Mark P. McHenry School of Engineering and Information Technology, Murdoch University, Perth, Western Australia, Australia email: mpmchenry@gmail.com O U T L I N E Can Biochars Increase Livestock Growth Rates, or Provide a New Market for Semiarid Forestry? 453 A Comparison of Biochar Carbon Value for Different 454 Potential Income Streams Introduction 447 Theoretical Income Streams Renewable Energy and Fuel Generation Carbon Sequestration of Biochars and Carbon Markets 448 448 449 Conclusion 454 Agricultural Benefits 450 Disclaimers 455 Economic Analysis 451 Can Biochar Be a Cost-effective Fertilizer Substitute? 451 Can Biochar Be a Cost-Effective Approach to 452 Increase Grain Crop Primary Productivity? References 455 INTRODUCTION Rural biomass energy and carbon options seem to offer increased financial resilience to agricultural enterprises relative to fluctuating seasonal growing conditions and uncertain market prices of inputs, products, and exchange rates. The projected increases in farming costs from any future inclusion of the agricultural sector from carbon pricing may be offset by additional net income from such rural biomass-based sequestration and renewable energy activities. Cellulose, hemicelluloses, and lignin are the main components of wood and crop residues of known potential for bioenergy and stable carbon forms, and the management of which requires detailed agronomic, technical, and market information. Thus, there is a synergistic match between growing Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00026-7 food and growing biomass for energy and carbon in the same rural enterprise. Modern concepts of biochar-agricultural systems and their respective projected financial viabilities have been outlined in the existing literature (Lehmann and Joseph, 2009). These systems commonly incorporate complex semi-industrial operations with rural and forestry biomass as well as small-scale low-technology concepts with farm waste and domestic heating. To narrow research specificity, this work focused on the West Midlands of the Northern Agricultural Region of Western Australia (WA), and uses Australian dollars. (At the time of writing the Australian and US currencies were roughly parity.) To date, this region is one of the few regions of Australia that has exhibited economically encouraging agricultural responses from biochar addition, and has an established 447 Copyright Ó 2014 Elsevier B.V. All rights reserved.
448 26. BIOCHAR PROCESSING FOR SUSTAINABLE DEVELOPMENT IN CURRENT AND FUTURE BIOENERGY RESEARCH practice of profitable grazing leguminous fodder shrubs, which is a potentially large and sustainable biomass supply. THEORETICAL INCOME STREAMS The potential income streams in the West Midlands from above ground rural biomass include renewable energy and fuel generation, carbon sequestration of biochars, and agricultural benefits from the use of biochar and ash from energy and fuel generation or charring alone. It is likely that tradable sequestered carbon will be reliant on the supplies from bioenergy generation plants that are able to comply with both emission and biochar quality standards. However, a price on carbon may help offset the additional costs of the coproduction of electricity and biochar from biomass. (Table 26.1 outlines key benefits, costs and barriers to biochar compliance to carbon markets.) The Australian Farm Institute (2011) estimates an income reduction of between 1.4% and 1.6% from a carbon price based on electricity consumption for a WA mixed farming enterprise of 4900 ha (2400 ha cropped and about 2000 head of livestock, mainly sheep), assuming agriculture and transport fuels are excluded from any carbon liability (Australian Farm Institute, 2011). In contrast to concerns of a carbon price reducing agricultural profitability, this work presents the case that integrating new sequestration options into conventional production systems from low-cost biochars produced from agricultural wastes (with sufficient operational safety considerations) may offset costs in the West Midlands. The profitability of income streams (presented in Table 26.1) are highly sensitive to (and often dependent upon) government subsidies for renewable energy, a carbon price, and TABLE 26.1 also the location-specific demand for biochar and energy. Similarly, agricultural effects of biochar addition will vary more with soil type, seasonal conditions, and animal nutrition characteristics. In complex and uncertain circumstances, predictive modeling can become particularly challenging. However, the agricultural effects (where they occur) will likely provide a more solid basis for emerging industry development than the highly sensitive and evolving carbon and electricity markets. If agricultural benefits, initially at least, exhibit less risky investments to individual farms than bioenergy cooperatives or carbon sequestration pooling activities, then agricultural benefits may be a more suitable foundation for the establishment of biomass-based industrial developments than energy or carbon sequestration policies in the initial phases. Renewable Energy and Fuel Generation In simple terms, wood-fired stoves, barbeques, or water heaters are a biomass-based renewable energy system. Yet, the growing range of new medium- to large-scale bioenergy technologies include gasifier and pyrolysis power stations coproducing electricity, heat, and a range of biofuels. Nonetheless, all traditional and new technologies convert the complex hydrocarbon molecules in biomass to hydrogen, methane, carbon monoxide, carbon dioxide, and numerous other gasses, including polyaromatic hydrocarbons and dioxins. Some technologies also produce liquid and soil fuels (such as biochar) from the same biomass. In general, while small-scale and simple technology designs have less control and efficiency, they exhibit lower capital and operating costs, although they are usually more labor intensive per unit production of output (McHenry, 2012b). At the regional scale biomass power plant technology choices often Outline of Key Potential Income Streams from Rural Biomass in the West Midlands Income stream Carbon Form Benefits Costs Barriers Renewable Energy from Biomass Crop residues, woody harvest/coppice, manures Electricity, fuel Capital and running costs Competition from other renewable energy technologies Carbon Markets Trees > 2 m, soil carbon, biochar Carbon sequestration Establishment, practice change, manufacture/ purchase and application, monitoring Accreditation and acceptable methodologies Agricultural Benefits Biochar from crop waste or woody biomass or manures Increased grain yield at maintenance P, reduced fertilizer requirement, detannification of livestock feed, composting accelerant Manufacture/ purchase, application High soil P levels, crop and pasture benefits restricted to sand and gravelly soil types (these are more common in the West Midlands than some other regions)
449 THEORETICAL INCOME STREAMS TABLE 26.2 Performance of a Selected Range of Available Biomass Conversion Technologies that May be Suitable to Some West Midland Applications Technology Cost Electrical Output Application Challenges Gasifier Power Station (Waste to Energy) w$50 million w30 GWh/yr Regional landfill Biochar contamination, transport costs, gas cleaning Rainbow Bee-Eater (Crucible Carbon Slow Pyrolysis Design) e w1 MWh/t dry straw and 350 kg char Regional center near substation New technology but clean gas Slow Pyrolysis (BEST) w$15 million e Regional center near substation or customer Gas cleaning Updraft Gasifier (Big Char) w$0.25 million Nil Mobile plant for biomass conversion to biochar (w25% efficiency) Conversion rate and biochar quality? Woodgas Genset (Powerpallet) w$25,000 20 kW On farm Current price of diesel, biochar production rate, emissions? Simple Drum Kilns Low Nil On farm Biochar production rate, emissions, biochar quality include gasifiers (which optimize gas production), and slow pyrolysers (which optimize biochar production). A general outline of the variations in biomass renewable energy technologies are shown in Table 26.2. In terms of developing a regional energy/biochar industry, medium-sized biochar production units may address concerns of soil nutrient loss from harvested biomass. Despite the generally high costs of transporting timber trees, transporting returned biochar is relatively efficient on a weight basis, as the biochar mass is 70e80% less than the original dry biomass (Lehmann, 2007). Nonetheless, industrial biochar production and use will require a number of safeguards. Handling risks include flammability concerns, and the dusts can spontaneously combust in enclosed spaces and is comparable to the risk of handling some metals, foods (flour, etc.), coal, plastics, and woods (Joseph, 2007). TABLE 26.3 Carbon Sequestration of Biochars and Carbon Markets Biochars contain very stable forms of carbon (fixed carbon) (Blackwell et al., 2010). The proportion of the original biomass carbon as total solid carbon in biochar ranges from around 5% from gasification technologies to about 35% in slow pyrolysis technologies. Wellmanaged large-scale biomass power plants can produce biochar of a consistent quality, whereas often smallscale “low-tech” technologies tend to produce very variable quality biochars, and are highly dependent on the homogeneity of the heating regime and the duration of heating. (See Tables 26.3 and 26.4.) A low-cost highvolume supply of sustainable biochar with a high carbon fraction will be needed to generate meaningful climate change mitigation benefits. However, currently Conversion Efficiencies and Outputs of a Range of Technologies Producing Biochars Technology Power Output Biochar Yield Fixed C Limitations Gasification Maximum w5% e Some waste contamination. Biochar credit may be owned by the power plant operators Slow Pyrolysis Some w35% w65% Biochar credit may be owned by the power plant operators Crucible Carbon (Slow Pyrolysis) Some 35% 70% Project developing Kiln Pyrolysers (Big Char) Nil 25% e Simple Drum Kiln (Ogawa) Nil e Output rate Drum Kiln (TLUD) Nil e Output rate Pit Kiln Nil 12.5e30% (Brown, 2009) Output rate
450 26. BIOCHAR PROCESSING FOR SUSTAINABLE DEVELOPMENT IN CURRENT AND FUTURE BIOENERGY RESEARCH TABLE 26.4 Carbon Analysis of Biochars Used in Research in the West Midlands Biochar Source (Temp of Slow Pyrolysis) Total Carbon, % Fixed Carbon, % K, % P, % Wood Jarrah (600  C Simcoa)* 65 69 e e 45.7 76 e e 58.6 e 34 4 40.1 e 14 12  Wood Jarrah (600 C Wundowie)* y  Wheat Chaff (550 C) y  Chicken Manure (550 C) * Blackwell et al., 2010. y Krull, personal communication, Department of Agriculture, Fisheries and Forestry (a national initiative for biochar research), and Grains Research and Development Corporation (biochar for agricultural productivity), 2010. low carbon prices will not provide sufficient commercial incentives to simply apply biochar in soils for mitigation alone. The unacceptably high uncertainties of the direct and indirect influences and residence times of biochars and other organic carbon species in soils, their suitability for carbon markets (Intergovernmental Panel on Climate Change, 2000), and even the commercial incentives of high-volume production are fundamental barriers to widespread biochar use. Therefore, there is a growing need for researchers to quantify the net effect of specific biochars and application methods within niche agroecological systems (particularly grains and livestock) and to verify any stable sequestration of carbon fractions (McHenry, 2011). Furthermore, in terms of farm application risks, some biochars can contain toxic materials that are controlled by “permissible exposure limit” standards. The levels of these toxic materials in the biochar is dependent on both the biomass feedstock and the biochar manufacture process, thus no simple permissible exposure limit is available for biochar to date (Blackwell et al., 2009). Thus, the development of a secure and responsible biochar industry will require awareness of safe methods of handling agricultural inputs and will need to be justified economically, and be integrated with existing agricultural production systems (McHenry, 2011). AGRICULTURAL BENEFITS Available research to date has shown that biochar alters various soil properties in a number of ways. (See Table 26.5.) In the context of the siliceous sandy soils of the West Midlands, the most sought effects are improved microbial habitats and improved nutrient supplies from relatively low (w1 t/ha) rates of biochar use. (See Table 26.6 for crop and pasture research responses in the West Midlands.) It is clear that more research is needed on how various biochars influence the flows of nutrients through the soil profile (Lehmann et al., 2006; Laird et al., 2008), particularly under Australian conditions (McHenry, 2011). To date, the major claims have been related to biological immobilization of inorganic N, adsorption of dissolved ammonium, nitrates, P, and hydrophobic organic pollutants (Beaton et al., 1960; Gustafsson et al., 1997; Accardi-Dey and Gschwend, 2002; Lehmann et al., 2003; Bridle, 2004; Mizuta et al., 2004). However, the available research scope does not include an assessment of whether this adsorption could reduce some transport of agricultural fertilizers or other pollutants into ground and surface waters in agricultural catchments (Lehmann et al., 2006; Lehmann, 2007). Early work by Bridle (2004) suggested that biochar applications reduce nitrate leaching, as his research found levels of nitrate and ammonium did not change in soils for 56 days after application. The soil incubation study further revealed that in contrast, soil bicarbonate availability and plant available P levels would increase slowly (Bridle, 2004). The laboratory results suggested that biochar would provide a source of P for plant growth and could have applications on soils as a slow release form of P, yet some research suggest a reduced uptake of N. This may be more useful in deep sandy soils where P leaches from the surface into groundwater. Biochars are also hypothesized to slow the N cycle by increasing the carbon to N soil ratio, possibly due to increased soil aeration reducing anaerobic conditions (Lehmann et al., 2006). Rondon et al. (2005) found a significant reduction of nitrous oxide emissions, and a near-complete suppression of methane emissions in glasshouse environments at biochar additions of 30 g/kg of soil for some crops (Rondon et al., 2005; Lehmann et al., 2006). However, in some circumstances a high carbon to N ratio and abiotic buffering of mineral N may lead to low N availability (Lehmann and Rondon, 2006). Therefore, medium-scale crop biochar trials are required with regionally common soil biota and mineralogy, and also crop, pasture, and animals for greater understanding of commercial agricultural applicability in a particular region.
451 ECONOMIC ANALYSIS TABLE 26.5 Positive and Negative Effects on Plant Growth of Biochar Additions to Soil Effect (D) or (L) Process Rate of Application Appropriate Soils Neutralization (þ) Most biochars are alkaline and can adsorb Al3þ ions Often 10þ t/ha mixed in topsoil. Has worked on Krasnozems. Possibly very acid sands (Wodjils). Increase Water Holding (þ) Increased microporosity Effect proportional to rate and biochar character More for lower clay contents Increase Nutrient Holding (þ) Increased cation exchange capacity and anion exchange capacity? Effect proportional to rate and biochar character More for lower clay contents Increased Nutrient Supply (þ) Direct supply from biochar inorganic fraction Effect proportional to rate and biochar character All soils Reduced Mechanical Strength (þ) Lowers soil cohesion Higher rates Non sands Reduced N2O Emission (þ) Unclear Higher rates? 10 t/haþ? Especially poorly drained soils at risk of denitrification Microbial Habitat Improvement (þ) Micropores in the biochar help mycorrhizal fungi and bacteria survival, which use symbiosis to improve nutrient and water supply to plants which host them w1 t/ha banded Mainly low available P soils (Colwell <20 ppm) Phytotoxins () Carbon compounds such as phenols are retained on low temperature biochars and ones cooled with their own emissions Quenching, soaking and resting in the soil after incorporation may lower the concentration of these substances Depends on biomass and charring process Herbicide Adsorption (þ) and () Can increase application requirement and/or reduce leaching All rates All soils TABLE 26.6 Summary of Benefits to Crop and Pasture Production from Applied Biochars in the West Midlands (2007e2010) Crop Trial (Place) Yield/DM Increase Increase, % Fertilizer Soil Wheat (Mingenew) 0.23 t/ha 40 96 kg/ha super Colwell P 5 ppm Wheat (Mingenew) 0.45 t/ha 25 25 kg/ha DAP Colwell P 5.5 ppm Pasture (Irwin) 40 kg/ha DM 20 Nil Colwell P 10 ppm Clover (Tubes) 0.15 kg/ha at the rate of flowering 75 Nil Colwell P 5.5 ppm Clover (Badgi) Very visible Leached biosolids Gravelly sand DM, Dry matter. DAP, Di-ammonium phosphate. ECONOMIC ANALYSIS A recent analysis by Blackwell et al. (2010) on biochar effects on profitability of dryland wheat production in WA provided a perspective of the breakeven investment costs per hectare of different responses over the medium term. Table 26.7 shows that for the West Midlands area (high rainfall north) a 10% yield increase from 1 t/ha application of banded biochar with a declining response over 12 years would break even at $130/ha, based on the previous 12-year data. This breakeven cost included estimated biochar application costs of between $20 and $50/ha; thus a production/purchase and transport cost would need to be no higher than about $50e$100/t to enable some income from the biochar use, which encourages further work toward low-cost biochar production technology development (Blackwell et al., 2010). Can Biochar Be a Cost-effective Fertilizer Substitute? An analysis by McHenry (2012a,b) quantified the potential of using biochar as a soil amendment to displace annual applications of single superphosphate (SSP)
452 26. BIOCHAR PROCESSING FOR SUSTAINABLE DEVELOPMENT IN CURRENT AND FUTURE BIOENERGY RESEARCH TABLE 26.7 Breakeven Cost of Applied Biochar in Six Rainfall Regions in WA Trial Low Rainfall North (const.) Low Rainfall North (decl.) Medium Rainfall North (decl.) High Rainfall North (decl.) Low Rainfall South (decl.) Medium Rainfall South (decl.) High Rainfall South (decl.) 10% Yield Increase 170 100 140 130 100 120 140 70 40 50 50 40 40 40 240 140 190 190 150 160 180 50% P Fertilizer Reduction 10% Yield Increase and 50%, P Fertilizer Reduction With and without initial yield increase and/or P fertilizer reduction responses, which decline linearly to nit after 12 years (decl.) or are constant for 12 years (const.). (0% N, 8.8% P, 0% K, 11% S) in wheat cropping systems in WA. The analysis assumed two biochar applications over a 15-year period, applied in year zero, and year eight. The analysis ignored all production inputs and outputs, and only calculated the difference between using an average “full rate” of SSP (90 kg/ha), and a “half rate” of SSP with deep banded biochar equivalent to 1 t/ha. The 45 kg/ha year half-rate SSP application is approximately equivalent to an annual application of 4 kg of P/ha. The simplified analysis assumed that the use of either method would achieve an identical wheat yield, negating the requirement to model wheat prices. The application cost of deep banding the biochar (tons per hectare per application) was assumed to be $110. The annual application costs of both the rates of SSP were assumed to be $20/ha, goods and services tax (GST)1 inclusive. A range of biochar prices (delivered to farm, per ton) was analyzed: $0, $50, $100, $150, $200, $250, $300, $350, $400, and $450/t. Similarly, a range of SSP costs (delivered to the farm, per ton) were calculated: $250, $300, $350, $1250. A carbon price was included in the analysis, and was analyzed at intervals of $5 tCO2-e, between $0 and $100 tCO2-e. The analysis assumed a 0.8 carbon fraction recalcitrance. A real discount rate of 8% p.a. was used, and all capital and maintenance costs were based on average current prices and were GST inclusive. In summary, the results showed that without a carbon value the “half rate” of SSP (45 kg/ ha year) and biochar (1 t/ha application) were only cost competitive with the full rate of SSP (90 kg/ha year) when the biochar purchase price was unreasonably low (<w$20). At 2012 prices of SSP (generally between $200 and $450/t), the choice of using half SSP application rates with biochar additions at the above application rate assumptions were not an attractive option unless the biochar purchase price was practically zero. The net cost was also calculated assuming the carbon in the biochar was eligible in soil carbon markets, and the various potential prices of carbon were subtracted from the gross biochar purchase price. While the introduction of a carbon price would effectively subsidize biochar costs, very high carbon prices (>$100/t) were required for the sequestration value of biochar to simply equal the purchase price and cover costs of soil application (McHenry, 2012a). The low SSP price, the high market prices for biochar, and the high biochar soil application cost of deep banding relative to conventional broadcasting, all resulted in the option of halving SSP applications by using biochar an unattractive practice. Can Biochar Be a Cost-Effective Approach to Increase Grain Crop Primary Productivity? For comparison, a further analysis by McHenry (2012a,b) was undertaken of the value of applying biochar at 1 t/ha with the full rate of SSP as described above. This analysis was undertaken to explore the relative impact of using biochar to increase yield, as opposed to increasing fertilizer use efficiency. The analysis assumed that using the full rate of SSP (90 kg/ha year) with a 1 t/ha year application of biochar increased wheat yields by 15% on average over the 15 years relative to full SSP applications only, in the southwest of WA2. The baseline yield used for the scenario was 1.75 t/ha, an approximate average wheat yield for WA. The assumptions of the model, including a total area wheat return increase of $71.75/ha, were based on an increased production of an additional 15% wheat yield from the 1.75 t/ha at a constant value of $350/t over 1 The Australian GST (goods and services tax) is a value-added tax of 10%, paid only by the final consumer of a good or service. 2 This assumption requires basic research to verify, although some agronomic studies indicate that this may be possible for certain crops and soil types (Lehmann, J. and S. Joseph. 2008). Many individual studies are detailed in the book Biochar for Environmental Management: Science and Technology (2008) published by Earthscan.
ECONOMIC ANALYSIS the 15 years using the 8% real discount rate. The scenario did not include additional harvesting or transport costs for the additional wheat yield. The results indicated that the required carbon prices to recoup biochar purchase price costs were lower when biochar is used to increase yield, rather than reduce fertilizer use. When biochar purchase prices were below $250/t, the application of biochar was attractive without any carbon price, assuming the 15% yield is achieved (McHenry, 2012a). Therefore, these relatively simple analyses suggest that the most cost-effective on-farm use for biochar is to simply increase the wheat yield. The results confirm previous assertions that agricultural biomass production for the sole purpose of producing biochar for soil carbon sequestration may not be economically feasible (Lehmann et al., 2006). Can Biochars Increase Livestock Growth Rates, or Provide a New Market for Semiarid Forestry? It is now clear that forestry carbon offsets are resilient features of Australian climate change policies. To participate in such markets, farmers must be able to adequately measure, and verify the mitigation achieved (The CRC for Greenhouse Accounting & Tony Beck Consulting Services Pty Ltd, 2003). Forestry plantations that include some rotational harvesting for biochar or bioenergy will require more sophisticated carbon accounting than a simple revegetation project (Independent Pricing and Regulatory Tribunal, 2008). The establishment of tree fodder plantations has long offered a significant productivity option for some farmers (Sanford et al., 2003). Deferring the early grazing of annual pastures and reduce dry season hand-feeding has long generated interest (Patabendige et al., 1992; Cleugh et al., 2002), and perennial fodder tree plantations offer another source to supplement stock feed in the summer/autumn period (Sanford et al., 2003). Deep-rooted perennials are well known to use available water when annual pastures are dead, recover nutrients from deeper soils, reduce soil acidification, minimize erosion, and some leguminous species also fix nitrogen (Patabendige et al., 1992; Cransberg and McFarlane, 1994; Hatton and Nulsen, 1999; Wise and Cacho, 1999; Valzano et al., 2005). Adding value to these conventional applications in such regions is the use of tree woody wastes to produce biochar as a feed additive which may improve ruminant growth when fed on the trees (which may be of lower grade and/or be a “high tannin” content), and in the process sequester carbon in the soil (McHenry, 2010). The mechanism for this improvement is generally known as “detannification”, and may enable the use of potentially large resources of high-tannin fodder species (such as Acacia sp.) by increasing the availability of leaf protein (Van et al., 2006; Blackwell et al., 453 2009). Acacia sp. fodder plantations require annual pruning of the higher branches to provide fodder for grazing animals. Animals eat the leaves from the branches on the ground, leaving the inedible woody waste components in the paddock to dry and be collected as a potential source of biomass for biochar manufacture. The improved digestibility of some high-tannin fodder trees with biochar feed additives may expand their utility within agricultural production systems (McHenry, 2010). In particular, if an Acacia sp. biochar feed additive is effective in Australian semiarid production systems (such as the West Midlands), this might provide a further incentive to revegetate semiarid sandy soils suitable to many native Acacia sp. to attain a combination of positive benefits (Graetz and Skjemstad, 2003; Antle et al., 2007). These options are currently based on a 12-week experiment by Van et al. (2006) comparing goat growth rates fed on tannin-rich Acacia sp. fodder. The goats were either fed biochar (produced from bamboo) at a feed rate of <1 g per day per kilo of live weight, or no biochar for the control group. The experimental group exhibited notably higher growth rates (w20%) than the control goats that received no biochar feed additive on the same feed regime. Over the 12 weeks the experimental goats fed biochar weighed 5.2% heavier than their controls (Van et al., 2006). This may be a sufficient commercial incentive to drive demand and subsequent biomass conversion technology investment without a carbon price (McHenry, 2010). The work by Van et al. (2006) also presents a mechanism (via animal excreta) that may be assessed for efficacy when avoiding relatively expensive biochar soil application options such as deep banding, broadcasting, seeding application, topdressing, aerial delivery, or precision application to ailing plants (Blackwell et al., 2009). In addition to researching the efficacy of small biochar additions to the diet of grazing animals, the opportunity arises to simultaneously investigate the reported capacity and magnitude of numerous other biochar benefits (McHenry, 2010), including the ecologically delivered biochar to biosequester C; biologically immobilize inorganic N; retain soil N; increase soil pH; adsorb dissolved ammonium, nitrates, phosphate, as well as hydrophobic organic soil pollutants such as polycyclic aromatic hydrocarbons (Beaton et al., 1960; Gustafsson et al., 1997; Accardi-Dey and Gschwend, 2002; Lehmann et al., 2003). The remaining levels of biochar in the animal excreta would also determine the carbon fractions that survive the digestive system to determine the maximum available long-lived carbon species fractions to be sequestered in soils via the ecological delivery method (McHenry, 2010). Long-term soil testing may also be able to detect the stable fraction of the ecologically delivered biochar after being exposed to the soil environment. However, there is clearly
454 26. BIOCHAR PROCESSING FOR SUSTAINABLE DEVELOPMENT IN CURRENT AND FUTURE BIOENERGY RESEARCH much research required to verify a number of assertions and assumptions to provide a level of certainty acceptable to farmers and investors who collectively command much of Australia’s productive capacity (Intergovernmental Panel on Climate Change, 2000; Barker et al., 2007). A Comparison of Biochar Carbon Value for Different Potential Income Streams A simple analysis of potential value per unit of dry biomass associated with various potential production systems may help identify suitable uses of agricultural biomass (Table 26.8.) These theoretical comparative financial values are exclusive of costs, which are extremely variable according to the application and scale of operation. These basic scenarios seem to indicate that the highest values of agricultural residues are animal husbandry or cropping applicationsdonly if the biochar can increase conventional yields. This demonstrates that a key focus for the development of a sustainable biochar industry is the value of the product to an industry, rather than the cost of production per se. This also illuminates the aspects of supplying biochar with appropriate characteristics for the specific application, as it is likely that biochar applications will mature, TABLE 26.8 and standards for biochars will be sought by users assessing cost-effective product suppliers. In any case, it seems reasonable that small-scale waste-to-energy suppliers will be established at some point near rural settlements with the assistance of government subsidies in Australia. It also seems reasonable that various agricultural wastes will be co-fired, as well as potential adjustments installed to increase clean biochar production options. These projects can be a sound foundation to understand biomass-to-biochar technology by supplying sufficient volumes of relatively cheap and consistent biochars suitable for numerous medium- to large-scale research trials. It is further likely that bioenergy and biochar cogeneration at a regional level may be more cost-effective when agricultural wastes are leveraged by municipal solid waste resources, if quality control of municipal wastes is maintained. However, this will also require much evaluation and research for processing technology and downstream application suitability. CONCLUSION Taken in isolation, the cost and benefits of using biochar for only farm soil carbon sequestration may Comparative Values per Unit Wood Biomass for Different Mitigation and Sequestration Applications Product Product Value Value per Ton Dry Biomass Biomass Energy (Combustion) Mitigation $23 tCO2-e1 $32.20/t* Carbon Credit (Uncut Forestry) Sequestration $23 tCO2-e1 $42.16/ty Income Biochar Effect Biochar Used Grain Production 20% yield increase 0.3 t of biochar used on 0.3 ha at a rate of 1 t/ha Wheat yield at the rate of 2 t/ha $300/t (wheat), $23 tCO2-e1 $36 þ $17.71/tyy Detannification 20% increase in liveweight gain over 12 weeks relative to controls 0.75 g/kg liveweight Live goat growth at the rate of 10 kg/goat (liveweight) $1000/t liveweight, or $1/kg), $23 tCO2-e1 $359 þ $17.71/tx * 1 ton of wood with a 0.5 carbon dry fraction and a dry energy content of 21 MJ/kg can generate 1.75 MWh with a 30% conversion technology. If this technology was able to directly displace electricity from the South West Interconnected System, using the latest estimate for the scope 2 emission factor (0.80 kgCO2-e/kWh) this would mitigate up to 1.4 tCO2-e, or at $23 tCO2-e1 a gross mitigation value of up to $32.20. y If this same ton of wood remained in the paddock unharvested and was part of a carbon sequestration plantation, the theoretical sequestration of the carbon fraction would be 1.833 tCO2-e (500 kg C  3.666 tCO2-e t/C). At a price of $23 tCO2-e1, the gross value of the wood is now $42.16, or 25% more than the bioenergy option. yy In this scenario the ton of dry wood is assumed to be converted to biochar at an efficiency of 30%, with a stable carbon fraction of 0.7. Therefore, the 1 ton yields 300 kg of biochar containing 210 kg C or 0.770 tCO2-e (210 kg C  3.666 tCO2-e t/C). If the paddock yield for the wheat crop increases 20% above the 2 t/ha, in the 0.3 ha the biochar from the 1 ton of wood was applied to, the additional yield is 120 kg (2 t/ha  0.3 ha ¼ 0.600 t, 0.600 t  0.2 ¼ 0.120 t). Therefore, the additional 120 kg of wheat at $300/t is worth $36. The additional stable fraction of 0.770 tCO2-e at a carbon price of $23 tCO2-e1 is also theoretically worth $17.71. x In this scenario the ton of dry wood is also converted to biochar at an identical efficiency (30%), with a stable carbon fraction of 0.7. The 300 kg of biochar is fed to 10 kg (liveweight) goats as a detannification feed using the Van et al. (2006) methodology. The 300 kg would be sufficient for 40,000 daily doses for a goat weighing exactly 10 kg per head when given 0.75 g for each kg of liveweight each day. The 300 kg is the theoretical equivalent dose for 476 goats over the 12-week, or 84-day period (40,000/84). At a 20% growth increase relative to controls on the same diet of high tannin fodder of an assumed 9 g per day, this is an additional liveweight of 0.756 kg per animal over the interval, or 359 kg for the total gain of the 476 goats. Therefore, the 1 ton of wood (or the 300 kg of char) has a value of $359 when the value of goats is assumed to be $1 per kg of liveweight. In terms of carbon sequestration, the carbon fraction of 0.7 (assuming the digestion process does not influence the char), the additional $17.71.
