Author: Reddy J.N.  

Tags: mechanics  

ISBN: 047117985X

Year: 2002

Text
                    ENERGY PRINCIPLES
AND VARIATIONAL
METHODS IN APPLIED
MECHANICS
J. N. Reddy
Department of Mechanical Engineering
Texas A&M University
College Station, Texas
JOHN WILEY & SONS, INC.


CONTENTS PREFACE xv 1 INTRODUCTION 1 1.1 Preliminary Comments / 1 1.2 The Role of Energy Methods and Variational Principles / 2 1.3 Some Historical Comments / 2 1.4 Present Study / 4 References / 5 2 MATHEMATICAL PRELIMINARIES 8 2.1 Introduction / 8 2.2 Vectors / 9 2.2.1 Definition of a Vector / 9 2.2.2 Scalar and Vector Products / 11 2.2.3 Components of a Vector / 15 2.2.4 Summation Convention / 16 2.2.5 Vector Calculus / 19 2.2.6 Integral Relations / 24 2.3 Tensors / 29 2.3.1 Second-Order Tensors / 29 2.3.2 General Properties of a Dyadic / 32 * 2.3.3 Nonion Form of a Dyadic / 32 vii
CONTENTS 2.3.4 Eigenvectors Associated with Dyadics / 36 Exercises / 41 References / 47 REVIEW OF EQUATIONS OF SOLID MECHANICS 3.1 Introduction / 48 3.LI Classification of Equations / 48 3.1.2 Descriptions of Motion / 49 3.2 Conservation of Linear and Angular Momenta / 50 3.2.1 Equations of Motion / 50 3.2.2 Symmetry of Stress Tensor / 53 3.3 Kinematics of Deformation / 54 3.3.1 Strain Tensor / 54 3.3.2 Strain Compatibility Equations / 58 3.4 Constitutive Equations / 62 3.4.1 Introduction / 62 3.4.2 Generalized Hookers Law / 62 3.4.3 Plane Stress Constitutive Relations / 65 3.4.4 Thermoelastic Constitutive Relations / 66 Exercises / 69 References / 78 WORK, ENERGY, AND VARIATIONAL CALCULUS 4.1 Concepts of Work and Energy / 79 4.2 Strain Energy and Complementary Strain Energy / 84 4.3 Virtual Work / 95 4.4 Calculus of Variations / 102 4.4.1 The Variational Operator / 102 4.4.2 Functional / 106 4.4.3 The First Variation of a Functional / 106 4.4.4 Fundamental Lemma of Variational Calculus / 107 4.4.5 Externum of a Functional / 108 ' 4.4.6 Euler Equations / 110 4.4.7 Natural and Essential Boundary Conditions / 112 4.4.8 Minimization of Functionals with Equality Constraints / 117
CONTENTS JX Exercises / 123 References / 132 5 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS 133 5.1 Virtual Work Principles / 133 5.1.1 Introduction / 133 5.1.2 The Principle of Virtual Displacements / 133 5.1.3 Unit-Dummy-Displacement Method / 137 5.2 Principle of Total Potential Energy and Castigliano's Theorem! / 142 5.2.1 Principle of Minimum Total Potential Energy / 142 5.2.2 Castigliano's Theorem T / 146 5.3 Principles of Virtual Forces and Complementary Potential Energy / 151 5.4 Principle of Complementary Potential Energy and Castigliano's Theorem TI / 156 5.5 Betli's and Maxwell's Reciprocity Theorems / 164 Exercises / 169 References / 176 6 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE 177 6.1 Introduction / 177 6.2 Hamilton's Principle for Particles and Rigid Bodies / 177 6.3 Hamilton's Principle for a Continuum / 183 6.4 Hamilton's Principle for Constrained Systems / 190 6.5 Rayleigh's Method / 196 Exercises / 198 References / 203 7 DIRECT VARIATIONAL METHODS 204 7.1 Introduction / 204 7.2 Concepts from Functional Analysis / 205 7.2.1 General Introduction / 205 7.2.2 Linear Vector Spaces / 206 7.2.3 Normed and Inner Product Spaces / 211
X CONTENTS 7.2.4 Transformations, and Linear and Bilinear Forms / 215 7.2.5 Minimum of a Quadratic Functional / 216 7.3 The Ritz Method / 221 7.3.1 Introduction / 221 7.3.2 Description of the Method / 222 7.3.3 Properties of Approximation Functions / 225 7.3.4 Ritz Equations for the Parameters / 226 7.3.5 General Features of the Method / 230 7.3.6 Examples / 232 7.4 General Boundary-Value Problems / 245 7.4.1 Variational Formulations / 245 7.4.2 Ritz Approximations / 254 7.5 Weighted-Residual Methods / 259 7.5.1 Introduction / 259 7.5.2 Galerkin's Method / 262 7.5.3 Least-Squares Method / 263 7.5.4 Collocation Method / 263 7.5.5 Eigenvalue and Time-Dependent Problems / 264 7.5.6 Equations for Undetermined Parameters / 266 7.5.7 Examples / 267 7.6 Summary / 288 Exercises / 289 References / 297 8 THEORY AND ANALYSIS OF PLATES 299 8.1 Introduction / 299 8.1.1 General Comments / 299 8.1.2 An Overview of Plate/Shell Theories / 301 8.2 Classical Plate Theory / 305 8.2.1 Governing Equations of Circular Plates / 305 8.2.2 Analysis of Circular Plates / 312 , 8.2.3 Governing Equations in Rectangular Coordinates / 334 8.2.4 Navier Solutions of Rectangular Plates / 342 8.2.5 Levy Solutions of Rectangular Plates / 356 8.2.6 Variational Solutions: Bending / 362 8.2.7 Variational Solutions: Vibration / 379
CONTENTS Xi 8.2.8 Variational Solutions: Buckling / 384 8.3 Shear Deformation Plate Theory / 396 8.3.1 Governing Equations of Circular Plates / 396 8.3.2 Governing Equations in Rectangular Coordinates / 398 8.3.3 Exact Solutions of Axisymmetric Circular Plates / 401 8.3.4 Exact Solutions of Rectangular Plates / 405 8.3.5 Relationships Between Bending Solutions of Classical and Shear Deformation Theories / 412 8.3.6 Variational Solutions of Circular and Rectangular Plates / 421 Exercises / 425 References / 430 9 THE FINITE ELEMENT METHOD 433 9.1 Introduction / 433 9.2 Finite Element Analysis of Bars / 434 9.2.1 Governing Equation / 435 9.2.2 Representation of the Domain by Finite Elements / 435 9.2.3 Approximation over an Element / 436 9.2.4 Weak Form / 441 9.2.5 Finite Element Equations / 442 9.2.6 Assembly (or Connectivity) of Elements / 445 9.2.7 Imposition of the Boundary Conditions / 448 9.2.8 Calculation of Reactions and Derivatives of Solution: Postprocessing / 449 9.3 Finite Element Analysis of the Euler-Bernoulli Beam Theory / 454 9.3.1 Governing Equation / 454 9.3.2 Weak Form over an Element / 455 9.3.3 Derivation of the Approximation Functions / 456 9.3.4 Finite Element Model / 458 9.3.5 Assembly of Element Equations / 459 9.3.6 Imposition of Boundary Conditions / 461 9.4 Finite Element Models of the Timoshenko Beam Theory / 463 9.4.1 Governing Equations / 463 9.4.2 Displacement Finite Element Models / 464 9.4.3 Reduced Integration Element (RIE) / 466
Xli CONTENTS 9.4.4 Consistent Interpolation Element (CIE) / 467 9.4.5 Superconvergent Element (SCE) / 469 9.5 Finite Element Models of the Classical Plate Theory / 471 9.5.1 Introduction / 471 9.5.2 General Formulation / 472 9.5.3 Conforming and Nonconforming Plate Elements / 474 9.5.4 Fully Discretized Finite Element Models / 475 9.6 Finite Element Models of the First-Order Shear Deformation Plate Theory / 479 9.6.1 Governing Equations and Weak Forms / 479 9.6.2 Finite Element Model / 480 9.6.3 Numerical Integration / 483 Exercises / 493 References / 499 10 MIXED VARIATIONAL FORMULATIONS 502 10.1 Introduction / 502 10.1.1 General Comments / 502 10.1.2 Mixed Variational Principles / 502 10.1.3 Extremum and Stationary Behavior of Functionals / 505 10.2 Stationary Variational Principles / 506 10.2.1 The Minimum Total Potential Energy Principle / 506 10.2.2 The Hellinger-Reissner Variational Principle / 508 10.2.3 The Reissner Variational Principle / 513 10.3 Variational Solutions Based on Mixed Formulations / 514 10.4 Mixed Finite Element Models of Beams / 517 10.4.1 The Euler-Bernoulli Beam Theory / 517 10.4.2 The Timoshenko Beam Theory / 523 10.5 Mixed Finite Element Models of the Classical Plate Theory / 527 10.5.1 Preliminary Comments / 527 10.5.2 Mixed Model I / 527 10.5.3 Mixed Model II / 532 10.6 Closure / 539 Exercises / 539 References / 542
CONTENTS XiM ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 544 INDEX 583 ABOUT THE AUTHOR 591
PREFACE The increasing use of numerical and computational methods in engineering and applied sciences has shed new light on the importance of energy principles and variational methods. The number of engineering courses that make use of energy principles and variational formulations and methods has also grown very rapidly in recent years. In view of the increase in the use of the variational formulations and methods (including the finite element method), there is a need to introduce the concepts of energy principles and variational methods and their use in the formulation and solution of problems of mechanics to both undergraduate and beginning graduate students. This book, an extensively revised version of the author's earlier book Energy and Variational Methods in Applied Mechanics, is intended for senior undergraduate students and beginning graduate students in aerospace, civil, and mechanical engineering and applied mechanics, who have had a course in fundamental engineering subjects as well as in ordinary and partial differential equations. The book is organized into ten chapters and is self-contained as far as the subject matter is concerned. Chapter 1 presents a general introduction to the subject of variational principles. Chapter 2 contains a brief review of the algebra and calculus of vectors and Cartesian tensors. A review of the basic equations of linear solid continuum mechanics is included in Chapter 3. These equations are frequently referred to in subsequent chapters. Much of the material presented in Chapters 1 through 3 can be assigned as reading material, especially in a graduate class. Chapter 4 deals with the concepts of work, energy, and the basic topics from variational calculus, including Euler equations, the fundamental lemma of calculus of variations, essential and natural boundary conditions, and minimization of functionals without and with equality constraints. Virtual work and energy principles and energy methods of solid and structural mechanics are presented in Chapter 5. Chapter 6 is devoted to a discussion of Hamilton's principle for dynamical systems. Classical variational methods of approximation (e.g., the methods of Ritz, Galerkin, Kantorovich, etc.) are presented in Chapter 7. All of the concepts and methods xv
XVi PREFACE presented in Chapters 4 through 7 arc illustrated using bars and beams, although the methods discussed in Chapter 7 are readily applicable to field problems whose differential equations resemble those of bars and beams. Chapter 8 is dedicated to applications of (he energy principles and variational methods developed in earlier chapters to circular and rectangular plates. In the interest of completeness and for use as a reference for approximate solutions, exact solutions are also included. The finite element method is introduced in Chapter 9, with applications to beams and plates. Displacement finite element models of Euler-Bernoulli and Timoshenko beam theories and classical and first-order shear deformation plate theories are presented. A unified approach, more general than that found in most solid mechanics books, is used to introduce the finite element method. As a result, the student can readily extend the method to other subject areas of solid mechanics as well as to other branches of engineering. Lastly, the mixed variational principles of Hellinger and Reissner for elasticity are derived in Chapter 10. Mixed variational formulations, including mixed finite element models of beams and plates, are discussed. Each chapter of the book contains many example problems and exercises that illustrate, test, and broaden the understanding of the topics covered. A list of references, by no means complete or up-to-date, is also provided at the end of each chapter. Answers to selected problems arc included at the end of the book. The book is suitable as a textbook for a senior undergraduate course or a first-year graduate course on energy principles and variational methods taught in aerospace, civil, and mechanical engineering and applied mechanics departments. To gain the most from the text, the student should have a senior undergraduate or first-year graduate standing in engineering. Some familiarity with basic courses in differential equations, mechanics of materials, and dynamics would also be helpful. T have benefited by the professional works, encouragement, and support of many colleagues as well as students who have taught me how to explain complicated concepts in simple terms. While it is not possible to name all of them, without their help and support it would not have been possible for me to make some modest contributions to the field of mechanics through teaching, research, and writing. My sincere thanks are due to my teacher, Professor J. T, Odcn (University of Texas at Austin), for the many things I learned from him which have been useful ail my life. In the same vein I wish to thank Professor C. W. Bert (University of Oklahoma, Norman) for giving me the opportunity to teach and develop into what I am. I am very grateful for their mentorship, advice, and support. Deep gratitude is due to my wife for her patience and support while I am occupied with the writing of this book and others, and for her smile when I say every day "T have so much to do." Most books are not free of errors, especially those with many mathematical equations and numbers. I wish to thank in advance those readers who are willing to draw attention to typos and errors, using the e-mail address: jnreddy@hotmail.com. J. N. Reddy College Station, Texas
1 INTRODUCTION 1.1 PRELIMINARY COMMENTS The phrase "energy methods" in the present study refers to methods that make use of the total potential energy (i.e., strain energy and potential energy due to applied loads) of a system to obtain values of an unknown displacement or force, at a specific point of the system. These include Castigliano's theorems, unit-dummy-load and unit-dummy-displacement methods, and Betti's and Maxwell's theorems. These methods are often limited to the (exact) determination of generalized displacements or forces at fixed points in the structure; in most cases, they cannot be used to detennine the complete solution (i.e., displacements and/or forces) as a function of position in the structure. The phrase 'Variational methods," on the other hand, refers to methods that make use of the variational principles, such as the principles of virtual work and the principle of minimum total potential energy, to determine approximate solutions as continuous functions of position in a body. In the classical sense, a variational principle has to do with the minimization or finding stationary values of a functional with respect to a set of undetermined parameters introduced in the assumed solution. The functional represents the total energy of the system in solid and structural mechanics problems, and in other problems it is simply an integral representation of the governing equations. In all cases, the functional includes all the intrinsic features of the problem, such as the governing equations, boundary and/or initial conditions, and constraint conditions. 1
2 INTRODUCTION 1.2 THE ROLE OF ENERGY METHODS AND VARIATIONAL PRINCIPLES Variational principles have always played an important role in mechanics. Variational formulations can be useful in three related ways. First, many problems of mechanics are posed in terms of finding the extremum (i.e., minima or maxima) and thus, by their nature, can be formulated in terms of variational statements. Second, there are problems that can be formulated by other means, such as by vector mechanics (e.g., Newton's laws), but these can also be formulated by means of variational pri nciples. Third, variational formulations form a powerful basis for obtaining approximate solutions to practical problems, many of which are intractable otherwise. The principle of minimum total potential energy, for example, can be regarded as a substitute for the equations of equilibrium of an elastic body, as well as a basis for the development of displacement finite element models that can be used to determine approximate displacement and stress fields in the body. Variational formulations can also serve to unify diverse fields, suggest new theories, and provide a powerful means for studying the existence and uniqueness of solutions to problems. In many cases they can also be used to establish upper and/or lower bounds on approximate solutions, 1.3 SOME HISTORICAL COMMENTS In modern times, the term "variational formulation" applies to a wide spectrum of concepts having to do with weak, generalized, or direct variational formulations of boundary- and initial-value problems. Still, many of the essential features of variational methods remain the same as they were over 200 years ago when the first notions of variational calculus began to be formulated. Although Archimedes (287-212 B.C.) is generally credited with the first to use work arguments in his study of levers, the most primitive ideas of variational theory (the minimum hypothesis) are present in the writings of the Greek philosopher Aristotle (384-322 B.C.), to be revived again by the Italian mathematician/engineer Galileo (1564-1642), and finally formulated into a Principle of Least Time by the French mathematician Fermat (1601-1665). The phrase virtual velocities was used by Jean Bernoulli in 1717 in his letter to Varignon (1654-1722). The development of early variational calculus, by which we mean the classical problems associated with minimizing certain functionals, had to await the works of Newton (1642-1727) and Leibniz (1646-1716). The earliest applications of such variational ideas included the classical isoperimetric problem of finding among closed curves of given length the one that encloses the greatest area, and Newton's problem of determining the solid of revolution of "minimum resistance." In 1696, Jean Bernoulli proposed the problem of the brachistochrone: among all curves connecting two points, find the curve traversed in the shortest time by a particle under the influence of gravity. It stood as a challenge to the mathematicians of their day to solve the problem using the rudimentary tools of
SOME HISTORICAL COMMENTS 3 analysis then available to them or whatever new ones they were capable of developing. Solutions to this problem were presented by some of the greatest mathematicians of the time: Leibniz, Jean Bernoulli's older brother Jacques Bernoulli, L'Hdpital, and Newton. The first step toward developing a general method for solving variational problems was given by the Swiss genius Leonhard Euler (1707-1783) in 1732 when he presented a "general solution of the isoperimetric problem," although Maupertuis is credited with having put forward a law of minimal property of potential energy for stable equilibrium in his Memoires de VAcademic des Sciences in 1740. It was in Euler's 1732 work and subsequent publication of the principle of least action (in his book Methodus inveniendi tineas curvas ...) in 1744 by Euler that variational concepts found a welcome and permanent home in mechanics. He developed all ideas surrounding the principle of minimum potential energy in his work on the Elastica, and he demonstrated the relationship between his variational equations and those governing the flexure and budding of thin rods. A great impetus to the development of variational mechanics began in the writings of Lagrange (1736-1813), first in his correspondence with Euler. Euler worked intensely in developing Lagrange's method, but delayed publishing his results until Lagrange's works were published in 1760 and 1761. Lagrange used d'Alembert's principle to convert dynamics to statics and then used the principle of virtual displacements to derive his famous equations governing the laws of dynamics in terms of kinetic and potential energy. Euler's work, together with Lagrange's Mecanique analytique of 1788, laid down the basis for the variational theory of dynamical systems. Further generalizations appeared in the fundamental work of Hamilton in 1834. Collectively, all these works have had a monumental impact on virtually every branch of mechanics. A more solid mathematical basis for variational theory began to be developed in the eighteenth and early nineteenth century. Necessary conditions for the existence of "minimizing curves" of certain functionals were studied during this period, and we find among contributors of that era the familiar names of Legendre, Jacobi, and Weierstrass. Legendre gave criteria for distinguishing between maxima and minima in 1786, without considering criteria for existence, and Jacobi gave sufficient conditions for existence of extrema in 1837. Amore rigorous theory of existence of extrema was put together by Weierstrass, who, with Erdmann, established in 1865 conditions on extrema for variational problems involving corner behavior. During the last half of the nineteenth century, the use of variational ideas was widespread among leaders in theoretical mechanics. We mention the works of Kirchhoff on plate theory, Lame, Green, and Kelvin on elasticity, and the works of Bctti, Maxwell, Castigliano, Menabrea, and Engesser for discrete structural systems. Lame was the first in 1852 to prove a work equation, named after his colleague Clapeyron, for deformable bodies. Lame's equation was used by Maxwell [1] for the solution of redundant frame works using the unit-dummy-load technique. In 1875 Castigliano published an extremum version of this technique, but attributed the idea to Menabrea. A generalization of Castigliano's work is due to Engesser [2].
