Author: Rich Alexander   Hill Terrell L.  

Tags: molecular biology  

ISBN: 978-1-4612-9134-3

Year: 1987

Text
                    Springer Series in Molecular Biology

Series Editor: Alexander Rich


Springer Series in Molecular Biology Series Editor: Alexander Rich Yeast Genetics Fundamental and Applied Aspects J.F.T. Spencer, Dorothy M. Spencer, A.R.W. Smith, eds. Myxobacteria Development and Cell Interactions Eugene Rosenberg, ed. DNA Methylation Biochemistry and Biological Significance Aharon Razin, Howard Cedar, Arthur D. Riggs, eds. Cooperativity Theory in Biochemistry Steady-State and Equilibrium Systems Terrell L. Hill Molecular Biology of DNA Methylation Roger L.P. Adams, Roy H. Burdon Protein Compartmentalization Arnold W. Strauss, Irving Boime, Gunther Kreil, eds. Peptides of Poisonous Amanita Mushrooms Theodor Wieland Structure, Function and Genetics of Ribosomes Boyd Hardesty and Gisela Kramer Host-Parasite Relationships and the Yersinia Model Akira Wake, Herbert R. Morgan Linear Aggregation Theory in Cell Biology Terrell L. Hill Cytochromes c: Biological Aspects Graham W. Pettigrew and G.R. Moore
Terrell L. Hill Linear Aggregation Theory in Cell Biology With 119 Figures Springer-Verlag New York Berlin Heidelberg London Paris Tokyo
Terrell L. Hill National Institute of Diabetes and Digestive and Kidney Diseases National Institutes of Health Bethesda, Maryland 20892, USA Series Editor: Alexander Rich Department of Biology Massachusetts Institute of Technology Cambridge, Massachusetts 02139, USA Library of Congress Cataloging-in-Publication Data Hill, Terrell L. Linear aggregation theory in cell biology. (Springer series in molecular biology) Bibliography: p. Includes index. 1. Proteins-Synthesis. 2. Polymers and polymerization. 3. Biopolymers. 4. Statistical mechanics. I. Title. II. Series. QP551.H495 1987 574.87'6042 87-4867 ISBN-13: 978-1-4612-9134-3 e-ISBN-13: 978-1-4612-4736-4 DOl: 10.1007/978-1-4612-4736-4 © 1987 by Springer-Verlag New York Inc. Softcover reprint of the hardcover 1st edition 1987 All rights reserved. This work may not be translated or copied in whole or in part without the written permission of the publisher. (Springer-Verlag, 175 Fifth Avenue, New York, New York 10010, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in connection with any form of information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed is forbidden. The use of general descriptive names, trade names, trademarks, etc. in this publication, even if the former are not especially identified, is not to be taken as a sign that such names, as understood by the Trade Marks and Merchandise Marks Act, may accordingly be used freely by anyone. Typeset by Asco Trade Typesetting Ltd., Hong Kong. 9 8 7 6 5 432 1 ISBN-13: 978-1-4612-9134-3 Springer-Verlag New York Berlin Heidelberg
To Marie-France earlier and Marc Kirschner
Series Preface During the past few decades we have witnessed an era of remarkable growth in the field of molecular biology. In 1950 very little was known ofthe chemical constitution of biological systems, the manner in which information was transmitted from one organism to another, or the extent to which the chemical basis oflife is unified. The picture today is dramatically different. We have an almost bewildering variety of information detailing many different aspects of life at the molecular level. These great advances have brought with them some breathtaking insights into the molecular mechanisms used by nature for replicating, distributing, and modifying biological information. We have learned a great deal about the chemical and physical nature of the macromolecular nucleic acids and proteins, and the manner in which carbohydrates, lipids, and smaller molecules work together to provide the molecular setting ofliving systems. It might be said that these few decades have replaced a near vacuum of information with a very large surplus. It is in the context ofthis flood of information that this series of monographs on molecular biology has been organized. The idea is to bring together in one place, between the covers of one book, a concise assessment of the state of the subject in a well-defined field. This will enable the reader to get a sense of historical perspective-what is known about the field today-and a description of the frontiers of research where our knowledge is increasing steadily. These monographs are designed to educate, perhaps to entertain, certainly to provide perspective on the growth and development of a field of science that has now come to occupy a central place in all biological studies. The information in this series has value in several perspectives. It provides for a growth in our fundamental understanding of nature and the manner in which living processes utilize chemical materials to carry out a variety of activities. This information is also used in more applied areas. It promises to
V111 Series Preface have a significant impact in the biomedical field where an understanding of disease processes at the molecular level may be the capstone that ultimately holds together the arch of clinical research and medical therapy. More recently, in the field of biotechnology, there is another type of growth in which this science can be used with many practical consequences and benefit in a variety of fields ranging from agriculture and chemical manufacture to the production of scarce biological compounds for a variety of applications. This field of science is young in years but it has already become a mature discipline. These monographs are meant to clarify segments of this field for the readers. Cambridge, Massachusetts Alexander Rich Series Editor
Contents Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi I. Linear Equilibrium Aggregates 1. Statistical Thermodynamic Background . . . . . . . . . . . . . . . . . . . . . . . . ... ... ... ... 4 6 18 23 2. Attached Single-Stranded Polymer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 5. Attached Polymer at Equilibrium or Steady State . . . . . . . . . . . . . . . . 6. Attached Polymer in Transients. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7. Attached Polymer under a Force. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 42 51 3. Free Single-Stranded Polymer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 8. Free Polymer at Equilibrium. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9. Kinetic Aspects for a Free Polymer . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 90 1. 2. 3. 4. Canonical and Grand Partition Functions. . . . . . . . . . . . Aggregation and Osmotic Pressure Vi rial Coefficients . . . Partition Function for an Open, Independent Aggregate. The Macroscopic Aggregate as a Limiting Case. . . . . . . . . . . . ..... ..... ..... ..... . . . . 3
Contents x 4. Single-Stranded Polymer Modified by a Second Component, a Bound Ligand, or a Cap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 110 10. Two-Component Single-Stranded Polymer . . . . . . . . . . . . . . . . . . . . . 11. Single-Stranded Polymer with Bound Ligand or Cap. . . . . . . . . . . . . . 110 122 5. "Surface" Properties of Some Long Multi-Stranded Polymers. . . . . . 137 12. 13. 14. 15. ... . .. ... ... 137 144 156 167 6. Some Attached Multi-Stranded Polymers at Equilibrium and in Transients. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 16. 17. 18. 19. General Discussion of the Models. . . . . . . . . . . . . . . . . . . . . . . . . Equilibrium and Steady-State Properties of Aligned Models. . . . . Equilibrium and Steady-State Properties of Staggered Models. . . . Models with Dimers as Subunits. . . . . . . . . . . . . . . . . . . . . . . . . . Simple Dual Aggregation and the Vernier Effect. Dual Aggregation with Vernier Enhancement. . . A Further Example of Dual Aggregation. . . . . . . Aligned Tubular Models at Equilibrium. . . . . . . . . . . . . . . . .... .. . . .... .... . .. .. . ... ... . .. ... ... ... . .. .. . ... ... . . . . 174 184 190 193 II. Linear Steady-State Aggregates 7. Enzymatic Activity at Polymer Tips Only 199 20. Enzymatic Activity along the Polymer Length. . . . . . . . . . . . . . . . . .. 21. Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes.. 22. Enzymatic Activity at Polymer Tips Only: Length Distributions and Transients. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 23. Fluctuations in the Polymer Length Distribution. . . . . . . . . . . . . . . .. 200 203 8. NTP Caps and Possible Phase Changes at Polymer Ends. . . . . . . . .. 227 24. 25. 26. 27. Illustrative Biochemical Models that Generate Phase Changes. . . . . Attached Polymer with Phase Changes at the Free End. . . . . . . . . . Free Polymer with Phase Changes at the Ends. . . . . . . . . . . . . . . . . Simulation of Two "Phases" by Aggregation of One Component on Another. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219 223 .. .. .. 228 244 265 .. 284 Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 303
Introduction The coverage ofthis book is quite limited. The aim is to give a unified theoretical treatment of the basic physical and chemical principles involved in the reversible aggregation of independent linear polymer molecules. Generally these polymers are aggregates of protein monomers or subunits. The emphasis in the book is on basic principles as illustrated by simple models, not on particular applications or on particular polymers (though microtubules and actin are at present the only known examples of some fundamental phenomena to be described). "Reversible aggregation," above, refers to linear polymer molecules that can gain or lose subunits continuously from their ends (unless capped or blocked): aggregation occurs because of physical forces between subunits (van der Waals, electrostatic, etc.), not from more permanent chemical bonds (as in DNA, for example). The term "independent," used above, will limit the discussion (with a few exceptions) to solutions that are dilute in polymer molecules (so that interpolymer interactions can be neglected) or to noninteracting polymers attached to a surface or nucleating center. The aggregation properties of independent polymer molecules are not only fundamental but also are sufficiently complicated to provide more than enough material for a book of reasonable size. The emphasis is on equilibrium or steady-state behavior. Transients do not receive much attention. The choice of material outlined above parallels, not coincidentally, my own contributions to the subject. In fact, the book is essentially a cohesive account and extension of this work. I hope that this more systematic treatment will make the theory easier to follow and more useful in applications to both students and researchers. I would be especially gratified if the book serves to interest physical chemists, physicists, and theoreticians in this subject. This book is related in fundamental ways to two previous books: my own Thermodynamics of Small Systems, Part II (Benjamin, New York, 1964) and
xn Introduction Thermodynamics of the Polymerization of Protein by F. Oosawa and S. Asakura (Academic Press, New York, 1975). The present volume is not meant to supplant the pioneering work by Oosawa and Asakura, but rather to supplement it, in part by more recent advances. In fact, most space in the two books is devoted to rather different topics. A significant difference is that the present book uses statistical mechanics as the starting point. The book on small systems, mentioned above, provides a rigorous thermodynamic background for the present analysis (see Chapter 10, especially). Also, the pertinent partition function Y is introduced and applied to simple aggregation models. The first six chapters of the present volume, which are the most fundamental, are devoted to general physical aggregation systems. The treatment in these chapters is based, for the most part, on quite simple illustrative models (not on particular real polymers). "Physical" here signifies that the on and off transitions of the subunits at the ends of an aggregate do not involve any chemical reaction. In contrast, Chapter 7 relates to the aggregation of enzyme molecules. Although the enzyme molecules themselves aggregate because of simple physical interaction forces, the detailed catalytic properties of the enzyme (e.g., in the overall reaction E + S -+ E + P) are altered as a conseqnence of the aggregation. The modified enzymatic activity at and near the polymer ends has a number of interesting consequences. Actin filaments and microtubules are the two known examples of this kind of behavior at the present time (actin catalyzes ATP -+ ADP + Pi and tubulin catalyzes GTP -+ GDP + PJ Finally, Chapter 8 deals with a special case of the enzyme aggregation problem of Chapter 7. In this case, the subunits at and near a polymer end can exist in one of two "phases": practically all ES or practically all EP, with occasional changes from one phase to the other. If these two phases have very different subunit on and off rate constants, phase changes can have dramatic effects on polymer stability. So far, micro tubules are the only known example of this kind of two-phase activity. For greater generality, the free subunit activity a is used throughout the book in place of the more conventional concentration c whenever no additional complication ensues. The book contains much material not previously published. On the other hand, my original papers on this subject (some with collaborators) contain many details not included in the book. For the convenience of readers interested in pursuing these details, the following is a list of the pertinent references: 1. Molecular clusters in imperfect gases. 1. Chem. Phys. 23, 617-622 (1955). 2. Statistical Mechanics (McGraw-Hill, New York, 1956; also Dover, New York, 1987), Section 27 (Exact treatment of physical clusters). 3. Thermodynamics of Small Systems, Part II (Benjamin, New York, 1964), Chapter 10. 4. Theory of aggregation in solution. Biopolymers 12, 1285-1312 (1973). With Y. Chen. 5. Bioenergetic aspects and polymer length distribution in steady-state head-
Introduction 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. Xlll to-tail polymerization of actin or microtubules. Proc. N atl. Acad. Sci. USA 77,4803-4808 (1980). Steady-state head-to-tail polymerization of actin or microtubules. II. Two-state and three-state kinetic cycles. Biophys. J. 33, 353-372 (1981). Microfilament or microtubule assembly or disassembly against a force. Proc. Natl. Acad. Sci. USA 78, 5613-5617 (1981). Subunit treadmilling of micro tubules or actin in the presence of cellular barriers: possible conversion of chemical free energy into mechanical work. Proc. Nat!. Acad. Sci. USA 79, 490-494 (1982). With M.W. Kirschner. Bioenergetics and kinetics of microtubule and actin filament assemblydisassembly. Int. Rev. Cytol. 78, 1-125 (1982). With M.W. Kirschner. Regulation of microtubule and actin filament assembly-disassembly by associated small and large molecules. Int. Rev. Cytol. 84, 185-234 (1983). With M.W. Kirschner. Steady-state theory ofthe interference of GTP hydrolysis in the mechanism of microtubule assembly. Proc. N atl. Acad. Sci. USA 80, 7234-7238 (1983). With M.F. Carlier. Use of Monte Carlo calculations in the study of microtubule subunit kinetics. Proc. Natl. Acad. Sci. USA 80, 7520-7523 (1983). With Y. Chen. Interference of GTP hydrolysis in the mechanism of microtubule assembly: an experimental study. Proc. Natl. Acad. Sci. USA 81, 771-775 (1984). With M.F. Carlier and Y. Chen. Phase changes at the end of a microtubule with a GTP cap. Proc. Nat/. Acad. Sci. USA 81, 5772-5776 (1984). With Y. Chen. Introductory analysis of the GTP-cap phase-change kinetics at the end of a microtubule. Proc. N atl. Acad. Sci. USA 81, 6728-6732 (1984). Phase-change kinetics for a microtubule with two free ends. Proc. Natl. Acad. Sci. USA 82, 431-435 (1985). Monte Carlo study of the GTP cap in a five-start helix model of a microtubule. Proc. Nat!. Acad. Sci. USA 82,1131-1135 (1985). With Y. Chen. Theoretical treatment of microtubules disappearing in solution. Proc. Natl. Acad. Sci. USA 82, 4127-4131 (1985). With Y. Chen. Theoretical problems related to the attachment of micro tubules to kinetochores. Proc. Natl. Acad. Sci. USA 82, 4404-4408 (1985). A model for actin polymerization and the kinetic effects of ATP hydrolysis. Proc. Natl. Acad. Sci. USA 82, 7207-7211 (1985). With D. Pantaloni, M.F. Carlier, and E.D. Korn. Theoretical study of a model for the ATP cap at the end of an actin filament. Biophys. J. 49, 981-986 (1986). Effect of fluctuating surface structure and free energy on the growth of linear tubular aggregates. Biophys. J. 49,1017-1031 (1986). A theoretical study of cooperative dual linear aggregation and the vernier effect. Biophys. Chern. 25,1-15 (1986). Synchronous oscillations in microtubule polymerization. Proc. N atl. Acad. Sci. USA 84 (1987), in press. With M.-F. Carlier, R. Melki, D. Pantaloui, and Y. Chen.
xiv Introduction I am much indebted to Yi-der Chen, Marc Kirschner, Marie-France Carlier, Tim Mitchison, and Dominique Pantaloni for their stimulation and collaboration in this field over a number of years. Also, I received help on particular points from Bruce Nicklas, Attila Szabo, William Eaton, Harold Erickson, Robert Rubin, and Otto Berg. Finally, I thank Kathy Van Tassel for her prompt and expert typing of the manuscript.
I Linear EquilibriuIn Aggregates
1 Statistical Therm.odynam.ic Background By an "equilibrium aggregate" is meant one whose formation involves physical attractive forces only (e.g., van der Waals and electrostatic interactions between protein molecules). This is the simplest case and is the subject of Part I of this book. Part II treats "steady-state aggregates." This term refers to cases in which the monomers are enzyme molecules and a chemical reaction, catalyzed by the aggregated enzyme, accompanies or follows monomer attachment to the aggregate. Actin and micro tubules provide examples of this, in which the chemical reactions are the hydrolysis of ATP or of GTP, which are bound to the respective monomers. Furthermore, in the case of microtubules, at least, it is possible for a microtubule end to spend most of its time in either of two distinct states or phases: the subunits at and near the end have GTP bound (not yet hydrolyzed) or these subunits have GDP bound (i.e., the GTP has been hydrolyzed). In the first phase, the microtubule end is stable; in the second it is unstable. This two-phase behavior is novel and requires separate treatment (Chapter 8). An "equilibrium aggregate" can be at equilibrium, in a steady state (i.e., growing or shortening at a steady rate), or in a transient. The present chapter is concerned with general equilibrium background only. Applications and illustrations will be reserved for later chapters. Kinetic aspects will be introduced in particular cases, beginning in Chapter 2. If desired, the reader may regard Sections 1-3 as appendices, to be consulted only for results, as needed. In particular, although Section 2 provides the general and fundamental approach to equilibrium aggregation theory, many readers may find the details uninteresting or too complicated. These details are hardly needed for the remainder of the book. They are included primarily for readers with an intrinsic interest in fundamentals. Section 4 contains necessary reference material, mostly thermodynamic, that pertains to very large aggregates.
Statistical Thermodynamic Background 4 1. Canonical and Grand Partition Functions In this section, we review a few properties of the two most important partition functions, for a one-component system. Generalizations will be required later, but these will be introduced as needed. Canonical partition function. We consider an equilibrium thermodynamic system with independent variables N (number of molecules), V (volume), and T (temperature). The system might be gas, liquid, or solid. The temperature is fixed by a surrounding heat bath. We are interested in the connection between the macroscopic thermodynamic properties of this system, on the one hand, and its (quantum mechanical) molecular properties, on the other. Statistical mechanics provides the desired formal connection, which can be made explicit for systems that are not too complicated. If Ej(N, V) is the energy (eigenvalue) of the j-th energy eigenstate of the system, then the canonical partition function is Q(N, V, T) = I (1.1) e-Ej(N,V)/kT, j where the sum is over all statesj, and k is the Boltzmann constant. Degenerate energy levels are represented by several equal terms in the sum. In practice, in solution physical chemistry (where there is more than one component), usually the internal vibrational partition function of each molecule is separated out, and the remaining coordinates (translation, external and internal rotation) are taken care of by a classical phase integral over e- H / kT , where H is the Hamiltonian function of these coordinates. Details are given in Ref. 1, pp. 262-264. The connection with thermodynamics is made via the Helmholtz free energy: A(N, V, T) = - kTln Q(N, V, T) (1.2) and dA = -SdT - pdV + /ldN (1.3) or d( - A/kT) = - Ed(l/kT) + (p/kT) dV - (/l/kT) dN, (1.4) where S is the entropy, p the pressure, /l the chemical potential, and E the mean energy (E fluctuates). Differentiation of - kTln Q, Eq. (1.3), or In Q, Eq. (1.4), gives various thermodynamic properties of interest. An example, from Eq. (1.4), is - (81n Q) 81/kT E = - From Eq. (1.1), we then find V.N· (1.5)
Canonical and Grand Partition Functions L if = 5 Eje-Ej/kT (1.6) --"jI=-e----;;EJ--;;./k-OOT;- • j This shows that the probability that the system, at equilibrium, will be observed in state j is e-EJikT/Q. This is the famous Boltzmann probability distribution. Grand partition function. 2 Here the one-component system is open with respect to N: the system exchanges molecules with a reservoir of the same molecules at chemical potential /1 and temperature T Consequently, N fluctuates. The independent thermodynamic variables are now /1, V, T rather than N, V, T The corresponding partition function, introduced by Gibbs, is the grand partition function: 2.(/1, V, T) L = Q(N, V, T)e NJl /kT (1.7) Q(N, V, T)A N, (1.8) N~O L = N~O where A (the absolute activity) == eJl/kT, and the sum goes over all possible values of N. The notation QN(V, T) for Q(N, V, T) is used in the next section. Incidentally, when N = 0, Q(O) = 1 (unless the system includes an always present "background"). The term "grand" refers to the second, or higherorder, sum over N; the first sum is over j for each N, as in Eq. (1.1). The connection with thermodynamics is p V = kTln 2.(/1, V, T), d(pV) = SdT 2. = e PV /kT , + pdV + Nd/1. (1.9) (1.10) Of course Eqs. (1.3), (1.4), and (1.10) are all equivalent thermodynamic rela- tions, rearranged for convenience (according to the choice of independent variables). From Eqs. (1.7), (1.8), and (1.10), we find the important relations N = kT(Oln2.) 0/1 I A(Oln2.) OA T,V = T.V (1.11) NQ(N, V, T)A N N~O L Q(N, V, T)A N (1.12) N~O Equation (1.12) shows that PN = Q(N, V, T)AN _ c. (1.13) is the probability that this open system contains exactly N molecules. If Q(O) = 1, the probability that the system is empty (N = 0) is 1/2.. Equation (1.13) is the analogue of the Boltzmann probability distribution for a closed
Statistical Thermodynamic Background 6 system (see above). In other words,just as the separate terms in the canonical partition function, Eq. (1.1), give the relative probabilities that the closed system has various energy values Ej , so the separate terms in the grand partition function, Eq. (1.8), give the relative probabilities that the open system contains various numbers of molecules N. A further differentiation of Eq. (1.12), N8 L = NQ(N, V, T)2 N, N;;'O leads to the variance in N: (J~ -- (ON) = N 2 - N2 = (N - N)2 = 2 - 02 . (1.14) T,V In summary: the first differentiation ofln 8 with respect to 2 gives N, Eq. (1.11), and the second gives (J~, Eq. (1.14). For the special case of a very dilute gas (2 ~ 0), we have from Eqs. (1.8) and (1.11): 8 = ePVjkT = 1 + Q(I)2 pV/kT = Q(I)2 = N, J1 + ... e == N/V = Q(I)2/V = J1°(T) + kTln e J1°(T) = -kTln[Q(I)/V]. (1.15) (1.16) (1.17) (1.18) Equation (1.17) is the familiar thermodynamic expression for J1 as a function of concentration e and T. Equation (1.18) gives the expression for the thermodynamic standard chemical potential J1°(T) in statistical mechanical terms. When the gas is not dilute, J1(c, T) = J1°(T) + kTln a(e, T) (1.19) defines the concentration activity a(c, T). The activity coefficient y(e, T) is defined by a = ye. The activity has the property a ~ e as e ~ 0; also, in this limit, y ~ 1. The osmotic solution version of some of these equations appears in the next section. 2. Aggregation and Osmotic Pressure Virial Coefficients Aggregation or cluster formation in an imperfect one-component gas can be treated exactly. !, 2 The McMillan- Mayer solution theory!' 2 allows the formal methods of imperfect gas theory to be extended without change to a solution that is in an osmotic equilibrium. Thus, aggregation of solute in a solution can also be treated exactly, at least in principle. The formal procedure 3 is outlined in this section. The method is actually quite general and not limited to linear aggregates, the special topic of this
Aggregation and Osmotic Pressure Vi rial Coefficients 7 book. The aggregates may have any shape, but we shall always have linear aggregates in mind. The "solute," in a solution, is the aggregating species. We are interested, in this book, in cases in which the intermolecular forces between solute molecules, leading to aggregation, are quite strong, strong enough, in fact, to produce equilibrium linear aggregates of macroscopic length at a finite concentration c of the solute (called the "critical concentration," ce). In addition to these forces that produce the aggregation, there will be additional nonaggregating forces between the aggregates themselves, at the least "hard" interactions that arise from the fact that two aggregates cannot occupy the same space in the solution at the same time. Except in Section 8, we shall assume in this book that the large aggregates or polymers of primary interest (e.g., F-actin and micro tubules) are dilute enough so that hard interactions between polymers can be ignored: the strong aggregating forces completely dominate; the polymers can be treated as independent open systems. 4 This assumption is made to simplify the analysis and also to help limit the size of the book. It excludes aggregates such as tobacco mosaic virus 5 and HbS,6 solutions of which can actually separate into two macroscopic phases because of hard interactions between non-dilute polymers. The stacking of nucleosides or bases is an example of equilibrium linear aggregation in which the aggregating forces are not very strong, partly because the aggregate is single-stranded. The dominating aggregates are relatively small: dimers, trimers, tetramers, etc. For such systems, with weak aggregating forces, hard interactions between aggregates should not be ignored. The present section is especially pertinent for such systems but we shall not pursue this particular application 3 here. It is usually assumed, for simplicity, that the actual transitions that occur at the ends of a large equilibrium linear aggregate are the arrivals and departures of monomers. Section 5 will include the more general model in which these exchanges at polymer ends may involve not only monomers but also dimers, trimers, etc. The present section will provide the needed background theory. With this introduction, we turn now to the general theoretical approach. 3 The object is to show formally but exactly the way in which osmotic pressure virial coefficients En for an aggregating solute in a solution can be decomposed into contributions arising from aggregation (or "association") equilibria and from interactions (for brevity, this term will refer, in this section, to all nonassociating interactions between aggregates). The essential point is the formal mathematical identity between aggregation in a one-component imperfect gas and aggregation of solute molecules in an "osmotic" solution. Once we establish this relationship, we shall revert to the much simpler notation appropriate to a one-component gas with the understanding that we are in fact primarily interested in the osmotic solution case.
Statistical Thermodynamic Background 8 Consider a solution 2 at pressure p + II, temperature T, volume V, with solute molecules at chemical potential fl, and solvent molecules (any number of components) at fla' flp, .... This solution is in contact and in equilibrium, via a membrane permeable to all solvent components but not to solute, with a large solvent reservoir at p, T, fla' flp, .... Let QN.N.,Np,.JV, T) be the canonical partition function for the solution with numbers of molecules N (solute), Na , N p , .... Then the grand partition function of the solution is (2.1) where A = e fl / kT , Aa = e fl ./kT , etc. This is a generalization of Eqs. (1.8) and (1.9). Equation (2.1) can be rewritten as e(p+II)V/kT L = '¥N AN, (2.2) N~O where This is a sum over the numbers of all solvent components but with a fixed number N of solute molecules. The special case '¥o is the grand partition function for a volume V of the solvent reservoir: HI _ pV/kT _ " Q o,N.,Np, ... /LIN.1Np (2.4) TO - e L..., /Lp .•.. N.,Np, ••• ~ 0 a Then, from Eqs. (2.2) and (2.4), e IIV / kT = L ('¥N/'¥o}AN. (2.S) N':i30 Equation (2.5) may be compared with the grand partition function of a one-component gas, Eq. (1.8): e PV / kT = L QN(V, T)A N. (2.6) N~O The vi rial coefficients and related properties of the gas can be expressed 2 in terms of the QN with N = 1, 2, 3, .... There is no restriction on the nature of the gas molecules except that a vi rial expansion must exist. These same expressions hold for the solution 2 with II replacing p and '¥N/'¥O replacing QN' In the solution case, the influence of the solvent appears implicitly through the dependence of,¥Nand '¥0 on fla' flp, .... Also, the potential of mean force 2 between solute molecules plays the role of the intermolecular potential energy in the gas. If the solute molecules are charged, the solvent must include an electrolyte (in order to have a virial expansion). The osmotic virial expansion will diverge when macroscopic aggregates are formed, that is, at the critical concentration C = Ceo Hence, this procedure is valid only for C < ceo We use Eq. (2.5) as our starting point but from here on, for simplicity of notation, we denote '¥N/'¥O by QN' Since gas cluster theory 2 begins with
Aggregation and Osmotic Pressure Virial Coefficients 9 Eq. (2.6), we can take over the essentials ofthat theory, without change, in our consideration of solute aggregation in solution. Q2 (= '1'2/'1'0) is a partition function for exactly two solute molecules in solvent (V, T, J.1~, J.1P'···)· Q2 includes an integral over all possible positions of the two solute molecules in V and over all rotational orientations (Ref. 1, pp. 264, 271, and 277). If there are forces leading to aggregation, Q2 may be split into two parts (usually by a specified division of the translational-rotational configuration space), Q2 = Q200 ... + QOlO ... , (2.7) where Q200 is the partition function of two (interacting) monomers in V and is the partition function of one dimer in V This is an exact split up of Q2; no approximation is involved. In general, in QN, the boldface N == N1, N 2 , .,. represents a set of Nl monomers, N2 dimers, N3 trimers, etc., in V Q2 is the complete partition function for two molecules. It includes both "interaction" and "association" forces between the molecules. The dimer state is included implicitly in Q2' On the other hand, the right-hand side of Eq. (2.7) makes the existence of dimers explicit. The division of Q2 into Q200 and QOlO is a nonthermodynamic procedure that depends on the definition of a dimer. But in practice, with strong associating forces, there will be little ambiguity about the definition. 3 Figure 1-1 provides an idealized illustration. The solute monomers are uniform (hard) spheres except for sites A and B, at the two ends of a diameter, which have a very strong attraction for each other. Consequently, a linear chain of spheres can form by aggregation. An "association" between two monomers is considered to have been formed if the distance between sites A and B is less than some preassigned value (a few A). Figure 1-1(a) shows Q010 (a) (} (c) (b) Fig. 1-1. Illustration of linear association by strong interactions between sites A and B on spherical molecules. (a) Two monomers in a "hard" interaction. (b) Monomer and dimer. (c) Dimer and trimer.
10 Statistical Thermodynamic Background a hard interaction between two monomers, not an association. Figures l-l(b) and 1-1 (c) show, respectively, a hard interaction between a dimer and a monomer and a hard interaction between a trimer and a dimer. Corresponding to Eq. (2.7), we have for N = 3 and N = 4, + Q110 + Q0010 Q40 + Q210 + Q1010 + Q020 + Q00010' Q3 = Q30 (2.8) Q4 = (2.9) where Q30 is the partition function of three monomers in V, Qll0 refers to one monomer and one dimer, etc. On the right-hand side, all sets of subscripts are used such that N = Nl + 2N2 + 3N3 + .... A digression on notation is needed at this point. We have introduced the symbol QN above (instead of using QN) because eventually [following Eq. (2.52)] we shall want to use Q2 for QOlO (dimer), Q3 for Q0010 (trimer), Q4 for QOOOIO (tetramer), etc. That is, Q2 is only part of Q2' etc. However, for a monomer, there is no subdivision as in Eqs. (2.7)-(2.9), hence Q1 = Q1' and we shall always use Q1 for the partition function of a monomer in solvent (i.e., for 'Pd'Po). Equation (2.5), with QN == 'PN/'PO' includes aggregates implicitly. As in Eqs. (1.16) and (1.19), we define the concentration activity a == Q1A/V so that a --+ e when e --+ 0, where e = R/v (the concentration of solute). Then,2 on taking the logarithm of both sides ofEq. (2.5), after replacing A by aV/Q1, we obtain IT/kT = a + b2 a 2 + b3 a3 + "', (2.10) where the bj are related to the QN (in the first few cases) by2 v2 2!Vb2 = Z2 - 3!Vb3 = Z3 - 3VZ2 4!Vb4 = Z4 - 4VZ 3 + 2V 3 - 3z1 (2.11) + 12V 2 Z 2 ZN == QN VNN!/Qf· - 6V 4 (2.12) ZN is the so-called configuration integral (translation and rotation) for N molecules (see Ref. 1, pp. 264, 271, and 277). The bj are generally used in actual calculations rather than the QN' To obtain the osmotic virial expansion, IT/kT in powers of e rather than of a [Eq. (2.10)], we use the solution equivalent of Eq. (1.11): e iJIT/kT = a--- = aa L jbja . J, (2.13) j?3 1 where b1 == 1. This gives e as a power series in a. This series can easily be inverted (we omit details) to give a as a power series in e. Substitution of the latter series, aCe), in Eq. (2.10) then produces the desired result: (2.14)
11 Aggregation and Osmotic Pressure Virial Coefficients where B z = -bz, B4 = Bs = 112bi - B3 = 4bi - 2b 3, + 18bz b3 - 3b4 , 144bib3 + 32b z b4 + 18bj - (2.15) -20bi 4b s . The Bn are functions of T, fla' flp, •... The above relations, Eqs. (2.10)-(2.15), refer to the system of aggregating solute molecules but these equations do not display the aggregation explicitly: the forces leading to association appear implicitly in the calculation of the QN' Alternatively, we can treat aggregates explicitly by writing Eq. (2.5) in the form e nV/kT = " L... N;>O QNl\l lN 1 1Nz I\z ... (2.16) for a multicomponent mixture of solute aggregates of sizes s = 1, 2, 3, ... (monomers, dimers, etc.). These are the QN of Eqs. (2.7)-(2.9). In this case, we define a concentration activity for each species by (2.17) The identity of A1 and 2 follows from Eqs. (2.5) and (2.16) in the limit A --+ O. The corresponding concentrations are Cs = Ns/V with C1 + 2c z + 3c 3 + ... = (2.18) c. The activities have the property that as --+ Cs for all s when all Cs --+ O. The as differ from the Cs because of the interactions between aggregates. The aggregates of various sizes are in equilibrium with each other so that the As in Eq. (2.16) are not independent. The equilibrium condition is As = AS for all s ;;. 1 (i.e., fls = Sfl). For example, 2z = 2 2 , together with Eq. (2.17), leads to the relation (2.19) where K2 is the concentration equilibrium constant for dimer formation from monomers. K z is a function of T, fla' flp, .... The partition functions appear here in the conventional wayZ for a chemical equilibrium. A theoretical calculation of K z requires formulation of Q1 and QOlO (see Section 8 for an example). Similarly, from A3 = A3, - (QdV)3 - = A4 , etc., a 4 _ Q00010/ V = K a 3 _ QOOlO/V = K a3 A4 3, a4 - (QdV)4 - 4, etc. (2.20) Note that K3 is the association constant for trimers from monomers, etc.
Statistical Thermodynamic Background 12 The notation Ks is shorthand for a partition function quotient. The Ks should not be regarded here as empirical equilibrium constants. The more conventional equilibrium constants are + dimer <=± trimer; monomer + trimer <=± quad rimer; monomer K~ = K3/K2 K~ = K4/K3 (2.21) etc., or (2.22) We return now to Eq. (2.16). We replace the As by the as' using Eq. (2.17), take logarithms, and find the muiticomponent generalization of Eq. (2.10): II -k T = a 1 + a 2 + a 3 + ... . . ... + "L. bjaila~2 j (j1 + j2 + ... ~ 2) (2.23) where the boldfacej refers to the setj1,j2' ... and 2! Vb 20 = Z20 - V 2 3!Vb30 = Z30 - 3VZ20 + 2V 3 2! Vb020 = Z020 - V 2 1! 1! Vb 11 0 = (2.24) Z110 - V 2 2! 1! Vb 210 = Z210 - VZ 20 - 2VZ 110 + 2V 3, with ZN = QN V N 1+ N2+···N1 !N2 ! ... Nl N2 N3 Q1 Q010Q0010··· (2.25) ZN is the configuration integral for the set of solute molecules N = N 1 , N 2 , •••• Again, the bj are used in practice 3 rather than the QN. The bj are related to multicomponent virial coefficients Bn in a generalization of Eqs. (2.15) [see Eq. (89) of Ref. 3]. Whereas the Ks include the forces that lead to aggregation, the bj (or Bn) take into account the remaining forces of interaction between aggregates. At the least, the latter would include "hard" or "space-filling" interactions; ifthe aggregates are charged, and in an electrolyte solution, there would be a screened electrostatic potential of mean force as well, etc. Our aim here is to express the operational thermodynamic quantities Bn , of Eq. (2.14), in terms of the theoretical quantities Ks and the bj • We do this via the bj since the Bn and bj are rather simply related by Eqs. (2.15). If we substitute a 1 = a, a2 = K 2a 2, a3 = K 3a 3, etc., in Eq. (2.33), we will obtain a power series in the single variable a that must be identical with Eq. (2.10). We can then equate coefficients of like powers of a to find the bj in terms of the Ks and bj : b2 = K2 + b20
Aggregation and Osmotic Pressure Virial Coefficients 13 (2.26) Actually, these equations can be written down as easily as Eqs. (2.7)-(2.9), because both sets depend on the same subscript rules. Substitution of Eqs. (2.26) into Eqs. (2.15) gives: B 2 = -K2 - b20 B3 = B4 = (2.27) + 2K 2(4b 20 - bll0 ) + 4bio - 2b30 -3K4 + 18K 2K 3 - 20K~ + 3K 3(6b 20 - b1010 ) + 3K~(6bll0 - 20b20 - b020 ) + 3K2(6bll0b20 - 20bio + 6b 30 - b210 ) + 18b2o b3o - 20bio - 3b40 . - 2K 3 + 4K~ (2.28) (2.29) Thus, we have the beginning of the osmotic pressure virial expansion, Eq. (2.14), with the Bn expressed explicitly in terms of the association equilibrium constants, Eq. (2.20), and nonassociating interactions among aggregates, Eqs. (2.24). In principle, the procedure can be continued to higher Bn but in practice B5 is about the limit. As already mentioned, this virial expansion would diverge at e = ee (where indefinitely large aggregates are formed). When the aggregation forces are very strong, the bi in Eqs. (2.27)-(2.29) can be neglected, as a reasonable approximation. That is, the solute in solution is treated as an ideal mixture of aggregates. This approximation is discussed in some detail in the next subsection. When aggregation forces are weak, leading to small aggregates only, the bi should not be and need not be neglected. An explicit example of this type, including both the Ks and the bi , can be found in Ref. 3, pp. 1300-1311. The Ks part of this example is treated in Section 8. When association does not occur at all (all Ks = 0), Eqs. (2.27)(2.29) reduce to Eqs. (2.15). Note that in the "ideal" case (all bi = 0), values of the Ks can be deduced from the experimental Bn alone. But in the "real" case, where both the Ks and the bi are included, some theory (e.g., of the bi ) and/or nonthermodynamic experiments have to be invoked in order to divide the Bn between the Ks and the bi . This situation is to be expected because aggregation is an extrathermodynamic concept. It is of some interest to have equations for the as and the e., especially the latter, as functions of e. The dependence of a 1 (= a) on e is, like that of II on e, thermodynamic in nature (i.e., the explicit recognition of aggregates is not involved). The relation is aCe) where ~ <pee) = L... k;;' 1 = ee'P(cl, (k- k+-1) (2.30) Bk+l e k . (2.31 )
14 Statistical Thermodynamic Background The Bk+1 are the virial coefficients of Eqs. (2.14) and (2.27)~(2.29). Equation (2.31) may be derived as follows, using Eq. (2.13) and recalling that a = c when c --> 0: <p = a IC (a) fC In ~ = c=o d In ~ = 0 -_ fC (ddInIn ac - ) 1 dIn c ~(dIT/kT _ 1)dc -_ ~ ~ (~) Bk+l C k • dc o c k k;:' I The physical or experimental significance of <p(c) is <p = Because as In (a/c) = In y = (activity coefficient in osmotic solution). (2.32) Ksa s for s ~ 2, we have for the higher as' as(c) = KscSeS'P(C) (s ~ 2). (2.33) Using Eg. (2.31), we then find, to terms in c4 , a/c y = 1 + 2Bl c + GB3 = + 2Bi}c l + (4B4 + 3BlB3 + tBi)c 3 + ... (2.34) + 4Bl c + (3B3 + 8Bi}c l + ". 1 + 6Bl c + ". 1 + .... al/Klc l = 1 (2.35) a3/K3c3 = (2.36) a4/K4c4 = (2.37) To find the concentrations of aggregates of different sizes cs(c), s = 1,2, ... , we first put Eg. (2.23) into Cs = as o(IT/kT) and, after the differentiation, substitute Eqs. The result, to terms in c 4 , is cdc = 1 - 2K l c cZ/Klc l = c3/K3C3 = c4/K4C4 = + [8K~ (2.38) oas (2.34)~(2.37) - 3K3 - 2K l (b 110 - and 4blO )]C l (2.27)~(2.29). + [-40K~ + 30K l K 3 - 4K4 + 4K~(6bll0 - 20blO - bOlO) + 3K 3(6b lO - blOlO) + 2K l ( -20bio + 6b 30 - bllO + 6bllOblO)]C3 + ". 1 + (-4K z + b llO - 4b zo )c + [20K~ - 6K3 + 2K l (20blO - 6b llO + bOlO) + 20bio - 6b 30 + b210 - 6bllOb20]Cl + ". 1 + (-6Kl + bIOI - 6blO )c + ... 1 + . ". As a check, it will be noted that these equations satisfy Eq. (2.18): (2.39) (2.40) (2.41) (2.42)
Aggregation and Osmotic Pressure Virial Coefficients 15 Also of some interest are the "apparent" equilibrium quotients, in powers of c: c2/K 2ci = C3/K 3ci = C4/K4Ci = 1 + (b ll0 - 4b 20 )c + [2K 2(-2b llO + 4b 20 + b020 ) + 20bio - 6b 30 + b210 - 6bllOb20]C2 + ... 1 + (b l010 - 6b 20 )c + ... 1 + .... (2.43) (2.44) (2.45) Note that, as expected, these are all unity if the bj are all zero (ideal mixture of aggregates), or if c ---> O. Of course, the activity quotients ajKsa S == 1 under any conditions. Noninteracting Aggregates With very strong association forces between solute molecules, it may be a good approximation to neglect non-association interactions between aggregates. This is the "ideal mixture of aggregates" case already mentioned. However, even with this simplification we still have an imperfect solution of solute in solvent (i.e., fI/kT =1= c) because of the association interactions themselves. These interactions are contained in the Ks [Eqs. (2.19) and (2.20)]. In the equations above, we set all of the bj = o. This leads to relatively simple relations that we summarize here, for convenience. The osmotic pressure virial expansion is fI/kT = c + B2c 2 + B3C3 + ... with B z = -K z , B4 -3K4 = B3 = -2K3 + + 4K~, 18K 2K3 - 20K~, etc. (2.46) Equation (2.23) becomes fI/kT = a l Then Eq. (2.38) gives Cs + a 2 + a3 + .... = as for all s. In particular, because a 1 = a, we have (2.47) where c= Cl + 2c 2 + 3c 3 + .... Thus, although Cs = as for all s, we have c =1= a (for the reason mentioned at the beginning of this subsection). Using as = Cs == Ksa s [Eq. (2.20)], other expressions for fI/kT are fI/kT = Cl + C2 + C 3 + ... (2.48)
Statistical Thermodynamic Background 16 From Eqs. (2.34)-(2.37) or from (2.39)-(2.42), we have y = a/c = cdc = 1 - 2K 2c + + (8K~ (-40K~ + 30K 2K 3 - 4K 4)C 3 + ... + (20K~ 1 - 6K 2c + ... 1 + ... . c2/K 2c 2 = 1 - 4K 2e c3/K3e3 = c4/K4C4 = - 3K 3)C 2 - 6K3)C2 + ... (2.49) (2.50) (2.51) (2.52) To simplify notation, let us write Q2 for QOlO (a dimer), Q3 for QOOIO (a trimer), etc. Then, from Eq. (2.20), we have a = e = K as = Qs!V (Ql"l,)S = Qs AS. S S (Ql/VY S V (2.53) V Thus, the relative abundance of aggregates of size s is proportional to QsA s. The resemblance to Eq. (1.13) should be noted. We shall return to this result in the next section (where N is used for the size of an aggregate rather than s). Hard Spheres: Numerical Examples In this subsection we go to the opposite extreme and consider solute molecules that interact as hard spheres but do not associate except to form polymers. This rather paradoxical behavior is, in fact, exhibited by HbS. 7 ,8 Correspondingly, the critical concentration C e for polymer formation at 37°C is 2.5 mM, which is about 1000 times larger than the critical concentration of actin or tubulin (to form micro tubules). A very accurate semiempirical expression for flv/kT for hard spheres of volume v at a concentration c is 9 ve[1 flv + ve + (vef kT - (ve)3] The "density" ve = 1 is never reached; at close packing ve integrating the thermodynamic equation [see Eq. (2.13)J d (~) kT (2.54) (1 - ve)3 = ! (Dn/kT) e DC de (T constant) = 0.7405. By (2.55) T one can then deduce lny Table 1-1 gives values ofy concentration: 8 = ve[8 - 9vc = + 3(ve)2] (1- ve)3 (2.56) a/c for several values of vc. For HbS at the critical
Aggregation and Osmotic Pressure Virial Coefficients 17 Table 1-1. Hard-Sphere Activity Coefficients ve y = ale ve y = ale 0.001 0.01 0.02 0.04 0.06 0.08 0.10 1.00805 1.085 1.181 1.413 1.715 2.117 2.659 0.12 0.14 0.16 0.18 0.20 0.25 0.30 3.41 4.47 6.00 8.29 11.8 33.7 130.5 v = g-l 51.0 M- 1 = 0.16 g cm- 3 = 2.48 mM = Ce VC e 0.79 cm 3 (2.57) = 0.1264, y = ae/c e = 3.71. Thus the activity and concentration are quite different. For comparison, if we treat tubulin as a sphere and take Ce = 10 .uM, we have v = 0.736 cm 3 VC e = 7.36 X 10- 4 , g-I = 73.6 M- 1 Y = ae/c e = 1.0059. (2.58) Clearly, hard interactions are not significant at this low concentration. This suggests that any differences between a and C for tubulin and actin could be treated by the previous subsection. Incidentally, an alternate semi empirical expression for In y in Ref. 8 (in the form of a polynomial) gives essentially the same values of y presented here in Table 1-1 and in Eqs. (2.57) and (2.58). It is sometimes usefup·3 to employ the molality of solute (or mole ratio, which is practically the same thing) instead of C as the composition variable. In the simple special case 3 that the solvent has only one component (denoted a) and that the volume of solution is additive, (2.59) where v (volume of a solute hard sphere) and Va (volume of a solvent molecule) are constants, the relation between the two composition variables is Nv ( (2.60) 1+t vC=~=~- V ( is the volume ratio of the two components. In this special case, it is more natural to use ( than the mole ratio N/Na • On substituting Eq. (2.60) into Eq. (2.54), TIv kT (1 + 4( + 6(2 + 2(3) 1+ ( (2.61)
Statistical Thermodynamic Background 18 On using the Gibbs-Duhem equation (Ref. 2, pp. 364-368), the mole-ratio activity coefficient y' is found to be 2 [---In(I+~)-1+2~--~ 1 3 2J . 9 + lny , =6~+-~ 2 1+ 2 ~ (2.62) This expression is valid provided that ~ < 1. The quantity in brackets is of order ~3. Finally, we give the power series expansions of Eqs. (2.54)9 and (2.61): IIv - kT = vc +L 00 i=l (3i + i2)(vc)'+1 • (2.63) (2.64) 3. Partition Function for an Open, Independent Aggregate We begin with some needed statistical mechanical background that is a continuation of Section 1. Equation (1.8) gives the grand partition function 3(11, V, T) for a one-component system of volume V that is open with respect to the number of molecules, N. That is, 11 is the independent variable, whose value is determined by an outside reservoir of molecules,and N fluctuates. The size of the system is fixed by V We can go one step further and allow fluctuations in V as well as N (i.e., p becomes the independent variable, not V). In this case, the partition function 4 is 1(11, p, T) = L Q(N, V, T)2 N (e- p / kTlV • (3.1) N,V The independent variables are nominally 11, p, and T. However, for a onecomponent macroscopic system, only two of these three intensive variables can be independent. This leads to difficulties 1 in handling this partition function. However, the difficulties disappear if 1 is applied to a finite (small) system 4 rather than to a macroscopic system. For example, if l1e is the chemical potential of the macroscopic system at the given p and T, then we must choose 11 < l1e to make the system finite. With this limitation, 11, p, and T can all be independent variables. Whereas the macroscopic system at l1e' p, T has an indeterminate size (there is no independent extensive variable), the small system has well-defined values of N and V determined by p, T, and 11 < l1e' This kind of system is called "completely open" because no extensive variable has a fixed (or assigned) value. An example of such a system is a small crystallite at p and T in equilibrium with vapor at f.1 < f.1e'
Partition Function for an Open, Independent Aggregate 19 The fundamental statistical thermodynamic relations 4 are (3.2) tS'= -kTlnY=E- TS+pV-J1N dtS' = -SdT + Vdp - NdJ1 (3.3) N __ (atS') _kT(alnY) _A(alnY) aJ1 p.T aJ1 p,T aA p,T = ~ L Y N,V NQAN(e-pVlkTt, (3.4) (3.5) with analogous relations for V. tS' is a new thermodynamic function that does not appear in macroscopic thermodynamics. Equation (3.5) and the corresponding equation for V show that (3.6) is the probability that the completely open system will be observed to have the values N and V at given J1, p, and T. It then follows, on differentiating Eq. (3.5), that the variance in N is 2 (IN = -N2 - N-2 = kT (aN) - aJ1 p.T = A (aN) - aA p,T . (3.7) There is a similar expression for (J~. For small systems that are assumed to be incompressible (for example, one might reasonably use this approximation for the crystallite mentioned above), V is always proportional to N and hence is not a separate operational extensive variable. The properties of such a system do not depend on the pressure p. In this case, we have the simpler relations [using the notation QN(T) now in place of Q(N, T)]: Y(J1, T) = L QN(T)A (3.8) N N tS' = -kTin Y, N = A(aln Y) 8A dtS' T' = -SdT - NdJ1 (J~ = PN = QNANjY. A(aN) 8A T (3.9) (3.10) (3.11) The last equation gives the probability that the system contains N molecules. Equation (3.8) resembles Eq. (1.8) (grand partition function), but there is a fundamental difference: in Eq. (1.8) the size of the system is fixed by the specification of V whereas in Eq. (3.8) the system is completely open and its mean size N is determined by how close J1 is to J1e (as J1--+ J1e' N --+ (jJ). Fluctuations in N in Eq. (1.8) are normal (small), but they are large in Eq. (3.8) when J1 is near J1e' Equation (3.8) is fundamental for much of the remainder of the book.
20 Statistical Thermodynamic Background Independent Aggregate in Solution As at the end of Section 2, we consider equilibrium aggregates, in a solution of volume V, that are sufficiently dilute to be regarded as independent of each other (except for solute exchange with the same pool offree solute molecules). Aggregating forces are strong and the degree of polymerization (i.e., number of subunits, or monomers, or solute molecules in the polymer) is usually in the range 10 2 to 104 , or more. Because of the independence of the aggregates or polymers, we can select a single aggregate as the small or finite statistical thermodynamic system of interest (analogous to the crystallite above). The entire collection of aggregates in V then has the status of an ensemble of such systems. A time average of some fluctuating quantity over the states of a single system (polymer) will then be the same as an instantaneous ensemble average over the whole collection (in the limit that the time and ensemble size both approach infinity). Implicit in the above comments is the following point of view that we adopt in the remainder of the book (except in part of Section 8): our primary interest is in large polymers that we assume are dilute enough to be independent of each other; the large polymers are in contact with a pool or reservoir of ("free") solute molecules that may contain small clusters (dimers, trimers, etc.) or whose molecules may otherwise interact with each other (e.g., hard-sphere interactions), or both. That is, although the polymers are assumed independent of each other, we do not necessarily make the same assumption about the constituents of the pool or reservoir of free (i.e., non-polymer) solute molecules. See the last two subsections of Section 2 in this connection. We now denote the partition function Qs in Eq. (2.53) by QN' This is the partition function of a single aggregate of size N (degree of polymerization, number of monomers or solute molecules in the aggregate) in the volume V, with the properties ofthe solvent appearing implicitly in QN through J1a' J1p, ... [see Eqs. (2.5) and (2.6)]. QN includes (Sections 4 and 8) translation, rotation, vibration of the N subunits relative to each other, internal vibration within each subunit, and the intermolecular interactions between neighboring subunits. The translational partition function is proportional to V. Because the intrinsic properties of the polymer (e.g., its size distribution) have nothing to do with the volume V of the container in which the polymer moves, it is preferable to use QN/V instead of QN in our definition of Y: Y(J1, T) = L (QN/V)A N, (3.12) N where J1 (in A = e IL1kT ) is the chemical potential of solute molecules (Section 2). The solvent variables J1a' J1p, ... are implicit in QN and Y, and are not shown in Eq. (3.12). The pressure is determined by all the chemical potentials and T. Equation (3.12) is an example of Eq. (3.8). Hence, Eqs. (3.9)-(3.11) apply. The summand in Eq. (3.12) is in fact the same expression as in Eq. (2.53). Therefore, Y in Eq. (3.12) has the physical significance of LN eN (the total
Partition Function for an Open, Independent Aggregate 21 concentration of all polymers). A similar relationship, using different composition variables, is discussed in detail (in its thermodynamic aspects) in Ref. 4, pp. 120-135. As A is increased in Eq. (3.12), terms at large N will eventually no longer converge. This occurs at /l = /le' where N -+ 00 (the polymer becomes macroscopic). The Chemical Potential For simplicity, we shall often assume that noninteracting monomers are in equilibrium with a sizable polymer (the small system) and that small clusters (e.g., dimers and trimers) are insignificant. That is, a = c. In this case, in Eq. (3.12), A= /l eJl/kT = (V/Ql)C = /l0(T) + kTln c (3.13) /lO(T) = -kTln(QdV), where Eq. (2.17) has been used and c is the concentration of free monomers (all other species, including polymers, make a negligible contribution to the total concentration c of solute). A less drastic assumption (see above) is that very dilute (independent) polymers are in equilibrium with interacting monomers or with small solute clusters (e.g., monomers, dimers, trimers), which in turn are in equilibrium with each other. In this case [Eq. (2.17)] (3.14) where the activity a is related to the total solute concentration c by Eqs. (2.27)(2.29) and (2.34) (the polymers are so dilute that they make a negligible contribution to c). Implicit in the small-cluster case (see Ref. 4, p. 121) is the not unreasonable assumption that some intermediate-size cluster or clusters (e.g., N = 5 to 7) represent a minimum in stability, have very small concentrations, and would present a free energy barrier to be surpassed in the polymer nucleation process (see Section 9). The physical significance of Ql' and hence of /lo, is discussed in detail in Ref. 10, pp. 6-12. In summary, Ql has contributions of the following types: (a) Translation and external rotation ofthe single solute molecule (monomer), just as for a single solute molecule in the gas phase. The translational partition function is proportional to V; hence Ql is proportional to V. (b) Internal vibration (and rotation) within the solute molecule, including any perturbations caused by the solvent molecules surrounding the solute molecule. (c) Intermolecular interactions between the solute molecule and the surrounding solvent molecules. (d) Perturbation of all the degrees of freedom and of the related intermolecular
Statistical Thermodynamic Background 22 interactions ofthose solvent molecules in the immediate neighborhood of the solute molecule as it moves about the volume V. A simplified example is included in Section 8. Independent Aggregate Attached to a Surface An independent linear polymer or aggregate with one end attached to a surface is a very important special case in cell biology. The polymer is immersed in solvent as before, but it is no longer free to rotate or translate. These degrees of freedom for a free polymer become additional vibrational degrees offreedom for an attached polymer. The partition function QN for an attached aggregate of size N (Section 5) does not include a factor V. Hence, Eq. (3.8), for Y, applies to this case without modification. Although Y is different (Chapter 2) for the same kind of aggregate depending on whether it is attached or free to move in solution, divergence of Y will occur at the same f.1 = f.1e in both cases. This follows because f.1e is a property of the macroscopic polymer, for which free rotation and translation of the entire polymer, if present, are inconseq uen tial. A more general situation for an attached polymer is shown in Fig. 1-2(a). Here we assume that the polymer is a straight rod of length L and that an external (axial) force F acts on the "free" end. F is (arbitrarily) taken as positive if it is an extending force and negative if it is a compressing force (the opposite of the pressure p in a conventional thermodynamic system). Subunits can still exchange with the polymer, at one or both ends. The attached polymer discussed in the preceding paragraph is the special case F = O. The critical value f.1 = f.1e (where N -> 00) depends on F (Section 4). This is also a completely open system, but with independent variables f.1, F, T The appropriate partition function is [in Eg. (3.1) replace V by Land p by -F] Y(f.1,F, T) = I N,L QN(L, T)AN(eF/kT)L. (3.15) Also, f!= -kTlnY=E- TS-FI-f.1N F(fixed) (a) (b) (3.16) Fig. 1-2. (a) Linear rod-like polymer of length L under a fixed force F (positive in the direction shown). Double arrows indicate possible monomer exchange at the polymer ends. (b) Polymer between two rigid barriers a fixed distance L apart.
23 The Macroscopic Aggregate as a Limiting Case dt&' - _ -SdT - IdF - Ndp = (Oln Y) I N -..1.:11' 2 (iN = = F. T (J1l (ON) ..1.;;-;Ull 2 ,(iL = (3.17) Y) kT(oln :1 uF (3.18) !l. T (OL) kT - of F.T . (3.19) !l.T Fluctuations in N are large, as are fluctuations in L (because L is essentially proportional to N). For a given N, L has small fluctuations, associated with the vibration of subunits relative to each other (i.e., associated with the slight axial compressibility of the rod-like polymer). A detailed discussion of an example of dual fluctuations of this type, in a completely open system, is given in Ref. 4, pp. 90-94. Another system of interest is shown in Fig. 1-2(b). Here the polymer is attached to (or pushing against) surfaces (rigid barriers) at both ends. In this case, p > Pe (Section 7). The length L is fixed. The independent variables are p, L, T. This system is not completely open because the extensive variable L is fixed, as is V in Eq. (1.7). N -+ 00 is not possible. The grand partition function IS I 3(p, L, T) = QN(L, T)A N. (3.20) N Also, FL d(FL) = = -SdT - A(0In3) N = -kTln3 ~ + FdL 2 T.L ,(iN = (3.21) - Ndp (ON) aT A (3.22) T.L . (3.23) The fluctuations in N are normal (small). Large fluctuations occur only in completely open systems because the polymer size for such systems becomes indeterminate as the critical concentration of solute is approached. Many of the above relations will be needed and illustrated in Chapter 2. 4. The Macroscopic Aggregate as a Limiting Case At the critical concentration C e of solute, an open aggregate becomes macroscopic in size (N becomes very large but has no definite value). Thus, at C e , we have a conventional two-phase equilibrium between a one-dimensional solid (the aggregate) and free solute molecules. 4 • 11 The equilibrium conditions are the same whether the aggregate is free or attached to a surface at one end (because the six rotational and translational degrees of freedom of the free aggregate become negligible compared to all the other degrees of freedom
Statistical Thermodynamic Background 24 when N --+ 00). The concentration Ce may be regarded as the solubility of the macroscopic polymer in the solvent. Let 1l0(T) be the chemical potential per subunit (solute molecule) of the macroscopic polymer. A particular statistical mechanical model would provide an explicit expression for 1l0(T) (see the end of this section). Because of the equilibrium between polymer and free solute molecules, the chemical potentials of these molecules in the two phases must be equal: (4.1) where Eq. (3.14) has been used and the activity a e is a function of Ce as in Eqs. (2.27)-(2.29) and (2.34). The subscript e always refers to a macroscopic equilibrium; usually we are interested in a nonmacroscopic equilibrium involving finite aggregates at C < Ceo If we can neglect small clusters and other intersolute interactions in the free solute (as it is customary to do for simplicity), a e = C e and (4.2) If we use K(T) to denote the equilibrium constant for adding solute molecules to the macroscopic polymer, then solute(in solution) +:t solute(in polymer) K(T) = = e-I!.Go/kT = e[!'O(T)-!'o(T)]/kT (4.3) lla e ~ liCe· Besides depending on T, the equilibrium constant K also is a function of the solvent chemical potentials Ila' IIp, ... (Section 3), but we leave this as implicit. The critical concentration Ce marks a separation point between two polymer regimes along the C axis. If C > C e and C is held fixed, the polymer will grow steadily; this is a steady-state system, not an equilibrium one (Chapter 2). If C < C e (c is held fixed) and we start with a very large polymer, the polymer will shorten at a steady rate (this is also a steady-state system). However, if C < C e has a value rather close to Ce, the polymer will eventually stop shortening at some definite sizable mean finite size (depending on c). This is an equilibrium system (Section 3). Macroscopic Aggregate under a Force F We consider the macroscopic thermodynamics (Ref. 12, pp. 44-48) of the polymer in Fig. 1-2(a), which has a length L, is subject to an external force F, and contains N molecules (or subunits). As in Eq. (l.3), we have dA = -SdT + FdL + IlpdN, where the subscript p on II p refers to the polymer (Il p (4.4) = 110 when F = 0, as
25 The Macroscopic Aggregate as a Limiting Case above). Integration of Eq. (4.4), holding intensive properties constant, gives A = FL + /lpN, dG = - G == /lpN = A - FL (4.5) and then + /lpdN (4.6) d/l p = -(SjN)dT - (LjN)dF. (4.7) SdT - LdF The important result for present purposes is d/l p = -ldF (T constant) (4.8) where 1== LjN (length per subunit). We let 10 be the value of I when F = 0 (rest length per subunit; 10 = 6.15 Afor a microtubule). Whereas I is a function of F and T, 10 depends on T only. Again, as in Section 3, the solvent plays only an implicit role, through /la' /lp, .... We assume that the linear rod-like polymer is slightly compressible: the subunits in the polymer have an optimal spacing relative to each other, determined by the intermolecular forces, but this spacing can be altered somewhat by a compressing (F negative) or extending (F positive) force. The equation of state (i.e., the relation between F and L) for such a system can be written (4.9) Actually, we shall use only the first term in applications; the other terms (considered negligible) are included for generality and perspective. In fact, for many purposes even the first term is not needed: the polymer can be considered to be incompressible (h is very large, I ~ 10)' The coefficient h(T) is the Hooke's law constant. For simplicity, we assume that the polymer does not bend under a compressing force. In the case of actin, this implies use of a bundle of actin filaments. Our primary concern is to find the dependence of /lp on F (or I) because this effect will alter the critical activity a e or concentration C e [Eqs. (4.1) and (4.2)] when the polymer is in equilibrium with free solute. Because ofthe form of Eq. (4.9), the simplest procedure is to rewrite Eq. (4.8) as d/l p = - (l - 10 dF + 10) d(l _ 10) d(1 - 10)' Then, using Eq. (4.9) to obtain the derivative, integration from zero to I - 10 gIVes /lp = /lo - 10F - WF2 10)F - ih 2 (l- 10)3 - ... ~ /lo - 10F - 2h ~ /lo - 10F. (4.10) (4.11) The first correction term in Eq. (4.10), -loF, is of order hlo(l - 10)' the second
26 Statistical Thermodynamic Background is of order h(l - 10)2, the third is of order h2(l- U 3 , etc. Usually it suffices to use /1 p = /10 - 10F (as for an incompressible polymer). In some cases, the term -F 2/2h [Eq. (4.11)] provides a small correction. The ratio between 10F and F2/2h is 210/(l - U which is large. The chemical potential /1 p is increased [Eq. (4.11)] when the polymer is under compression (F negative). That is, the subunits in the polymer are less stable than at F = 0 and will have a greater "escaping tendency" (to relieve the compressive force). Conversely, /1 p is decreased when the polymer is subject to an extending force: subunits in the polymer are more stable than at F = 0; this encourages subunits in the solution to add to the polymer (which will relieve the extensive force). We therefore expect the critical concentration Ce to increase under compression (c e is the "solubility" of the polymer, a measure of its escaping tendency) and to decrease under extension. As in Eq. (4.1), the formal relation that determines the connection between C e and F is (4.12) where the activity ae is a function of C e [Eq. (2.34)]. If we denote the value of ae at F = 0 by a~, then /10 + kTln a~ = /10 and kTln(ae/a~) = -loF - (F 2/2h). (4.13) Usually we can omit the term in F2. Then In a e = In a~ - (IJ /kT). (4.14) As a further approximation, a e ~ C e and a~ ~ c~. Thus, at least approximately, In C e depends linearly on F: In C e increases when the polymer is under compression (F is negative), etc. The (linear) thermodynamic compressibility can be defined in the conventional way: K (IOF)-l = ~(OL) = L of N.T 01 [hlo + (h + 2h2/0)(l- 10) + ... r ~ 1/hlo· = (4.15) 1 (4.16) If h -+ 00, K -+ 0 (incompressible polymer). In Fig. 1-2(b), where the polymer has grown at subunit activity a (with a > a~) so that it is in contact with two rigid barriers, L is held constant. When the polymer first touches the two barriers, F = O. The number of subunits in the polymer at this point is denoted by No (thus, 10 = L/No). As N increases by addition of further subunits at one or both ends of the polymer, F decreases (compression of the polymer) until Eq. (4.14) is satisfied (with
27 The Macroscopic Aggregate as a Limiting Case the activity a, of free solute, in place of ae in this equation). The polymer at the final F and N is at equilibrium with free solute at a. The above suggests that an alternative form of F = h(l - 10 ) should be useful for open aggregates when L is constant: F = - const. x (N - No). The constant is easily found: F = h(1 - 10 ) = hlo --(N - No) No = hL (~ - ~J (L constant) (4.17) to the linear term in N - No. A dimensionless form of this equation is IJ/kT = (4.18) -y*(N - No)' where (4.19) is a force constant of sorts. The asterisk is used to avoid confusion with the activity coefficient y. Another way to write Eq. (4.18), using Eq. (4.16), is (4.20) This shows the three ingredients on which N - No depends, N - No being the number of subunits that must be inserted into a polymer of compressibility K, and originally with No subunits, in order to raise F = 0 to - F, keeping L ( = 10No ) constant. In Fig. 1-2(b), the value of - F is determined by the preassigned values of a (for a e ) and a~ in Eq. (4.14). Incidentally, the next higher term in Eq. (4.17) is easy to find: - N0 ) F = -hlo ( N No - N + (hl o + h2/~) (N No 0 )2 + ... (L constant) (4.21) Numerical Examples We illustrate some ofthe above quantities using a microtubule as an example. 12 The same qualitative conclusions would be drawn about polymerized HbS or F-actin bundles. The value of the Hooke's law constant, h, for a microtubule has been estimated as 1.1 x 10 4 dyn cm- l (Ref. 12, p. 47). The length per subunit (tubulin dimer) is 10 = 6.15 A. The largest force likely to be of interest is F = 5 X 10- 6 dyn (this extending force, per microtubule, suffices to stop chromosome movement in anaphase l3 ). Most in vivo forces are much smaller than this. In Eq. (4.13), the ratio of the second correction term to the first, using the above numbers, is F2/2h loF = F 2h1o = 0.0037.
Statistical Thermodynamic Background 28 Hence, even allowing for some error in the estimate of h, the second correction term in Eq. (4.13) can be neglected: in Eq. (4.14) the polymer is treated as incompressible (h does not appear). With the above values of F and 10, 10F/kT = 7.5 at 25°C. This corresponds to ae/a~ = 5.7 x 10- 4 in Eq. (4.14). For F = 10- 7 dyn, a more common order of magnitude, 10F/kT = 0.15 and ae/a~ = 0.86. For F = _10- 7 dyn (a compressing force), ae/a~ = 1.16. Similarly, a ratio ae/a~ = 5 corresponds to lJ/kT = -1.6 and F = -1.1 X 10- 6 dyn. Thus, typically, 101F1/kT is of order unity. From Eq. (4.17), N - No No -F hlo ' where N - No is the number of extra subunits that need to be inserted in a polymer of No subunits, held at constant length, in order to increase the force from zero to - F. This fraction will be very small in cases of interest because it is just twice the ratio of correction terms examined above. Thus, (N - No)!No is -0.0074 for F = 5 X 10- 6 dyn and it is 0.0016 for F = -1.1 X 10- 6 dyn. In the latter case, if No = 2 X 104 (a microtubule oflength 12.3 Jlm), N - No = 32. That is, insertion of only 32 subunits in this microtubule, held at constant length, would increase F = 0 to - F = 1.1 X 10- 6 dyn. The value of the compressibility K, from Eq. (4.16), is 1.5 x 10 3 dyn- 1 and the value of y* in Eg. (4.19), if we take No = 2 X 104 , is 0.05. Equation (4.18) or (4.20) can be used to check N - No = 32 for the example given above. The Critical Concentration of a Simple Einstein Model In Chapters 2 through 6, we shall use extensively a simple Einstein modeF of a linear crystal to represent finite aggregates with various structures. As an introduction, we consider here the limiting case of infinite (very long) linear aggregates. The Einstein model will lead to a simple expression for the critical activity a e (or concentration ce) via Eg. (4.1) or (4.2). The same approach applies to the Einstein model of a three-dimensional crystal in a solvent; in this case Ce is generally referred to as the solubility of the crystal. For illustrative purposes, we consider the five examples of linear polymers in Fig. 1-3. Because the polymers are assumed to be very long, it is immaterial whether they are free in solution or attached to a surface (i.e., translation and rotation of the polymer as a whole are negligible, thermodynamically, in any case). The subunit (monomer) in all these cases is assumed to be an essentially isotropic spherical protein molecule. A free monomer in the solvent has three translational, three rotational, and a large number of internal vibrational degrees of freedom.2 Together with interactions between protein and solvent, these degrees of freedom determine JlO(T) in Eg. (3.13) via Ql/JI, where, it will be recalled, Ql == 'l'd'l'o in Eg. (2.5) [for details, see Ref. 10, pp. 6-12 and the discussion following Eg. (3.14)].
29 The Macroscopic Aggregate as a Limiting Case z = :2 z= 3 z=4 (a) (b) (c) w w z=4 z=6 (d) (e) Fig. 1-3. Sections oflong idealized polymers comprised of isotropic spherical monomers. Nearest-neighbor interactions have a free energy w; each monomer has z neighbors. (a) Single strand. (b) Two strands, aligned. (c) Two strands, staggered. (d) Eight strands in a tube, aligned. (e) Four strands forming a 2-start helix (flattened for clarity). The dotted monomers are repeats of monomers in the left-hand strand. The protein molecules are packed together in various ways in the hypothetical polymers shown in Fig. 1-3. In Fig. 1-3(d), rings of eight molecules are aligned on top of each other. Figure 1-3(e) illustrates a polymer comprised of four strands with a 2-start helical structure such that each molecule has six nearest neighbors. For example, six open circles are nearest neighbors of the shaded circle. The value of z given in the figure in each case is the number of nearest neighbors. The lines in the figure represent the nearest-neighbor interactions of an arbitrary subunit. Each interaction has a free energy w, which is negative. That is, w is the isothermal reversible work necessary to bring two subunits together in the solvent, to the neighbor distance in the lattice, starting with the subunits widely separated in the solvent; w is also called the potential of mean force 1 ,2 between two subunits. The interaction free energy per subunit in each of the structures in Fig. 1-3 is then zw/2 (division by 2 is required because each interaction is shared by two subunits). The larger the value of z, the more stable the polymer is. In the Einstein model, 2 we assume that each subunit in the polymer has
Statistical Thermodynamic Background 30 a partition function q that represents: intermolecular interactions between the free subunit and the solvent; the internal vibrations (and rotations) within the subunit; three vibrational degrees of freedom of the center of mass of the subunit in the potential field of its lattice neighbors; and three rocking degrees of freedom, also in the potential field of its neighbors. The latter two contributions correspond to (i.e., are residual forms of) the unrestricted translational and rotational degrees of freedom, respectively, of the free monomer. An explicit example of q is discussed in Section 8. Of course, q would not be the same in all of the examples in Fig. 1-3 even if w is the same: the smaller the value of z, the looser the vibrational and rocking motion in the field of the neighbors, and hence the larger q. With the above ingredients, the complete canonical partition function for a large polymer with N subunits is then (4.22) From [Eq. (4.4)] dA = -SdT + 1l0dN A = -kTlnQ (4.23) we find JlnQ 110 = -kT~ = _kTln(qe-zw/2kT). (4.24) Then, from Eq. (4.1), (4.25) (4.26) In some cases, we can put ae ~ C e • The polymer is more stable and ae (c e is the "solubility" of polymer) is smaller when - w is large (e.g., actin and microtubules versus HbS), z is large, and q is large. A larger q would be associated with a smaller z, as already mentioned, but this is a second-order effect compared to the effect of z in ezw/2kT. On the other hand, a e is increased by a larger QdV (free monomer more stable). The contributions of the large number of internal vibrational degrees of freedom to Q1 and q would cancel, to a good approximation. Perhaps the most important conclusion from Eq. (4.26) is that, other things being equal, a polymer with larger values of - wand of z (more neighbor interactions) will have a smaller critical concentration. In a tubular polymer, the maximum value of z is 6 (close packed, two dimensions), as in Fig. 1-3(e). In HbS, a solid rod of 14 strands, not all subunits are equivalent but the average value of z is greater than 6. Of course, real polymers do not have spherical isotropic subunits. Equation (4.26) also applies (in the Einstein model) to a three-dimensional crystal with solubility Ce' If the lattice is close packed, z = 12. An explicit illustration of Eq. (4.26) is included in Section 8.
The MacroscopIc Aggregate as a LImItmg Case 31 We assume throughout the book, for simplicity, that subunits exchange at polymer ends only, not in the body of the polymer. To lose a subunit from the polymer body (thus creating a vacancy for an incoming subunit) requires that z neighbor "bonds" be broken. The number of bonds broken in the loss of a subunit from a polymer end is at least one less than z (the end may not be "smooth"; see Chapters 5 and 6). Each broken bond "costs" a free energy - w. For a polymer to exist in the first place, - w cannot be small. We therefore assume, on energetic grounds, that subunit exchange with the polymer body is negligible. References 1. Hill, T.L. (1956) Statistical Mechanics (McGraw-Hill, New York; also Dover, New York, 1987). 2. Hill, T.L. (1960) Introduction to Statistical Thermodynamics (Addison-Wesley, Reading, MA; also Dover, New York, 1986). 3. Hill, T.L. and Chen, Y. (1973) Biopolymers 12, 1285. 4. Hill, T.L. (1964) Thermodynamics of Small Systems, Part II (Benjamin, New York). 5. Onsager, L. (1949) Ann. N.Y. Acad. Sci. 51, 627. 6. Minton, A.P. (1977) J. Mol. BioI. 110, 89. 7. Ross, P.D. and Minton, A.P. (1977) J. Mol. BioI. 112,437. 8. Ferrone, F.A., Hofrichter, J., and Eaton, W.A. (1985) J. Mol. BioI. 183,611. 9. Carnahan, N.F. and Starling, K.E. (1969) J. Chem. Phys. 51, 635. 10. Hill, T.L. (1985) Cooperativity Theory in Biochemistry (Springer, New York). 11. Oosawa, F. and Asakura, S. (1975) Thermodynamics of the Polymerization of Protein (Academic, New York). 12. Hill, T.L. and Kirschner, M.W. (1982) Int. Rev. Cytol. 78, 1. 13. Nicklas, R.B. (1983) J. Cell BioI. 97, 542.