455 REFERENCES not be a profitable activity. Yet, the net sum over the agricultural system in terms of biochars increasing conventional productivity may prove to be a more cost-effective option than existing operations in some areas (Antle et al., 2007). Notwithstanding economic issues, the greater scientific challenge is determining the efficacy of biochar carbon species in a range of specific agricultural production systems over both the long and the short term (McHenry, 2009, 2011). Integrated agricultural production systems require suitably high-resolution data to determine the agricultural systems and regions that may be able to implement options cost-effectively and sustainably (McHenry, 2010). Thus, a coordinated and cross-disciplinary research approach will likely be the most effective means of utilizing existing biomass/bioenergy activities for new agricultural applications (Nabuurs et al., 2007). Providing greater scientific rigor and certainty to farmers, environmentalists, governments and the broader community require undertaking biochar research alongside their impacts on upstream and downstream activities (McHenry, 2011). Once this research becomes available, it may provide a form of indemnity to farmers before prematurely applying new systems and technologies that may be only cost-effective in highly specific situations. Conversely, if biochar feed additives prove effective, even in localized regions, a major source of biochar will be required, and as Acacia sp. are native to Australia, and also most major continents, this may have extensive global implications in arid, semiarid, and even some temperate regions (McHenry, 2010). Nonetheless, complex biological and agroecological production systems require high-resolution information to determine where the best opportunities are to integrate these new diversification options into their existing production systems. In conclusion, the author offers a selection of key biochar-related knowledge deficiencies for Australian agriculture in bullet points,3 and also numbered suggestions for groups in the West Midlands of WA: • The sensitivity of biochar industry to policy change and administrative changes; • Development of biochar research that aims to create major benefits to agricultural productivity. 1. Proceed with caution 2. Understand carbon credit ownership in biomass provided to regional power stations 3. Test cropping benefits with affordable biochar 4. Using appropriate safety precautions, experiment with on-farm production and application of biochar on a small scale 5. Encourage research into effects of biochar on crops, animal nutrition, and animal health 6. Monitor technical developments of small scale (2e20 MW) gasifier power units 7. Consider relationships with local waste-to-energy projects using landfill • Key sensitivities of biochars in major West Australian agricultural operations (grains and livestock); • Key sensitivities of biochar carbon sequestration in major agricultural operations; • Energy, material, and cost flows of various biochar/ bioenergy conversion systems; • Major feedstock availability in different regions, costs, and transportation logistics; • Efficacy and cost of various biochar application technologies for West Australian conditions; References 3 DISCLAIMERS This material has been written for Western Australian conditions, and many conclusions do not imply suitability to other areas. The inclusion of biochar/bioenergy products or trade names do not imply recommendation, the comparisons are simply for a general audience, and are not sufficiently detailed for commercial comparisons or technical appropriateness for any one or range of applications. The omission of any locally available technology is unintentional. Acknowledgments This chapter would not have been written without the considerable experience and expertise of Dr Paul Blackwell, Department of Agriculture and Food, Western Australia’s (DAFWA) Geraldton Regional Office. Dr Blackwell’s long-time contribution to this field of research in WA under particularly limiting funding and time allocations is an example to those of us following in his notable footsteps. Accardi-Dey, A., Gschwend, P.M., 2002. Assessing the combined roles of natural organic matter and black carbon as sorbents in sediments. Environ. Sci. Technol. 36, 21e29. Antle, J.M., Stoorvogel, J.J., Valdivia, R.O., 2007. Assessing the economic impacts of agricultural carbon sequestration: terraces and agroforestry in the Peruvian Andes. Agric. Ecosyst. Environ. 122 (4), 435e445. Australian Farm Institute, 2011. The Impact of a Carbon Price on Australian Farm Businesses: Grain Production. Australian Farm Institute, Sydney, New South Wales, Australia. Barker, T., Bashmakov, I., Alharthi, A., Amann, M., Cifuentes, L., Drexhage, J., Duan, M., Edenhofer, O., Flannery, B., Grubb, M., Hoogwijk, M., Ibitoye, F.I., Jepma, C.J., Pizer, W.A., Yamaji, K., These requirements are in addition to the major research projects underway regarding the influence of various biochar feedstocks and conversion technologies on the characteristics of the final biochar product.
456 26. BIOCHAR PROCESSING FOR SUSTAINABLE DEVELOPMENT IN CURRENT AND FUTURE BIOENERGY RESEARCH 2007. mitigation from a cross-sectoral perspective. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. In: Climate Change 2007: Mitigation. Cambridge University Press, Cambridge, United Kingdom and New York, USA. Beaton, J.D., Peterson, G.A., Bauer, N., 1960. Some aspects of phosphate adsorption by charcoal. Soil Sci. Soc. Am. J. 24, 340e346. Blackwell, P., Krull, E.S., Butler, G., Herbert, A., Solaiman, Z., 2010. Effect of banded biochar on dryland wheat production and fertiliser use in south-western Australia: an agronomic and economic perspective. Aust. J. Soil Res. 48, 531e545. Blackwell, P., Reithmuller, G., Collins, M., 2009. Biochar application to soil. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science and Technology. Earthscan, London, United Kingdom, pp. 207e226. Bridle, T.R., 2004. Use of Pyrolysis to Recover Energy and Nutrients from Biosolids. Cited 25.07.08. Available from: http://www.wef. org/NR/rdonlyres/7DA581D9-C0D3-4E5C-B127AC68B7ABA6DD/0/Bridle_Paper.pdf. Brown, R.C., 2009. Biochar production technology. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science and Technology. Earthscan, London, United Kingdom, pp. 127e145. Cleugh, H.A., Prinsley, R., Bird, R.P., Brooks, S.J., Carberry, P.S., Crawford, M.C., Jackson, T.T., Meinke, H., Mylius, S.J., Nuberg, I.K., Sudmeyer, R.A., Wright, A.J., 2002. The Australian national windbreaks program: overview and summary of results. Aust. J. Exp. Agric. 42, 649e664. Cransberg, L., McFarlane, D.J., 1994. Can perennial pastures provide the basis for a sustainable farming system in southern Australia? N. Z. J. Agric. Res. 37, 287e294. Graetz, R.D., Skjemstad, J.O., 2003. The Charcoal Sink of Biomass Burning on the Australian Continent. CSIRO Atmospheric Research Technical Paper No. 64. CSIRO, Aspendale, Victoria, Australia. Gustafsson, O., Haghseta, F., Chan, C., Macfarlane, J., Gschwend, P.M., 1997. Quantification of the dilute sedimentary soot phase: implications for PAH speciation and bioavailability. Environ. Sci. Technol. 31, 203e209. Hatton, T.J., Nulsen, R.A., 1999. Towards achieving functional ecosystem mimicry with respect to water recycling in southern Australian agriculture. Agroforestry Syst. 45, 203e214. Independent Pricing and Regulatory Tribunal, 2008. Compliance and Operation of the NSW Greenhouse Gas Reduction Scheme 2007. The New South Wales Government & The Independent Pricing Regulatory Tribunal of New South Wales, Sydney, Australia. Intergovernmental Panel on Climate Change, 2000. Land Use, Land Use Change, and Forestry. Cambridge University Press, Cambridge, United Kingdom. pp. 25e45. Joseph, G., 2007. Combustible dusts: a serious industrial hazard. J. Hazard. Mater. 142, 589e591. Laird, D., Fleming, P., Wang, B., Horton, R., Karlen, D.L., 2008. Impact of soil biochar applications on nutrient leaching. In: The 2008 Joint Annual Meeting of Black Carbon in Soils and Sediments: III. Environmental Function Symposium, Houston, Texas, USA. Lehmann, J., 2007. Bio-energy in the black. Front. Ecol. Environ. 5 (7), 381e387. Lehmann, J., Gaunt, J., Rondon, M., 2006. Bio-char sequestration in terrestrial ecosystems. Mitigation Adapt. Strategies. Global Change 11, 315e419. Lehmann, J., Joseph, S., 2008. Biochar for Environmental Management: Science and Technology. Earthscan, London, United Kingdom. Lehmann, J., Joseph, S., 2009. Biochar systems. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental Management: Science and Technology. Earthscan, London, United Kingdom, pp. 147e168. Lehmann, J., Pereira da Silva Jr, J., Steiner, C., Nehls, T., Zech, W., Glaser, B., 2003. Nutrient availability and leaching in an archaeological Anthrosol and a Ferralsol of the Central Amazon basin: fertiliser, manure and charcoal amendments. Plant Soil 249, 343e357. Lehmann, J., Rondon, M., 2006. Bio-char Soil Management on Highly Weathered Soils in the Humid Tropics. CRC Press, Boca Raton, Florida, USA. McHenry, M.P., 2009. Agricultural bio-char production, renewable energy generation and farm carbon sequestration in WA: certainty, uncertainty & risk. Agric. Ecosyst. Environ. 129, 1e7. McHenry, M.P., 2010. Carbon-based stock feed additives: a research methodology that explores ecologically delivered C biosequestration, alongside live-weights, feed-use efficiency, soil nutrient retention, and perennial fodder plantations. J. Sci. Food Agr. 90, 183e187. McHenry, M.P., 2011. Soil organic carbon, biochar, and applicable research results for increasing farm productivity under Australian agricultural conditions. Commun. Soil Sci. Plan 42, 1187e1199. McHenry, M.P., 2012a. Sensitive variables for applying biochar as a fertiliser substitute and a method to sequester carbon in soils: a wheat crop scenario. In: Ryan, B.J., Anderson, D.E. (Eds.), Carbon Sequestration: Technology, Measurement Technologies and Environmental Effects. Nova Science Publishers, Hauppauge, New York, USA. McHenry, M.P., 2012b. Small-scale (6 kWe) stand-alone and gridconnected photovoltaic, wind, hydroelectric, biodiesel, and wood gasification system’s simulated technical, economic, and mitigation analyses for rural regions in Western Australia. Renewable Energy 38, 195e205. Mizuta, K., Matsumoto, T., Hatate, Y., Nishihara, K., Nakanishi, T., 2004. Removal of nitrate-nitrogen from drinking water using bamboo powder charcoal. Bioresour. Technol. 95, 255e257. Nabuurs, G.J., Masera, O., Andrasko, K., Benitez-Ponce, P., Boer, R., Dutschke, M., Elsiddig, E., Ford-Robertson, J., Frumhoff, P., Karjalainen, T., Krankina, O., Kurz, W.A., Matsumoto, M., Oyhantcabal, W., Ravindranath, N.H., Sanz Sanchez, M.J., Zhang, X., 2007. Forestry, Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, USA. Patabendige, D.M., Scott, P.R., Lefroy, E.C., 1992. Fodder Trees and Shrubs for High Rainfall Areas of South Western Australia. Department of Agriculture Western Australia, Perth, Western Australia. Rondon, M., Ramirez, J.A., Lehmann, J., 2005. Greenhouse gas emissions decrease with charcoal additions to soils. The Third USDA Symposium on Carbon Sequestration. Baltimore, USA. Sanford, P., Wang, X., Greathead, K.D., Gladman, J.H., Speijers, J., 2003. Impact of Tasmanian blue gum belts and kikuyu-based pasture on sheep production and groundwater recharge in south-western Western Australia. Aust. J. Exp. Agric. 43 (8), 755e767. The CRC for Greenhouse Accounting & Tony Beck Consulting Services Pty Ltd, 2003. Opportunities for the Western Australian Land Management Sector Arising from Greenhouse Gas Abatement. Western Australian State Government, Perth, Western Australia. Valzano, F., Murphy, B., Koen, T., 2005. The Impact of Tillage on Changes in Soil Carbon Density with Special Emphasis on Australian Conditions. Report No. 43. National Carbon Accounting System, Australian Greenhouse Office, Canberra, Australia. Van, D.T.T., Mui, N.T., Ledin, I., 2006. Effect of method of processing foliage of Acacia mangium and inclusion of bamboo charcoal in the diet on performance of growing goats. Anim. Feed Sci. Tech. 130, 242e256. Wise, R., Cacho, O., 1999. A Bioeconomic Analysis of Soil Carbon Sequestration in Agroforests. Cited 16.07.12. Available from: http://www.une.edu.au/carbon/CC02.PDF.
C H A P T E R 27 Development of Thermochemical and Biochemical Technologies for Biorefineries Michael P. Garver, Shijie Liu* Department of Paper and Bioprocess Engineering, College of Environmental Science and Forestry, State University of New York, Syracuse, NY, USA *Corresponding author email: sliu@esf.edu O U T L I N E Introduction 457 Characteristics of Lignocellulosic Biomass 458 An Overview on Biomass Conversion 461 PretreatmentdBiomass Size Reduction by Physical or Mechanical Methods Mechanical PretreatmentdChipping, Grinding, Milling, Refining Irradiation Pretreatment by Electron Beam, Gamma Ray, or Microwave Ammonia Recycle Percolation Pretreatment Ozonolysis Pretreatment Organosolv Pretreatment Oxidation Pretreatment Ionic Liquid Pretreatment Sulfite Pretreatment to Overcome Recalcitrance of Lignocelluloses Hot Water 462 470 473 473 474 474 475 Hydrolysis 476 BioconversiondConverting Sugars to Products 477 465 465 465 466 466 467 Thermochemical Conversion Combustion Gasification Pyrolysis Direct Liquefaction 478 478 478 481 481 Conclusion 482 468 469 References 482 463 INTRODUCTION A biorefinery is a complex industrial system to convert raw biologically derived materials into usable and valuable products. The actual design of a biorefinery depends on the desired product, the raw materials available, and the method of conversion desired. For the purposes of this chapter, the raw material considered is woody biomass or more generally, Bioenergy Research: Advances and Applications http://dx.doi.org/10.1016/B978-0-444-59561-4.00027-9 Steam Explosion Ammonia Fiber Explosion Supercritical Carbon Dioxide Explosion Biological Pretreatment Acid Hydrolysis Alkaline Hydrolysis lignocellulosic biomass (LB). LB may originate from forest, herbaceous plants or organic waste streams such as sewage, food processing waste, or animal manure. LB is a source of energy that can reduce the consumption of fossil fuels. Energy independence is an important economic and political goal. Renewable sources of energy are also critical for a balanced ecological policy. Biorefineries may be designed to output a specific set of products and by-products. These products include 457 Copyright Ó 2014 Elsevier B.V. All rights reserved.
458 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES biofuels, adhesives, surfactants, biochemicals, biopolymers, food and medicine. This chapter will focus on some common products such as acetic acid, ethanol, butanol, acetone, hydrogen, and polyhydroxyalkanoates. These products stem from the fermentation of sugars derived from LB or they may be derived from thermochemical conversion processes. The first objective in any conversion is to reduce the size and increase the surface area of the raw material. This enables subsequent treatment methods to attack and exploit specific properties of LB more effectively to obtain sugars for bioconversion or obtain products from thermochemical conversion. Secondary treatment or conversion methodologies include some form of hydrolysis, fermentation or any of a variety of thermochemical conversion treatments. The objective of these methods is to break lignin and complex carbohydrates into either simple sugars or intermediate products or even down to CO and H2 (syngas) for further fermentation (bioconversion) or thermochemical conversion. Fermentation is usually followed by separations or filtrations as final steps in the acquisition of a desired product in a bioconversion. CHARACTERISTICS OF LIGNOCELLULOSIC BIOMASS Understanding the characteristics of LB is necessary for the effective application of a conversion technology or a pretreatment method. LB is the most abundant organic and renewable resource on the planet (Klass, 1998). Man has been producing chemicals, materials, and energy from LB since his origin. These activities continue to be the most promising activities to pursue in order to address our contemporary challenges. For (a) example, we currently look to use LB to reduce dependence on fossil fuels. There are three families of LBs: grassy plants, shrubs, and trees, each possess four primary components: cellulose, hemicellulose, lignin and extractives. Each species possesses these components in different proportions. Hardwood trees (angiosperms), shrubs and grassy plants (graminoids) usually possess less lignin than softwood trees (gymnosperms) (Liu, 2012). Figure 27.1 illustrates the variation in cellular structure between hardwood and softwood. Notice that in both images the structure is porous, where the pores are empty spaces and run longitudinally. Figure 27.2 is a compilation of sketches of wood cells. The cellular structure of each plant species is of different sizes and shapes and varies in the size and number of pores. From Figure 27.1, one can note that there is no space between the cells. Regardless of size and shape, each cell is glued tightly to its neighbors. The intercellular spaces are called middle lamellae. The great majority of the middle lamellae, over 80%, contain lignin (Liu, 2012). Lignin is the glue that binds the cells together and provides the rigid structure of wood. The remaining volume of the middle lamellae consists of hemicellulose and extractives. Conversely, the cell wall is mostly made of cellulose and to a much lesser degree contains lignin and hemicellulose. The great majority of biomass dry weight is derived from the cell wall. A much smaller portion of the biomass dry weight comes from the middle lamellae. Most of the total lignin content of LB comes from the cell wall. Despite the high concentration of lignin in the middle lamellae, over 60% of the lignin from LB comes from the cell wall portion of the material (Liu, 2012). Table 27.1 shows that cellulose is the largest portion of LB. Cellulose, which represents between 40% and 50% of the dry weight of wood, is a homopolymer of (b) Softwood FIGURE 27.1 Hardwood Typical structures of wood (a) Softwood (b) Hardwood. Source: Liu, 2012.