4 INTRODUCTION Among prominent contributors to the subject near the end of the nineteenth century and in the early years of the twentieth century, particularly in the area of variational methods of approximation and their applications to physical problems, were Rayleigh [3J, Ritz |4'J, and Galerkin [5]. Modern variational principles began in the 1950s with the works of Hellinger [6] and Reissner [7,8] on mixed variational principles for elasticity problems. A variety of generalizations of classical variational principles have appeared, and we shall not describe them here. In closing this section, we note that a short historical account of early variational methods in mechanics can be found in the book of Lanczos [9] and a brief review of certain aspects of the subject as it stood in the early 1950s can be found in the book of Truesdell and Toupin [10]; additional information can be found in Smith's history of mathematics [11] and in the historical treatises on mechanics by Mach [12], Dugas [13], and Timoshenko [14]. Reference to much of the relevant contemporary literature can be found in the books by Washizu [15] and Oden and Reddy [16]. Additional historical papers and textbooks on variational methods arc listed at the end of this chapter (sec [17-56]). 1.4 PRESENT STUDY The objective of the present study is to introduce energy methods and variational principles of solid and structural mechanics and to illustrate their use in the derivation and solution of the equations of applied mechanics, including plane elasticity, beams, frames, and plates. Of course, variational formulations and methods presented in this book are also applicable to problems outside solid mechanics. To equip the reader with the necessary mathematical tools and background from the theory of elasticity that are useful in the sequel, a review of vectors, matrices, tensors, and governing equations of elasticity are provided in the next two chapters. To keep the scope of the book within reasonable limits, only linear problems are considered. Although stability and vibration problems are introduced via examples and exercises, a detailed study of these topics is omitted. In the following chapter we summarize the algebra and calculus of vectors and tensors. Tn Chapter 3 we give a brief review of the equations of solid mechanics, and in Chapter 4 we present the concepts of work and energy, energy principles, and Castigliano's theorems of structural mechanics. In Chapter 5 we present principles of virtual work, potential energy, and complementary energy. Chapter 6 is dedicated to Hamilton's principle for dynamical systems, and in Chapter 7 we introduce the Ritz, Galerkin, and weighted-residual methods. In Chapter 8, applications of variational methods to the formulation of plate bending theories and their solution by variational methods are presented. For the sake of completeness and comparison, analytical solutions of bending, vibration, and budding of circular and rectangular plates are also presented. An introduction to the finite element method and its application to displacement finite element models of beams and plates is discussed in Chapter 9. The final chapter, Chapter 10, is devoted to the discussion of mixed variational principles,
REFERENCES 5 and mixed finite clement models of beams and plates. To keep the scope of the book within reasonable limits, theory and analysis of shells is not included. REFERENCES 1. Maxwell, J. C. "On the calculation of the equilibrium and the stiffness of frames," Phil. Mag. Set: 4, 27, 294 (1864). 2. Engesscr, E, "Uber sialisch unbestimmte Trager bei beliebigen Formanderungs Gesetze und uber den Satz von der Kleinsten Erganzungsarbcit," Zeitschrift des Architekten und Ingenieur-Vereins zu Hanover, 35, 733-744 (1889). 3. Lord Rayleigh, (John William Strutt), Theoiy of Sound, Macmillan, London, 1877. Reprinted by Dover, New York (1945). 4. Ritz, W., "Uber eine neue Melhode zur Losung gewisser Variationsproblemc der mathematischen Physik," Journal fur reine und angewandte Mathematik, 135, 1-61 (1908). 5. Galerkin, B. G, "Series solutions in rods and plates," Vestnik Inzhenerov i Tekhnikov, 19 (1915). 6. Hellinger, E., "Die allgcmeinen Ansatze der Mechanik der Kontinua," Enzyclopadie der Mathematischen Wissenschaften IV, 4, 654-655 (1914). 7. Reissncr, E., "Note on the method of complementary energy," /. Math. Phys., 27,159-160 (1948). 8. Reissner, E., "On a variational theorem for finite deformation,"./. Math. Phys., 32, 129— 135(1953). 9. Lanczos, C, The Variational Principles of Mechanics, 4th edition, Dover, New York (1986). 10. Tmesdell, C. A., and Toupin, R. A., "The classical field theories," Encyclopedia of Physics, S. Fltigge (ed.), Springet-Veilag, Berlin (1960). 11. Smith, D. E., History of Mathematics, Vol. 1 and II, Dover, New York (1951). 12. Mach, E., The Science of Mechanics, translated and published by Court 1L(1919). 13. Dugas, R., A History of Mechanics, translated into English by J. R. Maddox, Dover, New York (1988). First published by Editions du Griffon, Neuchatel, Switzerland in 1955. 14. Timoshenko, S. P., History of Strength of Materials, Dover, New York (1983). Originally published in 1953 by McGraw-Hill, New York. 15. Washizu, K., Variational Methods in Elasticity and Plasticity, Pergamon Press, New York (1967). 16. Oden, J. T., andReddy, J. N., Variational Principles in Theoretical Mechanics, 2nd edition, Springer-Vcrlag, Berlin (1983). 17. Argyris, J. H,, and Kelsey, S., Energy Theorems and Structural Analysis, Buttcrworths, London (1960). 18. Arthurs, A. M., Complementary Variational Principles, Oxford University Press, London (1967).
6 INTRODUCTION 19. Becker, M., The Principles and Applications of Variational Methods, MIT Press, Cambridge, MA (1964). 20. Biot, M. A., Variational Principles in Heat Transfer, Clarendon, London (1972). 21. Charlton, T. M., Principles of Structural Analysis, Longmans (1969). 22. Charlton, T. M., Energy Principles in Theory of Structures, Oxford University Press, Oxford (1973). 23. Davics, G. A. O., Virtual Work in Structural Analysis, John Wiley, Chichester (1982). 24. Denn, M. M., Optimization by Variational Methods, McGraw-Hill, New York (1969). 25. Ekeland, I., and Tcmam, R., Convex Analysis and Variational Problems, North-Holland, New York (1976). 26. Finlayson, B. A,, The Method of Weighted Residuals and Variational Principles, Academic Press, New York (1972). 27. Forray, M. J., Variational Calculus in Science and Engineering, McGraw-Hill, New York (1968). 28. Gregory, M. S,, An Introduction to Extremum Principles, Butterworths (1969). 29. Hestenes, M. R., Optimization Theory: The Finite Dimensional Case, John Wiley, New York (1975). 30. Hiidebrand, F. B., Methods of Applied Mathematics, 2nd edition, Prentice-Hall, Englewood Cliffs, NJ (1965). 31. Kantorovitch, L. V., and Krylov, V. J., Approximate Methods of Higher Analysis, translated by C. D. Benster, Noordhoff, Groningen (1958). 32. Leipholz, H., Direct Variational Methods and Eigenvalue Problems in Engineerings Noordhoff, Leyden (1977). 33. Lippmann, H., Extremum and Variational Principles in Mechanics, Springer-Veriag, New York (1972). 34. Lovelock, D., Tensors, Differential Forms and Variational Principles, Wiley, New York (1974). 35. McGuire, W., Gallagher, R. H., and Ziemian, R. D., Matrix Structural Analysis, 2nd edition, John Wiley, New York (2000). 36. Mikhlin, S. G, Variational Methods in Mathematical Physics, Pergamon Press (distributed by Macmillan), New York (1974). 37. Mikhlin, S, G, Mathematical Physics, An Advanced Course, North-Holland, Amsterdam (1970). 38. Mikhlin, S. G, The Numerical Performance of Variational Methods, Wolters-Noordhoff, Groningen (1971). 39. Oden, J. T., and Ripperger, E. A., Mechanics of Elastic Structures, 2nd edition, Hemisphere/McGraw-Hill, New York (1981). 40. Parkus, H., Variational Principles in Thermo- and Magneto-Elasticity, Lecture Notes, No. 58, Springer-Verlag, New York (1972). 41. Petrov, L P., Variational Methods in Optimal Control Theory, translated from the 1965 Russian edition, Academic Press, New York (1968).
REFERENCES 7 42. Prenter, P. M., Splines and Variational Methods, John Wiley, New York (1975). 43. Reddy, J. N., An Introduction to the Finite Element Method, 2nd edition, McGraw-Hill, New York (1993). 44. Reddy, J. N., and Rasmussen, M. L., Advanced Engineering Analysis, John Wiley, New York (1982); reprinted by Krieger, Melbourne, FL (1990), 45. Reddy, J. N., Energy and Variational Methods in Applied Mechanics* John Wiley, New York (1984). 46. Reddy, J. N., Applied Functional Analysis and Variational Methods in Engineering, McGraw-Hill, New York (1986). 47. Reddy, J. N., Mechanics of Laminated Composite Plates, Theory and Analysis, CRC Press, Boca Raton, FL (1997). 48. Reddy, J. N., Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia (1999). 49. Rektorys, K., Variational Methods in Mathematics, Science and Engineering, Reidel, Boston (1977). 50. Richards, T. R, Energy Methods in Stress Analysis, Ellis Horwood (distributed by John Wiley & Sons), Chichester, UK (1977). 51. Rustagi, J. S., Variational Methods in Statistics, Academic Press, New York (1976). 52. Schechter, R. S., The Variational Methods in Engineering, McGraw-Hill, New York (1967), 53. Shames, I. H., and Dym, C. L., Energy and Finite Element Methods in Structural Mechanics, Hemisphere, Washington, DC (1985). 54. Shaw, F. S., Virtual Displacements and Analysis of Structures, Prentice-Hall, Englewood Cliffs, NJ (1972). 55. Smith, D. K., Variational Methods in Optimization, Prentice-Hall, Englewood Cliffs, NJ (1974). 56. Stacey, W. M., Variational Methods in Nuclear Reactor Physics, Academic Press, New York (1974).
2 MATHEMATICAL PRELIMINARIES 2.1 INTRODUCTION Our approach in this book is evolutionary; that is, we wish to begin with concepts that are simple and intuitive, and then generalize these concepls step by step into a broader and more abstract body of analysis. This is a natural inductive approach, more or less in accord with the development of the subject of variational methods. In analyzing physical phenomena, we set up relations between various quantities that characterize the phenomena (such as Newton's laws, energy conservation, etc.). As a means of expressing a natural law, a coordinate system in a chosen frame of reference can be introduced, and the various physical quantities involved can be expressed in terms of measurements made in that system. The mathematical form of the law thus depends upon the chosen coordinate system and may appear different in another type of coordinate system. The laws of nature, however, should be independent of the artificial choice of a coordinate system, and we may seek to represent the law in a manner independent of a particular coordinate system. A way of doing this is provided by vector and tensor analysis. When vector notation is used, a particular coordinate system need not be introduced. Consequently, the use of vector notation in formulating natural laws leaves them invariant to coordinate transformations. A study of physical phenomena by means of vector equations often leads to a deeper understanding of the problem in addition to bringing simplicity and versatility into the analysis. t The term vector is used often to imply a physical vector that has "magnitude and direction" and obeys certain rules of vector addition and scalar multiplication. In the sequel we consider more general, abstract objects than physical vectors, which are also called vectors. It transpires that the physical vector is a special case of what is known as a "vector from a linear vector space." Then the notion of vectors in modern 8
VECTORS 9 mathematical analysis is an abstraction of the elementary notion of a physical vector. While the definition of a vector in abstract analysis does not require the vector to have a magnitude, in nearly all cases of practical interest the vector is endowed with a magnitude, in which case the vector is said to belong to a normed vector space. Like physical vectors, which have direction and magnitude and satisfy the parallelogram law of addition, tensors are more general objects that are endowed with a magnitude and multiple direction(s) and satisfy rules of tensor addition and scalar multiplication. In fact, vectors are often termed the first-order tensors. As will be shown shortly, the stress (i.e., force per unit area) requires a magnitude and two directions—one normal to the plane on which the stress is measured and the other is the direction of the force—to specify it uniquely. In this chapter we review basic elements of vector and tensor analysis that are of use in the sequel. For additional reading, the books listed at the end of the chapter [1-27] may be consulted. 2.2 VECTORS 2.2.1 Definition of a Vector In the analysis of physical phenomena we are concerned with quantities that may be classified according to the information needed to specify them completely. Consider the following two groups: Scalars Nonscalars Mass Force Temperature Moment Density Stress Volume Acceleration Time Displacement After units have been selected, the scalars arc given by a single number. Nonscalars need not only a magnitude specified, but also additional information, such as direction. Nonscalars that obey certain rules (such as the parallelogram law of addition) are called vectors. Not all nonscalar quantities are vectors. The specification of a stress requires not only a force, which is a vector, but also an area upon which the force acts. A stress is a second-order tensor, as will be shown shortly. In written or typed material, it is customary to place an arrow or a bar over the letter denoting the vector, such as A. Sometimes the typesetter's mark of a tilde under the letter is used. In printed material the vector letter is denoted by a boldface letter, A, such as used in this book. The magnitude of the vector A is given by |A| or just A. Two vectors A and B are equal if their magnitudes are equal, |A| = |B|, and if their directions and sense arc equal. Consequently a vector is not changed if it is moved parallel to itself. This means that the position of a vector in space may be
10 MATHEMATICAL PRELIMINARIES B A+B=B+A ° B (a) (b) Figure 2.1 (a) Addition of vectors, (b) Parallelogram law of addition. chosen arbitrarily. In certain applications, however, the actual point of location of a vector may be important (for instance, a moment or a force acting on a body). A vector associated with a given point is known as a localized or bound vector. Let A and B be any two vectors. Then we can add them as shown in Fig. 2. la. The combination of the two diagrams in Fig. 2.1a gives the parallelogram shown in Fig. 2.1b. Thus wc say the vectors add according to the parallelogram law of addition so that C = A + B=B + A. (2.1) We thus see that vector addition is commutative. Subtraction of vectors is earned out along the same lines. To form the difference A — B, we write A - B = A + (~B) (2.2) and subtraction reduces to the operation of addition. The negative vector —B has the same magnitude as B, but has the opposite sense. With the rules of addition in place, we can define a (geometric) vector. A vector is a quantity that possesses both magnitude and direction and obeys the parallelogram law of addition. Obeying the law is important because there are quantities having both magnitude and direction that do not obey this law. A finite rotation of a rigid body is not a vector, although infinitesimal rotations are. The definition given above is a geometrical definition. That vectors can be represented graphically is an incidental rather than a fundamental feature of the vector concept. A vector of unit length is called a unit vector. The unit vector may be defined as follows: &A = 7. ' (2.3) A We may now write /^■iS? A = Ae^. (2.4)
VECTORS 11 Thus any vector may he represented as a product of its magnitude and a unit vector. A unit vector is used to designate direction. It does not have any physical dimensions. We denote a unit vector by a "hat" (caret) above the boldface letter. A vector of zero magnitude is called a zero vector or a null vector. All null vectors are considered equal to each other without consideration as to direction: A + 0 = A and 0A = 0. (2.5) The laws that govern addition, subtraction, and scalar multiplication of vectors are identical with those governing the operations of scalar algebra. 2.2.2 Scalar and Vector Products Besides addition, subtraction, and multiplication by a scalar, we must consider multiplication of two vectors. There are several ways the product of two vectors can be defined. We consider first the so-called scalar product. Let us recall the concept of work. When a force F acts on a mass point and moves through an infinitesimal displacement vector ds, the work done by the force vector is defined by the projection of the force in the direction of the displacement times the magnitude of the displacement (see Fig. 2.2). Such an operation may be defined for any two vectors. Since the result of the product is a scalar, it is called the scalar product. We denote this product as follows: F ds = Fds cos 0, 0 < 0 < jr. (2.6) The scalar product is also known as the dot product or inner product. To understand the vector product, consider the concept of the moment due to a force. Let us describe the moment about a point O of a force F acting at a point P, such as shown in Fig. 2.3a. By definition, the magnitude of the moment is given by M = FU F=|F|f (2.7) where / is the lever arm for the force about the point O. If r denotes the vector OP and 0 the angle between r and F as shown, such that 0 < 9 < n, we have I = r sin 0, and thus M = Fr sinO. (2.8) F / ! *—> ■ ds Figure 2.2 Representation of work.