2 Attached Single-Stranded PolYlIler In the remainder of the book, except for Chapters 5 and 6 and Section 24, we shall treat linear polymers formally as if they consist of a single strand only, as in Fig. 2-1 (a). To be more precise: we shall assume that there is only a single subunit attachment or departure site at a polymer end or if there are several such sites, that they are all equivalent. In effect, then, there is a single overall on rate constant for a polymer end and a single off rate constant, and these rate constants are constant (see below). This model would be exact for the structure in Fig. 2-1 (a) and it would also be exact in Figs. 2-1 (b) and 2-1 (c) if the intersubunit interactions in the polymer are so strong that there is always only one significant addition site (see the arrows in the figure) and only one significant departure site despite the fact that there is more than one strand. Figure 2-1(b) illustrates a I-start, 2-strand helical structure (as in actin) and Fig. 2-1 (c) shows a I-start, 3-strand helical structure (flattened). In the limiting case just mentioned, both structures would behave kinetically like a single helix (i.e., in effect, a single strand). This would be true of any I-start tubular helical polymer in the strong-interaction limit. In general, however, if a polymer has several strands, it will also have several on and off sites at an end, with associated fluctuating rate constants that depend on the instantaneous arrangement of subunits at the end (see Chapters 5 and 6). Such polymers will have average overall on and off rate constants but these rate constants will, in general, not be constant: they will depend on the free subunit concentration (Chapters 5 and 6). However, as an approximation, such polymers will be included in the single-strand model that we adopt here and use in most of the book. It should be added that this approximate treatment of multistranded polymers (e.g., actin, HbS, microtubules) as operationally single-stranded is, in fact, conventional throughout the research literature. This is, of course, done for simplicity, on the assumption that the approximation made is not very serious. As will be apparent from this book,
33 Attached Polymer at Equilibrium or Steady State Fig. 2-1. Polymer ends with only a single subunit attachment or departure site. The total interaction free energy between the terminal subunit and its neighbor or neighbors is w. The arrows refer to transitions that are possible (1) or not possible (-t). (a) Single strand. (b) 1start, 2-strand structure. (c) 1-start, 3-strand helix (flattened). Subunit A is shown twice. HH H IV IV IV HUH ..} ::.~ ..... (a) (b) (c) many of the problems that arise, even for an effective single-strand polymer, are already quite complicated. The approximation just mentioned is, however, avoided in Chapters 5 and 6 and Section 24. The additional complexity that ensues is considerable. 1. 2 We adopt, then, Fig. 2-1(a) as our simple working model ofa linear polymer. The model should not be taken literally in the sense that, with physical and not chemical forces holding the subunits together, such a structure (with z = 2 only, Fig. 1-3) would tend to be unstable. The free energy w in Fig. 2-1(a) is the interaction free energy between the terminal subunit and its neighbor (or neighbors-see below). It is also the interaction free energy per subunit in the bulk polymer. If Fig. 2-1 (a) is used as a working picture to represent Fig. 2-1(b) or Fig. 2-1(c), then w includes several interactions, as shown in the latter figures. If a complicated structure such as a microtubule (13 strands) is represented by Fig. 2-1(a), as an approximation, w should be considered to be the interaction free energy per subunit in the bulk polymer. This is one-half of the interaction free energy between one particular subunit and all of its neighbors [i.e., zw/2 in the notation of Eq. (4.22)]. On the average, when one subunit is added to the end of the polymer, the end does not change but one subunit has been added to the interior ofthe polymer. Hence the average interaction free energy change is just the interaction free energy per subunit in the bulk polymer. This chapter (Sections 5-7) is concerned with a single-stranded equilibrium polymer attached to a surface (e.g., grown from a nucleating site on the surface) or extending between two surfaces (Section 7). Chapter 3 treats the same kind of polymer free in solution, while Chapter 4 introduces the complications of two-component polymers, binding on subunits, caps on a polymer end, etc. 5. Attached Polymer at Equilibrium or Steady State We begin by considering the equilibrium properties 3 of the single-stranded polymer with N subunits shown in Fig. 2-2. The polymer is in contact with a pool offree subunits at activity a (or concentration c). The polymer is attached,
Attached Single-Stranded Polymer 34 Fig. 2-2. Polymer attached to a nucleating site on a surface. The notation is described in the text. with an interaction free energy w', to a nucleating site on a surface; the intersubunit interaction free energy is w. All subunits have a partition function q (Section 4) except the distal terminal subunit, which has q' (the motion of this subunit is somewhat less restrained than that of the others so we would expect q' > q). A slightly more general model (not used) would also assign a different q to the first subunit, the one in contact with the attachment site on the surface. We neglect the slight motion of the entire polymer that is the residue of centerof-mass translation and overall rotation of the free polymer (Chapter 3). The basic statistical thermodynamic equations for this model are Eqs. (3.8)-(3.11). For the model as just described [compare Eq. (4.22)], Qo = 1, QN = q' e-w'/kT(qe-w/kTt-l (5.1) where q' e-w'/kT qe w/kT . (5.2) C=--~ Presumably - w' ~ - wand q' > q (see above) so C > 1; however, we shall usually take C = 1 for simplicity and because C, as an end effect, is not important in any case when N is large. The partition function for this completely open system (the polymer) is Y = L 00 QNAN = 1 + C N=O = 1+C f N=l L (qJee-W/kT)N 00 N=l xN = 1- x + Cx (5.3) 1- x where (5.4) Note that x ex:. Je and Je ex:. a (or c). The series in Eq. (5.3) converges for x < 1. The probability of a polymer of size N (i.e., with N subunits) is QNJe N Y PN = - - = The probability of an empty site is CxN(l - x) 1 - x + Cx (N ~ 1). (5.5)
Attached Polymer at Equilibrium or Steady State Po 1 35 1- x =-=----- Y 1- x + Cx (5.6) Because x < 1, PN falls off exponentially with N. As a consequence of PN oc x N , large polymers dominate as x ---+ 1. There is an equilibrium between free subunits and bulk polymer at Xe = 1, which value defines the critical activity a e or critical concentration Ce: Xe = 1 = q2 e e- w / kT = q(VIQ1)e- w /kT a e (5.7) (5.8) where we have used Eq. (3.14) for 2. Equation (5.7) is the same as Eq. (4.26) for this case (z = 2). For an arbitrary x, x = q2e- w /kT = q(VIQ1)e- w /kT a (5.9) = alae ~ clc e· Thus the operational meaning of x is alae or, approximately, clc e. The size distribution PN oc (clcet, is wellknown,3.4 but it is usually assumed, incorrectly, to apply to free polymers in solution. Free polymers have a somewhat different distribution, the difference arising from free translation and rotation of the polymer as a whole (Chapter 3). The distribution PN oc (alae)N is a property of immobile linear polymers only. The distribution PN = QN2 NIY refers to the fluctuations in size of a polymer on a single site over a very long period of time or to the polymer size distribution observed in a large ensemble of independent and equivalent surface sites at one particular time. On differentiating Eq. (5.3) twice, Eqs. (3.10) give - Cx N=----,---:------:- (1 - x)(1 - x (J~ &2 + Cx) 1 - x 2 + Cx 2 Cx (5.10) (5.11) The mean polymer size, &, is shown as a function of x for small & in Fig. 2-3. Large values of & occur only when x is very near 1. For example, if x = 0.99, & = 99 for C = 1 and & = 99.9 for C = 10. The relative variance in Eq. (5.11) is large, of order unity rather than 1/&, which is its usual order for open thermodynamic systems. For a typical open system, PN has a Gaussian peak centered at N = & and has a standard deviation (IN of order &1/2. Also, the value of & is proportional to the fixed size of the system, determined for example by V In the present completely open system, with no fixed size, the distribution PN is very different: when x ---+ 1 and & is large, PN is almost flat. H is interesting that Eq. (5.10) has exactly the same form as the BrunauerEmmett-Teller (B.E.T.) isotherm for the physical adsorption of a gas on a surface. This is a consequence of a very unrealistic model (independent piles of molecules) for the gas adsorption problem.
Attached Single-Stranded Polymer 36 Fig. 2-3. Mean polymer size N as a function of x = alae for C = 1 and C = 10. 4 3 2 0.6 x 0.8 1.0 = alae As already mentioned, we shall usually take C properties simplify to 1 y=-1 - x' X N=--, I-x PN = xN(I - x) (J2 N - x)2' 1. In this case, the above (N;::' 0) (J~ X (1 _ = -&2 1 X (5.12) (5.13) As an equilibrium system, the polymer properties derived above are completely independent of the kinetic mechanism involved in the formation and maintenance of the polymer size distribution. Thus, the various kinetic assumptions made in the remainder of this section have no influence on these equilibrium properties. Rate Constants and Subunit Flux Subunits go on and off the free end of the equilibrium polymer in Fig. 2-2. We assume throughout the book, except in the next subsection, that this subunit exchange involves monomers only. We now introduce monomer rate constants that are functions of temperature only (the solvent chemical potentials are implicit, as usual), not functions of the free subunit concentration c. It is convenient to begin with subunit exchange at the free end (the ri end) of a very long (bulk) polymer at equilibrium, that is, at c = Ce and a = a e . The "off" rate constant ri'(T) is defined as the mean number of subunits that leave the bulk polymer end per unit time. This is a first-order rate constant with units s-1, ms- 1 , etc., which, it is reasonable to assume, is indeed independent of the free subunit concentration as required. We then define the second-order "on" rate constant ri(T) as ri'(T)K(T), where
37 Attached Polymer at Equilibrium or Steady State K(T) is the equilibrium constant for addition of subunits to bulk polymer, defined in Eq. (4.3). Thus we have the conventional relation between equilibrium constant and rate constants: K = a/a'. Furthermore, because K = l/a., we have aa e = a'. This expresses the off-on detailed balance at the a end of the bulk polymer at equilibrium: a' is the rate of departure, as already mentioned, and aa e is the mean number of subunits that add to the bulk polymer end per unit time. The product aa e is a psuedo-first-order rate constant. At an arbitrary activity a and concentration c, the first-order off and on rate constants for the polymer end are then a' and aa, respectively, because a' and a are independent of subunit concentration c. Of course, the conventional approximation is to use ac in place of aa. The on and off rate constants for the end of a finite polymer, as in Fig. 2-2, will still be aa and a', respectively, because the properties of the polymer tip are independent of N provided that N :;?; 2. The partition function q' has no effect because one q' is present in QN before and after any transition. However, if we assume that the on rate constant for the empty site (N = 0) is also aa (e.g., because this rate is diffusion controlled), then the off rate constant for N = 1 must be a'/C (i.e., if C > 1, a' is reduced because the lone subunit when N = 1 has extra stability, through w' and q'). The above conclusions about rate constants are summarized in Fig. 2-4, where the individual states shown represent the polymer with various sizes N. At equilibrium, for a finite polymer (a < a e ), there must be a detailed balance between any two successive states. That is, the transition rates in opposite directions must be equal. These detailed balance relations can be used to verify that the rate constant assignments in Fig. 2-4 are consistent with Eqs. (5.5) and (5.6), which were obtained without consideration of kinetics or mechanism. Thus, from Fig. 2-4, aaPo = (a'/C)P1 , aaPN = a'PN +1 (N:;?; 1). (5.14) Then we find (Xa (X' P + a ae -N -1 = ~ = Ka = ~ = x PN P1 ~ Po (XaC aC (x' ae = - - = KaC = - (5.15) (N:;?; 1) = Cx. (5.16) These results are in agreement with Eqs. (5.5) and (5.6). Note that Eq. (5.14) is consistent with aa e = a' for a bulk polymer because PN = PN + 1 in the limit x--+l. -- ow N= 0 ex'ie ~ ex exa cw 2 3 ••. ~ ex Fig. 2-4. Linear kinetic diagram, with rate constants, for a polymer with N subunits.
38 Attached Single-Stranded Polymer When the above discussion of rate constants for a single-stranded polymer is used, as an approximation, for a multi-stranded polymer, IXa and IX' represent total on and off rates for the entire polymer end, including contributions from all strands. If the polymer end grows (a > ae ) or shrinks (a < ae ) at a steady rate, the net mean subunit flux (in subunits per unit time) is (5.17) Ja = IXa - IX' = IX(a - ae ) = IXc[l ~ + 2Bzc + (!B3 + 2Bnc z + ... ] - IXC - IX', IX' (5.1S) (5.19) where we have used IX' = IXa e and Eq. (2.34) for a(c). Thus a single-stranded polymer that exchanges only monomers at the IX end will have a flux Ja that is linear in a, but Ja would, in general, be only approximately linear in c. Note that Ja = 0 (bulk equilibrium) when a = ae and c = Ce, where Ce is the value of c that makes a(c) equal to IX'/IX. If an experimental plot of Ja(c) is actually linear in c, it can be concluded that the osmotic virial coefficient corrections in Eq. (5.1S) are negligible. In order to apply Eq. (5.17) to a shortening polymer (a < ae , Ja < 0), the polymer would have to be rather long to begin with. Eventually the polymer would stop shortening at a finite equilibrium size, with fluctuations, depending on the value of alae. Equations (5.17)~(5.19) would also apply to one end (the IX end) of a very long polymer free in solution because IX and IX', like K, would be the same for a long polymer whether it is attached or unattached (see Section 4). Comparison of the kinetic properties of the two ends of a long free polymer will be reserved for Chapter 3. Exchange of Dimers, Trimers, etc. at a Polymer End We shall usually assume, for simplicity, that only monomers exchange at a polymer end. In fact, we shall often go further, as is conventional, and take a = c and Ja = (Xc - (X'. However, in this subsection we recognize that the free subunit pool may contain dimers, trimers, etc., and that these might exchange, with significant rate constants, with the polymer end. Of course it is possible that an appreciable population of dimers, trimers, etc., exists but that the on and off rate constants for these species are so small that only monomer exchange is significant kinetically. In this case, the treatment in the previous subsection applies but, because of the dimers, trimers, etc., we would have a =I- c (Section 2). We now consider dimer, trimer, etc., exchange. We recall from Section 2 that a small cluster (dimer, trimer, etc.) of size s, at equilibrium, has the properties
39 Attached Polymer at Equilibrium or Steady State As = AS, as as = Ksa s, K s = QsA,/V (5.20) = Q,/V (Ql/V)" (5.21) where the notation Qs was introduced in Eq. (2.53) and Ks is the equilibrium constant for forming an s-cluster from monomers. Also, from Eq. (5.20), Ps + kTln as> (5.22) == - kTln(Q'/V). (5.23) = p~ where p~ Equilibrium aspects of the exchange of an s-cluster at the end of a macroscopic (bulk) polymer can be summarized as follows [as in Eq. (4.3) for monomers]: s-cluster (in solution) f=± s subunits in polymer P; = t1G~ = sPo(T) - p~(T) + kTln a; = p~(T), K(S)(T) = (5.24) sPo(T) e-IlGUkT = l/a;. (5.25) The equilibrium constant K(s) refers to the addition of an s-cluster to the end of a bulk polymer and a; is the value of as at a = a e , C = Ceo The first-order rate constant for the loss of an s-cluster from the bulk polymer IX end is denoted IX~(T). This is the mean number of s-clusters lost per unit time. We then define the second-order on rate constant for an s-cluster by IX,(T) = IX~(T)K(S)(T) so that K(s)(T) = IXs(T) IX~(T) , IXsa; = IX~. (5.26) K(1), Ql, 1X1' and 1X'1 here are the same as K, Qe' IX, and IX' in the previous subsection (for monomers). The latter of Eqs. (5.26) expresses the detailed balance for s-clusters at the IX end of a long polymer at equilibrium (a = a e , C = ce). The first-order on and off rate constants for s-clusters at an arbitrary a and c are then IXsa s and IX~, respectively, because IXs and IX~ are independent of c (or a). The on rate constant can also be written as IXsKsas. For a long polymer at an arbitrary a and c, the total subunit on rate from clusters of all sizes (s = 1,2, ... ), expressed in monomers per unit time, is on rate + 2IX z a z + 31X 3a3 + ... 1X1 a + 2IX z K z a 2 + 31X3K3a3 + .... = 1X1 a = (5.27) Similarly, the total off rate is [using Eqs. (5.21) and (5.26)] off rate + 21X~ + 31X~ + ... 1X1 ae + 21X2K2a~ + 31X3K3a~ + .... = 1X'1 = (5.28)
Attached Single-Stranded Polymer 40 Equations (5.27) and (5.28) assume that the small-cluster equilibrium is maintained while the polymers are out of equilibrium. Thus, the net subunit flux, in monomers per unit time, is Ja = on rate - off rate = a 1 (a - a e ) + 2a 2 K 2 (a 2 - a;) + 3a3K3(a3 - a~) + ... . (5.29) The first term on the right is the net monomer flux, the second term the net dimer flux (measured in monomers), etc. Note that each separate net flux is zero at a = a e (a consequence of detailed balance at equilibrium). Clearly, if there is significant dimer, trimer, etc., exchange, Ja will not be a linear function of a, and even less so of c. The rate constant ratios and equilibrium constants are not independent of each other. From a~ = Ksa! and Eq. (5.26), we have a'a: = Ks (a'a: )S = KK: = K(s)' 1 (5.30) where K == K(l) as already mentioned. Another fundamental relation, from Eq. (5.18), is (5.31) This equation determines ae and C e if a and a' are specified. Expressions for B2 , B3 , etc., are given in Eqs. (2.27)-(2.29). Finally, a word about the magnitude of the cluster rate constants. For an actin-like structure (z = 4), as in Fig. 2-1 (b), the departure of a monomer (in the a'l process) requires the breaking oftwo neighbor "bonds." The departure of a dimer, trimer, etc., requires the breaking of three such bonds. The corresponding free energy in each case represents, essentially, the free energy barrier for escape. Hence we would expect a'l to be considerably larger than a~, a~, ... . For a close-packed tubular structure (z = 6), such as Fig. 2-1(c) or Fig. 1-3(e), departure of a monomer, dimer, trimer (triangular), etc., requires the breaking of 3, 5, 6, etc., bonds, respectively. Hence monomer departure is even more strongly favored here (relative to dimer, etc.). Thus the conventional implicit assumption that a~, a~, etc., are small compared to a'l is probably a good one. From the relation a~ = asKsa;, already used in Eq. (5.28), we have to conclude that if a~ is small (by the above argument), then either as or Ks or both are small. In any case, the nonlinear terms in Eq. (5.29) would be relatively small. The unrealistic nature of our working model, Fig. 2-1 (a), becomes especially evident if the above considerations are applied to it: only one bond is broken for any s = 1,2,3, ... ; hence, roughly speaking, the polymer is as likely to break at one place as at another (including loss of a monomer).
41 Attached Polymer at Equilibrium or Steady State Exchange of Monomers at the Attached Polymer End We now generalize Fig. 2-4 (monomer exchange at the free IX end of the polymer) to allow monomer exchange at the attached {3 end as well, as shown in Fig. 2-5. Presumably the rate constants {3 and {3' are significantly smaller than they would be if the attached end were also free. For example, a subunit escaping from the {3 end would have to break the additional bond (w') between subunit and surface, thus reducing the off rate constant {3'. Similarly, the on rate constant {3 would be reduced because of the difficulty of inserting a subunit between the surface and the polymer end. The model in Fig. 2-5 would actually be quite unrealistic for a single-stranded polymer as shown, because the entire polymer is likely to escape from the surface following a monomer departure from the {3 end. However, the model becomes more realistic if the polymer has many strands (one subunit is lost at a time so the other strands could maintain attachment to the surface). Still more realistic would be exchange of subunits at the end of a polymer that is attached to a surface by insertion in a sleeve (Sections 7 and 25). Gain or loss of a subunit from the attached end of a very long polymer at equilibrium actually alters the number of bulk subunits and not the state of the attached end itself. The same is true, of course, for gain or loss of a subunit from the free end. The polymer has the same initial and final states no matter which end gains or loses a subunit. Thus the standard free energy change and equilibrium constant are the same for subunit exchange at either end. Hence we have 1 ae {3 {3' rx K=-=-=IX' IXa e = IX', (5.32) (5.33) {3a e = {3'. In general rx =f. {3 and rx' =f. {3', but the ratios rx/rx' and {3/{3' must be equal. With subunit exchange possible at either end, the kinetic diagram changes from Fig. 2-4 to Fig. 2-6. Equation (5.16) is unaltered but Eq. (5.15) becomes PN+1 -- = PN (rx + {3)a rx' + {3' = a ae Ka = - Fig. 2-5. Introduction of subunit exchange at the attached end of the polymer. = x (N?:- 1). (5.34)
Attached Single-Stranded Polymer 42 (a + ~)a aa N=O ......._ __ a'ie (a • 2 c/ + {3' + ~)a • 3 ... a' +~' Fig. 2-6. Kinetic diagram, as in Fig. 2-4, when subunit exchange is possible at both ends. As we should expect, the equilibrium distribution PN is unaffected by the presence of a new kinetic pathway (/3, /3'). For a long attached polymer, the separate subunit fluxes at the two ends are Ja = aa - a' and Jp = /3a - /3'. (5.35) The total subunit flux (both ends) is then J = Ja + Jp = (a + /3)a - (a' + /3'). (5.36) These three subunit fluxes are illustrated in Fig. 2-7. All three fluxes are zero at the same point, a = a e • 6. Attached Polymer in Transients Most of this book examines equilibrium and steady-state properties of linear polymers, rather than transients. This section is one ofthe exceptions. We deal here with transients in the kinetic model in Fig. 2-4, but we take C = 1 to simplify the mathematics. Also, again for simplicity, we consider that the polymers are dilute enough so that a and c for the free subunits do not change J~ a b===-"(-~==--:::=======­a .i5 Fig. 2-7. Subunit flux as a function of activity a for C( and f3 ends separately and together (J).
Attached Polymer in Transients 43 with time. That is, the rate constants aa and a' are both constants. The resulting model is a classical case in stochastic theory, often referred to as the immigration-emigration process or simple queue (N is the population size or queue size). Another description is a random walk in one dimension (N ?: 0), with reflection at N = O. Even for this simple model, some of the mathematics is rather complicated. Generally, we give results only, without derivations, which are available in the references. Basic Equations We consider a large ensemble of attached polymers, especially the probability distribution in N of these polymers. The value of PN at an arbitrary time t is increased in the next interval dt by transitions of polymers of sizes N + 1 and N - 1 and decreased by transitions of polymers of size N. From Fig. 2-4, we can express this explicitly as dPN(t) , , ----;It = a PN+I(t) + aaPN-I(t) - (aa + a )PN(t) (N?: 1) dPo(t) ----;It = a , (6.1) (6.2) PI (t) - aaPo(t)· We need a special equation for N = 0 (empty nucleation site) because states with N < 0 do not exist. These are the so-called master equations that determine the time dependence of this system (the ensemble of attached polymer molecules). There is a stationary (equilibrium) solution of Eqs. (6.1) and (6.2) at t = 00 if aa < a', that is, if x = aa/a' < 1. This is found by setting all dPN/dt = O. The solution is that already given in Eqs. (5.12)-(5.16) with C = 1. If aa > a', there is no stationary solution at large times: the mean of the distribution, N, increases steadily with time, and the distribution spreads as well (see below). If we multiply each term in Eq. (6.1) by N and sum both sides over all N ?: 1, we obtain L 00 N=I dPN dN d Nd- =d-INPN=-d =aa-a t t t I + a Po (t). I (6.3) To obtain the right-hand side, special care must be taken with terms in N = 0 and N = 1. When aa < a', Po(CX») = 1 - x [Eq. (5.12)]. If this is substituted into Eg. (6.3), we find dN/dt = 0 at t = CX) when x < 1, as we should expect for a stationary equilibrium distribution. The expression for poet) is quite complicated (see below) so it would not be easy to obtain N(t) by integrating Eq. (6.3), starting from, say, a J function at an arbitrary N = No at t = O. However, we know the final value N( CX») to be x/(1 - x) [Eg. (5.13)]. If aa < a' and the initial No is much larger than the final N (00) (i.e., the initial polymers are very long), then poet) will be negligible for a long time. During
44 Attached Single-Stranded Polymer this period, dN dt = Ja = aa - + (aa N(t) = No a', - a')t. (6.4) The polymers shorten at a steady rate Ja = dN/dt < 0, as in Eq. (5.17). When aa > a', the term a'Po in Eq. (6.3) will quickly become negligible (ifit is not negligible to begin with) and Eqs. (6.4) will again hold, with Ja > O. The polymers grow indefinitely at a steady mean rate. If we multiply Eq. (6.1) by N 2 and sum over N, in the same way as above, we find dN 2 - dt _ 2(aa - a')N(t) = and then, from (J~ = N 2 a' Po(t), _ = aa + a' - a' [l + 2N(t)]Po(t). ~ dt (6.5) N2 and Eq. (6.3), - d(J2 + aa + a' - (6.6) When aa < a', the Po(oo) and N(oo) expressions used above give dN 2/dt = 0 and d(JMdt = 0 at t = 00, as expected. When Po(t) is very small (e.g., long polymers), d(J~ dt = aa +a, I (J~(t) = (J~(O) + (aa + a')t. (6.7) For an initial b function, (J ~(O) = O. The distribution spreads with time in a simple way. When the important values of N are large, or as an approximation, N as well as t can be treated as a continuous variable. Then the distribution function is denoted P(N, t) and IX) P(N, t)dN = 1. (6.8) In Eq. (6.1), we substitute the Taylor expansions P(N + 1) = P(N - 1) = ap 1 a 2p ap P(N) - aN 1 a 2p + 2" aN2 P(N) + aN + 2" aN2 + ... (6.9) + ... and arrive at the Fokker-Planck differential equation ap at = 1 2(exa I a2p + ex ) aN2 - I ap (exa - ex ) aN· (6.10) The physical significance of the combinations aa - a' and aa + a' has already appeared in Eqs. (6.4) and (6.7).IfEq. (6.10) is regarded as a diffusion equation,
Attached Polymer in Transients 45 the coefficient of az p/aN z is the diffusion coefficient D. The evolution of peN, t) with time, in this approximation, is now governed by Eq. (6.10). If we start with an ensemble of long polymers, all of size No at t = 0, we can ignore the reflection at N = 0 (except at long times, when x < 1). In this case, the solution of Eq. (6.10) is the Gaussian function peN, t) 1 = [21UTMt)] l/Z exp [N - N(t)]Z} { 20'~(t) (6.11) , where N(t) is given in Eq. (6.4) and O'~(t) in Eq. (6.7), with O'~(O) = O. The peak of the distribution, at N = N(t), moves at the steady rate J~ = aa - a'. At the same time, the Gaussian (with standard deviation O'N) spreads, with O'~ increasing at the steady rate aa + a'. If monomers can be exchanged at the attached f3 end of the polymers as well as at the free a end, as in Fig. 2-6, then in Eqs. (6.4) and (6.7) aa is replaced by (a + f3)a and a' by a' + f3'. The same changes are made in Eqs. (6.10) and (6.11). As a slightly more complicated case, suppose both monomers and dimers can exchange at the free polymer end only. The kinetic diagram now appears as in Fig. 2-8, where the notation is taken from Section 5. The master equation for N ~ 3 is - (a~ + a'l + ala + aZaZ)PN' (6.12) In the continuous-N approximation, peN + 2) = peN) ap azp aN aN + 2 - + 2 -z + ... (6.13) ' etc., so that The coefficient of oP/oN is the subunit flux J~ [Eq. (5.29)]. Fig. 2-8. Kinetic diagram for a polymer that can exchange monomers and dimers at the free end only. 4 ... 2 N= 0 3 ... '"
Attached Single-Stranded Polymer 46 PN(t) with Reflection at N =0 We return now to the master equations, (6.1) and (6.2). The general solution, starting with PNo = 1 at t = 0, is not easy to derive (Ref. S, pp. 249-2S0 and Ref. 6, pp. 193-196) and the final result is complicated. For any N ~ 0, PN(t) = x(N-No)/2e-(aa+a')t[IN_No(2t~) + X- 1/2 1N+No+l(2t~) + (1 - x) I j=2 x- j/2 1N+No+i2t~)], (6.1S) 1v(z) = Lv(z) (6.16) where x = aa/a', 1v(z) = 00 (z/2)" + 2, Jo r!(v + r)! -+ (2nz) 1/2 eZ (z (v ~ 0) -+ 00), (6.17) (6.18) fv(z) is a modified Bessel function. When x < 1, PN( 00) is given by Eq. (S.12). To see how this result follows from Eq. (6.1S), as t -+ 00, requires "appreciable further analysis" (Ref. 6, p. 196), which we do not pursue. Equation (6.1S) simplifies considerably if x = 1, that is, if a = a e and aa e = a'. In this case, the rate constants are the same in both directions (Fig. 2-4). Then (6.19) If the initial ensemble consists of empty nucleation sites (No = 0), and x = 1, (6.20) Figure 2-9 shows Po(t) and Pi (t) at early times, from Eq, (6.20). As t -> 00, Eq. (6.19) or (6.20) gives PN -+ 1/(na't)1/2 for any N. Because x = 1, the distribution tends to spread uniformly to zero over the entire half-space (N ~ 0). Equation (6.26), below, with t -+ 00 gives more detail about this asymptotic behavior. We turn now to the solution of the same problem (N = No at t = 0; reflection at N = 0) in the continuous-N approximation. Equation (6.10) has to be solved with the boundary conditions just mentioned. To simplify notation, we use here 1 D = "2(aa + a'), Ja = aa - a', (6.21) where D is the diffusion coefficient in N -space. The solution is found to be (Ref. 5, p. 52 and Ref. 6, pp. 223-225, with typographical errors corrected) 1 { [(N - No - Jat)2] P(N, t) = 2(nDt)1/2 exp 4Dt
47 Attached Polymer in Transients 0.2 2 3 4 2a't 6 5 7 8 Fig. 2-9. poet) (empty site) and PI (t) starting with all empty sites at t = 0, in the special case a = a e (x = 1), based on Eq. (6.20). (6.22) where ( ) - cp s = 1 (2n)I/2 fro e -y2/2 s d (6.23) y. When J~ < 0 (i.e., x < 1), Eq. (6.22) has a stationary equilibrium solution at t ...... co: (6.24) This can easily be seen to be consistent with PN = x N (1 - x) [Eq. (5.12)] if x is near unity: when J~ < 0, the continuous-N approximation is a good one at large t only if N( co) = x/(l - x) is large. Equation (6.22) is certainly easier to use than Eq. (6.15) but it is still moderately complicated. However, it becomes quite simple in the special case J~ = 0 (i.e., x = 1, a = ae , exa e = ex', D = ex'): P(N,t) = 2(nex~t)I/2 {exp [ (N ~,~o)2J + ex p [ - When No = 0 (empty sites at t = 0), and J~ = (N :a,~o)2J}. (6.25) 0, peN, t) = (nex,lt)I/2 exp ( - ~~t). (6.26)
Attached Single-Stranded Polymer 48 ° ° This is the right half of a Gaussian function, with peak at N = at all t, multiplied by two (for normalization). The <5 function at t = spreads over N > as t increases. When No = but rxa and rx' are arbitrary, Eq. (6.22) becomes ° ° _ 1 peN, t) - (nDt)1/2 exp [(N 4Dt - Jat)2] _ JaD e J.N/D <p (NJ2Dt + Jat) . (6.27) The polymers all start from empty nucleation sites and grow in accordance with this equation. If the initial distribution is not a <5 function at N = No but some broader normalized distribution over No, PNo' the results in Eqs. (6.15) and (6.22) have to be averaged over PNo • This simple procedure is appropriate because each polymer in the ensemble behaves independently (except for sharing the same pool of free subunits). Thus, if the left-hand side of Eq. (6.15) is now denoted by PN.N,(t), then (6.28) As an example, PNo might be the equilibrium distribution, Eq. (5.12), at one subunit concentration (and activity) and then the concentration is switched at t = so that the new activity is a. The new activity then prevails in the time evolution of the system. ° PN(t) with Absorption at N = 0 There are two circumstances, in the present context, for which absorption at is the appropriate boundary condition. In the first, we start (t = 0) with an ensemble of polymers at N = No and ask about the time it takes for some or all of these polymers to first dissolve completely (i.e., reach N = 0, an empty site). Of course, a given polymer that eventually dissolves both gains and loses subunits on the way: this is a stochastic process. Hence this problem is not directly related to the rate of label loss, if the original polymers are labeled and the free subunits are unlabeled (see Section 9). Furthermore, once a polymer first dissolves (reaches N = 0), subunits can be added at the same site and start a new polymer. However, in this problem, we have lost interest in and no longer count such a polymer: we keep track only of first disappearances of the original polymers that started at N = No. It is because of this that absorption at N = is the relevant boundary condition. The probability distribution in the time at which first disappearance occurs is the so-called first passage time distribution to N = 0, which we denote by g(t).1f Ja > (growing, on the average), the integral of get) over all t is less than unity because some polymers never reach N = 0. However, if Ja :::::; 0, this integral is unity. A second circumstance in which absorption at N = 0 is the appropriate boundary condition is the following. Suppose that, at t = 0, we start with an ensemble of polymers at N = No, and, in addition, we introduce a ligand that N = ° ° °
Attached Polymer in Transients - IXa N= 0 --:IX 49 IXa 2 3 '" ------:-- ~ IX IX Fig. 2-10. Kinetic diagram with absorption at N = O. binds rapidly and strongly on any nucleating site that becomes empty (N = 0) and "caps" the site. That is, the ligand effectively removes empty nucleation sites from the ensemble whenever they first appear. The boundary condition, absorption at N = 0, is especially obvious in this case. If we start with a b function at N = No, and have absorption at N = (Fig. 2-10), the probability distribution in the size of the surviving polymers at an arbitrary time is found to be (Ref. 5, p. 16) ° PN(t) x(N-No)/2e-("a+"·)'[IN_NJ2t~) - I N+NJ2tJaaa')]' = (6.29) where x = aa/a' as before. The sum of PN(t) over all N starts at unity and generally decreases with time because of the absorption at N = 0. The rate of absorption at N = is the first passage time probability distribution: ° g(t) = d:ro = a'P1 = Noclx-No/2e-("a+<X')'INJ2t~), (6.30) where we have used the recursion relation 1'-1 (z) - 1.+1 (z) = When J" < 0, the mean first passage time from No to from Fig. 2-10 to be (Ref. 7, p. 205) t = too tg(t)dt = (6.31) (2v/z)I.(z). No / - ° is easily found directly (6.32) J". This is what one would expect intuitively. We consider now the continuous-N approximation. The solution of Eq. (6.10) with a b function at N = No at t = 0, and with absorption at N = 0, is (Ref. 5, p. 52 and Ref. 6, p. 221) P(N, t) = 1 { [(N-No -J"t)2] 2(nDt)1/2 exp 4Dt _ ex p [ - J"No _ (N D + No 4Dt J"t)2]}. (6.33) ° This is the probability distribution in the size of the surviving polymers, that is, those that have never reached N = 0. In the special case J" = (i.e., a = a e , aa e = a', D = a'): P(N, t) = 1 { [(N - No)2] [(N + No?]} 2(na't)1/2 exp 4a't - exp 4a't . (6.34)
50 Attached Single-Stranded Polymer The integral of peN, t) in Eq. (6.33) over all N is the fraction of polymers still surviving at t: p*(t) == f" peN, t)dN -JNID (No-Jat) =rp ( -No-Jat) _eocorp fiDt (6.35) fiDt ' ° where rp is defined in Eq. (6.23). If Ja ~ 0, p*(t) --+ as t --+ 00. That is, all polymers eventually reach N = 0. However, if Ja > (the polymers are growing, on the average), p*( (0) = 1 - e-JocNo/D. (6.36) ° The fraction of polymers that reach N = 0, through stochastic fluctuations, is e-JocNo/D. The mean number of subunits in surviving polymers, per original polymer, IS N(t) = = I'" N peN, t)dN (Dt) 1/2 e-(No+Joct)2/4Dt n + (No + Jat)rp (- No - Jat) fiDt _e-JocNo/D[ (~tyI2 e-(No-Joct)2/4Dt - (No - Jat)rp(Nfi-it) ° J (6.37) At t = 0, N = No. If Ja < 0, N(00) = because the polymers shorten (on the average) until none survive. If Ja > 0, N increases steadily at large times (the surviving polymers grow) in accordance with dN = (1 - e-JocNo/D)J . dt a - ° (6.38) The situation when Ja = is anomalous because polymers disappear (reach N = 0) very slowly at large times while other polymers grow. In fact, insofar as the total number of subunits in polymers is concerned, the growth of survivors just offsets the shrinkage of other survivors and the absorption at N = 0, at all values of t: it is easy to show from Eqs. (6.34) and (6.37) that N(t) = No at any t. Thus, at large times, fewer and fewer survivors get longer and longer on the average, keeping N = No. An example will be discussed in Section 9. Here the first passage time (No to 0) probability distribution get) is clearly just - dp* /dt (the rate of polymer absorption at N = 0). A straightforward differentiation of Eg. (6.35) gives (see also Ref. 6, p. 221) t)2] . No [(No + Ja get) = 2(nDt 3 )1/2 exp 4Dt (6.39)
51 Attached Polymer under a Force When Ja = 0, g(t) approaches zero very slowly as t --+ 00: g(t) ex C 3/ 2 • This approach to zero is much faster for either positive or negative Ja . The integral of g(t) over all t is unity for Ja ~ 0 and is e- JaN./D for Ja > o. Using Laplace transforms (Ref. 6, pp. 211-212 and 221-222), the mean and variance in the first passage time distribution, when Ja < 0, are found to be [compare Eq. (6.32)] (6.40) The comment at the end of the preceding subsection applies to this subsection as well: if the initial distribution in polymer size is not a (j function, one has to average the above results over the initial distribution, as in Eq. (6.28). 7. Attached Polymer under a Force In this section, we consider a very long polymer under a force,s at equilibrium or steady state (i.e., steady growth or shortening). There are several possible arrangements of polymer and force. We begin with the most fundamental case, shown in Fig. 2-11 [also Fig. 1-2(a)]. For generality, we assume that subunit exchange (monomers only) can occur at both ends, with the first-order rate constants shown. The free subunit activity is a. The relevant equilibrium statistical thermodynamic expressions are Eqs. (3.15)-(3.19). As in Section 5, we introduce rate constants for the (long) polymer at equilibrium (a = ae ) and then use these rate constants at steady state as well (a # ae , J # 0). The principal question here is: how do the rate constants depend on F? As in Eq. (4.13), when F = 0 in Fig. 2-11, we have at equilibrium (7.1) From Eq. (4.12), when F is arbitrary (but the last term is negligible), kTlna e = (7.2) /10 - /10 - [oF. " OIQ Fig. 2-11. First-order rate constants for monomer exchange at both ends when a polymer is under a force F (positive for an extending force). 01 F
Attached Single-Stranded Polymer 52 The polymer at F = 0 is in equilibrium with free subunits at activity a~; at F, the polymer is in equilibrium with free subunits at a e • The equilibrium detailed balance relations at the two ends when F = 0 are (7.3) These are the same as Eqs. (5.33) but subscripts and superscripts have been introduced here to refer explicitly to the F = 0 case. These rate constants relate to the situation in Fig. 2-11 with F = 0, not to a polymer with free ends. Both a~/ao and [3~/[30 are equal to a~. The equilibrium constant for subunit addition is Ko = l/a~ [Eq. (5.32)]. The same would be true of the corresponding rate constant ratios for the two free ends (because the thermodynamic properties of a very long polymer do not depend on the status of the two ends), but the individual rate constants would be expected to be different. When the polymer is under a force F, and at equilibrium with free subunits at a e , the analogous detailed balanced relations are written [3a e aa e = a', = [3'. (7.4) The rate constants a, a', [3, and [3' are all functions of F, in general. When F ---+ 0, a ---+ ao , etc. The equilibrium constant for subunit addition is K(F) = l/a e [Eq. (5.32)]. On combining Eqs. (7.1)-(7.4), we obtain ~ = e-loF/kT = a~ a' ao aa~ (7.5) On rearrangement, (7.6) [3 [3' - = [30 1 F/kT . -eo [3~ (7.7) These equations show how the ratios a/a' and [3/[3' depend on F. Typically, 101FI/kT is of order unity (see Section 4). These relations were derived by consideration of an equilibrium polymer but they also apply to a steady-state polymer (a =I- a e ) because all of the rate constants in Eqs. (7.6) and (7.7) are independent of a or c (see Section 5). The rate constant ratios a/a' and [3/[3' in Eqs. (7.6) and (7.7) increase with F. Thus, aggregation is encouraged by extension (F > 0) and discouraged by compression (F < 0). This is a thermodynamic statement, already implicit in Eq. (4.14). The separate rate constants a and a' (the following remarks also apply to [3 and [3', of course) also depend on F but this is a kinetic, not a thermodynamic, matter. The effect of eloF/kT on a/a' is necessarily divided between a and a'. Whatever the division, it can be expressed conveniently in terms of a dimensionless parameter f~, as follows:
Attached Polymer under a Force 53 Fig. 2-12. Physical significance of /.; free energy barrier that determines the on-ofT rate constants in the case of compression. On (7.8) The corresponding equations for [3 and [3' are (7.9) Typically, fa and ffJ would fall between 0 and 1. In general, they would be functions of F (see below). Figure 2-12 illustrates the physical significance of fa in the case of attachment under compression (F < 0). The full curve in Fig. 2-12 shows the hypothetical free energy of interaction of a monomer with the end of the polymer and with the surface to which the polymer is attached, when F = 0, as a function of the distance of the monomer from its attachment site on the polymer. When the polymer is under compression, the attached state is less stable than when F = 0 [Eq. (4.11)], and the attachment free energy well is raised an amount -loF to curve C. At the same time, the rate constants change from ex o and ex~ to ex and ex'. The new (F) free energy curve from "off" to C might have various levels at the position of the transition state (maximum). If fa = 0, the transition state level is unaffected by F, and ex = ex o ' In this case, the full effect of F is in ex', which is increased over ex~ by a factor e- V /kT (because the free energy barrier to escape, on --+ off, is reduced). If fa = 1, the transition state level is increased by the full -loF. In this case, there is no effect of F on ex~(ex' = ex~) but ex is smaller than ex o by a factor eloF/kT (because the barrier to attachment is higher). If fa = 1/2, ex is smaller than ex o and ex' is larger than ex~, but both by less severe factors. When the polymer is under significant compression, it seems intuitively reasonable that it will be more difficult for an incoming monomer to squeeze between the polymer end and the attachment surface and also that the compression will tend to push end monomers out of the polymer. That is, ex will decrease and ex' increase. Hence we might expect an intermediate fa, say fa ~ 1/2, when F is large and negative. On the other hand, extension (F > 0) should make it easier for monomers to attach, though ex at large F cannot exceed the diffusion-controlled limit. Thus, for large F, we might expect fa oc l/F, so thatfJ becomes constant in Eq. (7.8). Figure 2-13 shows, qualita-
Attached Single-Stranded Polymer 54 Fig.2-13. Illustration of how fa might depend on the force F. Compress o Extend tively, the above-described F dependence of fa. This should be considered as merely a plausible possibility. Turning now to steady polymer growth or shortening at a =1= ae , using the rate constants shown in Fig. 2-11, the net rates of addition of mon6mers to the two ends, at arbitrary F, are la = exa - ex', lp = /3a - /3', (7.10) where the F dependence of the rate constants is given in Egs. (7.8) and (7.9). For example, (7.11) The total rate of addition is 1 = la + lp. The dependence of these rates on force is illustrated for la in Fig. 2-14. The three solid lines are for loF/kT = + 1, 0, -1, taking fa = 1/2 in the + 1 and -1 cases. Both slopes (ex) and intErcepts ( - ex') change with force. The broken line represents the loF /kT = + 1 case if we take fa = 0 for this F. In this case, the on rate is unaffected by F: ex = ex o • la = 0 for each line at a = ae; ae itself depends on loF/kT according to Eg. f. = Y2 loF/kT = + 1 Fig. 2-14. Illustration of the effect of F on J. for an equilibrium polymer.
Attached Polymer under a Force 55 Fig.2-15. The In a versus 10F/kT plane for an equilibrium polymer. On the line a = a., polymer is in equilibrium with monomer. In a III II o Compressing Extending (4.14). This latter dependence is shown in Fig. 2-15, where the heavy line is the logarithm of the critical activity, In a., as a function of the extending or compressing force [oF /kT If we were to include Jp lines in Fig. 2-14 (using, say, fa = /P), they would cross the corresponding Ja lines on the abscissa [Eqs. (7.4)], but the slopes would generally be different (i.e., rto "# /30)' Free Energy Transduction We can rewrite Eqs. (7.10) as Ja = r,(rta trtf --1 ) (7.12) ' where [see Eqs. (7.1), (7.4), and (7.5)] (7.13) with Afl defined by [compare Eq. (4.1)] Afl == f l - flo = flO + kTlna - flo = kTln(a/a~). (7.14) The expressions in Eq. (7.13) are alternative ways of writing e X / kT , where (7.15) is the total thermodynamic force (the same at rt and /3 ends) driving monomers from the solution onto the polymer ends when the polymer is under a mechanical force F. When free monomers in solution are in equilibrium with monomers in the polymer, a = a e , X = 0, and Ja = Jp = O. When X > 0, that is, when a > a e for a given F, Ja and Jp are both positive. In this case the polymer grows at the total rate
Attached Single-Stranded Polymer 56 J = Ja + Jp = (IX' + [3')(e X / kT - 1). (7.16) ° Thus the polymer grows for all points in the plane of Fig. 2-15 above the heavy line a = a e • Conversely, the polymer shortens when a < a e and X < (points in Fig. 2-15 below the heavy line). The two separate contributions to X are the intrinsic (or pure) monomer or subunit aggregation thermodynamic force 11/l and the mechanical force term loP' 11/l is the thermodynamic force driving monomers onto the polymer in the absence of a mechanical force F. 11/l [Eq. (7.14)] is positive above the horizontal line a = a~ in Fig. 2-15 and negative below this line. For any choice of a and F in region I of the plane of Fig. 2-15, the polymer is under compression (F < 0) yet it grows (X > 0, J > 0) against the compressive force because the intrinsic subunit thermodynamic force 11/l is large enough (a > ae) to counteract the opposing effect of loF. An illustration would be the growth of a bundle of actin microfilaments (ignoring ATPase activity for the moment), or HbS aggregates, against a cell membrance that resists the growth, thus pushing the membrane out and distorting the cell shape. This is an example of free energy transduction: mechanical work is done against the force F at the expense of the intrinsic subunit aggregation thermodynamic force 11/l. In the aggregation (growth) process, some subunit free energy is converted into mechanical work. The efficiency of the conversion is IJ = -loF/ 11/l. If -loF is close to 11/l (a ~ a e), IJ -+ 1. The rate of free energy dissipation is d.S Tit = JX = J(I1 /l + IJ) > 0, (7.17) where J [Eq. (7.16)] and X are both positive. Note that this free energy transduction process does not involve any NTPase activity; rather, part of the subunit aggregation free energy is converted into mechanical work. The other case of primary interest in Fig. 2-15 corresponds to region II. Here the subunit concentration is low enough (a < a e ) so that the polymer shortens (X < 0, J < 0) despite the extending force F > that opposes the shortening. The intrinsic subunit force - 11/l favoring depolymerization exceeds loF. Hence some of the subunit free energy -11/l is used to do work against the resisting mechanical force F. The efficiency of free energy transfer is IJ = loF /( - 11/l). The rate of free energy dissipation is again given by Eq. (7.17), but in the present case both J and X are negative. An example would be the shortening of a group of micro tubules (ignoring the GTPase activity) that are attached to a chromosome which offers viscous resistance to its movement (caused by the shortening). ° Maintenance of Attachment under an Extending Force Maintenance of contact between the two polymer ends and the surfaces to which they are attached, in Fig. 2-11, is not a problem under a compressing force (F < 0). However, when F is an extending force (F > 0), the question
57 Attached Polymer under a Force naturally arises as to how the polymer can maintain its attachment to a surface when subunits are being exchanged at the polymer end in contact with that surface. If the polymer does not penetrate the surface, but has only superficial contact, this trick would appear to be impossible for a single-stranded polymer, or even a double-stranded polymer like actin. It would seem to be possible though still improbable for a multi-stranded polymer (microtubule, HbS) or a tight bundle of actin filaments. In this case, the strands could take turns either holding onto the surface or exchanging subunits (Ref. 8, pp. 57-59), so that some contact is always maintained. Much more likely is that a subunit-exchanging polymer end, under an extending force, is held by a sleeve into which the polymer end is inserted, as shown schematically for the IX end in Fig. 2-16(a). Free subunits, at activity a, can be exchanged through the empty part of the sleeve. There are attractive intermolecular interactions between the subunits of the polymer outer wall and the wall of the sleeve. These interactions create the intermolecular force that induces the polymer to penetrate further into the sleeve (Ref. 8, p. 59). This penetration is opposed by the extending external force F and by a net loss of subunits from the IX end (such loss is the case of interest). The bestknown example is the apparent penetration of micro tubules into the outer layer of the kinetochore, while subunit loss at this end shortens the microtubules and pulls the attached chromosome toward the pole in anaphase. 9 ,lo For pedagogical purposes only, we shall compare hypothetical subunit exchange at the {3 end in Fig. 2-16(a), as already introduced in this section, with subunit exchange from the sleeve at the IX end. We begin by recalling the thermodynamic aspects of the following process: one subunit leaves the {3 end and enters the surrounding free subunit pool, which is at activity a. The respective chemical potentials of subunits in the polymer and in the pool are :. "'~U71 (a) ~ ... (b) Fig.2-16. (a) Polymer with IX end inserted in sleeve. An external force F pulls the sleeve structure to the left; a corresponding force is set up in the polymer. (b) Significance of variable n and of some rate constants. See text for details.
58 Attached Single-Stranded Polymer (Section 4) p,p = P,o - loF, p, = p,0 + kTln a. (7.18) The free energy change AG for the process is then AG = p, - p,p = Ap, + [oF = X, (7.19) as in Eqs. (7.14) and (7.15). The case of interest (region II in Fig. 2-15) is a < a e , AG = X < 0, Ap, < 0, and F > O. The subunit on and off rate constants at the 13 end are given in Eqs. (7.9). Also, from Eq. (7.13), f3a _ 13' a = _ = ae e!!Il-IkT eloF/kT = e X / kT . (7.20) The efficiency offree energy transfer is 1] = loF I( - Ap,)(see above) and the rate of free energy dissipation is JpX, where both Jp and X are negative. Implicit in these properties just summarized for the 13 end is the assumption that the relationship between the polymer end and the adjacent attachment surface is maintained without change when the subunit departs (in the above process). That is, operationally there is a single kinetic process only: the polymer end and adjacent surface readjust instantaneously as the subunit leaves. There is a single off rate constant (13') and a single on rate constant (13). Because there is no change in the state of the 13 end itself in the above process, the actual change is the loss of one subunit from the bulk polymer (under the force F). We examine now the same process as above, but at the r:t end: that is, loss of one subunit from the polymer to the pool, leaving the state of the r:t end unchanged. The net free energy change is again AG: the bulk polymer undergoes the same process (loss of a subunit) whichever end is used. However, there is not a single kinetic process at the r:t end; two separate processes are involved,10 which we consider after introducing the interaction free energy. Let w be the interaction free energy (negative) between one subunit in the polymer wall and the wall of the sleeve. The integer n in Fig. 2-16(b) is used to locate the tip of the polymer within the sleeve. The maximum value of n is M. If the polymer passes all the way through the sleeve, n ::::; O. The total free energy of interaction between polymer and sleeve is (M - n + l)w, if 1 ::::; n ::::; M. When n ::::; 0, this interaction free energy is Mw. The interaction free energy tends to pull the polymer into the sleeve (i.e., reduce the value of n, until n = 1). The corresponding mechanical force is - wllo • When the tip of the polymer is at n and one subunit is lost, the tip (not the polymer) moves to n + 1. To restore the r:t end to its former state n, the sleeve structure must move to the right, against the external force F, a distance 10 (one subunit). The order of these two subprocesses may be reversed. The free energy changes in the two subprocesses are Ap, - wand [oF + w, respectively, with sum AG = Ap, + [oF as in Eq. (7.19). In the first subprocess, one interaction is lost ( - w). Note also that the polymer tip is not under the tension F, as is the bulk polymer in Fig. 2-16(a). In the second subprocess, one interaction
Attached Polymer under a Force 59 is gained (w) and an amount of work [oF is done against the external force. When n ~ 0, the two subprocess free energy changes above become tl.J1 and [oF, because w is no longer involved. Each subprocess has its own pair of rate constants. For the subunit on and off rate constants we use rioa and ri~S, respectively, where S == e w /kT < 1. Here we are assuming that the full effect of w is felt by the off rate constant: the subunit has to pull away from the attraction of the sleeve wall. The constants rio and ri~ would be those of the free ri end if access of subunits to the polymer tip through the open end of the sleeve is considered to be uninhibited. Otherwise, rio and ri~ would be reduced somewhat, but by the same factor. The relations to thermodynamics are [compare Eq. (7.13)J (7.21) (7.22) When n ~ 0, sand ware omitted from Eq. (7.22) because the departing subunit is not in contact with the sleeve structure. The sleeve structure is assumed to undergo brownian motion with a rate constant K for discrete steps in either direction (tl.n = ± 1) of length [0' However, when the polymer is penetrating the sleeve (1 ~ n ~ M), a move of the sleeve structure to the left (tl.n = + 1) reduces the number of interactions w by one. This presents a potential barrier - w to movement (Fig. 2-17). The corresponding rate constant is thus reduced to KS. This effect is missing if n ~ 0. Another modification is required because of the presence of the external force F. A step by the sleeve structure to the right (tl.n = -1) requires a free energy increase of [oF. This free energy effect is assumed (because of symmetry) to be split evenly between the two K rate constants, giving KqJ to the right and KSqJ -1 to the left, where qJ2 == e-loF/kT < 1. This qJ is not related to the qJ in Eq. (6.23). Steps to the left (tl.n = + 1), leading to less penetration of the sleeve by the polymer, are favored by this effect. There is still another perturbation of these rate constants to be expected. A significant w implies close contact between polymer and sleeve. As a consequence of molecular "roughness" in both polymer and sleeve, especially the detach Fig. 2-17. Schematic free energy curve showing the origin of parameters s = e w / kT and r = e- b / kT • tI Free Energy o
60 Attached Single-Stranded Polymer subunit periodicity of the polymer surface, one would expect a resistance (potential barrier) to movement ofthe polymer in the sleeve, in either direction, that increases linearly with the extent of penetration of the sleeve by the polymer. This is indicated schematically in Fig. 2-17, where the unit potential barrier is b. We define r == e- b / kT < 1. For example, in Fig. 2-17, for n = M 1 -+ n = M, the barrier is 2b and the corresponding rate constant factor is r2. For n -+ n + 1, the factor is r M- n +1. When n :::;; 0, the factor is rM. This effect tends to reduce penetration in steady shortening of the polymer (a < a e ), but it would have no effect at equilibrium (a = a e ) because the reduction factor is the same in both directions. The final rate constants for motion of the sleeve structure, when 1 :::;; n :::;; M, are then Kr(n + 1) (7.23) (n + 1 -+ n) (7.24) n K((n) = Ksr M- n +1 qJ-l (n-+n + 1) = Kr M- n +1qJ and, for n :::;; 0, K((n) = Kr M qJ-l (n -+ + 1) (7.25) + 1) = (n + 1 -+ n) (7.26) Kr(n KrMqJ where the subscripts I and r refer to left and right. Corresponding to Eq. (7.22), we have for the second subprocess K((n) = ~ = Kr(n + 1) qJ2 K((n) ~c-----,-,-- Kr(n + 1) = 1 ~ qJ2 = e(l,F+w)/kT (1 :::;; n:::;; M) (7.27) eloF/kT (n :::;; 0). (7.28) If we now combine the two subprocesses [see Eq. (7.22)], the analogues of Eq. (7.13) are Kln) = Kr(n + 1) lXoa IX~ K((n) ---'--'---- = Kr(n + 1) e(!!I'-w)/kT e(loF+w)/kT e!!l'/kT eV/kT = = e X / kT eX/kT (1 (n:::;; 0). ~ n ~ M) (7.29) (7.30) Equations (7.20), (7.29), and (7.30) all agree in their final outcome, as we should expect because this is a thermodynamic result that refers to the same thermodynamic process. The efficiency of free energy transduction is again 1] = loF /( -11f.1). It is clear, though, that kinetic behavior at the IX end is much more complicated than at the f3 end, a subject we shall return to below. Although our primary interest is in steady shortening at the rx end, we digress here to consider the probability distribution Pn among all possible n values, at equilibrium (i.e., when a = ae ). This is the distribution in position of the sleeve structure on the polymer end; it is not concerned with the length distribution of the (long) polymer itself. The distribution is found most easily
Attached Polymer under a Force 61 not from rate constants but by using the free energy Gn of the sleeve structure in the presence of the external force F and of the polymer. If we arbitrarily assign Gn = 0 at n = 1, then Gn + loF) = -(n - 1)(w = -(n - 1)loF (n;:: 1) (7.31) 1). (7.32) (n ~ Here F > 0, W < 0 and, in real cases, Iwi is much larger than loF. The minimum in Gn is at n = 1, where there is full contact between polymer and sleeve. Gn increases linearly with n on both sides of n = 1, but more steeply on the positive n side. The equilibrium probability Pn of any n value is simply proportional to e- Gn / kT • Hence n = 1 (full penetration) has the largest Pn, with the distribution tailing off (exponentially; see below) on either side ofn = 1 (overpenetration, n < 1; under-penetration, n > 1). From Eqs. (7.31) and (7.32), we find easily Pn = (scp-2r!(1 - cp2)(1 - Scp-2) 1- S (cp2)1-n(1 - cp2)(1 - Scp-2) 1- s (n ;:: 1) (7.33) (n (7.34) ~ 0). Clearly we require s < 1, cp2 < 1, and Scp-2 < 1. We are assuming in Eqs. (7.31) and (7.33) that convergence is reached before n = M. The probabilities of any n ;:: 1 or of any n ~ 0 are then 1 - cp2 Pn?>! = -1-- (7.35) cp2 _ 8 Pn';;;O =-1--' (7.36) -8 -8 These latter quantities are of interest because from them we can calculate the average off rate constant at equilibrium. When n ;:: 1, the off rate constant is IX~S, and when n ~ 0 the off rate constant is IX~. Thus is rz , IX= rz (1 - cp2)IX~S + (cp2 - S)IX~ '2 , - / F/kT =IXcp =IXe 1-s 0 0 • (7.37) 0 This is just what we should expect from the relations (7.38) The last equation is the overall on-off detailed balance relation at equilibrium. It is easy to confirm from detailed balance between nand n + 1 at equilibrium that the on, off rate constants IXoa e , IX~S (n ;:: 1) and IXoa e , IX~ (n ~ 0), and Eqs. (7.23)-(7.26), are consistent with Eqs. (7.33) and (7.34). In other words, Eqs. (7.33) and (7.34) could also have been derived either from the on-off rate constants or from the K l , K, pairs.
62 Attached Single-Stranded Polymer o KsrMlp-1 01 KrMcp + Q'~S + Q'oa .. KSrM-1lp-l + Q'~S 2 KrM-1cp .. +Q'oa 1··· KSr2cp-l 3··· M - 1 01 Kr2qJ + o;~S + (ioa .. Ksrqr 1 + Q'~ S M .. detach Fig. 2-18. Kinetic diagram and rate constants for position of polymer tip in (or through) the sleeve. Equilibrium (a = a.) is actually a limiting case. Our primary interest here is in a polymer that is steadily losing subunits (a < a e ) at the r:t. end but is still held onto by the sleeve. (Here we consider the f3 end to be inactive.) Thus the sleeve structure, despite the resisting force F, is pulled toward the f3 end. The rate constants already discussed permit us to write the kinetic diagram for the states n at an arbitrary a ~ a e , as shown in Fig. 2-18. In steady shortening of the polymer, there will be a steady-state probability distribution Pn among the states in this diagram. This distribution relates physically to the location of the tip of the polymer within or through the sleeve. In some cases, the full distribution falls within the interval 1 < n < M, but in others n < 1 has significant probability. Obviously, if PM has a non-negligible value, the polymer will have a finite chance in a given time interval of being pulled out of the sleeve (Fig. 2-18). Because the kinetic diagram is linear, there will be a "detailed balance" solution of the rate equations (dPn/dt = 0) even though the system is at a nonequilibrium steady state. The solution cannot be expressed in closed form (except at equilibrium, a = a e) but it is easy to find the Pn numerically in any given case by starting, say, at Pl (to be determined at the end of the calculation by normalization) and then calculating, successively, P2, P3' ... ' and also Po, P-l, ... in terms of Pl. The n ~ 0 part of this calculation can be done analytically because the rate constants are independent of n. For n > 1 (Fig. 2-18), (7.39) etc. Typical distributions confined to n ~ 1 (see Fig. 2-20 and Fig. 3 of Ref. 10) are approximately Gaussian in shape, with peaks located mid-range in the interval 1 < n < M. When the peak of the Pn distribution is well into the n ~ 1 region, it is possible to find a closed expression 1 0 for nmaX' the location of this peak. Let n* be the value of n that gives equal forward and backward rate constants between n* and n* + 1, so that Pn* = Pn*+1. Then, from Fig. 2-18, (7.40)
63 Attached Polymer under a Force If we solve this equation for n*, then nmax = n* + (1/2). Thus we find (7.41) In this equation In r is negative and, in realistic cases, In [ ] is positive (K is a relatively large rate constant; see Section 25). We can see from Eq. (7.41) that nmax is increased (less penetration of the sleeve by the polymer) if r is decreased ("rougher" surfaces), if K is decreased, if qJ is decreased (stronger force F), if s is increased (weaker interactions), if rx~ is increased, or if rxoa is decreased. There are cases where Eq. (7.41) locates a minimum rather than a maximum. However, these have no real interest because in such cases the Pn will not converge before n = M is reached. If the full distribution falls within 1 < n < M, the rate of loss of subunits from the rx end is (Jp = 0, J = J,,) (7.42) where n* is given by Eq. (7.41) [n* = nmax - (1/2)]. The speed of movement of the sleeve structure toward the f3 end is in any case - J1o. If the Pn distribution has a significant component in n ~ 0, where the net off rate is larger, rx~ - rxA -J has to be found by averaging as in Eq. (7.37), but using the steady-state Pn distribution. If a is increased toward ae, we approach the equilibrium situation in Eqs. (7.31)-(7.38). The efficiency offree energy transduction, as already mentioned, is 11 = 10F/ (-11/1). The flux J does not enter this expression because it cancels from numerator and denominator. The rate of free energy dissipation is (7.43) where J is given by Eq. (7.42), or is averaged, including n ~ 0, if necessary. Both J and X are negative except at equilibrium (a = a e ), where J = 0 and x=o. We consider next a numerical example based on parameters appropriate for a microtubule attached to a kinetochore (see the end of Section 25 for further details on the initial parameters). We take M = 65, K = 1800 s-l, and rx~ = 340s- 1 . Nicklas 11 has found that an imposed force F = 5 X 10- 6 dyn, per microtubule, suffices to stop chromosome movement (J = 0) toward the pole in anaphase. This F gives 10F/kT = 7.474 at 25°C, with 10 = 6.15 A (microtubule). For an equilibrium polymer, J = 0 implies equilibrium. From Eqs. (7.3) and (7.5), = 340e- 7 . 474 = 0.1930 S-I. That is, we choose rxoa = 0.1930 S-1 so that J ---+ 0 as F ---+ 5 X 10- 6 dyn (we are interested, primarily, in J as a function of F, holding 1X0a constant). From
Attached Single-Stranded Polymer 64 cpZ = e- 7 .474 = 5.677 x 10- 4 , at equilibrium, we are led to choose s = 4.5 X 10- 4 in order to achieve safe convergence of Pn (i.e., well before n = M = 65) at equilibrium [note scp-z in Eq. (7.33), and see Fig. 2-19(b)]. This value of s corresponds to the rather large value w = -4.57 kcal mol- 1 and a mechanical force (Ref. 8, p. 59) - w/l o = 5.16 x 10- 6 dyn tending to pull the microtubule into the kinetochore. An external force F slightly larger than 5 x 10- 6 dyn, namely, F = - w/lo ' leads in the calculations below to scp-z = 1, nonconvergence of the steady-state Pn by M = 65, and detachment of the microtubule from the kinetochore. The final parameter, r, is chosen as follows. The (mean) rate of movement of an uninhibited chromosome toward the pole 11 is 0.73 pm min-l, which corresponds to a loss of 19.8 subunits S-l. The frictional resisting force to this movement is smaller than loF/kT = 7.474 (above) by a factor 1 1.1Z of about 7000. So we take J = -19.8 S-l at lJ/kT = O. The value ofr required to give J = -19.8 S-l at F = 0 is found, by computation, to be r = 0.9338. With the above set of parameters, - J has been calculated as a function of loF/kT, holding CXoQ constant, as shown in Fig. 2-19(c). -J is the speed of movement of the chromosome toward the pole expressed as the number of subunits lost from the microtubule per second. Figure 2-19(c) is the so-called force-velocity curve for this equilibrium polymer. The efficiency offree energy conversion (recall that there is no NTPase activity) is simply proportional to the force: '1= loF/kT -11f.1/kT lJ/kT 7.474 . 0.20 20 0.15 15 "," 0.10 rl' loF/kT ~ 0 (7.44) =-- 10 0.934 0.05 -1 0 4 8 12 16 20 /J n (a) (b) 24 6 8 Fig. 2-19. Numerical example discussed at length in text. (a) Pn distribution when F = O. (b) Pn distribution when lJ jkT = 7.474 (equilibrium; J = 0). (c) Force-velocity curve for this example.
65 Attached Polymer under a Force Thus 1J = 0 at F = 0 and 1J = 1 at J = 0 (equilibrium) whereas in muscle contraction 1J = 0 at both F = 0 and J = 0 (isometric contraction, but with ATPase activity). In muscle contraction, 1J refers to the partial conversion of ATP free energy of hydrolysis into mechanical work. There is a Pn distribution for each F. The two extreme examples (loF /kT = 0 and 7.474) are shown in Figs. 2-19(a) and 2-19(b). Figure 2-19(a) has very small fluctuations around the full insertion position (n = 1). Figure 2-19(b) is the equilibrium distribution, which is shifted considerably to higher n (less penetration) as a consequence of the large external force imposed. The population fraction in n ;;:, 1 is 0.99988. The value of Po is 1.2 X 10- 4 • If we take loF w = - - = -Ins = 7.7063 kT kT ' - (7.45) which is not a large increase over loF/kT = 7.474 in Fig. 2-19(b), the Pn curve is flattened considerably and convergence is not achieved by n = M = 65 (PdP65 = 3.07). Thus the microtubule would soon pull out of the kinetochore sleeve at this force. As a second numerical example, we alter some of the parameters above. We take lY.oa = 0 (subunits do not add to the polymer end) and s = 19.8/340 (so that the off rate when the polymer tip is within the sleeve will be 19.8 S-l). This value of s corresponds to w = -1.69 kcal mol- 1 and to a mechanical force - w/lo = 1.90 x 10- 6 dyn tending to pull the microtubule into the kinetochore. We then choose r = 0.925 so that, when F = 0, the Pn distribution (Fig. 2-20) is in the lower range of n values but does not penetrate significantly into the region n ~ O. In fact, the Pn distribution for F = 0 and r = 0.925 in Fig. 2-20 has Pn~ 1 = 0.99974 and J = -19.9 S-l. The value of Po is 2.5 X 10- 4 . As determined by Eq. (7.41), the peak of this distribution moves to larger n if F is increased, holding other parameters constant. At loF /kT = 2.5 (Fig. 2-20), 0.12 0.10 0.08 " Fig. 2-20. Two Pn distributions for a modified (see text) numerical example.
66 Attached Single-Stranded Polymer or F = 1.67 X 10- 6 dyn, the distribution just begins to reach n = M = 65 (P65 = 7.3 x 10- 4 ). The value of J is -19.8 S-l. For F values larger than this, detachment from the kinetochore would occur in due course. Corresponding to Eq. (7.45), we have here -In s = 2.84. In this second example, equilibrium is not a possible state because we have chosen Q(oa = O. Note that the value of J is essentially constant at -19.8 S-l from F = 0 to loF/kT = 2.5. There are a new additional comments on this sleeve model at the end of Section 25. Polymer between Two Rigid Barriers 8 In this subsection we consider cases in which the free subunit concentration c is high enough to cause a polymer with free ends to grow until, in the course of its growth, the polymer encounters obstacles or barriers at both ends. In another case the polymer may be anchored at one end to begin with and grows until the other end reaches a barrier. Even if, as we assume in this subsection, monomer exchange is still possible at both ends after the polymer has made contact with both barriers, net growth of the polymer will quickly cease. The physical reason for this is that addition of further monomers to the polymer (between barriers) will induce a rising compressive force F (negative) within the polymer that, in turn, will increase the critical concentration for growth of the polymer. When the critical concentration reaches c, the polymer will stop growing. Because these polymers are rather incompressible, not many additional monomers will be required to raise the compressive force enough to turn off the growth. As usual, we assume that the polymer is rod-like and does not bend. In the case of actin, we would be dealing with a bundle of actin filaments, to provide rigidity. In Fig. 2-21 [see also Fig. 1-2(b)], there are rigid barriers a fixed distance L apart. Somewhat elastic barriers are treated in the next subsection. The solution contains free subunits at an activity a that is arbitrary except that it is larger than the critical activity a~ for growth of the free polymer. In Fig. 2-22, which presents an illustrative case, the dotted lines labeled Ja* and Jt represent the growth rates of the two ends of the free polymer (Section Fig. 2-21. Equilibrium polymer that has grown up against rigid barriers a distance L apart. The free subunit activity is a > a~.
Attached Polymer under a Force 67 A'" ·T .."' 1 1 J' " 1 . ... C 1 J . : .a Fig.2-22. Subunit flux changes when an equilibrium polymer, with subunit activity a, encounters barriers at the two ends. 9). The points A and B on these lines are the rates of addition of monomers to the two ends of the free polymer at the particular subunit activity a. When the growing polymer first reaches the barriers, the force on the polymer is still F = but the growth rates change to points C and D on the lines J; and J3, with rate constants, lXo ' IX~, Po, p~. These rate constants refer to polymer in contact with barriers but with F = [Eq. (7.3)]. The critical activity is still a~ because the barrier merely acts as an inhibitory cap. At this point (F = 0), let No be the number of monomers in the polymer of length L [as in Eqs. (4.17)- (4.21)]. Then 10 = L I No • Because a > a~, a few more monomers (v) will add to the polymer until the compression (F < 0, N = No + v, I = L I N) is sufficient for the lines J; and J3 in Fig. 2-22 to shift (arrows) to the lines Ja and Jp with rate constants IX, IX', p, p'. Here the rate constants refer to a force F (negative) just sufficient to increase the critical activity from a~ to a, or just sufficient to move points C and D in Fig. 2-22 to point E on the abscissa. Of course, in the aggregation process at activity a, just described, only point E on the lines Ja and J(J is actually realized. The lines themselves represent hypothetical growth rates at the same F but at variable monomer concentrations. We now examine the quantitative aspects of the above discussion. The chemical potential of free subunits is ° ° J1. = J1.0 + kTlna. (7.46) For the polymer just in contact with the barriers, at F = 0, and in equilibrium with subunits at a~, the chemical potential of monomers in the polymer is [Eq. (4.13)]
68 Attached Single-Stranded Polymer flo = flo + kTlna~. (7.47) Because a > a~ and hence fl > flo, monomers at a will add to the polymer until F becomes sufficiently negative to raise fl for monomers in the polymer up to the value of fl in Eq. (7.46). At this point, the polymer will be in equilibrium with subunits at a [Eq. (4.12)]: (7.48) This equation determines F as a function of a. The independent thermodynamic variables here are T, L, and a (or fl in place of a). At the beginning of this section, on the other hand, F was the independent variable. On combining Eqs. (7.47) and (7.48), the explicit expression for F(a) is a In- = y*v. [oF -- = kT a~ (7.49) We have included y*v here from Eq. (4.18). If y* is known, v as well as F can be calculated from a. However, the much more important relation is F(a); this does not depend on knowledge, or an estimate, of y*. Because a/a~ would usually be of order 10 or less, the three expressions in Eq. (7.49) are usually of order 2 or less. For example, if y* v = 1 and y* = 0.05, v = 20. In this connection, see the numerical examples following Eq. (4.21). The individual rate constants shown in Fig. 2-21 are functions of F, as given in Eqs. (7.8) and (7.9), but here F is determined by the activity a offree subunits [Eq. (7.49)]. We turn now to a consideration of the equilibrium fluctuations in v, the number of monomers added to the polymer in passing from F = 0 to the final F in Eq. (7.49). The quantity v in Eq. (7.49) is more appropriately designated v (see below). The independent variables for the polymer in Fig. 2-21 are fl, L, and T, with fl determined by the subunit pool, Eq. (7.46). This system is open with respect to the number of subunits N in the polymer; that is, N fluctuates. The appropriate partition function for these variables is the grand partition function [Eq. (3.20)] 3(fl, L, T) = L Q(N, L, T)eNl'lkT, (7.50) N where Q is the canonical partition function (given below). This system, which is essentially a one-dimensional solid, has normal (small) fluctuations in N. The size ofthe system is set by L.1t suffices to use macroscopic thermodynamic functions below (No is of order 104 ). The finite aspect of the polymers dealt with in this book is important only for completely open systems. We need first an explicit expression for Q(N, L, T) for a macroscopic polymer with independent variables N, L, T We shall obtain Q from the free energy A and the relation Q = e- AlkT [Eq. (1.2)]. We start with the polymer with variables F = 0, N = No, L, and T, and then increase N from No to an arbitrary N, holding Land T constant (as in Fig. 2-21). The initial value of A is No flo· The change in A can then be found by integrating dA = flpdN [Eq. (4.4)], but
Attached Polymer under a Force 69 first we need f.1 p as a function of N. From Eq. (4.8) and F = h(l - U df.1 = -ldF = _l(OF) dl = -hldl p ol T . On integrating from 10 = (7.51) LINo to 1 = LIN, we have (7.52) Next we integrate dA = f.1 pdN from No to N, with L constant: A - Nof.1o = IN f.1p(N)dN (7.53) No A kT Nf.1o kT = + W _ y*v 2 kT = -2-' hl;(N - No)2 2NkT (7.54) (7.55) v = N - No· In writing Eq. (7.54), we have put N ~ No in the denominator. W is the isothermal work required to insert v subunits at the polymer ends, keeping L constant. From Eq. (4.18) we see that F= y*vkT dW 10 d( vlo ) (7.56) ---= --- is the force produced when v subunits have been inserted. The grand partition function can now be written as 3(f.1,L, T) = L (7.57) e-NlLo/kT e- r *v'/2 e NIL/kT. N Omitting factors in No, the probability of observing a given value of v is (7.58) where C is a normalization constant. As a slight approximation, we treat v as a continuous variable and Pv as a Gaussian distribution with mean v (to be determined). From InPv olnPv = InPv + (OlonPv ) v f.1- f.10 kT Jv - v) + !(02In/v) v=v 2 - - = - - - y* v = 0 (for v = v) ov ov (v - V)2 + ... (7.59) v=v (7.60) (7.61)
Attached Single-Stranded Polymer 70 we find [Eqs. (7.46) and (7.47)] y*v = J.l - J.lo kT P v = In ~ = a~ _loF (7.62) kT * )1/2 -y*(v-v)2/2 = (~ 2n e . (7.63) The variance in v is u; = v 2 - v2 = 1jy* = N 2 - f,P = (7.64) u~. Equation (7.62) agrees with Eq. (7.49). Also, Eq. (7.64) can be confirmed with the aid of macroscopic thermodynamics. From Eqs. (4.7) and (4.15): L (aF) [2 (aF) ( aJ.lp) aN L, T = - N aN L. T = N az T NK (7.65) We now apply this result to the present system, starting with Eq. (3.23) and using Eq. (4.16): 2 UN = kT (aN) aJ.l y* ' L.T (7.66) where we have introduced J.l = J.l p (equilibrium), N ~ No, and I ~ 10' If y* = 0.05, U v = UN = 4.5. Note that UN is very small compared to N (of order 104 ) but U v is significant compared to v (of order 30). The above discussion recognizes the fact that v (or N) fluctuates at equilibrium. Of course, v also fluctuates when the system is not at equilibrium, that is, in a transient. An example of a transient is the further addition of subunits after the polymer just comes in contact with the second barrier (Fig. 2-21). Let Pv(t) be the probability that the polymer has v extra subunits at t. The probability distribution Pv changes with time and approaches Eq. (7.58) as t --> 00. Pv represents an average over an ensemble of identical systems, or an average for a single system if the same experiment is repeated a large number of times. Figure 2-23 (see also Fig. 2-21) shows the rate constant notation we use in following transitions between different values of v (i.e., we follow the gain and loss of individual subunits). The rate constants depend on v because the work of inserting or removing a subunit depends on v. The differential equation in Pv(t) is, from Fig. 2-23, ... v - J v + J ... v Fig. 2-23. Kinetic diagram for excess number of subunits (v brium polymer between rigid barriers. = N - No) in an equili-
71 Attached Polymer under a Force (7.67) There is an equation like this for each value of v. At equilibrium (which is the only case we consider here), each pair of terms in parentheses must vanish separately because of detailed balance. We now consider the rate constants a v and a:+ 1 for the process v ~ v + 1 at the a end of the polymer. If there were no work or force involved in the addition or subtraction of a subunit, these rate constants would be ()(o and ()(~, as in Eq. (7.3). We need to correct these rate constants for the work of insertion. The work L1 W necessary to add one subunit to a polymer already with v excess subunits is given by [Eq. (7.55)] L1W kT y*(v + 1)2 y*v 2 2 2 (2v + l)y* (7.68) 2 This is also the value of -loF/kT in Eq. (7.56) at v + (1/2). Thus (7.69) where x == e- y*/2 < 1. This x is unrelated to the x in Eq. (5.4) and other equivalent uses of x. Equation (7.69) is the discrete analogue of Eq. (7.6). Note that when we deal with insertions explicitly, y* becomes a key parameter. For the separate rate constants we write, as in Eqs. (7.8), (7.70) Similarly, at the f3 end, (7.71) These are the rate constants one would use in Eq. (7.67) in studying a particular transient. A possible complication here is that ia and i p may vary with F and F in turn depends on v [Eq. (7.56)]. Thus, in general,fa and ill depend on v. At equilibrium, from Eq. (7.67), for v ~ v + 1 at the a end, (7.72)
72 Attached Single-Stranded Polymer where we have made use of Eq. (7.3) in the last form. This same result would be found on considering detailed balance at the f3 end. As a check on the above kinetic argument, let us also derive Eq. (7.72) directly from the equilibrium grand partition function. Pv is given in Eq. (7.58). For Pv +1' we replace v by v + 1. Then we find (7.73) having used Eq. (7.62). As a final point, we note that if 1Y.v!1Y.~+1 in Eq. (7.69) (the equilibrium constant for subunit addition after v subunits have been added) is averaged over the equilibrium distribution Pv in Eq. (7.63), the result is the same as in Eq. (7.6): IY. =~ IY.~ (y*)1/2 2n f {y* } exp --[2v+ 1 +(V-V)2] dv 2 (7.74) Such simple averaging over the separate rate constants IY. v and carried out because of the likely dependence of f~ on v. 1Y.~+1 cannot be Polymer between Slightly Elastic Barriers8 The polymer may grow, at monomer activity a > a~, against an elastic barrier or barriers, instead of against rigid barriers. Examples of this may be the plasma membrane or a cortical array of actin. In this case, as the compressive force F (negative) is built up in the polymer, the barriers are pushed back in accordance with some macroscopic law offorce appropriate to the particular barrier material or materials. In Fig. 2-24, the barriers are a distance Lo apart (lo = Lo/No) when the polymer first contacts the barriers (F = 0). After growth F F Fig. 2-24. Encounter of a growing polymer with slightly elastic barriers that yield to the extent La ..... L.