459 CHARACTERISTICS OF LIGNOCELLULOSIC BIOMASS Southern yellow pine Western hemlock Ray tracheid Softwood Redgum Parenchyma cells Hardwood 200 µm Birch Oak early wood Redgum Aspen Alder Eucalyptus Gmelina Oak tracheid Softwood fibers Tracheid Fiber Hardwood fibers Birch Aspen Oak late wood Hardwood vessel elements FIGURE 27.2 Diagrams of major cell types in softwood and hardwood. All the diagrams are shown at the same magnification to illustrate the relative sizes of these elements. Source: Parham, 1983. b-D-glucopyranose where dehydration of the b-Dglucose units forms a linear chain with a degree of polymerization (DP) between a few hundred and several thousand b-D-glycosidic bonds. The dehydration occurs between the one and four carbons of b-D-glucopyranose units and leaves an oxygen atom to join the two units, which is written as, b-1-O-4 glycosidic bonds. The formula for cellulose is He(C6H10O5)neOH, where “n” represents the DP. This highly ordered, tightly bound pattern is made of bonds that are quite strong and are difficult to break. Cellulose grows into microfibrils with crystalline and amorphous regions. The crystalline portions of the molecule line up side by side. Hydrogen bonds, between the hydroxyl groups, provide strong, sturdy and stable links between and within these crystalline units. When these microfibrils form macrofibrils and interact with noncellulosic material in the cell walls of plants, the result is strength and rigidity. While the crystalline regions are stable and strong, the amorphous regions provide an opportunity to break down the large structure into smaller saccharides. Solvents, reagents and enzymes may be used to penetrate and hydrolyze the structure. Hydration requires the addition of energy or a strong acid. Alternatively, enzymes, such as cellulase, may facilitate the conversion. Enzymatic hydrolysis tends to be much slower than acid hydrolysis. Reducing the chip size or increasing the exposed surface area of LB increases the effectiveness of these solvents, reagents and enzymes. Hemicelluloses compose another large portion of LB, between 20% and 30% of the dry weight of wood, see Table 27.1. These are heteropolymers, or heterosaccharides of five- and six-carbon sugars. They are found mostly in the cell walls of LB. Common hemicellulose sugars are D-glucose, D-mannose, D-galactose, D-xylose, L-arabinose, and to a lesser degree, L-rhamnose. Hemicellulose has a low DP, around 100e200, and thus is more easily hydrolyzed into their monomeric sugar components (Glaudemans and Timell, 1958; Goring and Timell, 1960; Koshijima et al., 1965; Timell, 1960). The structure of a hemicellulose tends to possess a primary backbone, off of which might hang a variety of residual units. These residual units are nonpolymeric acids and sugars. The degree of branching or number of residual units depends on the origin or species of the biomass. For hardwood, the backbone is xylan, containing b-linked bonds at carbons one and four, like cellulose. Unlike cellulose, residual units can hang off from the other carbon positions. These residues may include those of acetic acid, glucuronic acid, mannose, arabinose and galactose. Softwood is even more variable in that the backbone may be made of more diverse materials. The backbone is typically made of galactoglucomannan units or arabinoglucuronoxylan units. Galactoglucomannan is a polymer that is a primarily
460 TABLE 27.1 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES Major Components of Wood Distribution, wt% Type Cellulose OH O O HO OH OH HO O O O OH OH HO O OH HO O Softwoods Hardwoods 40e50 40e50 20e30 20e25 OH O OH OH Lignin Phenolic eOH Per C6C3 unit 0.2e0.3 0.1e0.2 Aliphatic eOH Per C6C3 unit 1.15e1.2 1.1e1.15 Methoxyl eOCH3 Per C6C3 unit 0.9e0.95 1.4e1.6 Carbonyl >CaO Per C6C3 unit 0.2 0.15 Hemicellulose 25e30 25e35 Galactoglucomannan (1:1:3) 5e8 0 (Galacto)glucomannan (0.1:1:4) 10e15 0 0 2e5 7e10 Trace Trace 15e30 5e8 2e4 The OH groups in the xylose units were partially substituted by OAc on the C-2 or C-3 positions, i.e. RaCH3CO (Ac) or H Glucomannan (1:2 to 1:1) OH OH HO O O O O HO OH OH O OH HO O OH O O HO OH O OH Arabinoglucuronoxylan Glucuronoxylan OH in the xylo-units were partially substituted by OAc on the C-2 or C-3 positions (about 7 in 10 xylo-units), i.e. RaCH3CO (Ac) or Extractives Aliphatic and alicyclic Terpenes, terpenoids, esters, fatty acids, alcohols, etc. Phenolics Phenols: p-cresol, p-ethylphenol, guaiacol, salicyl alcohol, eugenol, vanillin, coniferyl aldehyde, acetovanillone, propioguaiacone, salicylic acid, ferulic acid, syringaldehyde, sinapaldehyde, and syringic acid; stilbenes: pinosylvin, pinosylvin monomethyl and dimethyl ethers, 4-hydroxystilbene, 4-hydroxystilbene monomethyl ether; lignans; hydrolyzable and condensed tannins; flavonoids; isoflavones or isoflavonoids Carbohydrates Arabinose, galactose, glucose, xylose, raffinose, starch, pectic material Inorganics  Ca, K, Mg, Na, Fe, SO 2 4 , Cl , etc. Others Cyclitols; tropolones; amino acids, protein, alkaloids, etc. Ash Source: Fengel, 1989. 0.2e0.5 0.2e0.8
AN OVERVIEW ON BIOMASS CONVERSION linear and perhaps mildly branched chain. In hemicellulose, the residual units take the place of the strong hydrogen bonding that occurred with cellulose components. Recall that cellulose is highly ordered and tightly bound and thus resistant to hydrolysis. Hemicellulose is not. Hemicellulose tends to be more randomly organized with a more variable and loosely bound structure (amorphous). Therefore, it can be hydrolyzed by weaker or more dilute acids and bases, or at milder conditions. Lignin is the third largest component of LB at 25e35% of the LB dry weight (Boerjan et al., 2003). Lignin is a heteropolymer with methoxylated phenylpropylene alcohol units. Its structure tends to be amorphous and variable. These units are interconnected by stable ether and ester linkages. It is hydrophobic and aromatic. It covalently links to hemicelluloses and cross-links different plant polysaccharides giving mechanical strength to the cell wall (Mielenz, 2001). Additionally, lignin is highly resistant to biological degradation and thus it protects cellulose and hemicellulose from decay. Lignin from different plant families vary in their alcohol content and composition. These lignins are thus defined by these components into different types. The lignin precursor in gymnosperms is coniferyl alcohol. The precursor in angiosperms is p-coumaryl alcohol and sinapyl alcohol. The corresponding lignins are guaiacyl (G), p-hydroxyphenyl (H) and syringal (S), respectively. Grasses tend to contain G while palm trees contain mostly S (Sjöstrom, 1993). The next largest component of LB is the extractives. These make up between 2% and 8% of the total dry weight (Table 27.1). Extractives are compounds found in LB that are soluble in neutral organic solvents or water at standard temperatures and atmospheric conditions. Extractives vary in solubility. Some are lipophilic and others are hydrophilic. Lipophilic extractives that are soluble in nonpolar organic solvents are called resins. There is a large diversity in the number of extractives. Additionally, the concentrations of extractives are highly variable throughout the plant depending on the tissue type, i.e. root, stem, bark, branch, needle or leaf. It is important to note that over 70 metal, earth elements, and inorganic compounds may be found in LB. The extractives are the first components that can be extracted from wood. This is advantageous for using LB as a bioremediation for toxic soil and wastewater in addition to being a source for biofuel and other products. AN OVERVIEW ON BIOMASS CONVERSION Conversion refers to the collection of processes employed to modify a feedstock into desired product(s). 461 Given that LB is composed of a number of distinct components, there are a variety of treatment options available that one can use to change these components into fuel, chemicals and other products. With a harsh condition (high temperature, strong acid/base, strong solvents, or a combination of these agents), LB can be turned into small molecular units (such as C, CO2, CO, H2 and H2O) and then further converted to a desired product. Thermochemical conversion technologies usually employ this strategy to break down LB unselectively to accommodate further conversions either catalytically or biologically. Therefore, thermochemical conversion technologies can be versatile. The structure of wood (or LB in general) is sufficiently strong and complex that it is not feasible to attack the whole complex at mild conditions in a single step, nor is it feasible to isolate the components and attack them individually. When mild conditions are desired, one must attack at least one portion of the whole structure and weaken it. Follow up with another treatment to break down the first component or attack a second component. Continue to treat the biomass until the desired composition is obtained. These treatment options are classified into mechanical, thermal, chemical, or biological processes. These are not discrete classifications. In other words, a process can be considered to belong to more than one category. For example, if one were to saturate LB with water, then heat it under high pressure and rapidly release the pressure, the hot water could vaporize into steam and thus explode apart the woody cells it had penetrated. This is called steam explosion and it uses a thermal process to accomplish a mechanical breakdown of the woody material. Steam explosion will be discussed later in this chapter. While there are several LB conversion technologies available, this chapter will focus on biochemical conversion technologies with some discussions in thermochemical conversions. These treatments may be applied at various points across the process. The typical process to acquire fuel products using a bioconversion methodology is generally described in four parts: pretreatment, hydrolysis, fermentation and distillation, separation and filtration. As discussed previously, cellulose, hemicellulose and lignin are strong, stable structures. These structures are challenging for one to convert into fermentable components (Mielenz, 2001). Of these three components, hemicellulose is the most vulnerable and easiest to degrade. Recall that compared to cellulose, hemicellulose is a lower molecular weight and is less uniform as it is composed of a variety of sugar polymers and residual units. In bioconversions, the objective of pretreatment is to, as efficiently as possible, prepare LB for fermentation into products. The amount of energy required to break
462 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES FIGURE 27.3 Schemes of biochemical conversion to materials, chemicals and fuels: (1) sequential incremental deconstruction; (2) two-step saccharification and fermentation; (3) simultaneous saccharification and fermentation; (4) gasification and fermentation. Source: Wang et al., 2012. (For color version of this figure, the reader is referred to the online version of this book.) down LB is fixed. No matter what suite of treatment options used to convert LB to product, the thermodynamic barrier is the same. It requires the same amount of energy to completely biologically degrade wood as it does to chemically treat it or gasify it. One trades off time for allowing organisms to invest energy on one’s behalf versus applying heat or concentrated chemicals to accomplish the same task more quickly. Additionally, biological methods allow for greater selection of the portion of biomass to convert it into product, and usually by selecting a descending route of molecular chemical energy or intermediates that are not down to simplistic building blocks if possible (green chemistry). By selecting only a portion of the LB to convert, one lowers the amount of investment energy required. Biochemical processes operate at moderate or low temperatures. These milder conditions may be slightly more efficient than their thermochemical counterparts. However, a burden of biological or biochemical processes could arrive for the need of detoxification. One must often remove toxic components resulting from the pretreatment methods employed. Figure 27.3 illustrates a set of four treatment pathways to convert LB into various products. These are not the only methods available but merely an example of commonly used methods. This pathway represents one of the most popular biorefinery designs used to biochemically convert LB into biofuels and bioproducts. Pathway 4 shows a gasification process to produce syngas. This is a thermochemical process. The sugars in syngas are subsequently fermented into liquid fuels similar to those produced by the more biochemical methods. The four pathways shown vary in the number of steps, or time, required to acquire the product. Figure 27.4 provides a slightly more detailed look at pathway 1 from Figure 27.3. Different pretreatment methods have different desired characteristics (Limayem and Ricke, 2012). A summary of pretreatment methods to be discussed in this chapter and their characteristics is shown in Table 27.2. PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS The first and most important step in any conversion process is to reduce the physical size of LB. In order to obtain the high yields required for commercial success in bioconversion operations, it is vital to pretreat and reduce the biomass into an effective size (Mosier et al., 2005). Reducing the LB size from a log to wood chips to even fine powders improves mass and heat transfer as well as increases the surface area of the particle. Increasing the surface area exposes a higher percentage of the glycosidic or ester bonds to the agents in solution
463 PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS Lignocellulosic biomass Size reduction Hemicelluloses extraction (Pretreatment) Hemicelluloses extracts Hydrolysis Detoxification & neutralization Residual biomass Enzymatic hydrolysis Fermentation of hemicellulosic sugars Product recovery Fermentation of cellulosic sugars Ethanol butanol biopolymer ... Residue processing Co-products FIGURE 27.4 Schematic flow sheet for biomass conversion to bioproducts. Source: US DOE, 2006. (For color version of this figure, the reader is referred to the online version of this book.) (Mosier et al., 2005). Catalysts, such as a proton or an enzyme, can only access active chemical bonds when exposed at the solideliquid interface (Liu, 2003; Yang and Liu, 2005). Smaller particles translate into faster, more uniform reactions and a more complete conversion. The energy required to reduce the biomass into a treatable size depends on the density of the biomass source. Herbaceous materials do not require as much processing to achieve the needed particle size as it does to reduce wood (Cadoche and Lopez, 1989). Since LB reduction is much more energy intensive, it is imperative to adequately define the reduction process. This requires an understanding of the quality and condition of the source materials. Qualities such as moisture content, soil particles, foreign matter, and initial cut length will impact efficiency, energy requirements and downstream treatment conditions and requirements. Pretreatment is costly and greatly influences the cost and effectiveness of downstream operations. It affects fermentation toxicity, the rate of enzymatic hydrolysis, enzyme load, powder mix, product concentration, product purity, waste treatment requirements, energy requirements and a host of other process variables (Zhu and Pan, 2010). Thus, it is important to begin the design process with the end in mind. Much effort should be invested to design the whole process up front with specific source materials and conditions defined. Mechanical PretreatmentdChipping, Grinding, Milling, Refining At the time of harvest, an operation is performed in the field to presize the LB. Herbaceous biomass is prepared by shredding or forage cutting. Chipping is the preferred method for reducing the size of wood. Chipping reduces wood to 10e50 mm in two dimensions and 5e15 mm in the third (Zhu and Pan, 2010). This is the minimum treatment necessary to begin conversion. However, additional reduction is often performed. For example, wood chips may subsequently be refined to fibers such as that in fiber production, pulverized into wood fibers or wood flour (Zhu and Pan, 2010). Pulverization requires much more energy than chipping (Zhu and Pan, 2010). In addition to chipping and shredding, hammer milling, knife milling, disk or attrition milling, and ball milling are viable alternatives to reduce biomass sizes. Large-scale reduction operations have favored hammer and disk milling (Tienvieri et al., 1999). Chip refining is also an alternative as it can have a large throughput. Hammer milling is primarily used for making wood flours for composites and pellets. Disk milling is used for wood fiber production at a commercial scale, around 1000 tons per day. Disk milling operations are dependent on environmental conditions and the quality of source materials. The energy requirement and the wood particle size and shape depend on these operational parameters (Tienvieri et al., 1999). Milling operations have a significant impact on downstream energy requirements and the efficiency of enzymatic cellulose saccharification. Since the goal of a biorefinery is to optimize the conversion process, to reduce energy requirement and maximize the enzymatic cellulose saccharification, it is important to attend to the biomass size reduction portion of the process. Failure at this stage amplifies the cost of energy requirements and reduces the effectiveness of subsequent treatments. Since these mechanical processes can produce a range of particle sizes it is often necessary to control the
464 TABLE 27.2 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES Summary of Pretreatment Methods and Key Characteristics Pretreatments Key Characteristics References Dilute Acid (H2SO4, HCl (0.5e5%) • Practical and simple technique. Does not require thermal energy • Effective hydrolysis of hemicelluloses with high sugar yield • Generates toxic inhibitors • Requires recovery steps (Chandel et al., 2007; Chaudhary et al., 2012; Gamez et al., 2004; Li et al., 2012; Lloyd and Wyman, 2005; Schell et al., 2003; Um and van Walsum, 2012; Wyman et al., 2005; Yasuda and Murase, 1995) Hot Water • • • • (Banerjee et al., 2009; Hu et al., 2008; Kemppainen et al., 2012; Ladisch et al., 1998; Laser et al., 2002; Lynd et al., 2002; Mosier et al., 2005; Shupe and Liu, 2009; Weil et al., 1994) Lime • High total sugar yield including pentose and hexose sugars • Effective against hardwood and agricultural residues • High pressure and temperature hinder chemical operation • Commercial scalability problem (Kim and Holtzapple, 2006; Weil et al., 1994; Zhu and Pan, 2010) Ammonia Fiber Explosion (AFEX) • Effective against agricultural residues mainly corn stover without formation of toxic end products • Not suitable for high-lignin materials • Ammonia recovery • No wastewaters (Bisaria and Ghose, 1981; Dale et al., 1984; Hendriks and Zeeman, 2009; Jin et al., 2012; Lau and Dale, 2008; Speers and Reguera, 2012; Sun and Cheng, 2002) Ammonia Recycle Percolation (ARP) • High redistribution of lignin (85%) • Recycling ammonia • Theoretical yield is attained (Drapcho et al., 2008; Gupta and Lee, 2009; Kim and Dale, 2005) Steam Explosion with Catalyst • Effective against agricultural residues and hardwood • High hemicelluloses fraction removal • Not really effective with softwood (Bisaria and Ghose, 1981; Bura et al., 2009; Galbe and Zacchi, 2002; Kemppainen et al., 2012; Lloyd and Wyman, 2005; Monavari et al., 2009; Park et al., 2012) Organosolv • • • • • • (Cybulska et al., 2012; Koo et al., 2012; Monavari et al., 2009; Pan et al., 2005) Sulfite Pretreatment to Overcome Recalcitrance (SPORL) • Effective against high-lignin materials, both softwood and hardwood • Highest pretreatment energy efficiency • Minimum of inhibitors formation • Accommodate feedstocks versatility • Steam explosion combined to SPORL in presence of catalyst becomes effective against softwood materials • Cost-effective (Li et al., 2012; Shuai et al., 2010; Tian et al., 2011; Wang et al., 2009; Zhu et al., 2010a,c, 2009; Zhu and Pan, 2010) Ozone • Effectively remove lignin from a wide range of cellulosic material without generating inhibitors • Expensive (Garcı́a-Cubero et al., 2009; Mvula et al., 2009; Sun and Cheng, 2002) Alkaline Wet Oxidation • The combination of oxygen, water, high temperature and alkali reduce toxic inhibitors • High delignification and solubilization of cellulosic material • Low hydrolysis of oligomers (Chaudhary et al., 2012; Klinke et al., 2004; Monavari et al., 2009) Fungal Bioconversion • Environmentally friendly • Low use of energy and chemical • Slow bioconversion (Dashtban et al., 2010; Nguyen et al., 2000) The majority of hemicelluloses can be dissolved No chemicals and toxic inhibitors Average solid load Not successful with softwood High yield is enhanced by acid combination Effective against both hardwood and softwood Low hemicellulosic sugar concentration Formation of toxic inhibitors Organic solvent requires recycling High capital investment
PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS particle size used in the biorefinery. Size characterization is accomplished using sieves, screens and imaging analysis. The particle surface area is the most relevant determination of effectiveness, and thus, it is the quality to be controlled. Specific surface area correlates to energy consumption and the efficiencies of a variety of size reduction processes have been compared (Holtzapple et al., 1989). There is a limit to the effectiveness of size reduction. At this point, additional surface area increases, or particle size reductions will not improve substrate enzymatic digestibility. This critical size is proportional to the pore size in and along the wood cells. Refer to Figures 27.1 and 27.2. A common target size is one that maintains the cell structure while allowing for lignin removal from the middle lamellae. Size reduction below the cell size will provide a more efficient conversion. To reduce particle size to this smaller level is done by comminution (Vidal et al., 2011). Comminution of biomass, especially at the final sizing stage, is energy intensive and the product is of low value. Thus, there is much interest in finding the most efficient milling processes. To that end, ball milling has been extensively studied. It has been shown to deliver excellent results in terms of the hydrolysis rate and sugar yield. Additionally, this pretreatment method is clean and easy to do. Vibratory ball milling has been shown to be more effective at breaking down the crystallinity of cellulose and improving the digestibility of the biomass over ball milling alone (Millet et al., 1976). Mechanical milling requires long operation times and a large amount of energy (Lynd et al., 1996). The smaller the desired particle size the greater the comminution requirements will be in terms of time and energy (Cadoche and Lopez, 1989). Irradiation Pretreatment by Electron Beam, Gamma Ray, or Microwave Irradiation is an option for biomass size reduction. Following a gross procedure to reduce field supplies into at least chip size, one can employ high-energy radiation such as gamma radiation and/or microwave radiation to accomplish fine particle reduction (Wasikiewicz et al., 2005). Not only does a high-energy radiation treatment produce fine particles, but it can also favorably alter the physical and chemical properties of the biomass, depending on dosage (Bouchard et al., 2006). Irradiation has been shown to decrease the DP (Bouchard et al., 2006) and make microstructural changes to the irradiated cellulose pulp (Dubey et al., 2004). These changes include an increase in the carbonyl contents and an overall improvement in the vulnerability of the cellulose crystalline regions to reagents (Stepanik et al., 1998). This in turn leads to a higher 465 rate of enzymatic hydrolysis. Furthermore, irradiation leads to a significant increase in sugar yield (Yang et al., 2008). An electron beam cuts biopolymers such as cellulose, hemicellulose and lignin into smaller chains. Analysis by powder X-ray diffractometer and Fourier transform infrared spectroscopy confirm the electron beam treatments reduce the degree of crystallinity and improve the sugar yields from enzymatic hydrolysis from treated samples (Karthika et al., 2012). Electron beam irradiation is preferred over irradiation using a radioisotope. First of all, electron beam is safer. Turn off the power and the electron beam stops. A radioisotope is continuous and thus requires significant safety precautions to handle and dispose of. Furthermore, dosages delivered by high-energy electron beam can be controlled and they can provide more power per dose. This is a feature that would be useful in the continuous treatment of LB (Auslender et al., 2002). Compared with microwave and gamma ray treatments, treatment by electron beam is more energy effective. The larger particle sizes, that it can treat, significantly offset the negative effect of higher dosage. That said, there still remains the challenge of the limitation of electron beam source and the potential limitation on the scale of operations. Even though all these irradiation options reduce the particle size and reduce the DP, they are too expensive to use in full-scale operations. Currently, the prospects of engaging an irradiation treatment, even if in conjunction with other environmentally friendly treatment options, does not look promising due to the excessive energy requirements. Table 27.3 shows a comparison of these irradiation treatment methods on a variety of wood species. Ammonia Recycle Percolation Pretreatment When aqueous ammonia, 10e15 wt%, is percolated through biomass at temperatures between 150 and 170  C with a fluid velocity of 1 cm/min and a residence time of 14 min, lignin depolymerizes and the lignincarbohydrate linkages break. This process is known as ammonia recycle percolation (ARP) (Iyer et al., 1996). This process is advantageous in that it does not inhibit downstream biological processes. A water wash is therefore not necessary (Kumar et al., 2009a). Additionally, it is possible to recover and recycle the ammonia. On the downside, ARP is inefficient when used to pretreat softwood pulp (Mosier et al., 2005) where lignin had already been removed. Ozonolysis Pretreatment A pretreatment option that is appropriate for grassy biomass and some softwood is ozonolysis.
466 TABLE 27.3 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES The Effects of Cellulose Conversion by Different Irradiation Pretreatments Wood Species Maximum Dose Electron Beam Irradiation (Khan et al., 1986) Spruce 2 MGy 1e2 mm in thickness and 10e20 cm2 Gamma Irradiation (Betiku et al., 2009) Softwood (Triplochiton scleroxylon) 40 kGy 32e42 mesh 8 0.673  0.10 Gamma Irradiation (Betiku et al., 2009) Hardwood (Khaya senegalensis) 90 kGy 32e42 mesh 8 0.803  0.10 Microwave (Verma et al., 2011) Beech 400 W 30e42 mesh 48 Particle Size Enzymatic Hydrolysis Duration (h) Cellulose Conversion 72 0.89 0.561 Source: Wang and Liu, 2012. In this case, ozone is used to degrade the lignin and hemicellulose in bagasse, green hay, peanut, pine, cotton straw, wheat straw and poplar sawdust (BenGhedalia and Miron, 1981; Vidal and Molinier, 1988). The process primarily acts on the lignin component and only mildly affects the hemicellulose component, while cellulose is negligibly affected. Ozonolysis is notable in that it removes lignin effectively and the reactions take place at room temperature and standard pressure (Ben-Ghedalia and Miron, 1981). The most significant advantage is that following an ozone pretreatment where 60% of the lignin is removed from wheat straw, the rate of enzymatic hydrolysis increases by 500% (Ben-Ghedalia and Miron, 1981). The most notable drawback is that the process is expensive due to the large volume of ozone required (Sun and Cheng, 2002). Raising the operational temperature above 185  C eliminates the need for a catalyst, either for an inorganic acid or for an organic acid. At this condition, the amount of delignification is quite satisfactory. Adding acid yields a high quantity of xylose. Since the organic solvents inhibit downstream biological processes, such as organism growth and enzymatic hydrolysis, it is necessary to remove these solvents from the system. This is quite difficult as some quantity of solvent is likely to reside in the system even after efforts to remove them. Organic solvents tend to evaporate into the atmosphere and are hazardous to the environment and one’s health. Containing the solvent is another challenge. Given these challenges, an organosolv pretreatment is not necessarily ideal for large-scale or commercial operations. Oxidation Pretreatment Organosolv Pretreatment Take an organic or aqueous organic solvent such as formic acid, acetic acid, methanol, ethanol, acetone, ethylene glycol, oxalic acid, triethylene glycol or tetrahydrofurfuryl alcohol and combine it with an inorganic acid catalyst such as hydrochloric acid or sulfuric acid and one can eliminate the internal lignin and hemicellulose bonds. This is known as an organosolv process (Pan et al., 2006; Sarkanen, 1980; Thring et al., 1990). Alternatively, an organic acid, such as oxalic acid, acetylsalicylic acid and salicylic acid may be substituted for the inorganic catalyst. It has been observed that approximately 72% of xylose in untreated wood, in both its monomeric and oligomeric forms, could be recovered using an organosolv pretreatment process (Pan et al., 2006). Pan et al. (2006) also investigated a bioconversion of hybrid poplar to ethanol at 180  C, for 60 min, with 1.25% H2SO4, and 60% ethanol. They observed that nearly 74% of the lignin was removed as a precipitate in the ethanol extraction. Oxidation is a pretreatment option whereby an oxidizing agent, such as hydrogen peroxide or peracetic acid, is applied to LB. The result is the removal of hemicellulose and lignin and thus, an increased accessibility to cellulose to enzymatic hydrolysis. This result is the culmination of several reactions: electrophilic substitution, displacement of side chains, cleavage of alkyl aryl ether linkages, or the oxidative cleavage of aromatic nuclei (Hon and Shiraishi, 2001). Often the oxidative agent is not selective and a significant loss of hemicellulose and cellulose may occur. Additionally, there is a high risk of forming downstream inhibitors as soluble aromatic compounds are formed while the lignin oxidizes. When using hydrogen peroxide as the oxidative agent on sugarcane bagasse, the rate of enzymatic hydrolysis improves. In one study, hemicelluloses and approximately 50% lignin were solubilized by 2% hydrogen peroxide at 30  C over 8 h. This was followed by enzymatic hydrolysis, or saccharification, using cellulase at 45  C within 24 h. The result was 95% efficiency in glucose production (Azzam, 1989).
PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS In another study, peracetic acid was applied at ambient temperatures to a hybrid poplar and sugarcane bagasse mixture (Teixeira et al., 2000). It was determined that peracetic acid was very selective for lignin and in some cases, no significant carbohydrate was lost. When peracetic acid was applied at 21%, the enzymatic hydrolysis of cellulose increased from 6.8% for untreated biomass to 98% in the peracetic acid-pretreated biomass. Ionic Liquid Pretreatment Ionic liquids are strong solvents. They are able to dissolve the components of LB at ambient to moderate temperatures. Furthermore, ionic liquids are highly tunable through the selection of anion and cation. Beyond toxicity and corrosivity, other considerations affecting the selection of an ionic liquid include price, availability, water tolerance, biodegradability, and physical properties such as viscosity, melting point, dipolarity and hydrogen bond basicity. An effective wood dissolution is possible when both the ionic liquid and conditions are properly identified and employed. The most significant consideration for practical large-scale operations is the toxicity of the ionic liquid to be used. For example, 1-butyl-3-methylimidazolium chloride ([BMIM][Cl]) is a good solvent to use on cellulose as it is only moderately toxic compared to that of 1-ethyl-3-methylimidazolium chloride ([EMIM] [Cl]) (Swatloski et al., 2004; Wu et al., 2004). Corrosivity of the selected ionic liquid is also important. It plays a large role in the economics of a commercial operation. One can minimize corrosivity by selecting an ionic liquid that is halogen free. Good choices include 1-ethyl-3-methylimidazolium acetate ([EMIM][OAc]) (Liebert, 2010) and 1,3-dimethylimidazolium-dimethylphosphate ([MMIM][(MeO)2PO2]) (Zavrel et al., 2009). Balancing the physical properties and operational conditions is important to obtaining the most ideal dissolution of LB. For example, if the viscosity of an ionic liquid is high, it may be necessary to operate the pretreatment at a high temperature to obtain a practical dissolution. As a result, the reactions may become unstable and may give rise to undesirable reactions and by-products. A solution to this problem is to reduce the viscosity of the ionic liquid by combining it with a cosolvent. A good viscosity-reducing cosolvent is polyethylene glycol (Willauer et al., 2000). The dissolution rate is inversely proportional to wood chip sizes. For example, ball-milled wood powder produces a higher dissolution rate than does sawdust. The dissolution rate for TMP fibers is higher than that for sawdust, which is much greater than that of wood chips (Kilpeläinen et al., 2007; Sun et al., 2009; Zavrel et al., 2009). 467 In addition to particle size, dissolution efficiency is also highly sensitive to the water content. Water attenuates the dissolution effectiveness of an ionic liquid. Studies have shown that storing wood chips at warm temperatures, e.g. 50 C or 90 C, reduces the water content of the wood and thus improves the pretreatment effectiveness (Kilpeläinen et al., 2007; Sun et al., 2009). Reducing the water content improves the dissolution power of an ionic liquid regardless of the type of wood being treated. However, if the wood becomes too dry, the wood composition may change unfavorably. Determining the precise water content of LB is quite difficult and is complicated due to the diversity of environmental conditions of the regions from which the wood studied grew. Variables such as humidity and variances in species present a challenge when comparing literature on the subject (Wang et al., 2012). The type of LB, dissolution time, temperature and ionic liquid to wood ratio, are all factors that contribute to the dissolution power of an ionic liquid. That said, those ionic liquids that were effective at dissolving both lignin and cellulose were also excellent at overall LB dissolution. One of the best solvents for wood chips is the combination of 1-allyl-3-methylimidazolium chloride ([AMIM][Cl]) and [EMIM][OAc]. Ionic liquids derived from polycyclic amidine bases have been shown to dissolve aspen wood chips completely (D’Andola et al., 2008). The ionic liquids used in this study were 1,8-diazabicyclo[5,4,0] undec-7-enium salt, and 1,8-diazabicyclo[5,4,0] undec-7-enium chloride [HDBU] [Cl] (D’Andola et al., 2008). It has been observed that [AMIM][Cl] can effectively dissolve both hardwood and softwood wood chips. However, the same solvent only partially dissolved Norway spruce (Kilpeläinen et al., 2007). The efficiency of [AMIM][Cl] in dissolution of wood is due to the presence of p-electrons both in the alkenyl chain as well as in the imidazolium ring. Possible pep interactions may occur between the aromatic part of lignin and the ionic liquid (Hunter and Sanders, 1990; Kilpeläinen et al., 2007). The highest solubility of maple wood powder was achieved using [AMIM][Cl] and [BMIM][Cl] (Lee et al., 2009). [EMIM][OAc] can completely treat three types of wood chips. It is used to treat spruce, beech and chestnut. However, it only partially dissolves silver fir (Abies alba) wood chips (Zavrel et al., 2009). A plausible explanation for this difference is that silver fir contains more cellulose (50.3%) and lignin (27.7%) than the other wood species (Kuznetsov et al., 2002). When comparing the dissolution effectiveness of [EMIM][OAc] to [BMIM][Cl] and controlling for species, wood chip size, and temperature one can obtain a 3.6-fold increase in dissolution effectiveness using [EMIM][OAc] vs [BMIM][Cl]. In this case, southern yellow pine wood chips were treated at 110  C. This 3.6-fold increase in dissolution effectiveness is
468 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES FIGURE 27.5 A process for the dissolution of wood and regeneration of ionic liquid.Source: Sun et al., 2009. (For color version of this figure, the reader is referred to the online version of this book.) Wood chips Ground wood IL Cooking IL Recycle Wood/IL solution Cellulose-rich materials Filtration Regeneration Acetone/H2O Wash, dry Lignin in solution Evaporation of acetone Lignin attributable to the basicity of the acetate anion, higher than that of the chloride anion. Thus, [EMIM][OAc] is stronger at breaking the intramolecular hydrogen bonds (Fort et al., 2007; Sun et al., 2009). An opportunity for improvement in using ionic liquids is better recovery of solubilized cellulosic materials and lignin. A significant drawback is that much of the hemicellulose is washed away during the recovery process. Figure 27.5 illustrates the process. Following pretreatment with an ionic liquid, an enzymatic hydrolysis pretreatment is applied to produce the sugars for downstream fermentation. This pretreatment can recover as much as 90% of the cellulose for enzymatic hydrolysis. While cellulose is recovered at a high rate, the hemicelluloses are not as they are washed away. As Figure 27.5 illustrates, the ionic liquids are recovered and recycled for reuse. Even so, the problems of price, toxicity, and the lost hemicellulose persist, which inhibit wide adoption in industrial scale operations. Sulfite Pretreatment to Overcome Recalcitrance of Lignocelluloses Sulfites are found to be efficient agents for pretreating LB, in both hardwoods and softwoods (Zhu et al., 2009). In sulfite pretreatment to overcome recalcitrance of lignocelluloses (SPORL), the sulfite refers to any sulfite, bisulfate or combination. A combination may contain any two of the following three reagents: sulfite (SO2 3 ), bisulfite (HSO 3 ), and sulfur dioxide (SO2, or H2SO3). The specific combination to use depends on the pH of the pretreatment liquor and the temperature (Zhu et al., 2009). The first step in the process is to treat wood chips or another LB feedstock with a sulfite salt solution where the salt may be sodium, magnesium or calcium. This first step usually operates at a temperature between 160  C and 190  C and at a pH between two and four for 10e30 min. The second step is to fiberize the resultant biomass using a disk mill. This yields a fine fibrous substrate suitable for robust saccharification and fermentation (Shuai et al., 2010). The typical acid charge on oven-dried wood is 0.5e1% for hardwood and 1e2% for softwood. The typical bisulfite charge is 1e3% for hardwood and 4e8% for softwood (Zhu et al., 2010b). More than 90% of the cellulose was converted from SPORL-treated spruce chips. In this case the oven-dried wood chips were treated with an 8e10% bisulfate and 1.8e3.7% sulfuric acid combination at 180  C for 30 min. The resultant material was treated with enzyme hydrolysis for 48 h using 14.6 cellulase and 22.5 b-glucosidase per gram of substrate. Shuai et al., and Zhu et al., have performed comparative SPORL studies using dilute-acid pretreatments for both softwoods and hardwoods (Shuai et al., 2010; Zhu et al., 2010b). In these studies, it was observed that SPORL is better at saccharification of hexoses and pentoses than was a dilute acid (DA) treatment. In one case, where oven-dried spruce was treated at 180  C for 30 min with 1% H2SO4 at a 5:1 liquor-to-wood ratio, 87.9% of the hexoses and pentoses were recovered using SPORL versus a similar DA treatment where 56.7% of the saccharides were recovered.