12 MATHEMATICAL PRELIMINARIES O^ \Wp (a) (b) Figure 2.3 (a) Representation of a moment, (b) Direction of rotation. ^ J —/ 1 \ C=AxB\A X"" r_ 7 *s' \ r s \ (a) (b) Figure 2.4 (a) Axis of rotation, (b) Representation of the vector. A direction can now be assigned to the moment. Drawing the vectors F and r from the common origin O, we note that the rotation due to F tends to bring r into F (see Fig. 2.3b). Wc now set up an axis of rotation perpendicular to the plane formed by F and r. Along this axis of rotation we set up a preferred direction as that in which a right-handed screw would advance when turned in the direction of rotation due to the moment (see Fig. 2.4a). Along this axis of rotation we draw a unit vector &m and agree that it represents the direction of the moment M. Thus we have M = Fr sin 0 e> = r x F. (2.9) According to this expression, M may be looked upon as resulting from a special operation between the two vectors F and r. It is thus the basis for defining a product between any two vectors. Since the result of such a product is a vector, it may be called the vector product. The vector product of two vectors A and B is a vector C whose magnitude is equal to the product of the magnitude of A and B times the sine of the angle measured from A to B such that 0 < 9 < k. and whose direction is specified by the condition that C be perpendicular to the plane of the vectors A and B and' points in the direction which a right-handed screw advances when turned so as to bring A into B. The vector product is usually denoted by C = AxB = AB sin(A, B) e, (2.10)
VECTORS 13 BxC A Figure 2.5 Scalar triple product as the volume of a parallelepiped. where sin(A> B) denotes the sine of the angle between vectors A and B. This product is called the cross product, skew product, and also outer product, as well as the vector product (see Fig. 2.4b). Now consider the various products of three vectors: A(B-C), A-(BxC), Ax (BxC), (2.11) The product A(B • C) is merely a multiplication of the vector A by the scalar B • C. The product A -(BxC) is a scalar. It can be seen that the product A • (B x C), except for the algebraic sign, is the volume of the parallelepiped formed by the vectors A, B, and C, as shown in Fig. 2.5. We also note the following properties: 1. The dot and cross can be interchanged without changing the value: ABxC = AxBC= [ABC]. (2.12) 2. A cyclical permutation of the order of the vectors leaves the result unchanged: ABxC = CAxB = BCxA = [ABC], (2.13) 3. If the cyclic order is changed, the sign changes: ABxC = -ACxB = -C-BxA = -B AxC. (2.14) 4. A necessary and sufficient condition for any three vectors, A, B, C to be copla- nar is that A • (B x C) = 0, Note also that the scalar triple product is zero when any two vectors are the same. The product A x (B x C) is a vector normal to the plane formed by A and (BxC). The vector (B x C), however, is perpendicular to the plane formed by B and C, This means that A x (B x C) lies in the plane formed by B and C and is perpendicular to A (see Fig. 2.6). Thus A x (B x C) can be expressed as a linear combination of B andC: Ax(BxC)= miB + n\C. (2,15) p
14 MATHEMATICAL PRELIMINARIES Figure 2.6 The vector triple product. Figure 2.7 Plane perpendicular to A, and passing through the terminal point of B. Likewise, we would find that (A x B) x C = m2A + «2B. (2.16) Thus the parentheses cannot be interchanged or removed. It can be shown that m i = A • C, n i = —A • B, and hence that A x (B x C) = (A • C)B - (A ■ B)C. (2.17) Example 2.1 The equation of a plane perpendicular to a vector A and passing through the terminal point of vector B can be obtained without the use of any coordinate system (see Fig. 2.7). Let O be the origin and B the terminal point of vector B. Draw a directed line segment from O to Q, such that OQ is parallel to A and Q is in the plane. Then OQ = aA, where a is a scalar. Let P be an arbitrary point on the line BQ. If the position vector of the point P is r, then BP = OP - OB = r - B. i Since BP is perpendicular to OQ = or A, we must have BP OQ = 0 or (r - B) • A = 0, which is the equation of the plane in question.
VECTORS 15 The perpendicular distance from point O to the plane is the magnitude of OQ. However, we do not know its magnitude (or, a is not known). The distance is also given by the projection of vector B along OQ: OQ d = B — = B- e*. |OQ| *' where e^ is the unit vector along A, e^ = A/A. Example 2.2 Let A and B be any two vectors in space. Then vector A can be expressed in terms of components along (i.e., parallel) and perpendicular to B: the component of A along B is given by (A - e#), where e# = B/Z?. The component of A perpendicular to B and in the plane of A and B is given by the vector triple product e# x(Axeg). Thus, A = (A ■ e5)e5 + e# x (A x e5). (a) Alternately, using Eq. (a) with A = C = e^ and B = A, we obtain e# x(AxeB)=A" (eB - A)e/y. 2.2.3 Components of a Vector So far we have proceeded on a geometrical description of a vector as a directed line segment. We now embark on an analytical description of a vector and some of the operations associated with this description. Such a description yields a connection between vectors and ordinary numbers and relates operation on vectors with those on numbers. The analytical description is based on the notion of components of a vector. In what follows, we shall consider a three-dimensional space, and the extensions to n dimensions will be evident (except for a few exceptions). A set of n vectors is said to be linearly dependent if a set of n numbers jSi,ft,...,Ai can be found such that Pi Ai + fcA2 + ■ ■ • + AiA„ = 0, (2.18) where Pi, fh>. *.<, fin cannot all be zero. If this expression cannot be satisfied, the vectors are said to be linearly independent In a three-dimensional space a set of no more than three linearly independent vectors can be found. Let us choose any set and denote it as follows: e1.e2.e3. (2.19) This set is called a basis (or a base system). It is clear from the concept of linear dependence that we can represent any vector in three-dimensional space as a linear combination of the basis vectors (see Fig. 2.8): A = Aiei + A2e2 + A3«v (2.20)
16 MATHEMATICAL PRELIMINARIES Figure 2.8 Components of a vector. The vectors A\e], A2e2, and A3e3 are called the vector components of A, and Ai, A2, and A3 are called scalar components or measure numbers of A associated with the basis (e\, e2, 63), Also, we use the notation A — (A\, A2, A3) to denote a vector by its components. 2.2.4 Summation Convention It is useful to abbreviate a summation of terms by understanding that a repeated index means summation over all values of that index. Thus the summation 3 A = ]TA/e/ (2.21) 1=1 can be shortened to A = A,e;. (2.22) The repeated index is a dummy index and thus can be replaced by any other symbol that has not already been used. Thus we can also write A = A,e; = A,„em, (2.23) and so on, When a basis is unit and orthogonal, that is, orthonormal, wc have [eie2e3l=l. (2.24) In many situations an orthonormal basis simplifies calculations. For an orthonormal basis the vectors A and B can be written as A = Aiei + A2e2 + A3e3 = A/e; B = BLei + B2e2 + B3e3 = ^e,, where e,- = e,- (/ = 1, 2, 3) is the orthonormal basis, and A/ and Bj are the corresponding physical components (i.e., the components have the same physical dimensions as the vector).
VECTORS 17 It is convenient at this time to introduce the alternating symbol Sjjk for representing the cross product of two orthonormal vectors in a right-handed basis system. We define the cross product e,- x ey for a right-handed system as 6/ X ey = Sijkfya (2.25) where Sijk = 1, if /, j, k are in cyclic order and not repeated (i ^ j ^ /c), — I, if z, 7, k are not in cyclic order and not repeated (i ^ j ^ /c), 0, if any of /, j\ k are repeated. (2.26) The symbol Sijk is called the alternating symbol or permutation symbol In an orthonormal basis the scalar and vector products can be expressed in the index form using the Kronecker delta 8jj = e, • ey, and the alternating symbols: AB- (A/e*) • (Bjej) = AjBjSij = AtBh A x B = (Ajtj) x (Bj&j) = AiBfsUkek. (2.27) Further, the Kronecker delta and the permutation symbol are related by the identity, known as the 8-8 identity, Gijk&imn — °jmbhi ~ bjn°kw (2.28) The permutation symbol and the Kronecker delta prove to be very useful in proving vector identities. Since a vector form of any identity is invariant (i.e., valid in any coordinate system), it suffices to prove it in one coordinate system. In particular, an orthonormal system is very convenient because of the permutation symbol and the Kronecker delta. The following example illustrates some of the uses of Si j and £;77c. Example 23 We wish to express the vector operation (A x B) • (C x D) in an alternate vector form (to establish a vector identity): (A x B) • (C x D) = (AiBjEijk&k) ■ (C(nDne„wpep) = ^ /' Bj C m Dn Eijk £mnp Okp = A i Bj Cm On Sjjk Smnk — AjBjC}J1Dn(8jmSjn —8jn8j}11) = A j B j Cm On dim ojn — A; Bj Cm Dn o/;j bjm,
18 MATHEMATICAL PRELIMINARIES where we have used the e-S identity (2.28). Since Cm8jm = C{ (or A/5,-m = Anu etc.), we have (AxB).(CxD)=: AjBjdOj - AiBjCjDf ^AidBjDj -AjDiBjCj = (A . C)(B • D) - (A • D)(B ■ C). (a) Although the above vector identity is established in an orthonormal coordinate system, it holds in a general coordinate system. That is, the vector identity (a) is invariant. Example 2,4 Let e,- (/ = 1, 2, 3) be a set of orthonormal base vectors, and define new light-handed coordinate base vectors by (ei • e2 = 0) ei = 2eiH-2e2-he3 e2 = ei -e2 V2 ' and e3 = ei x e2 = cj + e2 - 4C3 3v/2 An arbitrary vector A can be represented in either coordinate system A — Aj&i = At*;. The components of the vector in the two different coordinate systems are related by {A} = Lj8]M}, ft,=c,-ey. (2.29) The coefficients 0/y are called the direction cosines. Note that the first subscript of 0/y comes from the barred coordinate system and the second subscript from the unbarred system. For the case at hand we have 0n =ei -ei = -, 021 =C2-ei = —, V2 i 031 = 63 *ei = 012 = ei -e2 = -, 022 = e2 • e2 = - 013 = ei • e3 = V5' 023 = e2 • e3 = 0, 3V2' 032 = e3 • e2 = 3a/T 033 = 63 • e3 = - 3^2* or m 3a/2 2-Jl 2V2 V2 3-3 0 1 1 -4
VECTORS 19 Z,jrJ (x,y,z) = (x\ x2,x>) ^^zA **uS%& y^ X, X' Figure 2.9 Rectangular Cartesian coordinates. When the basis vectors arc constant, that is, with fixed lengths (with the same units) and directions, the basis is called Cartesian. The general Cartesian system is oblique. When the basis vectors are unit and orthogonal (orthonormal), the basis system is called rectangular Cartesian, or simply Cartesian. In much of our study, we shall deal with Cartesian bases. Let us denote an orthonormal Cartesian basis by {ev,ey,ez} or {ei,e2,e3}. The Cartesian coordinates are denoted by (x, y9 z) or (x', x2, Jt3). The familiar rectangular Cartesian coordinate system is shown in Fig. 2.9. We shall always use right-handed coordinate systems. A position vector to an arbitrary point (,v, j>, z) or (x!, x2, x3), measured from the origin, is given by r = xsx + yey + zez = x] ei + A'2e2 + x%, (2.30) or, in summation notation, by r = xj%j. (2.31) < The distance between two infinitesimally removed points is given by dx • dx = (ds)2 = dx}dx* - {dx)2 + (dy)2 + (dz)2. (2.32) 2.2.5 Vector Calculus The basic notions of vector and scalar calculus, especially with regard to physical applications, are closely related to the rate of change of a scalar field with distance. Let us denote a scalar field by <f> — 0(r). In general coordinates we can write 0 = <p(qi,q2i q3)> The coordinate system (q\ q2, q3) is referred to as the unitary system.
20 MATHEMATICAL PRELIMINARIES We now define the unitary basis (ei, e2, e3) as follows: d* #r 9r ,^^x eis^' e2S^' e3S^- (Z33) Hence, an arbitrary vector A is expressed as A = A1 ex + A2e2 + A3e3, (2.34) and a differential distance is denoted by dr = dq[e[ + dq2£2 + dq3e^ = dg'e,-. (2.35) Observe that the A's and dq's have superscripts whereas the unitary basis (ei, e2, £3) has subscripts. The Jg1 are referred to as the contravariant components of the differential vector dr and A' are the contravariant components of vector A. The unitary basis can be described in terms of the rectangular Cartesian basis (ev, ey, ez) = (e\, e2, £3) as follows: V ~a^e'v a^' + a^ — = — e + —e + — fi a#2 a#2 ' a#2 ■ a*?2 a?3 ~ dq^ + Bq**'* dq*** *2 = T~2 = Trzix + Tr&y + ^e„ (2.36a) In the summation convention we have dr dxj e/s—r = —7e/, 1 = 1,2,3. (2.36b) Associated with any arbitrary basis is another basis that can be derived from it. Wc can construct this basis in the following way: Taking the scalar product of the vector A in Eq. (2.34) with the cross product ei x e2, we obtain A • (ei x e2) = A3e3 • (ei x e2) since ej x e2 is perpendicular to both ei and e2. Solving for A3 gives A3=A- eiXC2 X=A-^. (2.37a) e3 • (ei x e2) leie2e3] In similar fashion we can obtain the following expressions for A^A-^, A* = A-£2L!L. (2.37b) [eie2e3] [eie2e3]
VECTORS 21 We thus observe that wc can obtain the components A1, A2, and A3 by talcing the scalar product of the vector A with special vectors, which we denote as follows: cl = «* X e3 f c2 = e3 X et ^ c3 __ e» X e2 (238) [eie2e33' [eie2<J3]' feie2e3l' The set of vectors (e1, e2, e3) is called the dual or reciprocal basis. Notice from the basic definitions that we have the following relations: eiei=Sii\1' l=J . (2.39) } J\0, i^j It is possible, since the dual basis is linearly independent (the reader should verify this), to express a vector A in terms of the dual basis: A = Aie1 + A2e2 + A3e3. (2.40) Notice now that the components associated with the dual basis have subscripts, and A/ are the covariant components of A. By an analogous process to that above we can show that the original basis can be expressed in terms of the dual basis in the following way: e2 x e3 e3 x e1 e'xe2 61 = [ele¥]' e2=fe1cVjt e3 = [e*eV]' ( ° Of course in the evaluation of the cross products we shall always use the right-hand rule. It follows from the above expressions that A^A-e1, A2 = A-e2, A3=A-e3, or A'^A-e'', Ai = A ■ ei, A2 = A • e2, A3 = A • e3, or Ar- = A ■ e,-. (2.42) Returning to the scalar field <j>, the differential change is given by d<P = ^dql + ^dq2 + ?tdq\ (2.43) The differentials dq[,dq2, def are components of dr [see Eq. (2.35)], We would now like to write d(j> in such a way that we elucidate the direction as well as the magnitude of dr. Since e1 • ei = 1, e2 • e2 = I, and e3 • e3 = 1, we can write d<f> = e1 ^ ■ eydq1 + e2^ . e2dcj2 + e3^ ■ e3dq3 dql dql dq5
22 MATHEMATICAL PRELIMINARIES Let us now denote the magnitude of dv by ds = \dr\. Then e = dr/ds is a unit vector in the direction of dr, and we have The derivative (d(j>/ds)z is called the directional derivative of <f>. We see that it is the rate of change of 0 with respect to distance and that it depends on the direction e in which the distance is taken. The vector that is scalar multiplied by e can be obtained immediately whenever the scalar field is given. Because the magnitude of this vector is equal to the maximum value of the directional derivative, it is called the gradient vector and is denoted by grad<£: d0 d</> d<f> grad 0 = e' —T + ez —T + eJ —T dq1 dq2 From this representation it can be seen that dq 3' (2.46) dqv dq2' d<l> are the covariant components of the gradient vector. When the scalar function 0(r) is set equal to a constant, <£(r) = constant, a family of surfaces is generated. A different surface is designated by different values of the constant, and each surface is called a level surface (see Fig. 2.10). If the direction in which the directional derivative is taken lies within a level surface, then d<f>/ds is zero, since 0 is a constant on a level surface, Tn this case the unit vector e is tangent to a level surface. It follows, therefore, that if d<p/ds is zero, then grad <f> must be perpendicular to e and thus perpendicular to a level surface. Thus if any surface is given by <j> (r) = constant, the unit normal to the surface is determined by n = ± grad<£ I grad 0|* (2.47) grad 0 Figure 2.10 Level surfaces.