Attached Polymer under a Force 73 has ceased, the barriers are a distance L apart (L > L o' I = LIN), and the force F in the polymer is again determined only by a, from Eq. (7.49). Because of the mechanical equilibrium, this same force F acts on the barriers. To keep the remaining discussion simple, we now assume that the response of the combined barriers (when pushed by the polymer) follows another Hooke's law relation, F = A(Lo - L), and that the length change, L - L o' is small compared to Lo (e.g., L - Lo is several hundred A whereas Lo is of order 10 5 A). The constant A here is not related to J1 (as it usually is). In the final equilibrium state, a, F, and L here have the same significance for the polymer as in the preceding subsection, so most of the discussion of the equilibrium state there is still valid. However, there are now two contributions to v == N - No, which we consider below [y*v in Eq. (7.49) needs modification]. Fluctuations in Nand L are also different here; these will also be treated below. The activity a determines the equilibrium F, and F then determines L from F = A(Lo - L): kT a L=L +-Ino Alo a~' (7.75) where the last term is much smaller than Lo. With L available, N can then be found from the polymer equation -F = h(Lo _ L) = kT In~. No N 10 a~ (7.76) If we put because No » v, and use Eq. (7.75) for L, we then deduce from Eq. (7.76), kT v = ( AI; 1) a + y* In a~' (7.77) From Eq. (7.75) we can see that the new contribution to v, here is equal to (L - LoVio' which is of order 30 to 40. This is of the same magnitude as the term in y*-l. The new term in v is obviously due to the extra space made available to subunits when the barriers are pushed back (Lo --+ L). Aside from the extra contribution to v, elastic barriers do not introduce any really new features. Land F refer to the final state with stretched barriers. We now examine fluctuations in Nand L in the above system. We proceed as we did in deriving Eq. (7.57). Because the barriers obey the Hooke's law relation F = A(Lo - L), the associated barrier free energy is A(L - Lo)2/2 relative to an arbitrary zero at L = Lo. To find the free energy A ofthe polymer at arbitrary N, L, we start at No, Lo (with No = Lo/lo) and first increase No to N, holding Lo constant. We then increase Lo to L, holding N constant. In the
74 Attached Single-Stranded Polymer first step, Eq. (7.52) is still applicable except that 10 then leads to = Lo/No. Equation (7.53) (7.78) We then integrate [Eq. (4.4)] dA = FdL = h(~ -/o)dL (N, Tconstant) from Lo to L to obtain (7.79) On combining Eqs. (7.78) and (7.79), A = NfJo Nh (LN +2 -/0 )2 == NfJo + W'. (7.80) This is a rather obvious result, intuitively. We now construct a grand partition function [analogous to Eq. (7.57)] for the system polymer plus barriers with summand factors e- A/kT , e-).(L-Lo)2/2kT , and eN/l/kT ; S(fJ,Lo, T) = L e-N/lo/kTe-W'/kTe-).(L-Lo)2/2kTeN/l/kT. (7.81) N.L There are fluctuations in Nand L; the macroscopic size of the system is set by Lo. As usual, fJ = fJo fJ - fJo = + kTlna kTln(a/a~) = (7.82) -/oF. F is the final equilibrium force shown in Fig. 2-24, determined by the value of a. The probability of the system having particular values of Nand L, for given fJ (determined by a), Lo and T, is proportional to the summand in Eq. (7.81). We denote this summand by O. We shall treat 0 as a continuous Gaussian distribution in two variables. If we expand In 0 about N, L, where Nand L are the values of Nand L that satisfy 81nO/8N = 0, 81nO/8L = 0, (7.83) then we have, to quadratic terms, InQ(N,L) = 2 2 - + (8 - - + -1 (8 In Q(N,L) -1nO) 8 2 (N - Nf -1nO) 8 8 2 N N.r N L N,r - L) x (N - N)(L 2 + -1 (8- -1nQ) 2(L - L)2 2 8L -N,L + .... (7.84)
75 Attached Polymer under a Force Using InQ from Eq. (7.81), Eqs. (7.83) give (7.85) and A(Lo - L) = h(~ - 10). (7.86) Equation (7.85) is the same as Eq. (7.52) and Eq. (7.86) is simply the condition for mechanical equilibrium. That is, these are thermodynamic results already encountered. Equation (7.85) provides L/N as a function of /1. Equation (7.86) gives L as a function of L/N and hence as a function of /1. Finally, Land L/N determine N(/1). These expressions are: (7.87) (7.88) N - No = _ No (/1 - /10) ( h ) v= hi; 1 + ANo . (7.89) In all of these equations, /1 - /10 is given by Eq. (7.82). To simplify notation, we define w~, W~L' and wi by (a~~2Q) N,£ = - ~~, (~~~~) N.r = W~L' (a:~2Q) N,r = - ~i' (7.90) Then from InQ(N,L) [Eq. (7.81)], we deduce kTN 3 hL2 w~=-_-, w2 L kTN 2 W~L=---' hL (7.91) kTN ---= - h + AN Equation (7.84) can now be rewritten in Gaussian form as _ - [(N Q(N,L) - Q(N,L)exp N)2 + (N 2 2WN - N)(L - L) - (L - 2L)2] • 2 W NL (7.92) 2WL By completing the square in the quadratic form, the cross term can be eliminated. The necessary integrals are then easy to carry out. We find for the desired variances 2 WN - 2 (N - N) 2 = - = (v - v) l-1{1 (7.93)
76 Attached Single-Stranded Polymer (7.94) (7.95) where hjAN 1 + (hjAN)' Explicitly, (N - N)2 = kT!l3 (h + AN) hL2 AN ~~ y* (N - N)(L - L) ------,=-:-2 (L - L) = (1 + kTY*) AI; (7.96) = NkT AI = kT T' (v _ vy (7.97) (7.98) (7.99) These fluctuations are all small in magnitude. In Eq. (7.97), kTy* j AI; is of order unity [Eq. (7.77)]; the fluctuations in N and v are about twice as large as in the rigid barrier case [Eq. (7.64)], but they are still small. Finally, we note from Eqs. (7.82) and (7.89) that a loF - kT = In a~ = [1 y*v + (kTy*j.,U;)] ' (7.100) which differs from Eq. (7.62). However, both of these equations can be put in the form v= (N 2 - N 2)( -/oFjkT), (7.101) which is related to Eqs. (4.20) and (7.66): - V = No/(( -F) = No [/o(N2 - N2)] NokT (- F). (7.102) References 1. Chen, Y. and Hill. T.L. (1985) Proc. Nat!. Acad. Sci. USA 82, 1131. 2. Hill, T.L. (1986) Biophys. J. 49, 1017. 3. Hill, T.L. (1964) Thermodynamics of Small Systems. Part II (Benjamin, New York), Chapter 10. 4. Oosawa, F. and Asakura, S. (1975) Thermodynamics of the Polymerization of Protein (Academic, New York). 5. Goel, N.S. and Richter-Dyn, N. (1974) Stochastic Models in Biology (Academic, New York).
Attached Polymer under a Force 77 6. Cox, D.R. and Miller, H.D. (1965) The Theory of Stochastic Processes (Wiley, New York). 7. Karlin, S. (1969) A First Course in Stochastic Processes (Academic, New York). 8. Hill, T.L. and Kirschner, M.W. (1982) Int. Rev. Cytol. 78, 1. 9. Mitchison, T.J., Evans, L.M., Schulze, E.S., and Kirschner, M.W. (1986) Cell 45, 515. 10. Hill, T.L. (1985) Proc. Natl. Acad. Sci. USA 82, 4404. 11. Nicklas, R.B. (1983) J. Cell BioI. 97, 542. 12. Nicklas, R.B. (1965) J. Cell BioI. 25, 119.
3 Free Single-Stranded PolYlIler Chapter 2 dealt with single-stranded polymers attached to a surface. This chapter is a companion piece that is concerned with the same kind of polymer, but free in solution. The introduction to Chapter 2 applies here as well, and should be reviewed. In particular, it should be noted that the term "singlestranded" has a rather broad connotation. 8. Free Polymer at Equilibrium In this section we investigate the equilibrium length distribution in an ensemble of independent aggregating linear polymers, and related topics. The method used is statistical mechanics. Rate constants and kinetic behavior will not be considered until the next section. Explicit Model as an Example: Equilibrium Constants Primarily for pedagogical purposes, we introduce the subject of this section by a study of a simple explicit model of a linear aggregate. This model is rather unrealistic for the aggregation of protein molecules; it was designed originallyl for the stacking of bases, nudeosides, etc. (which, however, do not form the very long linear aggregates of primary interest in this book). Even this simple model must be treated approximately (essentially as an Einstein-like onedimensional crystal, as in Section 4). The advantage of the use of this model is that the reader can see typical but not very complicated expressions for partition functions, aggregation equilibrium constants, etc. These results will
79 Free Polymer at Equilibrium then serve as a foundation for our discussions of Y, PN , etc., in the following subsection. In this model, the monomers of mass m are considered to be rigid "checkers" (diameter d, height l) of uniform density that also stack like checkers. That is, an aggregate of N monomers has the shape of a right circular cylinder of diameter d, length Nl. The aggregation is caused primarily by strong attractive forces between adjacent monomers in a stack. We assume that each monomer adds to the stack in the same definite, preferred orientation. We write for the approximate partition functions QN of aggregates of sizes N = 1,2,3, ... , Ql = qP)q~l)qye-(W,+W;)/kT (8.1) QN = qlN)q~N)q~e-(W,+NW;)/kT q~-le-(N-l)W/kT. (8.2) Equation (8.1) is the partition function for a monomer; Eq. (8.2) applies to any N ~ 1. We now discuss the various factors appearing in these equations. The translational partition function for an N-mer in the solution of volume Vis qlN) = ( 2nNmkT)3/2 h2 V (N ~ 1). (8.3) Each monomer has a partition function qy for internal vibrational (and internal rotational) motion; further, qy is assumed to be unperturbed by stacking. The external rotational partition function q~N) of a cylinder of uniform density, diameter d, and length Nl, considered a rigid symmetrical top with symmetry number unity (in view of the complex molecular structure), IS (8.4) where IAN is the moment of inertia about the cylindrical axis and I BN ( = I cN ) is the other principal moment. These moments are easily found by integration to be md 2 N 8 I AN = - - IBN = mN(3d 2 + 4[2 N 2) 48 . (N ~ 1) (8.5) For a thin rod (dII-+ 0), IBN = mI2 N 3 112. The other factors in Eqs. (8.1) and (8.2) are less routine. The free energy of interaction, or potential of mean force, between the monomer and the solvent is represented roughly by Ws (contribution of top and bottom of cylinder) + w; (contribution of side of cylinder). For an N-mer, this becomes Ws + Nw;. This contribution would include, of course, not only direct solute-solvent interac-
Free Single-Stranded Polymer 80 tions but also the perturbation of solvent structure and of solvent-solvent interactions owing to the presence of the N -mer in the solvent. When two monomers are brought together to form a dimer, the six translational and six external rotational degrees of freedom of the two monomers become, for the dimer, three translational, three external rotational, and six vibrational, rocking, or hindered rotational degrees of the two rigid monomers relative to each other (in the neighborhood of their equilibrium configuration). These latter six degrees of freedom have a partition function that we denote by qtr; the subscript tr refers to the translation-rotation origin (in the monomers) of these degrees offreedom, which are essentially vibrational and much more restricted in the dimer. Similarly, for a trimer we introduce the factor q~ in Eq. (8.2) and, in general, we use q;:-l for an N-mer. This is an Einstein-like approximation. Actually, the relative motions of the monomers about equilibrium configurations at the N - 1 boundaries in an N-mer are not independent of each other, as implied by the simple product q;:-l, but should be treated by a combined normal coordinate analysis. Finally, in Eq. (8.2), W is the interaction free energy, between the two neighboring monomer surfaces at each pair contact in the N-mer, with the monomers in their relative equilibrium configuration. Equation (8.2) can be rewritten in more compact form (for N ~ 1): (8.6) where (8.7) The physical significance of w is that it is the potential of mean force (or free energy of interaction) between two neighboring monomers in the polymer, in their most stable positions, relative to infinite separation in the solvent. This follows because when two monomer surfaces are brought together, not only is the new monomer-monomer interaction W produced but also, in the process, these two surfaces lose their previous contacts with solvent (w s ). The partition function q in Eq. (8.7) is that of a single subunit in bulk polymer, as already introduced just before Eq. (4.22). The last factor in Eq. (8.6) is an end effect, the same' for any N. When N is very large (bulk polymer), In QN ~ N In(qe- w / kT ). (8.8) That is, the leading two factors (three degrees of freedom each) and the last factor in Eq. (8.6) are negligible for bulk thermodynamic purposes. Equation (8.8) is a special case of Eq. (4.22). In this special case, Eqs. (4.3) and (4.26) become qtr e- w / kT (qpl /V)q~l) , where K is the equilibrium constant for the process (8.9)
81 Free Polymer at Equilibrium solute (in solution) P: solute (in polymer). (8.10) As can be seen from Eq. (8.9), this equilibrium constant is made smaller by the greater freedom of motion (translation and rotation) of the solute when free in solution [(ql1) /V)q~1) compared to qtr] but it is made larger by the attractive potential of mean force between neighboring solute molecules in the polymer (e- W / kT ). Returning now to arbitrary N values, the equilibrium constant for formation of an N-mer from monomers is [Eqs. (2.20) and (8.6)] K _ QN/V _ (q(f) /V)q~N)(qtre-w/kTt-l N - (Ql/ V t [(qP)/V)q~l)]N (8.11) On using Eqs. (8.3), (8.4), and (8.9), this simplifies to KN = + 4N2]KN-l 3(d/lf + 4 N 3[3(d/l)2 (N ~ 2). (8.12) KN is important because it determines the activity aN of an N -mer [Eqs. (2.17) and (2.20)]: (8.13) For large N, KN is proportional to N 5 and to K N- 1 • If, as we almost always assume, polymers of significant size are very dilute and independent of each other [Eqs. (2.53) and (3.12)], QNAN N aN = CN = - - = KN a V (8.14) Y=IcN=IQNA N. (8.15) N N V The new feature here is that we have explicit expressions above for QN and KN (for this model). Equation (8.14) allows us to calculate CN' the concentration of polymers of size N, at a given activity a of free subunits. In many cases, we can use a ~ c. The equilibrium constant for adding a single monomer to an (N - I)-mer to form an N-mer is [Eq. (2.21)] I KN = KN K N- 1 = (N)3 [3(d/l)2 + 4N 2] N _ 1 . [3(d/l)2 + 4(N _ 1)2] . K, (8.16) where K (for similar addition to the bulk polymer) is given by Eq. (8.9). Equation (8.16) is valid for N ~ 2 (recall that K2 = K 2J Table 3-1 gives values of K~/K for d/l = 1 and d/l = 3. K~ differs from the bulk K-by more than 5% for N up to 100. This has obvious implications for the on and off rate constants at the two polymer ends (Section 5) for free N -mers of order N = 10 2 or smaller, as will be discussed in Section 9. The last column in
Free Single-Stranded Polymer 82 Table 3-1. Equilibrium Constant for Monomer Addition to an 1)-mer (N - N d/l = 1 d/l = 3 (N~1Y 2 4 6 8 10 20 50 100 200 1000 21.714 4.072 2.466 1.943 1.691 1.292 1.106 1.052 1.025 1.005 11.097 3.424 2.327 1.894 1.669 1.290 1.106 1.052 1.025 1.005 32.000 4.214 2.488 1.950 1.694 1.292 1.106 1.052 1.025 1.005 K~/K K~/K Table 3-1 shows that [N/(N - 1)]5 is a good approximation to K;'/K except for N < 10. The correction factors in Eq. (8.16) that convert K into K;' arise from translation and rotation of the finite polymers of sizes Nand N - 1. The two columns in Table 3-1 labeled K;'/K show that the early steps in the aggregation process (N < 10) are considerably enhanced thermodynamically from these effects, compared to aggregation onto bulk polymer. The enhancement tapers off as N increases. This effect should be included in the corresponding rate constants (Section 9). The qualitative origin of the above effect (Table 3-1) is the following. A free subunit loses less of its translational and rotational entropy when it adds on to a short aggregate (N = 2, 3, etc.) than when it adds on to the end of a very long polymer. Hence it is less reluctant to do the former and the corresponding equilibrium constant (K;') is relatively large. This effect tends to promote aggregation in the early stages and leads to an equilibrium distribution that is relatively depleted in small polymers (see the next subsection). Linear polymers may fragment or anneal. Without regard to the rate constants for these processes, which are presumably small enough to ignore in many cases, we may study the size dependence ofthe equilibrium constant 2 using the above analysis. Of course, if fragment -anneal equilibrium is reached in a polymer solution, this process will have no influence on the polymer length distribution, etc., because such equilibrium properties are independent of mechanism. For the fragmentation reaction N-mer +=1 N1-mer + N2 -mer, (8.17) (8.18)
Free Polymer at Equilibrium 83 Fig. 3-1. Plots of K' (dissociation), K (association), and Kr (dissociation equilibrium constant) as functions of N2 in a numerical example with n = 5, N = 1000, and all functions normalized to unity at the midpoint N2 = 500, about which there is symmetry [n refers to the exponent in Eq. (8.19)]. 3.0 2.0 N N ~ j ~ 1000 1000 - N2 n~5 1.0 f----------=~ 0.5 100 200 300 400 500 600 N2 with QN' etc., given by Eq. (8.6). Equation (8.18) is obviously consistent with Eq. (8.13). For polymers of significant size (N > 10), K ~C* f - N)5 (NN ' _1_2 (8.19) where C* is a collection of factors that do not depend on polymer size: C* = (2n 2 )5/2 d1 2 (mkT/h 2 )3 3qtr e w/kT --c:-----;,-;;;--- (8.20) Equation (8.19) predicts that the polymer is most likely to fragment in the middle, as illustrated in Fig. 3-1. In this example, N = 1000. The fragmentation constant K f (normalized to unity at its maximum) is plotted against N 2 , one of the fragment sizes (N constant). The curve has a Gaussian shape, symmetrical around the maximum at N2 = 500. Equation (8.19) also predicts that a longer polymer is more likely to fragment. Figure 3-2 shows how K f varies with N ( = 2N1 ) for breakage in the middle (N1 = N 2 ), normalized to unity at N = 1000. The size dependence shown in Eq. (8.19) is also a translation-rotation effect. If translation and rotation are ignored (as is common), one would mistakenly conclude that K f is independent of polymer or fragment size. As an appendix to this subsection, we give the generalization of Eqs. (8.5)
84 Free Single-Stranded Polymer 500 N Fig. 3-2. Plots of K' (dissociation), K (association), and Kr (dissociation equilibrium constant) as functions of N in a numerical example with n = 5, Nt = N2 = N /2 (i.e., symmetrical dissociation and association), and all functions normalized to unity at N = 1000. for a model more appropriate to a microtubule. This is a cylindrical rod with a cylindrical hole in the middle. The outer diameter is d and the inner diameter is do. Then one finds (S.21) IBN = mN(3d 2 For a microtubule we would use d + 3d; + 4[2 N 2 ) 4S = 300 A, do = (S.22) . 140 A, [ = 6.15 A, and m is the mass of a tubulin dimer. Thus each dimer is represented here by a very thin ring with the full microtubule cross section. The term in N 2 in Eq. (S.22) is ten times larger than the sum of the terms in d 2 and d;; when N = 147 (about 11 dimers per strand, which is similar to the condition N > 10 used above).
85 Free Polymer at Equilibrium Polymer Size Distribution at Equilibrium 3 In this subsection, we are interested in free polymers of significant size (IV is of order 10 2 , 103, or more, and a is near a e ) that are dilute enough to be considered independent of each other. In this case each polymer molecule may be treated as a completely open system with partition function Y, as in Eqs. (3.12) and (8.15). The collection of polymer molecules in the solution of volume V is the ensemble. The probabilities and averages we calculate may be regarded as long time averages for a single polymer molecule or ensemble averages over the collection of polymer molecules (in the limit of a very large ensemble). Because very small values of N are unimportant in the equilibrium distributions here (IV large), we need not confine ourselves to an explicit model as in the previous subsection. It is clear that any long linear polymer, regardless of its structural details, will have q:N) as in Eq. (8.3), IAN proportional to N, and IBN proportional to N 3 (mass x length 2, with both mass and length proportional to N). These three sources provide a factor N 3/2 . N l /2 • N 3 = N 5 in QN. Depending on the structure of the polymer, it is possible that the vibrational motion of the subunits relative to each other [corresponding to q~ in Eq. (8.6)] might introduce another factor in QN with N to some small positive or negative power (not necessarily an integer). For generality, we shall use N n where n is probably between 4 and 6; n = 5 is reasonable as a first approximation. Also, in order not to suggest any particular model or class of models, we can replace qe- w / kT (bulk polymer factors) in Eq. (8.6) by e-/l o/kT . This follows from Q = e- A/kT = e-N/lo/kT (8.23) for an incompressible bulk polymer [see Eq. (4.23)]. There will also be endeffect factors that depend on the particular case. Instead of Eq. (8.6), then, we use (8.24) where C' is a collection of factors that are independent of N. This is more general than Eq. (8.6) but, on the other hand, C' is not specified explicitly. The partition function Y is then [Eqs. (3.12) and (8.15)] Y = C' L Nn(e-/lo/kTt AN. (8.25) N:;'l Recalling that [see Eq. (4.1)] A = e/l/kT , fl fle = flO we have = flO + kTlna + kTln ae = flo, (8.26)
86 Free Single-Stranded Polymer (8.27) x == a/a e • The sum converges if x < 1, that is, if a < ae • Bulk polymer is formed at a = ae . The choice oflower limit in the sums here is not important because terms with very small N are negligible in any case (a near a e ; N large). For example, one might take the point of view, already mentioned, that the sum over N for polymers begins with the size of the critical nucleus; smaller values of N belong to free subunits, which may include small clusters (dimers, trimers, etc.). If n is an integer, the sum in Eq. (8.27) can be evaluated. For example, if n = 5, Y = C'x(l + 26x + 66x 2 + 26x 3 + x 4 ) (1 - X)6 (8.28) If x is near 1, we put x = 1 in the numerator and 1 - x = In(l/x) in the denominator: Y ~( 5!C' 1)6. (8.29) Inx Actually, Eq. (8.29) is a very close approximation to Eq. (8.28) even for x as small as 0.2 (the two equations give Y/C' = 6.9046 and 6.9043, respectively, in this case). Equation (8.29) is also obtained by treating N as a continuous variable in Eq. (8.27) and integrating: f CT(n ro y" C' 0 N"xNdN + 1) ~ (ln~)"H ' (8.30) where r(n) is the gamma function. When n is an integer, r(n + 1) = n! Because Eq. (8.30) is very accurate, we shall adopt this expression for Y. The probability that the polymer has size N is then, from Eqs. (8.27) and (8.30), 1 ( In~ )n+l NnxN r(n + 1) (8.31 ) The peak in this distribution, at N = Nm , can be located by using 0 In PN/oN = o. One finds n Nm=-l" 1nx (8.32)
Free Polymer at Equilibrium 87 For example, if x = 0.95 and n = 5, Nm = 97.5. Note that here we can calculate PN but not eN [Eq. (8.14)] because C' has not been specified (it depends on the particular model). Equations (3.1 0) give the mean and variance of the distribution: aln Y - n +1 (8.33) N=x--=-- ax aN 2 (J'N = X ax n = ( In~ x +1 1 1)2' In- (8.34) n+1 x Incidentally, by convention (for a polymer), wt. ave. == N 2 IN, no. ave. == N, (8.35) so that wt. ave. 1 ----1=--. no. ave. n+1 (8.36) As usual for a completely open system, the fluctuations in N are large: (J'~/N2 is of order unity rather than the usual order liN. Equation (8.31) can be written in a useful reduced form: PN = NnxN = PNm N::,x Nm [~exp(l Nm _ ~)Jn (8.37) Nm This function is plotted in Fig. 3-3 for n = 4 and 6. Note that PN is not symmetrical; this property has been implied already by the difference between Nand Nm • Equation (8.37) shows that different PN curves for the same polymer but with different N m values, obtained by varying a near a e , should superimpose if plotted in the above reduced form. Of course, actual observation of the true equilibrium probability distribution PN is a problem: it may take a 1.00 0.75 ~E ",< 0.50 0.25 Fig. 3-3. Theoretical polymer length distribution; reduced probability plots of PNIPNm versus N /Nm for n = 4 and 6. 1.0 1.5 2.0 N/N m 2.5 3.0 3.5
88 Free Single-Stranded Polymer very long time to reach the final distribution; and some methods may introduce artifactual results. The conventional belief is that PN ex x N (i.e., n = 0). This is correct for an attached rod-like polymer (Section 5) but it is not correct for a free polymer of this type. The translational and external rotational motion of a free finite polymer cannot be ignored. First Effect of Polymer Concentration on Polymer Size Distribution Elsewhere in this book we assume that polymer molecules are very dilute and independent of each other. This subsection is an exception. Here we consider the very first effect (second vi rial coefficient) of polymer-polymer interactions on the polymer size distribution, PN • The derivation of the main result, Eq. (8.44), is quite general. It applies not only to an aggregation equilibrium (Ref. 4, p. 370) but also to an equilibrium between isomers 5 or to a multiple binding equilibrium on a solute 6 (e.g., a protein). However, we shall use notation appropriate to an aggregation equilibrium. The starting point (Ref. 6, p. 46) is the activity aN of a polymer of size N expressed as a power series in the concentrations CN' of polymers of all sizes: aN = CN ( 1+2 ~ BNN,C N , + ... ). (8.38) This is a general thermodynamic expansion for a multicomponent solute, whether there is an equilibrium among solute components or not. The BNN , are osmotic pressure second virial coefficients; the osmotic pressure of the polymer mixture (or multicomponent solute) is (8.39) We define P~ as the equilibrium probability distribution among polymers in the absence of polymer-polymer interactions (i.e., with all BNN , = 0). This is the distribution in Eq. (8.31), where these interactions were not taken into account In this case aN = CN in Eq, (8,38), so that (8.40) We now rewrite Eq. (8.38) in the following forms: P~L CN aN' = - - - - - -N'- ' - - - - - (8.41)
89 Free Polymer at Equilibrium (8.42) 1 + 2 I BNN,C N, + ... N' P~(l + 2c N,N' I BNN'P~P~, + ... ) p = 'P~[l + 2C p (B - ~BNN'P~) + where the total polymer concentration is cp = ·.,l (8.43) (8.44) IN CN' (8.45) and we have introduced P~ for cN/c p in the sums because we are working only to the linear term in cp in Eq. (8.44). B has the physical significance of the second virial coefficient of the polymer mixture viewed, thermodynamically, as a single component with an internal equilibrium among substates N. From this point of view, Eq. (8,39) becomes II -k T = cp = Cp + c~ I N,N' .BNN'P~P~, + ... + Bc~ + .... (8.46) Equation (8.44) is the main result in this subsection, It will be illustrated below, If we multiply Eq. (8.44) by N and sum over N, we obtain N = N° - 2c p where N° is the mean of the calculation, one finds (J~ = (J,// - 2c p I I N,N' P~ (N - N°)BNN'P~P~, + ... , (8.47) distribution. Similarly, after a slightly longer [(N - NO)2 - (J7/]BNN'P~P~, N,N' + .... (8.48) The above results, Eqs. (8.44), (8.47), and (8.48), are quite general: they apply to any multicomponent solute with an equilibrium among all solute components. We now examine the special case of aggregating long rod-like polymers with hard (space-filling) interactions. For two cylinders with diameter d and lengths Nil and N 2 l, Onsager 7 showed that nd[2 NiN2 BN,N2 -- ---=----=4 (8.49)
90 Free Single-Stranded Polymer if N 1 /, NzI » d. Using this BNl N 2 , we find on carrying out the sums: ndl Z No B=--4 2 PN = (8.50) p,z:[ 1 + 2Bcp( 1 - ;0) + ... J - ( 2Bc N = N° 1 - n + ~ + .. .) (J02 (1 _ 4Bc p N n+1 + ... ) (Jz N = (8.51 ) (8.52) ' (8.53) where P~, N°, and (J~2 are given by Eqs. (8.31), (8.33), and (8.34), respectively. Equation (8.51) shows that PN is increased (compared to P~) for N < N° and decreased for N > N°. Correspondingly, N is less than N° as a result of the hard interactions. Similarly, (J~ is less than (J~2, but (J~/Nz is the same as (J~2 / No 2 (to the linear term in c p ). Let us use microtubules 8 as a numerical example, even though these are not equilibrium polymers. If we take n = 5, d = 300 A, 1= 6.15 A, N° = 15,000 (9.2 11m), and cp = 3 X 1011 cm- 3 , we find Bcp/(n + 1) = 0.100 for use in Eqs. (8.52) and (8.53). Actin provides another example. 9 We take n = 5, d = 104.5 A, 1= 28 A, N° = 923 (2.58 11m), and cp = 9.40 X 10 12 cm- 3 • These parameters lead to Bcp/(n + 1) = 0.0858. In both examples, linear terms are not sufficient in Eqs. (8.51)-(8.53). 9. Kinetic Aspects for a Free Polymer In this section, we introduce on-off rate constants for both long and finite polymers, examine rate constants for breaking and annealing a linear polymer, and then turn to several transient problems, including nucleation and growth. On-Off Rate Constants for Long Polymers; Steady State We have already discussed on-off transitions at the two ends of a long attached polymer in connection with Eqs. (5.32)-(5.36). Because the polymer is very long, the thermodynamic aspects are the same for an attached and a free polymer (e.g., the critical activity is a e for both). Even a free polymer has two different ends, because the subunits comprising the polymer are not isotropic. Thus, if the on-off rate constants at the two ends of a free polymer are 0:, 0:' and [3, [3' (usually the more active end is designated 0:), we have again, as in Eqs. (5.32), (5.35), and (5.36),
Kinetic Aspects for a Free Polymer 91 Fig. 3-4. Steady-state subunit flux, as a function of free subunit activity a, for rJ. end, f3 end, and both ends combined. a (9.1) = aa - a', Jp = f3a - 13' (9.2) = Ja + Jp = (a + f3)a - (a' + 13'). (9.3) Ja J The three steady-state fluxes are illustrated in Fig. 3-4. This figure differs from Fig. 2-7 in that Jp is much larger (on-off transitions at the 13 end are no longer inhibited by an attachment). If dimers, trimers, etc., can exchange at both polymer ends, as well as monomers, all of the thermodynamic relations in Eqs. (5.20)-(5.31) apply here also. The kinetic expressions among these equations (for the rJ. end) are also valid for the 13 end (replace the letter a everywhere by 13). For example, Jp = 131 (a - ae ) + 2f32K2(a2 - a;) + 3f33K3(a3 - a;) + ... 13; f3s = K s (f3~)S = a~ 131 as = K (a'l)S s a1 = Ks K S = _1_. K(s) (9.4) (9.5) The total subunit flux is N-Dependence of On-Off Rate Constants for Finite Polymers In the previous section we derived Eq. (8.31) for the equilibrium probability distribution PN when N is of moderate size or larger (a is near ae). In this case, values of N < 10 are not significant (at equilibrium). For example, if we take n = 5 and x = alae = 0.95, then N = 117.0, Nrn = 97.5, and PNjP10 = 990.
92 Free Single-Stranded Polymer Under these conditions, we also have, for the important values of N [Eq. (8.16) and Table 3-1J, K~=K(N~IY' (9.7) This is the equilibrium constant for adding a monomer to an (N - I)-mer to form an N-mer. Equation (8.31) for PN was derived from strictly equilibrium considerations, without mention of rate constants. Our object here is to relate this result to a kinetic approach in order to learn something about the N-dependence of the on and off rate constants. We let aN and /3N be the second-order on rate constants (at the two ends) for a polymer of size N and we let a~ and /3'" be the first-order off rate constants for the same polymer. Thus, when the free subunit activity is a, (9.8) is a piece of the kinetic diagram for a end transitions. For the /3 end, replace the letter a by /3 in (9.8). Here we are making the usual assumption that only monomers exchange at the two ends. There is detailed balance in these elementary transitions at equilibrium: /3N-l aPN- 1 = (9.9) /3'"PN· We then have, from Eqs. (8.31), (9.7), and (9.9), the interrelations ~ = (~)n ~ = (~)n Ka = K~a P N - 1 a N - 1 N- 1 e aN-1a a~ /3N- 1a (9.10) ----p;;-. Thus (N)n a( -N -)n - a' N - 1 N- 1 aK ,-K - N a~ N - 1 = /3N-l /3'" = (N)n K N - 1 = /3 ( N /3' N - 1 )n . (9.11) The rate constants a, a', /3, /3' refer to a very long polymer [Eq. (9.1)]. Equations (9.11) show that the ratios of inverse on and off rate constants for finite polymers are N dependent. The last column of Table 3-1, for N ~ 10, illustrates the order of magnitude of this effect. However, Eqs. (9.11) do not tell us how the factor [Nj(N - l)]n is divided between the individual rate constants a N- 1 and a~ (and between /3N-l and /3',,). Some assumption about this must be made.
Kinetic Aspects for a Free Polymer 93 As a rather general example, we begin by assuming that the on rate constant is diffusion controlled. As is well known (see the next subsection), the secondorder rate constant for a diffusion-controlled bimolecular reaction is proportional to the sum of the two diffusion coefficients. Risem'an and Kirkwood 10 found that for a linear string of N touching spheres, DN = D1lnN --=--N ' (9.12) where D1 and DN are the diffusion coefficients of monomer and N-mer. Thus the rate constant for a monomer adding to an N-mer is proportional to (9.13) When N --+ 00, the sum approaches D1 (i.e., the N-mer is essentially immobile). D1 can be expressed by Stokes law, if desired. For a slightly different geometry, we would expect an additional numerical factor of order unity in Eq. (9.12), which would depend on the particular case. In our example we take, in Eqs. (9.11 ), Cl. N - 1 and simHarly for PN-l' = CI. [ 1 + In(N_-~I)J _ (9.14) N - I It then follows from Eqs. (9.11) that CI.~ = CI.'[ 1 + In:_-/)J(N;; 1 y, (9.15) and similarly for p'". The ratio Cl.N-dCl.~ is, of course, consistent with Eqs. (9.11). Table 3-2 gives some numerical values of Cl.N-lICl. and CI.~ICI.', with n = 5. Compared to very long polymers, finite polymers have an increased on rate constant and a decreased off rate constant. The ratio of these two quotients is K~/K, the last column (in this approximation) in Table 3-1. The rate constants in Eqs. (9.14) and (9.15), or suitable modifications for Table 3-2. Size Dependence of On and Off Rate Constants in an Example, with n = 5 N rt.N-drt. r:t~/a' 10 20 50 100 200 1000 1.244 1.155 1.079 1.046 1.027 1.007 0.735 0.894 0.976 0.995 1.001 1.002
94 Free Single-Stranded Polymer slightly different models, should be used in transients as well as at equilibrium. However, for many transients (e.g., homogeneous nucleation and early growth) small values of N must also be considered. In this case, a refinement such as in Eq. (8.16) would have to be introduced as well as a modification of Eq. (9.12). The Fragment-Anneal Rate Constants The equilibrium constant Kr for fragmentation of a sizable N-mer into an N1 -mer and an N2 -mer, where N1 + N2 = N, is given in Eq. (8.19), for a particular model. More generally we shall write this as K ~ r - C* N)n (NN ' _1_2 (9.16) where C* would be an appropriate modification of Eq. (8.20) that would depend on the model. We shall designate the association (annealing) rate constant as K and the dissociation (fragmentation) rate constant as K', so that Kr = K' /K. All three of these quantities depend on N1 and N 2. Our procedure will be to give a simple and approximate derivation of K and then to obtain K' from KK r, because Kr is already available in Eq. (9.16). We use the general kind of argument 11 • 12 in Ref. 2 but with a significant qualitative improvement suggested by O.G. Berg (private communication). In a simple bimolecular association in solution, A + B ..... AB, the association rate constant is 4n(D A + Dn) R, where D A and Dn are the two translational diffusion coefficients and R is the small critical center-to-center approach distance (the bimolecular complex is considered to be formed whenever the center-to-center distance ~ R). The present problem, in which two rods (N1 , N 2 ) associate, adds several new features. In the first place, a given polymer can form an association at either end (the two ends are different). Whichever end is used, though, the associating partner must use its opposite end. Because of this two-end feature, the factor 4n above becomes 8n. Further, we assume that, in order for association to occur, two conditions must be satisfied. The angle 8 between the two rods must be less than some small maximum value 8m (Fig. 3-5); and the ends of the two rods must be within some small distance .... ~ ...... ...... . ..... Fig. 3-5. Two polymers aligned at the maximum angle Om that permits annealing (schematic).
Kinetic Aspects for a Free Polymer 95 b of each other. Then (9.17) Here sin 2 (8m /2) is the small solid angle fraction that is actually reactive and D ~ 2(DNl + DN,) = InNI 2Dl ( ~ InN2) + Ii; . (9.18) DNI and DN2 are translational diffusion coefficients as in Eq. (9.12) and the factor of 2 in Eq. (9.18) corrects approximately for rotational diffusion. 13 Finally, (9.19) The dependence of K and K' on Nl and N2 is confined to D and the factor (NIN2/N)n. To illustrate the N 1 , N2 dependence of the rate constants, K and K' are plotted in Fig. 3-1 as functions of N 2 , with n = 5 and N = 1000 = constant. That is, breakage (or the inverse) occurs at different places in a polymer of constant length. All curves are symmetrical about N2 = 500 and are normalized to unity at N z = 500. Thus K f and K' (dissociation) have maxima for symmetrical fragmentation (Nl = N z ) whereas K (association) has a minimum at Nl = N z . Figure 3-2 shows K and K' as functions of N for n = 5 and breakage or annealing in the middle (Nl = N z = N /2). All curves are normalized to unity at N = 1000. The association rate constant K decreases for longer polymers whereas the fragmentation rate constant K' increases. Finally, we give a hypothetical numerical example of the calculation of K from Eq. (9.17). We take Nl = N z = 2000, Dl = 7 X 10- 7 cm z s-1, b = 10 A, and 8m = 4°. This leads to K = 3.26 X 10- 17 cm 3 S-1 or to K = 1.96 X 104 M- 1 S-I. This happens to be the same order of magnitude found experimentally for microtubule annealing. 14 Mean First Passage Time in an Example As an interesting but only somewhat realistic illustration of a transient problem related to the N-dependent on-off rate constants introduced above, we derive expressions for the mean time it takes for a free polymer of size No + 1 to disappear (i.e., reach N = 1), when a < a•. The system is considered open, not closed, and a is a constant, independent of time. The kinetic diagram we use is shown in Fig. 3-6. This includes on-off transitions at both ends of the polymer. The mathematical problem is to find the mean first passage time to N = 1, starting at an arbitrary No + 1. As will be seen from the rate constants in this figure, two simplifications are being made here (in order to render the algebra tractable): the diffusion coefficient correction in Egs. (9.13)-(9.15) is
96 Free Single-Stranded Polymer (O! N~l + {Jla (O! .....~--­ + {J)a 4··· 3 d)"(c/+{J') (1)"(0!'+{J') Fig. 3-6. Kinetic diagram used in an example of a calculation of the mean time to disappearance of a polymer when a < a e • K'" omitted [thus the full effect of the N-dependence of in Eqs. (9.11) is placed on the off rate constants, rt'" and P~] and the resulting expressions for rt'" and P~ are used all the way down to N = 2, not correcting for small N as in Eq. (8.16). Despite these simplifications, this calculation will give a qualitative idea of the possible influence of N -dependent on -off rate constants. Figure 3-7 shows the conventional rate constant notation, used in books on stochastic theory, for absorption at i = O. This notation will be employed in this subsection only, to avoid confusion with prior usage (chemical potential, absolute activity). The mean time to absorption at i = 0, starting at i = 1, is (Ref. 15, p. 205) 1 0)1 = ~1 Similarly, starting at i + -,1,1- + ,1,1,1,2 ~1~2 ~1~2~3 + .... (9.20) 2, 3, ... , the mean time to absorption at i = 0)3 = 0)2 1 ,1,2 ,1,2,1,3 ~2 ~2~3 ~2~3~4 + - + -- + 0)2 = 0)1 + -1 + -,1,3- + ~3 ~3~4 + ... ,1,3,1,4 = 0 is (9.21) + ... (9.22) ~3~4~5 etc. Figure 3-6 is obviously a special case of Fig. 3-7 (actually a different case for each value of n). We shall use the notation t No +1 to denote the mean first passage time from N = No + 1 to N = 1 in Fig. 3-6. When n = 0, application of Eqs. (9.20)-(9.22) leads to _ t N o +1 x= N o (1 + x + x 2 + ... ) rt' + P' (rt rtf + p)a + P' rta ~ 0 pa a P' ae - rt' (9.24) - -- -- -Aj i (9.23) = -------- III 112 A3 A2 3 2 113 4··· 114 Fig. 3-7. First-order rate constant notation used, conventionally, in calculation of the mean time to absorption at i = O.
97 Kinetic Aspects for a Free Polymer Then t - No+! - No (a' J = (a + /3')(1 + f3)a No -- x) - -J (9.25) + /3'). (9.26) - (a' Here J is the total subunit flux (negative) for a long polymer molecule, as in Eq. (9.3). Equation (9.25) is essentially the same result as given in Eq. (6.32) for an attached polymer. Equation (9.25) has an obvious physical interpretation and serves as a useful reference point. When n = 1, we find from Eqs. (9.20)-(9.22) [ No (1 - x) t N~+l = + ( 1 + ~ + ~ + ... + ~J] (a' u + /3')(1 - x) (9.27) 2 The mean first passage time here is larger than in Eq. (9.25), as should be expected, because the off rate constants in Fig. 3-6 (leading toward N = 1) are smaller, especially for small values of N. Clearly, in Fig. 3-6, as n increases in the sequence n = 0, 1,2, ... , we should expect t No+! to increase at each stage because of reduced off rate constants. When No is large, Eq. (9.27) simplifies to - t No+! No [ y + In No ] ~ -J 1 + N o (1 _ x) , (9.28) where y = 0.577216 is Euler's constant. When n = 2, the analogue of Eq. (9.27) is [No (1 - X)2 t No+! + 2(1 L - x) (k- 1 ) + (1 (a' + /3')(1 _ X)3 = + x) ~)k-2)] (9.29) ' where both sums are from k = 1 to k = No, as in Eq. (9.27). When No is large, 2 ) -+ ((2) = n 2 /6, where ((5) is the Riemann zeta function. Then, for large No, L(k- t C'::: No+l - No [1 -J + 2(y + In No) N o (1 - x) (1 + X)((2)] + N o(1 - X)2 . (9.30) The same procedure can be carried out for n = 3, 4, ... , but the algebra becomes lengthy. We shall give the result for n = 5, from which the generalization to any integral n will be obvious: tNo+! = (a' + /3Tl(l - xf6[No (1 - X)5 + 5(1 - x)4I(k- 1 ) + 10(1 - x)3(1 + x)I(k- 2) + 10(1 - X)2(1 + 4x + x 2)I(k- 3) + 5(1 - x)(1 + llx + llx 2 + x 3)I (k- 4) (9.31) For large No,
98 Free Single-Stranded Polymer - ~ No [1 + t No +1 = -J + + 10(1 5(y + In No) N o (1 - x) + 4x + x 2 )((3) No (1 - (1 + X)3 10(1 + x)((2) N o (1 - X)2 + 5(1 + llx + llx 2 + x 3 )((4) No (1 - X)4 + 26x + 66x 2 + 26x 3 + X 4 )((5)] No (1 - X)5 ' (9.32) where ((3) = 1.20206, ((4) = n4/90, and ((5) = 1.03693. The binomial coefficients should be noted. The symmetrical polynomials that appear here are the same as those encountered in the sums I kixk. For example, ~ k3 L. k~l k _ x - x(1 + 4x +4 x 2 ) (1 - x) ( 1) x<. (9.33) The sum of the coefficients in the numerator is 31. See also Eq. (8.28). As a numerical example, if we take No = 1000 and x = 0 (i.e., a = 0) in Eq. (9.32), the quantity in the brackets [ ] becomes 1 + 0.0723. The correction (relative to n = 0) is not large. However, for No = 1000 and x = 0.5, the corresponding quantity is 1 + 2.420. Obviously, the correction becomes very large as x --+ 1. Disappearance of Polymers at the Critical Concentration We consider here an illustration of Eq. (6.35) in the special case J = O. That is, we start (t = 0) with free polymers, all with size No, at a = ae . The total on rate constant is ((X + f3)a e and the total off rate constant is (x' + f3'; in fact, these total rate constants are equal (a = ae , J = 0). We neglect the size dependence of free-polymer rate constants here, as an approximation, in order to be able to use Eq. (6.35). Of course, the rate constants for attached polymers (Section 6) are not size dependent. In Eg. (6.35) as used here, D = (x' + [3' because both ends are involved. The fraction of free polymers still surviving at t, p*(t), is then p*(t) -No) = qJ ( - - .j2Dt qJ (No) -- .j2Dt (9.34) Incidentally, in Section 6, we had to invoke a hypothetical ligand that capped empty sites in order to keep a new polymer from appearing on a site vacated by an original polymer that has dissolved. This is not necessary here. Because homogeneous nucleation is relatively slow, we can assume that free polymers
99 Kinetic Aspects for a Free Polymer 1.0 10 20 2DI/N~ Fig. 3-8. Theoretical calculation of the fraction of surviving polymers, as a function of time, at a = a e • that disappear do not reappear on the time scale of the process being considered. As explained following Eq. (6.38), the concentration offree subunits remains constant (even in a closed system) because surviving polymers grow longer at a rate just sufficient to incorporate those subunits made available by shrinking and disappearing polymers. In fact, Eq. (6.34) (with D = (x' + {3' in place of (X') gives the size distribution of surviving polymers at t. This distribution can be visualized (for N ~ 0) as a Gaussian function that starts as a 15 function at N = No and spreads symmetrically, from which is subtracted a similar "image" Gaussian centered at N = - No' Clearly, P(O, t) = 0 at all t (physically, because of the "absorption" at N = 0). Figure 3-8 gives a plot of p* as a function of t, from Eq. (9.34). After a relatively rapid drop (actually there is a very short lag period starting at t = 0, not visible in the figure), surviving polymers disappear quite slowly. The disappearance of sonicated F -actin at the critical concentration behaves in this way,16 with parameters No = 22 and D = (X' + {3' = 7 S-I. In this case p* = 0.50 at t = 77 s (see also Section 22). Rate of Label Loss from Polymer 17 We consider an ensemble of completely labeled large free polymers of size No (at t = 0) in a large solution of unlabeled monomers at constant activity a. The number of labeled monomers lost from the polymers to the solution is negligible compared to the original number of unlabeled monomers. Hence monomers that add to the polymers are always unlabeled. We are interested in the rate of loss of label from polymers in two cases: a < a e , so that the polymers are shortening, and a = ae , so that the polymers maintain a steady mean size. The case a > ae is not interesting because the label in a polymer molecule is quickly "capped" at both ends by added unlabeled monomers.
Free Single-Stranded Polymer 100 o = unlabeled 0= labeled Fig. 3-9. rx. end of a labeled polymer that is losing monomers to a solution of unlabeled monomers. a end 0000000 .. ·.. · I I L l_ _ _ _ _ _ _ _ _ _ _ __ q N Much of the notation in this subsection is rather special and is not used elsewhere. We consider first a < ae : the equilibrium polymer shortens at both ends. Figure 3-9 shows the a end of a single-stranded polymer at an arbitrary time at which the polymer has N labeled subunits and q unlabeled subunits at the a end. The /3 end has its own variable analogous to q. At t = 0 there are No labeled subunits and no unlabeled subunits (q = 0). Some of these labeled subunits are lost from both ends as time passes; thus N :( No. No is a large number whereas q is zero or a positive integer, usually small. N can decrease to N - 1 by a loss at the a end only when q = 0; if q ~ 1, the a end is temporarily "capped" (Section 27 is somewhat related). This system can be treated in detail stochastically but here we give a condensed and somewhat intuitive discussion. The activity at the a end can be summarized as follows: aa is the rate of adding unlabeled monomers; rl is the total rate of losing monomers (of which aa are unlabeled); and a' - aa is the rate of losing labeled monomers. Labeled monomers never return to the polymer, once lost. In effect, then, the kinetics of label loss can be treated as a unidirectional random walk on the integers N with total rate constant (a' + /3') - (a + /3)a, or -J, for each step N -+ N - 1. The temporary cap of the a end by unlabeled monomers (excursions in q with q ~ 1) is responsible for the effective reduction of a' to a' - aa as the a end rate constant for N-+N-1. Let P(N, t) be the probability of a polymer having N labeled monomers at t. Because N is large, we treat it as a continuous variable. Then, as in Eq. (6.10), P(N, t) satisfies the differential equation oP 1 (j2p at ="2( -J) oN 2 oP + (-J) oN' (9.35) This equation has the Gaussian solution [see Eq. (6.11)] [N - N(t)]2} 20"~(t) , (9.36) where mean and variance change with time according to N(t) = No - (-J)t " O"~(t) = (-J)t = No - N(t). (9.37)
101 Kinetic Aspects for a Free Polymer aa aa au _-3··· a' q = 0 ..........-:--_ a' Fig. 3-10. Kinetic scheme for Fig. 3-9 in following fluctuations in q. In summary, N(t) is the mean number of labeled monomers still remaining at t, per polymer, No - N(t) is the mean number of labeled monomers lost by t, and - J is the mean rate of loss of labeled monomers. We return now to the variable q (at the rx end). The excursions in q to q ?! 1 are very limited in extent so q must be treated discretely. Figure 3-10 shows the rate constant scheme that governs these excursions. This applies to every value of N. Shifting of N with t does not perturb the distribution in q. Consequently, after a brief transient, an equilibrium is set up among the q values. Let Pq be the equilibrium probability of q, irrespective of N value. Then, as in Eq. (5.12), we find Pq = x Q(1 - x) (q?! 0) x = rxa/rx' < 1. (9.38) We then have x rxa rx' - rxa (9.39) rxarx' (rx' - rxa)2 . (9.40) q=--=--1- 2 (J Q = X X (1 - X)2 = .,....,------,-;;- The q unlabeled subunits are like an attached polymer (Section 5), the attachment being to the rx end of the long labeled polymer (Fig. 3-9). The probability that the rx end has no unlabeled subunits is Po = 1 - x. The probability that it has at least one unlabeled subunit is 1 - Po = x. Hence the rate of losing labeled subunits is rx' Po or rx' - rxa, and the rate of losing unlabeled subunits is rx'(1 - Po) or rxa. These two rates were mentioned above without a real justification. We turn now to the more complicated case a = ae . We consider the rx end only; the treatment of the f3 end would be completely analogous. At the rx end, rxa e = rx': the on and off rates are the same; J~ = O. Figure 3-11 shows a sampling of the states passed through by the rx end of the polymer, in a hypothetical sequence, as time passes. The mean position of the end of the polymer remains at s = 0, because on and off rate constants are the same, but labeled monomers are gradually lost as a result of fluctuations that temporarily remove all of the unlabeled monomers. These fluctuations become less frequent, because they must be larger, as more labeled monomers are lost. We are interested in the amount of label lost by time t and in the rate at which label is lost at t. This problem is a random walk on the integers s (Fig. 3-11), starting at s = 0
Free Single-Stranded Polymer 102 F -2 -I 3 4 0:0:0:0:0 .. · 0000000··· 0 o 0 0 0··· 0 00000··· 000··· 000000··· 6 • • • 1=0- Q' =Iabclcd = unlabeled end Fig. 3-11. Hypothetical sequence of polymer states at the IX end in the special case of the system in Fig. 3-9 when the mean gain or loss of monomers at this end is zero. The variable s locates the polymer end, whether labeled or not. (with completely labeled polymer), and with rate constant ex' for steps in either direction. We are interested in cases with a large number of steps. The mean time between steps is 1/2ex'. The number of steps by time t is 2ex't. This is the maximum possible number of labeled monomers lost by t. Chandrasekhar (Ref. 18, Eq. 24) practically gives the solution, but in different notation. The probability that the walk has never gone beyond (to the right) a particular positive s value, after time t, is P (t) S = 1 fS (4nex't)1/2_00 [e-m2/4~'t - e-(2S-m)2/4~'t] dm. (9.41) After a change of variable in the second integral, this can be simplified to Ps(t) = 1 (nex't)1/2 IS e- m2 / 0 4a "dm. (9.42) It then follows that the integrand, Rm(t) = e- m2 /4a 'I/(nex't)1/2, (9.43) is the probability (density) that m is the largest value of s that has been reached in a walk lasting a time t. Thus, Rm(t) is the probability that m labeled monomers have been lost to the solution between t = 0 and t. The mean number of labeled monomers lost by t is then (9.44) The rate of loss of label, after the transient, is
Kinetic Aspects for a Free Polymer ~7 = (:~y/2 103 (9.45) This approaches zero slowly as t ~ 00. As a numerical illustration of Eq. (9.44), suppose a' = lOs-I. Then the mean number of labeled monomers lost from the a end, per polymer molecule, is 3.6 at t = 1 s, 11.3 at t = 10 s, 35.7 at t = 100 s, and 112.8 at t = 1000 s (16.7 minutes). As a postscript, the reverse problem should be mentioned. Suppose that the designation of the two types of monomers in Fig. 3-11 is simply exchanged. We are interested, then, in the rate at which label from solution penetrates into an end of an initially unlabeled polymer when the end has zero mean growth. Clearly, the mathematical argument just given applies and Eq. (9.44) is again the result. However, min Eq. (9.44) now represents the mean penetration of label, in monomer units, by time t. Nucleation and Growth of Polymers The classical homogeneous nucleation problem is the condensation of a liquid from its vapor via small droplets ("critical nuclei") which present a free energy barrier that must be surmounted. 19 The barrier is a consequence of an excess surface free energy term (compared to bulk liquid) that is proportional to N2/3. At the molecular level, this term arises from the fact that molecules on the surface of a droplet have missing intermolecular interactions. Other homogeneous nucleation problems in three dimensions are similar. In two dimensions, there is an analogous "surface" excess free energy term proportional to N 1/2 and a corresponding free energy barrier (critical nuclei). The above are dimensional (or geometrical) effects that do not arise in one dimension where the excess free energy term is a constant and there is no associated free energy barrier. However, in particular one-dimensional systems (see the discussion of actin below), there may be a free energy barrier of sorts at very small N owing to the individual characteristics of the early aggregation steps. A feature of two- and three-dimensional homogeneous nucleation systems, arising from the "dimensional" free energy barrier referred to above, is the existence of a critical supersaturation ratio l9 (clc e > 1) below which the rate of the condensation process is completely negligible, even though condensation is favored thermodynamically. The homogeneous nucleation of HbS polymerization is in essence a threedimensional problem, with critical supersaturation playing a crucial role. 20 This is complicated by subsequent heterogeneous nucleation on polymers once formed and by very large activity coefficient corrections (because of the relatively high concentration ofthe protein). Furthermore, translational, rotational, and vibrational partition functions of the various polymer species have been included in the theoretical interpretation of experimental results. Although this system is outside of the scope of the present chapter (on
104 Free Single-Stranded Polymer single-stranded or effectively single-stranded polymers, with one-dimensional "nucleation"}, the reader would do well to study Ref. 20 as a model of completeness in this field. The homogeneous nucleation of micro tubules from tubulin 2 1,22 is a problem somewhat related to HbS. This is essentially a two-dimensional nucleation system; again critical supersaturation is involved. The two-dimensional critical nucleus is a small piece (seven subunits in two strands) of the microtubule wall. 22 There is then a second-stage nucleation, similar in principle to the heterogeneous nucleation mechanism used for HbS in Ref. 20, that leads to a widening of the initial patch of microtubule wall. This is followed by tube formation and then conventional longitudinal growth. Computer simulation 22 leads to rather good agreement between theory and experiment. Presumably, in vivo, microtubules are almost always initiated by heterogeneous rather than by homogeneous nucleation (e.g., from the centrosome). That is, the microtubule is attached (as in Chapter 2) rather than free. This chapter is concerned with effective single:stranded polymers that are free in solution. Insofar as the initiation and growth of such polymers is concerned, this implies a one-dimensional system, and hence excludes microtubules (two-dimensional) and HbS (three-dimensional). We shall use the growth of actin, below, as a concrete one-dimensional example. However, first we consider, essentially as an exercise, a simple hypothetical example of growth in an open system (experimental studies on nucleation and growth are carried out in a closed system where the total number of subunits, free or in polymers, is held constant). In the remainder of this subsection, we use concentrations in place of activities (as an approximation) because of the need to count subunits. As a further more serious approximation, we use on-off rate constants that are independent of polymer size. In this example, we start, at t = 0, with a monomer-dimer equilibrium, and no polymers. We assume that a fast monomer-dimer equilibrium is maintained throughout the transient. From Eq. (8.13), C2 = K2C 2 is the concentration of dimers for an assigned value of c. K z is very small. The monomer concentration c exceeds Ce; hence J> o. "Polymers" start with dimers. The rate (concentration per unit time) at which trimers are formed from dimers is (0( + fJ)cc 2 • Because the system is open, c and C 2 are constant concentrations, independent of time. Thus trimers are formed steadily, at the above rate. To follow these trimers as time passes, we shall use the continuous-N approximation, introduced in Section 6. In this approximation, we denote a trimer as N = 1 and a dimer as N = 0. Once a. trimer is formed, Eqs. (6.33)-(6.40) are applicable, with No = 1; these equations apply to absorption at N = 0. Absorption (rather than reflection) is the appropriate boundary condition when trimer -+ dimer because a constant supply (C2) of dimers is already provided by the fast monomer-dimer equilibrium. Thus, when a trimer becomes a dimer by loss of a subunit, this or some other dimer would become two monomers, in order to maintain the equilibrium concentration C z .
Kinetic Aspects for a Free Polymer 105 Consider the (a + f3)cc 2 dt' trimers formed from the pool of dimers between t' and t' + dt'. Equation (6.37) with No 1, = t t', -H - J = (a D = t[(a + f3)c + f3)c - (a' + 13') > + a' + 13'] 0 (9.46) gives the mean number of polymer subunits at t, N(t - t'), per polymer (trimer) initially formed at t'. Thus the total concentration of subunits in polymers at t, c;(t), is c;(t) = (a + f3)K 2 c3 I (9.47) N(t - t')dt'. After an initial transient, the integrand is [see Eq. (6.38)] N(t - t') ~ (1 - e-J/D)J' (t - t'). (9.48) The first factor on the right is the fraction of initial polymers (No = 1) never absorbed at N = 0 [see Eq. (6.36)]. Then Eq. (9.47) becomes, on integration, (9.49) Thus, approximately, in this open system, c; ex t 2 , and this quadratic time dependence continues indefinitely. If this same system were closed, before the initial c changes significantly one would again expect (after an early transient) c; ex t 2 • This would be temporary behavior only, however, because of the decrease of c with time (eventually, c ~ Ce in a closed system). Wegner and Engel 9 applied the above model, with a fast monomer-dimer equilibrium, to the growth of actin (treated as an equilibrium polymer-but see Section 22) in a closed system. Because c changes with time, it is more convenient to return to a discrete approach rather than to use the continuousN approximation. Because of the assumed fast equilibrium, we have c2 (t) = K 2 c(t)2 at any t. For polymers, starting with trimers, we have the master equations [compare Eqs. (6.1)] dC 3 dt = dC dt4 = etc., for CS, c6 , (a + f3)K 2 c 3 (a •••• + f3)CC3 - (a' - (a' + f3')C3 + (a' + f3')c 4 + {3')c 4 + (a' + {3')c s - - (a (a + {3)cc 3 + {3)cc 4 , (9.50) (9.51) The total concentration of subunits, a constant, is Ctot = C + I j:;,2 jCj > Ceo (9.52) Wegner and Engel 9 solved the full set of equations numerically, by computer, in a few cases as a check on approximations that they found to be accurate and relatively easy to handle. However, we omit details of this analysis and turn instead to the later paper of Wegner and Savk0 23 in which a modified model is used to obtain better agreement with experiment.
Free Single-Stranded Polymer 106 In the modified model, there are two generalizations: (a) the fast equilibrium is between monomers and small aggregates up to size n (effectively, the "critical nucleus"; n = 3 or 4 in the actual cases studied) and (b) polymers may fragment [see Eq. (9.19)] with rate constant K'. It is assumed that K' has the same value for breakage at any bond between polymer subunits (i.e., K' is independent of Nl and N 2 ). Annealing is not included in the model. Omitting fragmentation for the moment, Eq. (9.50) becomes (using Cn = Knc n) dC n+1 = (a dt + fJ)K nc n+1 - (a' + fJ')cn+1 + (a' + fJ')c n+ 2 - (a + fJ)CC n+1' (9.53) The higher equations are obvious from Eq. (9.51). There is no way, in the work being described, to separate a and fJ or a' and fJ'; Wegner and Savko use the notation k = a + fJ and k' = a' + fJ'. Based on the computer tests made by Wegner and Engel,9 approximations were used instead of solving (numerically) the full set of master equations. We sketch the approximate argument here. The small aggregates N = 2 (dimer), ... , n are increasingly unstable thermodynamically, owing to not well understood details of molecular packing. The "critical nucleus" N = n is the least stable of all aggregates (including polymers, N ~ n + 1) and has the minimum concentration (c n). Because this is a minimum, Cn+ 1 ~ Cn (after an early transient). Also, after the early transient, the sum C z + ... + Cn is small compared to Cn + 1 + Cn +2 + ... , and can be neglected. Conservation relations are: C tot ~ c*p = C + c~ (9.54) L L jCj (9.55) cj • (9.56) j?; n+1 Cp = j~n+l Obviously, cp is the concentration of polymers (N ~ n + 1) and c~ is the concentration of subunits in polymers. In this closed system, Ctot is a constant, but the other concentrations in Eqs. (9.54)-(9.56) change with time. The time derivative of cp is -dc p dt ~ ( a + fJ) CC n - = (' a + fJ') Cn +1 + K , cpo* (9.57) The first term on the right represents the formation of new polymers from n-mers by monomer addition, the second term refers to the loss of polymers (n + 1 --+ n) by monomer departure, and the third term gives the rate of production of new polymer molecules by breakage (which might occur between any polymer subunits). Using Cn ~ Cn+ 1 (see above) and Cn = Knc n, Eq. (9.57) becomes (9.58)
107 Kinetic Aspects for a Free Polymer ~.........._ c',o, = A . I 101M 20 ~=..",.,....... IS 17.7 11.7 9.4 6.9 o 1X IO~ I Cs) Fig.3-12. Calculated "best agreement" polymerization curves 23 (solid) for actin in the presence of Ca 2 +. Broken curves are optimal when fragmentation is not allowed (K' = 0). See text for further details. Finally, the net rate ofloss of monomers (by polymer on-off transitions) is ~~ ~ - [(0: + [3)c - (0:' + [3')] cpo (9.59) Equations (9.58) and (9.59) for C and c p can be solved numerically, starting with c = C tot and c p = 0 at t = O. Wegner and Savk0 23 studied the growth of actin in the presence of (separately) K +, Ca 2+ , and Mg2+. There was no detectable fragmentation in the K + case (i.e., seven plots of c~ as a function of time, with Ctot ranging from 7.4 to 20 11m, could be fitted very well taking K' = 0). However, with Ca 2 + and Mg2+, the introduction of fragmentation improved the agreement between theory and experiment considerably. This is shown in Figs. 3-12 and 3-13, respectively. In these figures, the experimental points are omitted but the solid curves (only 5 out of 7 and 3 out of 7 are shown) represent the data very well. 23 The optimal parameters are = (0:' + [3')1(0: + [3) = 2.0 11M, =4 (9.60) X 10 5 M- 3 s-2} Ca 2 + (0:' + [3')2 K n = 32 • (0:' + [3')K' = 2.5 X 10- 9 S-2 (9.61) (0:' + [3')2 Kn = 1.9 X 10 5 M- 3 s-2} M 2+ (0:' + [3')K' = 7.5 X 10- 8 S-2 g. (9.62) Ce n in both cases, and also Recall that c~ = Ctot - C so that these c~(t) curves essentially give c(t) as well (e.g., the asymptotic value of c~ is C tot - ce). The dashed curves are best-fitting
Free Single-Stranded Polymer 108 ig.3-13. 25 20 :::E arnea in ig.3-12 except for actin in the pr ence I3 of Mgl . 15 14.9 .3 O Co t.; 6. 7 5 --------0 I X Q'I 1 _x 1Q'I 3X 10' I( ) curves in the special case K' = 0 (no fragmentation). In this case, n = 3 for Ca 2 + and n = 2 for Mg2+. The dashed curves are rather different from the solid curves (and are therefore unsatisfactory), especially for Mg2+ (note that the K' parameter is 30 times larger for Mg2+ than for Ca 2+ ). Incidentally, the K + curves mentioned above but not shown, with n = 3 and K' = 0, resemble the Ca2+ family qualitatively. Although the agreement between theory and experiment is very satisfactory, it should be recalled that the size dependence of on-off rate constants and of K' (fragmentation) has been omitted from the model. Also, annealing has not been included. However, the authors 23 have made it clear that their object was to find the simplest model that would account for their data. References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. Hill, T.L. and Chen, Y. (1973) Biopolyrners 12, 1285. Hill, T.L. (1983) Biophys. J. 44, 285. Hill, T.L. (1980) Proc. Natl. Acad. Sci. USA 77, 4803. Hill. T.L. (1981) Biophys. J. 33, 353. Hill. T.L. (1961) J. Chern. Phys. 35, 303. Hill, T.L. (1968) Thermodynamics for Chemists and Biologists (Addison-Wesley, Reading, MA), pp. 46-49. Onsager, L. (1949) Ann. N.Y. Acad. Sci. 51, 627. Mitchison, T. and Kirschner, M.W. (1984) Nature 312, 237. Wegner, A. and Engel, J. (1975) Biophys. Chern. 3, 215. Risernan, J. and Kirkwood, J.G. (1950) J. Chern. Phys. 18,512. Hill, T.L. (1975) Proc. Natl. Acad. Sci. USA 72, 4918. Hill, T.L. (1976) Proc. Natl. Acad. Sci. USA 73, 679. Doi, M. (1975) Chern. Phys. 11, 115, Eq. 16. Rothwell, S.W., Grasser, W.A., and Murphy, D.B. (1986) J. Cell BioI. 102, 619.
Kinetic Aspects for a Free Polymer 15. 16. 17. 18. 19. 20. 21. 22. 23. 109 Karlin, S. (1969) A First Course in Stochastic Processes (Academic, New York). Carlier, M.F., Pantaloni, D., and Korn, E.D. (1984) J. BioI. Chern. 259, 9987. Hill, T.L. and Kirschner, N.W. (1982) Int. Rev. Cytol. 78, 1. Chandrasekhar, S. (1943) Rev. Mod. Phys. 15, 1. Abraham, F.F. (1974) Homogeneous Nucleation Theory (Academic, New York). Ferrone, F.A., Hofrichter, J., and Eaton, W.A. (1985) J. Mol. BioI. 183, 611. Erickson, H.P. and Pantaloni, D. (1981) Biophys. J. 34, 293. Voter, W.A. and Erickson, H.P. (1984) J. BioI. Chern. 259, 10430. Wegner, A. and Savko, P. (1982) Biochemistry 21, 1909.
4 Single-Stranded Polytner Modified by a Second Cotnponent, a Bound Ligand, or a Cap In this third and final chapter on single-stranded, or effectively single-stranded, polymers, we consider perturbations produced by a second molecule. In Section 10, the "second molecule" is a second component that aggregates along with the first component. In Section 11, the "second molecule" is a ligand that binds to the aggregating species all along the polymer or, in other cases, the ligand binding is confined to the ends of the polymer. Section 27 deals with a somewhat related topic. 10. Two-Component Single-Stranded Polymer In the first subsection of this section we extend the treatment, in Section 8, of the polymer size distribution at equilibrium to a two-component polymer. Some kinetic considerations for this type of polymer are then discussed in a second subsection. We shall have a free polymer in mind for generality. An attached polymer is accommodated as the special case n = 0. To avoid complexity of a rather inconsequential sort (in the translational and rotational partition functions), we shall assume that the aggregating two components, 1 and 2, have, approximately, the same shape and mass. Thus, for a polymer with Nl arid N2 subunits of the two components, the simple factor N n in Eq. (8.24) is retained, where N == Nt + N 2. That is, Nt and N2 do not appear separately in this factor, as they would in a more general model. Actually, this assumption is not very restrictive because copolymerization 1 (i.e., co crystallization in one dimension) presumably would not occur unless the two components have similar shapes. Further, two proteins with similar shapes will generally have similar masses.