469 PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS TABLE 27.4 Pretreatment (180  C, 30 min) Comparison of Chemical Pretreatment Method on Lodgepole Pine Wood Chips Initial Liquid pH Untreated Disk-Milling Energy (kWh/ton wood) Size Reduction Energy Savings (%) 699 (%) 1.7 Hot Water 5.0 680 2.7 16.0 Acid 1.1 412 41.0 41.6 SPORL 4.2 594 15.0 75.1 SPORL 1.9 153 78.1 91.6 About 92.5% of the cellulose was recovered in an SPORL process utilizing a 9% sodium sulfite (w/w of wood) and 77.7% for the DA (Shuai et al., 2010). In another study of aspen, or Populus tremuloides, a comparison was made between an SPORL pretreatment using a combination of sulfuric acid and sodium bisulfite and a dilute sulfuric acid (DA) pretreatment. It was observed that nearly 60% more ethanol was produced from the SPORL-treated wood than from the DA-treated wood. In both cases enzymatic hydrolysis was conducted using 10 FPU cellulase per gram glucan for 120 h (Zhu et al., 2010b). Table 27.4 highlights a handful of comparisons between treatment methods and their effectiveness where the pretreatment conditions were L/W ¼ 3, disk-milling solids loading ¼ 30% (the solid contents of pretreated wood chips), and disk plate gap ¼ 0.76 mm. Sodium bisulfite charge was 8% on oven-dried wood for the two SPORL runs; sulfuric acid charge was 2.21 (w/w) on oven-dried wood for the DA and low pH SPORL runs, and 0 for the hot water and high pH SPORL runs. SPORL is an attractive pretreatment method due to several features. SPORL generates much less furfural and hydroxymethylfurfural (HMF) than does a simple DA pretreatment. SPORL significantly enhances fermentation yields by weakening the hydrophobic relationship between lignin and enzymes and enhancing saccharification of cellulose. One of the products of SPORL is a sulfonified lignin, which has potential economic value as a directly marketable coproduct within existing markets and for opening new markets. The energy consumption in the size reduction process is reduced by an order of magnitude. Lastly, SPORL has demonstrated commercial scalability with low technological and environmental risks (Zhu et al., 2010b). Hot Water Utilizing liquid water by itself, as the only pretreatment reagent, is an option of interest as it is environmentally friendly and inexpensive compared to other pretreatment methods (Amidon et al., 2008; Liu, 2010; Mosier et al., 2005). High pressure is applied to keep the water in a liquid state while it is at elevated temperatures (Hendriks and Zeeman, 2009). This enables the water to penetrate the cell structure of the biomass and thus hydrate the cellulose and remove the hemicelluloses. Another feature of water is that it has a high dielectric constant. This facilitates ionic substances to disassociate and allows for the dissolution of hemicelluloses and a portion of the lignin. When the water temperature exceeds 150  C, the hemicellulose begins to solubilize. The degree to which this occurs is determined by thermal, acid and alkali stability of the hemicellulose, which is dependent on the composition of the hemicellulose backbone and the branching groups. Temperature of the water can selectively solubilize hemicelluloses. A 75% maximum xylan solubilization in the hot water extract of sugar maple was obtained at 175  C after 2 h, whereas only 30% of the initial xylan was removed from a 2 h treatment at 152  C (Mittal et al., 2009). When the water temperature exceeds 180  C an exothermal reaction begins. It is most likely related to the solubilization of the hemicelluloses (Brasch and Free, 1965). Another result of the thermal process is that the pH of the extract decreases to 3e4 (Gregg and Saddler, 1996a). Portions of the hemicelluloses are hydrolyzed, which form acids such as acetic acid. These are released from acetylated polysaccharides in the wood. These acids lower the pH and catalyze the additional hydrolysis of hemicellulose (Liu and Wyman, 2003; Liu, 2008; Tunc and van Heiningen, 2008; Zhu et al., 2005). Depending on the intensity of the hot water extraction, sugars may dehydrate. When hexose sugar dehydrates HMF, also known as HFM or 5-hydroxymethyl2-furaldehyde, is formed. When pentose sugar dehydrates, furfural is formed. In addition to solubilizing hemicellulose, hot water treatment can lead to solubilization of portions of lignin (Ramos, 2003). Regardless, the produced compounds are usually phenolic heterocyclic compounds such as vanillin, vanillin alcohol, furfural and HMF. This is especially true in strong acidic
470 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES conditions. Additionally, these compounds tend to inhibit or toxify bacteria, yeast, methanogens and archae. This is a significant disadvantage in using hot water to extract cellulose and hemicellulose (Brownell et al., 1986). Hot water extracts can be converted to desired products as well, i.e. via separation and fermentation (Liu et al., 2009; Shupe and Liu, 2009). Fermentation is also strongly inhibited when a hydrolysate is produced from a treatment containing 3% or more of solids or the treatment temperature exceeded 220  C for 2 min. These conditions likely yield furfural or soluble lignin compounds. At temperatures in excess of 250  C pyrolysis begins to take place (Laser et al., 2002). Therefore, one should avoid these high temperatures. Another undesirable effect of thermal pretreatment is that it may increase the crystallinity index (CrI) of cellulose (Weimer et al., 1995). It is important to remove the soluble lignin compounds quickly. Since lignin is highly reactive, the disengaged lignin will recondense and precipitate onto the biomass (Liu and Wyman, 2003). This seems to be more prevalent in cases where severe pretreatment conditions are used. In these cases, more condensation and precipitation of lignin compounds takes place and sometimes, soluble hemicellulosic compounds such as furfural and HMF are also produced (Mittal et al., 2009) and polymerized (condensed) and deposited onto the extracted biomass. Despite the undesirable effects above, when compared to other pretreatment methods, liquid hot water wood extraction still has a major advantage. Since a large volume of water is used the solubilized hemicelluloses and lignin compounds appear in lower concentrations. As a result, the risk of undesirable degradation products is reduced. The substances in the extract can be separated and converted to desired products. Figure 27.6 illustrates the three methods of liquid hot water reactors. They are differentiated by their configurations. One is cocurrent, another is countercurrent and the third is a flow-through reactor. Briefly, in cocurrent pretreatment, the biomass and water are heated and held at the desired conditions for a specific residence time prior to allowing it to cool. In the countercurrent design, water and lignocellulosic material flow in opposite directions through the reactor. The flow-through reactor is designed such that hot water is passed over a stationary bed of LB and carries the hydrolysate and dissolved lignocellulosic components out of the reactor (Hendriks and Zeeman, 2009). Steam Explosion LB is bulky and much of the volume is “empty” or saturated by air. Air in LB can be replaced by liquid water at high temperature and pressure. When water-saturated LB is suddenly exposed to low pressures, liquid water suddenly expands when vaporized forcing LB to disintegrate into fine particles. The Masonite process was invented in 1926 (Mason, 1926) employing this water to steam explosion process. Since then, the steam explosion pretreatment (SEP) has been a common technique (Mason, 1928). SEP is used to break the crystalline and lignocellulosic structure of biomass into its three major components, cellulose, hemicellulose and lignin. SEP enhances the resultant cellulose’s susceptibility to enzymatic hydrolysis. High-pressure, saturated steam is applied to biomass for a brief period and then allowed to rapidly decompress to atmospheric pressure, hence the term explosion. The explosion breaks up solid particles and is used as a standard practice in chemical pulping operations. The steam is vented and the biomass is discharged to a larger vessel for rapid flash cooling (Mosier et al., 2005). SEP is as much a mechanical process as it is a thermal process (Holtzapple et al., 1989). Regardless, the explosion per se, whether it causes particle disintegration or not, does not play a significant role in producing a product that is easily digested by enzymes (Brownell et al., 1986). A more likely mechanism at play is the treatment’s effect in removing hemicellulose (Mosier et al., 2005). Applying acid catalysts, usually SO2 (or sulfite), enhances this effect by reacting, in conjunction with water, within the interstitial spaces to form sulfuric acid and thus catalyze hemicellulose degradation (Gregg and Saddler, 1996b). SEP effectiveness and the chemical changes that take place depend on residence time, temperature, chip size, and moisture content. Effectiveness is determined by the amount of hemicellulose solubilized and the rate of subsequent enzymatic hydrolysis. Optimal outcomes are obtained when pretreatment occurs at either high temperature or short residence time, such as 270  C for 1 min, or at lower temperature and longer residence time, such as 190  C for 10 min. Generally, initial treatment pressures range from 0.69 to 4.83 MPa and treatment temperature ranges from 160  C to 260  C. At high temperatures water acts as an acid. Thus, during the treatment time, the hemicellulose hydrolyzes into soluble sugars. The hemicellulose is considered to autohydrolyze as a result of exposure to the acetyl groups in the organic acids formed at these high temperatures. Acetic acid is formed from the acetylated hemicelluloses. The pH during SEP is kept quite low, near pH 3e4. SEP degrades a significant portion of the hemicellulose (Sun et al., 2005). However, degradation of hemicellulose may not stop at this point. If the treatment conditions are severe, the solubilized hemicellulose may undergo a series of secondary reactions that yield furfural and HMF. These severe conditions may be high temperature or a long incubation time. Furfural and HMF are
PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS (a) 471 Co-current reactor Check valve Pretreated biomass Water Biomass Steam Insulated coil (b) Counter-current reactor Water with dissolved extracts Biomass Jacketed reactor Pretreated biomass Water (c) Flow-through reactor Water Jacketed reactor Biomass Water with dissolved extracts FIGURE 27.6 Schematic diagram illustrating three types of liquid hot water reactors: (a) cocurrent, (b) countercurrent, (c) flow through. (For color version of this figure, the reader is referred to the online version of this book.) undesirable products that inhibit enzymatic hydrolysis and limit the effectiveness of fermentation. Meanwhile, lignin is partially depolymerized, some lignin is redistributed within the material and some may be removed completely from the fibers, each of which contribute to an improved exposure of the cellulose domains (Chen and Qiu, 2010). The reduction in hemicellulose and partial removal of lignin exposes the cellulose surface and thus improves the ability of the enzyme to attack the cellulose microfibrils (Alvira et al., 2010). If the treatment conditions are severe, some degradation of cellulose to glucose can occur. One study reported an enzymatic hydrolysis efficiency of 90% over 24 h using poplar chips using SEP. This was significantly better than the control where the enzymatic hydrolysis efficiency was 15% from untreated
472 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES poplar chips (Grous et al., 1986). That said, SEP is more effective on agricultural residues than in wood as a result of the lower acetic acid content in the hemicellulose portion of the biomass. Adding a supplemental acid to the SEP reduces both residence time and temperature. Adding an acid such as H2SO4 (or SO2) or CO2, typically 0.3e3% (w/w), improves hydrolysis, decreases the production of inhibitory compounds and leads to a more complete removal of hemicellulose (Kumar et al., 2009a). For the effective treatment of softwoods, adding an acid catalyst is essential to make the substrate susceptible to enzymatic hydrolysis. Adding a supplemental acid also improves the enzymatic hydrolysis of the residual solids and decreases the production of inhibitory compounds (Morjanoff and Gray, 1987). These three parameters, the level of H2SO4 (or SO2) or CO2, the residence time and the temperature, are the most influential parameters on total sugar yield. For SEP treatment of sugarcane bagasse, the optimal conditions are 1% H2SO4, 220  C; 30 s residence time, and a water-to-solids ratio of 2:1 (Holtzapple et al., 1989). After SEP treatment under these conditions sugar production was determined to be 65.1 g sugar/100 g. A two-step SEP is a good pretreatment for softwood (Söderström et al., 2003). In this case, the first step is to optimize the amount of hydrolyzed hemicellulose by employing low severity conditions where the biomass is treated at 180  C for 10 min with 0.5% H2SO4. In the second step, the solid material from the first step is washed and impregnated again with H2SO4. SEP is applied again using more severe conditions. This time the biomass is treated at 180  Ce220  C for a longer time, between 2 and 10 min, and with a higher concentration of acid catalyst, 1e2% H2SO4. These treatments appear to hydrolyze a portion of the cellulose and make it more accessible to enzymatic attack (Sassner et al., 2008). The most favorable conditions for Salix wood is to impregnate it with 0.5% H2SO4 at 200  C for between 4 and 8 min. The yield is thus 55.6 g glucose and xylose per 100 g dry biomass (Sassner et al., 2008). If one uses SO2 as the impregnating agent in spruce chips, the sugar yield is almost independent of impregnation time and slightly increases with decreasing chip size (Monavari et al., 2009). Shorter impregnation times result in slightly lower mannose yields in larger chips. The optimum pretreatment conditions when using SO2-catalyzed SEP for lodgepole and Douglas fir pine is 200  C for 5 min with 4% SO2 (w/w) (Ewanick et al., 2007; Kumar et al., 2010). Another option for an impregnating agent is to use a weak organic acid, in particular, lactic acid (Monavari et al., 2011). It was observed that it was not efficient and resulted in lower sugar yields in spruce, with or without the addition of SO2. However, using a weak organic acid is more environmentally friendly than using an inorganic acid as it would biodegrade in a waste stream or be used for production of a biogas such as methane. Particle size, by itself, is not a significant contributor to SEP effectiveness. Some studies report that larger particle size may improve the outcomes from SEP (Cullis et al., 2004). In these studies, pretreated Douglas fir, a softwood, was milled to three particle sizes: <0.422 mm screenings, 1.5  1.5 cm and 5  5 cm. They were then steam exploded using SO2. It was observed that the largest particle size suffered less from pretreatment severity and had the highest cellulose recovery. It had larger quantities of solubilized carbohydrate and contained fewer furan degradation products. The smaller particle sizes produced outcomes containing more solubilized hemicellulose and lignin. If the resultant biomass is further refined to particles of a finer size by plate milling, with a 0.178 mm gap, the initially larger particle size showed a higher lignin removal with peroxide washing and a greater rate of enzymatic hydrolysis. This is likely due to the reduction in lignin redeposition as a result of treatment severity. These findings were substantiated in studies of steam-exploded pine, another softwood (Ballesteros et al., 2000). The largest of the sizes exhibited a higher cellulose recovery and also a higher content of solubilized hemicellulose. Conversion of cellulose to glucose was only slightly higher from the larger particle sizes. However, the total recovered glucose, including the solubilized glucose from the steam explosion, was much higher when starting with the largest particle size (Ballesteros et al., 2000). Starting with a herbaceous feedstock, such as Brassica carinata residues, produced different results. Although the cellulose recovery was still higher for the 8 and 12 mm fractions, the smaller particle sizes performed better during enzyme hydrolysis. The 5e8 mm and 2e5 mm fractions yielded 100% while the 8e12 mm fraction produced 85% (Ballesteros et al., 2002). This suggests that lignin condensation is not as influential in herbaceous feedstock. It is a critical factor when pretreating softwood. In softwood, the larger particle size produces a higher maximum glucose yield, over 80%, compared to smaller particle sizes with yields under 70% (Ballesteros et al., 2002). SEP-treated hardwood exhibited no difference in either enzyme digestibility or ethanol yield between disparate particle sizes, in particular between 2e5 mm and 12e15 mm (Negro et al., 2003). This indicates that the severity of the treatment plays a larger role than particle size when using softwood. In these cases, smaller particles increase the lignin condensation and recalcitrance to enzyme hydrolysis.
PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS Overall, SEP is an attractive pretreatment method considering the low amount of energy required to reduce the biomass size compared to mechanical comminution. Conventional mechanical size reduction methods require 70% more energy than SEP to achieve the same particle size reduction (Mittal et al., 2009). Additionally, SEP is attractive because there are no recycling or environmental concerns. This too lowers cost. SEP is thus recognized as one of the most cost-effective pretreatment methods for hardwoods and agricultural wastes. It has been extensively tested on a wide array of lignocellulosic feedstock. It has been observed that SEP is less effective for the pretreatment of softwoods. The most significant limitations include the partial destruction of xylan, incomplete disruption of the lignin-carbohydrate matrix, low lignin removal, and lignin redistributes over the surface of cellulose (Chen et al., 2010). Additionally, there is a risk of producing undesirable compounds such as furfural, HMF and other soluble phenolic compounds. These undesirables inhibit microbial growth and enzymatic hydrolysis. Thus, prior to fermentation, SEP-treated LB must be washed with water to remove these undesirable materials along with water-soluble hemicellulose. Unfortunately, this wash lowers the overall effectiveness as it washes away around 20e25% of the initial dry matter and a portion of the soluble sugars (Sun and Cheng, 2002). Ammonia Fiber Explosion Another explosion pretreatment is the ammonia fiber explosion (AFEX) process. Instead of using liquid water under high pressure, liquid ammonia is used. AFEX is an effective and somewhat economically attractive method to increase the yields of fermentable sugars from LB (Holtzapple et al., 1991; Holtzapple et al., 1992). In this method LB is exposed to liquid ammonia, not ammonium hydroxide (i.e. no water/moisture), at moderate temperatures and elevated pressures for a longer period of time. After the appropriate residence time, the system is rapidly vented allowing the liquid to vaporize and literally explode the fibrous material. Typically, 1e2 kg of liquid ammonia is used for each kg of dry biomass. The system operates at temperatures below 100  C, pressures above 3 MPa, and is quite tolerant of pH. Any pH under 12 appears suitable. The residence time is between 10 and 60 min. Under these conditions, the system forms few degraded sugar products yet gives a high yield of desirable sugar products (Mosier et al., 2005). AFEX is an attractive treatment method for a variety of herbaceous crops and grasses as it significantly improves the saccharification rates. It has been tested on a variety of LB including aspen chips, softwood and 473 kenaf newspaper, alfalfa, wheat chaff, wheat straw, barley straw, rice straw, bagasse, coastal Bermuda grass, switchgrass, corn stover, and municipal solid waste. One of the benefits is that AFEX only solubilizes a trivial amount of solid material. Also, compared to acid pretreatment and acid-catalyzed steam explosion, very little hemicellulose or lignin is removed. Lastly, the structure of the material changes such that the result is an increase in water-holding capacity and improved digestibility. Although physically modified, the chemical composition of the material following AFEX pretreatment is essentially unchanged from its original condition. The benefit is illustrated as follows: over 90% hydrolysis of the cellulose and hemicellulose may be obtained after AFEX pretreatment of Bermuda grass where 5% of that is lignin. The result is similar for bagasse except 15% of the hydrolysate is lignin (Holtzapple et al., 1992). These low-lignin containing biomasses readily hydrolyze at near theoretical yields of sugars. The resulting sugars ferment rapidly with a high yield into a variety of desired products. Since the AFEX treatment produces very few inhibitors to the downstream biological processes, a water wash is not necessary (Dale et al., 1984; Mes-Hartree et al., 1988). Materials with a high lignin content, around 25%, have proved to be recalcitrant to AFEX. Therefore, AFEX is a less effective pretreatment method for hardwood chips, some newspaper material, and nut shells (Teymouri et al., 2005). AFEX does not require a small particle size for it to be an effective treatment option (Larson and King, 1986) like steam explosion and hot water treatments. The most significant cost is that associated with recycling the ammonia following pretreatment (Kumar et al., 2009a). Since pure ammonia is used in the process, more stringent environmental and recovery procedures are required. Thus, recycling is necessary to reduce the environmental impact and the cost of the procedure. To recover the ammonia, a superheated ammonia vapor, at temperatures upward of 200  C, is used to vaporize and strip the residual ammonia from the pretreated biomass. The evaporated ammonia is then drawn off the system by a pressure controller for final recovery (Holtzapple et al., 1990). Using this recovery method has demonstrated that over 99% of the ammonia can be recycled successfully. Even so, the overall capital and operating costs are higher than other comparable methods. Supercritical Carbon Dioxide Explosion To address the expense of the AFEX method, the supercritical carbon dioxide explosion method was developed. Compared to steam explosion, the supercritical CO2 explosion method produces fewer inhibitory
474 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES compounds. Additionally, CO2 is much more environmental friendly than organic solvents used in the organosolv method and the ammonia used in the AFEX method. Because carbon dioxide is nontoxic, physiologically safe and inexpensive it is used in a variety of industries, for example, in food and pharmaceutical production. The critical temperature of CO2 is 31.1  C and its critical pressure is 73 atm. The term supercritical refers to a fluid that at standard temperature and pressure would exist in its gaseous state. However, when compressed using high pressures and at temperatures above the critical point, the gas condenses into a liquidlike density. In this state it retains the characteristics of mass transfer that are “gaslike” but with the solvating power that is “liquidlike” (Kim and Hong, 2001). Carbon dioxide molecules are small, like water and ammonia, and thus it penetrates the small pores of LB. It is believed that CO2 forms carbonic acid and thus it should increase the hydrolysis rate. Furthermore, at low temperatures it is thought to prevent significant decomposition of the monosaccharides by the weak acid. However, the primary effect of supercritical carbon dioxide explosion is from the explosion whereby it disrupts the biomass structure and increases the surface area and improves its vulnerability to enzymatic attack (Conner and Lorenz, 1986; Zheng et al., 1998). Despite these advantages, the operating and capital costs of the supercritical carbon dioxide explosion pretreatment option remain prohibitive. Biological Pretreatment Perhaps the most natural pretreatment of biomass is a purely biological method. Nature commonly employs lignin-degrading microorganisms such as white, brown or soft-rot fungi (Lee, 1997; McMillan et al., 1999). A study that investigated the effect of high-yield concentrated recombinant MnP (rMnP), produced from the yeast Pichia pastoris on the biobleaching of kraft pulps found that rMnP applied at 30 U/g pulp for 24 h followed by alkali extraction removed a significant quantity of lignin from both hardwood and softwood unbleached kraft pulps (Xu et al., 2010). The rMnPtreated pulp was more susceptible to subsequent peroxide bleaching compared to the control pulp. More than 60% of the kappa number was reduced by sequential rMnP treatments and alkaline extractions. When using white-rot fungi, such as Ceriporiopsis subyermispora, to treat sugar maple chips, the amount of extracted hemicellulose can be increased (Barber, 2007). The biotreatment alters the physical and chemical structures of the LB and removes a portion of the noncarbohydrate mass. Because biological pretreatment is safe, environmentally friendly and energy saving it is gaining more attention (Okano et al., 2005). The downside is that biological pretreatment is too slow for some industrial applications and some material is lost to the microorganism as it is a consumer of hemicellulose, cellulose and lignin (Bohlmann, 2006). The microorganisms are also susceptible to poisoning by lignin derivatives (Hamelinck et al., 2003). Biological pretreatment by itself may not be the best solution but it could provide value when employed in conjunction with other pretreatment options. Acid Hydrolysis Water is a weak acid by itself; however, adding a salt to water will enhance the activity of the acid. Aqueous acids, especially those with a salt, autoseparate into hydrogen cations and hydroxyl anions, where one side of the cleaved sugar polymer receives the hydrogen cation and the other receives the hydroxyl group. One can apply acid hydrolysis either as a pretreatment or as a main hydrolysis step. A variety of acids act well at ambient temperatures to pretreat LB and prepare the material for anaerobic digestion. LB is eventually hydrolyzed into monosaccharides, furfural, HMF and other volatile products. The lignin, however, condenses and precipitates out as a result of the pretreatment (Esteghlalian et al., 1997; Liu and Wyman, 2003; Shevchenko et al., 1999). Concentrated acids are quite powerful, act at mild temperatures and result in rapid reactions. However, H2SO4, H3PO4 and HCl are highly toxic, corrosive, and hazardous. Reactors for acid hydrolysis need to resist corrosion. Furthermore, recovering the concentrated acid from the hydrolysis effluent is important to reduce the negative environmental consequences and to reduce costs. Hydrolysis using a DA is an effective pretreatment for LB (Hinman et al., 1992). It produces high sugar yields from some hardwoods, like poplar and aspen. In one study, poplar wood was pretreated with a 2% sulfuric acid at 190  C for 1.1 min and this was followed by an enzymatic hydrolysis (Wyman et al., 2009). In this particular study, the xylose yield was 18.5% and the glucose yield was 64.3% where the raw material contained 25.8% xylose and 74.2% glucose (Wyman et al., 2009). In another study of aspen wood, the wood was pretreated with a 1.1% sulfuric acid at 170  C for 30 min and followed by enzymatic hydrolysis (Tian et al., 2011). The xylose and mannose yield was 13 wt% (18 wt% theoretical contents) and the glucose yield was 85% following treatment (Tian et al., 2011). See Table 27.5 for a comparison of concentrated and DA pretreatments. There are essentially two classes of DA pretreatment processes: high-temperature continuous-flow and low
PRETREATMENTdBIOMASS SIZE REDUCTION BY PHYSICAL OR MECHANICAL METHODS TABLE 27.5 Concentrated Acid Dilute Acid 475 Comparison of Concentrated and Dilute Acid Pretreatments Wood Species Particle Size Acid Reaction Condition Effects and Results Hybrid poplar (Zhang et al., 2007) 40e60 mesh H3PO4 83e85.9% 50  C, 30e60 min Enzymatic cellulose digestibility 97% for 24 h Douglas fir (Zhang et al., 2007) 40e60 mesh H3PO4 83e85.9% 50  C, 30e60 min Enzymatic cellulose digestibility 75% for 24 h Spruce 1e5 mm H2SO4 70 wt% and 30 wt% 70 wt% H2SO4 (50  C, 2 h); 30 wt% H2SO4 (80  C, 6 h) 74% pentoses, 69% hexoses Birch 1e5 mm H2SO4 70 wt% and 30 wt% 70 wt% H2SO4 (50  C, 2 h); 30 wt% H2SO4 (80  C, 6 h) 73% pentoses, 68% hexoses Poplar (Wyman et al., 2009)  6.35 mm H2SO4 2% 190  C, 1.1 min 25.8% xylose 74.2% glucose Aspen (Tian et al., 2011) 6e38 mm H2SO4 1.1% 170  C, 30 min 72.2% pentoses 84% glucoses  Loblolly pine (Marzialetti et al., 2008) 35e60 mesh Trifluoroacetic acid pH 1.65 150 C, 60 min 70.3% pentoses, 22.9% hexoses Pine (Orozco et al., 2011) 1 mm H3PO4 2.5 wt% 175  C, 10 min 100% xylose 13% glucose Source: Wang and Liu, 2012. temperature batch processes. High-temperature systems operate at temperatures over 160  C and are appropriate for solutions with a low concentration of solids, between 5% and 10%. Low-temperature systems operate under 160  C and are appropriate for solutions with a high concentration of solids, between 10% and 40%. Even though a simple acid pretreatment significantly improves the rate of a hydrolysis process, it costs higher than other physicochemical pretreatment processes. One such process is steam explosion and was discussed previously in this chapter. Another consideration for an acid hydrolysis pretreatment is that one must neutralize the hydrolysate prior to subsequent enzymatic hydrolysis or fermentation (Sun and Cheng, 2002). Alkaline Hydrolysis Alkaline pretreatment is viewed as a viable treatment method because of its low energy requirement and low capital equipment and operational costs (Zhao et al., 2008). This process operates at lower temperatures and pressures than other pretreatment methods. However, at these conditions, the process is measured in hours or days vs. minutes or seconds for high-temperature, high-pressure methods (Karr and Holtzapple, 2000). Additionally, one may recover or regenerate many of the caustic salts. Alkaline pretreatment may follow an SEP and may be followed by an enzymatic hydrolysis pretreatment (Montane et al., 1994; Pan et al., 2006). The initial reactions of alkaline pretreatment involve solvation and saponification. Solvation, similarly associated with dissolution or diffusion, is where the solvent surrounds an ion, typically sodium dissolved in water. Traditional NaOH treatment requires high temperatures to be effective (Zhao et al., 2008). It may be supplemented with urea to lower operational temperatures and improve dissolution (Zhao et al., 2008). The alkaline solvent then saponifies the intermolecular ester bonds that cross-link xylan hemicelluloses and other components including lignin and other hemicelluloses. Removing these cross-links increases the porosity of the lignocellulosic materials. This improves the penetrability of the material to the solvent and swelling of the biomass follows. The swollen biomass is thus more vulnerable to enzymatic and bacterial activity. Compared with acid hydrolysis, alkaline hydrolysis generally causes less sugar degradation. That said, dissolution or solubilization of LB increases with alkali concentrations. At strong alkali concentrations, peeling of end-groups may occur. This leads to alkaline hydrolysis and degradation of the dissolved polysaccharides. Furthermore, this may also produce unwanted byproducts. However, there may be a downstream advantage in subsequent conversion treatments. It increases the internal surface area, decreases the DP, decreases crystallinity and separates linkages between lignin and carbohydrates causing an overall disruption of the lignin structure (Fengel, 1984). This provides opportunity for increased enzymatic and bacterial activity in downstream processes. An alternative process to improve sugar content is to use aqueous potassium hydroxide, which selectively removes xylan. Keeping the temperature low, at or
476 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES below room temperature, prevents peeling (Hon and Shiraishi, 2001). It appears that monomeric forms of hemicelluloses are easily degradable to other volatile compounds. Glucomannans and xylans are particularly vulnerable to peeling. However, by pretreating with a 3% NaOH and 12% urea at 15  C one can achieve a 60% glucose conversion (Zhao et al., 2008). Calcium hydroxide, or slake lime, is yet another effective alkaline pretreatment agent. It is one of the least expensive and it is highly recyclable (Karr and Holtzapple, 2000). Using common lime kiln technology, one can recover calcium hydroxide by regenerating it from insoluble calcium carbonate. Lime pretreatment removes lignin and hemicellulose and increases the CrI. Pretreatment with dilute NaOH decreases the lignin content within a range of 24e55% to 20% and increases the digestibility of NaOH-treated hardwood from 14% to 55%. No effect was observed for softwoods with lignin content greater than 26% (Bjerre et al., 1996). Dilute NaOH pretreatment causes swelling, which, as stated previously, has downstream benefits. The overall effectiveness of alkaline pretreatment depends on the lignin content of the biomass. Furthermore, it changes the cellulose structure such that it is less dense and more thermodynamically stable than native cellulose (Hendriks and Zeeman, 2009; Liu and Wyman, 2003). HYDROLYSIS Hydrolysis is generally defined as the depolymerization of a substance via hydration. An aqueous acid’s ions act to cleave long polymers like cellulose, hemicellulose and lignin into smaller chains. Pretreating LB to undergo hydrolysis or converting polysaccharides into monosaccharides will enhance later fermentation by improving TABLE 27.6 the ability of anaerobic organisms to digest the resultant, simpler sugars. Hydrolysis requires extended residence time. Unfortunately, monosaccharides degrade into other nonsugar molecules when subjected for extended times to relatively high temperatures and acid conditions (Hsu, 1996; Wyman et al., 2005). The hydrolysis reaction rate accelerates when either a chemical or an enzymatic catalyst is used and when the material to by hydrolyzed is concentrated. Enzyme hydrolysis is highly specific and relatively fast. Using an enzyme to act on its target polysaccharide will convert it rapidly into its component monomers. Additionally, this will convert the insoluble polysaccharide into a soluble monomer. Enzymatic hydrolysis is best applied after other pretreatment methods that leave cellulose as a major component. The most common method of saccharification is enzymatic hydrolysis following acid hydrolysis (Harun, 2010). The enzyme cellulase converts cellulose into glucose. Cellulases are so specific that they only affect cellulose and do not treat hemicelluloses in the LB (Wang et al., 2012). There are five general types of cellulases. They are classified by the reactions they catalyze. These five cellulases are endocellulases, exocellulases, cellobiases, oxidative cellulases and cellulose phosphorylases (Bayer et al., 1998). Table 27.6 summarizes the effectiveness on the hydrolysis of wheat straw of Cellulase, alphaGlucosidase and Xylananse from T. reesei, A. niger, and T. longibrochiatum after various pre-treatments. The high yields and mild conditions are attractive for commercial applications. These enzyme structures are complex and can be found in various bacteria as organized supramolecular complexes called cellulosomes (Bayer et al., 1998). These enzymes are commonly found in fungi such as Trichoderma reesei and Aspergillus niger and in bacteria such as Clostridium cellulovorans (Arai et al., 2006). These source organisms are either aerobic or anaerobic and Sugar Yield in the Enzymatic Hydrolysis of Wheat Straw after Various Pretreatments Pretreatment Enzymes Mixture Source of Enzyme Hydrolysis Condition Sugar Yield, g/g-DM % Max. Theoretical Dilute H2SO4 Impregnation þ Cellulase a-Glucosidase T. reesei A. niger 40  C; pH 5.0 96 h 0.612 99.6 0.75% (v/v) H2SO4, 121  C Cellulase a-Glucosidase Xylananse T. reesei A. niger T. longibrochiatum 45  C; pH 5.0 72 h 0.565 74 2.15% (v/v) H2O2 35  C Cellulase a-Glucosidase Xylananse T. reesei A. niger T. longibrochiatum 45  C; pH 5.0 120 h 0.672 96.7 Fine Grinding þ Wet Oxidation Cellulase a-Glucosidase T. reesei A. niger 50  C; pH 5.0 24 h 0.638 92 Source: Talebnia et al., 2010.