VECTORS 23 The plus or minus sign appears because the direction of n may point in either direction away from the surface. If the surface is closed, the usual convention is to take ii pointing outward. It is convenient to write the gradient vector as grad^(e'A+e2_£_+e3_L^ (2.48) and interpret grad<£ as some operator operating on <£, that is, grad0 = V0. This operator is denoted by and is called the del operator. The del operator is a vector differential operator, and the "components" d/dq[, S/Bq2, and 3/q3 appear as covariant components. It is important to note that whereas the del operator has some of the properties of a vector, it does not have them all, because it is an operator. For instance V • A is a scalar (called the divergence of A) whereas A • V is a scalar differential operator. Thus the del operator does not commute in this sense. In Cartesian systems we have the simple form r\ q f\ Vse.f+ey- +cz —. (2.50a) ox - ay oz or, in the summation convention, we have V = e/—. (2.50b) The dot product of del operator with a vector is called the divergence of a vector and denoted by V-A = divA. (2.51) Tf we take the divergence of the gradient vector, we have div(grad <p) = V ♦ V</> = (V - V)0 = V2</>. (2.52) The notation V2 = V ♦ V is called the Laplacian operator. In Cartesian systems this reduces to the simple form 2 . p* a2* a24> d24> The Laplacian of a scalar appears frequently in the partial differential equations governing physical phenomena.
s Ihe del opcialor opeiaung on a veclor by means 1A = VxA (2 51.0 —w*|£ aw Example 25 Using ihe index summation Dotation we piove Ihe following identity x(Vx»).7(V v)-V2v «w.»..i«(»i«..) = V(V v)-VJv neumes used as Ihe definition of Ihe Laplacian ol a vecloi that is V2v = grad(divv)-cuiluiilv cyhndncal and sphcncal cooidinale systems die shown in Tablp 2 2 226 Usehil
(•i «2 «3) ai» th» CartMMn un AxB A IBxCI Ax(BxC)-B(A C) C(A B) Cajiettln Component Fram 10 " 12 13 14 U .« 17 18 19 20 V (AxB) B (VxA) A (7x1) V ((/A) = (/v° A + v°(/ A V x (l/A) = Vl/ x A + 1/V x A V(t/A) Vl/A + l/VA Vx(AxB)-A(7 B)-B<V A) + B VA-A VB (VxA)xB = B 1VA (VA)T| 7 (Vl/) = V2l/ 7 (VA) = VJA V X V x A - V(V A) - (V V)A (A V)B A(V B) r„j—<A,B,) IL^ «y/jf(l/Aj)e e,^-<lMje,) •i/WrfljJ-Mi'/fcf «V*«« •'l-fiT* 4>U l2Vc 8'A, *"'"'" 5737"' ^ir« *«£ s=> -face S Lai nl or sin face and n Ihe uml oulwaid normal anu • it V : i differential volume element The following lelauoas are lalcen horn advanceu
26 MATHEMATICAL PRELIMINARIES Table 2.2 The del and Laplace operators in cylindrical and spherical coordinate systems Cylindrical coordinate system (i?, 0, z) x = R cos <f> y = R sin 0 z = z eR = cos 0 eA- 4- sin 0 ey e^ = — sin 0 eA- + cos 0 ey ez =ez 9e/e . A A -— = - sin 0 eA- + cos 0 ey = e^ o<p d*(f> ,, . . „ -—J- — — cos 0 e,T — sin 0 ey = —e# 30 All other derivatives of the base vectors are zero. "'■s 30^ a^ Spherical coordinate system (r, 0,0) * = r sin 0 cos 0 y == r sin 0 sin 0 z = r cos 0 e> = sin 0 cos 0 e,r + sin 0 sin 0 ey + cos 0 ez &0 = cos 0 cos 0 ex + cos 0 sin 0 e-y — sin 0 ez e^ = — sin 0 ev + cos 0 e-y 3e, do = e$ = sin 0 e# = -er = cos 0 e^ 30 3e# a^ —7- = — sin 0 er — cos 0 e# 90 All other derivatives of the base vectors are zero. 3r r 30 rsin0 ^ 30 1 . 3 v2F=^(r2f)+^^HS)+ 1 azF r2sin20 30: ,2 1 Gradient Theorem I grd&(j>dV = <A n<MS. (2.54a)
VECTORS 27 Divergence Theorem. Curl Theorem f diwAdV = Sn-AdS, / curl A dV = A n x \dS. (2.54b) (2.54c) Now let A = grad </> in Eq. (2.54b). Then the divergence theorem gives / div(grad0) dV = f V2(j> dV = (j) h grad <t> dS. (2.55) Jv Jv Js The quantity ii - grad </> is called the normal derivative of <p on the surface S, and is denoted by 90 — = n • grad </) = n • V</>. an (2.56) In a Cartesian system this becomes 8(f) 8<p d</> d4> dn Bx dy Bz (2.57) where nx, ny, and nz are the direction cosines of the unit normal n = nxex + nyey + nzez. (2.58) Example 2.6 This example illustrates the relation between the integral relations (2.54a-c) and the so-called integration by parts, Consider a rectangular region R = {(x, y): 0 < x < a, 0 < y < b] with boundary C, which is the union of line segments C\, C2, C3, and C4 (see Fig, 2,11). ^ * Q nJt = -l,n=-*x| •**x dS=~dx D / C R A ^ P dS=dx )ix= 1, ii =et Figure 2.11 Integration over rectangular regions.
28 MATHEMATICAL PRELIMINARIES Suppose that wc wish to evaluate the integral fR V2</>dxdy, From Eq. (2.55) we have / V20 dxdy = f V • (V0) dxdy = <L^dS. Jr Jr Jc 3« The line integral can be simplified for the region under consideration as follows (note that in two dimensions, the volume integral becomes an area integral): Jc &* id in Jc2 dn J a 9n Jc4 on = r(-BJ)\ **+?(-)\ * Jo \ tyj lv-0 Jo Ydx/ \x=a J a \^y/\y=b Jb \ dx) i-dy) x=0 The same result can be obtained by means of integration by parts: = f(»A\ndy+r(°A)rdx Jo \9x/ LY=0 Jo \dyj \y=o =r[Gi)„r(S)J^ Thus integration by parts is a special case of the gradient or the divergence theorem.
TENSORS 29 2.3 TENSORS 2.3.1 Second-Order Tensors To introduce the concept of a second-order tensor, also called a dyadic, we consider the equilibrium of an element of a continuum acted upon by forces. The surface force acting on a small element of area in a continuous medium depends not only on the magnitude of the area but also upon the orientation of the area. Tt is customary to denote the direction of a plane area by means of a unit vector drawn normal to that plane (see Fig. 2.12a). To fix the direction of the normal, we assign a sense of travel along the contour of the boundary of the plane area in question. The direction of the normal is taken by convention as that in which a right-handed screw advances as it is rotated according to the sense of travel along the boundary curve or contour (see Fig. 2.12b). Let the unit normal vector be given by ii. Then the area can be denoted by S = Sn. If we denote by AF(n) the force on a small area hAS located at the position r (see Fig. 2.13), the stress vector can be defined as follows: t(n) Hm AF(n) AS (2.59) We see that the stress vector is a point function of the unit normal n which denotes the orientation of the surface AS. The component oft that is in the direction of ii is called the normal stress. The component of t that is normal to n is called a shear stress. Because of Newton's third law for action and reaction, we see that t(—n) = —t(fi). \ (a) (b) Figure 2.12 (a) Plane area as a vector, (b) Unit normal vector and sense of travel. Figure 2-13 Force on an area element.
30 MATHEMATICAL PRELIMINARIES Figure 2.14 Tetrahedral element In Cartesian coordinates. At a fixed point r for each given unit vector n there is a stress vector t(n) acting on the plane normal to n. Note that t(ii) is, in general, not in the direction of ii. It is fruitful to establish a relationship between t and n. To do this, we now set up an infinitesimal tetrahedron in Cartesian coordinates as shown in Fig. 2.14. If —ti, —t2, —13, and t denote the stress vectors in the outward directions on the faces of the infinitesimal tetrahedron whose areas are ASi, AS2, AS3, and AS, respectively, we have by Newton's second law for the mass inside the tetrahedron, tAS -1, ASi - t2AS2 -13Aft + pAVf = pAVa, (2.60) where A V is the volume of the tetrahedron, p the density, f the body force per unit mass, and a the acceleration. Since the total vector area of a closed surface is zero [see the gradient theorem; set <j> — 1 in (2.54a)], we have ASii - AStei - AS2e2 - AS3e3 = 0. (2.61) It follows that ASi = (ft ■ ci)AS, AS2 = (ii - e2)AS, AS3 = (n • e3)AS. (2.62) The volume of the element A V can be expressed as Ah AV = —AS, (2.63) where Ah is the perpendicular distance from the origin to the slant face. The result in (2.63) can also be obtained from the divergence theorem (2.54b) by setting A — r. Substitution of Eqs. (2.62) and (2.63) in (2.60) and dividing throughout by AS reduces it to t - (n eOti + (ft • e2)t2 + (ft • e3)t3 + /Aa - f). (2.64) In the limit when the tetrahedron shrinks to a point, Ah -> 0, we are left with t = (ft • ei)t| + (ft • e2)t2 + (ft • <b)t3 = (ft.e,)t/. (2.65)
TENSORS 31 It is now convenient to display the above equation as t = fi.(eiti +e2t2 + e3t3). (2.66) The terms in the parenthesis are to be treated as a dyadic, called stress dyadic or stress tensor a: a =Mi +e2t2 + e3t3. (2.67) The stress tensor is a property of the medium that is independent of the ft. Thus, we have t(fi) = n • a (2.68) and the dependence of t on ft has been explicitly displayed. It is useful to resolve the stress vectors ti,t2, and t3 into their orthogonal components. We have t/ = aa ei + 07262 + <*i3*3 = auej (2.69) for z = 1, 2, 3. Hence, the stress dyadic can be expressed in summation notation as a =e/t/ = oye/ej. (2.70) The component 07/ represents the stress (force perunit area) on an area perpendicular to the ith coordinate and in the yth coordinate direction (see Fig. 2.15). The stress vector t represents the vectorial stress on an area perpendicular to the direction ft. <-> Equation (2.68) is Icnown as the Caiichy stress formula, and a is termed the Caiichy stress tensor. Figure 2.15 Definition of stress components in Cartesian rectangular coordinates.
32 MATHEMATICAL PRELIMINARIES 2.3.2 General Properties of a Dyadic Because of its utilization in physical applications, a dyad is defined as two vectors standing side by side and acting as a unit. A linear combination of dyads is called a dyadic. Let Ai, A2,..., A„ and Bi, B2,..., B,7 be arbitrary vectors. Then we can represent a dyadic as O = A1B1 + A2B2 + • • • + A„B„. (2.71) Here, we limit our discussion to Cartesian tensors. For a Cartesian tensor the basis vectors are constants and thus do not take roles as variables in differentiation and integration. One of the properties of a dyadic is defined by the dot product with a vector, say V: * . V = A, (Bi • V) + A2(B2 ■ V) + - • + A„ (B„ • V), (2.72) V • * = (V • A,)Bi + (V . A2)B2 + ■■• + (¥- A„)B„. The dot operation with a vector produces another vector. In the first case the dyadic acts as a prefactor and in the second case as a postfactor. The two operations in general produce different vectors. The conjugate, or transpose, of a dyadic is defined as the result obtained by the interchange of the two vectors in each of the dyads: T % =B1Ai+B2A2-f- — + B,TA,J. (2.73) It is clear that we have v-* = * -v, (2.74) 2.3.3 Nonion Form of a Dyadic Let each of the vectors in the dyadic be represented in a given basis system. In Cartesian system, we have A,- =Aijtj, (2.75) B, = Bikek. The summations on j and k are implied by the repeated indices.
TENSORS 33 We can display all of the components of a dyadic O by letting the k index run to the right and the j index run downward: * = 01iMl + 012^1 *2 +013M3 + 02ie2ei + <t>22&2&2 + <tee2$3 + 03l%ei + 032£3^2 + 033 $3 ^ (2.76) This form is called the nonion form, Equation (2.63) illustrates that a dyadic in three- dimensional space, or what we shall call a second-order tensor, has nine independent components in general, each component associated with a certain dyad pair. The components are thus said to be ordered. When the ordering is understood, such as suggested by the nonion form (2.76), the explicit writing of the dyads can be suppressed and the dyadic written as an array: m = 011 012 ^13 021 022 023 031 032 033 and ♦» $ = 'sr $2 ie3j 1 > L*]' ei 1 £2 e3J (2.77) This representation is simpler than Eq. (2.76), but it is taken to mean the same. In the general scheme that is thus developed, vectors are called first-order tensors and dyadics are called second-order tensors. Scalars are called zeroth-order tensors. The generalization to third-order tensors thus leads, or is derived from, tricidics, or three vectors standing side by side. It follows that higher-order tensors are developed from polycidics. Example2.7 With reference to a rectangular Cartesian system (*i,*2»*3)» the components of the stress dyadic at a certain point of a continuous medium are given by \<r} = 200 400 300 400 0 0 300 0 -100 psi. We wish to determine the stress vector t at the point and normal to the plane, x\ + 2x2 + 2jc3 —6 = 0, and then compute the normal and tangential components of the stress vector at the point. First we should find the unit normal to the plane on which we are required to find the stress vector. The unit normal is given by [see Eq. (2.47)] n = VP P(x\, X2, X3) = x\ + 2a'2 + 2x3 - 6, n = -(ei +2e2 + 2e3).
34 MATHEMATICAL PRELIMINARIES The components of the stress vector are displayed in an array h ti ti > = 200 400 400 0 300 0 300 0 -100 1 — i 3 1 2 2 1600 400 100 psi or t(n) = -(1600ei + 400e2 + 100e3) psi. The normal component tn of the stress vector t on the plane is given by ., . 2600 . /„ - t(n) • n = -jp psi, and the tangential component is given by (the Pythagorean theorem) ts = ^/itp - tl = 7(256+16+1)9-26x26 psi VT78I = 100- =468.9 psi. 9 A second-order Cartesian tensor O may be represented in barred and unbarred coordinate systems as The unit base vectors in the barred and unbarred systems are related by ^ OX j £ A /N A (2.78) where #/ denote the direction cosines between barred and unbarred systems [see Eq. (2.29)]. Thus the components of a second-order tensor transform according to hi = 4>ijfaPij or [4>] = [£]W>][£]T. (2.79) Tn right-handed orthogonal systems the determinant of the transformation mati'ix is unity, and we have m~l = [/?]T. (2.80) The unit tensor is defined as I = e/e,- (2.81)
TENSORS 35 With the help of the Kronecker delta symbol, this can be written alternatively as I=Sijtiej. (2.82) Clearly the unit tensor is symmetric. The sum of the diagonal terms of a Cartesian tensor is called the trace of the tensor: trace* = 0|7. (2.83) The trace of a tensor is invariant, called the first invariant, and it is denoted by I\; that is, it is invariant with coordinate transformations (0// = 0,7). The first, second, and third invariants of a Cartesian tensor are given by h = 0//, h = ^(0/;0,7 ~ 4>ii4>jj)> h = det[0] = |0|. (2.84) The double-dot product between two dyadics is very useful in many problems. The double-dot product between a dyad (AB) and another (CD) is defined as the scalar: (AB) : (CD) = (B ■ C)(A . D). (2.85) The double-dot product, by this definition, is commutative. The double-dot product between two dyadics is given by * : f - (0/ye/ey) : (irmnem*n) = 0ij^niii(e/ -e/7)(e; • ew) = 4>ijirmn&in8jm Note that the double-dot product of a Cartesian tensor * with the unit tensor I produces its trace /[ = 0//. We note that the gradient of a vector is a second-order tensor: dxj hitj. (2.87) It can be expressed as the sum of VA = - —± + —- ) e,-e, + - I —^ - —M e/ei. (2.88)
36 MATHEMATICAL PRELIMINARIES Analogously to the divergence of a vector, the divergence of a (second-order) Cartesian tensor is defined as div $ = V • $ = e/ 7 ' VVmn &m e/i) OXj = ——(e/ -e,,,)^ " 3* *" Thus the divergence of a second-order tensor is a vector. The integral theorems of vectors presented in Section 2.2.6 are also valid for tensors (second-order and higher): AdV = <p&AdS9 (2.89a) / grad f div Z d V = <f> n • Z dS, (2.89b) / curl WV = inx$^, (2.89c) / n . curl Z dS = <f> dS • $, (2.89d) where CS denotes 'closed surface'. It is important that the order of the operations be observed in the above expressions. 2.3.4 Eigenvectors Associated with Dyadics It is conceptually useful to regard a dyadic as an operator that changes a vector into another vector (by means of the dot product). In this regard it is of interest to inquire whether there are certain vectors that have only their lengths, and not their orientation, changed when operated upon by a given dyadic or tensor. If such vectors exist, they must satisfy the equation I A = XA. 1 (2.90) The vectors A are called characteristic vectors, or eigenvectors, associated with $. The parameter X is called an eigenvalue, and it characterizes the change in length (and possibly sense) of the eigenvector A after it has been operated upon by #.
TENSORS 37 Since A can be expressed as A = I -A, Eq. (2.90) can also be written as (*-A?) A = 0. (2.91) When written in matrix form for Cartesian components, this equation becomes <j)\\ - X 012 013 021 022 - A. 023 031 032 033 - A. A2 A3 = 0. (2.92) Because this is a homogeneous set of equations for A[y A2, A3, a nontrivial solution will not exist unless the determinant of the matrix [$ — A I ] vanishes. The vanishing of this determinant yields a cubic equation for X, called the characteristic equation, the solution of which yields three values of A, that is, three eigenvalues X{, A2, and A3. The character of these eigenvalues depends on the character of the dyadic O. At least one of the eigenvalues must be real. The other two may be real and distinct, real and repeated, or complex conjugates. In the preponderance of practical problems, the dyadic 0 is symmetric, that is, 0 = 0, Of course, $ is always real in our considerations. For example, the moment- of-inertia dyadic is symmetric, and the stress tensor a is usually but not always symmetric. We limit our discussion to symmetric dyadics. The vanishing of the determinant assures that three eigenvectors are not unique to within a multiplicative constant, however, and an infinite number of solutions exist having at least three different orientations. Since only orientation is important, it is thus useful to represent the three eigenvectors by three unit eigenvectors e*, ej, e|, denoting three different orientations, each associated with a particular eigenvalue. Suppose now that Ai and A2 are two distinct eigenvalues and Ai and A2 are their corresponding eigenvectors: <& • A] = AiAj, $ -A2 = A2A2. (2.93a) (2.93b) Scalar product of the first equation by A2 and the second by Ai, and then subtraction, yields A2 • * - Ai - Ai • * • A2 = (Xi - A2)A) • A2. (2.94) Since # is symmetric, one can establish that the left-hand side of this equation vanishes. Thus 0=(Xi-X2)Ai.A2. (2.95)
38 MATHEMATICAL PRELIMINARIES Now suppose that A i and A2 are complex conjugates such that X\ — A2 — 2* A1/, where / = V^T and Xa is the imaginary part of X\. Then A] ♦ A2 is always positive since Ai and A2 are complex conjugate vectors associated with X\ and X2. it then follows from Eq, (2.95) that Aj,- =0 and hence that the three eigenvalues associated with a symmetric dyadic are all real. Now assume that Ai and X2 are real and distinct such that Ai — A2 is not zero. It then follows from Eq. (2,95) that Ai - A2 = 0. Thus the eigenvectors associated with distinct eigenvalues of a symmetric dyadic are orthogonal. If the three eigenvalues are all distinct, then the three eigenvectors are mutually orthogonal. If X\ and X2 are distinct, but X3 is repeated, say A3 = A2, then A3 must also be perpendicular to Aj as deducted by an argument similar to that for A2 stemming from Eq, (2,95), Neither A2 nor A3 is preferred, and they are both arbitrary, except insofar as they are both perpendicular to A1. It is useful, however, to select A3 such that it is perpendicular to both Ai and A2. We do this by choosing A3 = Ai x A2 and thus establishing a mutually orthogonal set of eigenvectors. This sort of behavior arises when there is an axis of symmetry present in a problem. In a Cartesian system the characteristic equation associated with a dyadic can be expressed in the form A3 - /1A2 - /2A - h = 0, (2.96) where I\, I2, and I3 are the invariants associated with the matrix of <t>. The invariants can also be expressed in terms of the eigenvalues, h = Ai +A2 + A3, h = -(XJA2 + A2A3 +A3A,), h = A1A2A3. (2.97) Finding the roots of the cubic Eq. (2.96) is not always easy. However, when the matrix under consideration is of the form [011 0 0 1 0 022 023 , L 0 032 033 J one of the roots is Ai = 0n, and the remaining two roots can be found from the quadratic equation (022 - A)(033 - A) - 023032 = 0, That is, ^2,3 = r ± -y (022 + 033)2 ~ 4(022033 ~ 023032)- (2.98) In cases where one of the roots is not obvious, an alternative procedure given below proves to be useful.