Two-Component Single-Stranded Polymer 111 Fig. 4-1. Linear array of components 1 and 2 showing nearest-neighbor free energies of interaction wij' Because an attached polymer does not rotate and translate (n = 0), the above assumption is not needed in this special case. Although the two components are assumed to have similar shapes, they may have rather different intermolecular interaction free energies wll , W 12 ' and W 2 2 (Fig. 4-1) because of local contacts. Note that we take W 12 = W21 (a simplifying but not necessary assumption 2 ). The model we use (Fig. 4-1) is strictly single-stranded, in order to lead to tractable algebra when these different interactions are included in the statistics. Distribution in Size and Composition at Equilibrium There are two kinds of free subunits, with activities a 1 and a 2 . Component 1, by itself, would form infinite polymer at ae1 and component 2 at a e2 • Equations (8.26) become l1ei = 11~ + kTln aei = (i = 1,2) (10.1) 110i' The third equation here refers to pure (one component) bulk polymer i in equilibrium with free subunits at activity aei . The values of a 1 and a 2 of interest here produce, at equilibrium, a twocomponent polymer that is of significant size (N of order 10 2 , 10 3 , or more). The polymer is a one-dimensional two-component solid solution. Following Eqs. (8.23) and (8.24), we shall proceed by writing, for a polymer with N1 and N2 subunits of the two kinds, (10.2) where N = N1 + N2 (see above) and Q = e- A / kT refers to bulk properties for a mixture of N l , N2 molecules. The actual finite size of the system appears in C' and N n . Fortunately, the statistical mechanical problem in evaluating Q, including the statistical effects of different values of W ll , W 12 , and W 22 , has already been solved exactly: this is the classical one-dimensional Ising problem. The result is (Ref. 3, pp. 379-380) A = -kTlnQ = A1 + A2 + AAmix + AAexcess (10.3) (10.4)
Single-Stranded Polymer Modified by a Second Component 112 (10.5) = NkT[(1 - 8)ln(1 - 8) (R + 1 - 28) NkT [ (1 - 8)ln(R + 1)(1 _ 8) AAexcess = 8 == N 2 /N, + 81n8] + 8ln R == [1 - 48(1 - 8)(1 _ (R - 1 + 28)J (R + 1)8 y)]1/2 (10.6) (10.7) W < 0 and y > 1 correspond to positive cooperativity: subunits tend to cluster together (in the linear chain) with other subunits of the same kind. In these equations, Al and A2 refer to pure polymers, AAmix is the free energy of random mixing of the two components, AAexcess is the excess free energy arising from the intermolecular interactions (note that this term is zero if W = 0 and hence y = 1 and R = 1), and 8 is the mole fraction of component 2 in the polymer (1 - 8 is the mole fraction of component 1). Incidentally, Ref. 3 refers to the above results as the "quasi-chemical approximation." In one dimension, however, they are exact. Equation (10.3) then yields (10.8) The open-system partition function corresponding to Eq. (8.25), for a polymer in which NI and N2 are free to fluctuate (by on-off transitions), is Y = L (QN1N2) A~l A~2 N 1 ,N2 V L NnQA~lA~2 C' L Nn[ xl(R + 1) J1[ x = C' = N R +1- 28 2 (R + 1) R - 1 + 28 J2 N (10.9) (10.10) In writing Eq. (10.9), we have used Eqs. (10.1). The probability PN1N2 that the open equilibrium polymer contains N I , Nz subunits of the two kinds is proportional to the summand in Eq. (10.9). However, the variables N, 8 are more convenient than N 1 , N2 (N gives the total size of the polymer and 8, the mole fraction of component 2, specifies the composition). With these variables, P N, ° oc N n { [ Xl (R + 1) JI-O[ x 2(R + 1) JO}N. R +1- 28 R - 1 + 28 (10.11) Note that if W = 0 and hence R = 1 (random mixing, "ideal" solid solution), PN,o oc N n )l-O(X8 )OJN [( 1 X ~8 2 . (10.12)
Two-Component Single-Stranded Polymer 113 Incidentally, in the nonideal solid solution,3 R (R +1+ 1)(1 R - 1 + 28 (R + 1)8 28 - 8) and (10.13) are the mole fraction activity coefficients for components 1 and 2, respectively. Equation (10.11) could be rewritten in terms of these activity coefficients. At a given 8, the N dependence of PN,o has the same form as in Eq. (8.31). Hence fluctuations in N are large [see Eq. (8.34)]. However, fluctuations in 8 at a given N are normal, that is, small, as will be seen below. The most probable value of 8, which we denote by 8*, can be found from aIn PN • o/a8 = 0 and Eq. (10.11). After some algebra and cancellation, one finds that the equation R* + 1- 28* R* - 1 + 28* (10.14) determines 8* as a function of Xl and X2 (actually, as a function, say, of x2/xd. Here R* means R(8*) in Eq. (10.7). The value of 8* does not depend on N. It should be noticed that, in view ofEq. (10.14), the two expressions in square brackets in Eq. (10.11) are equal to each other at 8 = 8*. Let us denote X2/X l by X2l ' for brevity. Then, from Eq. (10.14), we can deduce the useful relations 8*(1 - 8*) 8* = 1 + (x 21 = X 2l /B (10.15) - 1)yl/2 B- 1/2 2 1 - _(x-=2::.::1________ - 1)yl/2 B-_ l /2 1 - 8* = _ 2 R*2 = y(l B + X21 )2/B == 4X2l + y(1 (10.16) (10.17) (10.18) - x2d 2. (10.19) Equation (10.16) gives 8* explicitly as a function of X2 l = X2/X l ' When W = 0 and hence y = 1, these relations simplify to B=(1+X 2l f, 8*= 1 X2l + X2 I , R*=1. (10.20) As already mentioned, the two expressions in square brackets in Eq. (10.11) are equal at 8 = 8*. Substitution of Eqs. (10.16) and (10.18) into either square bracket leads to (10.21) This quantity is now denoted by X for convenience. When Q 2 = 0, X = Xl; when Q I = 0, X = X 2 ; and when W = 0, y = 1, we have l X = Xl + X 2 • The probability distribution in N at the most probable composition 8 = 8* is then,
114 Single-Stranded Polymer Modified by a Second Component from Eq. (10.11), (10.22) which has the same mathematical form as Eq. (8.31) for a one-component polymer. The two-component polymer becomes very long (macroscopic) when X --+ 1. Correspondingly, Y diverges (the 0 = 0* terms are the largest in Y). A finite but large two-component polymer exists when X is less than but near X = 1. The most probable and mean values of N are given by Eqs. (8.32) and (8.33) with X in place of x. When X > 1, the polymer will grow indefinitely (see the next subsection). For a macroscopic polymer in equilibrium with solution, only one of Xl and X 2 can be independent. This is a consequence of the phase rule. Alternatively, if the composition 0* of the bulk polymer is chosen as independent variable, both Xl and X 2 are determined (these are related to the solubilities of the two components). If we put X = 1 in Eq. (10.21), we can find, say, Xl as a simple function of X 2 : (10.23) This function is plotted in Fig. 4-2 for three values of y. When y = 10 (positive cooperativity), W = -1.36 kcal mol- l at 25°C. A finite polymer has one more degree of freedom.4 Hence, although, in the y = 10 case, a macroscopic polymer is confined to the y = 10 curve in Fig. 4-2, a band of points inside of but near to this curve (X < 1) can represent finite polymers. If we take pairs of Xl' x 2 values along the y = 10 curve, and substitute X 21 = X 2 /X I from these in Eq. (10.16), we obtain Fig. 4-3, which shows Xl and X 2 as functions of 8* for the bulk polymer. Of course, for a bulk polymer, 0* is the only value of 0 of interest (because of very small fluctuations). The straight guidelines in Fig. 4-3 also represent the case y = 1. Curves like those in Fig. 4-3 will be familiar to students of solution physical chemistry. Fig. 4-2. Plot of Xl as a function of from Eq. (10.23) for a macros co pie two-component polymer at equilibrium, for three values of y. X2
Two-Component Single-Stranded Polymer 115 1.0 Fig. 4-3. Plot of Xl and X 2 as functions of e* for a macroscopic twocomponent polymer at equilibrium, for y = 10 and y = 1 (straight lines). ~-----------~ 0.8 0.6 0.6 0.4 0.8 1.0 0' We now consider fluctuations in () about () expansion In P = In P(()*) = ()*. +"21 (02Inp) ----afj2 0* (() - To do this we use the ()*)2 + .... (10.24) The first derivative is omitted because it is equal to zero. Then P = P(()*)e-(0-0*)2/2"5 1 _ (Ji - - (02Inp) ~ 0.' (10.25) (10.26) where (Ji is the variance in (). We start with Eq. (10.11) and obtain a 1nP 2 ----afj2 2 (Jo N R()(l - ()) yl/2X21(1 = + X21) NB 3/2 (10.27) (10.28) Thus (Ji is of order liN; fluctuations in () are practically negligible for a sizable polymer. When W = 0, y = 1, we have (10.29) Finally, then, we can write for PN,o in Eq. (10.11), using Eqs. (10.22) and (10.25), PN,O oc N nxN e-(0-0*)2/2"5 , (10.30) with (Ji given by Eq. (10.28). Normalization of PN,o is easy, by first integrating over ()((Ji involves N) and then over N, but we omit the result. Finally, for completeness we mention that the two chemical potentials in
Single-Stranded Polymer Modified by a Second Component 116 the bulk polymer are 3 III = 1101 112 = 1102 + kTln(1 - 8*) (R* + 1 + kTln (R* + 1)(1 + kTln 8* + kTln 28*) _ 8*) (R* - 1 + 28*) (R* + 1)8* . (10.31) (10.32) The last terms are activity coefficient terms [see Eq. (10.13)]. Then, from Eqs. (10.1), Xl = R* + 1 - 28* , R* + 1 X2 = R* - 1 + 28* R* +1 (10.33) Thus, the bulk polymer has a somewhat stronger form of Eq. (10.14): R* + 1- 28* R* - 1 + 28* R* + 1. (10.34) For a finite polymer, 8* determines X 2 /X l but not Xl and X 2 • For a bulk polymer, 8* determines Xl and X 2 (with the same X 2 /X l as in the finite case). Kinetic Topics A two-component linear aggregate is a rather special system from a kinetic point of view. The uniqueness arises from the fact that the components can interchange at a given position in the polymer, and hence introduce fluctuations in composition, only via on-off transitions at the polymer ends. Hence, the given position must reach one of the ends in order ever to change state (i.e., exchange components). A system in which a ligand can bind on each subunit of a one-component polymer (thus each subunit has two possible states) is quite different kinetically: interior polymer subunits can change states readily; the mechanism does not require use of a polymer end (see Section 11). Because a two-component polymer does not have an always present dynamic equilibrium in which interior subunits can change states (components), under most conditions the polymer cannot sample the full range of possible arrangements of the two components. Hence, such a polymer will generally find itself in a metastable configuration (arrangement) that depends on its past history. For example, a steadily shortening long polymer will shorten at a rate determined by a composition and component distribution that would have been established in a prior session of growth. An exception is equilibrium at X < 1, in which case finite polymers form. Fluctuations in polymer size (N) are large [Eq. (10.22) and Fig. 3-3] so that an ensemble of these polymers should attain a true equilibrium distribution in a reasonable time through on-off end transitions. On the other hand, a growing polymer (X> 1) will tend to have a random sequence of components, determined by the random order of addition of the
Two-Component Single-Stranded Polymer 117 two components from the surrounding solution. To the extent that they occur, off transitions will introduce some order into the otherwise random sequence. Because of this complication of metastability, kinetic aspects of twocomponent polymers are not very amenable to analytical discussion. We consider two rather limited topics here: first, detailed balance at the ends of an equilibrium polymer when X ::::; 1 and, second, the kinetics of growth (X > I) when the growth is fast enough to produce a random, rather than ordered (by molecular interactions), sequence of components. We examine on-off detailed balance at the (I. end (the f3 end has a completely analogous treatment) of an equilibrium two-component polymer with X ::::; 1. X is related to Xl and X 2 by Eq. (1O.21). Let (1.1 and (1.'1' for component 1, and (1.2 and (I.~, for component 2, be the on and off rate constants at the (I. end of bulk pure polymer 1 or pure polymer 2, respectively, as in Eq. (9.1). Then (10.35) For finite pure polymers of size N, we use the notation and relations (10.36) as in Eqs. (9.11). Note that, at this point, we are adopting the model introduced in the previous subsection (the two components have essentially the same size and mass). Consider, in the equilibrium ensemble of two-component polymers at X [i.e., Xl and X 2 are specified, in Eq. (10.21)], those polymers with size N. Let Pii be the fraction of these that have component i at the tip position at the (I. end and component j at the next position (Fig. 4-4). The sum of the four Pii is unity. Also, the probability that the tip position is component 2 is denoted (J1. Then (10.37) We omit an index N on Pii and 81 because these quantities turn out to be independent of N (for N not too small). Now consider the detailed balance in which a molecule of type 1 adds, at the (I. end, to a tip molecule of type 1 in a polymer of size N - 1 to form a tip 11 ... (Fig. 4-4) in a polymer of size N; the inverse process is the removal of the 1 from the 11 ... tip. The rates of these two microscopic processes must be equal at equilibrium (detailed balance). The on and off rate constants will Fig.4-4. Four possible states of the two tip positions at the rx end of a two-component polymer at equilibrium.
118 Single-Stranded Polymer Modified by a Second Component be the same as for pure 1 polymer of the same size. Hence the two rates are (10.38) When we introduce Eqs. (10.22), (10.35), and (10.36), this simplifies to (10.39) Next, consider the addition, at the a end, of a 1 to an N - 1 polymer with a 2 at the tip, forming a polymer 12 ... of size N. In this case an interaction free energy W12 is produced when the subunit attaches rather than Wll' as in the previous case. If this free energy difference appears entirely in the off rate constant, the off rate constant in the present case will be a'l.N(Yll/Yl2); Yij == e-wij/kT. Actually, the free energy factor Yll/Yl2 might be divided between on and off rate constants,2 but such a division would have no influence on the equilibrium properties we derive. The detailed balance expression in this case is then a1.N-l a l 8l PN-1.o* = a'1.N(Yll/Yl2)P12 PN.O*, (10.40) which simplifies to (10.41) In similar fashion, by examining the attachment of a type 2 molecule to a 2 or to a 1, we find, respectively, pzz = 8 1(X 2/X), PZ1 = (1 - 81)(xZ/X)(Y12/Y2Z)· (10.42) Note that P12 =1= P2l. At this point we introduce the simplified notation Y1 Yl2 ==-, Yll Yl2 Yz==-, Y22 Y -1 _ = (10.43) Y1YZ, In this notation, P21 = ~2Y2(1 - 81), P22 = ~281· (10.44) If we substitute Eqs. (10.44) into Eqs. (10.37), we find two expressions for 81 : 1 - ~2 + ~2Y2· (10.45) From the two right-hand members, we deduce (1 - ~1)(1 - ~2) = ~1~2y-l. (10.46) This is a generalization of Eq. (10.23) that is equivalent to Eq. (10.21). In other words, we have here a self-consistency check.
119 Two-Component Single-Stranded Polymer We note that 81 at the tip is not the same as 8* in the interior of the polymer [Eq. (10.16)] because 8* depends on Y = YllY22/yi2 whereas 81 depends on Y12/Yll and Y12/Y22 separately [~1 and ~2 in Eq. (10.45) are functions of y]. However, if all of the Yij are equal (as would be the case, for example, if component 2 differs from component 1 only by carrying a permanent nonperturbing label), X2 81 -- 8* -_ Xl (10.47) + X2 In general, though, there is an end effect: the tip composition 81 is not the same as the interior composition 8*. Substitution of one of the two expressions for 81 from Eq. (10.45) into Eqs. (10.44) gives explicit alternative expressions for the four Pij. These expressions (which we omit) depend on Y12/Yll and Y12/Y22 separately (not on Y alone, as for the interior PO; see below). When the Yij are all equal, it follows from Eqs. (10.47) that (10.48) where 81 = 8* [Eq. (10.47)]. These represent a random distribution both at the tip and in the interior, as expected. Incidentally, in the interior the pair probabilities are 3 [see Eqs. (10.11) and (10.21)] P!l= (1 - 8*)(R* + 1 - 28*) R*+l =(1-8*)~1 P*22 -- O*(R* - 1 + 28*) -- 8*;;"'2 R* + 1 pt2 = pil = (10.49) 28*(1 - 8*) 1 - (1 R* + 1 = 8*)~1 2 - 8*~2 Let 82 be the probability that the second subunit from the Then \J. . end is type 2. (10.50) From Eqs. (10.44) and (10.45), 82 = (1 - ~d(~2 + ~lYl) 1-~1+~lYl = ~~Y2 + (1 - ~d(1 - ~2). 1-~2+~2Y2 (10.51) One can verify that the two right-hand members are consistent with Eq. (10.46). The equilibrium detailed balance relations in Eqs. (10.38) and (10.40) can easily be extended from subunits in positions one-two at the tip (as in the above equations) to two-three, three-four, etc. For example ("two-three"), if we add a 1 to a 21 ... tip, (10.52)
Single-Stranded Polymer Modified by a Second Component 120 This leads to P121 = P21~IYl' There are seven other such Pijk expressions. From these eight equations we deduce, using Eqs. (10.44), PlIl = ~W - 8d, P211 = ~1~2Y2(1- 81 ), P112 = ~IY181 P212 = P121 = P221 = ~1~2y-181' P122 = ~1~2Y181' ~1~2y-l(1 - 8d, ~~Y2(1 - 8d (10.53) P222 = ~~81' These equations allow us to verify the expressions for 81 and 82 given above and also to obtain, for the third position from the polymer end, + P212 + P122 + P222 [~lYl(~1 + ~2) + ~~ + (1 - ~d(l - ~2)](1 1 - ~1 + ~lYl ~~Yz + (1 - ~d(l - ~2)(~1 + ~2 + ~2Y2) 1 - ~2 + ~2Y2 83 = Pl12 - ~1) (10.54) (10.55) Again one can show that Eqs. (10.54) and (10.55) are consistent with Eq. (10.46). Two examples at the next stage are (10.56) From the eight such expressions with 2 for the fourth subscript, we find 84 = g~Yz + (1 + (1 + ~1 ~2 + ~~ + ~2Y2(~1 + 2~2)] ~2 + ~2Y2)' (10.57) - ~d(l - ~2)[~I - ~1)2(1 - ~2)2}/(1 - On examining Eqs. (10.53) and (10.56), one can infer the general rule involved in writing such expressions, for arbitrarily large groups of n subunits at the tip [n = 4 in Eqs. (10.56)]. There is a close relation to the form of the terms in an Ising grand partition function (Ref. 2, Chapter 8). Using this general rule it is easy to see that, for large n, 8n ~ 8* as expected. It is easy to verify that [see Eq. (10.46)J ~1 1 - ~1' (10.58) ~1~2 p!lPi2 p!~ Pi2 P!2 = (1 - ~1)(1 - ~2) = y. (10.59) The last is a so-called quasichemical relation (Ref. 2, p. 331). One can show from the Pij' Pijk' and Pijkl that, at least as far as n = 4, analogues of Eqs. (10.58) and (10.59) hold also for pairs at positions (from the tip) 1, 2, at 2, 3, and at 3, 4. Presumably this extends to arbitrary n, n + 1. For example, for positions 1, 2: PlI P12 ~1 1 - ~1' P22 P21 ~2 1 - ~2 (10.60)
Two-Component Single-Stranded Polymer 121 PllP22 --=y. P12P21 (10.61) A numerical example may be informative. Let us take Xl = 0.75, x 2 = 0.70, = 0.2, Y2 = 0.5, and Y = 10 (as in Figs. 4-2 and 4-3). Then ()* = 0.4458, X = 0.9555, and N m = 109.8 (most probable size). The values of (}1 to (}4 are then found to be 0.5780, 0.5142, 0.4812, and 0.4641, respectively (converging toward 8*). Also, Pll = 0.3312 P!l = 0.4350 Y1 P12 = 0.0907 P!2 = 0.1192 P21 = 0.1546 Pil = 0.1192 P22 = 0.4235 Pi2 = 0.3266. (10.62) Although we started with detailed balance (kinetic) relations above, we deduced equilibrium properties that do not depend on rate constants. Thus all of these properties would be the same at the f3 end as at the iX end. In summary: use of detailed balance kinetic equations has allowed us to deduce equilibrium properties of the polymer tip at the two ends. These properties differ from corresponding quantities in the polymer interior. Finally, in this subsection, we turn to the problem of steady growth. The difficulty with the use of on or off rates, as in Eqs. (10.38) and (10.40), for a polymer end that is growing steadily (X > 1, N large) is that 81 and the Pij are no longer known and cannot be calculated analytically. Of course, this problem does not even exist in the steady growth of a one-component polymer. In steady growth of a two-component polymer, in contrast to equilibrium, single detailed balance relations are no longer available. For example, in Eq. (10.40), P12 (in an off process) is related to (}1 (on) (at equilibrium), but at steady state, in addition, P12 (in on processes) is related to Pl12 and P2l2 (off). Furthermore, this extra complexity propagates itself (to Pijkl, etc.). We therefore abandon the general subject and conclude with an approximate, limiting example. If we put intermolecular interaction effects entirely into the off rate constants, we have for the two steady growth rates at the iX end of a long polymer, (10.63) Ja2 = iX 2 a 2 - " iX2P22 - Y22 iX 2 - P 2 l ' Y12 where the four Pij are not known in general. However, if on transitions are much more frequent than off transitions, the subunit sequence will be approximately random. Thus Eqs. (10.48) can be used for the Pij in Eqs. (10.63), where (10.64)
122 Single-Stranded Polymer Modified by a Second Component This equation may be used to find calculated. (J1' and then Ja1 ((J1) and Ja2 ((Jd may be 11. Single-Stranded Polymer with Bound Ligand or Cap In this section we treat, rather briefly, a number of topics 5 related to the possible attachment of a ligand M (we have already used extensively the letter L, conventional for "ligand," as a length) to the ends or along the whole length of a long, effectively single-stranded, polymer. Because of the new molecular species and the related rate constants that arise, special notation is introduced. Also, to reduce other complications, we shall use concentrations rather than activities throughout this section. Steady-state polymers are treated and particular examples of real molecules M are mentioned in Ref. 5. We confine ourselves here to fundamentals for equilibrium polymers. Capping by M We consider a linear polymer in equilibrium with monomers. The monomer or subunit exists in two forms that are designated A(solution) or A(s) and A(polymer) or A(p). The polymer is assumed to be polar, with ends rx and /3. The second-order association rate constant on the rx end is rx and on the /3 end it is [3. The first-order dissociation rate constant on the rx end is rx' and on the /3 end it is /3'. The subunit flux at the two ends can be different but a single equilibrium constant Ce (the critical concentration) governs the equilibrium at both ends [Eq. (9.1)]. In this subsection we introduce a ligand, M, that is confined to the polymer ends. We ask what effect the presence of M will have on the aggregation thermodynamics and kinetics. This same question is asked in later subsections as well, where M plays different roles. We assume that M can interact both with the free monomer, A(s), and with the monomers A(p) at the ends of the polymer. We assume that each subunit A(s) can bind one M to form AM(s), with binding constant Ks. The concentration of free M in solution is CM. The total concentration of free subunits is c. Because of the binding of M, this is divided into two parts: and (11.1 ) for A(s) and AM(s), respectively. An end subunit of the polymer can be in state A(p) or AM(p). All non-end subunits of the polymer are in state A(p). That is, M is confined to the ends (hence the term "capping"). To maintain this restriction on M, we assume that
Single-Stranded Polymer with Bound Ligand or Cap 123 Fig. 4-5. Allowed transitions in equilibrium capping model. CA eM I\(s) + M Ks •• cA.M J\M (s) ~H~~tKI (I -'7 0 ) J\(p)+M • ~ J\M (p) ('70) J\(p) J\(p) J\(p) J\(p) either A(s) or AM(S) can attach to an end A(p) but neither can attach to an end· AM(p), for this would make the AM(p) a non-end subunit. Also, M can bind on an end A(p) or M can leave an end AM(p), but M cannot bind on a non-end A(p). The allowed transitions (a end) for this model, with rate constants introduced below, are shown in Fig. 4-5. The binding constants for M on an end A(p) are designated K~ and Kp. Because the free energy change on binding M to an end subunit does not involve the bulk polymer, but only the ends, we expect K~ #- Kp (the ends are different). The binding of M to (or release from) either A(s) or an end A(p) is assumed to be relatively fast: M binding is always at equilibrium, as in Eq. (11.1). The fractions of a and f3 end subunits with M bound, i.e., the fractions in state AM(p), are denoted by I]~ and I]p. Then 1'/ = ~ K~CM 1 + K~CM ' 1'/p = Kpc M 1 + Kpc M (11.2) We consider now the subunit kinetics at the a end (only notational changes are required for the f3 end). The mean rate at which A(s) adds to the a end of a polymer molecule is ac A(1 - 1],,), where 1 - 1]" is the fraction of the polymer ends in state A(p) and a is a rate constant. The corresponding mean off rate for A(p) from the a end of one polymer molecule is a'(l - I]~). For the binding of the liganded form, AM' the on rate is KCAM(l - I]~) and the off rate is K'I]~, where K and K' are the rate constants for these two processes (Fig. 4-5). The a and K "on" expressions, above, both include the factor 1 - I]~ because the end subunit must be A(p) in order to receive either A(s) or AM(S). The mean net rate of gain of subunits at the a end, at arbitrary c, is denoted by J~. Then (11.3) At equilibrium, there is detailed balance: the two pairs of terms in Eq. (11.3) are separately equal to zero. Then, using Eq. (11.2), a' e CAM ' (11.4) where cX is the critical concentration for the unliganded form of A and CXM is that for the liganded form. The critical concentration eX is the same whether
124 Single-Stranded Polymer Modified by a Second Component M is present or not. However, the concentration of the unliganded form is reduced by binding M to A [Eq. (11.1)]. From Eqs. (11.1) and (11.4), we find the relation K el K~ K' el' Ks· (11.5) The ratios ~/el' and K/K' are equilibrium constants for attaching A(s) or AM(S) to an el end. The critical subunit concentration, Ce, for polymer aggregation [A(s) --+ A(p)] is, from Eqs. (11.1) and (11.4), (11.6) An increase of CM increases the required C e because binding of M to A(s) reduces the concentration of A(s) available for bulk aggregation [Eq. (11.1)]. Note that Eq. (11.6) is a "bulk" thermodynamic equation; K~ is not involved. Similar considerations at the f3 end lead to 1 f3 el -=-=c~ f3' el' and V K Kp -=_.v' K' K~· (11.7) The rate constants v and v' at the f3 end are analogues of K and K' at the el end. Three different critical concentrations are used in Eqs. (11.4), (11.6), and (11.7). We shall generally use Ce, below. This is what is often measured experimentally. The choice is a matter of convention. We return now to Eq. (11.3) to make J~ more explicit. From Eqs. (11.1) and (11.2), (el' (1 + K' K~CM) + K~CM) . (11.8) This relation also leads to Eq. (11.6) for C e on putting J~ = 0 and using Eq. (11.5). When no M is present (c M= 0), J~ = elC - el'. On the other hand, if CMis very large, J~ = (K/K~CM)C - K'. The effective on rate constant here, K/K~CM' is small because an end subunit of the polymer is seldom in the "receiving" state A(p). An interesting feature of the fluxes J~ and Jp is that they depend on K~ and Kp, which can in general be different. Thus, although the critical concentration of the two ends must be the same for an equilibrium polymer, owing to the requirement for detailed balance in each reaction, away from equilibrium the on rates or off rates at the two ends can be differentially affected by M. Thus, for example, M can decrease the on rate at the el end more than it decreases the on rate at the f3 end. In general, therefore, a capping protein can differentially affect the polymerization or depolymerization rates of the two ends of an equilibrium polymer even though it would not affect the critical concentration at the two ends.
Single-Stranded Polymer with Bound Ligand or Cap 125 If M is a fairly large ligand, or if M in AM alters the shape of A, so that attachment of AM is inhibited, we would have K and Ka very small [see Eq. (11.5)]. In this case, J = a o(c 1 + Ksc M - 0('. (11.9) Here the rate of subunit addition and loss would be governed by 0( and 0(', and the concentration of unliganded A would simply be reduced by the factor 1 + Ksc M • At the other extreme, if M is relatively small and A and AM behave the same kinetically, so that K = 0( and K' = 0(' [note, from Eq. (11.5), that this implies Ka = K.], then again we obtain Eq. (11.9). In Eq. (11.9), CMlowers the slope of Ja but does not affect the intercept. Equations completely analogous to Eqs. (11.8) and (11.9) may, of course, be written for the f3 end. Although the effect of CMon Ce is quite simple [Eq. (11.6)], CM may have a variety of effects on the slope and intercept of the linear Ja(c) relation in Eq. (11.8), depending on the relative magnitudes in the pairs Ks and K a , 0( and K, and 0(' and K'. Figure 4-6 illustrates the effect of M on aggregation kinetics in a special case, Eq. (11.9) (applicable at both ends). Finally, we relate the ratios of rate constants to chemical potentials. Let 110 be the chemical potential per subunit A(p) of the bulk polymer (the ends make a negligible contribution). At equilibrium between A(s) and A(p), the two chemical potentials are equal: IlX + kTln cX = 110' (11.10) J Fig. 4-6. Illustration of effect of ligand M on aggregation kinetics of a polymer, according to Eq. (11.9).
Single-Stranded Polymer Modified by a Second Component 126 where IlX is a standard chemical potential for A(s). From Eq. (11.4), a a' (11.11) kTln- = IlX - 110' We can replace a/a' here by P/P'. From Eq. (11.5) we also have K kTln- = IlX - 110 K' For v/v', replace K~ 'K~ + kTln-. Ks (11.12) by Kp. Small Hybrid Structural Cap Two types of capping elements have been described: large macromolecular structures, which we call structural caps, that probably interact with a number of subunits at the end of actin and microtubules; and small molecules or proteins that interact with single subunits, which we call molecular caps. Kinetochores and centro somes for micro tubules or membrane insertion sites for actin filaments are potential examples of structural caps. In some cases the structure is large enough to be visible in the light microscope. Small drugs like cytochalasin B or colchicine that interact stoichiometrically with monomers at the end of the polymer are examples of molecular caps. Chapter 2 is concerned with structural caps, without using this terminology. The major difference between the two types of capping elements is that, by remaining attached to the filament at all times, the structural cap can more effectively influence the further assembly of subunits, usually by inhibiting it. The molecular cap, as we have defined it, once removed diffuses away from the end of the filament and does not influence the assembly reaction. There is now considerable discussion of subunit insertion where filaments meet structural elements such as at the kinetochore, or centrosome, or membrane attachment sites for actin filaments (see Section 7, for example). In the previous subsection we have treated only molecular caps. In this subsection we present the analysis of a simple hybrid model of a small structural cap that illustrates how the special properties of structural caps arise. In this hybrid model we suppose that M is a rather large molecule that does not bind to a free subunit A(s) (Ks = 0) but it may bind, in a rapid equilibrium, to an end subunit A(p) [Eq. (11.2)] to form an end AM(P). The binding constant is K~. A free subunit A(s) may add onto an end A(p) but not onto an end AM(P), just as above. An end A(p) may leave the polymer but an end AM(p) cannot leave. The rate constants for A attachment-detachment at the a end are a and a', as above. So far this is just the model in the preceding subsection with Ks = 0, K' = 0, and no free AM(S). Consequently, the rate constants K and K' are not involved. The new feature here is that we also assume that a subunit A(s) may insert itself between M and the end subunit, with rate constant k,
Single-Stranded Polymer with Bound Ligand or Cap 127 c Fig. 4-7. Allowed transitions in a model of incomplete blockage by M of the (J. end of a polymer. 1\ (s) c a ItA! t u (1)0) Ka M k ~ k (1-1)a)l\(p)+M::;:=I\(p) I\(p) I\(p) 1\ (p) 1\ (p) I\(s) and the end subunit itself may detach with rate constant k', leaving M behind. Thus there is incomplete blockage of attachment-detachment. The allowed transitions are summarized in Fig. 4-7. Figure 4-8 presents an explicit but hypothetical example to illustrate exactly how subunits might insert while M remains attached. This is the a end of an actin-like polymer with two strands and a I-start helix (there is only one site for addition or loss of a subunit at each end). The polymer end can exist in six possible states, A through F. An M bound to the polymer may be in binding state I or II. There is a rapid equilibrium between I (i.e., B + E) and II (i.e., C + F); I is strongly favored in the I +2 II equilibrium because it has a lower free energy. The two equilibrium state probabilities are denoted PI and Pn, with PI + Pn = 1 and PI » Pn' The overall binding constant of M to the polymer end is Ka; the hypothetical separate binding constants are PIKa and PnKa. The elementary rate constants for the possible transitions are shown in the figure. In the a; transition in B, state B --+ F. Similarly, for the transition an in C, C --+ E; for a; in E, E --+ C; and for an in F, F --+ B. The mean rate constants k and k', introduced above, are then expressed in terms of the elementary rate constants, in this example, by k = anPn and k' = a;PIThat is, when M is bound, a subunit can add to state II only and a subunit can be lost from state I only. Thus, in this example, we have an explicit model for incomplete blockage by M. Returning now to the general argument, the detailed balance rates at equilibrium are [compare Eq. (11.3)] (11.13) with '1a determined by CM in Eq. (11.2). The critical concentration of A(s) is ~. ~~ ··· ... A B c D E F Fig. 4-8. Explicit model of incomplete blockage of a polymer end by M, as in Fig. 4-7. Elementary rate constants and states are shown.
128 Single-Stranded Polymer Modified by a Second Component C e a' a k' =-=- k . (11.14) Bulk polymer is in equilibrium with A(s) by two different mechanisms; however, the free energy change and equilibrium constant are the same. At arbitrary c, the rate of growth at the a end is then (11.15) It is easy to see that this gives the same Ce as in Eq. (11.14), if we put Ja = 0, because 1 - '1a and '1a occur in the same way in the two brackets, [ ]. Consequently the apparent '1a dependence of Cc drops out. If when M binds it precludes any further subunit exchange (i.e., the rate of subunit addition in the presence of a cap is zero), then k = 0 and k' = O. This is the same system as in the previous subsection with Ks = 0 and K' = 0 [see Eq. (11.3) and Fig. 4-5]. If, however, the cap is not present (c M= 0 and '1a = 0), Eq. (11.15) reduces to Ja = ac - a', as in simple aggregation. If CM--+ 00 and '1a --+ 1, we have partial blockage of subunit exchange by an alwayspresent cap, M: Ja = kc - k'. By varying CMbetween CM= 0 and CM= 00, we pass continuously from the a, a' system to the k, k' system, but with a constant critical concentration Ce (see Fig. 2-7). Capping with No Exchange of M on Polymer The present case is a variation on the first subsection. No known example exists. However, since both micro tubules and actin polymers are always isolated in equilibrium with their subunits, it is difficult to know if a cap binds first to a free subunit before binding to the polymer. In this subsection we assume that M cannot bind reversibly to an end subunit on the polymer, but can bind reversibly to the free monomer. When the complex binds to the filament it blocks further subunit addition, as in the cases treated before. We now assume that M cannot bind on an end A(p) nor can M be released from an end AM(P). Otherwise the present model is the same as in Fig. 4-5. Thus A(s) and AM(S) can attach to A(p) but not to AM(p), etc. The allowed transitions are shown in Fig. 4-9. The present case could arise, for example, if the binding site for M on A is covered or blocked by a neighboring polymer subunit when A or AM attaches to a polymer end, thus preventing M from entering or leaving the binding site. A rapid equilibrium of M on A(s) in solution, with binding constant K s' is still assumed. The binding constants Ka and K p are still well defined, e.g., by Eq. (11.5), even though the rate constants for binding M on an end A(p) or for the release of M from an end AM(p) are essentially zero. The equilibrium properties of the polymer are necessarily unchanged because, in this new model, we have merely eliminated some transitions from the kinetic mechanism without making any states inaccessible.
Single-Stranded Polymer with Bound Ligand or Cap Fig. 4-9. Allowed transitions when M cannot exchange with the polymer end. 129 C,\ eM A(s) + M (1-1)0) -K, ~ A'J (s) A(P) AM(p) (1)0) A(p) A(p) A(p) A(p) Equilibrium Binding of M on Bulk Polymer Only For micro tubules several proteins have been described that interact stoichiometrically with the bulk mass of protein in the polymer. Tropomyosin was the first actin-associated protein that was found to bind along the length of the actin helix. It is also associated with actin in nonmusc1e cells. To describe these and other ligands that interact with the bulk polymer, M is no longer confined to the ends but permeates the bulk polymer. We assume here that M binds on subunits of the polymer, A(p), but not on free subunits, A(s). Thus there are molecules ofM and A(s) free in solution but no AM(S) (Ks = 0). The attachment -detachment transitions at the polymer ends involve A only, not A complexed with M. The binding and release of M on or from the polymer are assumed to be relatively fast transitions; there is always binding equilibrium. M is presumably but not necessarily a large molecule (e.g., tropomyosin on actin or tau protein or MAPs on microtubules) that binds to n polymer subunits, where n ~ 1. In the case of tropomyosin, n = 7. The fraction of polymer subunits occupied or covered by bound M is called 0; the range of 0 is 0 ~ 0 ~ 1. The model is shown in Fig. 4-10. Bound M influences attachment-detachment in this system by lowering the chemical potential of A(p). This stabilizes the polymer and lowers the t A(s) ~ t~, eM K O:;f: 0 A(p) A(p) A(p) Alp) OA A(p) Fig. 4-10. Allowed transitions in model in which M (shown as a rectangle) binds to the bulk equilibrium polymer but not to free subunits. In this case n = 2. (1') A(p)
Single-Stranded Polymer Modified by a Second Component 130 critical concentration Ceo In order to deal with this effect quantitatively, we consider the polymer to be an equilibrium solid solution (see Section 10 and Ref. 3, Chapter 20). The solid solution is in contact with another phase containing M and A(s) in water as solvent. There is an analogy with osmotic equilibrium that will be pointed out below. The total number of subunits in the polymer is called N, NA is the number of unoccupied subunits, and NM is the number of bound molecules M. Because each bound M covers n subunits, the number of occupied subunits is then nNM, and N = NA + nNM. Also, 8 = nNM/N. To begin with we choose Nand NM as composition variables for the binary solid solution (the polymer). Then the respective chemical potentials /lA and /lM are defined in terms of the Gibbs free energy G by dG = /lA dN + /lM dNM G = /lAN (T, p constant) + /lMNM' (11.16) The chemical potential /lM is relevant in the binding equilibrium of M because in this process NM changes with N held constant. Also, /lA is relevant in the attachment-detachment equilibrium of A at the polymer ends, because in this process NA changes with NM held constant, which is equivalent to a change in N with NM constant (because N = NA + nNM, dN = dNA when NM = constant). An alternative choice of variables is NA and N M: dG = /lA(dNA + ndNM) + /lMdNM (11.17) G = /lANA + /lAMNM, where /lAM == n/l A + /lM' Incidentally, Eq. (11.17) confirms our interpretation of /lA above. At binding equilibrium of M, /lM(polymer) = /l~ + kTln cM(solution). (11.18) To maintain generality, we assume that the extent of binding, 8, is measured as a function of CM' to give 8(c M) or the inverse function cM(8). If n = 1 and the binding of M on A(p) is noncooperative, this relation is simple and wen known: (11.19) where Kp is the binding constant for M on subunits of the polymer. But, in general, 8(cM) might be quite complicated (n > 1; cooperativity). For some models (see below) cM(8) is available explicitly, but not 8(cM). With 8(c M) available (experimentally, or from a model), we can use /lM to obtain /lA' required below, at arbitrary CM or 8, as follows. From Eqs. (11.16), we have the Gibbs-Duhem equation
Single-Stranded Polymer with Bound Ligand or Cap N dl1A + NM dl1M = 0 131 (p, T constant) and then (11.20) Note that I1A always decreases as CM increases. On integrating between CM and CM' we obtain I1A = + kTln H, 110 = 0 (11.21) where 110 is the same as in Eq. (11.1 0) and l1 In H(c M ) == -- n cM 0 8(c~) d In c~. (11.22) The term kTlnH, which is always zero or negative (i.e., H ~ 1), is the correction to 110 owing to the binding of M on the polymer, with In H = 0 (i.e., H = 1) when CM = and In H --+ -00 (i.e., H --+ 0) when CM --+ 00. In case cM (8) is available analytically, but not 8(cM ), an alternative form (after an integration by parts) is ° InH(8) = ~[ -8Inc M (8) In the special case ofEq. (11.19) with n H = = 1 + Kpc M + J: In cM (8' )d8J (11.23) 1, = 1- 8. (11.24) If the binding of M is on sites that are in independent one-dimensional grooves of the polymer, with nearest-neighbor interactions and n ;:, 1, as is apparently the case with tropomyosin on actin,6 then the McGhee-von Hippel theory (see Ref. 2, pp. 344-350, for a review) provides a moderately complicated but exact and explicit expression for cM (8). However, when n > 1, In H would have to be found in a particular case by numerical integration of Eq. (11.22) or (11.23). When n = 1, this binding model is the same as the usual one-dimensional Ising problem. In this case (Ref. 3, Eq. 14-16), H R + 1 - 28 R +1 ' = -=-----:-- (11.25) where (11.26) and w is the interaction free energy between two M molecules bound on nearest-neighbor subunits (or w might be due to altered interactions between the subunits themselves). Equations (11.21) and (11.25) are equivalent to Eq. (10.31). When w = 0, Eq. (11.25) reduces to Eq. (11.24). The binding isotherm corresponding to Eq. (11.25) is [Ref. 3, Eq. 14-13 and Eqs. (10.33)]
Single-Stranded Polymer Modified by a Second Component 132 K C eW/kT(R - 1 + 1- + 28) =------- R P M 28 (11.27) . This reduces to Eq. (11.19) when w = O. Turning now to the subunit exchange (A, not AM) at the two ends of the equilibrium polymer, with bound M, we have, corresponding to Eqs. (11.4), (11.10), and (11.11), I1X + kTln Ce = 110 + kTln H. rxce = rx', (11.28) Here rx and rx' are mean rate constants for the rx end of the polymer. Then rx kTin rx' If we use the notation = rxoc~ = rx~ I1X - 110 - kTin H(c M )· in the absence of binding (c M (11.29) = 0), then I1X + kTlnc~ = 110 (11.30) Recalling that H < 1 when CM > 0, we verify that binding of M stabilizes the polymer and reduces Ceo For a single small ligand (n = 1) showing no cooperativity of binding [Eq. (11.24)], the result of these calculations is that C e = c°(1 - 8) = e CO 1 e + Kpc M (11.31) Equations (11.28)-(11.30) also apply at the p end of the polymer (replace the rxs by ps) because H is a bulk property of the polymer, not an end property. The thermodynamic equation Ce = c~H, for example, is not concerned with the ends at all. A close analogy to osmotic pressure may be of interest. Binding of M reduces I1A [Eq. (11.21)] whereas a compressive force on the polymer ends, - F (F negative), increases I1A [Eq. (4.11)] by an amount -IJ, where 1o = LIN and L is the polymer length. The force - F just necessary to maintain I1A at the value 110 and Ce at the value c~, in the presence ofa small amount of binding 8 = nKpc M , is easily seen from Eq. (11.22) to be kT8 -/oF=-kTlnH=- n or (-F)L=NMkT. (11.32) The latter equation has the familiar osmotic form; a small amount of binding is a "colligative" property. There are obvious resemblances and differences between the above treatment and Section 10. In both cases the polymers have two components but the two components are permanent in Section 10 whereas here the two components are rapidly interchangeable via the binding equilibrium.
Single-Stranded Polymer with Bound Ligand or Cap 133 If there are two different molecules M1 and M 2 , at concentrations C 1 and c2 , that bind on the polymer, the above treatment is changed very little in its fundamental aspects. In fact, the formal extension to any number of bound species is easy. If N1 molecules of species 1 are bound, each occupying n 1 subunits, and similarly for component 2, then (11.33) (11.34) The Gibbs-Duhem equation in this case is N df.1A + N1 df.11 + N2 df.12 = 0 (p, T constant) (11.35) so that (11.36) Here (]1 and (]2 are both measurable functions of C 1 and between C 1 = C2 = 0 and c 1 , c2 , we obtain f.1A = f.10 C2' + kTlnH On integrating (11.37) (11.38) The path used in this integration (and corresponding binding experiment) is arbitrary, because H is a state variable that can depend only on the final concentrations C 1 and c 2 • For example, C 1 could be raised from C 1 = 0 to C 1 holding C2 = 0 constant, and then C 2 could be increased from C2 = 0 to C 2 holding C 1 constant. The important point is that kTln H, using this extended definition of In H, is the correction to f.10 (to obtain f.1A) irrespective of whether one, two, or more kinds of molecules bind on the polymer. Equations such as (11.28)-(11.30) are unchanged; only the calculation of H is different. In the simple special case n 1 = 1, n 2 = 1 and noncooperative, competitive binding of the two species M1 and M2 on the subunits of the polymer, it is easy to see that (11.39) H where Kp1 and subunits. Kp2 = 1- (]1 - (]2 = - - - - - - - 1+ K p1 C 1 +K are the binding constants for M1 p2 C 2 ' and M2 (11.40) on the polymer
Single-Stranded Polymer Modified by a Second Component 134 Equilibrium Binding of M on Bulk Polymer and on Free Subunits In the preceding subsection we had in mind, primarily, molecules M that are so large that they do not bind significantly to free subunits in solution but they do bind on a lattice of subunits (the polymer). Here we consider much smaller molecules M that bind, at equilibrium, either on free subunits in solution [Eq. (11.1), binding constant Ks] or on individual subunits of the polymer (n = 1, binding constant Kp). The simple binding in Eq. (11.19) is presumably the most important case, but the binding of M on the polymer could be cooperative [Eq. (11.27) is an example]. To retain generality in this respect, we use In H as in Eq. (11.22) (with n = 1). We assume that there is exchange of both A and AM between polymer ends and the solution. The model is shown in Fig. 4-11. The two polymer chemical potentials that are relevant in these equilibria are J1A and J1AM in Eq. (11.17) where J1AM == J1A + J1M' Incidentally, this same relation between chemical potentials holds in solution because of the equilibrium binding of M on A(s) to form AM(S). As in Eqs. (11.10) and (11.28), in the A(s) ~ A(p) equilibrium, acX = a', J1A + kTln cX = J10 + kTln H. (11.41) Both a and a' are averages over the a end composition, A(p) or AM(P). We then obtain in this case Eq. (11.6) for Ce and Eq. (11.29) for ala' or PIP'. If we use Eq. (11.1) to relate cX to Ce' and denote the criticalconcentration at CM = 0 by c~ (H = 1 here), then we find [compare Eq. (11.30)] ce _ - 0 ce (1 + KscM)H --+ c~(l 1 + KscM ) + Kpc M , (11.42) where the last form is for the usual simple binding case. In this case, CM has no effect on Ce if Ks = Kp. In the AM(S) ~ AM(p) equilibrium, (11.43) where J1M is given in Eq. (11.18). Both c" Ids) eM + ~ M ---- ;;: Ks K and K' are also averages over the a ci\M A M(s) A(p) or AM(p) A(p) eM K P Alp) M~AM(P) AM(p) A (p) Fig. 4-11. Allowed transitions for a model of a polymer in which M can bind on individual free subunits as well as on subunits (n = 1) of the polymer. In general, both A(s) and AM(S) may attach to either A(p) or AM(p) at the polymer end.
Single-Stranded Polymer with Bound Ligand or Cap 135 end composition, A(p) or AM(P). If we use the conventional relation (11.44) .uXM -.uX -.u~ = -kTlnKs to eliminate .uXM from Eq. (11.43), we then obtain for for the polymer) K/K' (the affinity of AM (11.45) The same equation holds for v/v' at the f3 end. On comparison with Eq. (11.29), we find (11.46) This is consistent with Eq. (11.1). In the Eq. (11.19) special case, if we put H = 1- e, CM = (1 _ e e)Kp' (11.47) Eqs. (11.29) and (11.45) become IX kTln- =.uX IX' K kTln-, K = -.uo - 0 /I ""A - ""0 /I - kTln(1 - e) eKs kTln-. K ( 11.48) p The binding constants here take care of the relative stability of M in AM(S) and in AM(p) [compare Eq. (11.12)]. The terms in eand 1 - eare expected from the statistical thermodynamics of ideal binary solutions (Ref. 3, Section 20-1). The rate constants in this problem are influenced by bulk thermodynamic effects, not by end effects (as in the capping models). Simple and plausible possibilities are IX' ex 1 - e, K' ex e, with IX and K independent of e. Equilibrium Binding of M on Free Subunits Only (Kp =F 0) In this system M binds rapidly on A(s) to form AM(S) but there is no exchange of M with polymer subunits. Both A and AM can enter and leave the polymer at the ends, so that the bulk polymer contains both A(p) and AM(p) (again we consider M to be a small enough molecule so that n = 1). The inability of M to exchange directly with polymer subunits has a kinetic, not a thermodynamic, origin: the binding site for M on A(p) is buried or blocked by neighbors in the polymer lattice. The equilibrium polymer, at equilibrium, has the same properties as in the previous subsection: though direct M exchange with A(p) does not occur, the polymer is, in principle at least, able to reach the same final equilibrium state (with M bound to bulk polymer) through AM exchange at the polymer ends.
Single-Stranded Polymer Modified by a Second Component 136 (' A f\(s) eM + M ~ :x AM(s) A(p) or AM(p) Fig. 4-12. Allowed transitions for a model of a polymer in which M can bind only on individual free subunits, even though Kp "# O. A(p) A(p) AM(p) AM(p) A(P) Consequently Eqs. (11.41)-(11.48) all still apply. The model is illustrated in Fig. 4-12. The binding constant Kp is still well defined, for example, using f) = Kpc M , at equilibrium, when f) and CM are small and have been measured (i.e., Kp = f)/c M ). There is, however, the same practical limitation on reaching equilibrium in some circumstances as has been discussed already in Section 10. References 1. Oosawa, F. and Asakura, S. (1975) Thermodynamics oj the Polymerization oj Protein (Academic, New York). 2. Hill, T.L. (1985) Cooperativity Theory in Biochemistry (Springer, New York). 3. Hill, T.L. (1960) Introduction to Statistical Thermodynamics (Addison-Wesley, Reading, MA; also Dover, New York, 1986). 4. Hill, T.L. (1964) Thermodynamics oj Small Systems, Part II (Benjamin, New York). 5. Hill, T.L. and Kirschner, M.W. (1982) Int. Rev. Cytol. 84, 185. 6. Wegner, A. (1979) J. Mol. BioI. 131, 839.
5 "Surface" Properties of Som.e Long Multi-Stranded Polym.ers This chapter 1 and the next deal explicitly with simple multi-stranded polymers. Hitherto some multi-stranded polymers have been included only as "effectively" single stranded. We use simple, illustrative models rather than realistic ones (e.g., a microtubule); even some of these simple models require Monte Carlo calculations. This chapter is specialized in the sense that it is not concerned with the full polymer molecule but rather only with its "surface" or tip: we consider the structure and properties of one end of a long polymer molecule. It is immaterial whether the other end is attached or free. The concentration here on properties of the polymer tip or surface is reminiscent of a similar preoccupation in the second part of Section 10 (on two-component polymers). The full open partition function for finite attached polymers, for some of the simpler models treated here, will be considered in Section 19 (Chapter 6), after a new method for handling such partition functions has been introduced. However, in this chapter, only a surface partition function appears. 12. General Discussion of the Models This chapter provides an introductory treatment of the surface free energy, which is related to molecular roughness, at an end of a long tubular linear polymer or aggregate comprised of s strands. Structural roughness arises as a consequence of gain or loss of individual subunits at concentration c and activity a. The polymer end may be in a dynamic equilibrium with the free subunits (a = a e ) or may be gaining or losing subunits at a steady rate (a # a e ). The inspiration for-this study is the aggregation of micro tubules, which have
"Surface" Properties of Some Long Multi-stranded Polymers 138 s= 3. Case 1:3 s = 3, Case 0:3 _---1l2 ::: --------1/1 ',·/13 . '" +1 - ------1112 =0 =- 1 =+ 1 s= 3 h h v .... f---+--+---1 1 i (b) =3 2 3 ( c) Fig. 5-1. (a) Tip of three-stranded (i = 1,2,3) "staggered" tubular polymer with 1/3 vertical rise as i increases. Dotted subunits (i = 3) show tubular structure. Vertical neighbor interaction v; horizontal interaction h. Height of a strand (nJ is measured in thirds of a subunit height, relative to n1 = 0. (b) Cross-sectional view of three-stranded tubular polymer. (c) "Aligned" tubular polymer with three strands. The m i measure strand height in full subunit heights, relative to ml = 0. Arrows indicate missing horizontal interactions. 13 strands (s = 13). However, much simpler cases are examined here. These simpler cases have their own intrinsic interest because surface roughness would be a general property of multi-stranded aggregates. Also, these cases demonstrate that the usual linear subunit flux equation, J = aa - a', Eq. (5.17), would not generally be expected to hold for multi-stranded polymers (s> 1); i.e., J(a) is nonlinear. This is a second source of nonlinearity in J(c), if a =F c [see Eq. 5.18]. We consider simple tubular aggregates constructed from hypothetical isotropic subunits or blocks (e.g., protein molecules) that are generally staggered helically [Fig. 5-1 (a)] but may be aligned horizontally as a limiting case [Fig. 5-1 (c)]. Figure 5-1 (b) is a transverse section showing, for s = 3, that i = 3 is a neighbor of i = 1 (to form a tube). Figure 5-1(a) is the case s = 3 in which the principal neighbor to the right of any subunit is raised 1/3 of a subunit height, thus forming a I-start right-handed helix (which we designate 1: 3). Viewed as a left-handed helix, this same structure would be designated 2: 3. Hence the right-handed cases 1: 3 and 2: 3 would have the same kinetic and thermodynamic properties; we need not consider both cases. The same is true for any pair s': sand s - s': s (e.g., 5: 13 and 8: 13 in a microtubule). We shall assume in all models that there are no lattice vacancies (each strand is solid to its end), that subunits do not migrate from one strand to another [i.e., surface structures, as in Fig. 5-1 (a) and 5-1 (c), change only by subunit on and off transitions], and that vibrational partition functions of subunits do not contribute significantly to the surface free energy differences introduced below, and hence can be ignored. For any case s': s, in the bulk polymer (tube) each subunit [e.g., see the heavy
General Discussion of the Models 139 subunit in Fig. 5-1 (a)] interacts with both horizontal (h) and vertical (v) nearest neighbors. Let W h and Wv be full subunit-subunit nearest-neighbor interaction free energies (these are negative quantities relative to a zero at infinite separation). Any given bulk subunit [e.g., the heavy subunit in Fig. 5-1(a)] is involved in interactions with total free energy 2wv + 2wh , but only half of this can be assigned to the given subunit. Thus, bulk polymer with N subunits has a total interaction free energy N(wv + wh ). This should be compared with Eq. (4.22) where z = 4 and Wv = Wh = W [see also Fig. 1-3(c)]. If a long polymer with 2N subunits is broken in half to form two new ends and surfaces, s vertical interactions and 2m horizontal interactions are lost, where m ~ 0 depends on the surface structures created in the break; m may be a fraction. After the break, the total polymer interaction free energy is 2N(wv + wh ) - SWv - 2mwh · This is larger (wv and Wh are negative) than the pre-break free energy 2N(wv + wh ). The difference in the two quantities, per end, is the surface free energy of one (either) end: (12.1 ) This is a positive quantity. The vertical contribution to G" - swv!2, is the same for every break, but the horizontal contribution depends on m. Thus the term - mWh is the interesting part of Gs ; m is an index of molecular roughness at the polymer end. For example, if the surface structure in Fig. 5-1(c) is thought of as having been formed by a break, two horizontal interactions are missing (arrows): 2m = 2 and m = 1. Incidentally, the minimum possible value of m for this polymer is m = O. In Fig. 5-1 (a), missing horizontal interactions at the surface are 1/3 (between strands i = 1, 2), 2/3 (i = 2, 3), and 1/3 (i = 3, 1). Thus, 2m = 4/3 and m = 2/3. This is actually the minimum possible value of m for this case (1 : 3). For any given surface structure, m may be calculated systematically as follows. In the special (aligned) case 0: s [as in Fig. 5-1 (c)], we arbitrarily select strand i = 1 as the reference strand and assign the position of its end the value m1 == O. The position (height) of the end or tip of each strand i = 2, ... , s is then measured relative to the end of strand i = 1, in units of subunit height, and denoted mi' which may be a positive or negative integer, or zero. For example, in Fig. 5-1(c), m2 = 1 and m3 = O. Then (12.2) The sum here is a measure of the amount of exposed vertical suface at the polymer end and hence a measure of missing horizontal interactions and of molecular roughness. In other cases s' : s, with s' > 0 (staggered), for convenience we first measure the amount of vertical surface not in integral units (as above) but in fractional units l/s. Just as m i measures the height of the end of strand i relative to the
140 "Surface" Properties of Some Long Multi-stranded Polymers end of strand 1 in units of subunit height, ni is used here to express the same quantity in units of lis of the subunit height. As before, n i == O. The ni are positive or negative integers, or zero. Hence (12.3) The division by s here corrects for the fractional units (lis) used. That is, In21 = slm21, etc. As an example, in Fig. 5-1(a), the sum in Eq. (12.3) is 1 + 2 + 1 = 4 and m = 4/6 = 2/3 (as already found). Equilibrium Surface Partition Function When the polymer end is in equilibrium with free subunits (a = a e ) via on and off transitions, the polymer end will pass, stochastically, through (in principle) an infinite number of discrete surface structures, each with a definite value of m, as calculated from Eq. (12.2) or Eq. (12.3). There is a Boltzmann probability distribution at equilibrium among these structures, a structure with m having a relative weight emwh/kT = pm, where 0 ~ p == e wh/kT ~ 1. The limiting case p = 1 (i.e., Wh = 0) corresponds to independent strands. When the strands are independent, fluctuations in strand height differences become indefinitely large. Also, when Wh ~ -00, P ~ O. The larger the value of m, the rougher the surface, the larger the surface free energy [Eq. (12.1 )], and the lower the relative weight of the state. Note that Wy is not involved here because the Wy term in Eq. (12.1) is the same for all stuctures. In staggered cases, to avoid fractions, it is usually convenient (but not necessary) to introduce n = sm and Yf == pl/s so that pm = pols = Yf". In realistic cases both p and Yf lie between 0 and 1. All possible surface structures can be generated by allowing each of m 2 to ms in Eq. (12.2) (aligned cases) or each of n2 to ns in Eq. (12.3) (staggered cases) to take on all possible values from -00 to +00. Several or many structures so generated may have the same value ofm. Let R(m) be the number ofstructures (degeneracy) with a particular m. Then the surface partition function, the sum over all surface states, is Qs == I m R(m)pm = I (12.4) S(n)Yf", " where these sums are over all possible values of m ~ 0 or n ~ 0 and R(m) = Sen) is the degeneracy. The subscript s on Qs or Gs refers to "surface," not to the number of strands. In nontrivial cases, early R(m) or Sen) can be found by computer enumeration. In many cases Qs can be expressed in closed form (see Sections 13 and 14). It should be noticed that we are not concerned here with the length of the polymer, which has large fluctuations, but only with the distribution in surface structures (at the polymer end), which have different values ofm corresponding to different degrees of molecular roughness.
General Discussion of the Models 141 The probability that the surface has a particular value of m or n is P = R(m)pm = P* = S(n)l'/n Qs m n (12.5) Qs· The asterisk is used when the index is n = sm. From this it follows that _ p dQs 1 1'/ dQs m=--=--- Qs dp (12.6) s Qs dl'/ 2 din 1 din am = p dp = ~ 1'/ dl'/ ' (12.7) where in is the mean value of m and a;' is the variance in m. Because all strands are equivalent (at equilibrium or steady state), the mean values of the separate terms in the sums in· Eqs. (12.2) and (12.3) must all be equal. Hence _ sq r m=-=2 2' (12.8) where q and r are used h~re to represent anyone of the terms in the sum in Eq. (12.2) or (12.3), respectively, for example, Im21 or In21. Of course r = sq. Although it is easy to calculate the mean values q orr [from in, Eq. (12.8)], the determination of the complete probability distribution in q or r requires further details, which we turn to below. This distribution is of some interest because Pm or Pn* gives the distribution in molecular roughness for the complete surface but the probability distribution in q or r refers to the individual elements of roughness (height difference) between two neighboring strands. Each of the R(m) surface structures with the same m value has the same weight in the equilibrium distribution. Also, each of these structures has s values of q [Eq. (12.2)], all with the same weight. Some of these q values may be repeats. The total number of q values for a given m is then sR(m). Of these, let W(q, m) be the number with a particular value of q. Then Lq W(q, m) = sR(m). The probability of a given m and also a given q is then R(m)pm W(q, m) Qs sR(m) W(q,m)pm (12.9) sQs Then the probability of a given q, irrespective of the m value, is 1 Pq = -Q L W(q, m)pm. (12.l0) Ssm This is the desired probability distribution in q (e.g., 1m2 I)· Values of W(q, m), for q and m not too large, can be found by computer enumeration. Closed expressions for the Pq can be obtained in some cases (see Sections 13 and 14). With the Pq available, at least in principle one can calculate, for example, the mean q [already known from in, Eq. (12.8)] and the variance Note that the a;.
142 "Surface" Properties of Some Long Multi-stranded Polymers series in Eq. (12.10) has p to the power m, not q. Hence equations like Eqs. (12.6) and (12.7) are not applicable. r Exactly the same argument applies if we use (for staggered cases) the ni and [Eq. 12.3)] in place of the m i and q [Eq. (12.2)]. The result is P:( = pq ) 1 = -Q L U(r, n)'1 n , s (12.11) s n where the asterisk is used for the index r = sq, and U(r, n) [ = W(q, m)] is the number of occurrences of a particular value of r among the sS(n) terms in Eq. (12.3) for the S(n) surface structures with a given value of n. Just as r = sq, we also have (); = S2();. The two modes of calculation (mi> q; ni , r) are completely equivalent; the choice is a matter of notational convenience only. Rate Constants and Detailed Balance Consider, at equilibrium, a particular surface structure or state A and another structure or state v that is reached from A on the addition of one subunit from solution (a = a e ) onto the end of one of the s strands of the polymer originally in state A. Aside from the change in surface structure, in effect one subunit has been added to the bulk polymer. However, this feature does not involve any free energy change because of the equilibrium between polymer and solution. The values of m in the two states are designated m(A) and m(v). The change in surface free energy in the process A -4 V is then [Eq. (12.1)J L1Gs = [ -m(v)whJ - [-m(A)whJ = [m(A) - m(v)]wh. (12.12) The ratio of the probability of the two states (at equilibrium) is (12.13) Let a",v be the second-order rate constant for the on process, A -4 v, and let a:", be the first-order off rate constant for v -4 A. Because of detailed balance between the two states at equilibrium, (12.14) Reference on and off rate constants, a and a', are defined and apply when .1.Gs = O. Examples with .1.Gs = 0 are shown in Fig. 5-2. The constants a",v and a:", differ from a and a', respectively, because of the surface free energy change. When states A and v are such that .1.Gs = 0, aa e P1 = In the general case (.1.Gs =1= a'p~, p~ = p~, and (12.15) 0), then, p~ = e- f1G ,/kT = a"'vae a:", p1 aa e = a'. a",v/a a:",/a' . (12.16)
General Discussion of the Models 143 Case 1:3 Case 1: 2 (a) (b) Fig. 5-2. Two examples in which addition of a subunit (arrow) does not change the surface free energy. If we put IX AV = IXb;.v and IX~A = IX' b~A' then bAV and b~A are factors that perturb the reference rate constants that arise from the surface free energy change ,1Gs in the process A ---+ v. The factor e-A.G,/kT is the corresponding perturbation of the equilibrium ratio p~/p~ (which is equal to 1 if ,1Gs = 0); e-A.G,/kT is split between the rate constant perturbation factors bAV and b~A in such a way that e-A.G,/kT = bAv!b~A [Eq. (12.16)]. A practical way to express this will be introduced in Section 13. At equilibrium (a = ae) or steady state (i.e., steady growth or steady shortening of the polymer when a f= ae), the subunit on rate for the particular process A ---+ v is IXAVapA [Eq. (12.14)]. The PA are state probabilities at steady IXAVapA' where the sum is over the state. The total on rate for state A is then s possible states v (i.e., a subunit may be added to the end of anyone of the s strands in state A). If we now sum over all individual surface states A, we obtain the mean subunit on rate for the s-stranded polymer: Iv (12.17) where ri is the mean (operational) on rate constant per strand. Similarly, on summing the subunit off rate IX~APV over v and A, we obtain the mean off rate constant per strand: a:, "L.. V,A ' IXVAPV - = IX"s, IX 1 =- ", L.. IXVAPV' s V,A (12.18) The mean subunit flux per strand is then J(a) = ria - IX'. We shall find in Sections 13 and 14 that, in general, ri and depend on a and hence that J(a) is nonlinear. Except in very simple cases, the steady-state probabilities PA' needed above, cannot be found analytically. Instead we use (Sections 13 and 14) Monte Carlo simulation to obtain J(a) and other properties ofthe polymer end. If, at equilibrium, we sum both sides of the detailed balance Eq. (12.14) over a:
"Surface" Properties of Some Long Multi-stranded Polymers 144 v and A, we obtain (12.19) Thus [Eq. (12.15)], a' ae = - = IX cx~ =-. (12.20) IXe Incidentally, if tip subunits moved from strand to strand rapidly compared to on and off transitions (we are assuming the opposite in this chapter), the equilibrium distribution p~ among the surface states would be maintained even when a "# a e • In this hypothetical case, J = flea - IX~ would be linear. However, in general, the steady-state p,\ depend on a and hence fl and r2 depend on a. The remainder of the chapter is devoted to a number of special cases, including some numerical results. 13. Equilibrium and Steady~State Properties of Aligned Models In this section we consider, at equilibrium and steady state, cases of the type O:s, as illustrated in Fig. 5-1(c). We start with these systems at equilibrium (a = a e ). For s = 2 [also regarded as a tube, as in Fig. 5-3(a)], one strand always ends at m 1 = and the other ends at m 2 = 0, ± 1, ± 2, ... [Fig. 5-3(b)]. The ° s= 2, Case 0:2 5=2 .---+---1- --111 1 = 0 i= 1 i (a) = 1 2 (b) Fig. 5-3. Two-stranded tubular aligned model showing (a) cross-sectional view and (b) a tip structure.
Equilibrium and Steady-State Properties of Aligned Models 145 value of m [Eq. (12.2)] is simply Imzl; that is, m is the height difference between the two strands, measured in subunit heights. The horizontal interaction free energy between two subunits is 2Wh [Fig. 5-3(a)]. This assignment is made in order to include s = 2 in the sequence of tubes s = 3,4, .... For a model with two piles of aligned cubic blocks (not a tube), the properties are the same as derived below if the horizontal interaction free energy between two blocks is called 2wh. All possible surface structures or states are enumerated by mz . The state with m = mz = 0 (flat surface) is the most stable and has a term 1 in Qs [Eq. (12.4)]. The next most stable states are m z = ± 1, each with a term p = e Wh / kT in Qs' etc. The complete Qs is 1+p Qs = 1 + 2(p + pZ + p3 + ... ) = - - . 1-p (13.1) Probabilities of the various m values are [Eq. (12.5)] 1 Po = - , Qs 2pm Pm = Qs (m ~ 1). (13.2) From Eqs. (12.6) and (12.7), the mean and variance are _ 2p m = 1 _ pZ' z 2p(1 + pZ) (Jm = (1 _ pZ)Z . (13.3) The bottom curve in Fig. 5-4 shows iii for this case as a function of - wh/kT. Strong horizontal attractions in the polymer (p ---+ 0) cause Po ---+ 1 and iii ---+ 0 (flat surface); weak horizontal attractions (p ---+ 1, independent strands) lead to iii ---+ 00. Vertical interactions (wJ are not involved in surface roughness. For s = 3 [Fig. 5-1(c)], for m1 == 0 and each ofmz = 0, ±1, ±2, ... , one can sum pm from m3 = -00 to +00. These series are then summed (over mz) to obtain Qs 1 + 4p = + p2 (1 _ p)Z (13.4) An alternative procedure for finding Qs' practical for s = 3, 4, 5 but not beyond, is to express Qs as the trace of a matrix product. This is possible because this is a type of linear nearest-neighbor interaction problem.2 I am indebted to Dr. T. Tsuchiya for confirming Qs in this way for s = 3, 4, 5. The most practical method, however, up to about s = 8, is to enumerate individual surface states by computer, letting m z , m 3 , ... range as far as necessary on either side of zero. The value of m is calculated (in the computer program) for each state [Eq. (12.2)] and tallied, to provide the R(m) values in Eq. (12.4). If R(m) extends accurately to large enough m, Newton's forward interpolation formula applied to successive R(m) values can be used to find R(m) as a polynomial in m. Summation over m [Eq. (12.4)] then gives a closed expression for Qs' The s = 3 case is expecially simple (Newton's formula is not needed):
"Surface" Properties of Some Long Multi-stranded Polymers 146 9 8 7 6 5 IE: 4 3 2 6 10 8 0.001 0.0001 Fig. 5-4. Curves of if! as a function of - wh/kT (p = e wh/kT) for four different cases, all at equilibrium. Qs = 1 + 6p + 12p2 + 18p 3 + 24p4 + ... = 6p 1 + (1 _ p)2' (13.5) which again leads to Eq. (13.4). Probabilities of different m values for s Po 1 = Q: Pm = 3 are [Eqs. (12.5) and (13.5)] 6mp m = Q. (m ~ 1). (13.6) From Eqs. (12.6) and (12.7), _ m= (J 2 m = -------_:_ 6p(1 + p) (1 - p)(1 + 4p + p2) (13.7) 6p(1 + 2p + 6p2 + 2p 3 + p4) (1 _ p)2(1 + 4p + p2)2 (l3.8) ---'----'--_-'---c~-=--------'-~_=_'_--'- The curve m( - wh/kT) is included in Fig. 5-4. The mean height difference between neighboring strands for s = 3 is q = Im21 = 2m/3 [Eq. (12.8)]. For example, at p = 0.5, m = 2.769 [Eq. (13.7)] and q = 1.846. To find the probability distribution in q [Eq. 12.10)], lV(q, m) is needed. These numbers can be obtained in the same computer program
147 Equilibrium and Steady-State Properties of Aligned Models Table 5-1. Initial Values of W(q,m) for the Case 0:3 m= 0 q=O 3 6 12 1 2 3 4 2 3 4 6 12 18 6 12 12 24 6 12 12 12 30 mentioned above by tallying the 3R(m) values of q for each m according to q value. Table 5-1 gives W(q, m) for this case for small values of q and m (the pattern is obvious from this fragment but the computing went well beyond Table 5-1). From Eg. (12.10), then, we obtain 1 _ p2 Po = 1 + 4p P = q 2[(q (13.9) + p2 + l)pq - (q - l)pq+1] (1 1 + 4p + p2 - p) (q ~ 1). (13.10) As a check, one finds Lq Pq = 1. Numerical values of the Pq are given in Table 5-2 for p = 0.5. The most likely neighbor height differences are q = 0, 1, and 2, but convergence in Pq is rather slow at large q. For s = 4, we find by the computer enumeration method described above, Qs = 1 + 12p = 1 + 2p + 42p2 + 92p3 + 162p4 + 252p 5 + ... L 00 (6 k=O 1 + 9p + 9p2 (1 _ p)3 + 10k + 5k 2)pk (13.11) + p3 (13.12) Then, from Eg. (12.6), Table 5-2. Values of Pq for the Case 0:3 at p = 0.5 q Pq q Pq 0 1 2 3 4 5 0.2308 0.3077 0.1923 0.1154 0.0673 0.0385 6 7 8 9 10 11 0.0216 0.0120 0.0066 0.0036 0.0020 0.0011
148 "Surface" Properties of Some Long Multi-stranded Polymers Table 5-3. Initial Values of W(q,m) for the Case 0:4 m= 0 q=O 4 1 2 3 4 5 72 96 168 144 120 120 120 216 192 168 144 168 2 24 24 48 120 96 80 72 48 3 4 5 _ 12p(1 + 3p + p2) (1 - p)(1 + 9p + 9p2 + p3)" (13.13) m=-----,---:------:--'-----::-~'----;;-:- Table 5-3 contains values of W(q, m) for this case (s = 4), obtained by computer (see above). For each m, the total number of q values is 4R(m), where R(m) is the coefficient of pm in Eq. (13.11). The q = 0 row and the diagonal require separate treatment. Otherwise, first differences in the rows are all 48. One finds from Eq. (12.10), + 4p + p2)(1 - p) 1 + 9p + 9p2 + p3 (1 Po = Pq = pq [( Qs q + 1)( + 2) + 6p(q + 2 - (1 - pf q (13.14) qP)J (q ~ 1). (13.15) Lq q Again, P = 1 as required. It will be noticed in Eqs. (13.1), (13.4), and (13.12) (s = 2,3,4) that the coefficients in the numerator of Qs are the squares of binomial coefficients. The above computer enumeration method has been used to confirm this property for s = 5 and 6. A less complete calculation has also been made to confirm the squared binomial coefficients for s = 7 and 8, as far as the central binomial coefficients (symmetry would then account for the remaining coefficients). Based on the special cases s = 2 to 8, the general expression for arbitrary s thus appears to be 1 Qs s-l [ = (1 - py-l k~O (s - I)! k!(s - 1 - k)! J2 k (13.16) p. From Eq. (12.6) we then find for arbitrary s, L2 _ (s - l)sp m = -----'-~= (1 - p) Ll ' where L1 is the sum in Eq. (13.16) and L2 = Sf (s - 1)!(s k=O k!(s - 1 - k)!(k The first few L2 are (13.17) 2)!pk + 1)!(s - . 2 - k)! (13.18)
Equilibrium and Steady-State Properties of Aligned Models s = 2, L2 = 1; s L2 = 1 S L2 = 1 + P + 3p + p2 L2 = 1 + 6p + 6p2 + p3 L2 = 1 + lOp + 20p2 + lO p3 + p4. = 4, S = s = 3, 5, = 6, 149 (13.19) L2 The sum of the coefficients in are the Catalan numbers 1,2,5, 14,42, .... The limit s ~ 00 is of theoretical interest. This refers to an infinitely wide tube with local fluctuations in roughness along the one-dimensional surface of s subunits. To study this case, we can use the familiar maximum term method in statistical mechanics (Ref. 3, Chapter 2). Let the summand in L1 [Eq. (13.16)J be denoted t k • Then 0 In tdok = 0, using s ~ 00, leads to k* = Sp1/2 (13.20) 1 + p1/2 ---:-= as that value of k which maximizes t k. Using only this tk in In Qs we obtain [Eq. (12.6)] Sp1/2 _ sp Sp1/2 (s ~ 00) (13.21) m=--+--~ 1 - P 1 + p1/2 I-p where the term spl(l - p) arises from the factor (1 - p)s-l in Qs. Note that m is an extensive thermodynamic property (prol?()ttional to s). The mean nearest-neighbor height difference along the surface is Ii = 2m - s 2p1/2 = -- 1- p (s -+ 00). (13.22) At P = 0.5, Ii = 2.828. For the same p, Ii = 1.333 for s = 2, 1.846 for s = 3, and 2.095 for s = 4. The variance in m is 2 (Jm dm = P dp = sp1/2(1 + p) 2(1 _ p)2 (s -+ 00). (13.23) Thus (J;;,lm 2 is of order lis ~ 0, which is normal for a fluctuating extensive thermodynamic property. The probability distribution in m is very sharp; indeed, this was already assumed [from the form of Eq. (13.16)] in using the maximum term method. A method for extending some of the above results to finite aligned polymers at equilibrium (a < a e ) is considered in Section 19. Aligned Models at Steady State Because aligned models are rather unrealistic (i.e., they have a simple square lattice), we consider in this section only s = 2 and s = 3. The s = 2 case (also designated 0: 2) can be handled analytically and hence provides a convenient
"Surface" Properties of Some Long Multi-stranded Polymers 150 introduction to steady-state systems. The s = 3 case (i.e., 0: 3) requires Monte Carlo calculations to obtain steady-state properties. The 0: 2 case is illustrated in Fig. 5-3. If we could add one subunit to the end of the polymer without changing Gs (i.e., ~Gs = 0), the reference on and off rate constants iX and iX' would be applicable and the interaction free energy change would be Wv + Wh (bulk polymer increases by one subunit). The equilibrium constant for this process would be K = iX/iX' = 1/ae (Section 5). Actually, for the 0:2 case, this process is hypothetical; it is necessary to add one subunit to each strand in order to obtain ~Gs = O. The equilibrium constant is then (iX/iX')Z and the free energy change is 2(wv + wh ). From either point of view, (13.24) This shows how iX/iX' depends on Wv and Wh ; we are not concerned here with other contributions. Equation (13.24) should be compared with Eq. (4.26) with z = 4, or with Eq. (8.9) (where z = 2). In actual transitions (~Gs =1= 0), there are two categories, as shown in Fig. 5-5. In Fig. 5-5(a), adding a subunit [2 -+ v in Eq. (12.12)] increases m by 1. Thus ~Gs = -Wh and [Eq. (12.16)] (13.25) Similarly, in Fig. 5-5(b), adding a subunit decreases m by 1. Hence, and e- wh / kT 0'1 1 p =- iXz/iX iX z iX 1 a~/a' ' iX~ iX' P ~Gs = Wh (13.26) (Xi It: Q' 1 Q'~ Jt: "- rn (a) .-- "'2 "'2 \d "'2 ffi "'2 I--- I (b) Fig. 5-5. Two different pairs of rate constants, shown in (a) and (b), for the case 0:2. Dotted subunit is added, with rate constants IX! and 1X 2 . Removal of this subunit has rate constants IX'! and IX~.