BIOCONVERSIONdCONVERTING SUGARS TO PRODUCTS are either mesophilic or thermophilic. Commercial production of cellulase is focused on fungal sources because bacterial sources tend to be anaerobic and thus are slow to grow (Duff and Murray, 1996). It appears that at least three classes of enzymes act together, synergistically, to hydrolyze cellulose: endocellulase, exocellulase, and cellobiase. Endocellulase (EC 3.2.1.4) randomly breaks internal (b-D-1,4) bonds at amorphous sites that create new chain ends. Exocellulase (EC 3.2.1.91) cleaves two to four units from the ends of the exposed chains produced by the endocellulase and results in tetrasaccharides or disaccharides. Lastly, the cellobiase (EC 3.2.1.21), otherwise known as b-glucosidase, hydrolyzes the exocellulase products into individual monosaccharides (Coughlan and Ljungdahl, 1988; Galbe and Zacchi, 2002; Rabinovich et al., 2002; Zhang et al., 2006). The cellulase action occurs in three steps. The first is adsorption of cellulase onto the surface of the cellulose. The second is biodegradation of cellulose into fermentable sugars. Lastly, desorption of cellulase occurs completing the catalytic cycle. Enzyme activity is affected by a variety of environmental and substantive conditions. Temperature and pH are known to affect enzyme activity. Most cellulose enzymes show an optimum activity at temperatures in the range of 45e55  C and at pH values between 4 and 5 (Galbe and Zacchi, 2002). For LB applications, the optimum pH is shifted upward to between 5 and 6.5 due to the presence of lignin in the system (Lucas et al., 2012). These are mild operational conditions. These mild conditions lower the overall operational costs compared to purely chemical hydrolysis methods. Additionally, substrate concentration, product concentration, activators, inhibitors and cellulose structure are also significant determiners of enzyme effectiveness (Detroy and Julian, 1982). Cellobiase is itself an inhibitor to endo- and exocellulases. Thus, the b-glucosidase activity is crucial for the efficiency of the hydrolysis process (Coughlan and Ljungdahl, 1988; Galbe and Zacchi, 2002; Rabinovich et al., 2002; Zhang et al., 2006). The structure of cellulose affects the rate of hydrolysis. The cellulose features known to affect the rate of hydrolysis include (1) molecular structure of cellulose, (2) crystallinity of cellulose, (3) surface area of cellulose fiber, (4) degree of swelling of cellulose fiber, (5) DP, and (6) associated lignin or other materials (Detroy and Julian, 1982). The purer and more refined the cellulose is, the more ideal the cellulase activity will be. Higher enzyme activity lowers the enzyme load and cost for the enzymatic hydrolysis process. Lastly, even under ideal conditions, the activity of the cellulase enzyme is affected by the age of the enzyme itself. The overall activity of the enzyme decreases rapidly 477 and slows the rate of enzymatic hydrolysis. There is currently much research devoted to improving the overall yield and maintaining a high rate of hydrolysis (Sun and Cheng, 2002). Supplementing cellulase enzymes with other enzymes is another area of current focus. Conjugating the action of cellulases and hemicellulases is known to increase the rate of enzymatic hydrolysis and result in an overall higher sugar yield. Cellulose is a homopolysaccharide, hemicelluloses are heteropolysaccharides. To obtain a more complete hydrolysis of LB one must consider a multiple-enzyme system and reap the yield of the combined activities. BIOCONVERSIONdCONVERTING SUGARS TO PRODUCTS Following hydrolysis, converting the resultant sugars to products is the next step. Fermentation is a biological option and is the focus of this section. Both chemical/catalytic and biochemical conversions are common. At this point, the pretreatment and hydrolysis activities were designed and executed all with the intent of optimizing and preparing for fermentation, the capstone process of the bioconversion (Gamage et al., 2010). Fermentation is referred to as anaerobic digestion. Fermentation is the chemical breakdown of a substance by bacteria, yeasts, or other microorganisms to produce ethanol or other alcohols, lactic acid, lactose, and hydrogen (Chandel et al., 2007; Wheals et al., 1999). One of the most significant factors in fermentation is the choice of organism or modification to an organism to acquire a desired product. Some organisms only metabolize hexoses while others may metabolize both hexoses and pentoses. Saccharomyces cerevisiae is an old and very popular strain of yeast used throughout the food and fuel industries. When added to a batch of material, it will metabolize the glucose component, almost exclusively, into ethanol and carbon dioxide. It will generally follow the Embden-Myerhof pathway under anaerobic conditions when the temperature is controlled around 30  C (Limayem and Ricke, 2012). S. cerevisiae grows optimally at this temperature and it also resists high osmotic pressure and it is tolerant of pH as low as 4.0 and it is tolerant of many inhibitory products (Hahn-Hagerdal et al., 2007). S. cerevisiae remains popular because of its high ethanol yield from hexose sugars; it generates 12.0e17.0% w/v, which is 90% of the theoretical maximum (Bayrock and Ingledew, 2001; Claassen et al., 1999). Despite all its great characteristics, S. cerevisiae cannot metabolize both hexoses and pentoses and thus it is not a great organism for converting LB. In
478 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES LB, there is a significant portion of the hydrolysate containing hemicelluloses, pentose sugars such as D-xylose, which may potentially enhance yields (Martin et al., 2002). Identifying and employing an optimal organism is a great opportunity in fermentation. The optimal organism ought to be high yielding, able to metabolize both hexose and pentose sugars, tolerant to high ethanol concentration and tolerant to chemical inhibitors left over from pretreatment and hydrolysis. There are numerous naturally occurring organisms that possess a subset of these characteristics, but none are ideal. To develop a more advantageous organism one might have to genetically modify an organism to achieve one’s goals. Table 27.7 lists several naturally occurring organisms and their features and liabilities (Limayem and Ricke, 2012). Reducing operating costs and product inhibition is another important goal. There are strategies that combine hydrolysis and fermentation together. Simultaneous saccharification and fermentation (SSF) is one strategy that has just that in mind. The needed enzyme(s) and the corresponding organisms are added together so that enzymatic saccharification of cellulose and subsequent fermentation of the resultant sugars takes place at the same time in the same reactor (Dowe and McMillan, 2008). However, SSF requires an overall compromise between saccharification and fermentation, usually resulting in a less optimum operation. Another strategy is to employ an organism that is capable of making its own enzymes for hydrolysis and of fermenting the resultant sugars. Consolidated bioprocessing lowers the cost of bioconversion by reducing enzymatic saccharification and fermentation into a single step and eliminates the need for cellulase enzymes (Ladisch et al., 2010; Lynd et al., 2005). Despite the number of prokaryotic and eukaryotic microorganisms that convert sugars to ethanol, most remain limited in terms of cofermentation, ethanol yields, and tolerance to chemical inhibitors, high temperatures and ethanol. THERMOCHEMICAL CONVERSION The process of converting LB to products using primarily heat as the engine of conversion is thermochemical conversion. Thermochemical processing appears more promising than bioconversion of the lignin fraction of the LB in that it serves as a source of process energy and the coproducts have benefits in a life-cycle context; however, it has a detrimental effect on enzymatic hydrolysis (Lynd et al., 1999, 2005; Lynd and Wang, 2004). This method differentiates on how much air is supplied to the conversion, as shown in Figure 27.7. If LB is heated in the presence of excessive amounts of air, specifically oxygen, then the biomass will combust. If the amount of air or elements of air is limited then gasification will occur. Lastly, if no air is allowed then pyrolysis or hydrothermal liquefaction is the outcome. Combustion Combustion is a result of a complicated network of exothermic chemical reactions. The reaction generates copious amounts of heat and radiation. The reaction tends to be self-perpetuating and continues spontaneously due to the large amount of heat generated by the reaction. Specifically, combustion is when carbon, hydrogen, combustible sulfur, and nitrogen in LB react with oxygen. The process includes fusion, evaporation, pyrolysis, a gas phase, and surface reactions. Combustion of solids can take place in many forms including evaporation combustion, decomposition combustion, surface combustion, and smoldering combustion. Evaporation combustion is when fuel, with a relatively low fusing point, fuses and evaporates by heating, and reacts with gaseous oxygen and burns. When gasses such as H2, CO, CmHn, H2O, and CO2 are produced by thermal decomposition and react with O2, it is called decomposition combustion. A common by-product of evaporation and decomposition combustion is char. Char burns by surface combustion. Surface combustion occurs when the material only contains carbon and small amounts of volatile compounds such as charcoal, oxygen, carbon dioxide, or steam. When these compounds diffuse into pores inside or on the surface of the solid they burn in a reaction of the surface of the material. Lastly, smoldering combustion is a slower and lower temperature reaction. It is a form of thermal decomposition that takes place at a temperature below the ignition temperature of the volatile components of the LB. If the temperature is forced up, smoke might be produced or the reaction may ignite. If it ignites the reaction is referred to as flammable combustion. Industrial LB conversion processes commonly employ decomposition or surface combustion. Gasification Gasification is the conversion of LB from its solid form to fuel gas or syngas. Syngas is simply a chemical feedstock in its gas phase. These might be CO, H2, CH4, CO2 and N2 as well as char (Balat, 2008b; Demirbas, 2004; Li and Suzuki, 2009). Gasification is a broad treatment method and produces a variety of different results depending on how it is controlled. Manipulating pressure, temperature, heating method, and conversion agent produces specific results. Pressure is usually controlled for either normal pressure (0.1e0.12 MPa) or high-pressure conditions
TABLE 27.7 Advantages and Drawbacks of Potential Organisms in Lignocellulosic-Based Bioethanol Fermentation Characteristics Advantages Drawbacks References Saccharomyces cerevisiae Facultative anaerobic yeast • Naturally adapted to ethanol fermentation • High alcohol yield (90%) • High tolerance to ethanol (up to 10% v/v) and chemical inhibitors • Amenability to genetic modifications • Not able to ferment xylose and arabinose sugars • Not able to survive high temperature of enzyme hydrolysis (Gamage et al., 2010; HahnHagerdal et al., 2007; Jorgensen, 2009; McMillan, 1994; Rogers et al., 2007; Talebnia et al., 2010) Candida shehatae Microaerophilic yeast • Ferment xylose • Low tolerance to ethanol • Low yield of ethanol • Require microaerophilic conditions • Does not ferment xylose at low pH (Banerjee et al., 2009; Ligthelm et al., 1988; McMillan, 1994; Sun et al., 2011; Zaldivar et al., 2001) Zymomonas mobilis Ethanologenic gram-negative bacteria • Ethanol yield surpasses S. cerevisiae (97% of the theoretical) • High ethanol tolerance (up to 14% v/v) • High ethanol productivity (fivefold more than S. cerevisiae volumetric productivity) • Amenability to genetic modification • Does not require additional oxygen • Not able to ferment xylose sugars • Low tolerance to inhibitors • Neutral pH range (Balat and Balat, 2008; Herrero, 1983; Liu et al., 2010; McMillan, 1994) Pichia stipites Facultative anaerobic yeast • Best performance xylose fermentation • Ethanol yield (82%) • Intolerant to a high concentration of ethanol above 40 g/L (Jeffries et al., 2007; McMillan, 1994; Nigam, 2001; Shupe and Liu, 2012; Zaldivar et al., 2001) 479 (Continued) THERMOCHEMICAL CONVERSION Species
Advantages and Drawbacks of Potential Organisms in Lignocellulosic-Based Bioethanol Fermentationdcont’d Species Characteristics Drawbacks • Able to ferment most of cellulosic-material sugars including glucose, galactose, and cellobiose • Possess cellulase enzymes favorable to SSF process • Does not ferment xylose at low pH • Sensitive to chemical inhibitors • Requires microaerophilic conditions to reach peak performance • Reassimilates formed ethanol • Low yield of ethanol • Require microaerophilic conditions • Does not ferment xylose at low pH References Pachysolen tannophilus Aerobic fungus • Ferment xylose Escherichia coli Mesophilic gram-negative bacteria • Ability to use both pentose and hexose sugars • Amenability for genetic modifications • Repression catabolism interfere to cofermentation • Limited ethanol tolerance • Narrow pH and temperature growth range • Production of organic acids • Genetic stability not proved yet • Low tolerance to inhibitors and ethanol (Gamage et al., 2010; Liu et al., 2010; Weber and Boles, 2010; Zayed and Meyer, 1996) Kluveromyces marxianus Thermophilic yeast • Able to grow at a high temperature above 52  C • Suitable for SSF/CBP process • Reduces cooling cost • Reduces contamination • Ferments a broad spectrum of sugars • Amenability to genetic modifications • Excess of sugars affect its alcohol yield • Low ethanol tolerance • Fermentation of xylose is poor and leads mainly to the formation of xylitol (Banat et al., 1992; Kumar et al., 2009b; Weber et al., 2010) Thermophilic bacteria • Thermoanaerobacterium saccharolyticum • Thermoanaerobacter ethanolicus • Clostridium thermocellum Extreme anaerobic bacteria • Resistance to an extremely high temperature of 70  C • Suitable for SSCombF/CBP processing • Ferment a variety of sugars • Display cellulolytic activity • Amenability to genetic modification • Low tolerance to ethanol (Georgieva et al., 2008; Kumar et al., 2009b; Lynd et al., 2002; Shaw et al., 2008; Zeikus et al., 1981) CBP, consolidated bioprocessing. (Zaldivar et al., 2001; Zayed and Meyer, 1996) 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES Advantages 480 TABLE 27.7
481 THERMOCHEMICAL CONVERSION FIGURE 27.7 Overview of thermochemical conversion technologies. Source: Bain, 2004. Thermochemical conversion Excess air combustion Partial air gasification No air pyrolysis hydrothermal liquefaction (0.5e2.5 MPa). Temperatures are usually either operated under low-temperature (<700  C) or high-temperature (>700  C) conditions. High-temperature decomposition may occur at the ash fusion point or above. Indirect gasification occurs when heating the raw material and gasification agent using an external heat source. Direct gasification occurs when heat generated from the reaction of partial gasification of raw material and oxygen is used as the heating source. The gasification agent is another variable with significant influence on the product. An agent is any combination of air, oxygen, or steam. Additionally, carbon dioxide may be used for a set period of time to affect the product. There are a variety of gasifiers in use today. Fixed bed, flow bed, circulating flow bed, entrained bed, mixing bed, rotary kiln, twin tower, and molten furnace are examples of industrial gasifiers (Yokoyama, 2008). Another method, supercritical water gasification (SCWG), is interesting because water under supercritical conditions has properties that facilitate the transport processes of compounds while remaining a benign media for processing. It even acts as a catalyst for acid/ base reactions under these conditions (Calzavara et al., 2005; Nolen et al., 2003; Savage et al., 1995). Of note in SCWG is that it takes place in either high- or lowtemperature conditions (Matsumura et al., 2005). However, if one adds an alkali catalyst to the processing at low temperatures, it increases the conversion into oil and gas. Additionally, the catalyst lowers the temperature at which the cellulose decomposes, or the onset of the gasification process (Minowa et al., 1997, 1995, 1998a,b, Murakami et al., 2012). Pyrolysis The conversion of LB by heating is pyrolysis (Balat, 2008a; Bridgwater, 2003; Mohan et al., 2006). Depending on the desired product, one chooses the operational conditions for pyrolysis. Factors such as heating rate, reactor temperature, residence time and composition of the feedstock determine the product. The goal of pyrolysis is to execute the process in the absence of oxygen and thus avoid the burning of the feedstock and instead break down the lignocellulosic bonds and crystalline structure. By doing so under these conditions, new compounds are formed. Compounds such as char, bio-oil, and gasses are produced (Thuijl et al., 2003). The biooil formed by pyrolysis is not easily stored because it is unstable and requires additional processing prior to long-term storage (Adam, 2005). There are three categories of pyrolysis: conventional, fast and flash. Conventional pyrolysis produces solids, liquids and gasses. Fast and flash pyrolysis produces primarily liquid and gaseous products (Demirbas, 2005). Direct Liquefaction Direct conversion of LB to biofuels is liquefaction. Typical products include biodiesel and heavy oils that are typically very viscous. Adding an alkali to the conversion will enhance the liquefaction process (Itoh et al., 1994; Demirbas, 2005). Hydrothermal liquefaction is an application where thermal depolymerization, or hydrous pyrolysis, is accomplished using superheated water under pressure. Drying feedstock in unnecessary when using hydrothermal liquefaction. Consequently, it is suitable for converting any biomass regardless of its moisture content. Aquatic biomass, garbage, organic sludge as well as LB are all good feedstock candidates for hydrothermal liquefaction. At the boiling point of water, 100  C, extraction of aqueous soluble components is possible. At temperatures above 150  C hydrolysis begins and biomass polymers, such as cellulose, hemicellulose, proteins, and so on, degrade into monomers. Then at 200  C and 1 MPa solidlike biomass changes into a slurry, a process called liquidization. At higher temperatures below the critical point of water, around 300  C and 10 MPa, liquefaction takes place and oily product is obtained. If one changes the reaction conditions such as reaction time or the catalyst, the main product can be changed to char, a process referred to as hydrothermal carbonization. Finally, at a temperature around the critical point and in the presence of a catalyst, the biomass will gasify.