TENSORS 39 In the alternative method we seek the eigenvalues of the so-called deviatoric tensor associated with 4>: tij^frj-^kkSij. (2,99) Note that 4>ii = 4>li ~ <Pkk = 0. (2.100) That is, the first invariant l[ of the deviatoric tensor is zero. As a result the characteristic equation associated with the deviatoric tensor is of the form {XfY - If2Xf -73=0, (2.101) where Xf is the eigenvalue, and 1^ and /^ are the second and third invariants of the deviatoric tensor. The eigenvalues associated with [4>] itself can be computed from A = A' + Ife, Ii = fyu4>v> /3 = l«'|. (2.102) The cubic equation in (2.101) is of a special form that allows a direct computation of its roots. Equation (2.101) can be solved explicitly by introducing the transformation l' = <5'0 1/2 COS Of, which transforms (2.101) into /l \3/2 21-/2) [4cos3a-3coso| = 1^. The expression in square brackets is equal to cos 3a. Hence 3/2 cos 3a -m (2.103) (2.104) (2.105) If a\ is the angle satisfying 0 < 3a 1 < n whose cosine is given by Eq. (2.105), then 3a 1, 3ai + 2jt, and 3a 1 — 2n all have the same cosine, and furnish three independent roots of Eq. (2.103), where a\ = cos -l 1/2 cos a/, i = l,2,3, (2.106a) 2tt <*2 = <*1 + "3"' <*3 a\ - 2tt 3 ' (2.106b) Finally we can compute A; from Eq. (2.102).
40 MATHEMATICAL PRELIMINARIES Example 2.8 We wish to determine the eigenvalues and eigenvectors of the matrix [01- r2 1 \ o 1 4 1 Ol 1 2J The characteristic equation is obtained by setting det ($,•/ — A 5//) to zero: = (2 - A)[(4 - A.)(2 - A.) - 1] - 1 • (2 - A.) = 0, 2-A. 1 0 1 4-A. 1 0 1 2-A or (2-A)[(4-A-)(2-A.)-2] = 0. Hence At = 3 + V5 = 4.7321, A.2 =3-V3= 1.2679, A.3 = 2. Alternatively, r*'i = -f i o r 4-i i 2-5 , l2 = M$} ~ <l>ii<l>'jj) = -tijtij '(-l),+(-I),+G),««' 10 T 52 27' /^=dct(^) = -. From Eq. (2.106b), 1 «i = cos ,-1 '52 /^y/2' 54 V10 J 11.565°, a2 = 131.565°, a3 = -108.435°, and from Eq. (2.103), X\ = 2.065384, A(, = -1.3987, A'3 = -0.66667. Finally, using Eq. (2.102), we obtain the eigenvalues A. 1 = 4.7321, X2 = 1.2679, A3 = 2.00.
EXERCISES 41 The eigenvector corresponding to A3 = 2, for example, is calculated as follows. From ((pjj - Xi$jj)Aj = 0 we have "2-2 J 0 1 4-2 1 0 1 2-2 Ax" A2 ,A3. ■ = 1 This gives A2 = 0, Ai=-A3. Using A] + A\ + A\ = 1, we obtain A3 = ±-=(1,0,-1), V2 for A3 = 2. Similarly, the eigenvectors corresponding to X 1,2 = 3 ± \/3 are given by A, = ±(3"j2 (l. (l + V3) , l) , for X! = 3 + V3, A2 = ±^±^(l,(l-V3),l), forl2 = 3-V3. EXERCISES 2.1 Find the equation of a line (or a set of lines) passing through the terminal point of a vector A and in the direction of vector B. 2.2 Find the equation of a plane connecting the terminal points of vectors A, B, and C. Assume that all three vectors arc referred to a common origin. 2.3 Prove the following vector identity without the use of a coordinate system (see Example 2.2): A x (B x Q = (A ■ C)B - (A • B)C. 2.4 If e is any unit vector and A an arbitrary vector, show that A = (A - e)e + e x (A x e). This identity shows that a vector can be resolved into a component parallel to and one perpendicular to an arbitrary direction e. 2.5 Verify the following identities: (a) 5,7 =3. (c) SijSjk = fyk- (e) £ijk£ijk = 6. (b) 8jj8jj =<5//. (d) EuijkGnjk = 2<5„ (0 AiA]eijk={).
42 MATHEMATICAL PRELIMINARIES 2.6 Using the index notation, prove the identity (A x B) • (B x C) x (C x A) = (A • (B x C))2. 2.7 Prove the following vector identity in an orthonormal system using index- summation notation: (A x B) x (C x D) = [A • (C x D)]B - [B - (C x D]A. 2.8 Determine whether the following set of vectors is linearly independent: A = ei-t-e2j B = e2 + e4, C = e3-he4, D = 6) + e2 + e3+e4. Here e/ are orthonormal unit base vectors in a four-dimensional space. 2.9 Determine whether the following set of vectors is linearly independent: A = 2ej — $2 + 63, B = e2 — £3, C = — ej + §2. Here e,- are orthonormal unit base vectors in Sfi3. 2.10 Determine which of the following sets of vectors span tfi3: (a) A = e, + 3e2 - 83, B = -4ei + 3e2 - 5e3, C = 2e} + e2 + e3. (b) A = ei-f-e2, B = ei + e2 - 2e3, C = 61-63. Here e,- are orthonormal unit base vectors in *R3. 2.11 Consider two rectangular Cartesian coordinate systems that are rotated with respect to each other and have a common origin. Let one system be denoted as a barred system, so that a position vector can be written in each of the systems as r = jc/e/, = xjeJ> where {£/} and {!/} are the respective orthonormal Cartesian bases in the unbarred and barred systems. By requiring that the position vector r be invariant under a rotation of the coordinate systems, deduce that the transformation between the coordinates is given by x\ — a\\x\ -f a\2X2 + 013*3 *2 = 021*1 + #22*2 + 023*3 / *3 = 031*1 + 032*2 + ^33*3 or, more compactly, Xj =ajjxj, ij = 1,2,3,
EXERCISES 43 where the terms au can be identified as the direction cosines ciij = if • ey = cos(e/, e/). Deduce further that the basis vectors obey the same transformation e/ = aj e j, and that the following orthogonality conditions hold: 2X1 Determine the transformation matrix relating the orthonormal basis vectors (ei, e2, S3) and (e'p e^ ^)» when e^ are given by (a) e\ is along the vector ei — 62 -I- $3 and e'2 is perpendicular to the plane 2x\ +3*2 + *3 -5 = 0. (b) e\ is along the line segment connecting point (1 > — 1, 3) to (2, -2, 4) and e£ = (-ei +e2+2e3)/V6. (c) 63 = S3, and the angle between x\ -axis and ;ei-axis is 30°. 2.13 The angles between the barred and unbarred coordinate lines are given by ei e2 h ei 60° 150° 90° h 30° 60° 90° £3 90° 90° 0° Determine the direction cosines of the transformation. 2.14 The angles between the barred and unbarred coordinate lines are given by xx X2 *3 *1 45° 60° 120° X2 90° 45° 45° *3 45° 120° 60° Determine the transformation matrix. 2.15 In a rectangular Cartesian coordinate system, find the length and direction cosines of a vector A that extends from the point (1, — 1, 3) to the midpoint of the line segment from the origin to the point (6, —6, 4). 2.16 The vectors A and B are defined as follows: A=3i-4k, B = 21 - 2j + k, where (i, j, k) is an orthonormal basis.
44 MATHEMATICAL PRELIMINARIES (a) Find the orthogonal projection of A in the direction of B. (b) Find the angle between the positive directions of the vectors. 2.17 Prove the following identities (see Eq. (2.13) for the definition of [ABC]): rfA^Al r d\d3\l dt dt2 J ~ [ dt dP J ' 2.18 Let r denote a position vector r = x\ e/ (r2 = jc,-jc,- ) and A an arbitrary constant vector. Show that: (a) grad(r) = }?. (b) grad(r") = nrn~2 r. (c) V2(r,!) = n(n + l)r""2. (d) grad(r • A) = A. (c) div(r x A) = 0. (f) curl(r x A) = -2A. (g) div(rA) = |(r • A). (h) curl(rA) = i(r x A). 2.19 Let A and B be continuous vector functions of the position vector r with continuous first derivatives, and let F and G be continuous scalar functions of r with continuous first and second derivatives. Show that: (a) curl(grad F) — 0. (b) div(curl A) = 0. (c) div(grad Fx grad G) = 0. (d) grad(FG) = F grad G + G grad F, (e) div(FA) = A- grad F + FdtvA. (0 curl(FA) = F curl A - A x grad F. (g) grad(A B) = A grad B + B grad A + A curl B + B curl A. (h) div(A x B) = curl A • B - curl B . A. (i) curl(AxB) = B-VA-A. VB + A divB-BdivA. (j) (V x A) x A = A • VA - VA • A. (k) V2(FG) = F V2G + 2VF VG + G V2F. 2.20 Show that the vector area of a closed surface is zero, that is, i s 2.21 Show that the volume enclosed by a surface S is ' volume = - (p grad(r2) • n dS,
EXERCISES 45 or volume = - (b r ■ n dS. lis 2.22 Let 0(r) be a scalar field. Show that where 'd^/'dn = n • grad<£ is the derivative of 4> in the outward direction normal to the surface. 2.23 In the divergence theorem (2.54b), set A = 0 grad ^ and A = \jf grad <f> successively and obtain the integral forms (a) / [0V2f + V0 . V^ldG = <& cj)^- rfr, (b) J [^vV - ^v2tf>] da = <£ U^ - v^l rfr, (c) £ [*vV - v20vV] *o = £ [*^(vV) - vV^l dr, where £2 denotes a (2D or 3D) region with boundary I\ The first two identities are sometimes called Green's first and second theorems. 2.24 Determine the rotation transformation matrix such that the new base vector e i is along ei — £2 + e3, and £2 is along the normal to the plane 2x\ + 3*2 4- *3 = 5. If T is the dyadic whose components in the unbarred system are given by 7ii = 1, Tn = 0, TX3 = -1, 722 = 3, T23 = -2, and T33 = 0, find the components in the barred coordinates. 2.25 Show that the characteristic equation for a second-order tensor 07/ can be expressed as where A3-/iA.2-/2A.-/3 =0, h = -(IVijVjkVki — 3<yjj0jj<Jkk + OiiVjjVkk) = dct(or,y) 0 are the three invariants of the tensor.
46 MATHEMATICAL PRELIMINARIES 2.26 Find the eigenvalues and eigenvectors of the following matrices: (a) (c) (e) r4 -4 0" -4 0 0 L° ° 3. fl 0 0 " 0 3 -1 [0 -I 3 _ r3 5 8" (5 10 8 0 2 (b) (d) (f) 2 -V3 0 -V3 4 0 0 0 4 "2 -11] -10 1 1 1 2 J "1 -1 Ol -1 2 -1 0 -1 2 2.27 Evaluate the three invariants of the matrices in Exercise 2.26 and check them against the invariants obtained by using the eigenvalues. 2.28 The components of a stress dyadic at a point, referred to the („yi , X2, x$) system, are (in ksi = 1000 psi): (i) 12 9 9 -12 0 0 (HO 0 0 6_ ' 1 J. (ii) -3 3 1 2 —s/2 9 0 12 V2 1 ~v/2 4 0 12' -25 0 0 16 Find the following: (a) The stress vector acting on a plane perpendicular to the vector 2ei — 2e2+e3 passing through the point. (b) The magnitude of the stress vector and the angle between the stress vector and the normal to the plane. (c) The magnitudes of the normal and tangential components of the stress vector. 2.29 Verify that I .* = *• I =*. 2.30 If A is an arbitrary vector and <J> is an arbitrary dyadic, yerify that (a) (T x A) • Z = A x Z. (b) (A x T) • Z = A x 5. (c) (vt> x A)1 = -A x <D
REFERENCES 47 REFERENCES 1. Aris, R., Vectors, Tensors, and the Basic Equations in Fluid Mechanics., Prentice-HalL Englewood Cliffs, NJ (1962). 2. Bedford, F. W, and Dwivedi, T. D., Vector Calculus, McGraw-Hill, New York (1970). 3. Bellman, T. E., Introduction to Matrix Analysis, 2nd edition, McGraw-Hill, New York (1970). 4. Borisenko, A. I., and Tarapo, I. E., Vector and Tensor Analysis with Applications, R. A. Silverman (translator and editor), Prentice-Hall, Englewood Cliffs, NJ (1968). 5. Bourne, D. E., and Kendall, P. C, Vector Analysis and Cartesian Tensors, 2nd edition, Academic Press, New York (1977). 6. Bowen, R. M., and Wang, C. C, Introduction to Vectors and Tensors, Plenum Press, New York (1976). 7. Campbell, H. G, An Introduction to Matrices, Vectors and Linear Programming, Prentice- Hall, Englewood Cliffs, NJ (1977). 8. Chambers, L. G, A Course in Vector Analysis, Chapman and Hall, London (1969). 9. Chisbolm, J. S. R. Vectors in Three-Dimensional Space, Cambridge University Press, Cambridge (1978). 10. Chorlton, R, Vector and Tensor Methods, Halsted Press, New York (1964). 11. Coburn, N., Vector and Tensor Analysis, Macmillian, New York (1964). 12. Eisele, J. A., Applied Matrix and Tensor Analysis, Wiley-Intersciencc, New York (1970). 13. Gantmachcr, E R., The Theory of Matrices, Chelsea, New York (1959). 14. Graham, A., Matrix Theoiy and Applications for Engineers and Mathematicians, Halsted Press, New York (1979). 15. Haskell, R. E., Introduction to Vectors and Cartesian Tensors, Prentice-Hall, Englewood Cliffs, NJ (1972). 16. Hauge, B., An Introduction to Vector Analysis for Physicists and Engineers, 6th edition, revised by D. Martin, Methuen, London (1970). 17. Jeffreys, H., Cartesian Tensors, Cambridge University Press, London (1965). 18. Karamchcti, K., Vector Analysis and Cartesian Tensors, Holden-Day, San Francisco (1967). 19. Lanzcos, C, "Tensor calculus," in Handbook of Physics, E. U. Conder, and H. Odishaw (eds.), Ch. 10, pp. Mil to 1-122. McGraw-Hill, New York (1958). 20. Lichncrowicz, A., Linear Algebra and Analysis, Holden-Day, San Francisco (1967). 21. Moon, P. H., and D. E. Spencer, Vectors, Van Nostrand, Princeton, NJ (1965). 22. Phillips, H. B., Vector Analysis, Wiley, New York (1933). 23. Rcddy, J. N., and Rasmussen, M. L., Advanced Engineering Analysis, John Wiley, New York (1982); reprinted by Krieger, Melbourne, FL (1991). 24. Sokolnikoff, I. S., Tensor Analysis, John Wiley, New York (1933). 25. Wills, A. P., Vector Analysis with an Introduction to Tensor Analysis, Dover, New York (1953). 26. Wrede, R. C, Introduction to Vector and Tensor Analysis, Dover, New York (1972). 27. Young, E. C, Vector and Tensor Analysis, Marcel Dekker, New York (1978).
3 REVIEW OF EQUATIONS OF SOLID MECHANICS 3.1 INTRODUCTION 3.1.1 Classification of Equations The objective of this chapter is to record the governing equations of a solid continuum in a form suitable for use later in this book. The word continuum needs to be explained first. Microscopically, a medium occupied by a matter, solid or fluid, is made of discrete particles of protons, neutrons, and electrons. Macroscopicaliy, the medium is assumed to contain no gaps or voids between material points of the medium so that it can be divided indefinitely into smaller and smaller parts without encountering a void. This concept allows us to shrink an arbitrarily small region to a point so that all spatial derivatives of various quantities associated with the medium can be defined. For example, if mass density p(x> t) of the medium is defined to be the mass per unit volume, we assume that the limit ><*•'> %teo!£ (3-1} exists (i.e., p is finite). Here Am denotes the mass of the matter occupying an infinitesimal volume AV, xis the position vector, and t denotes time. Of course, the continuum assumption may be violated when one considers matter at nanometer (10~9 m) or atomic scale, and the continuum equations to be derived in this chapter may not be applicable to problems at such scales. t The governing equations of a continuum are derived using the following conservation principles (or laws of physics): 1. Principle of conservation of mass. 2. Principle of conservation of linear momentum. 48
INTRODUCTION 49 3. Principle of conservation of angular momentum. 4. Principle of conservation of energy. The above principles do not explicitly account for geometric changes or mechanical response of the continuum. Without these, the equations derived from the conservation principles are insufficient to determine the total response of the continuum. Therefore, we must consider the following additional two sets of equations: 5. Kinematics (strain-displacement equations). 6. Constitutive equations (e.g., stress-strain relations). Kinematics is a study of the geometry of motion and deformation without con- sideration of the forces causing the motion. The constitutive equations describe the mechanical behavior of the continuum and relate the dependent variables introduced in the kinetic description (i.e., conservation of momenta and energy) to those in the kinematic description. Since the resulting equations involve spatial (i.e., position) coordinates and time, appropriate boundary and initial conditions are needed to determine the solution. For a detailed derivation of the equations, the reader may consult the books on continuum mechanics [1-31, elasticity [4-6], and fluid mechanics [7,8]. 3.1.2 Descriptions of Motion There are two alternative descriptions used to study the motion of a continuum. In the first, one considers the motion of all matter passing through a fixed spatial location, as shown in Fig. 3.1a. Here one is interested in various properties (e.g., velocity, pressure, temperature, density, and so on) of the matter that instantly occupies the fixed spatial location. This description is called the spatial description or the Eulerian description. In the second, one focuses attention on a fixed set of material particles, irrespective of their spatial locations (see Fig. 3. lb). The relative displacements of these particles and the stress caused by external forces and temperature arc of interest in this case. This description is known as the material description or the Lagrangian description. The Eulerian description is most commonly used to study fluid flows and coupled heat transfer and fluid flow, while the Lagrangian description is generally used to study heat transfer, stress, and deformation of solid bodies. (a) (b) Figure 3.1 (a) Fixed region in space and mass moving through it. (b) Motion of a fixed collection of material particles.