Equilibrium and Steady-State Properties of Aligned Models 151 2 • Cl' ,fl- 1 P + Cl'ap -f, - Fig. 5-6. Kinetic diagram for case 0: 2 in terms of m values of surface structures. To split AGs =1= 0 between a and a' [see Eq. (12.16)] in a manner consistent with Eqs. (13.25) and (13.26), we introduce the formalism [see Eq. (7.8) and Ref. 2] al = apJI, a'l = a' pJI-l a2 = ap-h, a~ = a'pl-h, (13.27) where 11 and 12 are constants, usually but not necessarily between 0 and 1. The factors, involving p, that modify a and a' in Eqs. (13.27) all arise from surface free energy effects. The most realistic assumption is probably 11 = 12 = O. That is, the on rate constant a is, say, diffusion controlled and is not influenced by neighbor interactions (wh). In this case, a'l = a' p-l and a~ = a' p; the full effect of AGs is felt by the off rate constants (neighbor interactions must be broken for a subunit to escape). Figure 5-6 shows the kinetic diagram, at an arbitrary a, that relates the different m values. The rate constants are unchanged at larger m values. There is only one surface state at m = 0 (flat surface) but there are two surface states (exchange the strands) for each of m = 1, 2, .... The steady-state probability of m is denoted Pm. Because the diagram in Fig. 5-6 is linear, there is a simple "detailed balance" solution 2 for the Pm. One finds 1- z p. - - 0-1+z' p. __ 2zm(1 - z) m 1+z ( m~ 1) (13.28) where z(a) == aapJI + a'pl-h + aap J a' p J1- 1 2 aa = a' p + a'pl-J,-h 1 + aap - J J' I 2 (13.29) The z here is, of course, not to be confused with z as the number of nearest neighbors (Section 4). In the special case a = a e = a'/a (equilibrium), z = p. Equations (13.2) and (13.28) are then the same. Because Pm OC zm, in analogy with Eq. (13.2), 2z m=-1-' -z 2 (13.30) as in Eq. (13.3). The average surface free energy at steady state is, from Eq. (12.1), -Wv - mWh with m given by Eq. (13.30). From Fig. 5-6, we see that the total (both strands) on rate is 2aap J, when
"Surface" Properties of Some Long Multi-stranded Polymers 152 m = 0 and is aapf. + aap-h when m mean on rate constant a(a) is = 1,2, .... Thus, as in Eq. (12.17), the (13.31) Similarly, + p1-h) 1+z rl(zpf.-1 (13.32) Both a and a' depend on a because z is a function of a. The subunit flux per strand is J = aa - ri. At equilibrium (a = ae ), z = p and _ a(pf' + p1-h) 1+P ae = (13.33) a'(pf. + p1-h) a' = -'---------'---- 1+ e P This confirms Eq. (12.20) for this particular case. In the important special case 11 = 12 = 0, Z = + a' p + a' p l ' aa aa _ a = a, ----; a = a'(z p(1 + p2) + z) . (13.34) We might illustrate the involvement of Eq. (13.24) here. Usually a dependence is the feature of main interest, with p = e kT assigned a fixed value. However, if the full effect of variations in W h is to be studied, Eq. (13.24) cannot be ignored. That is, the reference rate constants also depend on Who In this special case (fl = 12 = 0), it is natural to take, in Eq. (13.24), a as a constant and a' = a~p, where a~ is the value of a' when Wh = 0 (we assume that Wv is held constant). When Wh = 0, the two strands are independent. Then Wh / z = a' = + a~p2 _ a= a aa + a~ , aa + p2) + 2a~p2J 2aa + a~(1 + p2) a~[aa(1 (13.35) ----'-"'----'--------'-------oc'-'-- We shall defer numerical results until the case 1: 2 in Section 14. This case (1: 2) is very similar to 0: 2 and is more realistic (in fact, polymerized actin is an example).
Equilibrium and Steady-State Properties of Aligned Models 153 Fig. 5-7. Three types of rate constant pairs for case 0:3. - _ i ~ ......._ . 1 2 3 We turn now to s = 3, that is, case 0:3 [Fig. 5-1(c)] at steady state. There are three categories of rate constants in this case, as illustrated in Fig. 5-7. In the solid part of the figure, m 2 = - 2, m3 = -1, and m = 2 [Eq. (12.2)]. If a subunit is added to strand 1, m increases by 1. As in Eqs. (13.25) and (13.27), (13.36) If a subunit is added to strand 2, m decreases by 1: (13.37) If a subunit is added to strand 3 (to give m 2 = - 2, m3 = 0), there is no change in m; the rate constants are IX and IX'. It is not difficult to construct the kinetic diagram for this case, analogous to Fig. 5-6. However, here the diagram is a two-dimensional array of states associated with the possible values of m 2 = 0, ± 1, ... and of m3 = 0, ± 1, .... Unfortunately, it does not seem possible to deduce the steady-state state probabilities, analytically, from the diagram. We turn, therefore, to the Monte Carlo approach. In all of the Monte Carlo calculations in this section (case 0: 3) and in Section 14 (cases 1: 3 and 2: 5), we take all J; = 0 (diffusion-controlled attachment; all interaction effects in IX'). All on rate constants are then IX and the off rate constants are calculated as follows, in the course of the computer simulation. In the simulation, we follow the stochastic succession of detailed surface states (characterized by the mJ passed through by the end of the polymer. In a particular state of the sequence, a subunit might add to any of
154 "Surface" Properties of Some Long Multi-stranded Polymers the strands (with first-order rate constant aa) or a subunit might be lost from anyone of the strands. The off rate constant must be calculated for each strand. Let a* be the off constant for an arbitrary strand. To find a*, first th€ value of m is calculated from the mi [Eq. (12.2)] in the initial state. This is m(v) in Eq. (12.12). With a subunit removed from the arbitrary strand, m is recalculated from the new set of mi to give m(A). Then Am == m(v) - m(A) is related to AG. in Eq. (12.12) by AG. = - WhAm. Note that A is defined in both cases in the direction of adding a subunit. Then Eq. (12.16) becomes (because the on rate constant is alway a) (13.38) Examples are a'l = a' p-t, a~ = a' p, and a~ = a' in Eqs. (13.36) and (13.37). For any particular surface state in the stochastic sequence of states, there are (for the 0: 3 case) six possible transitions (three on, three off), each with a definite first-order rate constant. The reciprocal of the sum of these six rate constants gives the mean lifetime of the state, which is used in time-averaging various quantities of interest (e.g., m, m 2 , Im21, etc.) over all states in the sequence. The actual transition that is selected to produce the next state in the sequence is determined by a random number generator; the probability of a given transition (one of six) is proportional to the corresponding firstorder rate constant. The mean (time averaged) off rate constant is calculated for each strand; these in turn are averaged over all strands (they differ because of fluctuations) to give ri. Then J = aa - ri. As a check on the program, J is also calculated by the actual counting of individual on and off transitions; this latter J fluctuates more and is therefore less reliable. The number of transitions used to obtain averages varied from 10,000 to 40,000 (in Section 14,50,000 forl:3 and 60,000 for 2:5), depending on the extent of fluctuations in the different cases (values of a and p). Monte Carlo results at a = a e could be compared with exact equilibrium properties, as a further check on the program. All steady-state calculations (in this section and Section 14) were made using the reference values a = 1 and a' = 2 (a e = 2). Because a' = a~p [Eq. 13.35)J, this means in effect that when p was changed a~ was also adjusted to keep a~p constant. This procedure was adopted in order to expose surface effects rather than bulk effects. Figure 5-8 gives Monte Carlo J(a) results for the 0: 3 case at three different p values. The asymptotic straight lines aa - a' (p ~ 1; independent strands) and J = 0 (p ~ 0) are included for reference. J(a) is decidedly nonlinear at p = 0.1 and 0.01 because ri(a) is not a constant. This is contrary to conventional wisdom. Figure 5-9 shows m(a) in the same examples. These are essentially also q(a) curves because q = 2m/3 [Eq. (12.8)]. Surface roughness, as measured by mor q, increases with a because on transitions become more dominant and these occur at random on the three strands. As p ~ 0, m~ 0 (the
Equilibrium and Steady-State Properties of Aligned Models 155 (, S 0:3 1, =12 =0 4 3 J ~ o ~--~~~==~------------~ p=o 3 4 a 5 7 Fig. 5-8. J(a) Monte Carlo results at steady state for 0 : 3case with II = 12 = 0, rx = 1, and rx' = 2. o 0:3 p = 0.5 II =11 =0 iii IJ Fig. 5-9. Results for m(a) corresponding to Fig. 5-8.
156 "Surface" Properties of Some Long Multi-stranded Polymers Table 5-4. Monte Carlo SteadyState Properties for the Case 0: 3 (J;'/m 2 (J;/q2 65.4 16.9 9.92 6.21 3.81 97.9 26.3 14.8 9.53 6.02 p a 0.01 0 2 3.5 5 8 0.1 0 2 3.5 5 8 5.48 1.74 1.12 0.977 0.688 8.40 2.91 1.85 1.60 1.25 0.5 0 2 3.5 5 8 0.731 0.560 0.518 0.464 0.567 1.31 0.964 0.945 0.826 1.05 surface approaches flatness). Of course J and mat a = a e = 2 have equilibrium values [Eq. (13.7) and Fig. 5-4]. Table 5-4 contains Monte Carlo values of (J;,lm 2and (J; Iq2 (the separate m, and hence q = 2m13, values are shown in Fig. 5-9). The large values in Table 5-4 at p = 0.01 and a = 0 arise from m = 0.01522, (J;' = 0.01515, q = 0.01021, and (J; = 0.01021. The probability distributions in m and q were also calculated but are omitted to save space. 14. Equilibrium and Steady-State Properties of Staggered Models We begin, in this section, with a sampling oftubular staggered (helical) models ofthe type shown in Figs. 5-1(a), 5-1(b), and 5-2, at equilibrium. In many cases we merely give the surface partition function Qs; mand (J;' are easy to derive, if desired, from Eqs. (12.6) and (12.7). We give most details about the cases 1: 2 (related to actin), 1: 3, and 2: 5. The fractional stagger in the latter two cases (1/3 and 215) is similar to that in a microtubule (5/13 = 0.385). We begin with the simplest case, 1:2 [Figs. 5-2(a), 5-3(a), and 5-1O(b)]. However, first we digress to point out that the 1: 2 model is formally identical to a realistic model for polymerized actin. In Fig. 5-1O(a), each subunit in the actin-like structure has two interactions Wh with neighbors. The same is true in the 1:2 case [Fig. 5-10(b)] because of the assumed tubular configuration
Equilibrium and Steady-State Properties of Staggered Models ._j AClin Ca . C"s~ I : ~ 157 I:~ :... .: ---II~ ~ +3 ; ES---"' .., II'h -- ---",'0 IIIh II'h II' , III, . (a) ee) (b) Fig.5-10. (a) Model of actin structure, with significance ofw h and W (b) Two-stranded staggered case (1: 2) with neighbor interactions (Wh and wvJ shown. (c) Illustration of n z values for case 1: 2 (n! = 0 always). y • [Fig. 5-3(a)]. Thus, in bulk polymer, both structures have an interaction free energy of W h + Wy per subunit. All of the properties given in this section for the 1: 2 case apply as well to the model of actin in Fig. 5-1O(a). As explained in Section 12, it is convenient to use pm = YJn here, where n = 2m and p = e kT = YJ z. As can be seen from Fig. 5-1O(c), n1 == 0 and nz = ± 1, ±3, .... The surface can never be flat. In Eq. (12.3), n = 2m = Inzl. Hence Wh / 2pl/2 (14.1) 1-p Then, from Eqs. (12.6) and (12.7), _ m l+p p)' = 2(1 _ 2 am p = (1 _ p)z' (14.2) From Eqs. (12.5) and (14.1), the probability of a given n is Pn* = When p -+ 0 and YJ a;' -+ 0, and Pt -+ 2YJn Q. (n = 1,3,5, ... ). (14.3) -+ 0 (i.e., strong attractive interactions, Wh -+ -00), in -+ 1/2, 1. That is, the only important surface structures in this limit are n2 = ± 1 [Fig. 5-10(c)]. Successive subunits go on to or off of alternate strands of the polymer; the polymer behaves like a single helix (I-start helical growth). It is generally assumed that this is the situation in actin. The determining factor here is the magnitude of Wh compared to kT, not Wh compared to Wv (the value of Wv has no influence on surface structure). We consider next case 1: 3, already illustrated in Figs. 5-1 (a), 5-1 (b), and 5-2(b). As mentioned at the beginning of Section 12, this system has the same
"Surface" Properties of Some Long Multi-stranded Polymers 158 properties as 2: 3. We use p = e whl kT = '1 3 and n = 3m (Section 12). Qs can be found analytically by summing pm = '1n over n3 = ... , -4, -1,2, 5, ... for each of n2 = ... , - 5, - 2, 1, 4, ... (with n 1 = 0 in all cases). One finds from these series the closed expression (14.4) Either by expansion of Eq. (14.4) or by computer enumeration, we also have Qs = I n S(n)'1n = 3('1 2 + 2'15 + 3'1 8 + ... + '14 + 2'17 + 3'1 10 + ... ). (14.5) This provides Pn* [Eq. (12.5)], the probability of a given n. The leading term in Eq. (14.5) is 3'1 2, which has n = 2. There are three structures with n = 2. Thus, the smallest possible value of m is 2/3 (i.e., m = n/3). The structures in both Figs. 5-1(a) and 5-2(b) have m = 2/3; the surface cannot be flat. From Eqs. (12.6) and (12.7), m= 2 (Jm = ~~~~~~~~ 2(1 + 2'12 + 2'1 3 + '15) 3(1 - '1 3)(1 + '12) (14.6) 2'12(2 + 9'1 + 14'1 3 + 9'15 + 2'16) ---'---'-----::-9--:-(1--'_-'1-:;C3)-O;2'c-:(1-+-'1";;"2)-0;2----'----'- (14.7) The curve mas a function of - wh/kT is included in Fig. 5-4. It will be recalled [Eq. (12.8)] that r represents anyone of the terms in the sum in Eq. (12.3), for example, In21. This is the height difference between nearest-neighbor strands measured in thirds. The mean value of r is related to m by r = 2m [Eq. (12.8)]. Hence Fig. 5-4 gives essentially r as a function of -wh/kT The probability distribution in r, p~, follows from U(r, n) [Eq. (12.11)], which can be obtained by computer. The U(r, n) table in this case (not shown) is very simple (almost all entries are 0 or 6). From Eq. (12.11), we obtain pT = p~ = 2'12 Qs(1 _ '13) [(r (14.8) + 1)/3] '1'(1 - '1 3) + 2'1'+2 Qs(1 _ '13) [(r - 1)/3]'1'(1 - '1 3) + 2'1'+1 Qs(1 - '1 3 ) (r = 2,5,8, ... ) (r = 4,7,10, ... ) (14.9) (14.10) with Qs given by Eq. (14.4). In the limit p --+ 0 and '1 --+ 0 (strong horizontal interactions), Qs --+ 3'1 2, m--+ 2/3, if --+ 2, r --+ 4/3, Ft --+ 1, pt --+ 2/3, p! --+ 1/3. (14.11) This set of limits corresponds physically to the use of only three surface struc-
159 Equilibrium and Steady-State Properties of Staggered Models tures: ni = 0, 1, 2; ni = 0, 1, -1; and ni = 0, - 2, -1. The first two structures are those in Figs. 5-2(b) and 5-1 (a), respectively. The polymer gains or loses subunits only via a single right-handed I-start helix. In the remaining cases considered in this section, Qs was obtained from Eq. (12.4) after computer tabulation of S(n). Several series had to be summed in each case, some ofthem rather complicated. Details are omitted. Considerable computer time is needed for s ): 8. The cases included here are 1: 4, 2: 4, 1 : 5, 2 :5, 1: 6, 2: 6, 3: 6, and 4: 8. These cases, with 1: 2 and 1: 3 above, comprise a complete set of staggered tubes for s = 2 through 6. For cases of type 1 :s, we already have 1:2 [Eq. (14.1)] and 1:3 [Eq. (14.4)]. In addition, Qs = 41]3(1 + 31]3 + 1]6) (1 _ 1]4)3 _ 51]4(1 Qs - Qs = 61]5(1 (14.12) (1: 4, P = 1]4) + 61]4 + 61]8 + 1]12) (1 _ 1]5)4 . _ 5 (14.13) (1. 5, P - 1] ) + 101]5 + 201]10 + 101]15 + 1]20) (1 _ 1]6)5 (1: 6, P = 1]6). (14.14) The expressions for in and (J;' are easy to derive [Eqs. (12.6) and (12.7)]. The coefficients in the Qs sums are the same as in Eqs. (13.18) and (13.19). In the limit P ~ 0, 1] ~ 0 (strong horizontal interactions), Qs ~ S1]S-1 for the general case 1: s. The surface has regular steps as shown in Fig. 5-2(b) (for 1 : 3). The lowest step can be at any strand; hence the degeneracy factor s in Qs. With regular steps, it is easy to see that n = s - 1. In order words, in this limit for any s, the polymer gains and loses subunits only via a single righthanded I-start helix. As one would expect, it is possible, using a modification of Eq. (13.18) in Qs for arbitrary s, to show that Eq. (13.21) holds for in here in the limit s ~ 00. Another connected series of cases is 1: 2, 2: 4, 3: 6, and 4: 8. The Qs for 1: 2 is given in Eq. (14.1). For the other cases, Qs= Qs= 41]9(5 21]4(3 + 41]2 + 31]4) (1-1]4)3 + 151]3 + 231]6 + 151]9 + 51]12) (1-1]6)5 21]16(35 (14.15) (2:4,p=1]4) (3:6,p = (14.16) 1]6) + 1681]4 + 3991]8 + 5121]12 + 3991]16 + 1681]20 + 351]24) Qs= (1-1]8)7 (4:8,p = 1]8). (14.17) The remaining two examples studied are 2: 5 and 2: 6. In the latter case, 31]8(5 Qs = + 21]2 + 201]4 + 151]6 + 151]8 + 201]10 + 21]12 + 51]14) (1 _ 1]6)5 (2:6,p = 1]6). (14.18)
"Surface" Properties of Some Long Multi-stranded Polymers 160 In the former case, which will also be considered at steady state (below), Qs = 5rJ6(2 + rJ2 + 4rJ3 + 4rJ5 + rJ6 + 2rJ8) (1 _ rJ5)4 (2:5,p = rJ5). (14.19) From Eq. (12.6) for 2: 5, ~ 4(3 + 2rJ2 + 9rJ3 + 18rJ5 + 3rJ6 + 3rJ7 + 18rJ8 + 9rJl0 + 2rJll + 3y 13) 5(1 - rJ5)(2 + rJ2 + 4rJ3 + 4rJ5 + rJ6 + 2rJ8) m=~--~----~--~----~--~--~~~~~~~--~~ (14.20) In the limit p --+ 0, rJ --+ 0, we have m--+ 6/5. Figure 5-4 includes mas a function of -wh/kT From a rather complicated U(r, n) table (in the 2: 5 case), obtained by computer, we find for the [Eq. (12.11)], P: p! 2rJ6(3 = * _ ~{k(k + l)(k + 2)rJr Pr - Qs + 12 rJ6[(k + l)(k + 2rJ2 + 2rJ3 + 3rJ5) Qs(l _ p)3 2rJr+9 - p)3 + (1 + 2)pk(1 + * _ ~{j(j + l)(j + 2)rJr Pr - Qs + 12 rJ9[(j 2rJr+6 + (1 _ p)3 + + 1)(j + 2)pi(1 3rJr+6[2 + k(l - p)] (1 _ p)3 - p)2 + 2pk+l<k (1 _ p)3 k +2- (k + l)P)]} (14.22) = (r - 3)/5, r = 3,8,13, ... 3rJr+4[2 + j(l - p)] (1 _ p)3 - p)2 + 2pi+1<j (1 _ p)3 ° (14.21) +2- (j + I)P)]} (14.23) j = (r - 7)/5, r = 7,12,17, ... When p --+ and rJ --+ 0, Qs --+ lOrJ6. There are ten different surface structures with the five r values 2, 2, 2, 3, 3. These r values, on summing, give n = 12/2 = 6. Also, , = 12/5, m = ,/2 = 6/5, p! = 3/5, and pj = 2/5. Two of the ten surface structures are shown in Fig. 5-11 (anyone of the five strands can have the lowest position). Staggered Models at Steady State We consider three cases, 1: 2, 1: 3, and 2: 5. The first can be handled analytically. The other two require Monte Carlo calculations. As already mentioned, 1: 2 is equivalent to polymeric actin whereas 1: 3 and 2: 5 resemble microtubules (5: 13) in the fractional extent of stagger (1/3, 2/5, 5/13). The algebra in the 1: 2 case is similar to that for 0: 2 (Section 13). Figure
Equilibrium and Steady-State Properties of Staggered Models r = :2 i ~ t 1 2 3 3 161 3 r~ 2 3 t 4 i ~ 1 2 3 2 3 t ~ 4 (b) (a) Fig. 5-11. The two types of most stable surface structure for the case 2: 5. In (a), r = 2,2, 2, 3, 3; in (b) r = 2,2, 3, 2, 3. In both examples i = 1 is the lowest strand but any of the others can be. Arrows show subunit additions or departures that leave the surface free energy unchanged. 5-12 shows the three sets of rate constants that arise for 1 :2. In Fig. 5-12(a), there is no change in surface free energy. Hence the rate constants are the reference constants (X and (X'. In Fig. 5-12(b), Lln = 2 and Llm = 1 (on addition ofa subunit). In Fig. 5-12(c), Lln = -2and Llm = -1. Therefore, Eqs. (13.25)(13.27) all apply to the present model. The kinetic diagram in terms ofn values is shown in Fig. 5-13. The rate constants are unchanged between pairs of successive states in the linear diagram. Each state (n value) is doubly degenerate (exchange strands). The rate constants (X and (x' apply between the two substates of n = 1 [Fig. 5-12(a)], but they do not appear in Fig. 5-13 (however, ex ex' (a) ill (b) (e) Fig. 5-12. Three types of rate constant pairs for case 1: 2.
"Surface" Properties of Some Long Multi-stranded Polymers 162 aa/ I + a'p I f l 11=1 .. .. 3 5 • a'/l-l+aap-h p') p') p'I ... Fig. 5-13. Kinetic diagram for case 1 : 2 in terms of n values of surface structures. Rate constants for 3 +:i: 5, etc., are the same as those for 1 +:i: 3. see Ci and rt.' below). Because the diagram is linear, there is a "detailed balance" solution at steady state (as for Fig. 5-6). The successive states n = 1, 3, 5, ... have relative probabilities 1, z, Z2, ... , where z is the same as in Eq. (13.29). Hence the normalized probabilities are (n = 1,3,5, ... ). Pn* = z(n-1)/2(1 - z) (14.24) At equilibrium (eta = a'), z = p = 1'/2, thus recovering Eq. (14.3). From Eq. (14.24), using n = 2m, we find m= 1+z 2(1 _ z)' z (1 _ Z)2' 2 (Jm = (14.25) These are generalizations of Eqs. (14.2). The total subunit on rate for both strands is aa + aapf' when n = 1 and is aapf' + aap-h for all other values of n. The corresponding off rates are a' + a' p1-h for n = 1 and a' pI. -1 + a' p1-h otherwise. Using Pi = 1 - z and 1 - Pt = z, we find for the mean rate constants per strand Ci(a) = (a/2)[(1 - z)(l a'(a) = (a' /2) [(1 - z)(l + + pf') + p1-h) z(pf, + + p-h)] (14.26) + (14.27) z(pfl-1 p1-h)]. These then give J = Cia - a', the mean subunit flux per strand. At equilibrium (z = p), (14.28) a' = (a'/2)(1 ° + pf' _ P + p1- h ). ° (14.29) ° Thus Ci/a' = a/a' [Eq. (12.20)]. When p -+ and!l and!2 are between and 1, z -+ and Ci -+ a/2, a' -+ a'/2. The polymer behaves as ifit has only one strand (n2 = ± 1); subunits alternate strands in their arrivals and departures, using the reference rate constants a and a' [Fig. 5-12(a)]. In the important special case!l = !2 = 0, z is given in Eq. (13.34) and Ci = a, a' = (a'/2)(1 - p - z + Zp-1). (14.30) To illustrate some of the above results, we plot J(a) in Fig. 5-14 and mea) in Fig. 5-15 for a = 1, a' = 2,11 = 12 = 0, and p = 0.5,0.1, and 0.01. The J(a)
163 Equilibrium and Steady-State Properties of Staggered Models 11./1 0.1 0.0.5 6 5 4 .... 0.0:0 . . 0.5 1.1 0.5.0 1.0 1.2 11 = h = 0 J C o ~----~~------------------------ 5 7 " Fig. 5-14. Curves are analytical results for lea) in case 1: 2 with 11 = 12 = 0, IX = 1, and IX' = 2. Short arrows at a = 8 show values of 1 at a = 8 for p = 0.5 but with different pairs 11, 12 . curve in Fig. 5-14 for p = 0.1 is noticeably nonlinear. The p = 0.01 curve is close to the asymptotic line discussed above (p --+ 0), J = (eta - a')/2. Also, the p = 0.5 curve is not far from J = exa - ex' (for p --+ 1; independent strands). The behavior of mea) in Fig. 5-15 is similar to that already seen in Fig. 5-9 (0: 3) except that m = 1/2 is the lower limit (the surface cannot be flat). The mea) curves are linear in a [Eqs. (13.34) and (14.25)]. The mea) curves in the other examples in this chapter (Figs. 5-9, 5-17, and 5-19), all for 11 = 12 = 0, are also at least approximately linear. In the 0: 2 case [Eqs. (13.30) and (13.34)], mea) is also approximately linear, but it is not exactly linear. Although most choices are very unrealistic, the effect of varying 11 and 12 is also included in Figs. 5-14 and 5-15 for the case p = 0.5. In Fig. 5-14, we merely show (short arrows) the value of J at a = 8 for p = 0.5 and various pairs 11' 12' Of course, all curves (not shown) pass through J = 0, a = 2. In Fig. 5-15, the complete mea) curves are given for p = 0.5 and the same set of 11 J2 choices. In the 1, 1 case, surface free energy effects are confined entirely to the on rate constants; the off rate constant is always a' [see Eqs. (13.27)]. Hence, when a is small and random off transitions dominate, fluctuations in the strand that loses the subunit can lead to large values of m. The equilibrium value of iii (i.e., at a = 2) is not influenced by the kinetic parameters 11,12'
"Surface" Properties of Some Long Multi-stranded Polymers 164 I ! I , i 0.0 ! .I \ I \ i \ \ -- -- ------ -- 0.0.5 _ - - - - 0.5.0 o .0. 0.1 1.0 - -- ' - '- ' - ' - ' - · - 1.1 p=o .l (, " Fig. 5-15. Results for m(a) corresponding to Fig. 5-14 (solid curves'!1 =!2 = 0). The non-solid curves are for p = 0.5 and different pairs!I'!2' (> p = 0.5 5 ..j 1.3 II =h = 0 J 2 O~----~~~------------------------- 5 7 a Fig.5-16. J(a) Monte Carlo results at steady state for 1: 3 case with!1 =!2 = 0, er; = 1, and er;' = 2.
Equilibrium and Steady-State Properties of Staggered Models 165 Fig. 5-17. Results for mea) corresponding to Fig. 5-16. p = 0.001 3 4 5 " a Finally, we give Monte Carlo results for 1: 3 and 2: 5, taking all i = 0, rt = 1, and rt' = 2. The discussion of the Monte Carlo approach before and after Eq. (13.38) applies here as welL Figures 5-16 and 5-17 show lea) and mea), respectively, for 1: 3. The lea) curve for p = 0.1 is very nonlinear. The asymptotic lea) lines are rta - rt' (for p -+ 1) and (rta - rt')!3 (for p -+ 0). In the latter limit, growth is via a single right-handed helix [Eq. (14.11)]. Figures 5-18 and 5-19 gives lea) and mea) for 2: 5. Distinctly nonlinear lea) curves are found for p = 0.1 and 0.01. The limiting lea) line for p -+ 0 is not (2/5)(rta - rt') as one might expect, but (3/10)(rta - rt'). The reason for this is the following. When p -+ 0, only ten surface structures are used [Qs -+ 10'1 6 in Eq. (14.19)], all with minimal surface free energy (n = 6). Five are of the type in Fig. 5-11 (a) and five of the type in Fig. 5-11 (b). In Fig. 5-11 (b), there are two addition and two departure sites that leave the surface structure within the class of ten (i.e., still with n = 6). However, in Fig. 5-11 (a), there is only
"Surface" Properties of Some Long Multi-stranded Polymers 166 7 6 2:5 5 1.=11=0 4 J 3 3 4 6 7 a Fig.5-18. J(a) Monte Carlo results at steady state for 2: 5 case with fl = f2 = 0, rx = 1, and rx' = 2. one such site of each type. Hence 2-helix growth cannot be maintained at the nominal rate (2/5)(lXa - IX'); 2/5 and 1/5 are averaged to give the factor 3/10 above. A related problem has been discussed for microtubules. 4 • 5 Tables 5-5 and 5-6 give Monte Carlo values of O-;;.!if!2 and(Ji /q2 (as in Table 5-4) for 1: 3 and 2: 5, respectively. The separate if! values (also, q = 2if!/s) are in Figs. 5-17 and 5-19. These tables differ qualitatively from Table 5-4 at small p because if! -40 and q -4 0 as p -40 in the Table 5-4 case (0: 3) but in Tables 5-5 and 5-6, if! and q are always finite. Note that because r = sq and (Jr2 = S2 (Ji,
Models with Dimers as Subunits 167 Fig. 5-19. Results for m(a) corresponding to Fig. 5-18. 14 12 10 III 6 p = 0.01 p = 0.0001 p= o J 4 5 6 7 a u; u; q2 ,2' (14.31) Conclusion: consideration of fluctuating microscopic surface structures and microscopic rate constants in a class of multi-stranded "equilibrium" polymers shows that the assumption of a linear J(a) curve is not justified except in special cases. The corresponding J(c) curve would have a second source of nonlinearity if a -=1= c [Eq. (5.18)].
168 "Surface" Properties of Some Long Multi-stranded Polymers Table 5-5. Monte Carlo Steady-State Properties for the Case 1: 3 p a (J;,/m 2 (Ji /([2 0.001 0 2 3.5 5 8 0.000342 0.0138 0.0251 0.0368 0.0681 0.D1 0 2 3.5 5 8 0.1 0.5 Table 5-6. Monte Carlo Steady-State Properties for the Case 2: 5 P a (J;,/m 2 (Ji/([2 0.125 0.142 0.157 0.169 0.222 0.001 0 2 3.5 5 8 0.000133 0.00338 0.00664 0.00905 0.0192 0.0422 0.0585 0.0719 0.0874 0.109 0.00620 0.0752 0.125 0.165 0.258 0.133 0.228 0.296 0.347 0.501 0.D1 0 2 3.5 5 8 0.00963 0.0541 0.0792 0.111 0.157 0.0755 0.229 0.327 0.413 0.537 0 2 3.5 5 8 0.111 0.307 0.357 0.407 0.403 0.278 0.564 0.632 0.708 0.738 0.1 0 2 3.5 5 8 0.0774 0.156 0.190 0.225 0.232 0.300 0.546 0.699 0.758 0.808 0 2 3.5 5 8 0.490 0.445 0.489 0.501 0.482 0.860 0.796 0.866 0.871 0.895 0.5 0 2 3.5 5 8 0.216 0.246 0.235 0.247 0.252 0.759 0.801 0.785 0.795 0.997 15. Models with Dimers as Subunits This section is a generalization of Section 14 in which, mimicking micro tubules, we assume that the subunits of the polymers are dimers. This leads to more complicated horizontal neighbor interactions. The treatment we give is rather brief. We consider structures 1 : 3 and 2: 5 only. The two components of a dimer are denoted a and /3 in Fig. 5-20. It is assumed that subunit neighbor interactions in the polymer are accounted for by aa, a/3, /3a, and /3/3 interactions, as indicated by the labeled lines in the figure. These lines show all of the neighbor interactions in which the central subunit is engaged. Lines v, 1, ... represent interaction free energies W Wi' •.•. Because interactions 1, 2, and 3, on either side of the central subunit, always occur as a group, we define Wd == Wi + W2 + w3 • Then the three interactions we have to contend with are Wv (as in Section 12), Wd , and W 4 . The interaction free energy per subunit in the bulk polymer is Wv + W4 + Wd . What we called Wh in Section 12 is W4 + Wd here: V, (15.1)
Models with Dimers as Subunits 169 Fig. 5-20. Part of a staggered polymer with dimers (ap) as subunits. 1 refers to WI' etc. Wd is defined as Ci WI + W2 + W3 · Ci .6 Ci {3 Ci {3 {3 Ci Ci Ci {3 Ci {3 {3 Whereas, in Sections 12-14, we had to count missing interactions Wh (or fractions thereof) at the tip of a polymer to obtain the surface free energy Gs [Eq. (12.1)], here we have to count missing interactions W4 and Wd separately. We turn now specifically to the case 1 :3, at equilibrium to begin with. As shown in Fig. 5-1(a), there are three n i values: n 1 = 0, n 2 , and n 3 . We denote the possible values of ni +1 - ni (for i = 1,2, 3, where i + 1 == 1 when i = 3) by Ui' Possible values of the U i are ... , - 5, - 2, + 1, + 4, .... For example, in Fig. 5-1(a), the three values are U 1 = + 1, u2 = - 2, and U 3 = + 1. Comparison with Fig. 5-20 shows that when U i = + 1, the missing horizontal interaction between strands i and i + 1 is W 4 • Similarly, when U i = - 2, the missing interaction between strands i and i + 1 is Wd . It is easy to see by drawing the structures that for the sequence Ui = + 1, + 4, + 7, ... , the missing interactions are W4 , 2W4 + Wd , 3w4 + 2wd , ••• , and for the sequence U i = - 2, - 5, - 8, ... , the missing interactions are Wd , W4 + 2wd , 2W4 + 3wd , •... The total of missing horizontal interactions for any complete surface structure is obtained by adding the separate missing interactions associated with Ul> U 2 , and U 3 . For the structure in Fig. 5-1(a), with ni = 0,1, -1 and U i = 1, -2, 1, this total is W4 + Wd + W4 = 2W4 + wd. In fact, this is the minimum possible total (it also occurs with ni = 0, 1,2 and 0, - 2, - 1). The corresponding horizontal contribution to - Gs [Eq. (12.1)] is half ofthis: W4 + (Wd/2). Because of this result, we shall count, usingj and k, in units ofw4 and Wd/2, respectively, and define (15.2) Then Qs = I i,k Q(j, k)cpiljJk is the generalization of Eq. (12.4), where j = 1, 2, 3, ... , k (15.3) = 1, 3, 5, ... , and QU, k) is the number of structures with givenj and k. Gs for a given structure
"Surface" Properties of Some Long Multi-stranded Polymers 170 with j and k is (Wd) 2' . - k G = - -3 W -]W 2 s Ui 4 v (15.4) The formal connections between j and k for a given structure and its three values are (15.5} Because there are only three strands, it is not difficult to enumerate aU possible states, as in Eg. (14.4). From such an enumeration, we find n = 3,6, 9, ... for j, k = 1, 1; 2, 3; 3, 5; ... and also the same set of n values for j, k = 1, 3; 2, 5; 3, 7; .... All other values of n are zero. Equation (15.3) then gives Qs ~ 3cpl/l(1 + 1/1 2) (1 _ cpl/l2)2 (15.6) Qs diverges when p ~ 1, as in Eg. (14.4). It is conceivable (e.g., electrostatic repulsion) that cp > 1 or 1/1 > 1, but cpl/l2 < 1. As a check on self-consistency, we note that the 1: 3 case in Section 14 corresponds here to the choices (15.7) Substitution of cp = 1/1 = '1 in Eg. (15.6) produces Eg. (14.4), as expected. Because of the form of Eq. (15.3), it follows that -;- aIn Qs ] ='Cpaq;-' 2 (Jj = (15.8) oj cp acp , We then have, from Eg. (15.6), -;- 1 + cpl/l2 1+P --- 1 - cpl/l2 - 1 - p (15.9) ]- _ k = 2 _ (J. J 1 + 31/12 + 3cpl/l2 + cpl/l4 (1 - cpl/l2)(1 + 1/1 2) ---'---~,----'------,=--- 2cpl/l2 2p (1 _ cpl/l2)2 (1 _ p)2 -,-----'--'--c-c~ (15.10) (15.11) 41/1 2_ (1 _ +_ 2cp_ +_ 2cpl/l2 + 2cpl/l4 + cp21/1 4) (J2- _ ~--~=---_ k(1 _ cpl/l2)2(1 + 1/1 2)2 (15.12) The mean surface free energy is then -(Wd) 3 -;G=--w-]w-ks 2 v 4 2' (15.13)
Models with Dimers as Subunits 171 Next, we give some Monte Carlo results for 1: 3 at steady state. We take all !; = 0, a = 1, and a' = 2, as before. As a reasonable guess, we take the relative magnitudes of WI' ... , W4 (Fig. 5-20) to be 0.75, 1.0,0.75, 1.0, respectively. This gives Wd = 2.5w4 and !/J = e wd/ 2kT <p = p2f7, !/J = <p5/4 (15.14) = p5/I4. Thus, each choice of p determines values of <p and !/J. Table 5-7 presents numerical results for three values of p and of a. Exact values from Eqs. (15.9)-(15.12) are given for a = a e = 2; these provide a check on the Monte Carlo program. The surface roughness, without regard to the distinction between W4 and Wd , is shown by the values of q [Eq. (12.8)] and in the last two columns of the table. The flux values (J) are virtually the same as in Fig. 5-16, at the same p and a. We consider now the case 2: 5 at equilibrium. Possible Ui values are ... , - 8, -3, +2, +7, .... For example, the structures in Fig. 5-11 (with minimum surface free energy) use U i = - 3 and + 2 only. Drawing a few possible 2: 5 structures shows that the missing interactions for the sequence U i = + 2, + 7, ... are W4' 2W4 + Wd , 3w4 + 2wd , ••• , and for the sequence U i = - 3, - 8, ... , the missing interactions are Wd , W4 + 2wd , 2W4 + 3wd , ..•• The minimum missing interactions for a complete structure (Fig. 5-11) are 3w4 + 2Wd' The corresponding contribution to - Gs is then 3(w4 /2) + Wd • Thus the simplest integers j and k in this case are obtained if we define (J; (15.15) These definitions are not the same as in Eq. (15.2). The surface partition function is then Qs = I i,k (15.16) QU, k)cpi!/Jk, Table 5-7. Monte Carlo Steady-State Properties for the Case 1: 3 (Dimers) p a 0.001 0 2 8 1.0002 1.0020a 1.016 0.000155 0.02004 a 0.01617 0.1 0 2 8 1.040 1.2222a 1.952 0.04116 0.2469 a 1.436 0.5 0 2 8 1.732 3.000a 7.200 j (J2 ) 1.032 4.000a 23.28 a Exact equilibrium values. k k q 1.0004 1.0183 a 1.086 0.000857 0.03639 a 0.1705 0.4444 0.4506 0.4667 0.02476 0.02846 0.04393 1.164 1.768 a 3.518 0.3222 1.530a 6.608 0.4886 0.6541 1.215 0.06585 0.2371 1.211 3.006 5.7574a 14.27 (J2 4.810 16.941a 93.6 1.066 1.901 4.698 (J2 q 0.8726 3.020 18.08
"Surface" Properties of Some Long Multi-stranded Polymers 172 Table 5-8. Initial values of Q(j, k) for the Case 2: 5 (Dimers) k= j=3 5 7 9 11 10 5 2 3 20 60 20 5 90 180 50 4 5 6 7 20 240 400 100 50 500 750 100 900 175 where j = 3, 5, 7, ... and k = 1, 2, 3, .... Other subsidiary relations for this case are (15.17) j = ± lUi + 31, i=l 5 k = ± lUi - i=l 10 21. (15.1S) The values off.! have been found by computer enumeration (Table 5-S). The diagonal series in this table can be summed using variations on Eq. (9.33). The result is (15.19) The 2: 5 case in Section 14 is the special case here of <p = pl/5 = 1], 1/1 = p3/5 = 1]3. (15.20) Substitution of <p = 1] and 1/1 = 1]3 in Eq. (15.19) recovers Eq. (14.19). Equations (15.8) and (15.19) lead to j 6 + 5cp2 + 121jJ + 31jJ2 + 30cp21jJ + 30cp21jJ2 + 3cp41jJ + 12cp41jJ2 + 5cp21jJ3 + 6cp41jJ3 (1- cp21jJ)(2 + cp2 +41jJ +4cp21jJ + 1jJ2 + 2cp21jJ2) (15.21) The variances can be found from these expressions [Eqs. (15.S)], but are not given here. Steady-state calculations have been made for J; = 0, IX = 1, and IX' = 2. The relative values of Wi' ... , W4 were taken as 0.5, 1.0,0.5, 1.0, respectively. Thus Wd = 2W4 and <p = eW4/2kT <p = pl/6, = 1/1 1/4 1/1 = p2/3. (15.23)
Models with Dimers as Subunits 173 Table 5-9. Monte Carlo Steady-State Properties for the Case 2:5 (Dimers) p a 0.01 0 2 8 3.060 3.291 a 4.070 0.1175 0.5578 2.050 0.1 0 2 8 3.424 4.364" 7.554 0.7880 2.872 12.708 0.5 0 2 8 6.687 11.821 " 27.78 j (J2 J 9.431 34.67 182.3 k q 1.0231 1.1353" 1.602 0.02303 0.1317 0.5723 0.4906 0.5342 0.7137 0.02141 0.07200 0.2831 1.230 1.8221 " 3.591 0.2166 0.8303 3.393 0.5663 0.8095 1.481 0.1010 0.3957 1.856 k 3.135 5.7882" 13.81 (J2 2.479 8.63 45.50 1.309 2.083 6.154 (J2 q 1.378 3.451 33.81 "Exact equilibrium values. Numerical results are given in Table 5-9. Exact values of J and k from Eqs. (15.21) and (15.22) are included for a = ae = 2. These values provided a check on the Monte Carlo program. The flux values (J) are essentially unchanged from those in Fig. 5-18. Incidentally, the minimum possible value of q is 1'/5 = 0.48 [see the paragraph following Eq. (14.23)]. Recent work 6 suggests that the conventional 5-start helix model of a microtubule 7, used in this book, may need modification. References 1. Hill, T.L. (1986) Biophys. J. 49, 1017. 2. Hill, T.L. (1985) Cooperativity Theory in Biochemistry (Springer, New York). 3. Hill, T.L. (1960) Introduction to Statistical Thermodynamics (Addison-Wesley, Reading, MA; also Dover, New York, 1986). 4. Hill, T.L. and Kirschner, M.W. (1982) Int. Rev. Cytol. 78, 1. 5. Chen, Y. and Hill, T.L. (1985) Proc. Natl. Acad. Sci. USA 82,1131. 6. Mandelkow, E.-M., Schultheiss, R., Rapp, R., Miiller, M. and Mandelkow, E. (1986) J. Cell BioI. 102, 1067. 7. Amos, L.A. and Klug, A. (1974) J. Cell Sci. 14, 523.
6 Sotne Attached Multi-Stranded Polytners at Equilibriulll and in Transients Section 10 in Chapter 4 is concerned with a two-component single-stranded polymer: the two components are mixed together in the same strand (see also Section 27). Most of this chapter (Sections 16, 17, and 18) deals with a somewhat related problem: 1 two different components grow separate pure strands from a surface, but the strands are adjacent to each other and interact laterally so that there is some degree of cooperativity between them. This problem has a certain intrinsic interest in relation to self-assembly of various structures. In addition, the possibility of a "vernier effect" (Ref. 2, p. 126) can be studied using this model; this is the subject of Sections 16 and 17. A method introduced in Section 16 for the formulation of completely open partition functions, and extended in Sections 17 and 18, is also applicable 1 to some of the simpler models in Chapter 5. The new method allows a more general treatment than that given in Chapter 5. This topic is the subject of Section 19. 16. Simple Dual Aggregation and the Vernier Effect A well-known problem in biophysics is the fact that some linear biopolymers seem to know when to stop growing. One suggestion, in the case of side-byside dual linear aggregation of two components of different unit lengths, is that a vernier effect (Ref. 2, p. 126) may operate to terminate the growth. That is, growth will cease, because of extr~. stability, when the two adjacent aggregates happen to achieve exactly the same length. This section and the following one consider some aspects of this problem. To anticipate: it is found
Simple Dual Aggregation and the Vernier Effect 175 1=3/2 ~ K':x '" / \ '\.1)2 WI , / '\ \ 11=1 I I 1 1 I 1 2 I\LI___----'I NZ = 3 IV Fig.6-1. Illustration of a simple dual polymer with 1= 3/2. The interaction free energy w (see shaded area) refers to the interaction of one molecule of type 2 with the neighboring one and one-halftype 1 molecules. The first vernier package is completed at n = 1. that a simple vernier effect is probably inadequate for the purpose mentioned; some special source of extra stability needs to be involved at the "vernier length." In this model, two species of monomers, 1 and 2, at activities a 1 and a 2 , form a side-by-side linear aggregate at a nucleating site on a surface, as shown in Fig. 6-1. The molecules (or subunits) usually have different lengths: 12 = 1 (2 is assigned unit length and is always longer, if the lengths are unequal) and 11 ~ 1. We always take 11 to be a ratio of integers, i2/i1 (11 = 2/3 in Fig. 6-1). The ratio 12/11 = idi2 is denoted 1 and is used to designate the different cases (Fig. 6-1 is the case 1 = 3/2). The number of molecules in strand 1 is N1 and the number in strand 2 is N 2 • The two strands have the same length if N 1 11 = N 2 12 , that is, if N 1 : N z = i1: i2, 2i1 : 2i z , etc. It is assumed that the two strands are continuous: there are no subunit vacancies. The strands might wrap around each other in helical fashion, but this is not shown. The interaction free energy (Fig. 6-1) between nearest neighbors in strandj is Wj; the interaction free energy with the surface is wi (j = 1,2). The lateral interaction free energy between the two strands is assumed to be simply proportional to the amount of overlap. This is measured in units of length of component 2, with a free energy W per unit, and is expressed as Nw. Thus, in Fig. 6-1, N = 8/3. With this definition, N is the lesser of N2 and Ndl. It is convenient to define y == e- w /kT and Yo == y1/1. Of course Wj' wi, and ware all negative, and y ;?! Yo ;?! 1. In the special case w = 0 and y = Yo = 1, the two strands aggregate independently of each other. A large value of y tends to favor vernier structures (3:2, 6:4, etc., in Fig. 6-1) because in these structures there are no missing lateral interactions. A value y = 1000 corresponds to w = - 4.1 kcal mol- 1 at 25°C. To avoid extra and unimportant (for present purposes) algebra, we shall take wi = Wj (j = 1,2), as if the nucleating site on the surface already contains a molecule of each type (or the equivalent) on which to build.
176 Some Attached Multi-stranded Polymers at Equilibrium and in Transients Partition Function for an Open System We begin by considering the dual polymer when it is in equilibrium with monomers at activities a 1 and a2 such that the polymer has a finite, though possibly large, mean size. To introduce notation, let us review first the one-component special case a2 = O. The reader will find the necessary background in Eqs. (4.1), (4.22)(4.26), and (5.1)-(5.13) (where C = 1). The canonical partition function for strand 1 with a given size N1 is (16.1) where q1 is the partition function of a single subunit in the polymer (ignoring possible minor end effects). Strand 1, in the presence of monomers at a 1 < aL is an open system with a finite mean size. The appropriate partition function for the open system is 1 = L <Xl QNJf' = N,=O en L (q 1 e-w,/kT A1 )N, N 1 =O (16.2) where x = q e-w,/kT A = ~ 1 - 1 1 a~ . (16.3) That is, Xl is proportional to a 1 . Analogous equations apply to strand 2, when a 1 = o. The dual polymer of definite size N 1 , N2 has the canonical partition function (16.4) where (see above) N = N2 if N1 > N21 (strand 1 is longer) or N = Ndl if N21> N1 (strand 2 is longer). In the former case yN = yN2; in the latter case, yN = y:'. The factor yN arises from the free energy of interaction Nw between the two strands. The partition function for the open dual polymer, in contact with monomers at a 1 , a2 , is 1= where X2 is defined in the same fashion as a dual polymer of size N 1 , N2 is (16.5) Xl in Eq. (16.3). The probability of (16.6) We shall see below that PN ,N2 has local peaks in cases of interest at the vernier structures N 1: N2 = i 1 :i2, 2i 1: 2i 2, ....
Simple Dual Aggregation and the Vernier Effect 177 The simplest way to find a closed expression for Y is the following. The terms in Eq. (16.5) are collected into three groups, according to whether the two strands are equal in length (vernier structures), strand 2 is longer, or strand 1 is longer. The three groups produce the following respective contributions to Y: (16.7) where II = X~l X~2 i2 [this is the partition function term ofthe smallest vernier package in Eq. (16.5)], R 1(X 1Yo) is an always positive polynomial in X1Yo of degree i1 - 1 with coefficients that are powers of X 2, and R 2(X 2Y) is an always positive polynomial in X 2Y of degree i2 - 1 with coefficients that are powers of x l ' An example will be given below. However, the general form of Eq. (16.7) suffices to show that Y diverges (the polymer becomes very large) if Xl --+ 1 or x 2 --+ 1 or II --+ 1. The case of primary interest in the present context is Xl « 1, X 2 « 1, Y » 1, and II --+ 1. That is, when Y is large (strong attractive interaction between the two strands), bulk dual polymer is formed (II --+ 1) even though both Xl and X 2 « 1. In other words, the aggregation is cooperative. It is easy to confirm, using Eq. (16.4), that the equilibrium condition (16.8) between bulk dual polymer and monomers, is equivalent to II = 1. More generally (i.e., II ~ 1), if X == i1fl1 + i2fl2 - flo is the thermodynamic force driving monomers to aggregate, then II = e X / kT . At equilibrium, X = 0. Cooperative aggregation, arising from lateral interactions, can also occur in one-component polymers [see the discussion of Fig. 1-3 and of Eq. (4.26)]. For example, a bundle of laterally touching actin filaments would aggregate at a lower critical concentration than would a single actin filament. Let us now use the case 1 = 4/3 to illustrate Eq. (16.7). In this example, II = xix~y3. The first term in Eg. (16.7) arises from all vernier terms in Eq. (16.5): 1 + II + II2 + .... The second term (strand 2 is longer) in Eq. (16.7) represents (a) a vernier structure of any size plus (b) 0, 1, 2, or 3 additional molecules of type 1, together with 1, 1,2, or 3 molecules of type 2, respectively (so that strand 2 is longer), plus (c) any number of additional molecules of type 2. The first of the above three components in the second term contributes thefactor 1/(1 - II), the second leads to R1 (x 1Yo), and the third gives 1/(1 - xz)' The explicit expression for R1 is (16.9) Note how the integers 0, 1, 2, 3 and 1, 1, 2, 3, above, occur as powers of Xl and xz, respectively, in R 1. The maximum number of type 1 molecules in R1 is i1 - 1 = 3 (in order not to make another i 1, i z package). The third term (strand 1 is longer) in Eq. (16.7) is similar. There can be (a) any number of vernier packages plus (b) 0,1, or 2 (i.e., iz - 1) type 2 molecules, together with 1, 2, or 3 type 1 molecules, respectively (so that strand 1 is
178 Some Attached Multi-stranded Polymers at Equilibrium and in Transients longer), plus (c) any number of additional type 1 molecules. The three contributions to the third term are then 1/(1 - II), Rz(xzy), and 1/(1 - xd, respectively, where (16.10) Note that the integers 0,1,2 and 1,2,3 appear here as powers of X z and Xl' respectively. Rl and R z are found in any example with arbitrary i l , iz , by comparing lengths as above. The simplest special case (the two molecules have the same length) is I = 1/1, II = X1XZy, Rl = Xz, and R z = Xl' If the two components are the same, Xl = Xz = X and II = XZ y, this case becomes essentially the two-stranded aligned model already treated to some extent in Section 13 (see also Section 19). The separate terms in 1 are proportional to state probabilities [Eq. (16.6)]. Hence the three terms in Eq. (16.7) are proportional to the probability that the strands are equal (vernier), strand 2 is longer, or strand 1 is longer, respectively. The separate terms in Rl and R z have a similar probability interpretation. Because 1 - II appears in all three terms in Eq. (16.7), the probability that the polymer has n complete vernier packages (irrespective of additional molecules) is IIn(l - II), as in Eq. (5.12). The mean value of n, ii, is II/(l - II). If the common factor 1/(1 - II) is removed from the three terms in Eq. (16.7), the remaining partition function is the same (for this model) as the "roughness" partition function 10' to be introduced in Section 19. In any given case, the values of Xfl Xf2 yN in Eq. (16.6) are easily calculated and presented in an N l , N2 table, as illustrated in Table 6-1. The probability of any vernier structure Nl = nil' N2 = ni 2 relative to its four immediate neighbors in the table is easily seen to be P N1N2 Xl PNl+l.N2 P N1N2 Xz P N1 • N2 +1 , P N1N2 P Nl -l,N2 , P N1N2 P N1 ,N2 -l = X1Yo, (16.11) = X2Y· Table 6-1. Values of X / 1xz N2 y N = YPN1N2 for a Case" with 1=3/2 N2 = Nl =0 0 2 3 4 0.0001 1 2 1.0000 0.1000 0.0100 0.0316 0.3162 0.3162 0.0010 0.0100 0.1000 0.0003 0.0032 3 0.0010 0.0316 1.0000 0.03i6 0.0010 4 5 6 0.0001 0.0032 0.0003 0.1000 0.0100 0.0010 0.3162 0.3162 0.0316 0.0100 0.1000 1.0000 "y = 1000, Xl = Yo -1 /2, X2 = y- 1/2, II = 1.
Simple Dual Aggregation and the Vernier Effect 179 In cases of interest for the vernier problem, all of these quantities are > 1. That is, any vernier structure is represented by a local peak in the probability table. The peak is symmetrical in both directions (l/Xl = X1Yo and 1/x2 = X2Y) for the particular values Xl = y;;-1/2 and X2 = y-l/2. For this choice of Xl and X2' n = 1. The example in Table 6-1 is ofthis type: I = 3/2, Y = 1000, Yo = 100, Xl = 0.1, and X2 = 0.03162. Because n = 1, Table 6-1 continues indefinitely with a repeating pattern (Y diverges but the relative probabilities are significant). The vernier values (nn = 1) are underlined. If Xl and X2 are chosen so that n < 1, the table converges (the vernier values are nn < 1). The highest peaks in Table 6-1 are at the vernier values but there are smaller peaks at: N2 = 1, Nl = 1 or 2; N2 = 3, Nl = 4 or S; etc. This occurs for any choice of 1: the state probability is relatively high if the two strands have almost equal lengths, though not exactly equal lengths. Because of this feature, simple vernier structures do not have the dominant uniqueness required to produce the kind of vernier effect mentioned at the beginning of this section; there is significant competition from near-vernier structures. Rate Constants and Kinetics We take the on and off first-order rate constants for a single strand [as in Eqs. (16.1)-(16.3)] to be ala l and a'l' respectively, for component 1 and a 2 a 2 and a~ for component 2. Using a l a l and a'l in the kinetic diagram for Nl = 0, 1,2, ... (see Fig. 2-4 with C = 1), at equilibrium (detailed balance), it is easy to see that PN1 ex (alal/a'd N1 • Comparison with Eq. (S.12) shows that Xl = alat/a'l and X2 = a2a2/a~, as in Eq. (S.1S). In the dual polymer, the attractive interactions between the strands must influence the on and off rate constants. We shall assume throughout this chapter that the full effect of these interactions appears in the off rate constants (e.g., the on rate constants are diffusion controlled). Let Nw be the lateral interaction free energy, as already defined, in an arbitrary state Nt> N 2. If N21 ;;: N l , then N = Nt/I; otherwise N = N 2. If a type 1 molecule is lost from the N l , N2 polymer, the value of N becomes, say, N(l). If now Nzi ;;: Nl - 1, then N(l) = (Nl - 1)/1; otherwise N(1) = N z . If a type 2 molecule is lost, the value of N becomes N(2). If (N2 - 1)1;;: N l , then N(Z) = Nt/I; otherwise N(Z) = N2 - 1. The two off rate constants are then ° (16.12) ° The above algorithm takes care of all possible cases except (a) a'l = if Nl = and (b) a~ = if N z = 0. The most common special cases are at = a'l if strand 1 is initially longer by at least one type 1 molecule, a! = a~ if strand 2 is initially longer by at least one type 2 molecule, at = a't/yo if strand 2 is initially equal in length or longer, and a! = a~/y if strand 1 is initially equal in length or longer. These latter two rate constants are much reduced (y » 1) because the departing subunit must pull away from the adjacent attracting °
180 Some Attached Multi-stranded Polymers at Equilibrium and in Transients strand. There are also various "fractional" cases (depending on I) in which only part of a molecular interaction is lost in the off transition. In principle, one could now proceed to fill in all the rate constants in a two-dimensional kinetic diagram with states Nb N2 (as in Table 6-1). If the kinetic system were allowed to come to equilibrium (11 < 1), the equilibrium distribution among states would be the same as in Eq. (16.6) (using Xl = a 1a 1/a'1 and X 2 = a2a2/a~). However, to study transients and cases in which the polymer grows (11 > 1), it is necessary to use Monte Carlo calculations. There are four principal steady regimes: (a) 11 < 1 and equilibrium; (b) 11 > 1 with strand 1 always longer and growing faster than strand 2; (c) 11 > 1 with strand 2 always longer and growing faster than strand 1; and (d) 11 > 1 with the two strands growing together, that is, maintaining equal lengths except for fluctuations. Regime (a) has been discussed already. In regime (b), the net mean subunit fluxes in the two strands are and J2 = a~ aza z - - . Y (16.13) The condition for strand 1 to be growing is a 1a 1 > a'l or Xl > 1. For strand 1 to be growing faster (so that strand 2 does not catch up), the condition is J 1 > lJ2 • This is equivalent to Zl > 1, where Zl == a1 a 1 + (Ia~/y) . + a'l la z a 2 The term in 1/y is usually very small because y » 1. In regime (c), a' J 1 = a 1a 1 - ~ and Jz = a 2 a2 - a~. Yo (16.14) (16.15) The condition for strand 2 to be growing is a Z a2 > a~ or X z > 1. For strand 2 to be growing faster, lJz > J 1 • This is equivalent to Zz > 1, where (16.16) F or the strands to grow together, regime (d), we need 11 > 1 and also < 1 and Z2 < 1. The latter two conditions insure that whenever one strand becomes longer through a fluctuation, the other strand will, on average, grow faster and catch up. The actual rate of dual growth is not easy to express analytically except in the simple case I = 1/1 (see below). If successive pairs of N 1 , N2 values are followed (i.e., one transition at a time) in the course of a Monte Carlo simulation in a typical case of this kind (e.g., 1= 8/3, y» 1), starting at N1 = 0, N z = 0, the dual polymer is seen to increase in size stochastically but with significant pauses in the neighborhood of each vernier structure, nil' niz (n = 1,2, ... ), where there is extra stability. Thus, the mean rate of growth of the dual polymer is determined essentially by the mean first passage times from one vernier structure n to the next (n + 1 or n - 1). A number of examples of the above four regimes have been followed by Zl
Simple Dual Aggregation and the Vernier Effect 181 Monte Carlo simulation, especially for the case I = 8/3 and with Y = 10 3 to 10 5 . Both in regime (a) transients with II ~ 1 and in regime (d) with II > 1, vernier structures (n = 1,2, ... ) do indeed exhibit the extra stability expected, but the additional lifetime spent by the kinetic system at and near the vernier structures (n = 1 is the most important structure for the vernier effect mentioned at the beginning of the section) does not appear to be sufficient for a particular such structure (e.g., n = 1) to have a good chance to be incorporated (in vivo) into some superstructure that would freeze the vernier structure permanently. Some additional source of stability of vernier structures seems to be needed. One possibility would be to have as a third component present in solution a ligand that binds very tightly to the end of both strands if both strands have exactly the same length. The bound ligand could then serve as a cap that prevents subunits of either type (l or 2) from leaving or adding to the polymer. Because, in the growth process starting from Nl = 0, N z = 0, the vernier structure n = 1 is encountered first, the cap would tend to stabilize this particular structure. Three other mechanisms for stabilizing vernier structures, without using a third component, are discussed in Section 17. Dual Aggregation in the Case 1= 1/1 It has been pointed out already that in Eq. (16.7), in the case I = 1/1, we have R1 = X 2 , R2 = Xl' and II = Xl X2Y' The aggregation is cooperative. Dual bulk polymer will form if II = X 1 X 2 Y ~ 1, even though Xl < 1 and X 2 < 1. This is not a vernier example, but it has its own intrinsic interest. In this case (l = 1/1), it is relatively simple to construct the two-dimensional kinetic diagram for the states Nl , N 2 • From this one can write the master equations in dPN,N,/dt (for N1 > N 2 , N2 > Nu and Nl = N z ) and then find the intuitively obvious subunit flux equations, (16.17) where PN2 <N, is the sum of PN1N2 over all N1 and over all N2 < N l , etc. These equations apply even in transients. For steady dual growth of type (d), discussed in the previous subsection, it is of interest to find the Ps in Eqs. (16.17) in order to deduce explicit expressions for J 1 and J 2 • This is easy to do using a one-dimensional kinetic diagram for the quantity m == N z - N l . This is shown in Fig. 6-2. Although the polymer is growing, Fig. 6-2 has a stationary solution for the probabilities Pm' That is, the tip of the growing polymer has a steady-state distribution of states m (Chapter 5). The probabilities Pm are found from the "detailed balance" between neighboring pairs of states in the linear diagram: 3
182 Some Attached Multi-stranded Polymers at Equilibrium and in Transients a2a2 m=N 2 -N I =···-2 .. O'la t + a} Q2 a2 ~ + a'r ~ -1 >II + Q'l Y O'ial 0 + Ql Y + Q'~ Y ~ >II O'lal el2a2 + ell - Y +1 + 0'2 Po P-I P-2 Qla2 ~ 4 alaI PI +2 ... + 0'2 P2 Fig.6-2. State diagram, with rate constants, for m = N2 - Nl in the case of an I = 1/1 dual polymer, both of whose strands are growing steadily and together. where Zl and Zz are given by Eqs. (16.14) and (16.16) with 1 = 1 and Yo = y. We then find, for use in Eqs. (16.17), (16.19) Equations (16.18) and (16.19) are valid provided that Zl < 1 and Zz < 1. Equations (16.17) become, then . J 1 = a 1Q 1 - Jz = azQz - a'l z l(1 - zz) 1a~zz(1 1- ZlZz - Zl) ZlZZ a'l(1 - zd - ----y(1 - Z1ZZ) - a~(1 - zz) y(1 - Z1ZZ) (16.20) . The two strands in Eqs. (16.20) (II> 1, Z1 < 1, Zz < 1) are in fact growing together. Hence we should expect J 1 = Jz. Substitution of Eqs. (16.14) and (16.16) (with 1 = 1 and Yo = y) confirms this, after considerable algebra. In fact, both fluxes simplify to the symmetrical expression + a~)(a1 Q 1azQzy - a'l a~) y(a 1Q 1a'1 + azQza~) + a'la~(1 + y) (a'l J 1 = Jz = -------'-'-----"------- (16.21) Note that the numerator is proportional to II-I, that is, to e X / kT - 1, where X is the thermodynamic force driving aggregation of component 1, 2 pairs [see the discussion of Eq. (16.8)]. Dual Aggregation in the Case I = 2/1 This is the next simplest case, in which type 1 subunits are exactly half as long as type 2 subunits. Even for this case there seem to be only a few simple analytical results.
Simple Dual Aggregation and the Vernier Effect ... Ct'2 a1 -2 • 183 Ct'2 Q 2 0 (X2/Y • 0'; Fig. 6-3. State diagram, as in Fig. 6-2, but for M = • +7 ... 2N2 - N1 in the case 1 = 2/1. In the equilibrium regime, n = xi XzY < 1, R1 in Eq. (16.7) is X z + XZ (x 1 Yo) and R z is Xl, with Yo = y1/Z. Equations (16.13)-(16.16), with 1 = 2, apply in regimes (b) and (c). In regime (d), the two strands grow together, with n > 1, Zl < 1, and Zz < 1. The one-dimensional kinetic diagram in M = 2Nz - N1 is shown in Fig. 6-3. The rate constants on the right and left of the figure persist at higher values of IM I; rate constants change only in the center ofthe diagram (M = - 1,0, + 1). Figure 6-3 is the analogue of Fig. 6-2 for I = 1/1. The two strands have the same length at M = O. There is a steady-state distribution of probability (PM) among the states of Fig. 6-3 even though the polymer is growing steadily. This distribution could be found numerically, by iteration, in any particular case, but a simple analytical solution probably does not exist. The asymptotic form of PM at large positive M is PM ex sr, where Sz < 1 for convergence. To find 8 z, we equate the steady flow into an arbitrary state M with the steady flow out of this state: rtzazsr- z + (rt'l/Yo)8r- 1 + rt 1 a 1 Sr+ 1 + rt~8r+z (16.22) Division by 8r- z gives a quartic in 8z . One root of this is 8 z 8 z - 1 out, a cubic in 8z remains: = 1. Factoring (16.23) This determines 8 z as a function of rate constants and Yo' One root of the cubic is positive. This root approaches unity, 8z ~ 1, when Zz ~ 1, that is, when [Eq. (16.16)] + (rt'dYo)~rtlal + 2rt~. = 0 and lJ('l = 0, 8z = (lJ(zaz/IJ(~)1/2. In the special case 2rt z a z In the special case rtla l rt 2 a 2 = 0 and rt~ = 0,82 = rt'drtlalyo' There is a different asymptotic solution at large negative M: PM ex 81M. The same kind of procedure leads to the cubic equation in 8 1 , rt z a 2 8i + (rt z a 2 + rt'l)8i The positive root 8 1 ~ 1 when Zl - [rtla l ~ + (rt~/Y)]8l - (rt~/Y) = O. (16.24) 1 [Eq. (16.14)]. In the special case rt 1 a 1 = 0
184 Some Attached Multi-stranded Polymers at Equilibrium and in Transients ° and a~ = 0, and a'l = 0, Sl = (a~/a2a2y)l/2. In the special case a 2a 2 = S1 = a 1ada'l' One can verify numerically that these asymptotic solutions at large IMI do not persist to the center of the diagram (Fig. 6-3). The formal equations for the subunit fluxes, the analogues of Eqs. (16.17), are (from Fig. 6-3) (16.25) In steady growth, J 1 = 2J2 • However, the Ps in Eqs. (16.25) are not available without a solution for all the PM' 17. Dual Aggregation with Vernier Enhancement In the discussion following Eq. (16.16) it was mentioned that, in simple dual aggregation, vernier structures are indeed favored as expected, but not in a very dominating way. A ligand that binds strongly to the tip of both strands of vernier structures only, essentially capping such a structure, could enhance the vernier effect considerably. In this section we discuss three other models for vernier enhancement that do not involve a third component (the ligand, above). The first of these models is actually unsatisfactory but is included for comparison. The other two models are rather similar. Kinetics and numerical results are discussed only for the third model. Partition Functions for an Open System The first model is shown schematically in Fig. 6-4(a), for the case 1= 3/2. Both components have terminal appendages that interact with each other, with free energy W < 0, at each of the n vernier points in the dual polymer. This special interaction supplements the lateral interaction between strands, Nw, already introduced in Section 16. We define Y = e- W / kT • Because each vernier package in the polymer (whether there are additional subunits or not) is further stabilized by a factor Y > 1, Eq. (16.7) for Y is modified by replacing II everywhere by IIY. The polymer becomes very large when IIY -+ 1: the cooperativity between the two strands is stronger because of Y; the monomer concentrations needed for "condensation" of bulk polymer are smaller. The equilibrium probability PN,N2 of a state N 1, N 2 , when IIY < 1, is xf' x~2yNyn/y. Although vernier packages i 1 , iz receive extra stability in this model, this extra stability persists when additional subunits are added [e.g., the structure
Dual Aggregation with Vernier Enhancement 185 NI =4 n =I N2 = 3 (a) e ~B ----H -A NI=3 N 2=" - ------ 2 (b) _l-d. KY w Monomers B H 2 KH ---"--~ )A A lA (e) Fig.6-4. (a) Model (the case I = 3/2 is shown) in which each vernier package has extra stability because ofthe interaction (W) between the aligned tips of the two components. (b) In this model the tips of the two components are flexible and can form (state B) a reversible bond, but only at the very end of the polymer (in a vernier structure). (c) A similar model in which component 2 (only) undergoes a conformational change (A ~ B) such that state B interacts strongly with the end of strand 1 in a vernier structure. in Fig. 6-4(a) is stabilized by a factor Y]. That is, an exact vernier structure is enhanced by a factor Y relative to slightly smaller structures but not relative to slightly longer structures. In the second model, Fig. 6-4(b), the distal tips of both components are assumed to have some flexibility, allowing a reversible bond to be formed (state B) between the two terminal subunits of a vernier structure. The bond is broken in state A. The equilibrium constant for A <=± B is K » 1. Because of the distortion at the subunit tips in state B, further subunits cannot add to this state. The polymer can grow beyond a vernier structure only from state A. This model stabilizes only the final vernier package of a vernier structure. The A <=± B equilibrium is possible only in vernier structures. In effect, in this model, there is a reversible internal cap. The contribution to Y of the vernier structure Nl = nil' N z = niz (with
186 Some Attached Multi-stranded Polymers at Equilibrium and in Transients n ~ 1) is rr n- 1 . rr(1 + K) or rrn(1 + K). The factor 1 + K arises from the A ~ B equilibrium. 3 Hence Eq. (16.7) becomes (17.1) The individual state probabilities are as in Eq. (16.6) except that each vernier structure (n ~ 1) has an extra factor 1 + K. Thus, if K is large, the vernier structures can be very dominant in a probability table such as Table 6-1. The third model is similar. Type 1 molecules have no special properties but the distal end of a type 2 molecule can undergo a conformational change [A <=± B in Fig. 6-4(c)], the main feature of which is that state B, in a vernier structure, interacts strongly (free energy W < 0) with the tip of strand 1. This stabilizes the vernier structure and forms a reversible internal cap (as in the second model). Monomers of component 2 are in an A <=± B equilibrium. The activities are a ZA and a ZB = KaZA, where K is an equilibrium constant of order unity. Strand 2, by itself(i.e., a 1 = 0), has the following properties. A new monomer (A or B) can add only to a terminal subunit of the polymer that is in state A. Also, only the terminal subunit can undergo the A ~ B transitions (again with equilibrium constant K). Hence all internal subunits in strand 2 are in state A. Corresponding to Eq. (16.2) [compare also Eq. (17.1)], the open partition function for strand 2 alone is then Y = 1+ co L xf2- 1 . xz(1 + K) N2=1 =1+ (1 + K)x z 1~ Xz 1 + Kxz 1 ~ Xz =--- (17.2) ' where Xz relates to a 2A (that is, X2 = a 2A /a2A) because the bulk polymer, when X z -+ 1, is virtually all A. When both strands are present, the A ~ B transitions at the tip of strand 2 are possible only if strand 2 is at least as long as strand 1 (so that there is room for the structural change). If the strands have the same length (vernier structure), the A ~ B equilibrium constant is KY, where Y = e- w /kT » 1. A vernier structure in state B cannot add a subunit of any type (1, 2A, 2B). A monomer of type 1 cannot add to the polymer if strand 2 is in state Band strand 2 is not longer than strand 1 by at least 11 (otherwise there would be interference with the strand 2 tip). Similarly, a type 2 monomer in state B cannot add to the polymer unless the addition will make strand 2 at least as long as strand 1. With this model, Eq. (16.7) is modified to Y = 1 + (1 + KY)rr + (1 + K)R 1 (X 1 Yo) + 1 ~ IT (1 ~ xz)(1 ~ IT) (1 Rz(xzY) ~ x 1 )(1 ~ (17.3) IT)
Dual Aggregation with Vernier Enhancement 187 Table 6-2. Values of YPN,N z for a Vernier-Enhanced Case" with 1=8/3 N2 = Nl =0 1 2 3 4 5 6 7 8 9 10 0 1.000 0.152 0.023 0.004 0.001 0.304 0.616 1.249 0.534 0.081 0.012 0.002 2 3 0.046 0.094 0.190 0.385 0.780 1.580 0.285 0.043 0.007 0.001 0.007 0.014 0.029 0.058 0.119 0.240 0.487 0.987 101.000 0.152 0.023 "y = 1000, Y = 100, K = 1, Xl = X2 4 5 0.001 0.002 0.004 0.001 0.009 0.001 0.018 0.003 0.037 0.006 0.074 0.011 0.145 0.023 0.304 0.046 0.616 0.094 1.249 0.190 ................. = 0.15199, II = 1. The values of K and Y have no effect on the "condensation" point II = 1. Corresponding to Eq. (17.3), the state probabilities in Eq. (16.6) are multiplied by 1 + KY for vernier structures (n ~ 1) and by 1 + K whenever strand 2 is longer than strand 1. Because K is of order unity and Y » 1, vernier structures have unusually high probabilities. We consider, as a numerical example, a case with 1= 8/3 (as in Ref. 2, p. 126). We take as reasonable parameters y = 1000 (hence Yo = 13.34), Y = 100, K = 1, and Xl = X z = 0.15199. These values of Xl and X z were chosen to give II = x~x~y3 = 1. Values ofYPN,N2 are given in Table 6-2 for small values of Nl and N z . However, because II = 1, the family of numbers between N l , N z = 0, 0 and 8, 3 repeats itself between 8, 3 and 16, 6, between 16, 6 and 24, 9, etc. The vernier value of YPN,N z is always 101.0 (n ~ 1). Without this model's vernier enhancement (put K = 0), all values in Table 6-2 above the dotted lines (strand 2 is longer) would be divided by two and 101.0 would become 1.0. It is apparent that vernier enhancement can produce the desired qualitative effect. Rate Constants and Kinetics The rate constants a l and a'l are the same as in Section 16. For component 2, we assume that a z and a~ apply to both A and B conformations (but in general aZA and aZB are different: aZB = KaZA)' We then have X z = aZaZA/a~. The new interaction parameter Y affects both aT and a! in Eq. (16.12) (we assume Y has no effect on a l and a z ). Because Y applies to vernier structures only,
188 Some Attached Multi-stranded Polymers at Equilibrium and in Transients ex! ex' =_1 YoY and ex' ex!=~. yY (17.4) That is, the off rate constants are further reduced by the necessity to break the W "bond." For other than vernier structures, a! and a! are unchanged [Eq. (16.12)]. When strand 2 is longer, the rate constant for A -4 B is k1 and for B -4 A is k 2. Hence K = k 1/k 2. When the strands have the same length, we take for the corresponding perturbed rate constants, k 'l = k1 y1/2 an d k'2 = y1/2· k2 (17.5) This divides the free energy effect (W) equally in the two directions. We need not be concerned with A +2 B monomer kinetics; we assume a large equilibrium reservoir of monomers with fixed a2A , a2B , and a 1 • The discussion of the four steady regimes in Section 16 still applies here except that in Eqs. (16.13)-(16.16) a2 is replaced by a2A . In Eq. (16.15), a2a2 must be divided by 1 + K because strand 2 can receive a new monomer only when the tip of the strand is in state A; this alters ex 2a 2 to a 2 a 2A . Some kinetic properties of the first vernier structure in a number of cases related to Table 6-2 (/ = 8/3) are given in Table 6-3. The first case in Table 6-3 is a reference case with the same thermodynamic parameters (y, Y, K, Xl' X2) as in Table 6-2 and in addition the kinetic parameter choices ex 1 = a 2 = 1 and a'l = a~ = k2 = 1 (in arbitrary units of time and concentration). With these choices of a i and a; (which are not varied), Xl = a 1 and X 2 = a 2A . Thus, in the reference case, a 1 = a2A = 0.15199. The value of k1 is not independent: k1 = k 2K. Thus, k1 = 1 in the reference case. Except for the last two cases in Table 6-3, where several parameters are changed, other cases in the table have only one parameter change from the reference set. These changes are indicated in column (1). For each ofthe 13 cases in Table 6-3, a Monte Carlo simulation was carried out to determine the number of transitions of any kind and the time required (a) for a polymer originally in state N1 = 0, N2 = 0 (i.e., empty aggregation sites) to reach the first vernier structure N1 = 8, N2 = 3 and (b) for a polymer originally in state N1 = 8, N2 = 3 to escape from the neighborhood of this vernier structure (which has a free energy minimum). Each of the 26 simulations was repeated between 100 and 200 times, and averages taken. To be more precise, in process (a) the vernier structure was considered to have been reached when the conditions N1 ;?; 8 and N2 ;?; 3 were first simultaneously satisfied. In process (b), escape from the neighborhood of the vernier structure was deemed to have occurred when either of the conditions N1 + N2 = 5 or N1 + N2 = 17 was satisfied (note that the first vernier structure, the starting point, has N1 + N2 = 11 and the second has N1 + N2 = 22). Column (2) in Table 6-3 gives the value of II for each case (recall that II > 1 would lead to indefinite polymer growth). Columns (3) and (4) give, in the first
189 Dual Aggregation with Vernier Enhancement Table 6-3. Mean First Passage Times for the Table 6-2 Case (1) Parameter change (2) n (3) Mean no. transitions (4) Mean time (5) Escape high 574 496 439 1901 0.574 0,0 ---> 8,3 Out of8, 3 a l x 0.8 0.168 1001 520 798 1997 0.408 a l x 1.2 4.30 293 465 216 1796 0.760 113 225 78 891 0.977 a l x 2.0 256 alA x 0.5 0.125 993 557 989 2270 0.268 alA x 2.0 8 317 327 181 1089 0.852 553 636 384 303 489 718 332 2450 0.581 395 2091 374 4144 0.577 314 549 239 2153 0.735 265 71 454 126 0.578 370 102 1175 325 0.600 446 92 750 161 0.570 K= 2 kl = 0.5 y= 200 Y = 2000 8 K=O K = 0, Y = 104 *al' *a lA K=O **a l , **a lA *a l = a 2A = 0.081113 **a l = Yo -1/2 = 0.2738, alA ° = 0.545 (6) Time ratio 4.3 2.5 8.3 11.4 2.3 6.0 7.4 5.3 11.1 9.0 0.28 0.28 0.21 y-1 /2 = 0.0316. line of each case, the mean number of transitions and the mean time required to reach state 8, 3 from 0, (i.e., to produce the vernier structure by aggregation, starting from nothing). The second line of each case gives the mean number of transitions and the mean time during which the first vernier structure survives in its own neighborhood (in the N I , N2 plane) before either the polymer dissolves (approaches Nl = 0, N2 = 0) or approaches the second vernier structure (Nl = 16, N2 = 6). The longer the dual polymer survives in the neighborhood of state 8, 3, the more chance the vernier structure can be incorporated into some superstructure that will confer an indefinite lifetime
190 Some Attached Multi-stranded Polymers at Equilibrium and in Transients on the vernier structure. Most of the time spent "in the neighborhood of" state 8, 3 is actually spent at 8, 3 itself: in the reference case, over a long period of time, the fraction of time spent at each of the states Nl , N z would be proportional to the numbers in Table 6-2 (at and around state 8, 3). In fact, this feature was used in several cases to check the Monte Carlo computer program. Column (5) records the fraction of escapes from state 8, 3 that occurred on the high side (Nl + N z = 17 was reached first). This fraction tends to be high for II > 1 and low for II < 1, as expected. Column (6) gives the ratio of the mean survival time at state 8, 3 to the mean formation time of this state, in each case. The parameter changes (from the reference set) in Table 6-3 are not meant to be exhaustive but rather are designed to show the influence of each parameter. The most favorable case listed (large survival time, large ratio in the last column) is Y = 200. This is not surprising because Y is the principal vernierenhancing parameter. The last three cases in the table (K = 0) refer to simple dual polymers without vernier enhancement (as in Section 16). State B does not exist. In the next to last case, we have K = 0, y = 104 , and a l = aZA adjusted to give II = 1 again. This is a simple dual polymer but with stronger lateral interactions. In the last case, we return to the reference value y = 1000 but choose a l and aZA as prescribed following Eq. (16.11). These three cases without vernier enhancement (K = 0) are seen to be the least satisfactory in the table. Vernier enhancement again seems necessary in order to obtain desired properties. 18. A Further Example of Dual Aggregation In this section, we obtain the equilibrium open partition function Y for another kind of dual aggregation. This model is not concerned with the vernier effect; it is suggested by virus self-assembly. The model is presented in Fig. 6-5. Figure 6-5(a) shows a transverse section through the dual polymer and Fig. 6-5(b) is a very schematic longitudinal picture. Molecules of the two components, 1 and 2, have the same length but in each such unit of length in the polymer there may be two molecules of component 1 that wrap around one molecule of component 2 [Fig. 6-5(a)]. This is a lateral version of the longitudinal 1= 2/1 case in Section 16. The lateral interaction free energies of types 1-1 and 1-2 are denoted wand u, respectively, as shown in the figure. The longitudinal interaction free energies are denoted Wl and W z [Fig. 6-5(b)]. There are three strands in the polymer, with numbers of molecules Nl l , N 1Z ' and N z [Fig. 6-5(b)J. The strands have no vacancies. The polymer is nucleated by sites on a surface. For definite values of Nll , N 1Z , and N z , the canonical partition function is [compare Eq. (16.4)J
191 A Further Example of Dual Aggregation (b) (a) I~ I. (I) 5. (I) V II 1. (1)11 6. (I) II IV III 3. (I) III 7. (I) III IV IV 4. (I) IV 8. (I) III V (c) (d) Fig. 6-5. (a) Transverse section in a simple dual aggregation model in which two components have the same length but each level may be comprised of two molecules of component 1 and one molecule of component 2. (b) Very schematic longitudinal diagram of the same model. (c) Possible types of combinations of molecules at any level of the polymer. (d) Possible sequences of combinations, each of which contributes a term to Y. (18.1) where y == e- w /kT , z == e- u / kT , N is the lesser of Nll and N 12 , and M is the number of 1~2 interactions [M = 5 in Fig. 6-5(b)]. The partition function for an open system (in contact with monomers at a l and a 2 ) is then (18.2) where the sum is over N ll , N 12 , and N2 (from 0 to (0), and Xl and X2 are defined as in Section 16. The probability of particular values of N ll , N 12 , and N 2 , in the open system at equilibrium, is (18.3) The easiest way to deduce a closed expression for Y in Eq. (18.2) is to use a modification of the method introduced for Eq. (16.7). At any level of the polymer, there are five possible types of combinations of molecules, labeled I to V in Fig. 6-5(c). For example, the successive levels in Fig. 6-5(b) are I I III
192 Some Attached Multi-stranded Polymers at Equilibrium and in Transients V. In such a sequence, because vacancies are not allowed, a I can be followed by anyone ofI, ... , V, a II can be followed by a II or a IV, a III can be followed by a III, IV, or V, and a IV or V only by a IV or V, respectively. These possibilities then allow the eight types of sequences shown in Fig. 6-5(d) where (I) indicates any number of successive I combinations, including zero, and II, III, ... indicate any number of successive II, III, ... combinations, not including zero. Each of these sequences contributes terms to Y. That is, Y = Y I + ... + Y 8, where 1 Xz Ys =--·--1 - II 1 - x z ' (18.4) Xz Y s = - 1_ · 2x l xzz . _ I-II 1-x l x z z 1-x z ' where II = xi xzY z ZZ and factors of 2 have been included where a degeneracy requires it. The quantity YjY is the probability of observing a sequence of type i. A more compact form for Y follows on combining 1, 2, 4, 6 and 3, 7 and 5, 8: Y = 1 + Xl (1 - II)(l - xiyz)(1 - xd + - - - -2x-l x-Z z- - - (1 - II)(l - xlxzz)(l - x z (1 + XIXZZ) (1 - II)(l - xlxzz)(l - xz) +--------------:- (18.5) xi xzY z ZZ --+ 1. Condensation of bulk dual polymer occurs when II = equations - Nz 81nY = -81--' nx z 81nY N=-81ny Z ' - 2Nll = Xl) - 2N12 81nY = -8-1nX 81n Y M=-8lnz The I (18.6) can be used to find the mean values indicated, when II < 1, but we do not pursue this. In the special case Xl = 0, we have Y = 1/(1 - Xz), as in Eq. (16.2). If Xz = 0, 1 + Xl Y--,-,---------:-c-:----='-------;;,-----;;-:- (1 - x l )(1 - xiyz)· (18.7)
193 Aligned Tubular Models at Equilibrium This is the 1= 1/1 case, with Xl = X 2 = X, already mentioned following Eq. (16.10), except that here there are two contacts, w, between the two strands. If there are no lateral interactions (y = 1,z = 1), Eq. (18.5) simplifies to Y= 1 (1 - xd 2 (1 - x 2 ) (18.8) ' as expected (the three strands are independent). 19. Aligned Tubular Models at Equilibrium In Chapter 5 we examined the "roughness" at an end of a long tubular aggregate of s strands where the subunits in the different strands might be either aligned or staggered relative to each other. Here we use the method of the previous sections (see especially Fig. 6-5) to extend some of the results given in Section 13 of Chapter 5 for aligned tubular models at equilibrium. There is only one type of subunit; the vernier effect is not involved. We shall continue to use the notation already introduced in the present chapter. We consider an aligned tubular aggregate of s strands, initiated by sites on a surface (as in the other sections of this chapter). The aggregate is at equilibrium and has a finite size. In Section 13, the only equilibrium state considered was at the critical activity a e of free subunits, where bulk polymer is formed. Here we generalize to a ~ a e • Figures 6-6(a) and 6-6(b) show transverse sections for s = 2 and s = 3. The interaction free energy W here was called W h in Chapter 5. Note that there are two lateral interactions, w, in Fig. 6-6(a). This is in contrast with the nontubular two-stranded model in Fig. 6-6(c) (the 1 = 1/1 case of Section 16, with Xl = x 2 ), which has only one lateral interaction w. We define, as usual in this chapter, y = e- wjkT and [Eq. (16.3)] \ =3 = :! '\ -= '" G~ @ DO 1\' (=) (=) - (==) (==)= X2 <==)<==)= X3 X3 X6 II ' II' (J) ~ (b) (e) hlJ (e) Fig. 6-6. (a) Cross section of a tube with two strands. (b) Cross section of a tube with three strands. (c) Cross section of a nontubular two-stranded aggregate. (d) Types of configurations in the s = 2 (tube) case. (e) Types of configurations in the s = 3 case.