482 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES CONCLUSION Utilizing LB for energy and fuel production is as old as mankind. However, modern sources of fuel are more cost-effective and convenient to drive and meet our contemporary energy demands. Liberating carbon that has accumulated over millennia in such an unnaturally condensed period of time threatens to alter current climate conditions. Our economic engine has created dependencies on fossil fuels and has encouraged unhealthy relationships between highly industrialized societies like the United States and China and some energy suppliers around the world. The cycle of growing and using LB for fuel and energy is a closed and therefore, sustainable system that consumes as much carbon dioxide as is liberated. Finding technologies that can return modernized societies around the globe to more sustainable and self-reliant sources of energy is critical to reduce the environmental impact and improve national security and sovereignty. To that end, this chapter has presented and reviewed a wide array of biochemical and thermochemical LB conversion options. These methods have developed into capable methods of converting LB into fermentable intermediates such as sugars or products. Much development has already taken place and there is still much more needed until utilizing LB for energy and fuel can compete with and replace more convenient and less expensive sources of fuel. Industrializing the process of converting LB into valuable materials is the function of a biorefinery and it employs a complex and diverse set of conversion technologies to accomplish its tasks. There are several pretreatment options available that either exploit thermal, mechanical and chemical mechanisms or use biological and chemical mechanisms. Ultimately the treatments chosen will depend on the desired product and desired specifications. These products may include liquid biofuels, biochemicals as well as steam, heat and electricity. Regardless, the objectives of the pretreatment are the same: to break down the strong crystalline lignin and cellulosic structures such that the biomass becomes vulnerable to conversion processes, such as hydrolysis treatment that yields fermentable sugars that are free from microbial growth inhibitors, or a thermochemical conversion such as gasification or liquefaction. No treatment option is ideal. There are often trade-offs between competing features and liabilities such as high yield and long retention time, low concentration of inhibitors and high cost, significant environmental and safety concerns and high efficacy. The biochemical and thermochemical pretreatments and conversions have developed and improved over time independently and in conjunction with a series of pretreatment methods. The goal of the development of these pretreatments is to improve the quantity and quality of materials available to conversion, whether it is bioconversion or thermochemical conversion. In order to compete with contemporary and convenient sources of energy it is important to maximize the pretreatment and conversion processes as well as to eliminate and recycle waste and energy. The biorefinery holds great promise to enable the usage of biomass as a sustainable and reliable source of energy and fuel. The biochemical and thermochemical conversion options described here are by no means an exhaustive representation of all the studied methods. Many more exist in the literature and maybe the best methods are yet to come. References Adam, J., 2005. Catalytic Conversion of Biomass to Produce Higher Quality Liquid Bio-fuels. Norwegian University of Science and Technology, Trondheim, Norway, pp. 142. Alvira, P., Tomas-Pejo, E., Ballesteros, M., Negro, M.J., 2010. Pretreatment technologies for an efficient bioethanol production process based on enzymatic hydrolysis. Bioresour. Technol. 101, 4851e4861. Amidon, T.E., Wood, C.D., Shupe, A.M., Wang, Y., Graves, M., Liu, S., 2008. Biorefinery: conversion of woody biomass to chemicals, energy and materials. J. Biobased Mater. Bioenergy 2, 100e120. Arai, T., Kosugi, A., Chan, H., Koukiekolo, R., Yukawa, H., Inui, M., Doi, R.H., 2006. Properties of cellulosomal family 9 cellulases from Clostridium cellulovorans. Appl. Microbiol. Biotechnol. 71, 654e660. Auslender, V.L., Ryazantsev, A.A., Spiridonov, G.A., 2002. The use of electron beam for solution of some ecological problems in pulp and paper industry. Radiat. Phys. Chem. 63, 641e645. Azzam, A.M., 1989. Pretreatment of cane bagasse with alkaline hydrogen peroxide for enzymatic hydrolysis of cellulose and ethanol fermentation. J. Environ. Sci. Health, Part B 24, 421e433. Bain, R.L., 2004. An introduction to biomass thermochemical conversion. In: Laboratory, N.R.E. (Ed.), DOE/NASLUGC Biomass and Solar Energy Workshops. US Department of Energy. Balat, M., 2008a. Mechanisms of thermochemical biomass conversion process. Part 1: reactions of pyrolysis. Energy Sources, Part A 30, 620. Balat, M., 2008b. Mechanisms of thermochemical biomass conversion processes. Part 2: reactions of gasification. Energy Sources, Part A 30, 636. Balat, M., Balat, H., 2008. Progress in bioethanol processing. Prog. Energy Combust. Sci. 34, 551e573. Ballesteros, I., Oliva, J.M., Navarro, A.A., Gonzalez, A., Carrasco, J., Ballesteros, M., 2000. Effect of chip size on steam explosion pretreatment of softwood. Appl. Biochem. Biotechnol. 84e86, 97e110. Ballesteros, I., Oliva, J.M., Negro, M.J., Manzanares, P., Ballesteros, M., 2002. Enzymic hydrolysis of steam exploded herbaceous agricultural waste (Brassica carinata) at different particle sizes. Process Biochem. 38, 187e192. Banat, I., Nigam, P., Marchant, R., 1992. Isolation of thermotolerant, fermentatitve yeasts growing at 52  C and producing ethanol at 45  C and 52  C. World J. Microbiol. Biotechnol. 8, 259e263. Banerjee, S., Mudliar, S., Sen, R., Giri, B., Satpute, D., Chakrabarti, T., Pandey, R.A., 2009. Commercializing lignocellulosic bioethanol: technology bottlenecks and possible remedies. Biofuels, Bioprod. Biorefin. 4, 77e93. Barber, V.A., 2007. State University of New York, College of Environmenatl Science and Forestry.
REFERENCES Bayer, E.A., Chanzy, H., Lamed, R., Shoham, Y., 1998. Cellulose, cellulases and cellulosomes. Curr. Opin. Struct. Biol. 8, 548e555. Bayrock, D., Ingledew, W.M., 2001. Changes in steady state on introduction of a Lactobacillus contaminant to a continuous culture ethanol fermentation. J. Ind. Microbiol. Biotechnol. 27, 39e45. Ben-Ghedalia, D., Miron, J., 1981. The effect of combined chemical and enzyme treatments on the saccharification and in vitro digestion rate of wheat straw. Biotechnol. Bioeng. 23, 823e831. Betiku, E., Adetunji, O.A., Ojumu, T.V., Solomon, B.O., 2009. A comparative study of the hydrolysis of gamma irradiated lignocelluloses. Braz. J. Chem. Eng. 26, 251. Bisaria, V., Ghose, T., 1981. Biodegradation of cellulosic materials: substrate, microorganisms, enzymes, and products. Enzyme Microb. Technol. 3, 90e104. Bjerre, A.B., Olesen, A.B., Fernqvist, T., Ploger, A., Schmidt, A.S., 1996. Pre-treatment of wheat straw using combined wet oxidation and alkaline hydrolysis resulting in convertible cellulose and hemicellulose. Biotechnol. Bioeng. 49, 568e577. Boerjan, W., Ralph, J., Baucher, M., 2003. Lignin biosynthesis. Annu. Rev. Plant Biol. 54, 519e546. Bohlmann, G.M., 2006. Process economic considerations for production of ethanol from biomass feedstocks. Ind. Biotechnol. 2, 14e20. Bouchard, J., Methot, M., Jordan, B., 2006. The effect of ionizing radiation on the cellulose of wood-free paper. Cellulose 13, 601e610. Brasch, D.J., Free, K.W., 1965. Prehydrolysis-kraft pulping of Pinus radiata grown in New Zealand. Tappi 48, 245e248. Bridgwater, A., 2003. Renewable fuels and chemicals by thermal processing of biomass. Chem. Eng. J. 91, 87e102. Brownell, H.H., Yu, E.K.C., Saddler, J.N., 1986. Steam-explosion pretreatment of wood: effect of chip size, acid, moisture content and pressure drop. Biotechnol. Bioeng. 28, 792e801. Bura, R., Chandra, R., Saddler, J., 2009. Influence of xylan on the enzymatic hydrolysis of steam-pretreated corn stover and hybrid poplar. Biotechnol. Prog. 25 (2), 315e322. Cadoche, L., Lopez, G.D., 1989. Assessment of size reduction as a preliminary step in the production of ethanol from lignocellulosic wastes. Biol. Wastes 30 (2), 153e157. Calzavara, Y., Joussot-Dubien, C., Boissonnet, G., Sarrade, S., 2005. Evaluation of biomass gasification in supercritical water process for hydrogen production. Energy Convers. Manage. 46, 615e631. Chandel, A.K., Chan, E., Rudravaram, R., Narasu, M.L., Rao, L.V., Ravindra, P., 2007. Economics and environmental impact of bioethanol production technologies: an appraisal. Biotechnol. Mol. Biol. Rev. 2, 14e32. Chaudhary, G., Singh, L.K., Ghosh, S., 2012. Alkaline pretreatment methods followed by acid hydrolysis of Saccharum spontaneum for bioethanol production. Bioresour. Technol. 124, 111e118. Chen, H.Z., Qiu, W.H., 2010. Key technologies for bioethanol production from lignocellulose. J. Biotechnol. Adv. 28, 556e562. Chen, Y., Dong, B., Qin, W., Xiao, D., 2010. Xylose and cellulose fractionation from corncob with three different strategies and separate fermentation of them to bioethanol. Bioresour. Technol. 101 (18), 6994e6999. Claassen, P.A.M., Lopez-Contreras, A.M., Sijtsma, L., Weusthuis, R.A., van Lier, J.B., van Niel, E.W.J., Stams, A.J.M., De Vries, S.S., 1999. Utilization of biomass for the supply of energy carriers. Appl. Microbiol. Biotechnol. 52, 741e755. Conner, A.H., Lorenz, L.F., 1986. Kinetic modeling of hardwood prehydrolysis. 3. Water and dilute acetic-acid prehydrolysis of southern red oak. Wood Fiber Sci. 18 (5), 248e263. Coughlan, M.P., Ljungdahl, L.G., 1988. Comparative biochemistry of fungal and bacterial cellulolytic enzyme system. FEMS Symp. 43, 11e30. 483 Cullis, I.F., Saddler, J.N., Mansfield, S.D., 2004. Effect of initial moisture content and chip size on the bioconversion efficiency of softwood lignocellulosics. Biotechnol. Bioeng. 85, 413e421. Cybulska, I., Brudecki, G., Rosentrater, K., Julson, J.L., Lei, H., 2012. Comparative study of organosolv lignin extracted from prairie cordgrass, switchgrass and corn stover. Bioresour. Technol. 118, 30e36. D’Andola, G., Szarvas, L., Massonne, K., Stegmann, V., 2008. Ionic liquids for solubilizing polymers. WO Pat. 2008/043837. Dale, B.E., Henk, L.L., Shiang, M., 1984. Fermentation of lignocellulosic materials treated by ammonia freeze-explosion. Dev. Ind. Microbiol. 26, 223e233. Dashtban, M., Schraft, H., Qin, W., 2010. Fungal bioconversion of lignocellulosic residues: opportunities & perspectives. Int. J. Biol. Sci. 5, 578e595. Demirbas, A., 2004. Current technologies for thermo-conversion of biomass into fuels and chemicals. Energy Sources 26, 715e730. Demirbas, A., 2005. Bioethanol from cellulosic materials: a renewable motor fuel from biomass. Energy Sources 21, 327e337. Detroy, R.W., Julian, G.S., 1982. Biomass conversion: fermentation chemicals and fuels. Crit. Rev. Microbiol. 10, 203e228. Dowe, N., McMillan, J., 2008. SSF experimental protocols - lignocellulosic biomass hydrolysis and fermentation. In: Office of Energy Efficiency and Renewable Energy NREL. Midwest Research Insitute, Battelle, Golden, CO. Drapcho, C., Nhuan, N., Wlaker, T., 2008. Biofuel Engineering Process Technology. McGraw Hill Companies, Inc, New York, NY, pp. 371. Dubey, K.A., Pujari, P.K., Rammani, S.P., Kadam, R.M., Sabharwal, S., 2004. Microstructural studies of electron beam-irradiated cellulose pulp. Radiat. Phys. Chem. 69, 395. Duff, S.J.B., Murray, W.D., 1996. Bioconversion of forest products industry waste cellulosics to fuel ethanol: a review. Bioresour. Technol. 55, 1e33. Esteghlalian, A., Hashimoto, A.G., Fenske, J.J., Penner, M.H., 1997. Modeling and optimization of the dilute-sulfuric-acid pretreatment of corn stover, poplar and switchgrass. Bioresour. Technol. 59, 129e136. Ewanick, S.M., Bura, R., Saddler, J.N., 2007. Acid-catalyzed steam pretreatment of lodgepole pine and subsequent enzymatic hydrolysis and fermentation to ethanol. Biotechnol. Bioeng. 98, 737e746. Fengel, D.W.G., 1984. Wood: Chemistry, Ultrastructure, Reactions. Walter de Gruyter, Berlin and New York. Fengel, D.W.G., 1989. Wood: Chemistry, Ultrastructure, Reactions. Walter de Gruyter, New York. Fort, D.A., Remsing, R.C., Swatloski, R.P., Moyna, P., Moyna, G., Rogers, R.D., 2007. Can ionic liquids dissolve wood? Processing and analysis of lignocellulosic materials with 1-n-butyl3-methylimidazolium chloride. Green Chem. 9, 63e69. Galbe, M., Zacchi, G., 2002. A review of the production of ethanol from softwood. Appl. Microbiol. Biotechnol. 59, 618e628. Gamage, J., Lam, H., Zhang, Z., 2010. Bioethanol production from lignocellulosic biomass. J. Biobased Mater. Bioenergy 4, 3e11. Gamez, S., Ramirez, J., Garrote, G., Vazquez, M., 2004. Manufacture of fermentable sugar solutions from sugarcane bagasse hydrolyzed with phosphoric acid at atmospheric pressure. J. Agric. Food Chem. 52, 4172e4177. Garcı́a-Cubero, M.A.T.M.A., González-Benito, G.G., Indacoechea, I.I., Coca, M.M., Bolado, S.S., 2009. Effect of ozonolysis pretreatment on enzymatic digestibility of wheat and rye straw. Bioresour. Technol. 100 (4), 1608e1613. Georgieva, T., Mikkelsen, M., Ahring, B., 2008. Ethanol production from wet exploded wheat straw hydrolysate by thermophilic anaerobic bacterium Thermoanaerobacter BG1L1 in a continuous immobilized reactor. Appl. Biochem. Biotechnol. 145, 99e110.
484 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES Glaudemans, C., Timell, T.E., 1958. The polysaccharides of white Brich (Betula papyrifera) IV, the constitution of the hemicellulose. J. Am. Chem. Soc. 80, 1209e1213. Goring, D., Timell, T., 1960. Molecular properties of six 4-O-methylglucuronoxylans. J. Phys. Chem. 64, 1426e1430. Gregg, D., Saddler, J.N., 1996a. A technoeconomic assessment of the pretreatment and fractionation steps of a biomass-to-ethanol process. Appl. Biochem. Biotechnol. 57, 711e727. Gregg, D.J., Saddler, J.N., 1996b. Factors affecting cellulose hydrolysis and the potential of enzyme recycle to enhance the efficiency of an integrated wood to ethanol process. Biotechnol. Bioeng. 51, 375e383. Grous, W.R., Converse, A.O., Grethlein, H.E., 1986. Effect of steam explosion pre-treatment on pore size and enzymatic hydrolysis of poplar. Enzyme Microb. Technol. 8, 274e280. Gupta, R., Lee, Y., 2009. Pretreatment of hybrid poplar by aqueous ammonia. Biotechnol. Prog. 25, 357e364. Hahn-Hagerdal, B., Karhumaa, K., Fonseca, C., Spencer-Martins, I., Gorwa-Grauslund, M.F., 2007. Towards industrial pentosefermenting yeast strains. Appl. Microbiol. Biotechnol. 74, 937e953. Hamelinck, C.N., van Hooijdonk, G., Faaij, A.P.C., 2003. Prospects for Ethanol from Lignocellulosic Biomass: Techno-economic Performance as Development Progresses. Utrecht University. Report nr NWS-E-2003-55, p. 35. Harun, R., Jason, W.S.Y., Cherrington, T., Danquah, M.K., 2010. Biomass as a cellulosic fermentation feedstock for bioethanol production. Renewable Sustainable Energy Rev. http://dx.doi.org/ 10.1016/j.rser.2010.07.071. Hendriks, A.T., Zeeman, G., 2009. Pretreatments to enhance the digestibility of lignocellulosic biomass. Bioresour. Technol. 100, 10e18. Herrero, A., 1983. End product inhibition in anaerobic fermentation. Trends Biotechnol. 1, 49e53. Hinman, N.D., Schell, D.J., Riley, C.J., Bergeron, P.W., Walter, P.J., 1992. Preliminary estimate of the cost of ethanol production for SSF technology. Appl. Biochem. Biotechnol. 34, 639e649. Holtzapple, M.T., Humphrey, A.E., Taylor, J.D., 1989. Energy requirements for the size reduction of poplar and aspen wood. Biotechnol. Bioeng. 33, 207e210. Holtzapple, M.T., Jun, J. H., Ashok, G., Patibandla, S.L., Dale, B.E., 1990. Ammonia fiber explosion (AFEX) pretreatment of lignocellulosic wastes. In: Proceedings of the American Institute of Chemical Engineers National Meeting. Chicago, IL, USA. Holtzapple, M.T., Jun, J.H., Ashok, G., Patibandla, S.L., Dale, B.E., 1991. The ammonia freeze explosion (AFEX) process: a practical lignocellulosic pretreatment. Appl. Biochem. Biotechnol. 28, 59e74. Holtzapple, M.T., Davison, R.R., Stuart, E.D., 1992. Biomass refining process. USA Patent 5, 171, 592. Hon, D.N.S., Shiraishi, N., 2001. In: Wood and Cellulose Chemistry, second ed. Marcel Dekker Inc, New York. Hsu, T.A., 1996. Pretreatment of biomass. In: Wyman, C.E. (Ed.), Handbook on Bioethanol, Production, and Utilization. Taylor & Francis, Washington, D. C. Hu, G., Heitmann, J.A., Rojas, O.J., 2008. Feedstock pretreatment strategies for producing ethanol from wood, bark and forest residues. Bioresources 3 (1), 270e294. Hunter, C.A., Sanders, K.M., 1990. The nature of pep interaction. J. Am. Chem. Soc. 112, 5525e5534. Itoh, S., Suzuki, A., Nakamura, T., Yokoyama, S., 1994. Production of heavy oil from sewage sludge by direct thermochemical liquefaction. Proceedings of the IDA and WRPC World Conference on Desalination and Water Treatment. 98 (1e3), 127e133. Iyer, P.V., Wu, Z.W., Kim, S.B., Lee, Y.Y., 1996. Ammonia recycled percolation process for pretreatment of herbaceous biomass. Appl. Biochem. Biotechnol. 57-58, 121e132. Jeffries, T.W., Grigoriev, I.V., Grimwood, J., Laplaza, J.M., Aerts, A., Salamov, A., 2007. Genome sequence of the lignocellulose bioconverting and xylose-fermenting yeast Pichia stipitis. Nat. Biotechnol. 25 (3), 319e326. Jin, M., Gunawan, C., Balan, V., Lau, M.W., Dale, B.E., 2012. Simultaneous saccharification and co-fermentation (SSCF) of AFEX(TM) pretreated corn stover for ethanol production using commercial enzymes and Saccharomyces cerevisiae 424A(LNH-ST). Bioresour. Technol. 110, 587e594. Jorgensen, H., 2009. Effect of nutrients on fermentation of pretreated wheat straw at very high dry matter content by Saccharomyces cerevisiae. Appl. Biochem. Biotechnol. 153, 44e57. Karr, W.E., Holtzapple, T., 2000. Using lime pretreatment to facilitate the enzymatic hydrolysis of corn stover. Biomass Bioenergy 18, 189e199. Karthika, K., Arun, A.B., Rekha, P.D., 2012. Enzymatic hydrolysis and characterization of lignocellulosic biomass exposed to electron beam irradiation. Carbohydr. Polym. 90 (2), 1038e1045. Kemppainen, K., Inkinen, J., Uusitalo, J., Nakari-Setälä, T., Siikaaho, M., 2012. Hot water extraction and steam explosion as pretreatments for ethanol production from spruce bark. Bioresour. Technol. 117, 131e139. Khan, A.W., Labrie, J.P., McKeown, J., 1986. Effect of electron-beam irradiation pretreatment on the enzymatic hydrolysis of softwood. Biotechnol. Bioeng. 28, 1449e1453. Kilpeläinen, I., Xie, H., King, A., Granstrom, M., Heikkinen, S., Argyropoulos, D.S., 2007. Dissolution of wood in ionic liquids. J. Agric. Food Chem. 55, 9142. Kim, S., Dale, B., 2005. Global potential bioethanol production from wasted crops and crop residues. Biomass Bioenergy 29, 361e375. Kim, S., Holtzapple, M., 2006. Delignification kinetics of corn stover in lime pretreatment. Bioresour. Technol. 97, 778e785. Kim, K.H., Hong, J., 2001. Supercritical CO2 pretreatment of lignocellulose enhances enzymatic cellulose hydrolysis. Bioresour. Technol. 77, 139e144. Klass, D., 1998. Biomass for Renewable Energy, Fuels, and Chemicals. Academic Press, New York. Klinke, H., Thomsen, A., Ahring, B., 2004. Inhibition of ethanolproducing yeast and bacteria by degradation products produced during pretreatment of biomass. Appl. Microbiol. Biotechnol. 66, 10e26. Koo, B.-W., Min, B.-C., Gwak, K.-S., Lee, S.-M., Choi, J.-W., Yeo, H., Choi, I.-G., 2012. Structural changes in lignin during organosolv pretreatment of Liriodendron tulipifera and the effect on enzymatic hydrolysis. Biomass Bioenergy 42, 24e32. Koshijima, T., Timell, T.E., Zinbo, M., 1965. The number-average molecular weight of native hardwood xylans. J. Polym. Sci. Part C: Polym. Symp. 11, 265e279. Kumar, A., Singh, L.K., Ghosh, S., 2009a. Bioconversion of lignocellulosic fraction of water-hyacinth (Eichhornia crassipes) hemicellulose acid hydrolysate to ethanol by Pichia stipitis. Bioresour. Technol. 100 (13), 3293e3297. Kumar, S., Singh, S., Mishra, I., Adhikari, D., 2009b. Recent advances in production of bioethanol from lignocellulosic biomass. Chem. Eng. Technol. 32, 517e526. Kumar, L., Chandra, R., Chung, P.A., Saddler, J.N., 2010. Can the same steam pretreatment conditions be used for most softwoods to achieve good, enzymatic hydrolysis and sugar yields? Bioresour. Technol. 101, 7827e7833. Kuznetsov, B.N., Kuznetsova, S.A., Danilov, V.G., Kozlov, I.A., Taraban’ko, V.E., Ivanchenko, N.M., Alexandrova, N.B., 2002. New catalytic processes for a sustainable chemistry of cellulose production from wood biomass. Catal. Today 75, 211e217.
REFERENCES Ladisch, M., Kohlmann, K., Westgate, P., Weil, J. Yang, Y., 1998. Processes for treating cellulosic material. USA Patent 5, 846, 787. Ladisch, M., Mosier, N., Kim, Y., Ximenes, E., Hogsett, D., 2010. Converting cellulose to biofuels. SBE special supplement biofuels. CEP 106 (3), 56e63. Larson, K.A., King, M.L., 1986. Evaluation of supercritical fluid extraction in the pharmaceutical industry. Biotechnol. Prog. 2, 73e82. Laser, M., Schulman, D., Allen, S.G., Lichwa, J., Antal, M.J., Lynd, L.R., 2002. A comparison of liquid hot water and steam pretreatments of sugarcane bagasse for bioconversion to ethanol. Bioresour. Technol. 81, 33e44. Lau, M., Dale, B., 2008. Cellulosic ethanol production from AFEX-treated corn stover using Saccharomyces cerevisiae 424(LNH-ST). Proc. Natl. Acad. Sci. USA. 106 (5), 1368e1373. Lee, J., 1997. Biological conversion of lignocellulosic biomass to ethanol. J. Biotechnol. 56 (1), 1e24. Lee, S.H., Doherty, T.V., Linhardt, J.S., 2009. Ionic liquid-mediated selective extraction of lignin from wood leading to enhanced enzymatic cellulose hydrolysis. Biotechnol. Bioeng. 102, 1368e1376. Li, C., Suzuki, K., 2009. Tar property, analysis, reforming mechanism and model for biomass gasification - an overview. Renewable Sustainable Energy Rev. 13, 594. Li, X., Luo, X., Li, K., Zhu, J.Y., Fougere, J.D., Clarke, K., 2012. Effects of SPORL and dilute acid pretreatment on substrate morphology, cell physical and chemical wall structures, and subsequent enzymatic hydrolysis of lodgepole pine. Appl. Biochem. Biotechnol. 168 (6), 1556e1567. Liebert, T., 2010. Cellulose solvents-remarkable history, bright future, cellulose solvents: for analysis, shaping and chemical modification. In: Liebert, T., Heinze, T., Edgar, K. (Eds.). ACS Symposium Series, vol. 1033. American Chemical Society, Washington, DC, USA, pp. 3e54. Ligthelm, M.E., Prior, B.A., Preez, J.C., 1988. The oxygen requirements of yeasts for the fermentation of D-xylose and D-glucose to ethanol. Appl. Microbiol. Biotechnol. 28 (1), 63e68. Limayem, A., Ricke, S.C., 2012. Lignocellulosic biomass for bioethanol production: current perspectives, potential issues and future prospects. Prog. Energy Combust. Sci. 38 (4), 449e467. Liu, S., 2003. Chemical kinetics of alkaline peroxide brightening of mechanical pulps. Chem. Eng. Sci. 58, 2229e2244. Liu, S., 2008. A Kinetic model on autocatalytic reactions in woody biomass hydrolysis. J. Biobased Mater. Bioenergy 2 (5), 135e147. Liu, S., 2010. Woody biomass: niche position as a source of sustainable renewable chemicals and energy and kinetics of hot-water extraction/hydrolysis. J. Biotech. Adv. 28, 563e582. Liu, S., 2012. Woody biomass conversion: sustainability and waterbased processes. J. Bioprocess Eng. Biorefin. 1 (1), 6e32. Liu, C., Wyman, C.E., 2003. The effect of flow rate of compressed hot water on xylan, lignin and total mass removal from corn stover. Ind. Eng. Chem. Res. 42, 5409e5416. Liu, S., Mishra, G., Amidon, T.E., Gratien, K., 2009. Effect of hot-water extraction of woodchips on the kraft pulping of Eucalyptus woodchips. J. Biobased Mater. Bioenergy 3 (6), 363e372. Liu, T., Lin, L., Sun, Z., Hu, R., Liu, S., 2010. Bioethanol fermentation by recombinant E. coli FBR5 and its robust mutant FBHW using hot-water wood extract hydrolyzate as substrate. Biotechnol. Adv. 28 (5), 602e608. Lloyd, T., Wyman, C., 2005. Combined sugar yields for dilute sulfuric acid pretreatment of corn stover followed by enzymatic hydrolysis of the remaining solids. Bioresour. Technol. 96, 1967e1977. Lucas, M., Hanson, S.K., Wagner, G.L., Kimball, D.B., Rector, K.D., 2012. Evidence for room temperature delignification of wood using hydrogen peroxide and manganese acetate as a catalyst. Bioresour. Technol. 119, 174e180. 485 Lynd, L.R., 1996. Overview and evaluation of fuel ethanol from cellulosic biomass: technology, economics, the environment, and policy. Annu. Rev. Energy Environ. 21, 403e465. Lynd, L.R., Wang, M., 2004. A product-nonspecific framework for evaluating the potential of biomass-based products to displace fossil fuels. J. Ind. Ecol. 7, 17e32. Lynd, L.R., Elander, R.T., Wyman, C.E., 1996. Likely features and costs of mature biomass ethanol technology. Appl. Biochem. Biotechnol. 57, 741e761. Lynd, L.R., Gerngross, T., Wyman, C., 1999. Biocommodity engineering. Biotechnol. Prog. 15, 777e793. Lynd, L.R., Weimer, P.J., van Zyl, W.H., Pretorius, I.S., 2002. Microbial cellulose utilization: fundamentals and biotechnology. Microbiol. Mol. Biol. Rev. 66, 506e577. Lynd, L.R., van Zyl, W.H., McBride, J.E., Laser, M., 2005. Consolidated bioprocessing of cellulosic biomass: an update. Curr. Opin. Biotechnol. 16, 577e583. Martin, C., Galbe, M., Wahlbom, C.F., Hahn-Hägerdal, B., Jönsson, L.J., 2002. Ethanol production from enzymatic hydrolysates of sugarcane bagasses using recombinant xylose - utilizing Saccharomyces cerevisiae. Enzyme Microb. Technol. 31 (2), 274e282. Marzialetti, T., Olarte, M.B.V., Sievers, C., Hoskins, T.J.C., Agrawal, P.K., Jones, C.W., 2008. Dilute acid hydrolysis of loblolly pine: a comprehensive approach. Ind. Eng. Chem. Res. 47, 7131e7140. Mason, W.H., 1926. Process and apparatus for disintegration of wood and the like. USA Patent 1, 578, 609. Mason, W.H., 1928. Apparatus for the process of explosion fibration of lignocellulose material. USA Patent 1, 655, 618. Matsumura, Y., Minowa, T., Potic, B., Kersten, S., Prins, W., van Swaaij, W.P., van de Beld, B., Elliot, D.C., Neuenschwander, G.G., Kruse, A., Antal Jr, M.J., 2005. Biomass gasification in near- and super-critical water: status and prospects. Biomass Bioenergy 29, 269e292. McMillan, J.D., 1994. Pretreatment of lignocellulosic biomass. In: Himmel, M.E., Baker, J.O., Overend, R.P. (Eds.), Enzymatic Conversion of Biomass for Fuels Production. American Chemical Society, Washington, D. C., pp. 292e324. McMillan, J.D., Newman, M., Templeton, D., Mohagheghi, A., 1999. Simultaneous saccharification and cofermentation of dilute-acid pretreated yellow poplar hardwood to ethanol using xylosefermenting Zymomonas mobilis. Appl. Biochem. Biotechnol. 79, 649e665. Mes-Hartree, M., Dale, B.E., Craig, W.K., 1988. Comparison of steam and ammonia pretreatment for enzymatic hydrolysis of cellulose. Appl. Microbiol. Biotechnol. 29, 462e468. Mielenz, J.R., 2001. Ethanol production from biomass: technology and commercialization status. Curr. Opin. Microbiol. 4, 324e329. Millet, M.A., Baker, A.J., Scatter, L.D., 1976. Physical and chemical pretreatment for enhancing cellulose saccharification. Biotech. Bioeng. 6, 125e153. Minowa, T., Ogi, T., Yokoyama, S.-Y., 1995. Hydrogen production from wet cellulose by low temperature gasification using a reduced nickel catalyst. Chem. Lett. 10, 937e938. Minowa, T., Fang, Z., Ogi, T., Varhegyi, G., 1997. Liquefaction of cellulose in hot compressed water using sodium carbonate: products distribution at different reaction temperatures. J. Chem. Eng. Jpn. 30, 186e190. Minowa, T., Fang, Z., Ogi, T., Varhegyi, G., 1998a. Decomposition of cellulose and glucose in hot compressed water under catalyst-free conditions. J. Chem. Eng. Jpn. 31, 131e134. Minowa, T., Zhen, F., Ogi, T., 1998b. Cellulose decomposition in hot compressed water with alkali or nickel catalyst. J. Supercrit. Fluids 13, 253e259.