50 REVIEW OF EQUATIONS OF SOLID MECHANICS In order to understand the difference between the material and spatial descriptions, consider a continuum and identify a region of the continuum for our study. Let X denote the position of an arbitrarily fixed point in the region (in both descriptions) at time t = 0, and let us label the material particle that occupies position X with X (a name given to the particle). For time / > 0, the point X in a spatial description remains the same, but denoted by x. Although the current position x and initial position X are the same in the spatial description, the particles occupying the position at these two times are not the same (i.e., particle X no longer occupies the position x at time t > 0), unless there is no motion. On the other hand, in a material description, where attention is focused on a material particle, the particle X occupying the position X at time / = 0 moves to a new position x at time t > 0. To gain further understanding, consider representing a scalar quantity such as the mass density p in the two descriptions. In a material description, the functional variation of p is described with respect to the coordinate X occupied by the material particle X in the region as P = p(X,/). (3.2) In spatial description, p is described with respect to the position in space, x, currently occupied by a material particle in the continuum at time t as P = P(x, r). (3.3) In Eq. (3.2) a change in time t implies that the same material particle X has a different density. The particle's current position x has been expressed in terms of its position X at time t = 0. Thus, in material description, attention is focused on the material particle. In Eq. (3.3), however, a change in time t implies that a different density is observed at the same spatial location x, now probably occupied by a different material particle. Therefore, in spatial description, attention is focused on spatial location. 3.2 CONSERVATION OF LINEAR AND ANGULAR MOMENTA 3.2.1 Equations of Motion The principle of conservation of linear momentum, or Newton's second law of motion, applied to a set of particles (or rigid body), can be stated as the time rate of change of (linear) momentum of a collection of particles equals the net force exerted on the collection. Written in vector form, the principle implies — (m\) = F, (3.4) at t where m is the total mass, v the velocity, and F the resultant force on the collection particles. For constant mass, Eq. (3.4) becomes d\ F = 772-— = ma, (3.5) at
CONSERVATION OF LINEAR AND ANGULAR MOMENTA 51 which is the familiar form of Newton's second law (i.e., force = mass x acceleration). To derive the equation of motion applied to a control volume (i.e., either a fixed region in space through which material flows or a fixed material body that is in motion), we must identify the forces acting on it. Forces acting on a control volume can be classified as internal and external. The internal forces resist the tendency of one part of the region/body to be separated from another part. The internal force per unit area is termed stress, as defined in Section 2.3. The external forces are those transmitted by the body. The external forces can be further classified as body (or volume) forces and surface forces. Body forces act on the distribution of mass inside the body. Examples of body forces are provided by the gravitational and electromagnetic forces. Body forces are usually measured per unit mass or unit volume of the body. Let f denote the body force per unit mass. Consider an elemental volume dV inside V. The body force of the elemental volume is equal to pdVf, Hence, the total body force of the control volume is / pfdV. (3.6) Jv Surface forces are contact forces acting on the boundary surface of the body. Examples of surface forces are provided by applied forces on the surface of the body. Surface forces are reckoned per unit area. Let t denote the surface force per unit area (or surface stress vector). The surface force on an elemental surface dS of the control volume is tdS. The total surface force acting on the closed surface of the control volume is £> t dS. (3.7) Since the stress vector t on the surface is related to the (internal) stress tensor a by Cauchy's formula [see Eq. (2.68)], t = n-cf\ (3.8) where n denotes the unit normal to the surface, we can express the surface force as n ■ o dS. *. i< fs Using the divergence theorem (2.89b), we can write <&A-adS= f V-ffdV. (3.9) The principle of conservation of linear momentum as applied to a solid continuum yields the result V.Z + pf = p^, (3.10)
52 REViEW OF EQUATIONS OF SOLID MECHANICS U-3 dxf> dxn TX3 *22^ ff»+Tg**- <Tll+^ii<fa, a21+3^1^-2 axz OT2 Figure 3.2 Stresses on a parallelepiped element. where u is the displacement vector. In the Cartesian rectangular system, we have (3.11) >°J<+pfi=p*« (/ = l,2,3). dXj dt2 Equations of motion (3.11) can be derived by directly applying Newton's second law of motion to a volume element. Consider the stresses and body forces on an infinitesimal parallelepiped element of a material body. Figure 3.2 shows the stresses acting on the various faces of the infinitesimal parallelepiped with dimensions dx[, dx2, and dx$ along coordinate lines (x\, X2, x$). The sum of all forces in the ^indirection is given by ( &u + ~—dx\ ) dx2dx3 - <T[\ dx2dx$ + I cr2i + -—dxj ) dx\dxs ( d&3\ \ — a-i\ dx\dx-\ -f ( cT3i H dx-\ I dx\dxi — a-\\ dx\dx2 + pf) dx\dx2dx3 \ dX3 ' ) (da\\ dcr2\ daw \ = h -r— + h PJ\ dxldx2dx?>. \ dx\ 0x2 0x3 } By Newton's second law of motion, the sum of the forces is equal to the product of mass and acceleration in the x\ -direction, (pdx\dx2d.X3)—y> 1 where p is the density. Thus, upon dividing throughout by dx\dx2dx3, we obtain dan 0(72i 3cr3i 3x[ 3x2 0x3 d2u\
CONSERVATION OF LINEAR AND ANGULAR MOMENTA 53 Similarly, the application of Newton's second law in the a*2- and x$-directions gives respectively dan %gii 9(J32 „ 3 ^2 ^— + ~H— + ^— + ^2 = ^1^2~* dX[ 3*2 3x3 3?z 3ai3 3(723 , 3(733 , ,. #2"3 9 a-i 3a'2 3a"3 3fz or, in index notation, 3xy 3rz which is the same as thai in Eq. (3.11). For static equilibrium, we set the time derivative terms to zero and obtain the equilibrium equations ?PL+pfi=Q 0 = 1,2,3). (3.13) dxj 3.2.2 Symmetry of Stress Tensor The principle of conservation of angular momentum states that the time rate of change of the total moment of momentum for a continuum is equal to the vector sum of the moments of external forces acting on the continuum. In the absence of body couples (i.e., volume-dependent couples), the principle leads to the symmetry of the stress tensor. That is, the matrix of the stress components is symmetric: ^23 = ^32, 0*31 = Ci3> 0\i = cr2i. Thus, there are only six stress components that are independent. The symmetry of the stress tensor can also be established using Newton's second law for moments. Consider the moment of all forces acting on the parallelepiped about the A3-axis (sec Fig. 3.2). Using the right-handed screw rule for positive moment, we obtain ( o-\2 + -7T^dxi ) dx2clx^ ~Y + ^12 dx2clx^ ~Y - ( 0*21 + -—dx2 J dx\dx2 -r tai dxidxi) — = 0. Dividing throughout by £ dx\dxidx^ and talcing the limit dx\ -> 0 and dx-2 -+ 0, we obtain 0"I2 - 0*21 = 0.
54 REVIEW OF EQUATIONS OF SOLID MECHANICS Similar considerations of moments about the jcj-axis and JC2-axis give, respectively, the relations 3.3 KINEMATICS OF DEFORMATION 3.3.1 Strain Tensor Kinematics is a study of the geometry of motion and deformation without consideration of the forces causing them. Under the action of applied loads, a material body gets displaced and deformed or strained. The word deformation refers to changes in the geometry of the body. A body is said to undergo rigid body motion if the distance between any two arbitrary points and the angle between any two infinitesimal line segments in the body remains unchanged. The rigid-body motion does not alter the shape of the body. Thus, a measure of the deformation is provided by the change in the distance between points and angle between line segments in the body. Here we develop a measure of straining of the body and introduce strain tensors using the material description. Consider a fixed mass occupying a region with volume V and closed surface S. For simplicity, we also denote the region with V. A typical material particle in V is labeled as X, and its location with respect to a rectangular Cartesian coordinate system is denoted as X. Under the action of applied forces (and subjected to geometric constraints), the body undergoes deformation and the particle X moves to a new location x. Since all material parlicles of the body move, possibly by different amounts, the body occupies a new region, as shown in Fig. 3.3. To determine if the motion involved is a deformation (or straining), we must compute the distance between two neighboring points after deformation. Let P: (X\,X2, X3) and Q: (X{ + dXv, X2 + dX2) X3 + dX3) denote two arbitrary material points in the body at time t = 0 (our reference configuration). Under the action of externally applied forces, _the body deforms, and the points P and Q move to new places: P: (x\9X2, x$) and Q: (x\ ~f dx\> x2 + dx2, x3 + dx3), respectively. The motion of a particle X occupying position X in the undeformed body (i.e., the reference configuration) to point x in the deformed body can be expressed by the displacement vector u = x —X or iti =xj — Xj. (3.14) in the material description, the current position x of a material point is expressed in terms of its position at time t = 0: i x = x(X, 0, Xi = xtiXi, X2, X3, 0 (/ = 1, 2, 3). (3.15) By assumption, a continuous medium cannot have gaps or overlaps. Therefore, a one-to-one correspondence exists between points in the undeformed body and points in the deformed body. Consequently, a unique inverse to (3.15) exists.
dX-dX=(dS)2 KINEMATICS OF DEFORMATION 55 x2~X2 dx = (ds)2 i-Xt Figure 3.3 Deformation of a material body. Next, we wish to determine the change in the distance between points P and Q in the deformed body and compare it with the distance between points P and Q in the undeformcd body. The distances between points P and Q and points P and Q are given, respectively, by (dS)2 = dX-dX = dXidXi, (ds)2 = dx ♦ dx = dx\dxu (3,16) where summation on repeated subscripts is implied. If dS ^ ds9 then we know that the body has strained. Since the length of a vector is obtained by computing the square of its length and then taking the square root of it, it is convenient to consider the square of the distance between points. We have (ds)2 - (dS)2 = dx • dx - dX . dX = dxmdxm - dXidXj - (SN (w?*') -"•"• dxm dx„ dXj dx,* = 2£ijdXidXh ■dXjdX j — SjjdXjdX j (3.17) where sy are the components of the Green strain tensor at point P, 7 = EijEiitj, (3.18a)
56 REVIEW OF EQUATIONS OF SOLID MECHANICS with components SiJ~ 2\dXiSXj iJJ' (3.18b) HereE/ denote the unit basis vectors in the rectangular Cartesian system (X[, X2, ^3). The fact that £// are components of a second-order tensor will be established shortly. Note that the Green strain tensor components are symmetric by definition. Also, the change in the length of a line element is zero if and only if £/y = 0. The strain components can be expressed in terms of the displacement components at point P using Eq. (3,14): — um + X„ fix,, dun dXi dXj + 5/1!/. (3.19) We have Sij = ;(£+<-)(£+*-H _ -J- / duj_ duj_ du„, du„, \ ~ 2 \dXj dXj dXjdXj)' In expanded notation, typical strain components are given by (3.20) du\ 1 /9»A2 (du2\2 /3«3' \dXi) \dxj \8Xi, i /du\ dii2 dil[ dll\ Sll2 dui 3«3 flW3 \ £n ^ 2 \dx~2 + Jx~i + Jx[ ax^ + Wv~dX2 + 'dx~\^x~2j (3.21) When the derivatives of the displacement components are small compared to unity, their products can be neglected: In this case, the strain components £/y become the infinitesimal strain components e\$: en « *,-/ = - ( —- + —L = - ( —L + — . (3.23) lJ 3 2 \8Xj dXf J 2 \3xj dxi J V J Example 3.1 Consider a rectangular block of dimensions a x b x h, where h is very small compared to a and /?. Suppose that the block is deformed into the diamond shape shown in Fig. 3.4. By inspection, the geomeUy of the deformed body can be described as follows: let (X[, X2, X$) denote the coordinates of a material point in the undeformed configuration. In the deformed configuration, the material point has
KINEMATICS OF DEFORMATION 57 >*, Figure 3.4 Undeformed and deformed rectangular block. the coordinates xx = Xx + -£*2, b a A'3 & X3 (because h <£ a or b). Thus, the displacement components of the material point are v e°v «3 = *3 - X3 = 0. The strains are given by 811 = 2(7) ' $13 = £23 = £33 = 0. £12= — + —, 2fr 2a £22 J_/fo\2 2Ui ' The same results can be obtained using (lie elementary mechanics of materials approach, where the strains are defined to be the ratio of the difference between the final length and original length to the original jength. For example, a line element A B in the undeformed body moves to position A B. Then the strain in the line AB is given by £11 = £ab = AB-AB
58 REVIEW OF EQUATIONS OF SOLID MECHANICS When eo/h and e$/a are small compared to unity, their squares can be neglected, and the infinitesimal strain components are given by eo en^O, e22^0, en = ttt (a + b). Lab Now consider a line element oriented at an_angle from the Jti-axis. For example, the diagonal line AD deforms and becomes A D in the deformed body. The normal strain in A D is given by , _ V(a + e0)2 + (b + e0)2 - Va2 + b2 Va^H?2 el+epja + b) a2 + b2 Since ¥ is a second-order tensor, we can use the transformation relation in Eq. (2.79) and obtain the same result. We have *ii =PuPtj£ij = fiupusn + Pnfii2£22 + PuP\2&\2 + /3i2/3ii£2i. where Pij = *i -ey. In particular, we have fin = cos#, fi\2 = sin0, fin — cos0, fti = — sin0 and 0 = tan_1(/?/a). Hence, s\, = cos2 0e j i + sin2 0^22 + 2 sin 0 cos Oen = £# - Substituting for cos 0, sin 0, and £// into ej., we obtain ' _ a2 1 /go\2 ^2 l (eo\2 ? ab epfc + fr) eil"a2+i22U/ a2 + b22\b) * a2 + b2 lab __ gp + en (a + ft) a2 + ft2 ' 3.3.2 Strain Compatibility Equations ' When a sufficiently differentiable displacement field is given, the computation of the strains is straightforward: one can make use of the equations in (3.20) or (3.23) to compute the strains. However, when the strain components are given, the determination of the displacements is not always possible because there are six strain 1 +2 N el + eo(a + b) a2 + b2
KINEMATICS OF DEFORMATION 59 components related to three displacement components. Stated in other words, there are six differential equations involving three unknowns. Thus, the six equations should be compatible with one another in the sense that any three equations should give the same displacement field. To understand the situation further, let us consider the two-dimensional case. The strain-displacement relations of the linear theory are given by dlt\ ox\ -r1 = e22, (3.24) 0X2 ^~ + T" = 2e[2' 0X2 OX\ If €ij are given as functions of x\ and X2> they cannot be arbitrary: e\\, £22, and e\2 should have a relationship such that the three equations are compatible. This relation can be derived as follows: differentiate the third equation with respect to A'i and X2 to obtain d3lt[ d3«2 ^ d2<?i2 + 9 =2 ' . (3.25) 3jci 9x\ dx^dx2 9x\ 9x2' The third derivatives of u\ and 112 needed in Eq. (3.25) can be computed from the first two equations in Eq. (3.24): du\ d2e)\ d2U2 ox2e22 Bx\dxl 9x\ ' dx\dx2 dx\ ' Substituting these relations into Eq. (3,25), we obtain a relationship between the derivatives of the strain components: *Jln + S^ = 2plL. (3.26) OX2 OX2 0X^X2 Equation (3.26) is called the compatibility equation for a two-dimensional strain field. The discussion given above can be generalized to three-dimensional elasticity. First we form the following derivatives of the strains defined in (3.23): £/.;,*/ = jiMijki + ujjkl)> ekljj = jfckjij + w/,JWj)»
60 REVIEW OF EQUATIONS OF SOLID MECHANICS Next, we add the first two equations and the last two equations separately: eij,kl + ekijj = ?(Ujjki + Ujjkl + uuu + «/,*//). We immediately note that the right-hand sides of the two expressions are the same (note that the order of the subscripts following the comma is interchangeable). Therefore, in Cartesian component form we have e*jM + ekijj = eijM + **/,//, (3.27) or in vector form (not shown here; see Exercise 3.33), V x (V x £) = % (zero dyad). (3.28) Equation (3.27) forms the necessary and sufficient conditions for the existence of a single-valued displacement field (when (he strains are given). Although Eq. (3.27) yields 81 equations for the three-dimensional case, only six of them are different from each other, and those remaining are either trivial or linear combinations of the six equations. These six equations arc given by 32eu t 82e22 _0 d2en 3*2 dx\ 3*10*2* 3*2 dx2 3*23*3' d2eu , 32<?33 0 d2ei3 -r _ n =2: dx] dx2 8*18x3' 3 dx — /_^23 3£)l3 3fl2\ _ d2€\\ *i \ 3*1 3*2 3*3/ 3*23*3' 3 / 3^13 3^i2 3g23\ _ 32e22 „ ^ 3*2 \ 3*2 3*3 3*1 / dx 13*3' 3 / 3^12 3*3 \ 3*3 3*23 3£n \ _ 32g33 3*1 3*2 / 3*i 3*2 For the two-dimensional case, the six equations reduce to a single equation t d2e\] 32<?22 32(?i2 —± + —^ - 2—±±_ = 0. (3.30) 3*2 3*2 3*i 3*2 It should be noted that the strain compatibility equations are satisfied automatically when the strains arc computed from a displacement field. Thus, one needs to verify