194 Some Attached Multi-stranded Polymers at Equilibrium and in Transients (19.1) The interaction free energy W1 is the same as Wv in Chapter 5. We first adapt the method used in Fig. 6-5(d) to the s = 2 case [Fig. 6-6(a)]. The open partition function Y for this aggregate has two contributions, shown diagrammatically in Fig. 6-6(d): y=_I_+ I-II 2x (I-II)(I-x)' (19.2) where II = x 2 y2. The first term represents any number of complete packages x 2 y2, from 0 to 00, attached to the surface. The second term includes this contribution plus at least one additional subunit (1 to 00) on one of the two strands. All possible configurations of the aggregate are included in the two terms in Y. Bulk condensation of the aggregate occurs when II -+ 1, that is, when x -+ 1/y (recall that x is proportional to the subunit activity a). Equation (19.2) is valid for II < 1 or x < l/y. Note that Eq. (19.2) is the same as Eq. (18.7), as expected. Also, l/y = e w /kT is the same as p in Chapter 5. Let us digress at this point to compare some critical activities ae • First, we recall that a e is proportional to the "absolute activity" Ae [Eq. (4.1)]: (19.3) For a single strand (with no lateral interactions), Y = 1/(1 - x) [Eq. (16.2)] and condensation occurs as x --+ 1. From Eq. (19.1) in this limit, (19.4) We use this as a reference result. For a two-stranded aggregate with only one lateral interaction [Fig. 6-6(c)], II = x 2 yin Eq. (19.2) and II --+ 1, x --+ y-l/2 determines the critical activity. Thus (19.5) That is, the critical activity is reduced (w < 0) by the lateral interactions, as expected, because these interactions increase the stability of the condensed phase (i.e., they lower its chemical potential). With two lateral interactions at each level [Fig. 6-6(a)], II -+ 1, x --+ llY (see above), which gives a still lower critical activity (for the same value of w): (19.6) Note that this is the same result as in Eq. (13.24) [see also Eq. (4.26)]. Equation (19.6) is, in fact, found for any aligned tubular aggregate of s strands (s > 1), because II = xSyS --+ 1, and x -+ l/y for any s. This simple result occurs because there is one lateral interaction per subunit in a complete tubular package of any size, s > 1. We return now to the main argument. One can factor 1/(1 - II) out of both terms in Eq. (19.2), leaving y =1+~=l+x o 1- x 1 - x' (19.7)
Aligned Tubular Models at Equilibrium 195 which is valid for x ~ l/y (see below). The physical significance of the partition function Yo is that it represents the sum over all aggregation states of the polymer beyond the last complete package (x 2 y2). This is, then, a "surface roughness" partition function (Chapter 5). Incidentally, the fact that 1/(1 - II) can be factored out of Y for all of these models (i.e., for any s) corresponds to the fact that the distribution of "rough" states is the same for any number n of underlying complete packages X S yS. In the simple s = 2 case, roughness can arise only from extra subunits on one strand or the other. The reason y does not appear in Eq. (19.7)is that no lateral interactions can occur with one strand only. The polymer becomes infinitely long (a ~ a e ) when x = l/y = e w /kT , but Yo, which we have denoted by Qs in this limit (Chapter 5), remains well behaved (i.e., the singularity has been removed from Y). The result (using x = e w / kT = p) is the same as Eq. (13.1). However, Eq. (19.7) is more general; it applies whenever x ~ l/y. From Eq. (19.7), the mean number of subunits beyond the last complete package is 2x _ 8lnYo m=---=--(19.8) 8lnx 1 - x2 · This is the same as Eq. (13.3) in the limit a ~ ae , x ~ p. We turn now to s = 3 [Fig. 6-6(b)], an aligned aggregate of three strands. All possible configurations of such an aggregate are represented by the four diagrams in Fig. 6-6(e). For example, the last diagram refers to configurations with any number (0 to 00) of complete packages x 3y3 as a base, supplemented by at least one level (i.e., 1 to 00) with subunits in two strands (three choices), and this supplemented by at least one further subunit on one of these two strands (two choices). The four diagrams in Fig. 6-6(e) correspond to four terms in Y: 1 3x 2 y 3x 6x 3 y 2 y= I-II + (I-II)(I-x y) + (I-II)(I-x) + (I-II)(I-x2y)(l-x)' (19.9) where II = x 3 y3. This equation is valid for x < l/y. The separate terms are proportional to the probability of observing configurations of the different types. The "roughness" partition function Yo follows on factoring out 1/(1 - II). After simplification, we find 1 + 2x + 2x 2 y + x 3 y (l-x 2y)(l-x) Y=---~-'------o (19.10) This agrees with Eq. (13.4) in the limit x = l/y = e w / kT = p but Eq. (19.10) is more general: it is valid for x ~ l/y. For any s ~ 3, Yo will be a function of x and y, as in Eq. (19.10). The derivative 8 In Y 0/8 In x then gives the mean number of subunits in the aggregate beyond the last complete package xSys. Also, 8lnYo/8lny gives the mean number of lateral interactions, w, in this part of the aggregate. If we subtract the latter from the former, we obtain
196 Some Attached Multi-stranded Polymers at Equilibrium and in Transients _ 0 In Yo 0 In Yo olnx olny , m=----- (19.11) where m is one-half the mean number of "missing" interactions III the aggregate. This terminology, introduced in Chapter 5, refers to the fact that in bulk (i.e., very long) polymer there is one lateral interaction per subunit (as in a complete package, X S yS), but in the "rough" part of the aggregate, beyond the last complete package, there are more subunits than lateral interactions. That is, some lateral interactions are missing, compared to bulk polymer; one-half the mean number of these is given by Eq. (19.11). A special case is Eq. (19.8) for s = 2, where 0 In Y In y = o. The rougher the end of the aggregate, that is, the more exposed surface, the larger the value of m. The two derivatives in the s = 3 case are % + x 2 y) + 6x 2 y(l - x)(l + x) - x)(l + 2x + 2x 2 y + x 3 y) o In Yo olnx 3x(1 - x 2 y)(l olnYo olny 3x 2 y(1 +x) (1 - x 2 y)(1 + 2x + 2x 2 y (1 - x 2 y)(1 + x 3 y)· (19.12) (19.13) Then _ 3x(1 - x 2 y)(l + x 2 y) + 3x 2 y(l - x)(l + x) (1- x 2 y)(1 - x)(l + 2x + 2x 2 y + x 3 y) . m=----,:-------------;::----;o-- (19.14) This agrees with Eq. (13.7) when x = l/y = p. The algebra becomes rather lengthy for the s = 4 case, so we abbreviate the discussion. There are now 12 diagrams of the type shown in Fig. 6-6(e), and there are 12 corresponding terms in Y. If we factor 1/(1 - IT) out of each term, where IT = x4y4, the 12-term expression for Yo can then be simplified somewhat to give 10 = [(1-x 4 y)(1 +3x+ 3x 2y+3X 3y 2 +x 4 y2)+x 2(1-x)(1 +xy+x 2y2 +x 3y 3)+3x 4y(y-l)] (1-x 3 y 2)(1-x 2y)(1-x 2)(1-x) . (19.15) This equation is valid for x ::;;; l/y. This reduces properly to Eq. (13.12) if x = l/y = P and to (1 - x 4 )/(1 - X)4 if y = 1. That is, Y = 1/(1 - X)4 in the latter case (four independent strands). References 1. Hill, T.L. (1986) Biophys. Chern. 25,1. 2. Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., and Watson, J.D. (1983) Molecular Biology of the Cell (Garland Publishing, New York). 3. Hill, T.L. (1985) Cooperativity Theory in Biochemistry (Springer, New York).
II Linear Steady-State Aggregates
7 Enzytnatic Activity at Polytner Tips Only In the first six chapters of this book, we have examined a simple aggregation process that involves physical interactions only. There is no chemistry. These are termed "equilibrium aggregates." Chapters 7 and 8 are more complicated in that we study the aggregation of enzyme molecules: in addition to the intermolecular (i.e., interenzyme) forces that produce the aggregation, the subunits of the polymer may be engaged in enzymatic activity. Such polymers are called 1 "steady-state aggregates." The principal subject of this chapter is covered in Sections 21 and 22. This is the idealized limiting case in which enzymatic activity is restricted to the terminal subunit at each end of a free polymer (again, as in Chapter 2-4, we consider single-stranded or effectively single-stranded polymers). For example, if E is the aggregating enzyme and E catalyzes S ~ P, the aggregating subspecies from solution might be ES, which very rapidly becomes and remains EP while in the polymer. In this case, the polymer is essentially all EP. It was generally believed, 2.3 for a few years, that the aggregation of both actin and tubulin followed essentially this pattern with S = ATP and P = ADP in the case E = actin, and with S = GTP and P = GDP in the case E = tubulin dimer. It is now realized that both of these systems are considerably more complicated than this; in fact the details in both cases are somewhat controversial. Despite these complications, the model treated in Sections 21 and 22 remains significant as a reference standard and as a relatively simple limiting case. On the other hand, because this model seems much less important than it once did, many published details of its theoretical properties will not be included here. The interested reader should consult the original papers l - 7 for this material. For completeness (and contrary to the title of this chapter), we include Section 20, which presents a very brief discussion of an aggregated enzyme
200 Enzymatic Activity at Polymer Tips Only that carries out enzymatic activity along the entire length of the polymer (unlike the models in Sections 21 and 22 and Chapter 8). Also, Section 23 is essentially an appendix that examines fluctuations in the polymer length distribution PN when PN is obtained from a finite sample of polymers. 20. Enzymatic Activity along the Polymer Length Consider an enzyme E that catalyzes S ---+ P by means of a three-state cycle: E ---+ ES ---+ EP ---+ E. Now, to simplify, suppose (as is conventional) that EP is a transient intermediate so that E operates, in effect, by means of a two-state cycle: the two states are E and ES. We suppose further that free molecules of E and ES in solution can produce a long single-stranded linear aggregate, each subunit of which can still carry out the same two-state enzymatic cycle. Note that this is the steady-state generalization of the next-to-last subsection of Section 11 (on the equilibrium binding of M on both free and polymer subunits). The first-order rate constants of the two-state cycle for free enzyme molecules in solution are indicated in Fig. 7-1. The constant concentrations of Sand P are embedded in these rate constants. For example, K is proportional to Cs (we use concentrations rather than activities in this section). States E and ES are also referred to as states 1 and 2, respectively. Each subunit in the polymer changes state stochastically, between states 1 and 2. There are nearest-neighbor interactions wij (i, j = 1, 2) in the polymer, with W 12 = W 21 • Because of these interactions, the instantaneous cycle rate constants for a given subunit depend on the instantaneous states of its two neighbors; as the neighbors change state in their cycles, the rate constants in the cycle of the given subunit also change, as described in detail in Chapters 9-11 of Ref. 8. The rate constants of free subunits in solution, shown in Fig. 7-1, are reference or unperturbed rate constants; rate constants of a polymer subunit are perturbed by interactions wij with neighboring subunits. We assume here (unlike Sections 21 and 22) that these interactions are the only source or cause of rate constant alterations when a subunit enters the polymer. In the last subsection of Section 10, we studied the influence ofintermolecular interactions wij on the mean composition near the end of a two-component polymer at equilibrium. The present problem is similar but more complicated. Fig. 7-1. First-order rate constants for two-state enzymatic cycle. The dominant direction is clockwise.
201 Enzymatic Activity along the Polymer Length There are again two interacting components in the polymer (E and ES) but the polymer is at a steady state rather than at an equilibrium. Even bulk steady-state polymer properties cannot be deduced analytically and exactly, except in special cases (see Chapters 9-11 of Ref. 8). The composition near a polymer end is even more complicated, especially because it is perturbed by on-off transitions from the pool of E, ES free in solution at a still different composition. Incidentally, there are two thermodynamic forces and fluxes in this system: the S ~ P force and flux and the subunit aggregation force and flux [see Eq. (7.14) and the discussion thereof]. The above problem can be handled exactly only by Monte Carlo calculations. Instead, we outline here a simple and very approximate treatment that will at least serve to introduce the various ingredients that are pertinent. The total concentration of free enzyme molecules is C = CE + CES' In view of the rate constants in Fig. 7-1, the fraction of these molecules that are in state 2 (that is, in state ES) at steady state is e s + y' + 1(' + y + y' I( (20.1) =------'----I( The subscript s refers to "solution." Thus CES = esc and C E = (1 - es)c. We use the Bragg-Williams 8 (or mean field) approximation for the (long) polymer. The probability that any subunit in the polymer is in state 2, at steady state, is denoted e. The essence of the BW approximation is that the neighbors of a given subunit are assumed to have a mean composition e rather than fluctuating stochastically between states 1 and 2. We assume, as part of the approximation, that the same eapplies to all positions in the polymer, including the ends. From Eqs. (47.8) and (47.9) of Ref. 8, the equation that determines e is {} = I( I( y fK Y Y1 Y fK + 1(' y fK + y' y 1- f, + y y f, + y' y1 ----;c-----:-----;---;-------;-----:---;--;c Y12 ==-, Yll yiO-O) == -2-0-' Y2 Y12 Y2==-' Y22 1 Yij f, == e-w;;/kT 1 YllY22 Y==--=--2-' Y1Y2 Y12 (20.2) (20.3) (20.4) The definitions in Eqs. (20.4) are the same as in Eq. (10.43). The Yexpressions in Eq. (20.2) are the rate-constant perturbation factors that arise in the polymer from intermolecular interactions [compare Eq. (20.1)]. Y contains e so that Eq. (20.2) is an implicit equation for the determination of e (numerically). The parameters fK and fy are constants that relate to the splitting of intermolecular free energies into forward and backward rate constant contributions [see Eq. (7.8) and Fig. 2-12]. Reasonable choices might be fK = 0 and fy = 0.5.
202 Enzymatic Activity at Polymer Tips Only We turn now to a consideration of the on and off rate constants at the (;( end of the polymer. The f3 end can be treated in the same way. The on and off rate constants for pure polymer 1 (i.e., pure E) and for pure polymer 2 (i.e., pure ES) are denoted (;(1' (;('1 and (;(2' (;(2' respectively. These are related to the two corresponding critical concentrations by [Eq. (10.35)J (20.5) We assume that (;(1 is the (diffusion controlled) rate constant for adding an E onto the (;( end whether the (;( tip subunit is an E or an ES. This implies, then, that the corresponding off rate constant includes the full intermolecular force effect, as follows. If an E departs from the (;( tip leaving behind another E at the tip (just as in pure polymer 1), the off rate constant is ex'I' Also, recall that ex'l is proportional to e w,likT because a 11 interaction has to be broken [see Eq. (5.8), where Ce = ex'/ex is proportional to e W / kT ]. However, if an E departs from the ex tip leaving behind an ES at the tip, the off rate constant will be proportional to e wulkT because a 12 interaction has to be broken. That is, the off rate constant in this case will be (20.6) Similarly, if an ES departs from an ES the off rate constant is ex2 whereas if it departs from an E the off rate constant is ex2Y21. The probabilities of the four kinds of ex end mentioned above (in the same order) are W ... 11 (1- ... 21 8(1 - 8) ... 22 82 .. . 12 8(1 - 8) . The rates of adding free subunits to the (;( end are (;(1 c(1 - 8s ) for E and ex 2c8s for ES. The net mean rate of adding subunits of either species (E or ES) to the (;( end is then on = Ja = on - off (20.7) ex 1 c(1 - 8s ) + ex 2c8s (20.8) off = ex'l(1 - 8)2 + ex'IY118(1 - 8) + ex282 + ex2Y218(1 - 8). (20.9) If desired, one can eliminate ex'l and ex2 from Eq. (20.9) by using Eqs. (20.5). Ja depends linearly on c (8 and 8s depend on Cs and C p but not on c). The value of c at which Ja = 0 is COO = a The superscript 00 off ex 1 (1-8s )+ex 2 8s refers to steady state, i.e., t = 00. (20.10) At this total concentration
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes 203 of E, the (I. end has zero mean growth. The concentration c:' is a steady-state analogue of an equilibrium Ceo Even though Ja = 0 at c = c:" the separate species (E and ES) do not, in general, have zero mean flux at this concentration. If we define Ja1 (and similarly for JaZ ) as the component 1 (i.e., E) terms in Eq. (20.7), then (20.11) Substitution of Eq. (20.10) for c in (1.1 c(1 - Os) leads to Ja1(C~) = -JaZ(c~) (l.1(1.~0(1 - 0.)[0 + Y21(1 - 0)] - (l.z(I.'10s(1 - 0)(1 - 0 (1.1 (1 - Os) + (1.2 Os + Yi10) (20.12) For example, if Os is small and 0 is near unity, it is possible to have a net flow of E on to the nongrowing (I. end and a net flow of ES off of the (I. end because of the compensating enzymatic activity on the polymer. This has a formal resemblance to "treadmilling," which is introduced in the next section. 21. Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes In Section 20 we discussed an enzyme that could carry out its activity when free in solution or, when aggregated, at any position in the polymer. However, nearest-neighbor interactions in the polymer perturb the rate constants in the enzymatic cycle. The model in this section l - 3 is much more specialized, or nonuniform. It was believed for a few years to apply to actin Z and microtubules, but is now generally thought to be too simple in both cases. The model is shown in general terms (i.e., for an arbitrary enzyme E that catalyzes S --+ P) in Fig. 7-2. Dotted lines represent transitions that are negligibly slow (in either direction). Arrows indicate the dominant direction in the diagram. Vertical lines are subunit attachment-detachment transitions. Only the polymer tip subunit (either end) is included in the figure because ES --+ EP is very fast at the tip, and interior (i.e., non-tip) subunits are assumed to be frozen in state EP. Thus, in this model, (a) no reactions occur in the polymer interior, (b) conversion of S --+ P does not occur in free subunits, and (c) exchange of S for P does not occur at the polymer tips. However, there is hybrid enzymatic activity (arrows) that makes use of some transitions from each ofthe two three-state (triangular) cycles as well as on and off (aggregation) transitions. In the hybrid cycle, on-off transitions are tightly coupled to S --+ P transitions. The net effect of one complete cycle is conversion of one S to one P in solution with no change in the amount of polymer aggregation: the thermodynamic force driving the
Enzymatic Activity at Polymer Tips Only 204 In Solution E /~" ~ G EP .•.•.•. .•... Off \ i. . . .J Fig.7-2. Kinetic diagram for an enzyme that operates in solution or on the tip of a linear aggregate. Solid lines show important transitions (either direction); arrows show dominant direction in composite cycle. The two dominant states are enclosed by squares. 0" ~. ES At Polymer Tip cycle is simply Ils - IIp. However, not all cycles are completed when subunits aggregate. Those subunits in the polymer represent only half-completed cycles (i.e., ES in solution to EP in the polymer). In this model, the presumption is that the transition S -4 P on E requires aggregation of ES in order to be activated (perhaps by a conformational change induced by the aggregation) and also that neither S nor P can escape from the enzymatic site once the subunit is aggregated (perhaps because of blockage by the next deeper neighbor in the polymer). As a further simplification, it is assumed that the only important states in Fig. 7-2 are ES in solution and EP at the polymer tip (i.e., EP and E are transient intermediates in solution and ES is a transient intermediate at the polymer tip). Thus the hybrid cycle (arrows) reduces, effectively, to a two-state cycle. If the assumption that ES -4 EP is very fast at the polymer tip (compared to on-off transitions) is not made, it is possible to have, on the polymer, a residual "cap" of ES units near the polymer ends that have not yet been converted to EP. This kind of model, 9 which is presumably more realistic for both actin and microtubules, is dealt with in Chapter 8. Throughout this section, the polymer is considered to be long. Finite polymers will be considered in Section 22. From this point on, we use notation appropriate to the special case that E is an NTPase so that the reaction S -4 Pis NTP -4 NDP + Pi. This will be abbreviated as T -4 D + P. In the case of actin, T is ATP, and in the case of microtubules, T is GTP. Following Section 11, we use the notation A(s), AT(s), AD(p), etc., to represent, respectively, subunits or monomers (a) in solution with no bound T,D, or P, (b) in solution with T bound, (c) on the polymer with D bound, etc. It should be mentioned that in the case of microtubules a subunit or "monomer" is actually a tubulin dimer. A four-state cycle for the hydrolysis of NTP by a monomer in solution is shown above the horizontal line in Fig. 7-3 (which is the four-state analogue of Fig. 7-2). This is the minimal cycle required to include all necessary mole-
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes Fig. 7-3. A possible NTPase cycle involving a combination of monomer states from free monomers in solution and from monomers on the end of a polymer. T = NTP; D = NDP; P = Pi. The squared species are dominant in the overall NTPase cycle. 205 In SolutIOn A -_-- /1 On AD •••••••••••• ~ •••• AJ).p .... •• A ••••• ~ ••/•••••••••• AT Off 5J .... .... : : - - _ - A D•P At Polymer Tip cular events. The same four-state cycle, at the bottom of the figure, refers to NTPase activity on a terminal (tip) monomer. The vertical transitions in Fig. 7-3 represent reversible attachment and release of monomers, in either the AD or AT form, from either end of the polymer. The hybrid six-state cycle that remains if we ignore the dotted lines, operating in the clockwise direction, is a complete NTPase cycle: NTP is hydrolyzed, and products are released. This complete cycle is put together from parts of two four-state cycles that are separately ineffective in hydrolyzing NTP. Incidentally, this feature of combining complementary partial enzymatic cycles, along with on <=± off transitions of the enzyme, is not novel. Essentially the same concept is used in most muscle contraction models: part of the complete myosin-ATPase cycle is traversed when myosin-ATP is free and part when it is bound to an actin monomer of a thin filament. We now simplify the above kinetic model considerably (as in Fig. 7-2) by assuming that AD and A in solution and AT and A D.p on either polymer end are unimportant transient intermediate states. Hence the only significant states remaining are AT in solution and AD on the polymer (as indicated by the boxes in Fig. 7-3): the six-state NTPase cycle becomes a two-state cycle. If AD in solution is also considered to be a significant species, 1 0 the reduced cycle is a three-state rather than a two-state cycle. Treatment of the threestate case 4 (omitted here) is more complicated but the final results are not very different. Figure 7-4 shows, schematically, the polymer with its two-state enzymatic activity at each end. The rate constant notation for the two different ends is shown in Fig. 7-4(a) while the corresponding NTPase events (from Fig. 7-3) are given in Fig. 7-4(b). Thus (Xl and /31 are second-order rate constants for the overall process (Fig. 7-3) that leads from AT in solution to AD on the polymer, (21.1) while (X2 and /32 are first-order rate constants that refer to the transition from
206 Enzymatic Activity at Polymer Tips Only AT(solution) "'~ ~"' AT(solution) :::(0 }, AD Polymer AD AD AD AD AD AD AD AD AD AD AD AD "t ~" :: ( [ AD ~ AT (solution) (a) Polymer AD (AD ) ( P Q AT (solution) (b) Fig. 7-4. (a) Rate constant notation for two-state NTPase cycles at the two ends f3 below) ofa steady-state polymer. (b) Gain and loss of ligands in the course of the two-state NTPase cycles. (Ill. above, AD on the polymer to AT in solution, (21.2) The inverse rate constants (with primes) are probably negligible under ordinary conditions, where the concentration of AD(s) is small compared to that of AT(S) and the hydrolysis of NTP on AT(P) is fast compared to the release of AT from the polymer. However, these inverse rate constants must be included in order to understand the connections between the kinetics and the thermodynamics of the problem. The two inverse pairs of rate constants a, a' and /3, /3', each operative at one end of an equilibrium polymer (Section 9), are replaced here by two NTPase cycles, with two inverse pairs aI' a'l and a2, a~ at the a end and two inverse pairs /31' /3~ and /32' /3; at the /3 end. There are now two biochemically distinct modes of attachment and detachment of monomers at each end (e.g., at the a end, a 1 and a~ for attachment and a z and a'l for detachment). Also, there is a thermodynamic force (the negative of the NTPase free energy of hydrolysis) dirving each of the NTPase cycles. Thus, there are altogether new conceptual features present in the steady-state polymer compared to the equilibrium polymer. z. 3 WegnerZ was the first to recognize that "treadmilling" (see below) is a consequence of these features.
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes 207 Thermodynamic Force and Rate Constants The eight individual rate constants for the processes shown in Fig. 7-4 are related to the free energy of NTP hydrolysis. In this subsection we examine this relationship explicitly. Because we make the simplifying assumption that the concentrations of Ao and A in solution are negligible, the free monomer activity a now refers to AT in solution. As usual, we denote the chemical potential of AT, at a, by /1I1.T /1XT = + kTln a. (21.3) Similarly, the chemical potentials of NTP, NDP, and Pi in solution are written /1T = /10 = /1p = + kTln aT /1'0 + kTln ao /1~ + kTln ap • /1~ (21.4) The only thermodynamic force driving the cycles in Figs. 7-3 and 7-4 is the free energy of hydrolysis of NTP. At activities aT' ao, and ap (the activity a of AT is not involved), this force is XT _ = /1T - /10 - /1p = 0 0 0 aT /1T - /10 - /1p + kTln--. aoa p (21.5) This quantity is usually of order 12-14 kcal mol-I. The standard free energy ofNTP hydrolysis is /1'0 + /1~ - /1~, which is of order -7 kcal mol-I. X T can obviously be varied by changing aT' ao, and a p • The monomers in the polymer are in state Ao. The chemical potential of these monomers is denoted /111.0' which is a property of the bulk polymer, not of its ends. The chemical potential /111.0 is the analogue of /10 in Section 4. Because AT occurs only in solution and Ao only in the polymer, we have omitted sand p in /1I1.T(S) and /1l1.o(p), above. We are now in a position to derive the fundamental relation between the NTP force X T and the rate constants in Fig. 7-4(a). We first consider the hypothetical but possible situation in which all transitions in Fig. 7-4(a) are blocked by inhibitors except the (inverse) a l and a'l processes, that is, AT(s) <=± Ao(p) + Pi. Ifthe activity of AT in solution is now adjusted (a == a~l») so that there is equilibrium between AT and the polymer, at some fixed activity a p of Pi' then we have (21.6) The only new feature here, compared to Eq. (4.1), is the extra /1p term on the left, which arises because, on attachment of a AT subunit to the polymer, AT becomes Ao (thus increasing the number of bulk Ao by one) and one Pi is released to the solution. Also, because there is a true monomer-polymer attachment equilibrium and detailed balance, we have al a~l) = a'l. If we use
Enzymatic Activity at Polymer Tips Only 208 this relation to eliminate a~1) from Eq. (21.6), we obtain (21.7) This is a relation between intrinsic molecular properties that obviously does not depend on the free monomer activity a. Although the condition a = a~1) was used as a convenient device to derive Eq. (21.7), the equation provides a property of the rate constants 1X1 and 1X'1 that is valid under any conditions, including steady states and transients. If we apply the same argument to the /31, /3~ pair at the other end of the polymer, the free energy difference on the right-hand side of Eq. (21.7) and hence a~l) [Eq. (21.6)] are necessarily the same as at the IX end because initial (AT) and final (AD, PJ states in the attachment process [Eq. (21.1)J are the same. (It is not necessary that the intermediate biochemical details be the same at the two ends.) Therefore, 1X 1/1X'1 = /3d/3~. This is, of course, equivalent to IX/IX' = /3//3' in Eq. (9.1), even though there is some biochemistry involved in the attachment process in 1X1' 1X'1 and /31 /3~ [Eq. (21.1)]. The key point here is that each end participates in the same equilibrium reaction and hence by detailed balance the equilibrium constant must be the same for both ends. As in Section 9, once a subunit (AD) has been added to the polymer, free in solution, at either end (IX or /3), by the same overall biochemical process, the polymer "cannot tell" which end was used-it is in the same state in either case. The 1X2' IX~ (and /32, /3;) transitions, for the process AD(p) + T ~ AT(s) + D, can be treated in essentially the same way. At given activities aT and aD' we start with /lAD + /IT = /l"t.T + /lD = /l~T + kTln a~2) + /lD' (21.8) instead of Eq. (21.6), and lX~a~2) = 1X2' and obtain (21.9) where X T is the NTP thermodynamic force defined in Eq. (21.5). By the same argument as above, we also have 1X2/1X~ = /32//3;. If we now add Eqs. (21.7) and (21.9), which apply to the two biochemically different [Eqs. (21.1) and (21.2)] parts of the NTP cycle, there results 1X11X2 /31/32 kTln-,-, = kTln/3' /3' = X T· 1X11X2 1 2 (21.10) These are the desired relationships between NTP cycle rate constants and XTo They hold under any conditions (transients, steady state, etc.); they are selfconsistency requirements of the two-state aggregation model we are using. Figure 7-5 expresses Eqs. (21.7), (21.9), and (21.10) graphically, in terms of
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes Fig. 7-5. Schematic basic free energy levels for the two-state NTPase cycle at either end of a steady-state polymer. kTIn "'I 209 i"'; = kTin {Jji{J'1 !\D(P) + P ji kTln"'2i"'2 XT = kTln{J2/{3; ~:~~c Energy r !\T(s)+P+D-T iii free energy levels for a single cycle. The total free energy drop XT in a cycle is broken down into its two component parts, for the two steps in the cycle [Eqs. (21.1) and (21.2)J. The levels i, ii, and iii in Fig. 7-5 may be termed "basic free energy levels," for convenience, because of their analogy to the basic free energy levels 11 introduced for the biochemical cycles of independent enzyme molecules. However, special treatment has been required here because of the aggregation of the enzyme molecule. Because X T is usually of order 12-14 kcal mol- 1 , e XT / kT is usually of order 109 or 10 10 . The separate dimensionless factors in ()(la.~ = /31 a .~ = ()('1 ()(~ a /3~ /3; a e XT /kT (21.11 ) are then perhaps of order 104 to 106 , though they obviously depend on a. Consequently, the reverse rate constants (primed) are presumably negligible for kinetic (not thermodynamic) purposes. Equation (21.5) shows that XT depends on the activities ofNTP, NDP, and Pi through the term kTln(aT/aDa p ). Thus, from Eq. (21.10), we have ()(1()(2 ()(1()(2 /31/32 aT oc - - . /31/32 aDa p -,-, = -,-, (21.12) Correspondingly, for the reactions (partial cycles) represented by Eqs. (21.7) and (21.9), or Fig. 7-5, (21.13) If a particular detailed biochemical scheme is adopted, such as the one shown in Fig. 7-3, then it is possible to relate the individual rate constants ()(1' ()('l' etc. of the two-state cycle (which considers certain states negligible) to the more elementary rate constants (not shown in Fig. 7-3) of the complete biochemical scheme. In this way the separate explicit dependences of ()(l, ()('l, etc., on aT' aD'
Enzymatic Activity at Polymer Tips Only 210 kTlIl "'] u/",'] = kTln {3]a/{3'] iii Ao(p)+P kTln "'2/"'2a XT = kTln {32 /{32 a Gross Free Energy r AT(s)+P+D-T iii' Fig. 7-6. Same as Fig. 7-5 but for gross (actual) free energy levels. and ap can be found. Examples have been published elsewhere 3 • 4 but we shaH not pursue this subject here. Comparison of Eqs. (21.10) and (21.11) suggests an alternative breakdown of the overall free energy drop X T into two constituents. If we rewrite Eqs. (21.7) and (21.9) as (or replace rx by 13) (21.14) (21.15) then these free energy differences represent the actual free energy changes in the two parts of a cycle at an activity a of AT(S). These new free energy levels are shown schematically in Fig. 7-6. By analogy with the treatment of in de pendent enzyme molecules,11 these levels might be called "gross free energy levels." The locations of the top and bottom levels in Fig. 7-6 depend on a but the difference between these two levels is always X T [Eq. (21.11)]. Figures 7-5 and 7-6 refer to the same processes and system; by adding kTln a to Il'h, we are simply changing from a standard chemical potential for AT(S) in Fig. 7-5 to the actual chemical potential in Fig. 7-6. Figure 7-6 brings out an important new property possessed by steady-state polymers but not by equilibrium polymers: the polymer can simultaneously be more stable than free subunits (via part 1, i' --+ ii' in Fig. 7-6, of the NTPase cycle) and also less stable than free subunits (via part 2, ii' --+ iii', of the NTPase cycle). That is (Fig. 7-6), IlAT(S) > IlAD(P) IlAD(P) + IIp(s) > + IIp(s) (part 1) IlAT(S) - X T (part 2). (21.16) This is possible because of the availability of the free energy of hydrolysis of NTP, - X T • Thus, in principle (with suitable controls), even at a single end of a steady-state polymer at a fixed activity a, the polymer end can aggregate spontaneously and also disaggregate spontaneously. As of this writing, it seems very likely that this is the fundamental basis of the transitory nature of
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes 211 many microfilaments (actin) and microtubules 12 , 13 in cell biology. Also, this property provides a justification for the involvement of an NTP in the aggregation of actin and tubulin; otherwise such an involvement would appear to be very wasteful. We shall return to this topic at the end of this section and especially in Section 25 of Chapter 8. Treadmilling Rate and NTP Flux The rate of tread milling (defined below) and the rate of NTP hydrolysis can be related to the individual rate constants. It is simplest to consider these rates as a composite of contributions from the four individual reversible pairs of transitions [Fig. 7-4(a)]. These contributions are introduced by means of the illustration in Fig. 7-7, where several fluxes are plotted against the monomer activity. The two pairs of broken lines in Fig. 7-7 correspond to reversible transitions between AT(S) and Ao(p) and are analogous to the Ja , Jp pair of lines in Fig. 3-4. Because a'l and /3~ are very small, the lines a 1a - a'l and /31 a - /3~ intersect at a~1) ~ O. Correspondingly, because a~ and /3; are very small, the lines a;a - a2 and /3;a - /32 are almost horizontal and intersect at a very large value of a, a~2). As in Fig. 3-4, flux curves for the same reaction at the two ends of the polymer must intersect at the same critical activity. The rate of adding monomers at the a end of the polymer is simply the sum of the fluxes for the a 1, a'l and a2, a; reactions: ao ./1: a~ a (3'2 a - 2 (3 , , ( a(2) : " / '''/ ----------------------Y / : / , :/ ----------~~~~---------~/ Fig. 7-7. Various equilibrium and steady-state fluxes, as described in the text, for a steady-state polymer with free ends. The critical activity a~l) is very small, but not zero. Note the scale change needed because a~2) is very large.
Enzymatic Activity at Polymer Tips Only 212 (21.17) (21.18) where Eq. (21.18) is presumably a very accurate approximation, as explained above. la is shown as a solid line in Fig. 7-7, with slope (ad the same as for the line ala - a'l and intercept (-a 2) the same as for the line a~a - a 2. Similarly, at the f3 end, lp = (f31 + f3;)a - (f32 + f3'1) (21.19) (21.20) ~f31a-f32· lp is also included as a solid line in Fig. 7-7, with properties analogous to la. la = 0 at a = aa' where aa ~ a 2/a 1. This can be called the critical activity for the a end. Also, lp = 0 at a a p, where a p ~ f32/ f31, which is the critical activity for the f3 end. Unlike a~l) and a~2), aa and a p are not equilibrium properties because a 1 and a 2 do not pertain to inverse transitions, nor do f31 and f32 [see Fig. 7-4(a)]. There are no thermodynamic relations that determine the ratios a2/a 1 and f32/f31 [compare Eqs. (21.7) and (21.9)]. Whereas a~l) and a~2) have extreme values, aa and a p are in the measurable range, usually of order 1 tLM (a 1 and f31 are of order 10 tLM-1S-1; a 2 and f32 are of order 10 S-l). Because aa < a p, the a end in the illustration in Fig. 7-7 is the so-called + end of the polymer and the f3 end is the - end: in the free monomer activity range of primary interest, aa .::;; a .::;; a p, la is positive (monomers add to the + or a end) and lp is negative (monomers depart from the - or f3 end). Thus, in this activity range there is a net flux of monomers through the polymer. = The net total rate of addition of monomers to the polymer is 1 =la + lp (21.21) ~ (a 1 + (21.22) (31)a - (a 2 + (32)· This is included as a solid line in Fig. 7-7. 1 intersects lp at aa, where the rate of addition to the a end is zero. 1 intersects la at ap , where the rate of addition to the f3 end is zero. 1 = 0 at a = aoo , where la = -lp (vertical line). That is, this is the definition of aoo • Necessarily, aa < a oo < a p. Because 1 = 0 at a = aoo , the mean number of subunits in a polymer remains constant with time, at this activity (except for large fluctuations-see Section 22), though monomers are being added at the + (a) end and are being lost at the - (f3) end. This phenomenon 2 is usually referred to as treadmilling (or head-to-tail polymerization). The explicit expression for a oo , which is a joint steady-state property of both ends of the polymer, from Eqs. (21.17), (21.19), and (21.21), is a = 00 a2 + f32 a 1 + f31 + a'l + f3~ + a; + f3; --=---'--=------=---'----=- (21.23) (21.24)
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes 213 If in a closed system with many polymer molecules we start with a > aO'), there will be net growth (1 > 0) of the polymers and the free monomer activity a will decrease to aO'). Similarly, if a < aO') at the outset, there will be net loss of monomers from the polymers and a will increase to aO'). Thus aO') is the stable value of a in a closed system of this type. We shall usually use the term "treadmilling" to refer to the particular case 1 = 0 (polymer of constant mean length). But, from a more general point of view, treadmilling is a meaningful concept in the entire monomer activity range aa < a < a p, where 1a is positive and 1p is negative. In this range, excluding fluctuations, a monomer added to the If. end will make its way through the polymer and leave at the f3 end. However, the monomer itself is not moving; rather, the If. end is growing and the f3 end receding. The action is similar to that of the caterpillar tread on a tractor. The rate at which a monomer, newly added at the If. end, approaches the f3 end is called (here) the treadmilling rate, 1m. This is, of course, also the rate at which the f3 end approaches the added monomer, that is, the rate at which the f3 end recedes. Thus Jm (in units of monomers S-l per polymer molecule) is defined as -Jp, in the interval aa < a < a p. Hence 1m ~ f32 - f31 a. This is shown as a heavy line in Fig. 7-8, which represents the same hypothetical system as in Fig. 7-7. The value of 1m at a = aO') is of special interest; this is denoted 1;:; (also called the monomer flux). Here the "tractor" maintains a tread of constant length. At a = a(X" Fig. 7-8. Some fluxes taken from Fig. 7-7, together with additional fluxes defined in the text.
214 Enzymatic Activity at Polymer Tips Only J:;: == -Jp(a oo ) = Ja(a oo ) ((l.1PZ - (l.zPd(l - (21.25) e-XTlkT) (1.1 + PI + (1.2 + P; '" (1.1 pz - (l.ZP1 (1.1 + PI (21.26) (21.27) = where Eq. (21.26) follows on substituting Eq. (21.23) into Eq. (21.17) or (2-1.19), and using (1.1 PI (l.z a~ p~' a~ pz p;' (21.28) and Eq. (21.10). The treadmilling rate J:;: [Eq. (21.26)] is zero if the two ends of the polymer are alike ((1.1 = P1'(l.Z = (2) or if (hypothetically) aT' an, and a p have values such that X T = 0 (i.e., at NTP hydrolysis equilibrium). Thus, a nonzero NTP driving force XT is a necessary condition for steady-state treadmilling in solution. The approximate Eq. (21.27) follows because usually (1.2' P;, and e-XTlkT (of order 10- 1 °) are all negligible. J:;: is necessarily positive because we chose aa < ap: If a polymer is either growing or shrinking, the total net flux (from both ends) for subunits participating in part 1 of the NTP cycle, J10 involving the addition of AT(s) [Eq. (21.1)], is not equal to the total net flux, J 2 , in part 2 of the cycle, involving the dissociation of An(P) [Eq. (21.2)]. The definitions of the fluxes corresponding to the two parts of the NTP cycle are: + (PIa - PD == Ja1 + Jp1 (l.2 a) + (P2 - p;a) == Ja2 + Jpz J 1 == ((1.1 a - (I.'d J2 == ((1.2 J = J 1 - Jz. (21.29) (21.30) The subdivision into Ja1 , etc., will be needed in the next subsection. Under conditions when the polymer is either growing or shrinking and J1 i= J2 , the NTP flux, JT, is defined·as the lesser of J 1 and Jz , because it is only this amount of the flux that refers to completed NTP cycles (i.e., part 1 and part 2). In the usual excellent approximation, we have (21.31) The J 1 flux is proportional to the free monomer activity; the Jz flux is constant. These two lines are included in Fig. 7-8. Because the NTP flux, JT , is defined as requiring a complete NTP cycle [addition of AT(S) and removal of An(P)], JT increases with monomer activity until aoo • At this point subunits add and come off at the same rate. Above this activity JT follows Jz: subunits are being added faster than they come off; JT is independent of monomer activity. The
Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes 215 heavy portions of the lines J 1 and J2 in Fig. 7-8 represent the two branches of JT . Explicitly, (21.32) The superscript (-) refers to the shortening case, whereas (+) refers to lengthening. The first branch of JT, J+-l, is parallel to J. It is simple to verify from the definitions that J 1 = J2 when JI1. + Jp = 0; that is, J 1 = J2 at a = aoo. At a = aoo, JT is denoted J;: J; = J 1(a oo ) = Jz(a oo ) (a 1 + /3d(a z a1 ~ az + /3z)(l - e-XTlkT) + /31 + a~ + /3; + /3z, (21.33) (21.34) (21.35) where we have used Eqs. (21.10), (21.23), (21.28), and (21.29). If the two ends of the polymer are the same, J; is still positive (unlike J:;:, which is zero in this case). However J; = 0 when X T = 0 (NTP hydrolysis equilibrium). In the example shown in Fig. 7-8, based on Fig. 7-7, the flux of subunits through the polymer, Jm , is significantly smaller than the number of complete cycles of assembly and disassembly, JT • The ratio J:;: s == - j; a 1/3z - a z /31 (a 1 + /31)(a Z + /3z) = ------- (21.36) at aoo is of particular interest. It should be noted that Eq. (21.36) follows from Eqs. (21.26) and (21.34) without the usual approximation that the reverse transitions can be neglected. A similar relation was obtained by WegnerZ without including reverse transitions, but was expressed in terms of aoo. Because the terms a 1/3z and a z /31 appear in both numerator and denominator in Eq. (21.36), necessarily s < 1. One can regard s as a kind of kinetic (not thermodynamic) efficiency: the treadmilling rate of the subunits (at a = aoo) relative to the total rate of NTP turnover at both ends of the polymer. In the steady-state kinetics of independent enzyme molecules with multicycle kinetic diagrams,ll a considerable conceptual simplification is realized if one regards the observable fluxes as being made up of separate contributions from the various cycles of the diagrams. The same is true here, at a = aOO" We have so far referred to the two NTP cycles at the ends of the polymer. Actually, there are four NTP cycles, as shown in Fig. 7-9. Cycles c and d make mixed use of the two ends. All of these cycles have the same force, XT • The relations between force and rate constants are (using the cycle designations in Fig. 7-9): (21.37)
Enzymatic Activity at Polymer Tips Only 216 Polymer c "'1 (3) ~ i32 r-.. ~I "'2 I I c ~ "') Fig.7-9. Component cycles that contribute to the steady-state NTPase flux. i32 .....--.... I a ~ d '--'" (3) I The flux J; in Eq. (21.34) may now be regarded as a superposition of separate contributions from the four cycles: + J b + Jc + Jd lX'llX~)/D, J b = Uh f32 (21.38) J; = Ja Ja = (1X11X2 - Jc = (1X1f32 - lX'lf3;)/D, D == 1X1 - f3~ f3;)/D Jd = (1X2f31 - 1X~f3~)!D (21.39) + f31 + IX~ + f3;. The general form of the connection between Eqs. (21.37) and (21.39) is conventional; 11 only the composition of D is different (because of enzyme aggregation here). Neglecting reverse rate constants, the relative contributions of the separate cycles a, b, c, and d to JT are proportional to 1X11X2' f31f32' 1X1f32' and 1X2f31' respectively. Furthermore, we see from Eq. (21.26) that (21.40) In view of Fig. 7-9, this is just what we should expect: cycles a and b make no contribution to the tread milling motion or rate; cycle c makes a positive contribution (a subunit is added at the IX end and another is removed at the f3 end); and cycle d makes a negative (wrong-way) contribution. Rate of Dissipation and Storage of Free Energy We return now to Fig. 7-6 for more detailed discussion. Figure 7-10(b) is an extension of Fig. 7-6 in which the free energy levels are plotted as functions ofln a (one particular value of a was chosen in Fig. 7-6). For comparison, Fig. 7-1O(a) shows the corresponding plot for an equilibrium polymer [see Eq. (4-1) for notation]. Figures 7-7 and 7-8 should also be compared with Fig. 7-1O(b): Figs. 7-7 and 7-8 refer to the same system as in Fig. 7-1O(b) but various fluxes are plotted as functions of a (rather than In a). In Fig. 7-1O(a), the polymer is either less stable than, more stable than, or in equilibrium with free subunits, depending on the value of a. The situation in Fig. 7-10(b) for the steady-state polymer is more complicated. For values of a between a~l) and a~2), the polymer is both more stable than free subunits (via part 1 of the NTPase cycle) and less stable than free subunits (via part 2 of the cycle). The definitions of !J./11 and !J./12 are obvious from the labels in
217 Enzymatic Activity at Polymer Tips Only: Bioenergetics and Fluxes E c: " o "'OJ u E ..c: " U Subunits In a-+ (a) (b) Fig.7-10. (a) Plot offree subunit and polymer chemical potentials, for an equilibrium polymer, as functions of In a. (b) Same for the steady-state polymer of Figs. 7-6, 7-7, and 7-8. the figure (see also Fig. 7-6): 11/11 = (/1h 11/12 + kTln a) - (/1AD + /1p) = (/1AD + /1p) - (/1h + kTlna 11/11 + 11/12 = XT• XT) (21.41) (21.42) (21.43) These expressions apply, of course, at both ends of the polymer. Between a~l) and a~2), both 11/11 and 11/12 are positive (because a~l) is very small and a~2) is very large, this property holds essentially over the entire range of a of any interest). Figure 7-10(b) is a strictly thermodynamic diagram. Figure 7-7,
Enzymatic Activity at Polymer Tips Only 218 which has to do with related kinetics, shows that aa' aaJ' and ap all fall between a~l) and a~Z), in the order given. For a between aa and a p, the CI. end grows and the f3 end shortens. At the CI. end, Ja1 [Eq. (21.29)J > 0 (a gain of subunits) because LlJ11 > 0 and JaZ > 0 (a loss of subunits) because LlJ1z > o. Both are spontaneous processes, in accordance with the second law of thermodynamics. The net Ja = Ja1 - JaZ [see Eq. (21.30)J is positive; the CI. end grows. To the extent that the CI. end has net growth, free energy is being stored in the CI. end subunits because part 2 of the cycle, with LlJ1z > 0, has not occurred yet for these subunits. At the f3 end, Jp1 > 0 (a gain of subunits) because LlJ11 > 0 and J pz > 0 (a loss of subunits) because LlJ1z > o. The net J p = Jp1 - Jpz is negative (the f3 end shortens). To the extent that the f3 end shortens, there is an extra dissipation of free energy (LlJ1z per subunit) from previously polymerized subunits. For a > ap, both ends grow and store free energy as above for the CI. end. For a < aa' both ends shorten and dissipate extra free energy as above for the f3 end. The total rate of free energy dissipation, at any a, is d·S Tit + JpdLlJ11 + (JaZ + JPZ )LlJ12 = (Ja1 = J1LlJ11 + J2 Ll J1Z· (21.44) (21.45) In the special case a = aaJ' where the polymer has no net growth and as many cycles are completed (part 2) as started (part 1), Eq. (21.45) simplifies, as expected, to (21.46) in view of Eqs. (21.33) and (21.43). Incidentally, it is easy to show that each of the four terms in Eq. (21.44) is always positive or zero, as required by the second law. We use the first term as an example. We have (21.47) (21.48) where the LlJ11 expression follows from Eqs. (21.7) and (21.41). Obviously, then, Ja1 and LlJ11 always have the same sign and hence their product is positive (or zero at a = a~l»). The overall rate of storage of free energy (positive or negative) in the polymer, in the sense used above, can be written in several ways: J LlJ12 = (J1 - J2)8fJ,2 = J 1X T - (J1 LlfJ,l There is zero storage at a = aoo , J = O. + J28J12)· (21.49)
Enzymatic Activity at Polymer Tips Only: Length Distributions and Transients 219 The above discussion applies to the "fast hydrolysis at the tip" model being used in this section. In Chapter 8, where delayed hydrolysis is introduced, this topic takes on more interest and significance. Finally, we mention that a steady-state polymer with one end anchored, or treadmilling between restraining barriers, can convert some of the NTP free energy of hydrolysis into mechanical work, if a resisting force is attached to the subunits of the treadmill. In this case, not all of JTXT is dissipated, as in Eq. (21.46), which applies to free treadmilling polymers in solution. These topics are discussed in Ref. 1 but they are not included here because they are probably of only academic interest. 22. Enzymatic Activity at Polymer Tips Only: Length Distributions and Transients The previous section was concerned with long polymers. Here we consider finite polymers of the same type. There are two subtopics: polymers attached at one end and free polymers. Attached Polymers In the model in Fig. 7-4, we now assume that the f3 end is attached and the a end is free. There is no subunit exchange at the f3 end. The number of subunits in the polymer is N. Figure 2-4 is the corresponding kinetic diagram for an equilibrium polymer. Here there are two on-off routes, via part 1 of the cycle and via part 2 of the cycle (Fig. 7-4). Usually a'l and a~ are negligible. The two off rate constants when N = 1 (i.e., departure from the surface attachment site itself) are taken as a'l/C l and a 2 /C 2 , corresponding to a'IC in Fig. 2-4. Then the kinetic diagram for the attached steady-state polymer is that shown in Fig. 7-11. The rate constants are unchanged for N > 2. Because the diagram is linear, there is a "detailed balance" at steady state. The analogues of Eqs. (5.15) and (5.16) are (22.1) (<>1 N= 0 " + <>~)a Q'; -+(\'2 C2 (<>1 0( (\'2 + <>2)a • 2 ... + Q'i C1 Fig. 7-11. Kinetic diagram for an attached steady-state polymer. See text for details.
220 Enzymatic Activity at Polymer Tips Only (22.2) (22.3) In Eq. (22.1), aa follows from Eq. (21.17) when Ja = O. If the primed rate constants are negligible, (22.4) Because of the formal agreement between Eqs. (22.1) and (22.2) on the one hand and Eqs. (5.15) and (5.16) on the other, Eqs. (5.5), (5.6), and (5.10)-(5.13) all hold for the steady-state polymer at steady state. Thus the probability distribution PN formally is the same as for an equilibrium polymer. However, it should be emphasized that these are not equilibrium relations and they cannot be deduced from a partition function, as in Section 5: the only method available is steady-state kinetics (Fig. 7-11). In Eq. (22.4), aa ~ a z /a 1 is the critical concentration (x --+ 1) for formation of the attached infinite steadystate polymer (IV --+ 00), but a 1 and a z are not inverse rate constants: a 1 and a z refer to two different chemical processes [Figs. 7-3 and 7-4(b)]. Unlike ala' in Eq. (5.15), a 1 /a z is not equal to an equilibrium constant (but ada'l and az/a~ are). In brief, a 1 and a z are independent rate constants. Specification of the kinetic diagram, as in Fig. 7-11, is equivalent to specification of the differential equations for dPNldt, as in Eqs. (6.1) and (6.2), for use in transients. In fact, all of Section 6, except Eqs. (6.12)-(6.14), applies to an attached steady-state polymer if we take C notational changes: = 1 and make the following (22.5) Although the distinction between equilibrium and steady-state polymers should be kept in mind, the fact remains that steady-state polymers with fast two-state enzymatic activity at the tips behave formally like equilibrium polymers, both in steady properties and in transients. This will be seen again in the next subsection. Free Polymers The principal simplification we make here is that the transitions of the twostate cycles in Fig. 7-4 can be treated as elementary transitions in assigning the N-dependence of the rate constants. This is an approximation, though probably a good one, because the true elementary transitions are (for example) those shown in Fig. 7-3. Equilibrium polymer N-dependences, as in Eqs. (9.14) and (9.15), can be applied to the true elementary transitions but these de-
Enzymatic Activity at Polymer Tips Only: Length Distributions and Transients 221 pendences become more complicated when the elementary transitions are "collapsed," as in Fig. 7-4, because of assumed transient intermediates. This complication is dealt with in Ref. 4 but we omit it here. We begin, then, by assuming that Eqs. (9.14) and (9.15) apply to each of the four on-off inverse pairs of rate constants in Fig. 7-4(a). N effects (i.e., size effects) are, of course, the same at the a and [3 ends and for parts 1 and 2 of the cycles. The equilibrium polymer of Eqs. (9.8), (9.9), (9.14), and (9.15) has a kinetic diagram, including both ends, of the form ... N-l( + [3) [N]a 'N ... , + [3') [N] (N) (a (a' (22.6) where [N] == 1 + In(N - 1) N -1 (N) (22.7) (N- l)n == ------;:;- . (22.8) The individual rate constants in Eq. (22.6) are related by K = a/a' = [3/[3'. Correspondingly, with the assumption mentioned at the beginning of this paragraph, the kinetic diagram for the steady-state polymer, including both ends, is + a~ + [31 + [32) [N] a IN. . . . (a 2 + a'l + [32 + [3~)[N](N) ... N - 1 ( (a 1 (22.9) The rate constant relations here are [Eq. (21.28)] [31' ' (22.10) a ,2 The equilibrium polymer has the same critical activity a' [3' a' + [3' 1 a =-=-=---=e a [3 a+[3 K (22.11) at both ends, but this is not true of the steady-state polymer (Fig. 7-7): aa = a 2 + a'l [32 , # a p = [3 a1 + a2 1 + [3~ + [3'2 . (22.12) Although relationships among rate constants are less simple in the steadystate case, the steady-state PN distribution for free polymers will have the same formal appearance 3 ,4 as in the equilibrium case [Eqs. (8.31) and (9.10)] because Eqs. (22.6) and (22.9) differ only by the notational changes + [3 -+ a 1 + a~ + [31 + [3; a' + [3' -+ a 2 + a'l + [32 + [3~ a ~ a 1 + [31 (on) ~ a 2 + [32 (off). (22.13)
222 Enzymatic Activity at Polymer Tips Only Thus we have P (N)n N - 1 x, N PN - 1 = a aGO (22.14) X=- (22.15) a2 + a'l aGO = a1 + /32 + /3~ a2 + /32 + a~ + /31 + /32 ~ a 1 + /31 . (22.16) The critical activity aGO was introduced in Eq. (21.23). The various properties of the PN distribution are the same as found in Section 8. Again we see that this model of a steady-state polymer exhibits quasiequilibrium properties. The mean value of anyone of the on rate constants in Eq. (22.9), say a 1 , averaged over N at steady state, is (22.17) Similarly, for anyone of the off rate constants, say a 2 , a2 = a2 I N [N](N)PN = a 2x I N [N]PN- 1, (22.18) where the last form follows from Eqs. (22.8) and (22.14). Consequently, if we sum [Eq. (22.9)] (a 1 + a~ + /31 + /3'z)[N]aPN- 1 = (a2 + a'l + /32 + /3~)[N](N)PN (22.19) over N, we find (a 1 + a~ = a2 + /31 + /3'z)a + a'l + /32 + /3~ = + /31 + /3'z)a I [N]PN- 1 + a'l + /32 + /3Dx I [N]PN- 1· (a 1 + a~ = (a2 N N (22.20) The two expressions on the right lead to Eq. (21.23) (as a check) and the two expressions on the left give (22.21) which is the analogue of Eq. (21.23) when a < aGO (finite polymer). Beginning on p. 95, Section 9 treats a few transient problems for free equilibrium polymers. Using Eqs. (22.13) and ae -> aGO' these results can also be taken over for steady-state polymers. The subsection on "Rate of Label Loss from Polymer" (p. 99) is an exception in that the two ends must be
223 Fluctuations in the Polymer Length Distribution considered separately (because aa oF ap for a steady-state polymer). The appropriate discussion of this case is given in Ref. 1, pp. 104-108 and will not be repeated here. 23. Fluctuations in the Polymer Length Distribution We have repeatedly discussed equilibrium or steady-state probability distributions PN • The problem we consider here is the following (the discussion applies to a wide class of probability distributions; it is not limited to polymers). Suppose PN is the normalized probability of observing N for a given polymer or system. For an infinitely large ensemble of systems, PN would be the fraction of systems in the ensemble with N. The mean of the distribution PN is N and the variance is N 2 - N2 • Now in a particular experiment, suppose only a finite sample M of these polymers or systems is observed, say M = 200 or 300 rather than M = 00. In this sample, let the number of systems with k be Mk (we use k as the index for the finite sample M, and N as the same index for the "true" or M = 00 distribution). Let M represent the complete set of numbers MkCLkMk = M). For the set M, the mean and variance are denoted I< and v. A practical question is: how much are I< and v likely to differ from N and N 2 - N2 , respectively? It seems unlikely that these questions are new. We merely sketch the necessary derivations. For a sample of size M, the normalized probability of observing the particular set M is p(M) = M!Il pr' Il kMk! . (23.1) k If M k, Mt, etc., are averaged over all sets M, we find the well-known results Mk = L MkP(M) = MPk (23.2) Mt = MPk + M(M - 1)Pk2 (23.3) MkMI = M(M - 1)PkPI (k of- l) (23.4) M etc. To answer the first question above, we average (I< - N)2 over all sets M: LP(M) ( L kMk- M k M _)2 N = I L kMkM l -2- k,l M N 2• (23.5) The sum is over all k and l. We can use the expression in Eq. (23.4) in this sum,
224 Enzymatic Activity at Polymer Tips Only but there is an extra contribution M Pk from Eq. (23.3) when k = ,. Thus we find (k _ N)2 = N 2 ~ jlP This is the simple final result for the mean value Eq. (8.36), (23.6) k. In our polymer example, 1- (k - Nf (n N2 (23.7) + I)M Thus, if M = 200 and n = 5, the root-mean-square deviation of k from jJ is 0.029N. Turning now to the variance, we first calculate the mean of v(M), averaged over all sets M: = N2 - I k,l klMkMl M2 (23,8) (23.9) Our primary interest is not in P itself but in the variance in the variance, v2 - p2. For this, we need v2 : (23.10) These sums are handled essentially as above (though they are more complicated), taking due care whenever two or more indices are equal. We find for the three sums: (23.11) I klmn = (M - I)(M - 2)(M - 3) jJ 4 M3 ( ) + 6(M - I)(M - 2) N2jJ2 M3 (23.13) The final result can be put in the relatively compact form (for M ~ 3)
225 Fluctuations in the Polymer Length Distribution 2 V - -2 1 (1- M 1 )2 (N - N) 4 - M 1 (1- M 1) ( 1 - M 3 ) [(N - N) 2] 2 V = M (23.14) (23.15) Using the distribution in Eq. (8.31), v2 1)2 - 2M- 1 [en + 4) - v2 ~ 2(n - M(n M-1(n + 7) (1-Ml)2(n+l) + 3M- 2] + 4) + 1)" (23.16) (23.17) For example, if M = 200 and n = 5, the root-mean-square deviation from is 0.1221) = 0.122(N2 - jJ2). A slightly different quantity is I) (23.18) This is the mean square deviation of v from the true variance. This is easily shown to be the same as v2 - 1)2 in Eq. (23.14) except for a small added term (23.19) which makes no contribution to Eq. (23.15). A final deviation we consider, from the true PN , is the average over all sets M of Lk(Mk - MPk)2. This is the sum of the squares of the differences, over the entire distribution (all k), between each Mk and the corresponding "true" value MPk • This average is easily found to be L p(M) L (Mk M MPk)2 = M(l - k L P 2). k (23.20) k To normalize, both sides of the equation should be divided by M2, because the absolute magnitude of the deviation Mk - M Pk should be compared with M Pb the sum of which over k is M. With the distribution in Eq. (8.31), we find 1 + 1)lnX = [rcn + 1)]222n+l . rc2n f " 2 Pk (23.21) Because In (l/x) is a small quantity, this sum is almost negligible in Eq. (23.20). For example, if n = 5, "L.,.Pk2 k 1 =0.1231n-. x If x = 0.95, the right-hand side of Eq. (23.20) is 0.994M. (23.22)
Enzymatic Activity at Polymer Tips Only 226 References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. Hill, T.L. and Kirschner, M.W. (1982) Int. Rev. Cytol. 78, 1. Wegner, A. (1976) 1. Mol. BioI. 108, 139. Hill, T.L. (1980) Proc. Natl. Acad. Sci. USA 77, 4803. Hill, T.L. (1981) Biophys. 1. 33, 353. Hill, T.L. (1981) Proc. Natl. Acad. Sci. USA 78, 5613. Hill, T.L. (1982) Proc. Natl. Acad. Sci. USA 79, 490. Hill, T.L. and Kirschner, M.W. (1983) Int. Rev. Cytol. 84, 185. Hill, T.L. (1985) Cooperativity Theory in Biochemistry (Springer, New York). Carlier, M.F. and Pantaloni, D. (1981) Biochemistry 20,1918. Niedl, C. and Engel, 1. (1979) Eur. 1. Biochem. 101, 163. Hill, T.L. (1977) Free Energy Transduction in Biology (Academic, New York). Mitchison, T. and Kirschner, M.W. (1984) in Molecular Biology of the Cytoskeleton, eds. Borisy, G.G., Cleveland, D.W., and Murphy, D.B. (Cold Spring Harbor Laboratory, Cold Spring Habor, NY), p. 27. Hill, T.L. (1985) Proc. Natl. Acad. Sci. USA 82, 4404.
8 NTP Caps and Possible Phase Changes at Polyltler Ends The principal topic in Chapter 7 was a treatment of aggregation of NTP subunits accompanied by fast NTPase activity at the polymer tips. Actin and micro tubules were believed, at one time, to behave in this relatively simple way. As of this writing, it seems clear that, in both cases, NTP subunits actually penetrate (or survive) into the polymer ends by virtue of addition of subsequent subunits: conversion of an added NTP subunit into an NDP subunit is in fact not fast compared to on-off transitions.! Thus there may be a collection or "cap" of surviving NTP subunits at each polymer end. Though this much seems clear, further details are in the process of being worked out and are by no means generally agreed upon. For this reason, in this chapter we bypass biochemical details (except for illustrative examples in Section 24) and devote the bulk of the chapter to a consideration of two-phase behavior at the polymer ends. The two phases referred to are a polymer end either with or without an NTP cap. This subject can be dealt with without a commitment to a particular detailed biochemical model; we merely assume that some unspecified biochemical mechanism exists that generates two-phase activity (see Section 24 for examples). Actually, so far there is evidence for two-phase activity in microtubules 2 - 4 but not in actin. Theoretical aspects of two-phase behavior at polymer ends are treated in Sections 25 and 26. Section 24 provides an introduction to this subject. Section 27 is essentially an appendix that deals with a related topic.
228 NTP Caps and Possible Phase Changes at Polymer Ends 24. Illustrative Biochemical Models that Generate Phase Changes We shall designate a subunit with an NTP bound by T and one with an NDP bound by D. We assume that the only important free subunits are Ts, as in Chapter 7. If Ts newly attached to a polymer end do not hydrolyze (T --+ D) quickly, which is what we postulate throughout this chapter, a polymer end may contain a number of still surviving Ts. The interior of a long polymer, however, would be all Ds because the reverse rate constant for D --+ T is considered to be negligibly small as is the rate constant for interior exchange ofNTP for NDP, which also would convert D to T. The Ts at a polymer end will be referred to as an NTP cap, or simply as a cap. There are two especially simple assumptions that can be made about NTP hydrolysis in this cap: (a) hydrolysis (T --+ D) occurs at a significant rate only at the boundary (deepest T) between a solid cap of Ts and the all-D interior or (b) hydrolysis is equally likely to occur at any surviving T in the cap (hence the cap will not be a solid-T cap: Ds will be scattered among the Ts). To illustrate two-phase behavior at a polymer end, we consider first a simple model of type (a) above that can be handled analytically and then turn to three models of type (b) (two of which are somewhat more complicated than indicated above). As an aside, it should be mentioned that the NTP cap might, in fact, be an NDP' Pi cap. That is, when an NTP subunit adds to a polymer end, hydrolysis per se {NTP --+ NDP' PJ may indeed be fast but with the subsequent release of Pi (NDP' Pi --+ NDP + PJ relatively slow. Such appears to be the cases for actin' A TP at least. For brevity and generality, however, we shall continue to refer to polymer Ts and to an NTP cap. Hydrolysis at the Cap Boundary Only This simple, pedagogical model was suggested by the Appendix of Ref. 6 and by Ref. 7 and 8. No real system is known to behave in precisely this way. We consider the a end of a long, effectively single-stranded, polymer with a solid string of N Ts at the a end, where N = 0, 1, 2, .... The polymer interior (beyond the cap) is comprised of solid Ds. When N = the entire polymer is D (except possibly for a cap at the f3 end, which we are not considering). The successive positions at the a end ofthe polymer are designated n = 1,2, ... [Fig. 8-1 (a)]. Hydrolysis (T --+ D) can occur only at position n = N, provided that N ? 1. At position n = 1 only (i.e., at the polymer tip), when N = (i.e., when the polymer is all D), D can be converted back to T by exchange ofNTP for NDP: ° ° subunit· NDP --+ subunit --+ subunit· NTP (n = 1). (24.1) This is the same exchange reaction that is assumed (Chapter 7) to occur very rapidly whenever a D leaves the polymer end to become free in the solution.