486 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES Mittal, A., Chatterjee, S.G., Scott, G.M., Amidon, T.E., 2009. Modeling xylan solubilization during autohydrolysis of sugar maple and aspen wood chips: reaction kinetics and mass transfer. Chem. Eng. Sci. 64, 3031e3041. Mohan, D., Pittman, C., Steele, P., 2006. Pyrolysis of wood/biomass for bio-oil: a critical review. Energy Fuels 20, 848e889. Monavari, S., Galbe, M., Zacchi, G., 2009. Impact of impregnation time and chip size on sugar yield in pretreatment of softwood for ethanol production. Bioresour. Technol. 100, 6312e6316. Monavari, S., Galbe, M., Zacchi, G., 2011. Influence of impregnation with lactic acid on sugar yields from steam pretreatment of sugarcane bagasse and spruce, for bioethanol production. Biomass Bioenergy 35, 3115e3122. Montane, D., Salvado, J., Farriol, X., Jollez, P., Chornet, E., 1994. Phenomenological kinetics of wood delignification - application of a time-dependent rate-constant and a generalized severity parameter to pulping and correlation of pulp properties. Wood Sci. Technol. 28, 387e402. Morjanoff, P.J., Gray, P.P., 1987. Optimization of steam explosion as a method for increasing susceptibility of sugarcane bagasse to enzymatic saccharification. Biotechnol. Bioeng. 29, 733e741. Mosier, N., Wyman, C.E., Dale, B., Elander, R., Lee, Y.Y., Holtzapple, M., Ladisch, M., 2005. Features of promising technologies for pre-treatment of lignocellulosic biomass. Bioresour. Technol. 96, 673e686. Murakami, K., Kasai, K., Kato, T., Sugawara, K., 2012. Conversion of rice straw into valuable products by hydrothermal treatment and steam gasification. Fuel 93, 37e43. Mvula, E.E., Naumov, S.S., von Sonntag, C.C., 2009. Ozonolysis of lignin models in aqueous solution: anisole, 1,2-dimethoxybenzene, 1,4-dimethoxybenzene, and 1,3,5-trimethoxybenzene. Environ. Sci. Technol. 43 (16), 6275e6282. Negro, M.J., Manzanares, P., Ballesteros, I., Oliva, J.M., Cabañas, A., Ballesteros, M., 2003. Hydrothermal pretreatment conditions to enhance ethanol production from poplar biomass. Appl. Biochem. Biotechnol. 105e108, 87e100. Nguyen, D., Naotsugu, N., Tamikazi, K., 2000. Effect of radiation and fungal treatment on lignocelluloses and their biological activity. Radiat. Phys. Chem. 59, 393e398. Nigam, J.N., 2001. Ethanol production from wheat straw hemicellulose hydrolysate by Pichia stipitis. J. Biotechnol. 87 (4), 17e27. Nolen, S., Liotta, C., Eckert, C., Glaser, R., 2003. The catalytic opportunities of near-critical water: a benign medium for conventionally acid and based catalyzed condensations for organic synthesis. Green Chem. 5, 663e669. Okano, K., Kitagaw, M., Sasaki, Y., Watanabe, T., 2005. Conversion of Japanese red cedar (Cryptomeria japonica) into a feed for ruminants by white-rot basidiomycetes. Anim. Feed Sci. Technol. 120, 235e243. Orozco, A.M., Al-Muhtaseb, A.H., Albadarin, A.B., Rooney, D., Walker, G.M., Ahmad, M.N.M., 2011. Dilute phosphoric acidcatalysed hydrolysis of municipal bio-waste wood shavings using autoclave parr reactor system. Bioresour. Technol. 102, 9076e9082. Pan, X., Xie, D., Gilkes, N., Gregg, D., Saddler, J., 2005. Strategies to enhance the enzymatic hydrolysis of pretreated softwood with high residual lignin content. Appl. Biochem. Biotechnol. 121, 1069e1079. Pan, X., Gilkes, N., Kadla, J., Pye, K., Saka, S., Gregg, D., Ehara, K., Xie, D., Lam, D., Saddler, J., 2006. Bioconversion of hybrid poplar to ethanol and co-products using an organosolv fractionation process: optimization of process yields. Biotechnol. Bioeng. 94 (5), 851e861. Parham, R.A., 1983. Wood structure-hardwood, in pulp and paper manufacture. In: Kocurek, M.J., Stevens, F. (Eds.), Properties of Fibrous Raw Materials and Their Preparation for Pulping. CPPA, Montreal, Canada. Park, J.-Y., Kang, M., Kim, J.S., Lee, J.-P., Choi, W.-I., Lee, J.-S., 2012. Enhancement of enzymatic digestibility of Eucalyptus grandis pretreated by NaOH catalyzed steam explosion. Bioresour. Technol. 123, 707e712. Rabinovich, M.L., Melnik, M.S., Boloboba, A.V., 2002. Microbial cellulases (review). Appl. Biochem. Microbiol. 38, 305e321. Ramos, L.P., 2003. The chemistry involved in the steam treatment of lignocellulosic materials. Quim. Nova 26, 863e871. Rogers, P., Jeon, Y., Lee, K., Lawford, H., 2007. Zymomonas mobilis for Fuel Ethanol and Higher Value Products. Springer-Verlag, Berlin. Adv Biochem. Eng. Biotechnol 108, 263e288. Sarkanen, K.V., 1980. Acid-catalyzed delignification of lignocellulosics in organic solvents. Prog. Biomass Convers. 2, 127e144. Sassner, P., Mårtensson, C.G., Galbe, M., Zacchi, G., 2008. Steam pretreatment of H2SO4-impregnated Salix for the production of bioethanol. Bioresour. Technol. 99, 137e145. Savage, P., Goplan, S., Mizan, T., Martino, C., Brock, E., 1995. Reactions at supercritical conditions: applications and fundamentals. AIChE J. 41, 1723e1778. Schell, D., Farmer, J., Newman, M., McMillan, J., 2003. Dilute-sulfuric acid pretreatment of corn stover in pilot-scale reactor: investigation of yields, kinetics and enzymatic digestibilities of solids. Appl. Biochem. Biotechnol. 105, 69e85. Shaw, A., Jennery, F., Adams, M., Lynd, L., 2008. End-product pathways in xylose fermenting bacterium Thermoanaerobacterium saccharolyticum. Enzyme Microb. Technol. 42, 453e458. Shevchenko, S.M., Beatson, R.P., Saddler, J.N., 1999. The nature of lignin from steam explosion/enzymatic hydrolysis of softwoods: structural and possible uses. Appl. Biochem. Biotechnol. 77, 867e876. Shuai, L., Yang, Q., Zhu, J.Y., Lu, F.C., Weimer, P.J., Ralph, J., Pan, X.J., 2010. Comparative study of SPORL and dilute-acid pretreatments of spruce cellulosic ethanol production. Bioresour. Technol. 101, 3106e3114. Shupe, A., Liu, S., 2009. Ethanol fermentation from hot-water wood extract hydrolysate with a modified Pichia stipitis strain HWR. AIChE Annu. Meet. (AIChE 2009). Shupe, A.M., Liu, S., 2012. Ethanol fermentation from hydrolysed hotwater wood extracts by pentose fermenting yeasts. Biomass Bioenergy 39 (0), 31e38. Sjöstrom, E., 1993. Wood Chemistry: Fundamentals and Applications. Academic Press, New York. Söderström, J., Pilcher, L., Galbe, M., Zacchi, G., 2003. Two-step steam pretreatment of softwood by dilute H2SO4 impregnation for ethanol production. Biomass Bioenergy 24, 475e486. Speers, A.M., Reguera, G., 2012. Consolidated bioprocessing of AFEX-pretreated corn stover to ethanol and hydrogen in a microbial electrolysis cell. Environ. Sci. Technol. 46 (14), 7875e7881. Stepanik, T.M., Rajagopal, S., Ewing, D., Whitehouse, R., 1998. Electron-processing technology: a promising application for the viscose industry. Radiat. Phys. Chem. 52, 505e509. Sun, Y., Cheng, J.Y., 2002. Hydrolysis of lignocellulosic materials for ethanol production: a review. Bioresour. Technol. 83, 1e11. Sun, X.F., Xu, F., Sun, R.C., Geng, Z.C., Fowler, P., Baird, M.S., 2005. Characteristics of degraded hemicellulosic polymers obtained from steam exploded wheat straw. Carbohydr. Polym. 60, 15e26. Sun, N., Rahman, M., Qin, Y., Maxim, M.L., Rodriguez, H., Roger, R.D., 2009. Complete dissolution and partial delignification of wood in the ionic liquid 1-ethyl-3-methylimidazolium acetate. Green Chem. 11, 646e655. Sun, Z., Shupe, A., Liu, T., Hu, R., Amidon, T.E., Liu, S., 2011. Particle properties of sugar maple hemicellulose hydrolysate and its influence on growth and metabolic behavior of Pichia stipitis. Bioresour. Technol. 102 (2), 2133e2136.
REFERENCES Swatloski, R.P., Holbrey, J.D., Memon, S.B., Caldwell, G.A., Caldwell, K.A., Rogers, R.D., 2004. Using Caenorhabditis elegans to probe toxicity of 1-alkyl-3-methylimidazolium chloride based ionic liquids. Chem. Commun. 6, 668e669. Talebnia, F., Karakashev, D., Angelidaki, I., 2010. Production of bioethanol from wheat straw: an overview on pretreatment, hydrolysis and fermentation. Bioresour. Technol. 101 (13), 4744e4753. Teixeira, L.C., Linden, J.C., Schroeder, H.A., 2000. Simultaneous saccharification and cofermentation of peracetic acid-pretreated biomass. Appl. Biochem. Biotechnol. 84e86 (1e9), 111e127. Teymouri, F., Laureano-Perez, L., Alizadeh, H., Dale, B.E., 2005. Optimization of the ammonia fiber explosion (AFEX) treatment parameters for enzymatic hydrolysis of corn stover. Bioresour. Technol. 96 (18), 2014e2201. Thring, R.W., Chorent, E., Overend, R., 1990. Recovery of a solvolytic lignin: effects of spent liquor/acid volume ratio, acid concentration, and temperature. Biomass 23, 289e305. Thuijl, E., Roos, C., Beurskens, L., 2003. An Overview of Biofuel Technologies, Markets, and Policies in Europe. Energy Research Centre of the Netherlands, Amsterdam, Netherlands. Tian, S., Zhu, W., Gleisner, R., Pan, X.J., Zhu, J.Y., 2011. Comparisons of SPORL and dilute acid pretreatments for sugar and ethanol productions from aspen. Biotechnol. Prog. 27 (2), 419e427. Tienvieri, T., Huusari Jr., E.S., Vuorio, P., Kortelainen, J., Ystedt, H., Artamo, A., 1999. Thermomechanical pulping. In: Sundholm, J., Gullichsen, J., Paulapuro, H. (Eds.). Mechanical Pulping, vol. 5. Papermaking Science and Technology. Fapet, OyJyvaskyla, Finland, pp. 157e221. Timell, T.E., 1960. Determination of the degree of polymerization of reducing pentose and hexose oligosaccharides. Sven. Papperstid. 63, 668e671. Tunc, M.S., van Heiningen, A.R.P., 2008. Hemicellulose extraction of mixed southern hardwood with water at 150  C: effect of time. Ind. Eng. Chem. Res. 47 (21), 7031e7037. Um, B.-H., van Walsum, G.P., 2012. Effect of pretreatment severity on accumulation of major degradation products from dilute acid pretreated corn stover and subsequent inhibition of enzymatic hydrolysis of cellulose. Appl. Biochem. Biotechnol. 168 (2), 406e420. US DOE, 2006. Breaking the biological barriers to cellulosic ethanol: a joint research agenda. Report from the December 2005 Workshop. DOE/SC-0095. Verma, P., Watanabe, T., Honda, Y., Watanabe, T., 2011. Microwaveassisted pretreatment of woody biomass with ammonium molybdate activated by H2O2. Bioresour. Technol. 102, 3941e3945. Vidal, P.F., Molinier, J., 1988. Ozonolysis of LignindImprovement of in vitro digestibility of poplar sawdust. Biomass 16, 1e17. Vidal, B.C., Dien, B.S., Ting, K.C., Singh, V., 2011. Influence of feedstock particle size on lignocellulose conversionda review. Appl. Biochem. Biotechnol. 164, 1405e1421. Wang, Y., Liu, S., 2012. Pretreatment technologies for biological and chemical conversion of woody biomass. Tappi 11 (1), 9e16. Wang, G., Pan, X., Zhu, J., Gleisner, R., 2009. Sulfite pretreatment to overcome recalcitrance of lignocellulose (SPORL) for robust enzymatic saccharification of hardwoods. Biotechnol. Prog. 25, 1086e1093. Wang, Y., Liu, S., Liang, L., Buyondo, J.P., Wang, Y., Garver, M., 2012. Biochemical conversion. In: Stuart, P.R., El-Halwagi, M.M. (Eds.), Integrated Biorefineries: Design, Analysis and Optimization. CRC Press, New York, pp. 591e650. Wasikiewicz, J.M., Yoshii, F., Nagasawa, N., Wach, R.A, Mitomo, H., 2005. Degradation of chitosan and sodium alginate by gamma radiation, sonochemical and ultraviolet methods. Radiat. Phys. Chem. 73, 287e295. Weber, C., Boles, E., 2010. Sugar-hungry yeast to boost biofuel production. Sci. News Sci. Dly. 92, 881e882. 487 Weber, C., Farwick, A., Benisch, F., Brat, D., Dietz, H., Subtil, T., Boles, E., 2010. Trends and challenges in the microbial production of lignocellulosic bioalcohol fuels. Appl. Microbiol. Biotechnol. 87, 1303e1315. Weil, J., Westgate, P., Kohlmann, K., Ladisch, M., 1994. Cellulose pretreatments of lignocellulosic substrates. Enzyme Microb. Technol. 16 (11), 1002e1004. Weimer, P.J., Hackney, J.M., French, A.D., 1995. Effects of chemical treatments and heating on the crystallinity of celluloses and their implications for evaluating the effect of crystallinity on cellulose biodegradation. Biotechnol. Bioeng. 48, 169e178. Wheals, A., Basso, L., Alves, D., Amorim, H., 1999. Fuel ethanol after 25 years. Trends Biotechnol. 17, 482e487. Willauer, H.D., Huddleston, J.G., Li, M., Rogers, R.D., 2000. Investigation of aqueous biphasic systems for the separation of lignins from cellulose in the paper pulping process. J. Chromatogr. B Biomed. Sci. Appl. 743, 127e135. Wu, J., Zhang, J., Zhang, H., He, J., Ren, Q., Guo, M., 2004. Homogeneous acetylation of cellulose in a new ionic liquid. Biomacromolecules 5, 266e268. Wyman, C.E., Dale, B.E., Elandert, R., Holtzapple, M., Ladisch, M.R., Lee, Y.Y., 2005. Coordinated development of leading biomass pretreatment technologies. Bioresour. Technol. 96 (18), 1959e1966. Wyman, C.E., Dale, B.E., Elander, R.T., Holtzapple, M., Ladisch, M.R., Lee, Y.Y., Mitchinson, C., Saddler, J.N., 2009. Comparative sugar recovery and fermentation data following pretreatment of poplar wood by leading technologies. Biotechnol. Prog. 25, 333e339. Xu, H.W., Scott, G.M., Jiang, F., Kelly, C., 2010. Recombinant manganese peroxidase (rMnP) from Pichia pastoris. Part 1: kraft pulp delignification. Holzforschung 64, 137e143. Yang, L., Liu, S., 2005. Kinetic model for kraft pulping process. Ind. Eng. Chem. Res. 44, 7078e7085. Yang, C.P., Shen, Z.Q., Yu, G.C., Wang, J.L., 2008. Effect and aftereffect of g radiation pre-treatment on enzymatic hydrolysis of wheat straw. Bioresour. Technol. 99, 6240e6245. Yasuda, S., Murase, N., 1995. Chemical structures of sulfuric-acid lignin. 12. Reaction of lignin models with carbohydrates in 72-percent H2SO4. Holzforschung 49, 418e422. Yokoyama, S., 2008. The Asian biomass handbook. In: Yokoyama, S., Matsumura, Y. (Eds.), A Guide for Biomass Production and Utilization. The Japan Institute of Energy, Japan. Zaldivar, J., Nielsen, J., Olsson, L., 2001. Fuel ethanol production from lignocellulose: a challenge for metabolic engineering and process integration. Appl. Microbiol. Biotechnol. 56, 17e34. Zavrel, M., Bross, D., Funke, M., Buchs, J., Spiess, A.C., 2009. Highthroughput screening for ionic liquids dissolving (ligno-)cellulose. Bioresour. Technol. 100, 2580e2587. Zayed, G., Meyer, O., 1996. The single-batch bioconversion of wheat straw to ethanol employing the fungus Trichoderma viride and the yeast Pachysolen tannophytus. Appl. Microbiol. Biotechnol. 45, 551e555. Zeikus, J., Ben-Bassat, A., Ng, T., Lamed, R., 1981. Thermophilic ethanol fermentation. Basic Life Sci. 18, 441e461. Zhang, Y.H.P., Himmel, M.E., Mielenz, J.R., 2006. Outlook for cellulase improvement: screening and selection strategies. Biotechnol. Adv. 24, 452e481. Zhang, Y.H.P., Ding, S.Y., Mielenz, J.R., Elander, R., Laser, M., Himmel, M., McMillan, J.D., Lynd, L.R., 2007. Fractionating recalcitrant lignocellulose at modest reaction conditions. Biotechnol. Bioeng. 97, 214e223. Zhao, Y., Wang, Y., Zhu, J.Y., Ragauskas, A., Deng, Y., 2008. Enhanced enzymatic hydrolysis of spruce by alkaline pretreatment at low temperature. Biotechnol. Bioeng. 99 (6), 1320e1328.
488 27. DEVELOPMENT OF THERMOCHEMICAL AND BIOCHEMICAL TECHNOLOGIES FOR BIOREFINERIES Zheng, Y.Z., Lin, H.M., Tsao, G.T., 1998. Pretreatment of cellulose hydrolysis by carbon dioxide explosion. Biotechnol. Prog. 14, 890e896. Zhu, J.Y., Pan, X.J., 2010. Woody biomass pretreatment for cellulosic ethanol production: technology and energy consumption evaluation. Bioresour. Technol. 101 (13), 4992e5002. Zhu, Y., Lee, Y.Y., Elander, R.T., 2005. Optimization of dilute-acid pretreatment of corn stover using a high-solids percolation reactor. Appl. Biochem. Biotechnol. 121, 1045e1054. Zhu, J.Y., Pan, X.J., Wang, G.S., Gleisner, R., 2009. Sulfite pretreatment (SPORL) for robust enzymatic saccharification of spruce and red pine. Bioresour. Technol. 100 (8), 2411e2418. Zhu, J., Pan, X., Zalesny Jr, R., 2010a. Pretreatment of woody biomass for biofuel production: energy efficiency, technologies, and recalcitrance. Microbiol. Biotechnol. 87, 847e857. Zhu, J., Pan, X., Zalesny Jr, R., 2010b. On the energy consumption for size-reduction and yield from subsequent enzymatic saccharification of pretreated lodgepole pine. Bioresour. Technol. 101 (8), 2782e2792. Zhu, J.Y., Zhu, W.Y., O’Bryan, P., Dien, B.S., Tian, S., Gleisner, R., Pan, X.J., 2010c. Ethanol production from SPORL-pretreated lodgepole pine: preliminary evaluation of mass balance and process energy efficiency. Appl. Microbiol. Biotechnol. 86 (5), 1355e1365.