KINEMATICS OF DEFORMATION 61 the compatibility conditions only when the strains are computed from stresses that are in equilibrium. <■* Example 3.2 Given the strain tensor E = Errerer + Eoo^o^e in an axisymmetric body (i.e., Err and Egg arc functions of r and z only), we wish to determine the compatibility conditions on Err and Eoo- Using the vector form of compatibility conditions, Eq. (3,28), we obtain £ V7 S (dE™ , Eee-Err\ , , ^ 8£rrA A aEM A A F = V x E = I -j— + I e2e<9 + -j^Wr ~ -^-*r*e* n /SnT 9 /9E*0 , Eee-Err\f^ A.A Vx(F) "V~ + ^~J(erXe*)ez *2 9 £<96> /. „ x . , 1 /QEge , £00 - Err\ /A BeA A -w(e'x^)e'M^+^~Jrx^Jez ^ 19£,, A 9e,\. 1 9£rr ,. „ . fle* -rarrxWe'- + 9H^+^~)^xe^ a2^. /A A. A a2^ /A Noting that 9er . 366, ~^" — ^ ■^" = ~e''> e,. xe0=ez, eexe2=e,., ezxer = e<>, 0(7 00 and that a tensor is zero only when all its components are zero, we obtain ... _3_ (3E$8 EBe-Err\ 1 /9£e# £<?,?-£,,A = Q z dr \ dr r J r \ dr r ) d2E0e 1 9 eA: - -r^+ -■=-(£«•-£oe) = 0. 9r9z r az 9 /8£M , Eeo-E„\ . . d2E,r A A 92/^ w "ai2" ' ere'': ~~9?~ Example 3.3 We wish to check if the following two-dimensional strain field is compatible: en = c\x\ {x\ + xl), en = -c2x\, eyi = C3*J*2»
62 REVIEW OF EQUATIONS OF SOLID MECHANICS where c\, c*2, and 6*3 are constants. Using Eq. (3.30), wc obtain d2en , d2e22 n Pen 0 , n , —5- + -—5- - 2 = 2ci*i + 2c2Jri - 4c3jf|. dx\ dx\ dx\dx2 Thus the strain field is not compatible, unless C[ + ci — 2c3 = 0. 3.4 CONSTITUTIVE EQUATIONS 3.4.1 Introduction The kinematic relations and the mechanical and thermodynamic principles are applicable to any continuum iiTespective of its physical constitution. Here we consider equations characterizing the individual material and its reaction to applied loads. These equations are called the constitutive equations. The formulation of the constitutive equations for a given material is guided by certain rules (i.e., constitutive axioms). We will not discuss them here but will review the linear constitutive relations for solids undergoing small deformations. A material body is said to be homogeneous if the material properties are the same throughout the body (i.e., independent of position). In a heterogeneous body, the material properties are a function of position. An anisotropic body is one that has different values of a material property in different directions at a point, i.e., material properties are direction-dependent. An isotropic body is one for which every material property in all directions at a point is the same. An isotropic or anisotropic material can be nonhomogeneous or homogeneous. A material body is said to be ideally elastic when, under isothermal conditions, the body recovers its original form completely upon removal of the forces causing deformation, and there is a one-to-one relationship between the state of stress and the state of strain. The constitutive equations described here do not include creep at constant stress and stress relaxation at constant strain. Thus, the material coefficients that specify the constitutive relationship between the stress and strain components are assumed to be constant during the deformation. This docs not automatically imply that we neglect temperature effects on deformation. We account for the thermal expansion of the material, which can produce strains or stresses as large as those produced by the applied mechanical forces. The dependence of the constitutive properties on temperature and strains can be accounted for if required. Here, we review the basic constitutive equations of linear elasticity (i.e., generalized Hooke's law) for small displacements. ( 3.4.2 Generalized Hooke's Law The generalized Hooke's law relates the six components of stress to the six components of strain with respect to the coordinate axes (x\, X2, ^3) aligned with the material
CONSTITUTIVE EQUATIONS 63 planes as aU - Cijkf£kl* (3.31) where cr,y are the stress components, £// the strain components, and C/y*/ are the elastic coefficients referred to the material coordinates. Alternatively, Eq, (3.31) can be expressed in matrix form as \ax <T2 1^3 \a4 H5 [as: k = i ^11 C\2 C13 Ci4 C15 Ci6 C2\ C22 C23 C24 C25 C26 Qm ^32 ^33 ^34 Q35 <^36 C41 C42 C43 C44 C45 C46 C51 ^52 C53 C54 C55 C56 Ql Q2 Cfi3 Q>4 C65 Q)6 (3.32) The single subscript notation used in Eq, (3.32) for stresses and strains is called the contracted notation or the Voigt-Kelvin notation: (71=0-11, 0-2 = 0-22, 0-3=0-33, 04 = 023, 0-5=0-13, 0-6 = CT12, £l=£ll, £2 = £22, £3 = £33, £4=2^23, «5 = 2fii3, £<> = 2si2. (3.33) The two-subscript components Qy are obtained from C,y*/ by the following change of subscripts: 11 —^ 1, 22^2, 33-»3, 23^4, 13-► 5, 12-► 6. The resulting Cjj are also symmetric (Cjj = Cy /) by virtue of the assumption that there exists a potential function Uq = l/o(£//)» called the s/rawt energy density function, whose derivative with respect to a strain component determines the corresponding stress component: an = SUp (3.34) Such materials are termed hyperplastic materials. For hyperelastic materials, there are only 21 independent coefficients of the matrix [C]. When three mutually orthogonal planes of material symmetry exist, the number of elastic coefficients is reduced to 9, and such materials are called ortliotropic. The stress-strain relations for an orthotopic material take the form (3.35) \a[ cr2 |<T3 |(T4 1 5 [(re > = Cn cn Cn 0 0 0 Cm ■Cn. C23 0 0 0 Cn Cn C33 0 0 0 0 0 0 C44 0 0 0 0 0 0 c55 0 0 0 0 0 0 Q>6_ i e\ 1 £2 £31 £4 85 S6\
64 REVIEW OF EQUATIONS OF SOLID MECHANICS The inverse relations, strain-stress relations, are given by Ui r2 1*3 \s4 r5 [e*>. ► = 5ll Su S[J Sn S22 S23 5|3 523 533 0 0 0 0 0 0 0 0 0 1 ~e[ ~~e[ v\z Ei 0 0 0 _V2\_ 'I1 E2 V23 ~ E2 0 0 0 0 0 0 544 0 0 V31 £3 -M 0 f3 0 0 0 0 555 0 0" 0 0 0 0 566. 0-2 03 J 04 I &5 tcr6j 0 £3 0 0 0 0 1 G23 0 0 0 0 1 0 0 0 1 G12J (3.36) where 5// denote the compliance coefficients, [S] = [C]_1, E[, £"2, £3 are Young's moduli in 1,2, and 3 material directions, respectively, vy is Poisson's ratio, defined as the ratio of transverse strain in the jth direction to the axial strain in the /th direction when stressed in the /-direction, and G23, G13, G12 = shear moduli in the 2-3, 1-3, and 1-2 planes, respectively. Since the compliance matrix [5] is the inverse of the stiffness matrix \C\ and the inverse of a symmetric matrix is symmetric, it follows that the compliance matrix [S] is also a symmetric matrix. This in turn implies that the following reciprocal relations hold: V21 E2 V12 Eil V31 £3 V]3 E{; V32 Ei E2' or, in short, -^l — ~±- (no sum on /, j) Ei (3.37a) (3.37b) for z, v = 1, 2, 3. Thus, there arc only nine independent material coefficients, £l, #2> £31 G23, G13, G12, Vl2» Vi3, V23, / for an orthotropic material. When there exist no preferred directions in the material (i.e., the material has an infinite number of planes of material symmetry), the number of independent elastic coefficients reduces to 2. Such materials are called isotropic. For isotropic materials we have E\ = £"2 = E3 = £, G12 = G13 = G23 = G, and v\2 = V23 = ^13 = v.
CONSTITUTIVE EQUATIONS 65 Of the three constants (£\ v, G)y only two are independent and the third one is related to the other two by the relation E 2(1+v)' 3.4.3 Plane Stress Constitutive Relations (3.38) kplane stress state is defined to be one in which all transverse stresses are negligible. The strain-stress relations of an orthotropic body in plane stress state can be written as Sn 5.2 0 Sl2 S22 0 0" 0 S(6_ r -k «l\ cr6\ -L _M 0 Ei E2 Ei E2 G12J (3.39) 0 0 The strain-stress relations (3.39) are inverted to obtain the stress-strain relations (3.40) where the Qjj, called the plane stress-reduced stiffnesses, are given by <ri <*2 <?6 Gn G12 0 Qu Q22 0 0 0 g66 Qn = Ql2 = Q22 = S22 SuS22-Sf2 ~ I -V12V21' Sn _ VV1E2 S11S22 '12 Sn - fl2V21 E2 (3.41) Si I S22 - 5?2 ! ~" V|2V21 ' Q66 = -z— = G\2- ■>66 Note that the reduced stiffnesses involve four independent material constants, Ei, £2, v\2, and G12. The transverse shear stresses arc related to the transverse shear strains in an orthotropic material by the relations G44 0 0 Qss ]{:)• (3.42) where Q44 = C44 = G23 and Q55 = C55 = G13.
66 REVIEW OF EQUATIONS OF SOLID MECHANICS 3.4.4 Thermoelastic Constitutive Relations When temperature changes occur in the elastic body, we account for the thermal expansion of the material, even though the variation of elastic constants with temperature is neglected. When the strains, geometric changes, and temperature variations are sufficiently small, all governing equations are linear and superposition of mechanical and thermal effects is possible. The linear thermoelastic constitutive equations have the form <Tj=Cji[-ai(T-T0) + eil Sj = SjjCTi +ctj(T~ 7b), (3.43) (3.44) where a,- (i = 1, 2, 3) are the linear coefficients of thermal expansion, T denotes temperature, and 7b is the reference temperature of the undeformed body. In writing Eqs. (3.43) and (3.44), it is assumed that a\ and C\j are independent of strains and temperature. For an isotropic material, we have a\ = a2 = «3 = a. The plane stress constitutive relations for a thermoelastic case are given by (3.45) where AT = T - 7b is the temperature change from the reference temperature To and at (i = 1, 2) are the coefficients of thermal expansion of an orthotropic material in the x; -coordinate direction. In addition, Eq. (3.42) holds for the thermoelastic case. Example 3.4 Consider the problem of an isotropic cantilever beam bent by a load Fq at the free end (see Fig. 3.5). From the elementary beam theory [see Eqs. (4.35) and (4.36)], we calculate the stresses first: "1 <r2 i = 06\ Q\\ Qn 0 Cl2 Qn 0 0 0 666 £\ —a\AT e2 - 012AT S6 M3x2 Fqx\x2 VQ F0/72 2\ (3.46) h h Then the strains are given by (e\\ = o\\jE and eyi — o\2/2G = <ri2[(l H- v)/EJ): en = — F$X\X2 Eh ' ei2 = —veu = v- FqX)X2 Eh ' (3.47) M ^ 7*3 Figure 3.5 Cantilever beam bent by a point load, Fq.
CONSTITUTIVE EQUATIONS 67 where v is the Poisson ratio, EI3 is the bending stiffness (E is Young's modulus and I3 is the second moment of area about the *3-axis), and 2ft is the height of the beam. Next, we determine if these strains are compatible. Substituting e/y into Eq. (3.30), we obtain 0 + 0 = 0. Thus the strains satisfy the compatibility equations in two dimensions. Although the two-dimensional strains are compatible, the three- dimensional strains are not compatible. For example, using the additional strains, e33 = ~veu, £13 = ^23 = 0, one can verify that all of the equations except the last one in Eq. (3.29) are satisfied. Example 3.5 Consider the cantilever beam problem of Example 3.4. The displacement field in the beam according to the Euler-Bernoulli beam theory is given by (see Example 4.3) «i(*i,*2) = TFr(L2 " *?)*2, (3.48a) «2<*i, *2) = -^r {x\ - 3Z,2*, + 2L3), (3.48b) 6EI3 x and according to the Timoshenko beam theory (see Exercise 4.26), it is «i(*l, x2) = ^r(L2 - x\)x2, (3.49a) Zzii3 «2<*i. xi) = t~(x? - 3L2X] + 2L3) + -r^-{L - *,), (3.49b) where Ks is the shear correction factor, and A is the area of cross section. The transverse deflection u2 of the Timoshenko beam theory has an extra term due to the inclusion of the transverse shear force/strain effect. For a rectangular beam of height 2ft and width b, we have A = 2bh and h = 2ftft3/3. Since the strains are compatible in two-dimensional theory of elasticity, we wish to compute the displacement field u = eiwi -f- §2^2 of the beam using the known strains. Integrating the strains in Eq. (3.47), we obtain 3k 1 FoX[X2 F0xjx2 . - , . „ -n , en = — = =7— or u\ = ——^— + fi(x2)> (3.50a) dx\ Eh 2EI3 3u2 vFqxix2 vFoxxx* en = t— = —TT7— or u2 = ._. + /2(*i), (3.50b) 3*2 £/3 2E/3 where (/1, .ft) arc functions ofintegration. Substituting u\ and u2 into the definition of the shear strain 2ei2, we obtain = 3ki 3a2 = ft)-*? rf/i vP^l dh 6n 3x2 3xi ~~ 2EI3 dx2 2EH dx{
68 REVIEW OF EQUATIONS OF SOLID MECHANICS But this must be equal to that calculated earlier [see Eq. (3.47)]: Fqx\ df\ \>F<jx\ dh (1+V) „ /, 2 2\ ,, c,x ~iEh +dT2+-lEh + dTr~~^TFo(/l ~Xl)- a51) Separating the x\ and xt terms, we obtain _dft_ Jo_xz _ (1 + v)F0h2 = 4A _ (2 + i>)Fp 2 rfjcj 2E/3A'1 E/3 <fo2 2£73 *2' Since the left side depends only on X] and the right side depends only on X2, and yet the equality must hold, it follows that both sides should be equal to a constant, say co: dfx (2+v)F0 2 _ dfi , F0 2 (l + v)F0/t2 xt = en. Integrating the expressions for f[ and /2, we obtain (2 + v)F0 3^ /i fe) = —t^t—*2 + cox2 + ci, /2^l) 6e;3 ^0 3 d + ^oA2 a:i — x\ - cox] + C2, 6£/3 EJ3 where c\ and c2 are constants of integration that are to be determined. The displacements (u\, w2) are now given by u\(xi,x2) = «2(*1,*2) Fo 2 (2 + i;)F0 3j ■*1*2 H 7-^7 *2 + Cn-*2 + Cl, 2EI3 6EI3 (3.52a) (1 + v)F0h2 vFq 2 F0 3 fizT"*'*^1^^*1"^1"^ (3'52b) The constants co, ci, and c2 can be evaluated using the boundary conditions on the displacements. We impose the following boundary conditions: wi(L,0) = 0, M2(L,0) = 0. \dx2) = 0. (3.53) (^,0) Substituting the expressions for ui and «2 into the boundary conditions (3.53), we obtain «i(L,0)=0 -> c\ = 0 F0L3 u2a,o) = o \9x2) 6EI3 FqL2 = 0 -> — (L,0) 2£73 CtjL + C'2 = (l + y)F0/7.^L £J3 + co = 0 or co = FqL2 2Eh
EXERCISES 69 Finally, the displacement field of two-dimensional elasticity theory becomes ui{xuX2) = 2iML2 ~xl)X2+iJiinrxl (3-54a) + m^(L_xi) + ^_wl <3,4b) The values of the constants c, c\9 and ci depend on the boundary conditions used. For example, if one uses (dit2/dx\) = 0 in place of (du1/8x2) = 0 at jci = L and xz = 0, the resulting displacements would be ll](X] >X2) = sfM7*2 ~ x*)x2 ~ (^FirX2^h2 ~x^ (3,55a) ui{x\, x2) = ^-(xj - 3L2x{ + 2L3) + ^tXlxI (3.55b) ois/3 ' 2E/3 A comparison of the elasticity solution (3.54a,b) against the Timoshenko beam solution (3.49a,b) indicates that the elasticity solution has some higher-order terms that are negligible (x2 ~ h). Also one can see that (GA - 3Eh/2[(i 4- v)h2]) the shear correction factor is Ks = 2/3. EXERCISES 3.1 Let an arbitrary region in a continuous medium be denoted by R and the bounding, closed surface of this region be continuous and denoted by S. Let each point on the bounding surface move with the velocity v.y. It can be shown that the time derivative of the volume integral over some continuous function g(r, /) is given by j-j^Q(rj)dV^J(^-dV + £ Gv, . n dS. (a) This expression for the differentiation of a volume integral with variable limits is sometimes known as the three-dimensional Leibniz rule. Let each element of mass in the medium move with the velocity v(r, /) and consider a special region R such that the bounding surface S is attached to a fixed set of material elements. Then each point of this surface moves itself with the material velocity, that is, vs = v, and the region R thus contains a fixed total amount of mass, since no mass crosses the boundary surface S. To distinguish the time rate of change of an integral over this material region.
REVIEW OF EQUATIONS OF SOLID MECHANICS we replace djdt by D/Dt and write £tfRQ(r't)dvsLdldv+iQv-hds> (b> which holds for a material region, that is, a region of fixed total mass. Show that the relation between the time derivative following an arbitrary region and the time derivative following a material region (fixed total mass) is jt J Q(r, t)dV^^J 2(r, t)dV + j> Q(ys - v) • n dS. (c) The velocity difference v — \s is the velocity of the material measured relative to the velocity of the surface. The surface integral (f> Q(v-vs)>ndS (d) thus measures the total outflow of the property Q from the region R. Let Q = p(r, t) denote the mass density of a continuous region. Then conservation of mass for a material region requires that Show that, for & fixed region (\s = 0), conservation of mass can also be stated as dt JR pdV = -<p pvnclS (b) or /. -£dV = -<* py-ndS. (c) Interpret these equations physically. In the field description of a continuous variable <j> = 0(r, t), let the field position be a function of time such that r = r(f). Deduce that the total time derivative of <£ can be written as d<f> d(/) dr ^=a7 + ^,grad0- (a) This arbitrary total time derivative corresponds to a change in <f> following a change in r with time. If we let r correspond to the position of a fixed material element, then dr/dt = v corresponds to the velocity of the material element. To distinguish the time rates of change following a material element from other arbitrary changes, we write D<f> 3(f)
EXERCISES 71 This is the differential time rate of change of a field variable 0 (r, t) following a material element. It is referred to as the material derivative, the substantial derivative, or the Eulerian derivative. The corresponding material derivative for integrals was defined in Exercise 3.2. By means of vector identities, show that the continuity equation for mass conservation can be written as Dp —- + p div v = 0. (c) 4 The material derivative operator D/Dt corresponds to changes with respect to a fixed mass, that is, p dV is constant with respect to this operator. Show formally, by means of Leibniz's rule, the divergence theorem, and conservation of mass, that §; L>*dy-}.>%"■ w 5 Letting a finite volume R shrink to an infinitesimal volume dV, show, by setting <f> = 1 in Leibniz's rule and by use of the divergence theorem, that div v can be interpreted as divv= hrn J- —(rfV). (a) JV-+0dVDt The right-hand side can be interpreted as the rate of volumetric strain following a materia] particle, called the dilatation rate. 6 The acceleration of a material element in a continuum is described by Dv d\ , _^_+v.gradv. (a) Show by means of vector identities that the acceleration can also be written as Z)v dv (v2\ _ = _+grad^_j_vxcurlv. (b) This form displays the role of the vorticity vector, curl v. 7 Deduce that Dy D /yy\ Note that (v • v)/2 = v2/2 is the kinetic energy per unit mass of a material particle.