229 Illustrative Biochemical Models That Generate Phase Changes n ~ 1 2 3 4 5 6 f }N~3 D D D N~ alDa+k x 0 « alTa • alTa 2 a'm + Ko clIT + K (b) (a) N~ 0 --Phase 2 ~ --:--- - :k --:--- 2 « --- • 3··· a~T + K 3··· Phase 1 k' : (e) Fig. 8-1. (a) The IX end of a polymer with three Ts (N = 3) at the end. (b) Kinetic diagram for the polymer IX end in terms of N values. (c) Definition of phases and of rate constants between phases. We now list all possible transitions in the model, along with the corresponding first-order rate constants. All inverse (primed) rate constants are included, for thermodynamic completeness, though some of these can be ignored for kinetic purposes. The list follows: 1. 2. 3. 4. 5. 6. Hydrolysis at n = N = 1: Ko; inverse K~ Exchange at n = 1, N = 0: k x ; inverse k~ Hydrolysis at n = N ~ 2: K; inverse K' Addition of a T to a D (N = 0): IXlDa; inverse lX'lD Addition of a T to a T (N ~ 1): alTa; inverse a'lT Loss of a D from a D (N = 0): a 2; inverse a~a. (24.2) The notation for a rate constants is a generalization of Fig. 7-4(a). As usual, a is the activity of free T. The kinetically negligible rate constants are K~, k~, K', and a~. Required thermodynamic relations among the complete set of rate constants, corresponding to Eq. (21.10), are Kok x a 1o K o a 2 K~ k~ a'lD K~ a~ (24.3) Note that position n = 1, with N = 0 or 1, is special: the hydrolysis rate constant Ko (N = 1) is different from K (presumably Ko < K or Ko « K); the exchange reaction (kx' D --+ T) can occur only at n = 1 in the all-D polymer (N = 0); and the rate constant for addition of a T to a D (N = 0) is different from that for the addition of a T to a T (N ~ 1) (presumably a lD « a lT ). Position n = 1 is, in effect, the nucleating site for formation of a cap and state
230 NTP Caps and Possible Phase Changes at Polymer Ends N = 1 is a nucleus preceding formation of a significant cap (N ~ 2). If the polymer end is in the all-D state (N = 0), the first T (N = 1) can be established at the polymer tip (n = 1) either by exchange (k x ) or by addition of aT (C(lDa). Once the first T is in place, subsequent Ts have a better chance to add (and form a significant cap) because C(lTa » kx and C(lT » C(lD. For this model, the two phases referred to above will be defined as: phase 1, with a significant cap of Ts, N ~ 2, and phase 2, without a significant cap, N = 0 or 1. In order for this model to exhibit rather clear two-phase behavior [by this is meant that usually there are a large number of elementary transitions, Eq. (24.2), within each phase before a phase change, N = 1 ~ N = 2, occurs], the following conditions are necessary: (a) C(2 should be much larger than kx and C(lD so that an all-D polymer (phase 2) will lose many subunits before a significant cap (N ~ 2) grows from a nucleus (N = 1) and (b) once a significant cap is formed (phase 1), the cap should grow to reasonable size (N)> 1) so that a fluctuation back to N = 1 and N = 0 (phase 2) will be slow. The kinetic diagram for N, the number of Ts in the solid cap, is shown in Fig. 8-1(b). In this diagram, we have omitted the kinetically negligible rate constants K~, k~, K', and C(~. Also, C(2 does not appear in the diagram because the state N = 0 remains in state N = 0 following an C(2 transition (loss of a D from an all-D polymer). It should be noted that there are no restraints (i.e., required relations) among the rate constants in Fig. 8-1(b), including C(2: the omitted negligible rate constants provide the necessary flexibility required to satisfy Eqs. (24.3), whatever the choice of rate constants in Fig. 8-1 (b) (including C(2). In steady growth or shortening of the C( end of the long polymer, the cap will achieve a steady-state distribution among the N values in Fig. 8-1(b). This distribution is easy to find because the kinetic diagram is linear. From Fig.8-1(b), (24.4) (N ~ 1). From these we obtain, after normalization (LN PN = 1), 1- x Po=---1 -x + xo (24.5) (24.6) Thus PN falls off with N as X N- 1 • The mean value of N is
Illustrative Biochemical Models That Generate Phase Changes _ N X ro = 231 L NPN = (1 - x N=O + x o )(1 0 - x) (24.7) • Note that N -4 00 when x -4 1. The PN above converge provided that x < 1. The value of a at which x = 1 is obviously a critical value (denoted ao ): , a~T ao = +K ,x a lT a ao (24.8) =-. The above treatment is valid for a < ao . The variance in N relative to found, as in Eq. (24.7), to be 1 - x2 (J~ &2 + xox &2 is (24.9) Xo This is of order unity, corresponding to relatively large fluctuations [compare Eq. (5.11)]. Let Pn be the probability that position n has a T. Then Pn = Pn + Pn+1 + Pn+2 + ... (n ~ 1). (24.10) 1). (24.11) + P2 + P3 + .... (24.12) From Eq. (24.5), p = n x n- X 0 1- x l + Xo (n ~ An alternative expression for N [Eq. (24.7)] is N = Pl The steady-state subunit flux Ja can be written as a sum of contributions from each value of N. From Fig. 8-1(b), including a2 transitions, Ja = po(alDa - a 2) + Pl(alTa - a/lD) + (1 - Po - Pl)(alTa - a/lT)· (24.13) As in Chapter 7, we denote the polymer hydrolysis rate (rate of production of PJ by JT . From Fig. 8-1(b) again, JT = Pl Ko + (1 - Po - Pl)K x o [(1 - X)Ko + 1 - x + Xo XK] (24.14) This quantity is always positive. Using the detailed balance relations [Fig.8-1(b)] po(alDa + k x) = + Ko) PN+1(a /lT + K) Pl (a /lD PN(alTa) = (24.15) (N ~ 1) (24.16) one can rewrite Eq. (24.13) as Ja = -a 2Po - kxPo = -a 2Po + JT - + KoPl + K(1 kxPo, - Po - pd (24.17)
232 NTP Caps and Possible Phase Changes at Polymer Ends with Po given in Eq. (24.5) and JT in Eq. (24.14). The term -lXZPo is the D subunit flux whereas JT - kxPo is the T subunit flux: the net rate at which Ts are added to the polymer by on-off transitions must equal, at steady state, the net rate at which Ts disappear from the polymer by hydrolysis and exchange. When a = 0, x = 0, 1 kx x =--o Po = 1 + xo' lX'lD + /(0 xo Pl = 1 + xo (24.18) (24.19) (24.20) (24.21) When a > ao , the cap is always present but has no definite mean size: it increases steadily with time. The hydrolysis rate in this case is JT = /(, a constant, and the net rate at which subunits are added to the polymer end is Ja = lXlTa - lX'lT. The steady rate at which the cap grows is then Ja - JT > O. At a = ao , the a < a o regime and the a > a o regime have the same properties: (24.22) There is a discontinuity in the slope of Ja at a = ao • At a = a o +, the slope is lX lT . The slope at a = a o - is easy to derive but the expression is a little lengthy and is omitted. With the rate constants of interest here (especially Cl:zlarge and k., lXlD small), this latter slope is much larger than lXlT (see the numerical example below, in Fig. 8-5). So far we have introduced an explicit detailed or "microscopic" biochemical model for the behavior of the lX end of a hypothetical polymer. We now turn to an alternative way oftreating the same system, using "macroscopic" parameters that are composites of the microscopic or elementary rate constants of the original mechanism. The macroscopic parameters relate to a view of the polymer end in which the end is in one of only two "macroscopic" states or phases: either the polymer end has a significant cap (phase 1, N ~ 2) or it does not have such a cap (phase 2, N = 0 or 1). This point of view will not be useful for any arbitrary set of microscopic parameters but only for those sets for which transitions between phases (N = 1 <=± N = 2) are very infrequent compared to elementary transitions within each phase. Of course the present model is so simple that there would appear to be no need for an alternative approach to the system kinetics. The equations already given contain, implicitly at least, the essentials of the steady-state behavior of the system. However, for some parameter sets the two-phase treatment will bring out, explicitly, dominant kinetic behavior that would otherwise be missed (see below). Unlike the present simple model, for almost any realistic microscopic
Illustrative Biochemical Models That Generate Phase Changes 233 mechanism for this type of polymer, an analytical treatment of the system kinetics will not be possible: Monte Carlo calculations will be needed (see the next subsection). In these cases, if the Monte Carlo calculations indicate two-phase activity, the two-phase kinetic alternative provides a much simpler way of handling the kinetics of the system. In fact, many aspects oftwo-phase "macroscopic" kinetics can be studied analytically (Sections 25 and 26). It is also possible that suitable experiments might provide a full set of empirical two-phase macroscopic parameters without any consideration at all of the elementary (microscopic) kinetic mechanism. With the above introduction, we turn now to the two-phase formulation of the model in Eqs. (24.1)-(24.21), when a < ao • Figure 8-1(c) restates our definition of the two phases: phase 1 is the capped phase with N ~ 2 whereas phase 2 is the uncapped phase with N = 0 or 1. We define f1 and f2 as the probabilities that the rx end is in phase 1 or phase 2, respectively, at steady state. Then [Eq. (24.6)] (24.23) f2 = Po + P1 = 1 - f1 = + xo) . + Xo (1 - x)(l 1- x (24.24) When x --t 1 (i.e., a --t ao), f1 --t 1. When x --t 0 (i.e., a --t 0), f2 --t 1. The subunit flux is denoted J 1 when the rx end is in phase 1 and J 2 when the rx end is in phase 2. [Note that the J 1 , J2 notation here and in Section 25 is not the same as J 1 , J2 in Chapter 7; see Eqs. (21.29) and (21.30).J Thus Ja = f1 J 1 + f2 J 2 = f1 (rxlTa - rx'lT) (24.25) + po(rxlDa - rx 2) + P1 (rxlTa - rx'lD), (24.26) where Eq. (24.26) is essentially the same as Eq. (24.13). Comparison of Eqs. (24.25) and (24.26) gives (24.27) (rxlDa - rx 2) + xo(rxlTa - rx'lD) 1 + Xo (24.28) The relative weights of states N = 0 and N = 1 in phase 2 are 1 and xo. When rx 2 is relatively large, which is the case of primary interest, J2 ~ - rx 2/(1 + x o ). Thus, when a > rx'lT/rxlT, phase 1 (cap) is a growing phase (J1 > 0). The on and off rate constants for phase 1 are rx lT and rx'lT, respectively. The corresponding on and off rate constants for phase 2 are, from Eq. (24.28), (24.29)
NTP Caps and Possible Phase Changes at Polymer Ends 234 . off . , a(Z) - az = + x o a'1D 1 + Xo (24.30) These two rate "constants" are functions of a through Similarly, if we write Xo' JT = 11 Jil ) + IzJ+Z) (24.31) for the hydrolysis rate, then Eq. (24.14) leads to Jf!) = K (24.32) JT(Z) -_ P1 K o Po + P1 (24.33) for the hydrolysis rates in the two phases. The mean value of N in phase 1 is 2pz + 3P3 + ... N1 =-----pz + P3 + ... 2- X 1- x (24.34) Similarly, in phase 2, - Nz = P1 (24.35) --'-=-- Po + P1 Also, (24.36) Figure 8-2(a) is the two-phase macroscopic kinetic diagram we are in the t! -- frlTa +1 m+ 1 k' fr;T fr(2)a t!fr(2) m+ I k' fIt m k' - fr(2) at! fr(2) m fIt (a) ! -f2 m ! -f2 k m - I m -I k' Phase 1 +1 k' k m- I III k k m frlTat !fr'lT k k III m - I k' Phase 2 Phase 1 Phase 2 (b) Fig. 8-2. (a) Two-phase kinetic diagram with on and off rate constants in the two phases. Subunit gain or loss is counted by the integer m. (b) Simplification in which one-way composite rate constants are used in place of the on-off transitions in (a).
Illustrative Biochemical Models That Generate Phase Changes 235 process of developing. The subunit on and off rate constants for the two phases, just introduced, are included in Fig. 8-2(a). The integer m counts subunit gain or loss from the ex end, with an arbitrary origin (m = 0). We consider now the phase change rate constants k and k', which also appear in Fig. 8-2(a). At steady state, because k and k' apply to every m in Fig. 8-2(a) (k and k' depend on the NTP cap, not on the length of the polymer), k/1 = k'fz (detailed balance). That is k' k 11 Iz 1 - Po - PI Po (24.37) + PI XoX (1 - x)(1 (24.38) + xo)' Note that k = k' and 11 = t when XoX = (1 - x)(1 + xo ). If Xo is small, this relation is Xo ~ 1 - x. The two steady-state transition rates between states N = 1 and N = 2 in Fig. 8-1 (b) are also (by definition) the two steady-state rates between phases 1 and 2 [Fig. 8-1 (c)]. Furthermore, these two rates are equal, because of detailed balance. Thus, at steady state, rate phase 1 ~ 2 = (ex'l T + K)Pz = k/1 (24.39) k'f2' (24.40) = rate phase 2 ~ 1 = ex lT apl = From these relations we find (24.41) (24.42) Thus k decreases linearly with a and k k , = ~ cxlTa(pdlz) 0 as a -+ ao . Also, cx 1T ax o = --. 1+ Xo (24.43) Hence k' increases with a. Note that Eqs. (24.41) and (24.43) are consistent with Eq. (24.38). The quantities Pzll1 and pdlz are state probabilities within the separate phases. The form of k and k' is reminiscent of the Eyring rate theory. A representative part of the full kinetic diagram for an ex end in which we follow both the cap size (N) and the gain or loss of subunits from the end (m) is shown in Fig. 8-3. Figure 8-3(a) indicates the possible transitions in and out of the various states m, N and Fig. 8-3(b) provides the corresponding rate constant labeling. Because every m level in Fig. 8-3(a) has the same transitions and rate constants, the full two-dimensional diagram [Fig. 8-3(a)] can be compressed vertically, without approximation, into the one-dimensional diagram in Fig. 8-1(b). On the other hand, compression horizontally into the onedimensional two-phase diagram in Fig. 8-2(a) (the phase 1,2 order is reversed) is, in general, not exact because the N values within each phase are not
NTP Caps and Possible Phase Changes at Polymer Ends 236 - N=O m+ 1 m m -J " I~ I // " " • • • Phase 2 : Phase 1 --+ " I I I I I I I I • ~ • I I I I I J/ " 3 2 I I I // " {/ " I I (a) " jOilD~D k, 0i2 "0 • "" OilT~' OilT " OilT~ OilT (b) Fig. 8-3. (a) Detailed kinetic diagram, including both m and N, for the constant labeling for the transitions in (a). ~ end. (b) Rate equivalent. However, the approximation will become very accurate if the two phases are cleanly separated, that is, if phase 2 is concentrated at N = 0 and if phase 1 is dominated by states with large N (N » 2). Concentration at N = 0 in phase 2 requires xo« 1 and N »2 in phase 1 requires x near 1. An equivalent definition of "clean separation" is that when the two phases are about equally probable (fl ~ 0.5), there are, on the average, a large number of elementary transitions within each phase before a phase change occurs. Let us digress to be more explicit about the "not exact" comment above. Actually, at steady state, Eqs. (24.23)-(24.43) are formally exact by definition, whatever the choice of elementary rate constants in the model and whether or not there is clean separation between the two phases. However, to be useful and significant, a kinetic diagram like Fig. 8-2(a) must be valid not only at steady state but also in transients-e.g., any approach to steady state. However, Fig. 8-2(a) will be a good approximation to use for transients only if the elementary transitions within each phase are fast relative to transitions between phases so that each phase is practically in an internal steady state that is essentially unperturbed by the comparatively slow leakage between
Illustrative Biochemical Models That Generate Phase Changes 237 phases (i.e., phase changes). Thus, "not exact," above, refers to the general applicability of Fig. 8-2(a), not simply to steady states. We denote the mean number of elementary transitions during the mean lifetime of phase 1 by ~; %2 has a similar meaning for phase 2. For any state (N ~ 2) in phase 1, there are three possible elementary transitions: ell n IX1 Ta, and K. The mean time t1 between transitions is then _ t1 1 =------ IX'lT + K + IXlTa (24.44) The mean lifetime of phase 1 is k- 1 • Hence, from Eq. (24.39), (24.45) l+x I-x (24.46) can be large only when x is near 1. In phase 2, states N = 0 and N = 1 have relative steady-state populations 1 and Xo and mean times between transitions in these states: ~ (24.47) (24.48) Thus - t2 = tN=O - + xotN=l --'--~--'---'----=- 1 + Xo (24.49) and, using Eq. (24.43), 1 .A(=2 k't2 (IX'1O = Xo(1X1 Ta + Ko + IX lT a)(IX2 + IX10a + k x )(1 + x o )2 + XoI(2)' (24.50) (24.51) The approximation in Eq. (24.51) follows if IX2 is large, IX1 Ta is intermediate in magnitude, and Xo is small. In these circumstances, %2 » 1. We consider now a numerical example designed to produce two-phase behavior and to resemble an IX microtubule end qualitatively (see the next subsection). The microscopic rate constants chosen are:
NTP Caps and Possible Phase Changes at Polymer Ends 238 Table 8-1. Two-Phase Properties in an Example a (flM) k (S-l) k' (S-l) f1 %1 %2 N N1 6 7 8 9 9.5 9.6366 9.75 9.9 10 10.0 7.5 5.0 2.5 1.25 0.909 0.625 0.25 0.00 0.453 0.565 0.687 0.819 0.889 0.909 0.925 0.947 0.961 0.043 0.070 0.121 0.247 0.416 0.500 0.597 0.791 1.000 4.0 5.7 9.0 19.0 39.0 54.0 79.0 199.0 335.0 273.8 228.6 194.2 180.0 176.3 173.4 169.7 167.3 0.180 0.333 0.755 2.74 8.75 14.3 24.5 79.9 3.50 4.33 6.00 11.0 21.0 28.5 41.0 101.0 (XlT = 2.5 ,uM- 1 (X'lT = 0.25 (X2 = 200 K = S-l S-l 24.75 S-l 00 0.005 ,uM- 1 (XlD = (X'lD = 0.25 kx = S-l 0.04 S-l 00 00 S-l (24.52) S-l Note that (X2 is very large whereas (XlD and kx are very small. These rate constants give ao = 10 ,uM, a1 = 0.1 ,uM (the value of a at which 11 = 0), and (XlDao = 0.05 S-l (compared to kx = 0.04 S-l and (Xl Tao = 25 S-l). Of course x is simply a/a o; Xo increases linearly with a from 0.0178 at a = 0 to 0.04 at a = ao ; Xo is always small. Table 8-1 gives a number of calculated properties for a between 6 ,uM and 10 ,uM. The JIll and.JV; values tell us that the dominant phase (determined by 11) always survives for a large number of elementary transitions. At a = 9.6366 ,uM, where the two phases are equally probable (f1 = i), JIll = 54 and J112 = 176; thus both phases survive for many transitions, on the average, when 11 = i. The mean value of N in the capped phase (phase 1) is &1. Note that when &1 ~ 10 the value of JIll is approximately 2&1. Incidentally, the value of &2 increases approximately linearly with a from 0.0175 at a = 0 to 0.0385 at a = 10 ,uM. Over this entire range in a, Po » P1 [Eq. (24.35)]; recall that Ds are lost rapidly from state N = O. Figure 8-4 shows the phase change rate constants k (loss of a cap) and k' (formation of a significant cap) as functions of a: k = k' at a = 9.6366 ,uM, where 11 = i. Figure 8-5 shows (left-hand vertical scale) 11 , 12 , la, and lT as functions of a. Only phase 1 (cap) exists for a > ao. In this range (a > ao), lT is a constant and la = 11. Below a = ao, la is a composite of 11 and 12 , with relative weights [Eq. (24.25)] determined by II. At the bottom of Fig. 8-5,f1 and &dlOO are plotted, using the right-hand vertical scale. Below a = ao , the polymer end can be described as being either in a slowly growing phase (cap, phase 1), with subunit flux 1 1 , or in a rapidly shortening phase (no cap, phase 2), with subunit flux 12 • There are occasional jumps from
239 Illustrative Biochemical Models That Generate Phase Changes Fig. 8-4. Phase change rate constants as functions of a in a numerical example. 10 k,k' (5- 1 ) 2 4 8 6 10 J 1 ,J. J1 20 JT h 0 al 2 4 6 8 a(!1M) a. ao -." ~ II -100 0.8 0.6 NI -200 II'TOO 0.4 0.2 NI Fig. 8-5. Various properties discussed in the text for the same numerical example as in Fig. 8-4.
240 NTP Caps and Possible Phase Changes at Polymer Ends one phase to the other (k, k') that produce, over a sufficient period of time, the average value la. Without the two-phase analysis beginning with Eq. (24.23), we would have been able to calculate la(a) and IT(a) directly [Eqs. (24.14) and (24.17)] but we would have missed, in so doing, the dominant kinetic characteristic of this particular model: the system spends its time in either a capped phase or an uncapped phase, with relatively irifrequent jumps between phases. Thus the subunit kinetics [Eq. (24.2)] can be well described by the simplified ("macroscopic") diagram in Fig. 8-2(a). In fact, as we shall see in the next section, Fig. 8-2(a) can be simplified further, as a very good approxi·· mation, using composite one-way rather than two-way vertical transitions, as shown in Fig. 8-2(b). Thus the subunit kinetics of the polymer end can be represented by only four macroscopic kinetic parameters: k, k', 11 , and 12 , all of which would in general be functions of a (although 12 in Fig. 8-5 is almost constant). A very similar, though slightly more complicated, model that produces two phases has been discussed in Ref. 8. This same model, but with different numerical values for the microscopic rate constants, has been applied to actin in Refs. 7 and 8. Two-phase behavior is absent in the actin case. A second related model (for micro tubules) was discussed in Ref. 9. In this case 0::10 = 0 and K = 0 but the Ko and kx transitions at n = 1 can occur in a cap of any size (these rate constants are called K' and Kif, respectively, in Ref. 9). Thus, for any cap size, position n = 1 can be a D or a T. This leads to a double-tiered linear kinetic diagram; there are two tiers because n = 1 may be D or T. The model can be treated analytically but the algebra is a little more complicated than above. A third related model, a generalization of the one just described, has also been discussed briefly (see the Appendix of Ref. 6). In this case, we still have 0::10 = 0 but /( i= 0 (hydrolysis of the deepest T is possible, as well as of a T at n = 1). The algebra has still further complications, but an analytical solution is possible. Hydrolysis of Any T in the Cap We continue here our discussion of illustrative microscopic kinetic models that lead to two-phase behavior at the 0:: end of a polymer with an NTP cap. The new feature in this subsection is that we allow any T in the cap to be hydrolyzed to D, rather than restricting the hydrolysis to a T at n = 1 or to the deepest T in the cap (as in the previous subsection). This leads to non-solid caps: Ds are distributed among Ts in the cap, though the deep interior of the polymer is all-D. The mathematical consequence of this is that the exact or detailed structure of the polymer end (i.e., for each n, is there a D or aT?) must be followed, transition by transition. This is necessary primarily in order to know whether the departure of a D or a T from n = 1 (i.e., an off transition) will leave behind a D or a T at n = 1. The detailed structure is also needed if
Illustrative Biochemical Models That Generate Phase Changes 241 the model is to include cooperativity in the hydrolysis transition K (i.e., the value of K for a T at n depends on whether n - 1 and n + 1 have a D or aT). The conclusion that follows from the above is that if a microscopic model is used in which hydrolysis is possible at any T in the cap, the amount of necessary detail precludes an analytical treatment of such a model and necessitates a Monte Carlo simulation approach. Three microscopic models (based on micro tubules) that produce two-phase behavior will be discussed below. These are not put forward as models that are expected to withstand the test of time but rather only as illustrations. The object is simply to show that plausible microscopic models can produce macroscopic two-phase behavior. The next two sections will then be devoted to the properties of two-phase models, without regard to their microscopic origins. The first case we mention (see Table 4 of Ref. 10) has microscopic parameters [compare Eq. (24.52)] IXlT = 0.849 IlM- 1 lX'lT = 0.45 1X2 = K = 298 S-l IXlD = 0 lX'lD = 0.45 kx S-l 0.0579 S-l = S-l 0.204 S-l (24.53) S-l The value of K here is much smaller than in Eq. (24.52) because it applies to every T in the cap except a T at n = 1 (Ko and kx apply at n = 1, irrespective ofthe presence or absence of other Ts in the cap). Note that IXlD = 0; initiation of a new cap when the polymer IX end is all-D rests entirely on k x • A Monte Carlo simulation at a = 1.521lM ~ aa (as usual, Ja = 0 at a == aa) showed clean phase changes: the polymer end alternated, stochastically, between periods of (a) slow, steady growth (Ja > 0) with a cap present and (b) rapid loss of subunits (Ja < 0) from an all-D end. The mean number of Ts (including both phases) in positions n ~ 2 was 13.5. The cap (growing phase) was rather "porous": some Ds were scattered at random among the Ts of the cap. In the second example, the above model was altered only by introducing positive cooperativity in the hydrolysis rate constant K (the object was to cause the Ts of the cap to aggregate together). In this example, the rate constants chosen were (24.54) where, for example, the subscript TD refers to a T at any n ~ 2 that has for neighbors a T at n - 1 and a D at n + 1. In this case, Monte Carlo simulation (300,000 transitions) led to aa ~ 2.542 IlM, at which value of a the mean number of Ts in positions n ~ 2 was 28.6. After a IO,OOO-transition discard (to achieve steady state), the first 10 lifetimes (in numbers of transitions) of phases at this a were found to be 1550 (phase 1), 10,050 (phase 2), 1625, 350,
NTP Caps and Possible Phase Changes at Polymer Ends 242 0.09 I I I I I 0.08 I I 0.07 I I 0.06 I' ~ I I 0.05 ill 0.04 I I I I I I I 0.03 0.02 0.01 0.5 I I I I , 1.0 2.0 1.5 2.5 3.0 a, I'M Fig. 8-6. Variation of k, k', and J1 with a, from a series of Monte Carlo simulations. 9175, 1450,3725,225, 15,050, and 300. These are much larger than the values of %1 and %2 in Table 8-1. The phase changes were extremely clean-cut in this example (Ref. 10, Table 3). The statistically averaged value of K at a = 2.542 pM was the same as Kin Eq. (24.53) [in fact, this is how the value of Kin Eq. (24.53) was chosen]. In this second example, a simulation was carried out for 10 other values of a as well, in the range 0.75 to 3.0 pM. Clean phase changes occurred. Included in the printout in each case was the amount of time spent in each phase and the time-averaged subunit flux while in each phase. Thus we have values of fdf2 = k'jk, 1 1 , and 12 , Also, the number of phase transitions was counted in each simulation; this count was supplemented by additional simulation runs at some concentrations. This gives the mean time per pair of phase transitions (e.g., 1 ---+ 2 and 2 ---+ 1), which is equal to (k + k')jkk'. Thus k and k' can be calculated. The 11 points for each of 11 , 12 , k, and k' were then plotted as functions of a, and smooth curves were drawn; these curves are shown in Figs. 8-6 and 8-7. Figures 8-6 and 8-7 are analogues of Figs. 8-4 and 8-5 in the previous subsection. 12 at small a is taken from Eq. 28 of Ref. 9. 1a is a composite of 11 and 12 [Eq. (24.25)]. 1a = 0 at aa ~ 2.432 pM (the difference from aa ~ 2.542 pM, above, is simply a Monte Carlo fluctuation). The 11 curve is a quite accurate straight line: 11 = 0.873a - 0.583, a 1 = 0.668 pM. (24.55) Note that the rate constants in Eq. (24.55) are not simply an and a'n. The capped phase (1d and the uncapped phase (12) have very different fluxes. It is apparent that the cap (phase 1) protects the polymer from possible dissolution (phase 2). The steady-state distribution between the two phases (/1) causes the composite 1a(a) to shift from 11 (a) at high a to 12(a) at Iowa, with a resulting bend in 1a(a) near 1a = 0, but not a discontinuity.
Illustrative Biochemical Models That Generate Phase Changes 243 1.0 a, pM Fig. 8-7. Variation of 11,12,1., and J1 with a, from the same series of Monte Carlo simulations as in Fig. 8-6. 1. is a composite of 11 and 12 at any a; J1 is the steady-state probability of phase 1 (the capped growing phase). The balance between J 1 and J2 (at one polymer end) at a = aa is somewhat reminiscent of the balance between Ja and J p (at two ends) at a = aro in a tread milling conventional steady-state polymer (Section 21). In the third and final example in this subsection, the above treatment for a single stranded polymer was extended to a rather more complicated 5-start helical model of a microtubule. 11 The rate constants used apply to each of the five helices in the model: a1T = 2.0 IlM-1 a'lT = 0.20 S-l a2 = 140 S-l K= S-l a lO = a'lO = kx 0.01875 IlM-1 0.20 S-l = 0.10 S-l S-l (24.56) 1.0s- 1 . These rate constants were selected to simulate some of the experiments of Mitchison and Kirschner. 2,3 The four K values in Eq. (24.54) were used for each T in the cap (cooperativity within each of the five helices) combined with six additional cooperativity factors (details omitted) that relate to nearestneighbor subunit interactions between different helices. As already mentioned,
244 NTP Caps and Possible Phase Changes at Polymer Ends Table 8-2. Some Monte Carlo Results at Several Values of a a JT(S-l) IV No. of phase changes 0 0.384 0.556 0.978 1.98 6.70 9.75 10.8 13.2 14.7 15.9 18.6 20.0 0 0.125 1.07 6.89 69.2 111 129 172 198 214 258 286 0 0 2 8 23 50 54 54 59 62 48 25 1.5 3 4.5 6 7 8 9 10 11 13 15 the first cooperativity feature [Eq. (24.54)] has the effect of keeping the Ts in each helix rather compact. Similarly, the second set of six cooperativity factors was chosen to establish similar cap sizes in the different helices. Clean phase changes at the polymer end were again observed in the Monte Carlo simulations, over a wide activity range. Following the same procedure that led to Figs. 8-6 and 8-7, we obtain here Figs. 8-8 and 8-9. Also, Table 8-2 contains values of JT , N (number of Ts at n ;?: 2), and the number of phase changes that were observed in the 600,000 transitions used at each value of a. It will be noticed that the number of elementary transitions per phase change is very large. The JT and N values in Table 8-2 refer to the full polymer end (five helices). Thus, all of the examples in this subsection indicate that, for polymer ends that exhibit two-phase behavior, the microscopic kinetics can be replaced by an equivalent but much simpler macroscopic kinetic scheme [Fig. 8-2(b)] characterized by the four parameters k, k', J 1 , and J2 , all of which are functions of a. The polymer end is either growing slowly (11 > 0, phase 1, NTP cap) or receding rapidly (J2 < 0, phase 2, no NTP cap); there are infrequent transitions between these two "macroscopic" phases. 25. Attached Polymer with Phase Changes at the Free End In this section, we take the empirical point of view that two phases exist at the free (0:) end of an attached polymer, that one phase (phase 1) is a growing phase, that the other (phase 2) is a relatively rapidly shortening phase, and that phase changes (1 ~ 2) are relatively infrequent. The experimental work
Attached Polymer with Phase Changes at the Free End Fig. 8-8. Phase change constants k and k', as functions of a, found from a 5-start Monte Carlo example. 245 0.05 0.04 0.03 ~ :", ~ 0.02 0.01 10 15 100 LO 0.5 < -400~----~------~-------L--~0 10 15 a.{JM Fig. 8-9. Macroscopic rate constants J 1 and J2 (mean growth rate in the two phases) found in the 5-start Monte Carlo example! Ja is the composite growth rate of the microtubule end at steady state. The critical1l-ctivity is aa = 9.97 (the value of a at which Ja = 0). The 11 curve represents the fractiorl of time the microtubule end is in state 1 (growing, with cap), at steady state. At a = aa, 11 = 0.865.
246 NTP Caps and Possible Phase Changes at Polymer Ends of Mitchison and Kirschner 2 - 4 on microtubules is considered to be sufficiently convincing to justify a theoretical examination (in this section and in Section 26) of the consequences of the two-phase assumption (Mitchison and Kirschner refer to this model as "dynamic instability"). See also the discussion ofthe work ofHorio and Hotani and of Walker, et al. at the end of the chapter. The formal kinetic properties that are derived in the first part of this section do not depend on any assumption about the molecular nature of the two phases. However, in the second part of the section, we use the idea 6 ,9 that phase 2 consists essentially of an all-D end that is thermodynamically unstable whereas phase 1 has an NTP cap that stabilizes the polymer end. There is no requirement that the cap be large (contrary to the microscopic models in Section 24); the assumption of a large cap is made (below) only in the discussion of Fig. 8-16, and even there it is not a crucial part of the argument. Of course, if the cap is not large, some other microscopic mechanism would have to account for the apparent long survival times (small k) of phase 1. Before embarking on the subject of this section, we recall two earlier results needed for reference below. First, as in Eqs. (24.25) and (24.37), for the rx end of a very long polymer at steady state, the fraction of time that the polymer end spends in the two phases (f1'/2) is related to the phase change rate constants (k for 1 --+ 2 and k' for 2 --+ 1) by ("detailed balance") k' k (25.1) If J 1 and J2 are the steady-state subunit fluxes in the two phases, then the mean flux at the rx end is (25.2) The above results will be obtained in a more formal way in Section 26. Ja at a = aa' For this value of a, we have 11 f2 -J2 J1 k' k . = 0 (25.3) The second previous result needed is the mean size of an attached equilibrium polymer or of an attached steady-state polymer that has NTPase activity at the tip only [see Egs. (5.13) and (22.1)-(22.4), with C = 1]. As in Section 24, we continue to use m as the polymer size (N is the cap size). In the abovementioned equations, m= xl(1 - x), where x = alae (equilibrium polymer) or x = alaa (steady-state polymer). The mean value of m, excluding m = 0, is easily found from Eg. (5.12) to be 1 m=--. I-x (25.4)
247 Attached Polymer with Phase Changes at the Free End This kind of mean value (over m polymers. ~ 1) is used below for two-phase attached Approximate Size of an Attached Two-Phase Polymer As an introduction to the next subsection, we give here a simple semiquantitative argument 10 that some readers may find helpful. The next subsection treats the same subject in a more formal and much more complete way. The two-phase attached polymer (with free rx end) that we study in this section has the kinetic diagram shown in Fig. 8-10 [compare Fig. 8-2(a)]. State m = 0 is the bare attachment site (e.g., on a centrosome, for a microtubule). Polymers with sizes m ~ 1 can be in either phase 1 or phase 2. Pm and Rm are state probabilities; the sum of all these is unity. Large values of m are of primary interest so possible complications and refinements at very small m are ignored. Note that growth is initiated (from the bare site) only into phase 1. The first-order on and off rate constants shown in the figure (A, A', fl, fl') may have complicated relationships to microscopic rate constants (Section 24). In general, all of A, A', fl, fl' would be functions of the free subunit activity a. The subunit fluxes in the two phases are (25.5) Presumably fl is very small and may usually be neglected. Let 11 = 0, i.e., A = A', at a == a 1 • In order for significant polymer growth to be initiated from a bare site, it is necessary to have a > a 1 and 11 > O. On the other hand, when a > a~ and 1~ > 0 [Eq. (25.2)J, the attached polymer will grow indefinitely; a steady-state size distribution for finite polymers will not exist. The activity range a 1 < a < a~ is thus the one of interest: polymer growth is initiated and a steady-state polymer size distribution is eventually reached. Because k and k' in Fig. 8-10 are very small compared with 11 and -12' the most common stochastic behavior is slow growth in phase 1 at the mean rate 11 out to a rather large value of m, followd by a phase change to phase 2 and then rapid shortening at the mean rate - 12 until the polymer disappears ~ '//k h' (P 0) 0 , A' u''\. 11 -t- t-- tA A 1 ~ A - - A' k k' I' 1 A 2 ~ I' ~ 3 A' k k' I' 2 ~ JJ. ~ •• , Phase 1 A' k' I' 3 ~ •• , Phase 2 JJ. Fig. 8-10. Macroscopic two-phase kinetic scheme for the end of an attached polymer. The integers m = 0, 1, 2, ... refer to the size of the polymer. The rate constants Aand /l here are not to be confused with the absolute activity (A) and chemical potential (/l).
248 NTP Caps and Possible Phase Changes at Polymer Ends at state O. Growth can then start again in phase 1 from state O. Thus there is reflection, in the mathematical sense, at m = O. Occasionally there will be multiple phase changes (e.g., 1 --+ 2,2 --+ 1, 1 --+ 2) before state 0 is reached via state 2. Because of the extensive sessions in either phase between phase changes, the steady-state Pm and Rm values will be practically constant (except possibly for small m), falling off only very slowly with m. As a consequence of this very slow decrease of Pm and Rm with m, as an approximation let us treat Pm as a constant step function with the value P out to m = m and the value 0 for m > m, and similarly for Rm. We then equate the three net steady-state flow rates among the states in Fig. 8-10: the flow to the right in phase 1; the net flow downward, phase 1 --+ 2; and the flow to the left in phase 2. This gives PJ1 = kmP - k'mR = R( -J2 ). (25.6) Then P k R k' (25.7) PIR is the same as Idl2 in this case. We see then that 11/12 is equal to -J2/11 for an attached two-phase polymer (with reflection at m = 0 into phase 1) whereas 11/12 is equal to k'ik [Eq. (25.1)] for the same polymer end if the polymer is very long (because of detailed balance between the phases, and no reflection). These results are very different. Because - J2» J1, 11 is near 1 in the attached case. Figures 8-7 and 8-9 illustrate 11 in the long-polymer case. Because k and k' are small compared to J 1 and - J2 over a considerable range in subunit activity (see Figs. 8-6 to 8-9 for examples), mis large in this range [Eq. (25.7)]. Note that m--+ 00 as a --+ aa [compare Eqs. (25.3) and (25.7)]. The above behavior of an attached two-phase polymer is very anomalous. In the first place, one might expect infinite polymers to form for a above the critical value a 1 of the growth phase 1. This is not the case because, though k is small, phase 1 will eventually change to phase 2, and a previously growing polymer will begin to shorten rapidly. Thus mis finite above a 1 • At the same time, though, the long survival of phase 1 (because k is small) causes m to be large well below a = aa (even though m--+ 00 only at a --+ aa). This is very unusual: Eq. (25.4) for a conventional attached polymer gives a large m only when a is very close to aa' We pursue these matters in more detail in the next subsection. Attached Two-Phase Polymer at Steady State We return now to the model of a two-phase attached polymer in Fig. 8-10, and give a more systematic analysis. 12 The differential equations that govern the behavior of this kinetic system follow directly from Fig. 8-10:
249 Attached Polymer with Phase Changes at the Free End dPo dt = l' /c P1 + 11 R 1 , 1 (25.8) /cPo (25.9) (25.10) dR ---'" = IlR m-l + Il'R m +1 + kPm - (11 + 11' + k')R m (m dt ~ 2). (25.11) Except for brief mention of transients at the end of this subsection, we restrict ourselves here to steady-state solutions of Eqs. (25.8)-(25.11). For this purpose, the derivatives on the left of these equations are all set equal to zero. Because k and k' in Fig. 8-10, in cases of interest here, are small compared with the other rate constants, the steady-state solution is hardly affected (see below) if we replace A, A' by the unidirectional composite J 1 and/or 11, 11' by the composite - J2 , as shown in Fig. 8-11. Both J 1 and - J2 are positive. The model in Fig. 8-11 is especially useful because its mathematical properties are so simple. In practice, it is probably not possible to decompose J2 into 11 and 11' in any case. Also, as already mentioned, 11 would usually be negligible. We discuss and compare four cases below (the first three prove to be slight approximations to the fourth): I. Fig. 8-11, using composite J 1 and -J2 • II. Use composite -J2 , but retain both A and X. III. Use composite J 1 , but retain both 11 and 11'. IV. Fig. 8-10, unchanged (i.e., retain A, X, 11, 11'). Cases III and IV are included for completeness only; they are not of any practical interest. Also, the distinction between cases I and II is probably significant only for transients. We consider case I first (Fig. 8-11) and in most detail. In Eqs. (25.8)-(25.11) put X -+ 0, 11 -+ 0, .Ie -+ J 1 , and 11' -+ - J 2 • The steady-state solution of these equations is easily found to be _ Fig. 8-11. Simplified version of Fig. 8-10 in which composite Js are used. This is referred to as case I in the text. - J[ -J 2 ••• Phasel •.• Phase2
250 NTP Caps and Possible Phase Changes at Polymer Ends R J P. x m -J2 1 =_1_0_ _ m (m ~ 1) (25.12) -J2 (1 - x) Po = - - - " - - ' - - - (25.13) J 1 (k' - J2 ) x=-----J2 (k + J 1 ) (25.14) J1 - J2 The sum Pm + Rm is the probability of a polymer with m subunits. The normalization relation is Po + I m~l Pm + I m~ 1 Rm = Po + 11 + 12 = 1. (25.15) Note that both Pm and Rm resemble PN in Eqs. (5.12) and (22.1); x here is analogous to a/a~ in Eq. (22.1). However, because k and k' are relatively small in Eq. (25.14), x is near unity. To illustrate this, we use the same numerical example (from Monte Carlo printouts) introduced in Figs. 8-6 and 8-7. This provides activity dependences of all rate constants. The curve labeled x(I) in Fig. 8-12 shows x(a) calculated from Eq. (25.14) for this example. This x is very close to unity over a considerable range in a, unlike the corresponding 1800 1600 1400 1200 1.0 1000 0.8 800 0.6 600 0.4 400 0.2 200 IE: 0 ~ '" "" ~ 0 2.0 2.5 a,J,IM Fig.8-12. The two curves in the lower right corner and the a/a. line relate to Eqs. (25.4) and (25.23) for a simple polymer, with mo = 100 and a. = 2.432 liM. The remaining curves in the figure relate to a numerical example of a two-phase attached polymer, cases I and II [Eqs. (25.12)-(25.33)].
Attached Polymer with Phase Changes at the Free End 251 alaa in the same figure for the aggregation of a conventional polymer. Where x is near 1, Pm and Rm fall off very slowly with increasing m. This was the basis of the approximate argument leading to Eqs. (25.6) and (25.7), which provides a qualitative understanding of the odd behavior of this two-phase model. In this connection, note that Eq. (25.14) gives, for k and k' small, k k' 11 (-12) 1-x~----. (25.16) This is related to Eq. (25.7) for iii [see Eq. (25.21) below]. The probabilities Pm and Rm converge in Eq. (25.12) for x < 1. At x = 1, infinite polymers are formed. This locates the critical activity aa' On putting x = 1 in Eq. (25.14), we find that (25.17) as in Eq. (25.7) (iii -+ 00). Note, from Eq. (25.3), that la = 0 at a = aa for the end of a long polymer molecule in solution (i.e., the same free end would have the same aa whether the polymer is attached or not). Other properties in case I are as follows: f1 = '" Pox L. Pm = - 1- x (25.18) - -11 Po R -m -12 (1-x) (25.19) m;'l f 2-- '" L. m;'l Pm Rm f1 f2 k'-12 -12 k + 11 - 11 -=-=---~-- _ 1 m=-- 1 - x' (J~ -=X iii 2 (25.20) (25.21) (25.22) where iii is the mean size of polymers with m ;::, 1 and Pocc is the probability that m > mo. This quantity is of interest if, for example, it takes a polymer of size greater than mo to be detected visually (by electron microscope, say). The subscript occ refers to visual occupation of the site. Note, as already mentioned in the previous subsection, that f11f2 in Eq. (25.20) for the free end of an attached polymer is different (except at a = aa) from fdf2 = k'ik for an end of a very long polymer molecule in solution under steady conditions [Eq. (25.1)]. The above algebra is still simple if the first on rate constant (0 -+ 1) in Fig. 8-11 differs from the others (Jd. But we omit details. To illustrate numerically, Fig. 8-12 includes (for the numerical example referred to above) iii and Pocc (with mo = 100) as functions of a (aa = 2.432 flM). Pocc corresponds to Fig. 4 of Mitchison and Kirschner. 2 These curves are strikingly different from those in the lower right corner of Fig. 8-12 for a
252 NTP Caps and Possible Phase Changes at Polymer Ends conventional nucleated polymer with the same aa (chosen to facilitate comparison). The iii curve at the lower right corresponds to Eq. (25.4) with x = alaa and, from Eq. (5.12), (25.23) The two-phase behavior (with reflection into phase 1 at the origin) allows long but finite attached aggregates to exist in a range of a above a 1 (11 = 0) but far below a = aa' Incidentally, 11 ~ 1 (phase 1 dominates in the time average) over the whole range of calculation in Fig. 8-12 for an attached steady-state polymer [Eq. (25.20)], because - J2 » J 1 , but this is not. true of 11 for the corresponding end of a long polymer in solution, at steady state with respect to the phase transition [see Fig. 8-7 and Eq. (25.1)]. The resemblance of Pocc to Fig. 4 of Ref. 2 would be much closer if a much larger J 1 (a) had been used. Such a J 1 (a) was in fact fOtllld by Mitchison and Kirschner 3 (see the next example below). Qualitatively, what the analysis of the model suggests is that, well above a = a 1 and well below aa' a site can, at a slow rate, grow a long polymer because k is very small. But when the phase change (1 -4 2) finally occurs, the polymer shortens rapidly, quite possibly to disappearance (because k' is also small). The empty site thus formed can then start the process over again. The equations above provide the average steadystate properties of the polymer on a site, over a long period of time. Further comments on Fig. 8-12: x ~ ! when J 1 = k (because k' is very small); mo = 100 was chosen because this approximate size (for a single microtubule helix out of five) is needed for detection;2 Fig. 4 of Mitchison and Kirschner 2 does not reach full saturation at a = aa' as Pocc does in Fig. 8-12, probably because a true steady state has not yet been reached. As a second numerical example, for case I, we return to the Monte Carlo 5-start microtubule example in Figs. 8-8 and 8-9. The four functions k, k', J 1 , and J 2 are available from these figures. The value of aa is 9.97 pM, where 11 = 0.865. Calculated curves of x, 11 (attached), Pocc (solid curve; using mo = 500), and iii, as functions of a, are shown in Fig. 8-13. We use a five times larger mo here than in the previous example because there are five helices (this mo corresponds to a length of 0.3 pm). Compared with the corresponding curves in Fig. 8-12, we observe here that x is near 1 even for quite small alaa (because J 1 is much larger). Consequently, Pocc starts its upward trend at a much smaller alaa, despite the larger mo. The experimental curve (Fig. 4 of Ref. 2) starts up at about alaa = 0.2. Correspondingly, iii also reaches large values at relatively small alaa. It should be noted that the x curve must drop quite sharply for a < 2 pM because x ~ ! at J 1 = k and x = 0 at J 1 = O. Inspection of this Monte Carlo example shows that it is a quite good approximation to use (25.24)
Attached Polymer with Phase Changes at the Free End 253 9000 8000 7000 6000 5000 IE: 4000 3000 2000 1000 8 10 a, 11M Fig.8-13. Plots of X,fl' in, and P occ (solid curves) from Eqs. (25.14), (25.18), (25.21), and (25.22), using the Monte Carlo rate constant functions k, k', J 1 , and J2 in Figs. 8-8 and 8-9, and mo = 500. The dashed curve for Pocc is based on Eq. (25.24), using the Mitchison-Kirschner (M-K, Ref. 2) J 1 line. up to about a/aa = 0.7. The experimental 3 J 1 line (for the plus end) is significantly above the Monte Carlo 11 curve. If we make a hybrid calculation using the Monte Carlo k from Fig. 8-8 and the larger experimental J 1 (this is 5.364a - 0.37, with the a scale adjusted to our aa value), Eq. (25.24) leads to the dashed curve in Fig. 8-13 for Pocc. This curve starts up at about a/aa = 0.2, in agreement with experiment. 2 We turn now to case II. In Egs. (25.8)-(25.11) put /1-+ 0 and /1' -+ -J2 . Case II is similar to case I but the algebra is more complicated. The main properties are found to be as follows: (25.25) (1 - x)y (25.26) Po=--~ x+y /1 xy = X + y' 12 = /1 Pm 12 Rm X X +y -=~=y _ 1 1 - x' m=~~ (J;' m2 ~=X (25.27) (25.28) (25.29)
NTP Caps and Possible Phase Changes at Polymer Ends 254 (1 P occ = + y)x mo+1 X +Y , (25.30) where A'(k' - J2) - J2(k + Je) -2A'J2 vi' x = -----::-:-:----~- Y = A'(k' - J2) + J2(k 2A'k vi' == [Ji(k + Je)2 + 2A'J2(Je - (25.31) + Je) + vi' (25.32) k)(k' - J2) + A'2(k' - J2)2F2. (25.33) Note that Je and A' appear separately; J1 = Je - A'. The critical concentration a = aa' defined at x = 1, occurs when -J2k = (Je - A')k', as expected [here also Ja = 0 in Eq. (25.2) for the tI. end of a very long polymer]. In the numerical example in Fig. 8-12, the Pocc and mcurves are unchanged (on this scale), but there is an alteration in x, denoted x(II) in the figure, at Iowa (where k is no longer small). In Case III, there is no simple analytical solution and there is no x variable as above [e.g., in Eq. (25.25)J except asymptotically for large enough m. However, the Pm and Rm are easy to calculate successively, starting, say, with Po == 1 (normalization is postponed to the end). In case IV, the successive calculation of the Pm and Rm cannot be done directly as in case III; an iteration procedure must be used. Again there is an asymptotic x for large m. Cases III and IV were compared numerically with cases I and II in an example similar to Figs. 8-6 and 8-7 in which empirical analytical functions were used for all rate constants. Suffice it to say that, for all practical numerical purposes, in the range of a of interest (1.4 ~ a ~ aa), the four models are essentially equivalent. The conclusion is that the simple case I (Fig. 8-11) is an excellent approximation to all the other cases, including case IV, and might as well be used for steady states provided that k and k' are small (x near 1). Case II, as well as case I, might be important for transients. As an appendix, we mention that R.J. Rubin 13 has investigated the timedependent solution ofEqs. (25.8)-(25.11) in case I, starting with Po = 1 at t = 0 (bare site). The analysis is very elegant but too complicated to summarize here. The only simple closed expression occurs in the special case k' = 0: (m(t) = kJ + J 1 J1 X [ J1 - k 1 J2 - ( J 2 e-kt - J1 J1 ) J2 - k - k J1 - J2 ltJ -J e-(J1 2. (25.34) This is the mean value ofm including m = 0 [unlike Eq. (25.21)]. Incidentally, at steady state (t = (0), this mean value (with k' = 0 or k' i= 0) is related to m in Eq. (25.21) by (1 - Po)m, with Po given by Eq. (25.13).
Attached Polymer with Phase Changes at the Free End 255 Probability of Disappearance of Attached Polymer 12 If we start with an attached polymer of size m* in case I above, and it is growing (i.e., in phase 1), what is the probability that it will disappear (reach m = 0) after v = 1, 3, 5, ... phase changes? Similarly, if the initial polymer (m*) is shortening (phase 2), what is the probability that the polymer will disappear after v = 0,2,4, ... phase changes? We use an approximation to answer these questions, but it is a very good approximation because k and k' are small compared to J 1 and -J2. First, some background. Consider a polymer in phase 1 that started (t = 0) growing at rate J 1 at m = 0 (empty site). The rate constant for phase 1 - t phase 2 is k. The probability that phase 1 still survives at t is e- kt • The probability that the phase change 1 - t 2 occurs in the time interval between t and t + dt is e -kt . kdt. The mean size of the polymer at t is J 1 t. As an approximation, if we neglect fluctuations about this mean size, the probability that the polymer still survives at size m is e-(k/Jd m , where we have used m = J1 t. Similarly, the probability that the phase change occurs at m is k e-(k/Jdm._dm J1 k , )m, = _e-(k/J J1 (25.35) having put dm = 1. The exact expression for the probability that the phase change occurs at m is 1 Jm ( J J~ k )m-1 ( J k+ k ) = [ 1 + (kfJ1) (kfJd· 1 1 (25.36) That is, there must be m consecutive J 1 transitions followed by a k transition. Equation (25.35) is a good approximation to Eq. (25.36) if J1 » k. Incidentally, in cases II and IV, where both 2 and A' appear in the kinetic diagram (Fig. 8-10), one can show after a long calculation (see Section 27) that the exact probability the phase change occurs at m is (1 - z)zm-1, where J (25.36a) = (1 - 4P1P2)1/2 P1 = 2/(2 + A' + k) P2 = ..1.'/(2 + A' + k). Equation (25.36) is a special case: let A' - t 0 and put J 1 in place of 2. As an explicit example, let us calculate the probability (denoted at) that a growing polymer, initially of size m*, disappears after one phase change (i.e., reaches m = 0 in the first shortening session). Suppose the phase change occurs at m 1 ~ m*, and then shortening persists from m 1 to m = O. The
NTP Caps and Possible Phase Changes at Polymer Ends 256 probability of this sequence is (see above) ~e-(k/1tl(ml-m*). e-(k'/-J,)ml. J1 If we multiply this expression by dm1 and integrate from m 1 = m* to m 1 = 00, we obtain (25.37) where k k' a'=- b = -J' J1 ' - A 2 = a' + b. (25.38) This same kind of argument can be carried out for more complicated cases with a succession of phase changes. We use a' here to distinguish it from the activitya. For initially growing polymers, starting at m*, the probability of disappearance (reach m = 0) after v = 1,3,5, ... phase changes is found to be Q+ = ~(a'b)(1 (1 A A2 -bm* e (p L (1 p=O + 1)(20" - p)!(Am*)P ' P)".p. where 0" = (v - 1)/2. For initially shortening polymers, starting at m probability of disappearance after v = 2,4, ... phase changes is Q- = (1 a'b)(1+1 A (A2 * m e -bm * (25.39) (0" + 1)'(. 0" _ I -(20"-p)!(Am*)P - - - -__(1 p=O (0" + 1)!(0" _ p)!p!' = m*, the (25.40) where 0" = (vI2) - 1. For v = 0, the probability is e- bm *. The special case m* = 0 in Eq. (25.39) is especially important: + Q(1 a' (a' b)(1 = A A2 (0" (20")! + 1)!0"! (m* = 0). (25.41) For example, in Fig. 8-12 at a = 1.9 pM (where Pocc = 0.441 and in = 124.0), the probabilities of disappearance of a newly started (m* = 0) aggregate after 1, 3, 5, ... phase changes are 0.9731 (=Qt), 0.0255, 0.00134, 0.00009, .... Thus, the new polymer almost always disappears after only one phase change (1 -+ 2)-that is, in the first shortening session. The values of Qt at a = 1.6 and 2.2 pM are 0.9934 and 0.8525, respectively. At a = aa(2.432 pM), a' = b and Qt = 0.50. The probability of eventual disappearance (L(1 Q;) can be shown to be unity if a' ;::, b (a ~ aa) and a'ib if b > a' (a > aa). Attached Polymer That Encounters a Barrier Here we consider an attached polymer that cannot grow to an indefinite length because it encounters a barrier that limits the polymer size to v subunits. The particular example we have in mind is a microtubule (MT) that grows from
Attached Polymer with Phase Changes at the Free End I Jy p'0 -- -Jj P,:1 2 O_}kUk' kHk' 1 __ 2 -- v- I Jj -- -v-I -h -J2 -- kHk' -Jj Jj 257 -J2 R'm v +ko v -J2 Fig. 8-14. Kinetic diagram for a centrosomal microtubule that encounters the metaphase plate of chromosomes at m = v. The upper states (probabilities P.!) are in phase 1 (GTP cap); the lower states are in phase 2 (no GTP cap), with probabilities R!. J2 is negative. a centrosome (pole) until it reaches the metaphase plate of chromosomes. 14 We shall use language appropriate to this case but the model is of some general interest. The kinetic diagram for one of these MTs is shown in Fig. 8-14. Thus the usual diagram (Fig. 8-11) for an uninhibited centrosomal MT is cut off and turned around at m = v. The new rate constant ko is the reciprocal of the mean time it takes for the GTP cap of a MT that has reached the metaphase plate (and stopped growing) to deteriorate sufficiently by hydrolysis to allow MT shortening to begin (phase 2). This would be a relatively slow process, with ko «J1 • Because ko« J 1 , Pv* will be quite large (the asterisk refers to the truncated diagram in Fig. 8-14). Considered as a random walk problem, there is delayed reflection at m = v in Fig. 8-14. The time k;;1 spent at m = v (phase 1) gives the MT a good chance to be captured by a kinetochore, but we do not pursue this aspect of the problem here (see Ref. 14). The steady-state properties in this model differ considerably from case I above because of the reflection at m = v (case I properties are obtained in the limit v -+ 00). In particular, a steady-state probability distribution exists even if x > 1, because of the finite number of states (2v + 1) in Fig. 8-14. The steady-state probabilities are easy to find: P~ = P(jx m (m = 1,2, ... , v - 1) (25.42) (25.43) R! = P.*) x ( J~J: m- 1 (m = 1,2, ... , v) (25.44) (25.45) (25.46)
258 NTP Caps and Possible Phase Changes at Polymer Ends Note that x is defined in the same way as in Eq. (25.14). The ratios P* v (25.47) PV*-1 are important: these show that Pv* is much larger than the probability of the two neighboring states in the diagram. In a realistic numerical example 14 (MT, centrosome, metaphase plate), x is about 0.7, v about 1.2 x 104 , Pv* about 0.3, and the two ratios in Eqs. (25.47) about 9 x 10 3 and 8 x 104 , respectively. From the above state probabilities, . II P:+"'+Pv* 12 RT+"'+R~ -k o J2 x(1-x v - 1 )-JI J2 x v - I (1-x) koJI (1- XV) (25.48) (25.49) where iii is the mean value of m for m ~ 1. The above relations are valid for both x < 1 and x > 1. Exactly at x some special relations are required: P~ = -ko12 vko(J1 - 12) - 1112 II -ko12(v - 1) - 1112 12 k ol1 V = 1, (x = 1) (25.50) (x = 1) (25.51) (25.52) When v is very large, the last relation becomes iii -+ vj2. Simplified Bioenergetics of the Two-Phase Model 14 We have seen, both in Section 24 and in this section, that phase changes in both directions (1 ~ 2) at a polymer end can occur over a considerable range in free subunit activity. As will be shown in this subsection, this property is possible only because of the expenditure of the free energy of NTP hydrolysis (GTP hydrolysis in the case of microtubules, the only known example so far). This behavior (i.e., alternating phases over a range in activity) is distinctly different from that observed in an equilibrium system, where the two phases of an open system can interchange in both directions only at a precise value of the activity (if this is the intensive variable being altered). An example would be a substance precipitating out of a solution at a particular value of a (a = a e ) for the (dissolved) free molecules of the substance. This is illustrated in Fig. 8-15 (see also Fig. 7-10). For a> ae , the solid phase is the stable phase
259 Attached Polymer with Phase Changes at the Free End Fig.8-15. Chemical potentials for free molecules and a solid (ofthese same molecules), as functions of In a, where a is the activity of free molecules. The two phases are in equilibrium at a = a e , where the chemical potentials are equal. Free In a and the transition solid ---t free will not occur. For a < ae , the solid will dissolve and the transition free ---t solid will not occur. If the solid is a large but finite system (e.g., N = 10 4 ), the sharpness in Fig. 8-15 is dulled slightly,15 but on nothing like the scale referred to above where NTP hydrolysis is involved. Our primary aim in this subsection is to present an idealized and simple treatment of the basic bioenergetics of growing (capped) and shortening (uncapped) MTs. Despite the simplicity, this analysis is adequate for an understanding of the essentials of the problem. We use T to refer to GPT· tubulin and D to refer to GDP' tubulin. The chemical potentials of these two species, as monomers at activities aT and aD' are + RTInaT Il'b + R Tin aD' IlT = Il~ (25.53) IlD = (25.54) Similar expressions apply for IlGTP' IlGDP, and IIp, the chemical potentials in solution of GTP, GDP, and Pi' The activities of all of the above five species are considered as steady-state constants, below. If T monomer aggregates to form T polymer (i.e., a hypothetical MT in which GTP hydrolysis is inhibited), the chemical potential of the T polymer, denoted by Il p T, is independent of aT' Let a~ be the (critical) activity of T monomer at which T monomer and T polymer are in equilibrium. Then, at eq uili bri urn, Il p T = Il~ + R T In a~. If we use this relation in Eq. (25.53) to eliminate Il~, (25.55) we have (25.56) as in Fig. 8-15. An analogous equation applies to D monomer and polymer (see below). Figure 8-16 includes Il PTIRT (constant) and IlTIRT as a function ofln(aTla~). 1fT polymerizes at an activity aT > a~ (in the example in Fig. 8-16, aT = e 5a~ = 148.4a~), the process is spontaneous and the free energy change per mole (negative) is
NTP Caps and Possible Phase Changes at Polymer Ends 260 .' 35 30 .' .' .' .' .' .' .' .' ..... ,.' . .' ......... . ' " Jl CTP - Jl p =XT I'pT JlGDP T polymer 15 In(ulufJ Fig,8-16, Illustrative free energy levels, as functions of free subunit activities, for the cycle in Eq. (25.61). (25.57) The quantity 11/11 is shown in Fig. 8-16. With T polymer formed, we imagine the inhibition on GTP hydrolysis to be lifted so that T polymer transforms spontaneously into 0 polymer (i.e., a MT with 0 subunits), with release of Pi into the solution. The constant chemical potential of 0 polymer is denoted /1pD. The free energy change per mole (negative) for this process is (Fig. 8-16) (25.58) The fact that 11/12 is negative is a consequence of the hydrolysis of GTP on the polymer. Also included in A/12 is the presumed positive free energy change associated with the change in two-dimensional crystal structure, T polymer--+ o polymer. That is, T polymer is relatively stable with respect to its subunits (T) compared to 0 polymer with respect to its subunits (0). Perhaps there is
Attached Polymer with Phase Changes at the Free End 261 a conformational change in tubulin 4 as a consequence of GTP hydrolysis on the polymer. The relative instability of the D polymer is reflected in a much larger critical activity a~ (Fig. 8-16) for formation ofD polymer from D monomer, compared with a~ for the formation of T polymer from T monomer (in the example in Fig. 8-16, a~/a~ = e 10 = 2.2 x 104 ). As a consequence, at a typical activity aD < a~, D polymer will disassemble from the ends into D monomer at activity aD. In Fig. 8-16, aD = aT/1O is used, which is realistic. 16 The spontaneous process D polymer --+ D monomer at aD has a free energy change (Fig. 8-16) per mole (negative) of (25.59) The final process (A,u4) included in Fig. 8-16 is the spontaneous exchange ofGTP for GDP on D monomer at aD to produce T monomer at aT. The free energy change per mole (negative) is A,u4 = (,uT + ,uGDP) - (,uD + ,uGTP)· The four processes above [Eqs. (25.57)-(25.60)] comprise a cycle: 1 z T monomer --+ T polymer --+ 3 4 D polymer --+ D monomer --+ (25.60) (25.61) T monomer. In this cycle the tubulin returns unchanged to its initial state (T monomer at aT) but the cycle does have the net consequence that one mole of GTP in solution has been hydrolyzed to GDP + Pi in solution for each mole oftubulin converted into polymer. The total free energy change (negative) for the cycle, per mole of tubulin, is A,ul + A,uz + A,u3 + A,u4 = - X T == ,uGTP - ,uGDP - ,up. XT (25.62) (25.63) X T , a positive quantity, is the thermodynamic force that drives the cycle, Eq. (25.61). It is the force X T that makes it possible, in Fig. 8-16, to have both T polymer more stable than T monomer and D monomer more stable than D polymer. Furthermore, the cycle can occur, with each of the four steps spontaneous (downhill in free energy), for any aT and aD in the activity range between a~ and a~. The above analysis is obviously a generalization of the thermodynamic aspects of the discussion of Fig. 7-1O(b) in Section 21. Here, Fig. 8-16 has a more detailed subdivision of the free energy drop X T . The two T monomer lines and the D polymer line in Fig. 8-16 correspond to Fig. 7-1O(b). Also, A,ul + A,uz and A,u3 + A,u4 in Fig. 8-16 are the same as A,ul and A,uz, respectively, in Fig. 7-10(b). There is more free energy stored in T polymer than in D polymer by an amount A/12 (in Fig. 8-16) per mole of subunits.
262 NTP Caps and Possible Phase Changes at Polymer Ends Here we make a brief kinetic digression. The fact that the process T monomer ---+ T polymer is spontaneous implies J 1 > O. Furthermore, this growth will persist a relatively long distance because Jdk » 1 [see Eq. (25.36)]. The fact that the process D polymer ---+ D monomer is spontaneous implies J2 < O. Furthermore, this shortening will persist a relatively long distance because -J2 /k'» 1. In contrast, the persistence of growth or shortening fluctuations in an equilibrium polymer with forward and backward rate constants rxa (the analogue of Jd and /3 (the analogue of -J2 ) depends on the ratios rxa//3 and /3/rxa, respectively. These ratios are of order unity and the fluctuation persistence is small. As shown in Fig. 8-16, the initial free energy level of the system is /1T(a T) whereas the final free energy level is /1T(a T) - X T. The value of X T/RT chosen in Fig. 8-16 is 23.0, which corresponds to X T = 13.6 kcal mol- 1 at 25°C. We turn now to the more realistic case of MTs growing from centrosome sites until possible capture by a target, e.g., a kinetochore (see the previous subsection). Starting with an empty centrosome site, the initial section of polymer formed on the site is T polymer. The free energy change per mole is /)./11 [Eq. (25.57)]. After the lag period between aggregation and hydrolysis, 1 all further polymer growth in effect adds T to the free tip of the MT and simultaneously converts T to D (by hydrolysis) at the base of the GTP cap, the cap having been created by the "initial section of polymer" mentioned above. Thus, as this second and principal stage of growth proceeds, the cap maintains a constant size at the tip of the MT and the net process is T monomer ---+ D polymer. That is, the growth appears as lengthening D polymer, which extends from the centrosome site to the base of the T cap. The associated free energy change per mole in this second stage of growth is /)./11 + /)./12· If the tip of the MT loses its cap by a fluctuation (free energy change per mole of cap, /)./12) before the tip is captured, the MT will then shorten from the tip back to the empty centrosome site (free energy change per mole, 11/13 + /)./14). In this case an amount of free energy X T per mole of polymer formed has been expended to pay for the exploratory mission by the MT tip.4 If, however, the MT tip is captured by, say, a kinetochore, with cessation of growth, the GTP cap will be lost by hydrolysis (free energy change per mole, /).fL2)· The MT will then shorten (free energy change per mole, /).fL3 + /)./14), pulling the target to the centrosome. Again an amount of free energy X T' per mole of polymer formed, will have been expended. However, in this case, if the target offers a steady resisting force F to the attached polymer as it shortens, some of /)./13 (and X T ) is converted into mechanical work (see the discussion of region II of Fig. 2-15). The amount of this work, per subunit of polymer formed, is Flo, where 10 = 6.15 A. The overall efficiency offree energy conversion, F1o/XT (X T here is per molecule ofGTP, not per mole), is generally very small. The main value of the GTP free energy expenditure in MTs would appear to be to make possible both growth and rapid shortening, as in the Eg. (25.61) cycle.
263 Attached Polymer with Phase Changes at the Free End Nicklas 17 has found that if the force F, above, is increased artificially to a value of 5 x 10-6 dyn, microtubule shortening in anaphase is halted. It is interesting that, in the example shown in Fig. 8-16, an extending force of just this magnitude is required to stop the spontaneous shortening process D polymer ~ D monomer. This can be seen as follows. In Fig. 8-16, the chemical potential of D polymer, /1pD' is larger than the chemical potential of D monomer at aD by an amount - 11/13, given by Eq. (25.59). If an extending force F > 0 is imposed on D polymer, its chemical potential is lowered by an amount loF [Eq. (4.11)]. That is, the polymer is stabilized by the force. In order to halt the shortening process, the horizontal line /1pD + /1p in Fig. 8-16 must be lowered (by 10F) to the position of the lower dashed line in the figure, so that D polymer at F > 0 will be in equilibrium with D monomer at aD' Thus, the force needed is determined by (25.64) In the Fig. 8-16 example, we assumed 1 10' and aT 5 -=e. e aT (25.65) Hence we find (25.66) Using 10 = 6.15 A, Eq. (25.64) leads to F = 4.9 X 10-6 dyn at 25°C. Thus it seems possible that the Nicklas result, above, has a simple thermodynamic explanation. For completeness, we mention that a compressive force F < 0 could raise /1pT in Fig. 8-16 to the upper dashed line so that spontaneous growth (T monomer ~ T polymer) at aT would cease. The required F is determined by (25.67) This gives, in the present example, F = - 3.3 X 10-6 dyn. Finally, it might occur to the reader that equilibrium binding of a ligand on subunits might substitute for GTP hydrolysis in Fig. 8-16. It is easy to see that this will not work. Let T in Fig. 8-16 refer to a subunit with ligand bound and D to a subunit with no ligand bound. We suppose that ligand binds very strongly to free subunits but very weakly to polymerized subunits (because of some kind of neighbor interference). The dotted line at the top of the figure refers to D monomer. This line plus the next three lines (T monomer, T polymer, D polymer) form a plausible and complete set of free energy levels for the equilibrium system. However, unlike the steady-state case (with GTP hydrolysis), the two lines at the bottom ofthe figure (D monomer, T monomer)
264 NTP Caps and Possible Phase Changes at Polymer Ends are no longer present. Thus D polymer is the final, most stable state. It cannot depolymerize spontaneously: the process D polymer --+ D monomer now involves a large increase in free energy up to the dotted line at the top of the figure. In contrast, the cyclic GTPase activity provides an infinite set of repeating free energy levels 18 such that every step can be downhill in free energy. Appendix on the Sleeve Model in Figs. 2-16 to 2-20 In Section 7 we discussed a sleeve model for the maintenance of attachment of a microtubule to a kinetochore while the microtubule was shortening. The polymer was treated as an equilibrium polymer. It is now clear from the preceding subsection that the particular equilibrium polymer and process involved is D polymer --+ D monomer. There are three comments to make, supplementing the discussion in Section 7. (1) A purely thermodynamic argument was made above [Eq. (25.64)] showing that the Nicklas force, F = 5 X 10- 6 dyn, is of the magnitude required to stop the disassembly of D polymer. The calculation was based on Fig. 8-16, which is taken from Fig. 4 of Ref. 14. To some extent this figure was drawn arbitrarily, as an example. For comparison, it should be noticed that essentially the same result is implicit in the sleeve calculations in Section 7, though the thermodynamics was somewhat obscured there by the kinetic argument: aoa e = 0.1930 s-1, ~= aoa e aoa~ = a~ = a~ eloF/kT = ae F = 5.0 X = ~ 0.1930 340 S-l 1762 (25.68) 10- 6 dyn (25.69) = 10- 6 dyn. This corresponds to [Eq. (25.66)] ae aD ~ = lOe 5 = 1484 ' F = 4.9 X in Fig. 8-16. A slight adjustment in the figure would bring the two calculations into agreement. (2) Three parameters were introduced in Section 7 without explanation: M = 65, K = 1800 s-1, and a~ = 340 S-l. The first of these is based on the assumption that the outer layer of a three-layered kinetochore is 400 Athick. 19 Then M = 400/10 = 65. This is the maximum number of tubulin dimers (of a microtubule) that the sleeve can accommodate. The thickness and M may be larger than this. The argument leading to K = 1800 S-l is the following. The steps in the random walk in Fig. 2-18 (An = ± 1) are oflength 10 = 6.15 A. The diffusion coefficient of a chromosome is D = kT/C = KI;, where Cis the friction coefficient and K is the rate constant for discrete steps ofiength 10 , corresponding to
265 Free Polymer with Phase Changes at the Ends D. From Fch = (v, where Fch = 10- 8 dyn is the resisting force of a chromosome moving toward the pole with velocity v = 1 flm min -1 (Ref. 20), we find ( = 6 X 10- 3 g S-l and then K ~ 1800 S-l. The value IX~ = 340 S-l was taken from Table 1 of Ref. 3. This is the off rate constant in the shortening phase (phase 2) of a microtubule. (3) Mitchison 21 has suggested a modification of the equilibrium sleeve model in Section 7 in which a nondirectional ATPase in the wall of the sleeve is the component that attaches to the microtubule (the ADP form binds tightly, the ATP form binds weakly). This complication permits more flexibility in the assignment of rate constants. 26. Free Polymer with Phase Changes at the Ends We begin this discussion by considering general polymers and we end it with a particular application to micro tubules. One End of a Very Long Two-Phase Polymer l2 In this subsection, some general introductory relations for one end (either IX or f3) of a very long polymer molecule in solution are derived. The kinetic scheme is shown in Fig. 8-17. This is related to Figs. 8-2 and 8-10. The variable m counts subunits added to or lost from only the one end (e.g., we might start, at t = 0, with m = 0; m can be negative here). The states in phase 1 (or 2) have probabilities Pm (or Rm)· We define!l = Pm and!2 = R m, with!l + !2 = l. Thus'!l is the fraction of polymer ends in phase 1, etc. From Fig. 8-17 [see also Eqs. (25.1O) and (25.11)]. L: L: d:rm = APm- 1 + A'Pm+ 1 + k'Rm - (A + A' + k)Pm (26.1) (26.2) Phase I (cap) ••• A ~ m-I A' 11 Phase 2 (no cap) ••• ~ 11' kHk' m-I Fig. 8-17. Kinetic scheme for one end of a very long two-phase polymer in solution.