Index Note: Page numbers with “f” denote figures; “t” tables; and “b” boxes. A Ablative reactors, 245 Accelerated electron transfer, mediators for flavins, 142 phenazines, 143 Acetone butanol ethanol (ABE), 116 Acid-catalyzed transesterification reaction, 121 Acid-driven hydrolysis, 161e163, 162f Acidic lignin condensation, 321e322 Advanced reactor system, 9 AFEX. See Ammonia fiber explosion (AFEX) Agricultural residues, 50 Agroindustrial residues, for bioethanol production raw material, 50e52 cellulose-containing residues, 52 starch-containing residues, 51e52, 52t sugar-containing residues, 51, 51t sugar production and fermentation SHF, 52e55 SSF, 55 Algae biomass, 249 Alkali-catalyzed transesterification, 13 Alkaline neutralization, 195 Alkaline pulping, lignin reactions in, 321f Alkaline salts, in zeolites, 256 Alkaloids, 353 American Society for Testing and Materials (ASTM), 177 Amino acids, 110e112, 110f, 337e341 arginine, 340e341 aromatic amino acids, 341 glutamic acid, 338 lysine, 339e340 methionine, 340 threonine, 340 Amino-based products amino acids, 337e341 arginine, 340e341 aromatic amino acids, 341 glutamic acid, 338 lysine, 339e340 methionine, 340 threonine, 340 aspartame, 341 polyamines cadaverine, 348e349 putrescine, 346e348 poly amino acids, 341e345 cyanophycin/cyanophycin granule polypeptide, 342e343 poly-g-glutamic acid, 343e345 ε-poly-L-Lysine, 345 Ammonia fiber explosion (AFEX), 290e291, 473 Ammonia recycled percolation (ARP), 61e62, 290e291, 465 Anabaena cylindrica, 368 Anaerobic biodegradation, 204 Anaerobic digestion (AD), 14e15, 15f of microalgae, 164, 164f Animal-derived biolipids, 188e189 Animal fat wastes algae as biodiesel source, 12e13 biodiesel from, 11e12 beef tallow and chicken fat, 12 pork lard, 12 waste cooking oils, 12 Astaxanthin, 361 B Bacillus coagulans, 84 Bacteria, 77 Bacterial metabolism cytochromes (cell-bound), 140e142 shewanella oneidensis MR-1 bioelectrochemical machinery, 142, 144f riboflavin (vitamin B2) biosynthetic genes of, 142, 145f TFP, 139e140 Bidirectional hydrogenase, 371 Bioactive phytochemicals, 357, 357f Biobased fats, 197, 198f Biocathode chamber, microbial species in, 137e139, 138t Biocathodes, 136e137 Biochar processing agricultural benefits, 450 economic analysis, 451e454 cost-effective fertilizer substitute, 451e452 different potential income streams, biochar carbon value for, 454 grain crop primary productivity, costeffective approach to, 452e453 livestock growth rates, 453e454 postpyrolysis indirect application of, 440e442 489 container growth medium and container, 441e442 soil nutrient reclamation, 441 water filtration, 440e441 properties of, 436e438 chemical and physical, 436e437 microbiological effects and synergisms, 437e438 theoretical income streams, 448e450 biochars and carbon markets, carbon sequestration of, 449e450 renewable energy and fuel generation, 448e449 utilization of carbon sequestration, 438 plant nutrients and other pollutants, sorption of, 438e439 soil greenhouse gas emissions, 439e440 soil-specific biochar design, 440 Biochar overland flow filter, 441, 442f Biochemical (sugar) platform, 227e228 Biocrude, 166 Biodegradability, 204, 205t Biodiesel, 25 bioreactors for, 13e14 feedstocks for, 10e13 algae as source, 12e13 from animal fat wastes, 11e12 from pure vegetable oil, 10e11 waste cooking oils, 12 fuel properties of, 160t and petrodiesel, 160e161, 160t properties of, 196e197 Biodiversity, 374e375 Bioenergy production, technological routes for advanced biomass-to-biofuels development platform, 30e33 biomass pretreatment, 28e29 combined pretreatment/hydrolysis and fermentation strategies, 29 fermentation, 29 hydrolysis, 29 Bioenergy-related enzymes databases and web servers, 98e103 CAT and dbCAN, 101e102 CAZy database, 98e101 FOLy database, 102 plant coexpression network databases, 102e103
490 Bioenergy-related enzymes (Continued) purdue cell wall genomics and UC-riverside cell wall navigator databases, 102 plant biomass, 95e96 and regulation, 96e98 Bioenergy resources, 34 Bioethanol in Brazil, feedstock for, 25 ethanol production bioreactors in, 8e9 immobilization of cells for, 9 feedstock for, 3e4 fermentation, 7e8 lignocelluloses, pretreatment of, 4e7, 5f biological pretreatment, 5e6 chemical pretreatment, 6e7 physical pretreatment, 6 molecular biology trends in, 8 Bioethanol production process, 4, 4f, 161e164, 162f Biofuel production, representative microorganisms features for, 83t Biofuels development program, bioenergy resources and, 33e34 Biogas feedstock, 15 household digesters for, 15e17 fixed dome digesters, 15e16, 16f floating drum digesters, 16e17, 17f social and environmental aspects of, 17 Biohydrogen, 165e166, 165f Biohydrogen photoproduction by cyanobacteria, 372e375 H2 production strategies in, 373e375 by green algae hydrogen photoproduction, nutrientdeprived green algae by, 378e381 light-dependent hydrogen production pathways, 375e377 long-term H2 production by, 377e378 H2 photoproduction in, 377, 381e382 mechanisms of, 368e372 alternative nitrogenases, 369e370 hydrogenases, 370e372 nitrogenases, 369e370 oxygenic photosynthesis, 368e369 Biolipids biodiesel, properties of, 196e197 energy balance, 192e193 liquid fuels, biomass to, 197e198 cleaning process, 197 gasification, 197 synthesis, 197e198 processing of countercurrent extraction, 194 enzymatic hydrolytic maceration, 193e194 expression, 194 extraction, 193 hot continuous extraction, 194 maceration, 193 steam distillation, 193 supercritical fluid extraction, 194e195 ultrasound extraction, 194 INDEX pure plant oil, properties of, 195e196 alkaline neutralization, 195 bleaching, 196 degumming, 195 transesterification, 196 winterization, 195e196 sources of animal-derived biolipids, 188e189 edible lipids, 186e187 microalgae and other oleaginous microorganisms, 189e190 nonedible lipids, 187 plant-derived biolipids, 186 waste edible oil, 187e188 supply and projected/purrent volume, 190e192 Biological pretreatment, 5e6 microorganisms for, 73e79 bioconversion, practical applications in, 74e77 genetically modified microorganisms for, 77e79 Biomass addition of enzymes, 206 addition of microbes, 205e206 aquatic weeds, 209e211 biodegradability, 204, 205t digesters, loading and unloading of, 212 digestion systems, 211e212 family-size biogas plant, 211 scum layer digester, 211e212 solid biomass digester, 212 wet digesters, 211 energy crops, 207 food processing residues, 207e209 bagasse, 207e208 coffee husks and mucilage, 208e209 rice husks, 207 gasification of, 247e248 catalytic gasification, 247e248 hydrothermal liquefaction of, 248e251 catalyst, 250e251 feedstock, 249 reaction conditions, 249e250 solvent, 250 kitchen and garden waste, 209 longer retention times, 207 macro and micronutrients, 204e205, 206t pretreatments, 207 chemical pretreatment, 207 enzymes, biological pretreatment with, 207 hot water treatment, 207 mechanical pretreatment, 207 spent bedding, 209 use of methane, 213 utilization paths, 224, 227t volatile solids chemical conversion of, 213e214 thermal conversion of, 214 wet digesters digestate treatment in, 212e213 solids content in, 212 pyrolysis of catalytic pyrolysis, 244e245 fast pyrolysis, 244 reactors, 245e247 Biomass-based renewable energy system, 448e449 Biomass conversion, 461e462, 462f to bioproducts, 462, 463f Biomass-derived chemical building blocks, 297 Biomass pyrolytic conversion processes, 435e436, 436t Biomass sources, 33, 33f Biomethane, 25, 164e165 Biorefinery system, 32e33 and bio-based products, 221 bioconversion, converting sugars to products, 477e478 biomass conversion, overview on, 461e462 eco-efficiency, 231e236 and emerging opportunities for sustainable economy, 219e221 feedstock, 221e224 hydrolysis, 476e477 integration and multifunctionality in, 227, 232f lignocellulosic biomass, characteristics of, 458e461 mechanical pretreatment acid hydrolysis, 474e475 alkaline hydrolysis, 475e476 ammonia fiber explosion (AFEX), 473 ammonia recycle percolation pretreatment, 465 biological pretreatment, 474 chipping/grinding/milling and refining, 463e465 electron beam/gamma ray and microwave, irradiation pretreatment by, 465 hot water, 469e470 ionic liquid pretreatment, 467e468 organosolv pretreatment, 466 overcome recalcitrance of lignocelluloses, sulfite pretreatment to, 468e469 oxidation pretreatment, 466e467 ozonolysis pretreatment, 465e466 steam explosion, 470e473 supercritical carbon dioxide explosion, 473e474 physical/mechanical methods, pretreatment-biomass size reduction by, 462e476 platforms, 227e231, 232f structure of, 224e227 thermochemical conversion, 478e481 combustion, 478 direct liquefaction, 481 gasification, 478e481 pyrolysis, 481 Biopellets, 3 Bleaching, 196 Brown-rot fungi, 75e76 Bubbling fluid bed reactor, 245e246, 246f
491 INDEX C D H Calcium oxide (CaO), 256 Carbohydrate composition, 278e280, 279t Carbohydrate esterases (CEs), 98e100 Carbon dioxide pretreatment, 61 Carbon-negative bioenergy systems, 439f Carbon sequestration, of biochars and carbon markets, 449e450 Carboxymethyl cellulose (CMC), 78 Carotenoids, 353 Catalyst hydrophobicity, 259 Catalytic gasification, 247e248 Catalytic thermochemical processes, for biomass conversion gasification, 247e248 hydrothermal liquefaction, 248e251 pyrolysis, 244e247 CAZy database, 98e101 CAZyme Analysis Toolkit (CAT), 101 Cellobiose hydrolyzation, 161e163 Cellulose, into monomeric units of glucose, 161e163, 163f Cellulose raw material, 52 Cell wall model, pH on solubilization, 291e292, 294f Cell wall related (CWR), 96, 97t C9-formulae, 329e330 Chemical pretreatment acid hydrolysis, 6e7 alkaline hydrolysis, 7 ammonia hydrolysis, 7 ozonolysis, 7 Chemical pulp, productions of, 316, 316t Chlamydomonas, 166 Circulating fluidized beds, 245e246, 246f Clay minerals, 260e265 acid activation of, 262, 262f biodiesel production, acid-activated clay minerals in, 263e265 improving acidity, 262e263 treatment of, 262e263 Concept of eco-efficiency, 233 Confectionery industry waste streams, 429e430, 429f Conjugation, 393 Conserved domain database (CDD), 101 Contamination, 156 Corrosion, 121 Corn dry milling, 359 Countercurrent extraction, 194 CMC. See Carboxymethyl cellulose (CMC) Crude glycerol streams, 424 Crystalline cellulose, 115 Cyanobacteria, 165 biofuels, production system for, 393e403 aliphatic alcohols and alkanes, photosynthetic production of, 402e403 butyraldehyde and butanol, 401e402 ethanol, 398 ethylene, 398e400 hydrogen, 393e398 isoprene, 400e401 engineering, 392e393 strains/tools and methods, 392e393 Cyanothece, 373, 397 Dedicated bioenergy crops, 38e39 Dedicated energy crops (DECs), 15 Degree of lignin condensation (DC), 319 Degree of polymerization (DP), 458e459 Degumming, 195 DET. See Direct electron transfer (DET) Diacylglycerol acyltransferases (DGATs), 175e176 Digesters, 214 Digestion systems, 211e212 family-size biogas plant, 211 scum layer digester, 211e212 solid biomass digester, 212 wet digesters, 211 Dimethylallyl-diphosphate (DMADP), 400e401 Direct electron transfer (DET), 134, 134f, 137, 143f Direct hydrothermal liquefaction, 249 Hardwood cellular structure of, 458, 459f lignins, 318e319 Hemicellulose, 76 Heterocyst differentiation, 372e373 Heterocyst forming species, 395 Heterogeneity, in feedstock, 30e32 Heterogeneous catalysts clay minerals, 260e265 biodiesel production, acid-activated clay minerals in, 263e265 improving acidity, 262e263 heteropolyacids, 257e258 layered materials, 265e269 layered carboxylates, 266 layered double hydroxides, 265 layered hydroxide salts, 265e266 transesterification reactions, heterogeneous catalysts in, 266e269 polymeric catalysts, 269e272 zeolites, 258e260 Heteropolyacids, 257e258 Heteropolyoxometalate anions, 257 Heterotrophic microalgae, 155e156 High-efficiency algal-derived biocrude, second-generation biofuel from, 153, 154f biocrude, 166 biodiesel, 158e159 biodiesel and petrodiesel, 160e161, 160t bioethanol production process, 161e164, 162f biohydrogen, 165e166, 165f biomethane, 164e165 contamination, 156 culture techniques, 156 heterotrophic microalgae, 155e156 HTL, 167 hydrothermal catalytic liquefaction, 167 microalgae biomass/biofuel production-cultivation, 155 production of biodiesel from, 159e160 microalgal biofuels, 154e155 microalgal biomass to, 158 mixing, 156 nutrients, 156 open-pond culture, 157, 157f photobioreactors, 157e158, 157f phototrophic microalgae, 155 processing microalgal biomass for, 158, 158f subcritical water, properties of, 166e167, 167f Hot continuous extraction, 194 HOX. See Reversible hydrogenase (HOX) Hydrogen hydrogen bioproduction, 394 H2 production, heterocystous cyanobacteria by, 396e397 H2 production, nonheterocystous cyanobacteria by, 397e398 hydrogen-evolving enzymes, 394e395 hup-hydrogen uptake enzyme, 394e395 nitrogenase, gratuitous hydrogenase, 395 E Edible lipids, 186e187 Escherichia coli, 8, 8f Ethanol, 50 Ethylene, 398e400 bioproduction of, 399e400 microbial production of, 399 Energy balance, 192e193 Entrained-fow reactors, 245 Environmental impact ratio, 233e234, 235f Enzymatic hydrolytic maceration, 193e194 Enzyme-mediated transesterification, 119e120 Extracellular/free lipases, 121e122 F FAMEs. See Fatty acid methyl esters (FAMEs) Fatty acids, in animal fats, 188t Fatty acid methyl esters (FAMEs), 119e120, 159e160, 263e264 First-generation bioenergy, 408 Fish pond discharge (FPD), 180e181 Flavins, 142 Flavonoids, 353 FOLy database, 102 Fomitopsis pinicola, 76 Fossil fuel-derived petrodiesel, 160e161 Fossil fuels, 24e25 FPD. See Fish pond discharge (FPD) Fuel ethanol production, 25 Furfural, main outlets of, 299e300, 299f G Gasification, 197 Global energy production chart, 1e2, 2f Gloeophyllum trabeum, 76 Glycoside hydrolases (GHs), 96 Glycosyltransferases (GTs), 96 Green biorefineries, 227, 229f Greenhouse gases (GHGs), 1e2, 23e24, 65, 407e408
492 Hydrogen (Continued) reversible hydrogenase (HOX), 395 Hydrogen-evolving enzymes, 394e395 hup-hydrogen uptake enzyme, 394e395 nitrogenase, gratuitous hydrogenase, 395 reversible hydrogenase (HOX), 395 Hydrothermal liquefaction (HTL), 154e155, 167, 248e251 Hydrothermal pretreatment methods, 65 Hyper-isobutanol-producing strains, 113e114 I ILUC. See Indirect land use change (ILUC) Immobilization, 375 Indirect land use change (ILUC), 407e408 Insoluble inorganic salts, 260 Integrated biorefinery system, 236f Intracellular/immobilized lipases, 122, 122t Ionic liquid pretreatment, 467e468 Irradiation pretreatment, 465 Isobutanol production chemistry of fermentation, 112e113, 116 higher alcohol production, Keto acid pathways for, 110e112 microorganisms, metabolic engineering of, 113e114 overview, 109e110 sustainable feedstocks, bioenergy crops as, 114e115 technologies, 115e116 Isoprene, 400e401 J Jatropha curcas (Jatropha), 10, 172 K 2-ketoisovalerate (KIV), 111e112 biosynthetic genes, 113e114 L Layered carboxylates, 266 Layered double hydroxides, 265 Layered hydroxide salts, 265e266 Layered materials, 265e269 carboxylates, 266 double hydroxides, 265 hydroxide salts, 265e266 transesterification reactions, heterogeneous catalysts in, 266e269 Lewis acidity, 262 Light-dependent hydrogen production pathways, 375e377 Lignin biological degradation of, 73 classification of, 320, 320t degradation, 320 under acidolysis conditions, 321e322, 322f potential products from, 297, 297f production/properties and analysis, 318e332 advanced NMR methods, 330 analytical methods for characterization of, 323 INDEX molecular weight distribution, 330e331 C NMR analysis of, 327e330, 329t 31 P NMR reproducibility analytical techniques, 323e327, 326t structure-properties correlations in, 331e332 renewable raw material feedstock, 315e318 traditional and emerging applications emerging lignin applications, 332e333 traditional lignin applications, 332 valorization, 333 Lignocellulose-based chemical products C6 and C6/C5 sugar platform, 295e296 chemical transformation products, 296 fermentation products, 295e296 carbohydrate dehydration, 298e305 furfural production and applications, 298e301 hexose feedstock, 5hydroxymethylfurfural formation from, 301e305 cellulose, 280 chemicals and fuels, furans/aromatics as building blocks for, 297e298 hemicelluloses arabinogalactan, 282e283 arabinoglucuronoxylans, 282 arabinoxylan, 283 carbohydrate feedstocks, 283 complex heteroxylans, 283 galactoglucomannans, 282 glucomannan, 282 glucuronoxylans, 280e282 xyloglucans, 282 b-(1 / 3, 1 / 4)-glucans, 283 lignin, 283e286 lignin platform, 296e297 lignocellulosic biorefineries, 292e295 monoaromatic chemicals, technical lignins into, 305e309 acid-catalyzed depolymerization, 305 base-catalyzed depolymerization, 305 ionic liquids, 308 lignin aromatics, future perspectives of, 308e309 oxidative depolymerization, 306 pyrolysis, 305e306 reductive hydrodeoxygenation, 306e307 solvolysis, 307 sub-and supercritical water, 307e308 supercritical solvents, 308 occurrence and composition of, 278e280 storage carbohydrates, 280 structural carbohydrates, 280 pretreatment technologies, 286e290 alkaline (lime) pretreatment process, 289e290 dilute and concentrated acid pretreatment, 289 at laboratory/conceptual stages, 290e292 liquid hot water (LHW), 288 steam explosion, 286e288 wet oxidation, 288e289 13 Lignocellulose pretreatment, 64e65 Lignocellulosic bioenergy processes, 362, 362f Lignocellulosic biomass catalytically hydrothermal liquefaction of, 250, 250t lignin content and chemical structures of, 284, 284t pretreatment methods of, 286, 287t pyrolysis of, 244, 244t chemical composition of, 4e7, 5f ethanol production from, 25e26 pretreatment of, 58 microorganisms, biological pretreatment with, 73 nonbiological pretreatment, 72e73 pyrolysis of, 244, 244t utilization of, 50 Lignocellulosic energy crops, 50 Lignocellulosic feedstock biorefinery (LCFBR), 37e38, 224e227, 229f Lignocellulosic raw materials, 32e33 Lignocellulosic sources, cell wall compositions of, 30 Lignocellulosic waste materials, 50 Lignosulfonates, 322e323 Linear electron transport (LET), 368 Lipase-catalyzed biodiesel production acyl acceptors, 126 catalyzed transesterification done in, 121e122 chemical transesterification, disadvantages of, 120e121 chemistry of biodiesel, 120, 120f choice of enzyme, 124 feedstock, 123e124 algae oils, 124 animal oils/fats, 123 vegetable oils, 123 waste oils/fats, 123e124 historical background of, 121 immobilized, advantages of, 122e123 molar ratio, 124 reactor system, 126e127 solvents, 126 technical challenges, 123 temperature, 124e126 transesterification, 120, 120f using lipases, advantages of, 121 water content, 126 Lipid content, and biomass productivity, 190, 191t Liquid fuels, 25 biomass to, 197e198 cleaning process, 197 gasification, 197 synthesis, 197e198 Liquid hot water (LHW), 73 Liquefaction of biomass, 250 Long-term soil testing, 453e454 Low light-utilization efficiency, 381 Lowry acid sites, 259 Lysophosphatidate acid acyl-transferase (LPAT), 176
INDEX M Maceration, 193 Madhuca indica (Mahua), 11 Main technology platforms for bioenergy production, 26, 26f Marine biorefinery (MBR), 227 MCM-41 molecular sieves, 259 Mediated electron transfer (MET), 134e135, 137e139, 143f through exogenous redox mediators, 134e135, 135f through secondary metabolites, 135, 135f Membrane bioreactor, 9, 9f Metabolic engineering poses, 78 Metabolic flux analysis (MFA), 78e79 Methane emissions, 450 31 P-II methodology, 323 Methylene diphenyl diisocyanate (MDI), 317 Methyl esters, purification of, 121 MET. See Mediated electron transfer (MET) MFA. See Metabolic flux analysis (MFA) Microalgae biofuel, 154e155 biomass, 158 biofuel production-cultivation, 155 cultivation of, 40 and oleaginous microorganisms, 189e190 production of biodiesel from, 159e160 uses of, 41 Microbial conversion, of biomass biodiesel, 172 production, 177, 178t viable feedstocks for, 173 waste utilization for, 179e181 biological pretreatment, microorganisms for, 73e79 bioconversion, practical applications in, 74e77 genetically modified microorganisms for, 77e79 fatty acid methyl esters and fuel properties, 179 genetic engineering approach, 175e177 lignocellulosic biomass and pretreatment, 72e73 microorganisms, biological pretreatment with, 73 nonbiological pretreatment, 72e73 microbial pretreatment for biogas production, 80e81 for biomass conversion, 81e84 strategies of using, 79e84 petroleum fuel scenario, India, 172 potent strains, selection of, 173e175 RAS, 180f renewable energy, 171 Microbial electrolysis cells (MECs), 143e144 Microbial fuel cells (MFC) anode reaction in, 132e133, 132f biocathodes, electron transfer for, 136e139 DET for, 137 MET for, 137e139 bioelectrochemistry of, 132e139 biofilm electrochemistry for, 139e143 accelerated electron transfer in, 142e143 bacterial metabolism, 139 mediator-less factors affecting, 139e142 biofilms, electrogens in, 135e136 cathode reaction in, 132f, 133 concomitant electricity, wastewater treatment with, 143e147 reactor designs, 143e144 substrates used in, 145e147 electrode reactions in, 132e133 anode reaction, 132e133 cathode reaction, 133 electron transfer methods, 133e135 anodic biofilms, DET for, 134 anodic biofilms, MET for, 134e135 Microbial pretreatment for biogas production, 80e81 for biomass conversion, 81e84 strategies of using, 79e84 Microbial putrescine, and cadaverine production, 348t, 349 Microorganism platform, 227 Microwave-based heating, 63 Miscanthus species, 114e115 Miscanthus x giganteus, 114e115 Molecular hydrogen, 394 Multiple sequence alignment (MSA), 101 N Napier grass biomass, 114e115 Native lignins, structural moieties of, 318e319, 319f Negative performance, of bioenergy crops, 408 Neisseria gonorrhoeae, 139, 140f Nicotiana tabacum (Tobacco), 11 Nicotinamide adenine dinucleotide (NADH), 397e398 Nonbiological pretreatment, potential advantages over, 73 Nonedible lipids, 187 Nutrient recycling, 214 O Octahedral structural metals, 262e263, 262f Oilseeds, biodiesel production from, 424 Oil transesterification, 13, 13f Open-pond culture, 157, 157f Organic agriculture, 409e410 Organosolv pretreatment, 60e61 advantages of, 61 Oriented Strand Boards (OSB), 317 Oxidation pretreatment, 466e467 Oxygenic photosynthesis, 368e369 P Pelletization process, 3, 3f Penicillium digitatum, 399 Perennial grasses, 114e115 Perennial herbaceous energy crops, 24 Permissible exposure limit, 449e450 493 Petrodiesel, fuel properties of, 160t Phanerochaete chrysosporium, 5 Phenazines, 143 Phenoleformaldehyde (PF), 331 Photobioreactors, 157e158, 157f Photosynthetic electron transfer reactions, 368 Photosynthetic light reactions, 368 Phototrophic microalgae, 155 Physical pretreatment extrusion, 6 mechanical comminution, 6 steam explosion, 6 ultrasonic pretreatment, 6 Phytochemicals bioactive/pharmaceutical phytochemicals, 356e357 colorants, 355e356 coproduction of, 361e363 bio-oil/syn-gas and algal bioenergy processes, 362e363 fermentative phytochemicals production, colocation of, 363 phytochemical production by-products, utilization of, 363 plant oil-based bioenergy processes, 361e362 starch/sugar-based bioenergy processes, 361 dietary/nutraceutical/food or feed additives, 356 major groups of, 353e354, 354f for personal care, 358 production of, 358e361 algae via aquaculture, 361 cultured plant cells, 360 extraction and isolation from specific plants, 358 microbial fermentation, 360e361 staple crops, biorefinery processing of, 358e360 Phytosterols, 354 Plant biomass, 95e96 Plant coexpression network databases, 102e103 Plant-derived biolipids, 186 Plant-derived colorants, 358 Plant oil-based biodiesel processes, 360f, 361e362 Plant oil, properties of, 195e196 alkaline neutralization, 195 bleaching, 196 degumming, 195 transesterification, 196 winterization, 195e196 Polyamines cadaverine, 348e349 putrescine, 346e348 Poly amino acids, 341e345 cyanophycin/cyanophycin granule polypeptide, 342e343 applications for, 343 biodegradability of, 343 production of, 343 poly-g-glutamic acid
494 Poly amino acids (Continued) applications of, 344e345 biodegradability of, 344e345 production of, 344 poly-g-glutamic acid, 343e345 ε-poly-L-Lysine, 345 applications of, 345 degradation of, 345 production of, 345 Poly (3-hydroxybutyrate) (PHB) confectionery and bakery industry waste streams, production from, 429e430 production from whey, 430 Polyhydroxyalkanoates biodiesel industry by-products, valorization of, 424e427 biorefinery concepts, production integrated in, 421e430 food industries, by-product streams from, 427e430 second-generation bioethanol industry byproducts, 427 structure and properties of, 420e421 winery by-products, production from, 428e429 Polymeric carbohydrates, 354 Polymeric catalysts, 269e272 Polyoxometalates, 257 Polyvinyl chloride (PVC), 14 Pongamia pinnata (Karanja), 11 Postharvest processing, 360 Postpyrolysis indirect application, 440e442 container growth medium and container, 441e442 soil nutrient reclamation, 441 water filtration, 440e441 Preliminary bioreactor fermentations, 425e426 Pressurized high temperature entrained flow reactor (PiTER), 245, 245f Pretreatments, 207 enzymes, biological pretreatment with, 207 modeling, 65 trends in, 62e65 technologies, laboratory/conceptual stages AFEX, 290e291 ionic liquids, 291 lignocellulosic biomass pretreatments, 291e292 sub/supercritical treatments, 291 types of, 58e62 biological pretreatments, 58e59 chemical pretreatments, 60e61, 207 hot water treatment, 207 mechanical pretreatment, 207 physical pretreatments, 59e60 physicochemical pretreatments, 61e62 Proton exchange membrane (PEM), 132 Pseudomonas aeruginosa, 139, 140f, 143 Pseudomonas cepacia, 124 Purdue cell wall genomics, and UCriverside cell wall navigator databases, 102 Pyrolysis bioenergy industry, 436 INDEX R RAS. See Recirculatory aquaculture system (RAS) Recirculatory aquaculture system (RAS), 180, 180f Reed canary grass, 39 Renewable energy (REN), 24 Reversible hydrogenase (HOX), 395 Rotating cone reactor, 246e247, 246f Roundtable of sustainable biomaterials (RSB), 410, 412 S Saccharomyces cerevisiae, 29, 161 Saturated fatty acids, 179 Scenedesmus obliquus, 368 S-contained compounds, 354 Second-generation bioenergy, 408 Separate hydrolysis and fermentation (SHF), 52e55 SHF. See Separate hydrolysis and fermentation (SHF) Simultaneous saccharification and fermentation (SSF), 29, 52, 55 Single superphosphate (SSP), 451e452 Soaking aqueous ammonia (SAA), 61e62 Soft-rot fungi, 76e77 Softwood, cellular structure of, 458, 459f Soil fertility, 214 Soil greenhouse gas emissions, 439e440 Soil-specific biochar design, 440 Solid state cultivation (SSC), 73 Solid state fermentation approach, 9 Soybean processing, 358, 359f SSF. See Simultaneous saccharification and fermentation (SSF) Starch raw material, 51e52, 52t Steam explosion, 470e473 Steam distillation, 193 Steam pretreatment, objective of, 61 Stoichiometric models, 78e79 Sugarcane bagasse, 63 Sugar raw material, 51, 51t Sunflower-based biorefinery concept, 426e427, 426f Supercritical fluid extraction, 194e195 fluid parameters, 308, 308t Sustainability, 37f biodiesel, feedstocks for, 39e41 bioenergy feedstocks and dedicated biofuel crops, 37e39 dedicated bioenergy crops, 38e39 lignocellulosic feedstocks, 37e38 Sustainable farming bioenergy production biofuel production, sustainability criteria for, 410e412 biomass uses, 412 farm level, potential on, 414e415 global bioenergy potential, 413e414 ways of comparisons, 410 criteria for agricultural production, 409e410 conventional agricultural production, 409 food systems, 410 SWOT analysis, 233, 235t Synechococcus elongatus, 113 T Technoeconomic analysis, 66 TFP. See Type IV pili (TFP) Thermal depolymerization, 244 Thermochemical biorefinery (TCBR), 227 Thermochemical platform, 227e229 Thioesterase (TE), 176e177 Torrefaction, 64 Traditional fatty acid esterification processes, 256 Transesterification, 159e160, 160f, 196 heterogeneous catalysts in, 266e269 use of acid catalysis, 257 Transgenic organisms, enhance lipid biosynthesis in, 175e176, 176t Triacylglycerol (TAG), 175 chemical structure of, 120, 120f Trichodesmium, 373 Tricarboxylic acid (TCA), 399 Tubular photobioreactor (TPBR), 14, 14f, 396f Type IV pili (TFP), 139e140 U Ubiquitous phytocolorants, 355, 355f UC-riverside cell wall navigator databases, purdue cell wall genomics and, 102 Ultrasound extraction, 194 Unclassified biorefineries, 229 Unicellular cyanobacterium, 397 V Vegetable oil, production of, 360, 360f VFAs. See Volatile fatty acids (VFAs) Vegetative cells, 372, 372f Vibrio cholerae, 139, 140f Volatile fatty acids (VFAs), 145 Volatile solids chemical conversion of combustion, 213 gasification, 213e214 thermal conversion of, 214 V-nitrogenase, 369 W Waste edible oil, 187e188 Wet digesters digestate treatment in, 212e213 solids content in, 212 Wet milling yields, 359 Wet oxidation, 61 White-rot fungus application for animal feed, 75 biofiber, biomass for, 75 in biological pulping, 75 corn stover, 74 cotton stalks, 74 paddy straw, 75 rice straw, 74e75
495 INDEX Whole crop biorefinery (WCBR), 227, 229f Wine lees valorization, 428, 428f Winterization, 195e196 WIPO. See World Intellectual Property Organization (WIPO) Wood major components of, 459, 460t, 461 typical structures of, 458, 458f World Intellectual Property Organization (WIPO), 317e318, 318f X X-ray diffraction (XRD), 263 Z Zeolites, 258e260 Zeotype materials, 258