266 NTP Caps and Possible Phase Changes at Polymer Ends Because of the large range in m in cases of interest, the continuous version of Eqs. (26.1) and (26.2) is important: at = oP 8zP 8P D1 8mz - 11 8m + k'R 8R 8zR 8R DZom z - 1zom + kP - at = - kP (26.3) k'R, (26.4) where D1 = (2 + 2')/2, Dz = (f.1 + f.1')/2. (26.5) Equations (26.3) and (26.4) [compare Eq. (6.10)] are one-dimensional diffusion equations in P and R with added terms arising from phase changes (k, k'). The center of the P distribution moves to the right in Fig. 8-17 (11 > 0) and spreads as it moves (D 1 ); the R distribution moves to the left (1z < 0) and spreads (D z ); but this behavior is perturbed by the two distributions leaking into each other (k, k'). We return now to the more general Eqs. (26.1) and (26.2). By summing these equations over m as they stand, or summing after multiplying by m or mZ, a number of relations are easy to derive. Direct summation gives (26.6) The solution, if a is held constant, is f1 (t) = ft f1 + (flO k' OC! -~- - k +-k" - ft)e-(k+k')l (26.7) fZ l - f1, - where flO and ft are the initial and final values of fl' This confirms Eqs. (25.1). We define the mean values of m: (26.8) (26.9) The latter are the means in the separate phases. On multiplying Eqs. (26.1) and (26.2) by m and summing, we find dm dt = f1 J 1 + Iz 12 == 1 (26.10) (26.11)
267 Free Polymer with Phase Changes at the Ends kI1 _ dm2 _ m2)· at = J2 + 12 (m 1 - (26.12) If 11 and 12 have reached their steady-state values [Eq. (26.7)], Eq. (26.10) becomes + k12 (26.13) dt k + k' ' as in Eq. (25.2). 1 is now a constant and m = 1t + mO. Also, in Eqs. (26.11) and dm -=1 k'11 =-~--=- (26.12), k'f2 -+ ~ k, -kI1 -+ 11 At large t, both difference: m1 m2 and k' (t 12 behave like -+ ex)). (26.14) m(above) and maintain a constant (26.15) (26.16) (26.17) (26.18) It should be recalled that 11 is positive and 12 negative. Hence The variance in m, and related quantities, are defined by ~ 1L, "m 2Pm, ~ m 1 = fm2 1L, "m 2Rm = - f2 1 m m2 a;' = f 1af = f1mi m1 > m2 • (26.19) m + f2m~ + f2ai + fd2(m 1 - m2)2. Then, on multiplying Eqs. (26.1) and (26.2) by m 2 and summing over m, we derive (26.20) daf _ dt dai 2D1 +T dt = 2D2 + k'f2 2 2 - + (m1 - 2 - m2) ] (26.21) kf1 2 2 - 2 f2 [a1 - a2 + (m1 - m2) ]. (26.22) [a2 - a1
268 NTP Caps and Possible Phase Changes at Polymer Ends Very Long Polymer with Two Free Ends 22 If a polymer is very long (finite polymers are discussed in the next subsection), the two ends (0: and /3) are independent of each other. This is a limiting reference case. The treatment in the preceding subsection then applies to each end. The mean subunit flux J for the polymer as a whole (both ends) is J = Ja + Jp + f2rx J 2a, flpJl P + f 2p J 2P ' Ja = flJlrx Jp = (26.23) + f2rx = fl P + f2P = flrx 1 (26.24) 1, (26.25) where flrx (a function of t) is the fraction of 0: ends, in an ensemble of polymers, that are in phase 1 or is the probability that a given 0: end is in phase 1, and J lrx (a function of a) is the mean subunit flux for an 0: end in phase 1, etc. Phase 1 is the capped (slowly growing) phase; phase 2 is the uncapped (rapidly shortening) phase. When a steady state (dJ /dt = 0) is reached at a given a [Eq. (26.7)J, f lrx k' = k rx + k"rx rx f - lP - k' k P P + k'P, (26.26) where krx and kp are the rate constants for the phase change 1 ~ 2 (at end 0( or /3) and k~ and kp are those for the phase change 2 ~ 1. All of these rate constants are functions of a. Hence, at a steady state, flrx and fl P are also functions of a (Figs. 8-7 and 8-9). Whereas arx and ap are the activities at which Jrx = 0 and Jp = 0 at a steady state, respectively, a oo is the activity at which J = 0 at a steady state. At a = aoo , Jrx = -Jp. Thus Jrx and J p have opposite signs and there is formal "treadmilling" (Chapter 7). However, treadmilling is overshadowed here by the very different behavior of the two phases at each end. The notation in Eqs. (26.23)-(26.25) is unrelated to that in Eqs. (21.29) and (21.30) in Chapter 7. An alternative way to write Eq. (26.23) is (26.27) J 2l == -J2rx - J lP ' J 22 == -J2rx - J2p · (26.28) The subscripts i and j on Jij (i,j = 1,2) refer to the phases of ends 0: and /3, respectively. Because J lrx and J lp are small and positive (in cases of primary interest) and J 2rx and J 2P are large and negative, all of the Jij' as defined above, are positive. The four terms in Eq. (26.27) refer to the four possible phase states of the two ends. The probabilities of the four states are flrxfl P for state 11, etc. The polymer is shortening rapidly in three of these states and growing slowly in the fourth (state 11). At the a = a oo steady state, where the mean flux J is zero, we have
269 Free Polymer with Phase Changes at the Ends l----__---j~===J~ (al (b) (el Fig. 8-18. Three hypothetical pairs of steady-state J.(a) and Jp(a) curves (ex and the two ends), for very long polymers. fJ are (26.29) This balance between the one growing state and the three shortening states seems more significant, physically, than the equivalent balance 1. = -lp of the two ends. In Eq. (26.29), the lij on the right are large compared with 111 but f1.i1P is much larger than the other three phase-state probabilities. That is, the 11 growing state is most common. The probability f2.i2P is especially small (i.e., the probability that both ends of a polymer will be shortening at the same time is relatively small). Figure 8-18 shows three hypothetical pairs of steady-state 1.(a), lp(a) curves, based on Fig. 8-9 and Ref. 3. The last case [Fig. 8-18(c)] represents an unlikely coincidence (a. = ap). In each case, 1 (not shown) is simply the sum of 1. and Jp; J = 0 at a = a",. There is treadmilling at a = a", in Figs. 8-18(a) and 8-18(b). Related but more detailed sets of steady-state curves would be 110 , l 1p , 12 ., 12p or 1 11 , -112' -121, -122, all as functions of a. Finite Polymer with Two Free Ends22 The subsection above refers to a somewhat hypothetical or limiting case. In contrast, real polymers of finite length may shorten to a residual cap or to disappearance (see below). In this case, the two ends (1X and f3) are not independent because of their "interaction" when polymers become very short or disappear. We assume that there is no homogeneous nucleation of new polymers on the time scale of interest here. At this point we shall begin to refer explicitly to microtubules (MTs). Also, for practical reasons, we use concentration c in place of activity a. The kinetic diagram for a finite two-ended MT in solution is shown in Fig. 8-19. The diagram extends vertically (longer polymers) indefinitely. Am = + 1
270 NTP Caps and Possible Phase Changes at Polymer Ends m =2 m= 1 11 m=0 (b) Fig. 8-19. (a) Kinetic diagram for a finite MT in solution. The subscripts ij refer to the states of ends (X and p, respectively: state 1 is capped and growing slowly; state 2 is uncapped and shortening rapidly. ,1m = + 1 corresponds to addition of one subunit. (b) Rate constants within the square of states at each m level (m ;:, 1). These rate constants refer to phase changes at the two ends. corresponds to adding one subunit to the polymer. The state indices ij at the corners of squares refer to the phases of ends ()( and [3, respectively. The rate constants Jij are defined in Eqs. (26.28). For simplicity, we use one-way composite on-off transitions only (compare Figs. 8-10 and 8-11). The left column of states (11) corresponds to growth at both ends of the MT. The right column (22) refers to shortening at both ends. The last J22 transition (from m = 1) is assumed to lead to disappearance of the polymer. In each of the two central columns, one end is capped and growing slowly while the other end is uncapped and shortening rapidly (the net rate of shortening is J21 or J12 ). The last step in either of the central columns, from m = 1 to m = 0, corresponds to loss of the last subunit from the shortening end, leaving a polymer that consists only of the cap that was at the other end. This residual (average) cap is essentially a seed that can now grow from either end; hence, state m = 0 is placed in the left-hand (11) column. The transition from state m = 0 with rate constant y, leading to disappearance of the polymer, is included to allow for the possibility that some or many caps (we deal with an average) may be so small that they are unstable and hence disintegrate. No experimental information is available on this point at present. However, a straightforward random walk (~m = ± 1) to disappearance (as in Fig. 3-6) is very unlikely because the subunit concentration range of interest here is far above the critical concentration for state 11, where J11 = o.
Free Polymer with Phase Changes at the Ends 271 It is apparent from the above discussion that m refers to the number of subunits in the polymer in addition to the number in the residual cap, which we denote by me' For simplicity, we take me to be a nonfluctuating constant, independent of c. The total number of subunits in a polymer with m is then M = m + me' Very large values ofm generally predominate. Hence, the mean iii» me (iii is defined below). For simplicity, the diagram in Fig. 8-19 is uniform down to the smallest m values. Actually, the rate constants would be somewhat m dependent for small m (Chapter 3). Because large m predominates, this complication is probably not important. The really significant features at small m are the conversion of single-capped polymers into a residual cap (m = 0), the complete loss of polymer when uncapped at both ends, and the possible loss of the residual cap by disintegration. Mathematically, this model has some features of "absorption" at the origin (J22 , y) and some of "reflection" at the origin (J 12 , J21 followed by J 11 ). Because polymers disappear via J22 and y transitions, an initial ensemble of polymers plus free subunits, in a closed system, will never reach a true steady state. The surviving polymers will tend to grow off of subunits lost by shortening and disappearing polymers [compare Eq. (9.34)]. At t = 0, we start with a total of N° polymer molecules, in a volume V, distributed among the states of Fig. 8-19 with probabilities P8, p::" Q::', R::', S~ (m ~ 1), and with normalization P8 +I (P::' m + Q::' + R::' + S~) = (26.30) 1. Summation over m is understood, throughout this subsection, not to include m = O. At a time t > 0, N is the number of surviving polymer molecules in V and Pm.is the probability that an initial polymer has survived and is in state 11, m, and similarly for Qm, R m, Sm, and Po. We define F11 == I Pm' F12 m L == == Fll I m Qm, F21 == I R m, F22 m == ISm m + F12 + F21 + F22 + Po, (26.31) where I is the fraction of surviving polymers. The fraction of polymers lost is 1 Then I. IO dN dt = = dI. 1, N = (26.32) N°I N°Tt = N°( -J22 S1 - yPo)· (26.33) The last relation follows from Fig. 8-19(a). The total number of subunits in the form of polymer is N° M, where M = I m (m + mJ(Pm + Qm + Rm + Sm) + mcPo (26.34)
272 NTP Caps and Possible Phase Changes at Polymer Ends (26.35) m Note that in is the mean excess (over the residual cap) number of subunits per initial polymer molecule. If Ct is the total concentration of subunits in V, then (26.36) (26.37) where dM dt din dt dI J=-=-+m-. C dt (26.38) This subunit flux J is the mean rate of addition of polymer subunits per initial polymer molecule. If we assume dJ /dc > 0 (it is not obvious that this is true for all types of transients), then, because J and dc/dt have opposite signs [Eq. (26.37)J, J and dJ/dt would have opposite signs. Hence, when dJ/dc > 0, J moves monotonically toward J = 0 (where dc/dt = 0 or c = constant). This is not a true steady state, however, because polymers continue to disappear. The time dependence of the probability distribution is governed by differential equations that follow from Fig. 8-19: (26.39) (26.40) (26.41) (26.42) (26.43) where Eqs. (26.40)-(26.43) apply to m ~ 1. The coefficients in these equations are functions of c, and, in general, c changes with t [Eq. (26.37)]. If we sum each of Eqs. (26.40)-(26.43) over m, we find d~~l = kpF12 + k~F21 - (k/Z + kp)Fll + JllPO (26.44) (26.45)
273 Free Polymer with Phase Changes at the Ends (26.46) (26.47) On adding Eqs. (26.39) and (26.44)-(26.47), we recover Eq. (26.33). For very long polymers at steady state, F11 = f1ai1P' etc., as in Egs. (26.26) and (26.27) (because the two ends are independent). If the residual cap always disintegrates, state m = 0 drops out ofthe diagram (y ~ 00, Po ~ 0). In this case the J 12 and J21 transitions out of m = 1 are analogous to the J22 transition out of m = 1. If we multiply Eqs. (26.40)-(26.43) by m, sum each equation over m, and then add the four equations, the result is (26.48) Then, from Eqs. (26.33) and (26.38), J = dM dt = Jll(F ll + Po) - J 12 F12 - J 21 F21 - J22 F22 - ymePo - J 22 meS 1· (26.49) For a very long polymer [Eq. (26.27)], the terms in Po and S1 are negligible. In principle, the time dependence of this system is determined as follows. With CO (initial free concentration), V, N°, and me given, we start (t = 0) with an initial normalized probability distribution P3, p::" Q:;', R:;', S~. Equation (26.36) then determines the total subunit concentration Co which is held constant. The evolution of the probability distribution with time is then prescribed by Egs. (26.39)-(26.43), but the coefficients depend on c(t), which is found from Egs. (26.37) and (26.49). That is, Egs. (26.37) and (26.39)-(26.43) must be solved simultaneously. The four eguations that produce Eq. (26.48) give separately dm Fll d: l = Jll(Fll + Po) + kpF12(ml2 + k~F21(m2l dm 12 F12T = -J12 Fl2 = -J21 F21 mll ) - _ + kpFll(mll + k~F22(m22 dm21 F21 - dt - - - mll ) JllPOm ll _ - m12 ) ml2 ) + J 12 Ql m 12 _ + kaFll(m ll (26.50) (26.51) _ - m 2d (26.52)
NTP Caps and Possible Phase Changes at Polymer Ends 274 dm22 F22--;{t = -J22 F22 _ + k"F12 (m 12 - + kpF2l(m2l _ m22 ) - m 22 ) + J 22 S l m 22 · (26.53) In these equations, (26.54) That is, mIl is the mean excess number of subunits among the surviving polymers that are in state 11, etc. The relation of the mij to mis (26.55) The origin of the various terms in Eqs. (26.50)-(26.53) is rather obvious. If we define the variances m2 = 2 u 11 = -2- m 11 - -2 mIl' -2- m 11 = Im m 2 (Pm + Qm 1" 2 -F 1... m Pm, 11 m + Rm t Sm) (26.56) (26.57) etc. (26.58) then we find from Eqs. (26.40)-(26.43), on multiplying by m 2 and summing over m, (26.59) The first four terms are diffusion coefficient terms (Jij Also, for the separate U;], we find + k~F2l[uil - utI + (m11 + J ll Po [(mll - 1)2 - utI dur2 F12-;[t = J 12 F12 2 + kpFll[U 11 + k~F22[ui2 - Ur2 - 2 U1 2 + (m12 - J 12 Ql(mf2 - Ur2) = 2Dij), as in Eq. (26.20). - m2d 2] (26.60) - 1] + (m12 - 2 - m 11 ) ] - m22)2] (26.61)
275 Free Polymer with Phase Changes at the Ends + kpF22 [O"i2 - O"il + (m21 - m22)2] (26.62) - J21Rl(m~1 - O"id dO"i2 F22dt J22 F22 = 2 + k,J12[0"12 + kfJF21 [O"il - O"i2 2 - 0"22 + (m22 2 - m12 ) ] + (m22 - m21)2] (26.63) - J22 S 1(mL - O"i2)· An Experimental Example of a Finite Polymer with Two Free Ends23 The above formal equations provide some insight into the time dependences of the state probabilities, means, and variances but they are of little practical use because ofthe complexity of the system of equations. Instead, Monte Carlo calculations are called for in dealing with real data. The objective in this subsection is to apply the two-phase macroscopic kinetic model, for each end of a MT, to the experiment in Fig. 4 of Ref. 3. In this experiment MTs were grown from seeds in a solution with an initially high concentration of free T (GTp· tubulin). After shearing and then further rapid growth until the amount of tubulin in polymers became essentially constant, samples were taken over the next 55 min, from which were obtained the (decreasing) concentration of surviving MTs, the mean length (which increased) of surviving MTs, and the length distributions of these MTs.23 We analyze the above experiment theoretically by a Monte Carlo simulation on an initial group of 1550 MTs in a suitable small volume V, treated as a single kinetic system. The basic kinetic model used for each end is shown in Fig. 8-20 for the IX end. All ofthese rate constants are functions of c (see below), the concentration of free T. The variable m in this figure counts subunits added to or lost from the IX end. For a complete polymer, with two ends, the corresponding kinetic Phase 1 (Cap) Jlo """-ho """- Phase 2 (N 0 Cap) Fig. 8-20. Two-phase macroscopic kinetic model or diagram for the IX end of a MT.
276 NTP Caps and Possible Phase Changes at Polymer Ends diagram is given in Fig. 8-19. The discussion, above, of this figure should be reviewed. Note that in the y transition, me subunits become free in solution. In the last J22 transition, me + 1 subunits become free. Homogeneous nucleation of new polymers is assumed not to occur. The phase change rate constants shown in Fig. 8-19(b) apply to every m :?: 1. The values ofm usually populated are in the hundreds and thousands. Hence possible m dependences of rate constants at very small m are ignored. However, all rate constants are functions of c. In the Monte Carlo simulation, each of the surviving MTs does a random walk on its own kinetic diagram (Fig. 8-19). However, the whole collection of surviving MTs in V must be treated as a single system because gain or loss of polymer subunits in the Jij or y transitions alters the concentration of free subunits, which in turn alters the rate constants. The volume V of the solution is constant, as is the total number of subunits (free or in polymers). We turn now to the actual rate constants used. Analytical expressions are needed for each ofthese. Mitchison and Kirschner 2 found Coo = 14 JIM (critical concentration) but the true Coo ~ 10 JIM, because of inactive tubulin. Therefore we correct the constants 3.82 and 1.22 in Table 1 of Ref. 3 by a factor 14/10 to obtain J 1P = l.71c - 1.1. J 1a = 5.35c - 0.37, (26.64) The units throughout this subsection are S-1 for all first-order rate constants and JIM for c. At c = 0, J2a = - 340 and J2P = - 212, but there is no experimental information about the c dependence of these two quantities. We therefore use the shape of curve found in the five-helix Monte Carlo work (Fig. 8-9) but reduce the amplitude of the c dependence to keep all Jij positive at c = 20 (needed below). We use the empirical formulas J 2a = -378.7 J2P = -236.1 d = exp 138.1d + --d 1+ 86.1d + 1+d (26.65) C3~:). The four functions in Eqs. (26.64) and (26.65) are included in Fig. 8-21. We again have to resort to five-helix Monte Carlo results for the ks. For lack of other information, we assume both ends have the same k and k' [see, however, Eqs. (26.71)]. Empirical equations fitting the Monte Carlo simulations (Fig. 8-8) are as follows: ka = kp = k 0.841 = c·2474 (26.66)
277 Free Polymer with Phase Changes at the Ends ----==------, 1.0 ,~ 0.5 10 15 c,/lM Fig. 8-21. Steady-state subunit fluxes for (J. and fJ ends in phase 1 (cap) and phase 2 (no cap). J is the overall steady-state subunit flux (both ends) for very long MTs. The fraction of ends in state 1 (cap), at steady state, is fl' The k and k' functions in Fig. 8-22 are both smaller than those in Eqs. (26.66) by a constant factor of 2.5 (as explained below). Very long (i.e., nondisappearing) polymers with the above rate constants [Eqs. (26.64)-(26.66)] have the steady-state subunit flux l(c) included in Fig. 8-21, as calculated from Eq. (26.27). The critical concentration Coo, where 1 = 0, is Coo = 10.08 (consistent with Ref. 2). The curve 11 (c) at the bottom of Fig. 8-21 is k'/(k + k'), the steady-state fraction of ends in state 1 (i.e., with a cap). Other properties at c = Coo, from Eqs. (26.64)-(26.66), are 1la = 53.6, 1 1P = 16.1, 12a = - 254.1, 111 = 69.7, J 21 Ja = 9.09, k 112 = 104.8, 1p = -9.09, = 237.9, (c a = 9.75, 12p = -158.4 J 22 = 412.4 cp = 10.87) (26.67) = 0.00277, k' = 0.0164, 11 = 0.8554. Here Ja and 1p are the steady-state fluxes at the two ends (note that there is treadmilling), and Ca and c p are the critical concentrations of the separate ends (where 1a = 0 or 1p = 0). The 1ij are of order 10 2 S-l while k and k' are of order 10- 2 S-l, yielding a ratio of order 104 . Thus each 1ij transition in Fig. 8-19(a) is more probable than a phase change by a factor of order 104 . Because ofthis, only a very small error (which has been examined) is made if the 1ij transitions are assumed to occur in packages of g transitions, where we usually take g = 100. Using these g packages, the effective rate constants in the resulting modified diagram are 1ij/g and y/g (with no change in k and k'). Without this compression, the cost or time of a Monte Carlo simulation would be prohibitive because of the very large number of transitions. In two quite different cases, mentioned below, a
278 NTP Caps and Possible Phase Changes at Polymer Ends 0.020 0.016 0.012 ~ ~ 1.0 0.8 0.008 0.6 u 0 0.004 0.4 '" 0.2 0 2 C, 10 liM 15 Fig.8-22. Phase change rate constants (both ends) used in one example [Eq. (26.68)]. Also included is the curve Pocc(c) for the ex end based on Eqs. (25.22), (26.64), (26.65), and (26.68). simulation with g = 100 was repeated using g = 50; in both cases, as expected, the change in g had no noticeable effect on the results. Another somewhat similar simplification was introduced to reduce the cost of the computation. The number of free subunits in the small volume V used in the simulation is more than 10 7 • Corrections to the free concentration C were made only after each net gain or loss of 104 free subunits (i.e., a change in C of less than 0.1 %). The program (with Eqs. (26.64)-(26.66)] was tested first on the 5-min period of very rapid polymer growth after shear (Fig. 4 of Ref. 3), starting with c = 20 and an approximately equal number of subunits in the form of polymers. Details are omitted because no information was available on the initial length distribution and hence a guess was made that all initial polymers were in state 11, with a flat distribution in length of reasonable width about the mean length. The value of c decreased to Coo = 10.08 in 4.25 min and then began to oscillate slightly just below COO" After 5 min, polymers were in approximately the Coo steady-state distribution among states 11, 12, 21, and 22-i.e., in relative amounts f/, fl (1 - fl), etc., with fl = 0.8554 [Eq. (26.67)]. The length distribution was similar in all four states. One of the g = 50 versus g = 100 tests (see above) was made on this period of rapid growth. The main Monte Carlo calculations begin (this is designated t = 0) after
279 Free Polymer with Phase Changes at the Ends 100 o o 60 60 40 40 10 20 0~~--~--~--~~~LD50~-J60 0~~~7-~~~--~~~~ 60~ ( ) (b) (e) (d) 60 ~ 40 E " 20 Lcnlllh.1'1ll (e) (f) Fig. 8-23. Solid lines: experimental length distributions at t = 0 (a), t = 5 min (b), t = 15 min (c), t = 25 min (d), t = 40 min (e), and t = 55 min (f). Dashed lines: normalized Monte Carlo length distributions for the Eqs. (26.68) case. the above 5-min period of rapid growth. We use the experimental 3 starting (t = 0) distribution in length [Fig. 8-23(a)], smoothed out above 18 j1m. The connection adopted between subunit number and length is 5000 subunits = 3 j1m. In view of the results in the rapid growth period (above), we assume that C = Coo = 10.08 at t = 0 and that there is a steady-state distribution among states 11, 12, 21, and 22, all with length distribution of the same shape [Fig. 8-23(a)]. We start, for convenience, with 1550 polymers. The initial polymer concentration (Fig. 4 of Ref. 3) is 6.84 X 1011 ml- 1. Hence V = 2.26 X 10- 9 ml. The initial number of subunits in polymers [Fig. 8-23(a)] is 3.74 x 10 7 ; the initial number of free subunits is 1.37 x 10 7 • The total number of subunits is held fixed .
NTP Caps and Possible Phase Changes at Polymer Ends 280 70 7 60 6 I E. E 50 ::: .d 40 onc 4 -x cOJ 30 3 .,g :;; 2 ::t I ~ <!) 20 c e " <!) <) c Concentration 10 0 0 0 1 u 50 0 0 I,min Fig.8-24. Comparison of Mitch is on-Kirschner experimental points (0, .) with Monte Carlo points (0, .) in the Eqs. (26.68) case. The curves are for the Monte Carlo data. o and 0, mean length; • and ., concentration. The initial parameters chosen were Eqs. (26.64)-(26.66), g = 100, me = 30, and y = 104 . Because J11 is about 70 [see Eqs. (26.67)], this y implies that almost every residual cap disintegrates. This case gave somewhat too slow a decrease in MT number and too slow an increase in mean MT size. However, merely by dividing both k and k' by 2.5-i.e., on using k= 0.3364 C2 .474 , k' = 6.4 X 10- 6 c3 , (26.68) good agreement was obtained with the experimental results (Fig. 8-24). The functions in Eqs. (26.68) are those shown in Fig. 8-22. This change in k and k' does not affect J(c), Coo, and fi(c) in Fig. 8-21; also, Eqs. (26.67) are unaltered except for the k and k' values. In the Eqs. (26.68) case, almost 2 x 106 transitions (g packages, k, k') were needed to reach t = 55 min. At that time 337 polymers remained, distributed as 237 (state 11),49 (state 12),41 (state 21), and 10 (state 22). The final c = 10.11. The Coo steady-state distribution of 337 polymers would be 246, 42, 42, and 7. The initial steady-state distribution is soon distorted because state 21 polymers disappear faster than state 12 polymers [see Eqs. (26.67)]. But the steady-state distribution is approached again as the surviving polymers become longer. During the simulation, c dips below CO) (minimum c = 8.59 at 3.0 min) and performs damped oscillations primarily just below COO" This type of behavior was observed in every simulation. Correspondingly, the number of subunits in polymers is not quite constant with time. The values for every 6 min are shown in Table 8-3. The same case [Eqs. (26.68)] was run with g = 50; no effect other than apparently normal fluctuations was noted. Almost 4 x 106 transitions were required. The combined length distributions for these two cases (g = 100, g = 50), suitably renormalized, at t = 5, 15,25,40, and 55 min are compared
Free Polymer with Phase Changes at the Ends 281 Table 8-3. Number of Subunits in Polymers [Eq. (26.68)] t (min) No. x 10- 7 t (min) No. x 10- 7 t (min) No. x 10- 7 0 3 6 12 3.742 3.945 3.872 3.828 18 24 30 36 3.802 3.816 3.755 3.751 42 48 54 3.796 3.760 3.762 with experiment 3 in Fig. 8-23. The agreement appears to be within normal fluctuations. The experimental numbers of polymers in a sample range between 525 and 560; the Monte Carlo numbers of surviving polymers are 2228, 1399,1071,791, and 669, at the respective times. Of course, the separate Monte Carlo 11, 12, 21, and 22 length distributions are also available, but there are no experiments for comparison. It should be emphasized that the adjustment to k, k' [Eqs. (26.68)] was made solely to fit Fig. 8-24. The agreement between theory and experiment in Fig. 8-23 then followed automatically with no further adjustment of rate constants. The above case [Eqs. (26.68)] is by no means unique, although it represents the simplest alteration of the original information that produces agreement with experiment. Two other cases that agree with experiment (Figs. 8-23 and 8-24) essentially as well were found easily. In the first case k and k' in Eq. (26.66) are divided by the constant 5.0 and, in addition, y is reduced from 104 to 60 [the probability that a residual cap disintegrates, at C = Coo, is then 60/129.7 = 0.46, from Eqs. (26.67)]. A much smaller y, e.g., y = 10, leads to unsatisfactory bimodal length distributions because of the many new additions to state 11 at small m. For the second case, we start with y = 104 , J Za = -340 = constant, J ZfJ = -212 = constant (26.69) leaving k' as in Eq. (26.66). The coefficient in k(c) has been adjusted to give Coo = 10.08. In this case 11 (c) and J(c) are changed somewhat (results not shown). As with Eqs. (26.64)-(26.66), the decrease in MT number and the increase in MT size are found to be too slow. However, this is easily corrected by dividing k and k' by the constant 1.725 to yield k = 0.3643 CZ. 474 ' k' = 9.275 X 1O- 6 c 3 . (26.70) The value of Coo is still 10.08. This last case, with JZa and JZfJ constant [Eqs. (26.69)], is a limiting, actually unrealistic, example. It is not possible for JZa and JZfJ to be strictly constant, for this would mean that a MT without a cap (state 2) could never form a cap. A small amount of exchange of GTP for GDP at the MT end, or of attachment of T to D at the MT end, would
282 NTP Caps and Possible Phase Changes at Polymer Ends allow a cap to form and would alter J2a and J2P ' as for example in Fig. 8-2l. For an analytical expression of these points, see Eq. 28 of Ref. 9. In summary, the two-phase macroscopic kinetic model seems able to account very well for the experiments summarized in Fig. 4 of Ref. 3. As an appendix, the curve Pocc(c) (see Fig. 8-13) was calculated, using the rate constants for the IX end in Eqs. (26.64), (26.65), and (26.68), and is included in Fig. 8-22. Pocc is the probability at steady state that a nucleated site on a centrosome will be occupied by an observable MT (of more than 500 subunits). The calculation was made using Eq. (25.22). The corresponding experimental curve is Fig. 4 of Ref. 2. The theoretical curve (Fig. 8-22) starts up at the same c as the experimental curve (corrected for inactive tubulin) but the theoretical curve rises faster. In particular, Pocc (theoretical) = 1 at the critical concentration Ca = 9.75 11M. The discrepancy is to be expected because the theoretical curve refers to steady state but the experiments never reached steady state. Chronology of the Origin of the Two-Phase Model Some readers may be interested in a chronological summary and review of the origin of the two-phase model for the end of a MT. The previous view of MT aggregation (see Chapter 7), originating with the ideas of Wegner 24 and extended and summarized by Hill and Kirschner,25 was that hydrolysis of T to D is closely coupled to the aggregation of T from solution onto the end of a MT. Thus, a MT would consist entirely of D units. Treadmilling is an interesting consequence for a free MT in solution. However, this picture was upset by Carlier and Pantaloni,l who found that the hydrolysis ofT lags behind its aggregation. Hence, there can be a steady-state or transient cap ofT or mostly T units at each end of a MT, though the interior of a MT is all D. When these new kinetic details were incorporated by Hill and Carlier 9 and by Chen and Hill 26 into a theoretical analysis of the steady-state aggregation flux Ja(c) of, say, the IX end of a MT, where c is the concentration of free T, it was found that the theoretical Ja(c) has a discontinuity in slope or a sharp bend at or near the critical concentration c = Ca (where Ja = 0). This is a new feature that is a consequence ofthe GTP cap: the phase or regime above c = Ca is characterized by a significant mean cap of T whereas the phase or regime below c = Ca has a mean cap whose size decreases rapidly as c is decreased. These theoretical results led Carlier et al. 6 to carry out dilution experiments on MTs in solution that confirmed the predicted near-discontinuity in the slope of J(c) at the critical concentration c'" (where J = 0). Then transient Monte Carlo calculations, 6 using the same detailed or "microscopic" kinetic model as above, led to early-time shapes of Ja(c) curves in agreement with the dilution experiments. It was emphasized 6 that the GTP cap stabilizes a MT end and that the absence of a cap leads to instability and fast depolymerization, as exhibited by the relatively steep slope of J(c) below c = C w The next step, in this alternation between experiment and theory, was
283 Free Polymer with Phase Changes at the Ends the experiments of Mitchison and Kirschner 2 • 3 in which MTs formed from centrosomes, axonemes, or seeds were examined visually and individually to achieve a deeper level of detail. These experiments could be interpreted only by adding one further qualitative feature to the conclusions in Refs. 6, 9, and 26: not only were there two different phases (cap, stable; no cap, unstable) above and below C = Ca (for the If. end), as described above, but, over a range in C on either side of C = Ca , both phases can exist and interconvert (infrequently) at any given c. The capped phase dominates in these interconversions (phase changes) above C = Ca whereas the uncapped phase dominates below C = Cain agreement with the mean flux Ja(c) results already mentioned. As a consequence of the Mitchison-Kirschner experiments, previous (and new) Monte Carlo steady-state cases were examined 1 at very short time intervals, to look for these alternations in phase. At any C not too far from Ca , such phase changes were indeed found: in the time course ofthe steady-state Monte Carlo simulation, a MT end switches occasionally and cleanly from one phase to the other (cap or no cap). These Monte Carlo results then made it apparent that the detailed microscopic kinetic scheme 6 • 9 • 26 for each MT end, which formed the basis for the simulation, could be replaced by a much simpler quantitative macroscopic kinetic modepo based on the existence of the two phases. The two schemes are essentially equivalent and relate to the same system (the end of a MT) but, because of the clean phase changes seen in the simulation, the relatively simple macroscopic model is an excellent approximation to the much more complex microscopic kinetic scheme. The rate constants in the macroscopic model are, of course, effective composites of microscopic rate constants. Applications of the macroscopic (two-phase) model have been summarized in this book in Section 25 and in the present section. In very recent and impressive work, Horio and Hotani 27 have confirmed phase changes at the ends of MTs by direct visual (videotape) observation. The qualitative behavior of the MT ends is the same as described in this chapter. As one would expect, the rate constants k and k' are different at If. and f3 ends (only one, unspecified, concentration C was studied). The parameters found by Horio and Hotani, in the present notation, are (using 5000 subunits = 3 IlM, as above): J 1a = 17.5 s-1, J Za = -219 S-l ° ka = 0.00575 s-1, f1a = 0.906, J 1P = 5.56 s-1, kp = 0.00278 s-1, f1P = 0.992, 0.0556 S-l Ja = -4.70 S-l k~ = J2P = -422 kp = 0.333 S-l (26.71) S-l Jp = +2.02 S-l Ja + Jp = -2.68 S-l. Equations (26.23)-(26.26) have been used here. It will be noted that C must be close to but less than c'" because J is small and negative and also that cfJ < Ca J =
284 NTP Caps and Possible Phase Changes at Polymer Ends because Jp > 0 and Ja < 0 at c. The mean time per pair of phase changes (1 --> 2,2--> 1) is (26.72) This gives 192 s for the r:i end and 363 s for the f3 end. Similar experiments have also been reported by Walker, et al.,28 though the quantitative details are rather different. Finally, earlier et al. 29 have found large synchronous oscillations in turbidity in an ensemble of polymerizing microtubules; this is also a consequence of phase changes. 27. Simulation of Two "Phases" by Aggregation of One Component on Another This section is essentially an appendix to the present chapter. We consider the aggregation of one component on another, as shown in Fig. 8-25. These are equilibrium aggregates, not steady-state aggregates (there are no chemical reactions). Molecules of type A (also called component 1) can aggregate on a surface site and molecules of type B (also called component 2) can aggregate on the end ofthe A aggregate. However, in this model, B cannot attach directly to the surface site and A cannot aggregate on B. Thus, the number of A molecules is N1 ~ 0 and the number of B molecules is N2 = 0 if N1 = 0 and N z ~ 0 if Nl ~ 1. The treatment oflabelloss from a polymer, beginning on p. 99, is somewhat related. This model is of interest because B forms a cap on A. When B is present at the end of the polymer (N2 ~ 1), the A aggregate cannot grow or shorten (J1 = 0); when B is absent from the polymer (Nz = 0), the A aggregate can add or lose subunits (J1 =I- 0, except at equilibrium). Thus the A aggregate alternates between two "phases": it can grow or shorten when the B cap is absent but it is frozen in length (J1 = 0) when the B cap is present. The two "phases" here resemble superficially the two microtubule phases already discussed in this chapter, but there are important differences: the growing or shortening "phase," without the B cap, would generally he very short-lived because the attachment of a single B molecule freezes the A aggregate; the subunit flux in A, J 1 , can alternate between a nonzero value (no B cap) and zero (B cap), whereas the subunit flux in a microtubule can alternate between growing and N j =4 N2 = 3 Fig. 8-25. The A, B model, in which A (component 1) aggregates on a surface site and B (component 2) aggregates on A. The ws are interaction free energies.
Simulation of Two "Phases" by Aggregation of One Component on Another 285 shortening (it is this unusual feature, in a microtubule, that requires the expenditure of the free energy of GTP hydrolysis); and in the A, B model two different kinds of subunits (that are not interconvertible) may add to the polymer whereas, in a microtubule, only one kind of subunit (T) adds, but it can be converted into another kind (T --+ D). Thus, although the two-component equilibrium aggregate (Fig. 8-25) bears a certain resemblance to the two-phase microtubule end, the relationship is not deep. The essential difference is a consequence of the role ofGTP hydrolysis in the microtubule. Despite the above comments, the A, B model has significant intrinsic interest, which we pursue in the remainder of this section. Some readers may find this model pedagogically useful. As will be pointed out below, some of the stochastic properties of the model apply as well to the two-phase microtubule system. Equilibrium Properties We begin with equilibrium properties of the two-component polymer. Hence, the concentrations of both components are lower than the critical concentrations. The beginnings of Sections 5 and 18 should be consulted for similar notation and methods. We assume that the interaction free energies between subunits are as shown in Fig. 8-25 and that each A molecule in the polymer has a partition function ql. Similarly, each B molecule has a partition function q2. Let QNIN2 be the canonical partition function for the entire polymer, with Nl and N2 molecules. Then If we define Cl == e-w'tfkT/e-wtfkT, C2 Xl == qlAle-wtlkT, == X2 == e-w2/kT/e-W2/kT, q2A2e-W2/kT, (27.2) then the partition function for the completely open two-component polymer is (27.3)
286 NTP Caps and Possible Phase Changes at Polymer Ends This result should be compared with Eq. (5.3). Below the critical concentrations (as is required here), Xl < 1 and X z < 1. The probability that N z = 0 (for any N1 ~ 1) is PN 2 =O = 1 - 1 - Xz C' X + zX z (27.4) z as in Eq. (5.6). This is the probability, at equilibrium, that there is no B cap (when N1 ~ 1). For convenience, we define 1 - X z + C2X"i v == ---=-----=---=(27.5) 1 - Xz so that Y = 1 + uv. Then the mean number of molecules of each component in the polymer is N1 = aln Y = alnx 1 VC1X 1 (1 - x 1 )(1 - Xl + VC 1 X 1 ) - aln Y alnx z uCZ x 2 (1 - x z )[1 - X z + u(1 - x 2 Nz = - - = (27.6) + C2 X 2 )] . (27.7) N1 and Nz become large when Xl ---+ 1 and X z ---+ 1, respectively. These equations should be compared with Eq. (5.10). The effect of the cap is to stabilize component 1, that is, increase N1 [note that v > 1 and that vC 1 replaces C in Eq. (5.10)]. Simple Kinetics We turn now to some simple kinetic considerations. The assumed kinetic diagram and rate constants for the polymer in Fig. 8-25 are shown in Fig. 8-26. This is a direct generalization of Fig. 2-4. As in Section 5, the relations between the thermodynamic quantities Xl and X z and the rate constants in Fig. 8-26 are Xl = rx 1at/rx'l and X 2 = rx2aZ/rx~. In Fig. 8-26, the A aggregate increases in size along the top row of states 11 21 12 22 Fig. 8-26. The kinetic diagram, with rate constants, for the model in Fig. 8-25.
Simulation of Two "Phases" by Aggregation of One Component on Another 287 (N l increases). For any Nl ?: 1, there is a possible diversion into one of the (vertical) columns, with N z ?: 1. That is, a B cap may be formed at the tip of the A aggregate. The rate constants for B are the same in all columns (i.e., for an A aggregate of any size, Nl ?: 1). C l and Cz appear only in the final off transition (i.e., Ni = 1 --+ Ni = 0, with i = 1 or 2), because of the different stability of the lone Nl = 1 or N z = 1 subunit (see the discussion of Fig. 2-4). Consider a long A aggregate (Nl large) that is either growing steadily (Xl> 1) or shortening steadily (Xl < 1). The possible B cap is finite; that is, X z < 1 (if X z > 1, the B cap will in due course grow indefinitely and permanently block the A aggregate). When the B cap is absent (Nz = 0), the mean A subunit flux is J l = 1X1 a l - lX'l' When the B cap is present (N2 ?: 1), J l = O. The A subunit flux switches back and forth between these two values as the B cap comes and goes. When the cap has just been lost and J l =I 0, the mean lifetime w of this "phase" is simply 1/lX z az (i.e., attachment of one B molecule, N z = 1, is all that is required to end this period). Once the cap is formed (i.e., as soon as one B molecule has been added), J l = O. The mean lifetime w' of this second "phase" (i.e., of the B cap) is the mean first passage time to N2 = 0 in Fig. 8-27 starting from N2 = 1. From Eq. (9-20), this is found to be W , C2 =-----:- 1X~(1 - x2 ) (27.8) Note that if B were simply bound on the end of the A aggregate but did not aggregate itself (i.e., N z is confined to N z = 0 or N z = 1), we would have (27.9) Thus the factor 1 - X2 in Eq. (27.8) arises from excursions into the region N z ?: 2 in Fig. 8-27. These excursions become larger as x 2 --+ 1, increasing w'. Thus the A subunit flux alternates between the mean value J 1 = 1X1 a 1 - lX'l' with mean lifetime w, and the value J 1 = 0, with mean lifetime w'. Over a long period of time, the mean flux in A is (IX 1a 1 - lX'dw J1 = - - - - - w+w' (IX 1a 1 1- - lX'l)(1 - xz) Xz (27.10) + Czxz It should be noticed, from Eq. (27.4), that Eq. (27.10) can also be written as Fig. 8-27. Kinetic scheme for calculation of the mean lifetime of a B cap (mean first passage time from N2 = 1 to N2 = 0). 2 '"
288 NTP Caps and Possible Phase Changes at Polymer Ends (27.11 ) This is to be expected because, over a long enough time, the B cap can reach an equilibrium distribution even though the A aggregate is growing or shortening at a steady rate (i.e., even though A is not at equilibrium). Equation (27.11) simply states that the A aggregate grows or shortens only when there is no B cap. The equilibrium form of Eq. (27.11) arises because the B rate constants are the same for any Nl (Nl = 0 is excluded). An alternative way to look at the above discussion is the following. When the system arrives at some arbitrary state N l , 0 in the top row of Fig. 8-26, with Nl large (after a transition from Nl - 1, 0 or Nl + 1, 0), the system's mean lifetime at Nl before a top-row transition is made to Nl - 1,0 or Nl + 1, o is the mean first passage time t to the left-hand state in Fig. 8-28, starting from N l , O. That is, possible excursions into N2 ;?: 1 will delay transitions along the top row of Fig. 8-26. From Eq. (9.20) and Fig. 8-28, (27.12) The probability that the eventual top-row transition is to the right, N l , 0--+ Nl + 1, 0, is a l ad(a l a l + a'd. The probability per unit time for this transition is the effective mean top-row rate constant to the right: a l al a l a l (1 - x 2 ) + a'l)t (alai 1 - X2 + C2 X 2 (27.13) Similarly, the effective mean top-row rate constant to the left is a'l (1 - X2) a'l (alai + a'dt 1 - x 2 + C2 X 2 (27.14) These effective mean rate constants are consistent with Eq. (27.10). Simple binding of B (rather than aggegation) is a special case. Here 1 ill = - - , a2 a2 , C2 ill = ----;-, a2 Jl = (alai - a'l) , 1 + K 2 a2 (27.15) where K2 is the binding constant of B on the last A: (27.16) Fig.8-28. Kinetic scheme for calculation of the mean lifetime at Nl (mean first passage time from N1 , 0 to Nl - 1,0 or N J + 1,0).
Simulation of Two "Phases" by Aggregation of One Component on Another 289 Distribution in Arrival Times and Probability of Arrival In the remainder ofthis section we consider some questions that require more detailed analysis. If the A aggregate loses its cap (reaches N z = 0) at t = 0, what is the probability that the cap will be regained (i.e., N z = 0 --+ N z = 1) between t and t + dt? This is a simple first-order kinetic process with rate constant C( Za 2. As is well known, the required probability is C(zaze-a2a2tdt. The mean time at which the cap will be regained is w = 1/C(2az, as already mentioned. When the A aggregate acquires a cap (i.e., reaches N2 = 1 from N z = 0) at t = 0, what is the probability g(t)dt that the cap will be lost (first reach N z = 0) between t and t + dt? The answer to this question, in the special case C2 = 1, is given by Eq. (6.30) with No = 1: (27.17) This probability is not exponentially distributed in time, as in the simple C(z a 2 process above. However, in the binding special case (N2 limited to N z = 0 or 1), Eq. (27.17) reduces to (let C(2aZ --+ 0)C(~e-a2tdt. When Xz < 1, the mean value of t obtained by integrating 30 tg(t)dt from 0 to Cf) agrees with w' in Eq. (27.8) (with C z = 1); see also Eq. (6.32). When Xz ~ 1, the integral 30 of g(t)dt from 0 to Cf) is found to be unity. Physically this means that a cap is certain to disappear (i.e., reach N z = 0) eventually if X z ~ 1; Eq. (27.8) shows that the mean time to disappearance becomes very large as Xz --+ 1. However, if Xz > 1, the same integral of g(t)dt is l/x z < 1. This is the probability (which we shall call P) that the cap will eventually disappear; correspondingly, 1 - P is the probability that the cap never disappears (because X 2 > 1 and Jz > 0, the cap will grow indefinitely unless it happens to disappear by a fluctuation while still small). Another method can be used to find P when C z 0/= 1. In this method, we ignore the time but keep track of all locations, N z , and their probabilities, in a random walk, starting from N z = 1. At N z = 1, in Fig. 8-27, we designate the probability of a step to the left (i.e., to N z = 0) by q and a step to the right by 1 - q. From Fig. 8-27, rx~/Cz q=~,----- (C(z/C z ) +rxza z 1 + C 2XZ (27.18) Similarly, for any N z > 1, the probability of a step to the left (i.e., to N2 - 1) is p and to the right is 1 - p, where (from Fig. 8-27) (27.19) There is absorption at N z = 0 so there are no steps out of this location. Table 8-4 shows how a unit probability package at N2 = 1, before any transitions (r = 0), becomes distributed among the various values of N z as successive
290 NTP Caps and Possible Phase Changes at Polymer Ends Table 8-4. Table of Probabilities after r Transitions in Cap r N2 =0 N2 = 1 N2 = 2 N2 = 3 N2 =4 0 0 q 0 pq(1 - q) 1 0 p(1 - q) 0 0 l-q 0 p(1 _ q)2 +p(1 - p)(1 - q) 0 0 (1 - p)(1 - q) 0 0 0 0 (1 - p)2(1 - q) 1 2 3 transitions (r = 1,2, ... ) occur. Because of the absorption at N2 = 0, normalization requires that, for any r, the sum of entries in the N2 = 0 column down to and including r, plus the sum of entries at r for all Nz ;;?: 1 (i.e., in the r row), must add to unity. P, the probability that the cap will eventually reach N2 = 0 (i.e., disappear), is the sum of all entries in the Nz = 0 column from r = 1 to r = 00 (only odd values of r contribute). By extending Table 8-4 considerably (after the pattern of ps and qs is noted, it suffices to work with numerical coefficients only), one observes that (27.20) where p = p(l - q), Sl (1]) I] = (27.21) p(l - p) = 1 + I] + 21]2 + 51]3 + 141]4 + ... S2(1]) = 1 + 21] + 51]2 + 141]3 + 421]4 + .. . S3(1]) = 1 + 31] + 91]z + 281]3 + 901]4 + .. . S4(1]) = 1 + 41] + 141]2 + 481]3 + 1651]4 + ... (27.22) The recurrence relations among the Si are (27.23) etc. The coefficients in Sl (1]) are the Catalan numbers. Hence, s1 -- 1 + (1 2_ _ ~ 41])1/2 - P 'f I Xz < 1 ,p > 1 2 1. 1 1 _ p If x 2 > 1, P < 2' (27.24) The existence of two values of Sl persists for Sz, S3, etc., because ofEqs. (27.23). Thus there are two values of P, depending on whether Xz < lor Xz > 1. When
Simulation of Two "Phases" by Aggregation of One Component on Another 291 < 1, we find from Egs. (27.23) that Si = 1/pi and hence, from Eg. (27.20), that P = 1. This is the expected result, which provides a check on Egs. (27.20)-(27.23) (strictly speaking, these are only conjectures). When X 2 > 1, Egs. (27.23) lead to Si = 1/(1 - pf Then we find from Eg. (27.20) that X2 P q(l - p) = 1 - 2p + pq = 1 1 + C2 (X 2 1) - (1) X2 > (27.25) . When C2 = 1, P = 1/x 2 , as already encountered above. If X2 > 1 and we start with an uncapped, long A aggregate (Xl =1= 1, J 1 =1= 0) at t = 0, the A aggregate will certainly have one session (s = 1) of growth or shortening before being capped by B. The probability that it has only one such session is 1 - P, because this is the probability that once a cap is formed, it never disappears (i.e., grows indefinitely). P is the probability that this first cap does disappear, permitting a second session (s = 2) of growth or shortening by the A aggregate. Then P(l - P) is the probability that the aggregate has exactly two sessions of growth before being permanently capped. Clearly, the probability that the A aggregate has exactly s sessions of growth or shortening before permanent capping is P S - 1 (1 - P). From this probability distribution in s, the mean and variance of s are 1 S= 1_ P 2 p' as = (1 _ p)2' (27.26) . (27.27) where P is given by Eg. (27.25). Thus _ s = 1 + C 2 (X 2 - C2 (X 2 1) - 1) The mean number of sessions s --+ OCJ as X 2 --+ 1 (from above). The mean lifetime of a growth or shortening session is w = 1/cx 2 a2 • Hence, the total mean growth or shortening ofthe A aggregate (measured in molecules of A), before being permanently capped by B, is (27.28) Because J 2 = cx 2 a2 - cx~, this can be written as _ J 1 ws J1 1 + C2 (X 2 - 1) =-'--~-=----'- C2 X 2 J2 -J 1 J2 I'f C2 = (27.29) 1. ° The last result is especially simple: ws = 1jJ2 is the total mean time during which the A aggregate is free to grow or shorten. Recall that J2 > (because X 2 > 1) but that J 1 may be positive or negative. When X 2 --+ 1 (from above), J2 --+ and IJ1 wsl --+ 00. °
292 NTP Caps and Possible Phase Changes at Polymer Ends Distribution in Growth in One or More Sessions for a Long Polymer The primary question we address here is the following: if the A aggregate is long and loses its cap at t = 0, what will be the probability distribution in the amount of its growth (measured in molecules) before it is first capped again? "Growth" is used here in a general sense: it can be positive or negative. The mean of the growth distribution is obviously J 1 w (see above), but we are interested in the complete distribution. A secondary question: what is the distribution after s consecutive sessions of growth? We consider a long aggregate in order to avoid end effects at N1 = 0 or 1. The primary question above (but not the secondary) has relevance for a microtubule as well. Suppose a microtubule end attains phase 1 (or 2) at t = O. What is the probability distribution in the growth at this end before it switches to phase 2 (or I)? In phase 1 the growth is positive; in phase 2 it is negative (a long microtubule is needed in this latter case in order to avoid disappearance). Because this topic has some general interest, new noncommittal notation is used for the rate constants: li. (first order) for addition of a subunit; f3 for loss of a subunit; and k for the growth-termination transition. In the A, B model, li. = li. 1 Q 1' f3 = li.'1' and k = li.zQz. For phase 1 of a microtubule (Fig. 8-10), li. = A, f3 = X, and k = k. For phase 2 of a microtubule, li. = /1, f3 = /1', and k = k'. Also, we use N for the number of subunits (rather than N1 as in the A, B model). To simplify notation further, we shift the origin of Nand take N = 0 at t = 0, rather than the usual value N = No. Thus N may be positive or negative. We are interested here in the amount of growth, and not in the amount of time, in the first growth session. In fact, the probability distribution in the time, as already mentioned above, is ke-ktdt. The kinetic system in Fig. 8-29 provides a convenient way to study the amount of growth in the first growth session. The rate constants in the figure have been introduced already. The polymer starts (t = 0) at N = 0 in the top row of states and does a random walk on the top-row integers N until a termination reaction (k) occurs at some N, putting the system into the bottom (U-2) (U-1) CI N ~ .. , --2 ~-I il N~'" ----- (Uo) CI il 0 kl kl kl -2 -I 0 (W_ 1 ) (W o ) (W_ 2 ) --- --(U 1 ) il (U2) CI CI 1 kl 2 ... il kl 2 ... (WI) (W 2 ) Fig. 8-29. The system starts at N = 0 in the top row and does a random walk on the top-row integers until the walk is terminated by a k transition. See the text for further details.
Simulation of Two "Phases" by Aggregation of One Component on Another 293 row of states at N (where it remains). UN(t) is the probability that the system is at N at time t in the top row (i.e., the growth session has not been terminated yet). At t = 0, Uo(O) = 1 (a () function). As time passes, this initial () function spreads and the mean moves (along the top row) with velocity ex - f3 [see Eq. (6.11)]. While the probability distribution UN(t) is evolving in this way, it simultaneously leaks or dissipates into the bottom row with rate constant k at any N. The cumulated probability in the bottom row at N at time t is called WN(t). Eventually all of the top-row distribution UN(t) will have leaked into the bottom row. The desired probability PN that the growth session ended at N (irrespective of the time of occurrence) is then given by PN == WN ( 00). The normalization conditions are +00 L N=-oo [UN(t) + WN(t)] +00 L N=-oo 1 (27.30) +00 L WN ( 00) = = PN = 1. N=-oo The mean and variance of the distribution PN are easy to find. From Fig. 8-29, the master equations are dUN = exUN- 1 dt - + f3UN+ 1 - (ex + f3 + k) UN (27.31) (27.32) We define N;(t) == LN 2 UN(t) Nw(t) == I NWN(t), N; (t) == LN N 2 WN(t). Nu(t) == L NUN(t), N N N (27.33) Then, if we let Nand N 2 refer to the PN distribution, N = Nw ( 00) and = N;( CJJ). N2 (27.34) If we sum Eq. (27.31) over all N, we find -dtd LN UN = -k LN UN' LN UN = e- kt • (27.35) If we multiply Eq. (27.31) by N and then sum over all N, the result is dNu dt = (ex - f3)e -kt- - kNu Nu(t) = (ex - f3)te- kt . The same operation on Eq. (27.32) gives (27.36)
NTP Caps and Possible Phase Changes at Polymer Ends 294 dN dt - ~=kN U a-fJ fUt) = -k-[1 - (1 + kt)e- kt ] a-fJ N = Nw(oo) = -k-' (27.37) N is the mean of the first-session growth distribution PN[(a - fJ)/k is the same as J 1 ill in the A, B model notation, above]. Next, we multiply Eq. (27.31) by N 2 and sum over N. This leads to dN 2 dt U 2(a - fJ)2 te- kt = + (a + fJ)e- kt - -- kNu2 (27.38) Similarly, dN; Tt= N; (t) 2 --2 kNu = (a ~/)2 [2 _ (k2 t2 + 2kt + 2)e-kt] + (a : fJ) [1 _ (1 + kt)e-kt] _ -2 _ N - Nw(oo) - 2(a k2 fJf a + fJ + -k-' (27.39) The variance in the PN distribution is then 2 (IN = N2 _ N2 = (a - Ilf k2 + a + fJ k' (27.40) A (j function that starts from N = 0, moves with velocity a - fJ, remains a (j function as it moves (i.e., deterministically rather than stochastically), and leaks into the WN states would produce a WN variance of (a - fJ)2 /k2, as in Eq. (27.40). The second term in (J~, above, arises from the fact that the surviving UN distribution actually spreads as it moves. Note that the first term in (J~ dominates if k is small. Next, we find PN. Let VN(t) be the probability distribution (with Vo = 1 at t = 0) that satisfies (27.41) This is the same as Eq. (27.31) except that k = O. Then it is easy to show that UN(t) = VN(t)e- kt is the solution of Eq. (27.31). This is physically obvious because the leakage rate constant is k for every N. It then follows from Eq. (27.32) that
Simulation of Two "Phases" by Aggregation of One Component on Another PN = WN( ex)) = k LX) VN(t)e- 295 kt dt. (27.42) (~r/2 IN(2tJ""~Ji)e-(~+P)t, (27.43) The explicit expression for VN(t) is 3l VN(t) = where IN is a modified Bessel function [Eq. (6.17)]. Substitution in Eq. (27.42) then yields 30 J P3 ZN PN = (N ~ 0) (27.44) P3y-N (N ~ 0) where Z= y 1-J 2P2' 1-J Y=------'--2Pl / _ 1/2 _ a -(1-4P1P2) , Pl- a +[3+k [3 P2 = a + [3 + k' PI (27.45) k P3 = a + [3 + k + P2 + P3 = 1. At any state in the top row of Fig. 8-29, PI is the probability that the next step is to the right, P2 the probability that it is to the left, and P3 the probability of a termination step (down). It is easy to prove that z < 1 and y < 1, as is necessary for convergence. PN in Eqs. (27.44) falls ofT exponentially by on either side of N = O. Because zly = al[3, PN falls off faster on the negative side when a > [3. Equations (27.44) can be confirmed using the same method (Table 8-4) that leads to Eq. (27.20). Also, Eqs. (27.44) give the same N and a~ as in Eqs. (27.37) and (27.40), as expected. An alternative way to write the normalization factor in Eqs. (27.44) is b = ~ = (1 - z)(1 - y) 1- zy . -J In the special case [3 ---+ 0 or P2 ---+ (27.46) 0, Eqs. (27.44) simplify to PN ---+ (1 - pdpf PI = al(a + k). (N ~ 0) (27.47) Another important special case is k small. This is the situation for a micro-
296 NTP Caps and Possible Phase Changes at Polymer Ends tubule. In this case, with is small, one finds that rt. > 13, large values of positive N dominate. When Ie k Z-41--- a- P3 13' J k (27.48) --4-- a- 13' and hence that PN -4 a~f3exp( -rt.~f3) (N~O). (27.49) This has the same form as Eq. (25.35). The remainder of this subsection applies to the A, B model but not to microtubules (because of the change in phase). PN is the growth probability distribution after one session of growth, s = 1. If this distribution is frozen for a time by a cap but the first growth session is followed (after disappearance of the cap) by a second growth session of the same type as the first, what will be the cumulative growth probability distribution p~21 after the second session of growth? In general, after s consecutive such sessions, what is PiJl? The general scheme we use is to find piJl from PN (for s = 1) and PjJ-l), starting at s = 2: +00 PjJl= L PjJ-11PN _ M (s=2,3, ... ). (27.50) M=-oo Omitting the tedious details (which have been carried out through s = 5), the results for N ~ 0 and s = 2, 3, and 5 are b2 z N 1 - zy p~21 = - - [ ( 1 p~31 = b3 z N 2 [(1 2!(1 - zy) + (zy)2 (1 p~51 = b5 z N 4 [(1 4!(1 - zy) + 4zy(4 - - N)] + N)(2 + N) + 2zy(2 - N)(2 (27.51) + N) (27.52) - N)(2 - N)] + N)(2 + N)(3 + N)(4 + N) + N)(3 + N)(4 + N) + 6(zy)2(3 - N) + N)(4 + N) + 4(zy)3(2 - N)(3 - N)(4 - N)(2 x (4 - N)(3 x (4 + N) + zy(1 + N) + (zy)4(1 - N)(2 - N)(3 - N)(4 - N)]. N) (27.53) For N ~ 0, replace ZN by y-N and replace N everywhere by -N. Equations (27.51)-(27.53) suffice to show the pattern that presumably holds for all s ~ 2. Note the binomial coefficients within the square brackets. If for the A, B model we have X 2 < 1, there will be an indefinite number of growth sessions and the A aggregate will never be permanently capped by B. However, if X 2 > 1 (and J 2 > 0), the number of growth sessions before per-
Simulation of Two "Phases" by Aggregation of One Component on Another 297 manent capping will be finite. The probability of s sessions before permanent capping is f, = p s - 1 (1 - P), as already used in Eqs. (27.26). The mean growth distribution before permanent capping is then PN- I 1 PN + I 2 p(2) N + I 3 p(3) N + ... . (27.54) Unfortunately, a closed expression for PN is not easy to obtain (but see below for a special case). However, it is easy to find the mean and variance of the PN distribution. To make the argument more general, let PN be an arbitrary one-session growth probability distribution and let P},f) be the growth distribution of s consecutive such sessions. Also, let Is be the (arbitrary) probability of s sessions before permanent capping. The mean and variance of the PN distribution are denoted N and (J~, the mean and variance of the Is distribution are denoted sand (Js2, and the mean and variance of the P},f) distribution are then sN and s(J~. Let <N) and (J2 be the desired mean and variance of the PN distribution in Eq. (27.54). We then have <N) L NPN = = N 11 L NPN + 12 L NPk2) + ... N N =/1N + 2/2N + ... = Ns. (27.55) Similarly, <N 2 ) = L N 2PN = 11 L N 2PN + 12 L N 2Pk2) + ... N N N (27.56) Then (J2 = <N 2 ) - <N)2 = (J~S + N2(J;. (27.57) In the A, B model, we have N=IY.-I3 k s 2 ' 1 = 1 _ p' (IN = (IY.-I3f + k(1Y. + 13) (27.58) 2 (Js = (1 - P pf . Thus, for the A, B model, <N) (J 2 IY.-I3 = = (27.59) k(1 _ P) (IY. - f3f + k(1Y. + 13)(1 pf - P) -----------=-------0---k 2 (1 - (27.60) Equation (27.59) is the same as Eq. (27.28), but in different notation. Reverting to the original notation for the A, B model, with C2 = 1,
298 NTP Caps and Possible Phase Changes at Polymer Ends (J2 (a l a l <N)2 = 1 + ( N) = lll, 2 + a'dJ2 (27.61) lf This expression for <N) has already been given in Eg. (27.29). The variance of the PN distribution is large: (J2 j<N)2 is of order unity. We conclude this subsection by considering multi-growth sessions, and related averages, in the special case that k is small (and a > 13). In this case PN is given by Eg. (27.49). We can treat N as continuous and we need consider N ~ 0 only. In the A, B model this means that k = a2 a2 is small compared to a l a l and a'l. For brevity, we write [Eg. (27.49)J 0 == kj(a - PN = Oe-ON, Then, from Eg. (27.50), Pk2 ) = IN On continuing this process for pjj) = 13). (27.62) PMPN-MdM = 02 Ne- ON . 5 = (27.63) 3, 4, ... , we find easily that OS Ns-1e- ON (5 (5 - 1)! ~ 1). (27.64) We consider now Eg. (27.54), which is applicable only when X 2 > 1. Because > a~ and a 2 a 2 is small, a~ must also be small. Hence 12 = IX 2 a 2 - IX~ > 0 is small as well. On using Is = p s - 1 (1 - P) and Eg. (27.64) for Pjj), Eg. (27.54) leads to the simple result a2 a2 PN = e(l - P)e-O(l-P)N (N ~ 0). (27.65) Alternative expressions for the coefficient here are o 8(1 - P) = - = 3 1 ~- l 1 w3 = - 1 <N) = 12 C2X2 -.~~=----=~- 11 1 + C2 (X 2 - 1) (27.66) By assumption, 12 « 11 and <N) is large. Hence, the coefficient 0(1 - P) in Eg. (27.65) is small and the exponential fall-off with N is very slow. The distribution in Eg. (27.65) has mean and variance <N) = IX - 13 k(l-P)' (J2 = <N)2. (27.67) The second term in the numerator of Eg. (27.60) is lost because k ~ O. In this connection, see the discussion following Eg. (27.40). Distribution in Growth in One or More Sessions Starting from a Surface Site The main topic we examine here is the probability distribution in growth in the first growth session, starting from an empty surface site. In Fig. 8-26 (for the A, B model) we take C1 = 1 (the algebra is much more involved if C l i= 1).
Simulation of Two "Phases" by Aggregation of One Component on Another Fig. 8-30. Kinetic scheme, analogous to Fig. 8-29, but for a system that starts at an empty surface site ---- -CI N=O (N = 0). 299 CI 2 ... ~ ~ kl kl 2 t=;oo : PI P2 We do not consider the time explicitly. Instead we use the method introduced in Table 8-4. Figure 8-30 shows the relevant kinetic scheme. The rate constants are the same as in Fig. 8-29. We use these general rate constants again because the first-session treatment below is applicable to phase 1 of microtubules as well as to the A, B model. The case of interest is rt. > f3 so that the polymer has a tendency to grow indefinitely. In Fig. 8-30, the system starts at N = 0 and does a random walk on the top-row states, until terminated by a k transition (adding a B molecule in the A, B model or loss of the GTP cap in a microtubule). A time-dependent treatment, as in the preceding subsection, is difficult here because of the asymmetry at N = 0 (reflection at N = 0; no k transition at N = 0). Our main interest is in PN , the probability that the random walk will eventually terminate at N (irrespective of the time or number of transitions). The probabilities Pl' P2, and P3' introduced in Eqs. (27.45), are the basic ingredients of the analysis. We start with a unit probability package at r = 0, N = 0 in Table 8-5. The table shows how this package is dispersed after transitions r = 1,2,3, and 4. An underlined probability results from a termination reaction (there is always a factor P3 included). Such a probability is removed from further consideration in the evolution of the system. There is always reflection at N = 0 (N = 0 ~ N = 1). Termination may occur at N = 1 after transition r = 2,4,6, ... ; it may occur at N = 2 after r = 3,5, 7, ... ; etc. P l is the sum of all underlined probabilities in the N = 1 column, and similarly for P2 , etc. Thus Pl = P3 + P2P3(1 + pd + ... , etc. (27.68) The normalization relation among the PN is (27.69) because every random walk will eventually terminate at some N. If Table 8-5 is extended to about r = 8 or 9, the pattern of coefficients and polynomials (e.g., P2P3 and 1 + Pl' respectively, at N = 1, r = 4) in the underlined probabilities becomes clear. It is then possible, using obvious recurrence relations, to extend the table much further (say to r = 14), including now only the numerical coefficients in the underlined polynomials. In this way the complete expression for P l is found to be (27.70)
300 NTP Caps and Possible Phase Changes at Polymer Ends Table 8-5. Table of Probabilities after r Transitions in Fig. 8-30 r N=O 0 1 2 3 4 0 pz 0 0 pW P3 pz(l + Pl) PzP3(1 + pd + Pl) 2 3 4 0 0 Pl P1P3 P1PZ(l + 2pd 0 0 0 pi piP3 0 0 0 0 pi where the Si are defined in Eqs. (27.22) and (27.23). In place, of Eq. (27.24), however, we have here Sl = J 2 1+ J = 1-J (27.71) 2P1PZ' where is defined in Eqs. (27.45). From the recurrence relations, Eqs. (27.23), we then obtain Si = (1-J)i 2P1PZ (i = 1,2, ... ). Using this result, Eq. (27.70) becomes 2P3 P1 = 2Pl - 1 + J (27.72) (1 - J) 2pz 1- z =--'z=1-z z ' (27.73) where z is defined in Eqs. (27.45). In similar fashion we find Pz = P3 = + PZ S3(P1PZ) + P~S4(P1PZ) + ... J PiP3[S3(P1PZ) + PZ S4(P1PZ) + P~S5(P1PZ) + ... J, P1P3[SZ(P1PZ) (27.74) etc. From these we are led to the general result PN = 2p 3 2Pl -1 ZN +J = (1 - z)zN-l (N ~ 1). (27.75) This is close to PN in Eq. (27.44) for N ~ 0, as might have been anticipated. The mean and variance found from PN can be written as N Z (IN = = Pl - pz + A =--=--=--=--1 - Pl - pz Pl + pz - 4P1PZ + B Z ' (1 - Pl - pz) (27.76) (27.77)
Simulation of Two "Phases" by Aggregation of One Component on Another 301 where A = ~(J + 1 - 2pd B = -t[Pl + P2 - 4P1P2 - (27.78) (Pl - P2)JJ. A and B are end-effect terms arising from the asymmetry at N = 0; A --+ 0 and B --+ 0 when k is small. Aside from A and B, Eqs. (27.76) and (27.77) are the same as Eqs. (27.37) and (27.40), respectively. In the special case 13 --+ 0 (or P2 --+ 0), PN --+ (1 - Pl)pf- 1 (N ~ 1) (27.79) where Pl = al(a + k). This is the same as Eq. (25.36). When k is small compared to a and 13, as for a microtubule, k a-f3 z--+I--- I)J (27.80) k [k(N PN --+ --exp 13' a-f3 aThis is essentially the same result as in Eq. (25.35). Because large values of N dominate when k is small, one might as well drop the 1 in N - 1. Equations (27.79) and (27.80) are the same as Eqs. (27.47) and (27.49), except for the shift of the starting point of the termination (k) reaction (N = 1 compared to N = 0). Tn these two special cases, the reflection at N = 0 is, in effect, not used. The problem of successive sessions of growth is of interest for an A, B polymer starting from an empty site, but not for a microtubule. However, the exact treatment of this problem would be difficult and is not attempted here. The complication is that PN is the probability distribution for the first session (which starts at N = 0, with reflection at N = 0) but PN is not the correct form for the distribution in later growth sessions because these start at arbitrary values of N (but with reflection at N = 0). If one simply ignores this complication and treats all growth sessions as equivalent (i.e., a session starting at N has reflection at N and termination at N + 1, N + 2, ... ), then it is found that p(s) N - (N - 1)! (1 - Z)sZN-s (s - I)! (N - s)! and that, from Eq. (27.54) with Is PN = = + (1 (27.81) Ps-1(1 - P), (1 - cp)cpN-l cp == z (N >- s) "" - z)P. (N ~ 1) (27.82) If k is small, Eq. (27.80) is applicable (replace N - 1 by N, as a slight approximation) and reflection is, in effect, not used [see the discussion of Eq. (27.80)]. Thus, successive growth sessions are equivalent and the treatment in Eqs. (27.62)-(27.67) applies here as well.
302 NTP Caps and Possible Phase Changes at Polymer Ends References 1. 2. 3. 4. 5. 6. 7. 8. 9. to. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. Carlier, M.F. and Pantaloni, D. (1981) Biochemistry 20,1918. Mitchison, T. and Kirschner, M.W. (1984) Nature 312, 232. Mitchison, T. and Kirschner, M.W. (1984) Nature 312, 237. Mitchison, T. and Kirschner, M.W. (1984) in Molecular Biology of the Cytoskeleton, eds. Borisy, G.G., Cleveland, D.W., and Murphy, D.B. (Cold Spring Harbor Laboratory, Cold Spring Harbor, NY), p. 27. Carlier, M.F. and Pantaloni, D. (1986) Biochemistry 25, 7789. Carlier, M.F., Hill, T.L., and Chen, Y. (1984) Proc. Natl. Acad. Sci. USA 81, 771. Pantaloni, D., Hill, T.L., Carlier, M.F., and Korn, E.D. (1985) Proc. Nat!. Acad. Sci. USA 82,7207. Hill, T.L. (1986) Biophys. 1. 49, 981. Hill, T.L. and Carlier, M.F. (1983) Proc. Nat!. Acad. Sci. USA 80, 7234. Hill, T.L. and Chen, Y. (1984) Proc. Nat!. Acad. Sci. USA 81, 5772. Chen, Y. and Hill, T.L. (1985) Proc. Nat!. Acad. Sci. USA 82, 1131. Hill, T.L. (1984) Proc. Nat!. Acad. Sci. USA 81, 6728. Rubin, R.I., to be published. Hill, T.L. (1985) Proc. Nat!. Acad. Sci. USA 82, 4404. Hill, T.L. (1963) Thermodynamics of Small Systems, Part I (Benjamin, New York). Neidl, C. and Engel, 1. (1979) Eur. 1. Biochem. 101, 163. Nicklas, R.B. (1983) 1. Cell Bio!. 97, 542. Hill, T.L. (1977) Free Energy Transduction in Biology (Academic, New York). DeRobertis, E.D.P. and DeRobertis, E.M.F. (1980) Cell and Molecular Biology (Saunders, Philadelphia), 7th Ed., p. 391. Nicklas, R.B. (1965) 1. Cell Bio!. 25, 119. Mitchison, T. (1986) 1. Cell Science (Supp!. 5) (in press). Hill, T.L. (1985) Proc. Nat!. Acad. Sci. USA 82, 431. Chen, Y. and Hill, T.L. (1985) Proc. Nat!. Acad. Sci. USA 82, 4127. Wegner, A. (1976) 1. Mol. Bio!. 108, 139. Hill, T.L. and Kirschner, M.W. (1982) Int. Rev. Cyto!. 78, 1. Chen, Y. and Hill, T.L. (1983) Proc. Nat!. Acad. Sci. USA 80, 7520. Haria, T. and Hotani, H. (1986) Nature 321, 605. Walker, R.A., Pryer, N.K., Cassimeris, L.u., Soboeiro, M., and Salmon, E.D. (1986) 1. Cell Bio!. 103, 432a. Carlier, M.F., Melki, R., Pantaloni, D., Hill, T.L., and Chen, Y. (1987) Proc. Nat!. Acad. Sci. USA 84 (in press). Magnus, W. and Oberhettinger, F. (1949) Formulas and Theorems for the Special Functions of Mathematical Physics (Chelsea, New York), p. 33. Goel, N.S. and Richter-Dyn, N. (1974) Stochastic Models in Biology (Academic, New York), p. 10.
Index Actin polymer 7,56,90,99, 129, 199, 227 nucleation of 104 Activity 6, 11 critical 24, 26, 35, 194 Activity coefficient 6, 14, 17, 116 Aggregates equilibrium 3, 199 macroscopic 23 non-interacting 15 steady-state 3, 199 on surface 22, 33 Aligned tubular polymer 138, 144, 193 Anaphase 57, 63, 263 Arrival times 289 Association equilibrium constants 11, 13,39,81 ATP 199,204,205 Axoneme 283 Boltzmann distribution 5, 140 Boundary condition absorption 48,271, 287, 289 reflection 46,271,299 Bragg-Williams approximation 201 Brownian motion 59 Brunauer-Emmett-Teller theory 35 Cap 122, 128 ofB on A 284 GTP 257,275 NTP 227,244,282 residual 270, 281 structural 126 Catalan numbers 149,290 Centrosome 104, 126,257,262,282, 283 Chemical potential 21, 116 Chromosome movement 56, 63, 264 Clusters 6, 20, 38 Colchicine 126 Compressibility of polymer 23, 25, 26 Configuration integral 10, 12 Cooperativity 112,241, 243 Copolymerization 110 Critical concentration 7,26,35,98, 123 Critical supersaturation 103 Cycles 215 Cytochalasin B 126 Detailed balance 37, 52, 61, 71, 92 at polymer surface 142 of two-component polymer 117 Diffusion coefficient 45, 93, 95 Dual aggregation 174, 190 Dynamic instability 244
304 Einstein model 28 Ensemble average 20, 35 Exchange reaction 228 First-passage time 49,95, 188,287 Fluctuations in length distribution 223 Flux of subunits 36, 40 Fokker-Planck differential equation 44 Force on polymer 22,24,51,263 Force-velocity curve 64 Free energy levels b~sic 209 gross 210 Free energy storage 216 Free energy transduction 55 efficiency of 56, 58, 262 Growth distribution in 292 sessions of 291 GTP 199,204,258,259,285 Hard interactions 7, 10, 12 polymer-polymer 88 Hard spheres 9, 16 Hemoglobin S 7, 16,30,56 nucleation of 103 Hooke's law 25, 73 Ising problem III Kinetochore 57, 63, 126,257,262,264 Label loss, rate of 48, 99 MAPs on micro tubules 129 Master equations 43, 105, 249, 265, 272 McGhee-von Hippel theory 131 McMillan-Mayer solution theory 6 Mean field approximation 201 Microtubules 7, 27, 56, 63, 84, 90, 129, 137, 167, 199,227 Index nucleation of 104 five-start model of 243, 252, 276 Moment of inertia 79 Monte Carlo calculations 153, 165, 180, 188,241,275 Muscle contraction 65, 205 Nucleation of polymers 21, 103 Oscillations in polymerization 278, 280, 284 Osmotic pressure 6, 132 Partition function canonical 4 completely open 18,22,34,81,86, 176,191,194 grand 5, 68, 74 polymer "surface" 140, 145, 178, 193 , Phase change 227, 232, 241, 283, 284 bioenergetics of 258 macroscopic model of 233, 283 rate constants 235, 245 Phase rule 114 Polymers annealing of 82, 94 attached 33, 219, 244, 284 between barriers 66, 72, 256 binding on 129, 134 with bound cap 122 disappearance of 98, 255, 269, 275 free 78, 220, 265 fragmentation of 82, 94, 106 macroscopic 23 multi-stranded 137 single-stranded 32 "surface" of 137 two-component 110, 174, 190, 284 Potential of mean force 8, 79 Quasi-chemical approximation 112 Rotational motion 21, 23, 28, 79, 83
305 Index Size distribution 35, 86, 250 Sleeve, polymer in 57, 264 Small system thermodynamics 19 Solubility of polymer 24, 30 Staggered tubular polymer 138, 156 Stokes law 93 Surface free energy 140 Surviving polymers 50, 98 Tau protein 129 Thermodynamic force 55, 203, 207 Time average 20,35 Transients 49, 98, 222, 278 attached polymer 42 growth of actin 105 two-phase model 254 Translational motion 20,21,23,28, 79, 83 Treadmilling 206,211,268,282 Tropomyosin 129 Tubulin 199 Vernier effect 174 Vernier enhancement 184 Vibrational motion 21, 28, 79 Virial expansion 8, 10, 15 Virus self-assembly 190 Work of subunit insertion 71