72 REVIEW OF EQUATIONS OF SOLID MECHANICS 3.8 Deduce that /Dv\ DQ curl I — 1 = + Q div v - Q . Vv, (a) where Q = curl v is the vorticity vector, 3.9 Newton's second law of motion in its elementary form ma = F holds strictly for a point particle of fixed mass m. For a material region of continuously distributed mass, Newton's second law reads DiL Dt f pvdV = Ft (a) where F is the sum of all the forces acting on the material in R. Deduce by means of the results of Exercise 3.1 that the corresponding law of motion for a region of variable mass is — / pvdV = F + * pv(vv - v) • n dS, dt JR Js (b) where vv is the velocity of points on the surface S bounding the variable mass region R. 3.10 Newton's second law of motion applied to a continuum states that the rate of change of momentum following a material region of fixed mass is equal to the sum of all the forces on the region. When the forces are divided into surface forces and body forces, Newton's second law reads: ■J- / pvdV = <&A-adS+ [ ptdV, (a) Vt JR Js Jr where a- is the surface stress tensor, f is the body force per unit mass, p is the mass density, and v is the material velocity. Since the material particle mass p dV is constant with respect to the material time derivative D/Dt, make use of the divergence theorem and obtain the differential form of Newton's second law of motion for a continuum: p— = div a + pf. (b) 3.11 Multiply the continuity equation dp — + div(pv) = 0 (a) at i by the velocity v and add the result to the left-hand side of the momentum equation (Newton's second law) in Exercise 3.10. After use of vector identities, obtain the result d <* ~ Gov) + divGovv -<x)=pt, (b) at
EXERCISES 73 which is called the conservation form of the momentum equation. The combination pvv is called the momentum-flux tensor. 12 Take the scalar product with v of the equation for Newton's second law in Exercise 3.10 and obtain the equation of change for the kinetic energy of a material particle in a continuum: £(£)- v • div cr+pv- f. (a) How do you interpret the rate-of-work terms on the right-hand side? 13 Let e denote the thermodynamic internal energy per unit mass of a material. Then the equation of change for total energy of a material region can be written: -^- / p(e+ yW = &&.c.YdS+ f pt-ydV-iq-ndS. (a) The first two terms on the right-hand side describe the rate of work done on the material region by the surface stresses and the body forces. The third integral describes the net outflow of heat from the region, causing a decrease of energy inside the region. The heat-flux vector q describes the magnitude and direction of the flow of heat energy per unit time and per unit area. By suitable operations, obtain the differential form of the energy equation: p — I e -j- — J — div(<7 « v) -j- pi • v — div q. (a) Subtract the contribution from kinetic energy and obtain p—— = div(cr • v) — v • div a — div q. (b) This is called the thermodynamic farm of the energy equation for a continuum. 14 The total rate of work done by the surface stresses per unit volume is given by div(cr • v). The rate of work done by the resultant of the surface stresses per unit volume is given by v • div cr. The difference between these two terms yields the rate of work done by the surface stresses in deformation of the material particle, per unit volume. Show that this can be written as div(cr . vj — v • diva = a: (Vv) = a: Vv (a symmetric) = ^ a: [Vv -j- (Vv)1 ] (a symmetric), (a)
74 REVIEW OF EQUATIONS OF SOLID MECHANICS 3.15 Derive the transformation relations relating the normal and shear stresses, an and os, on a plane whose normal is ii = cosflei -f sinflej to the stress components an, (T22, and (7|2 = cT2i ontheei and 82 planes (seeFig. E3.15): an — an cos2 0 + 0*22 sin 0 -f 2an sin 0 cos 0, (a) as — (022 — en) sin 0 cos 0 -f 0*12(cos2 0 - sin2 0). Note that 0 is the angle measured from the positive xi-axis to the normal to the inclined plane. Then show that the principal stresses at a point in a two-dimensional body are given by <?11 + (T22 2 fan -<T22\2 . 2 (b) <y\\ + 022 _ //(TU -Q"22\ 1 W" 2 ) TU« -r O'min = \ I ~ 1 + a 2 and that the orientation of the principal planes is given by ^ = -tan~1 f 2aI2 1 \_cr\i -(T22J M, °22 Figure E3.15 A n 3.16 Determine whether the following stress fields are possible in a structural member free of body forces: x2 (a) gu = c\x\ + c2x2 + C3X[X2> <y\i = ™c"3y - ci*2* ^22 = t'4*l +C\X2- (b) <7[[ = X2 ~ 2^fiA'2 + CX3, <J\2 = —X]X2 + x\, ^13 = -*1*3, 0-22 = x\, (T23 = -*2*3» ^33 = (*1 + A'2)*3- (C) CX[i = 3X1 + 5^2, <Ti2 = 4^1 - 3X2, CT22 = 1X\ — 4,V*2.
EXERCISES 75 3.17 Given the following state of stress (au = 0,7): an = -2*^ on = — 7 4- 4*1*2 + *3> cri3 = 1 + x\ — 3*2, (722 == 3*f - 2*1 + 5*3, CT23 = 0, (T33 = -5 + ATI + 3*2 + 3*3, determine the body force components for which the stress field describes a state of equilibrium. 3.18 For the cantilever beam bent by a point load at the free end (see Fig, 3.5), the bending moment A/3 about the x-3-axis is given by A/3 = — Fo*j • The bending stress an is given by A/3*2 FqX]X2 (Til = = , h h where I3 is the moment of inertia of the cross section about the *3-axis. Starting with this equation, use the two-dimensional equilibrium equations to determine the stresses 0^2 and a\2 as functions of *i and *2. 3.19 Repeat Exercise 3.18 for the case in which the cantilever beam is bent by uniformly distributed load ^o/unit length applied at the top (i.e.. at *2 — —h). 3.20 For the state of stress given in Exercise 3.17, determine the stress vector at point (*i, *2, *3) on the plane x\ + *2 + *3 = constant. What are the normal and shearing components of the stress vector at point (1,1,3)? 3.21 Find the principal stresses and their orientation at point (1,2,1) for the state of stress given in Exercise 3.17. 3.22 Find the maximum principal stress and its orientation for the state of stress [*] = 3.23 Find the linear strains associated with the displacements u\ = [*i*2(2 - *i) - ci*2 + 6*2*2], *2 3.24 The two-dimensional displacement field in a body is given by it\ = *i \x\x2 + c\ (2(?2 + 3t*2*2 — *!)]> / 3 13 where C[ and C2 are constants. Find the linear and nonlinear strains. 3 5 8 5 1 0 8 0 2 psi. •
76 REVIEW OF EQUATIONS OF SOLID MECHANICS 3.25 Determine the displacements and strains in the (xi , X2) system for the bodies shown in Figure E3.25. 3.26 Determine the displacements and strains in the (x\, X2) system for the bodies shown in Figure E3.26. ► jr.,*. 3.27 Find the linear strains associated with the displacement field U [ = itf (X\, X2) + X301 (XI, X2), lt2 = u\{X\ , X2) + *3#2(*l> X2), K3 =Us(X[tX2). 3.28 Determine whether the following strain fields are possible in a continuous body: (a) [e] - (b) M = "(*?+*£) * X}X2 4 J' 'X3 (x\ + x\) 2X1X2X3 X3 .*2 XI 2X|X2X3 X3 Xl X^ 3.29 Find the normal and shear strains in the diagonal element of the rectangular block in Exercise 3.25. 3.30 The biaxial state of strain at a point is given by e\\ =800 x 10~6 in./in., £22 = 200 x 10~6 in./in., £\2 = 400 x 10~6 in./in. Find the principal strains and their directions.
EXERCISES 77 3.31 Find the axial strain in the diagonal element of Exercise 3.25, using (a) the basic definition of normal strain, and (b) the strain transformation equations. 3.32 Using the definition of V and the nonion form of a, show that the equations of motion in the cylindrical coordinate system arc given by dcfrr , 1 dcrro , darz 1 d2ur + IT- + ~z h -far - GOO) + PoA = A>" 8r ' r 30 " 3z r^ "^ ' ™'-™ dt2> 3arg 1 aa^ 3oj?z 2o># 32 wo 9arz 1 daez d<rzz arz d2uz dr r 36 dz r atL 3.33 Consider the displacement vector in a polar axisymmetric problem, u = arer + uzez. The strain tensor in the cylindrical coordinate system can be written in the dyadic form as s = err$rr + eee*e*e + szzezez + ero(*r*& + Mr) + ^z(ere2 + e2er) + e$z(eoez + eze#). Show that the only nonzero linear strains are given by du dltr Ur 1 f'dur dltz\ z_ dz ' 3.34 Using the definitions of V and u in the cylindrical coordinate system, show that the linear strains are given by dur 1 /1 dur duo u$\ 1 /dur duz\ £,.r = _ ffrt = _^-_ + ___j, 8n = -^— + —j, ur 1 duo 1 (due 1 9«*\ duz 3.35 Establish the vector form of the strain compatibility conditions V x (V x e)T = 0. where e is the infinitesimal strain tensor.
78 REVIEW OF EQUATIONS OF SOLID MECHANICS REFERENCES 1. Bird, R. B., Stewart, W. E., and Lightfoot, E. R, Transport Phenomena, John Wiley, New York (I960). 2. Malvern, L. E., Introduction to the Mechanics of a Continuous Medium, Prenlice-HalJ, Engiewood Cliffs, NJ (1969). 3. Reddy, J. N., and Rasmussen, M. L., Advanced Engineering Analysis, John Wiley, New York (1982); reprinted by Kricger Publishing, Melbourne, FL(1991). 4. Bonet, J., and Wood, R, D., Nonlinear Continuum Mechanics for Finite Element Analysis, Cambridge University Press, New York (1997). 5. Hjelmstad, K, D., Fundamentals of Structural Mechanics, Prentice-Hall, Englcwood Cliffs, NJ (1997). 6. Ogden, R. W, Non-Linear Elastic Deformations, Ellis Horwood, Chichester, UK (1984). 7. Batchelor, G. K., An Introduction to Fluid Dynamics, Cambridge University Press, Cambridge (1967). 8. Schlichting, H., Boundary-Layer Theory (translated by J. Kestin), 7th edition, McGraw- Hill, New York (1979).
4 WORK, ENERGY, AND VARIATIONAL CALCULUS 4.1 CONCEPTS OF WORK AND ENERGY Consider a material particle, moving from point A to point B along some path in space under the influence of a force F, which can be time-dependent. The position of the particle is measured from a fixed origin by position vector r. Then the work dW performed by the force F in moving the particle by an infinitesimal distance (or displacement) dr = du along the path over an interval of time dt is defined as (see Fig. 4.1a) dW = F • du = F[ du[ + F2du2 + F$du$. (4.1) In other words, work done is the product of the displacement and force in the direction of the displacement. The total work done, W, by the force F in moving the particle from point A to point B is given by W= f F-rfu. (4.2) By definition, work done is a scalar quantity, and it is positive whenever both displacement and force have the same direction and negative if they are in the opposite directions. Since u depends on the chosen reference frame, W also depends on the choice of the reference frame. Thus, work is a relative quantity. However, work done does not depend on the path but only the end points, W = Wr — Wa- If the reference frame is chosen such that Wa = 0, then W — W#. Analogous to the work done by a force, we can define the work done by a couple (i.e., a pair of equal, opposite forces) M in moving through an infinitesimal rotation 79
80 WORK, ENERGY, AND VARIATIONAL CALCULUS (a) **,v (b) Figure 4.1 (a) Motion of a particle under the action of a force, (b) Motion of a particle under the action of a moment. dO as dW =M- dO = Mi d0\ + M2d02 + M3 dO^ The total work done in a finite rotation from point A to point B is W -j: M • d6. (4.3) (4.4) From Eq. (4.1), it is clear that the rate of work done by force F in moving the particle through an infinitesimal distance dr is F • dr/dt. But dr/dt = v is the velocity of the particle. Hence, the rate of work done is equal to F« v. Now the work done by a force F in moving through the infinitesimal distance dr is rfW = F-dr = F. vd/. Hence, the total work done during the time interval (t\ > t2) is 'F-vrfA. ' W =i:< (4.5) The definition of work can be extended to material bodies, which can be viewed as a collection of material particles. If work is done on all particles of the body, the total work on the body is the sum of all such works. A particle of a deformable body
CONCEPTS OF WORK AND ENERGY 81 is generally subjected to internal and external forces. Then the total work done on a particle is the sum of the work done by internal and external forces: W = Wf + WE. (4.6) If a body is subjected to external point forces F], F2,.,., F" that displace the points of action by displacements Ari, Ar2,..., Ar„, respectively, then the work done by the forces on the body during the time interval Af is the sum of the work done by individual forces in moving through their respective displacements: n n WE = - J^F1' • Ar;- = - £V ■ v/ A/, where v/ denotes the mean velocity Ar,/A/. The minus sign indicates that the work is expended on the body as opposed to work stored in the body. If the time increment At approaches zero, the sum approaches a limiting value that is represented by an integral, as in Eq. (4.5). Iff = f (x) is the distributed force (measured per unit volume) acting on a particle occupying position x in a body with volume Q, and u = u(x) is the displacement of the particle, then the work done on the particle is f • u. The work done on all particles occupying the elemental volume d£2 is (f • u) dQ. Hence, the total work done on the body is the sum of work done on all particles occupying the body: WE = - I (t-VL)dQ. In calculating the external work done, the applied (external) forces (or moments) are assumed to be independent of the displacements (or rotations) they cause in a body. Wg is sometimes called the potential energy due to applied loads and is denoted by V. On the other hand, the internal forces generated inside the body due to the application of the external loads are related to the displacements they move through by the constitutive relations and strain-displacement relations. In general, the work done by internal forces in moving through their respective finite displacements is not equal to the work done by (or potential energy due to) external forces in moving through their respective displacements. However, in using energy principles and variational methods for solving problems, we only calculate the work done (internal as well as external) by forces/moments in moving through a set of imagined infinitesimal displacements/rotations. In this case, the sum of the internal work done and external work done, with the minus sign in place, is equal to zero. To understand further the difference between work done by external forces and internal forces, consider a spring-mass system in static equilibrium (see Fig. 4.2a). Suppose that the mass m is placed slowly (to eliminate dynamic effects) at the end of the spring. The spring will elongate by an amount eo, measured from its unde- formed state. In this case, the force is due to gravity and it is equal to F = nig. Clearly, F is independent of the extension eo in the spring, and F does not change
82 WORK, ENERGY AND VARIATIONAL CALCULUS F^mg Fi (b) F =&? "t * F =ke (c) de Figure 4.2 A spring-mass system in equilibrium, (a) Elongation due to weight mg. (b) External work done, (c) Internal work done. during the course of the extension e, going from 0 to its final value eo (see Fig. 4.2b). The work done by F is We = -(FeQy Next consider the force in the spring Fv, It goes from 0 to its final value as e goes from 0 to eo (see Fig. 4.2c). The work done by Fs in moving through de is Fsde. To obtain the total work done, we must integrate, because Fs depends on e, from 0 to e$\ Wi pea Fs (e) de. Tf the spring is assumed to be linearly elastic with spring constant fc, we have Fs = ke. Hence W, = l-kel t f At equilibrium we have F/ = F, where Fy denotes the final force in the spring. Now suppose that we wish to elongate the spring from eo to eo + Ae, Ae being infinitesimally small. Then the additional (or incremental) work done by F = mg and Fs = Fs (eo) are simply AWe = —FAe = —mgAe, AWj = FsAe = /ten Ae, which shows that AW/ — —AWe since F = Fs. Example 4.1 Figure 4.3 shows a rigid link in equilibrium. The initial, no-load position corresponds to 6 = 0, and the displacement to the equilibrium position is small (i.e., 9 is small). We wish to determine the work done by external forces as well as internal forces when the link is given an infinitesimal rotation A0 from its equilibrium position. We shall assume that the springs are linearly elastic.
CONCEPTS OF WORK AND ENERGY 83 Figure 4.3 A rigid link in equilibrium. The work done by external forces is given by AWE = -(PAv + MA9) = -(PL[sin(0 + A0) *-PL(6 + AG-G)- -sin6>] + MA0) MAO = -(PL + M)A0. Once again, the negative sign indicates that the work is done on the body. The work done by internal force Fs in the extensional spring k\ and internal moment Ms in the torsional spring ki is given by AW/ = FsAv + MsA0 = k]vAv + k20A0 *(k]L2 + k2)eAe. Note that AW/ = -AWe (because P - k\L6 and M = jfc20) or AWj + A We = 0. As we shall see later, this is known as the principle of virtual displacements. Energy is the capacity to do work. It is a measure of the capacity of all forces that can be associated with matter to perform work. Work is performed on a body through a change in energy. For example, consider Newton's second law of motion for a material particle, F = m(d\/dt), where m is the mass and v is the velocity vector. If r is the position vector from a fixed origin to the particle, the work done by F in moving the particle through an infinitesimal distance dv is d\ d\ dW = F • dv = m— • dr = m— • v dl dt dt d \ . — — I -mvl dt \2 dt. where v is the magnitude of the velocity vector v = dr/dt. Since the kinetic energy of the particle is K = (I/2)mi>2, it follows that dW = ^-dt = dK or W = AKy dt (4.7) where AK is the increment of kinetic energy resulting from work W. Equation (4.7) states that the work done by a force is equal to the increase in the kinetic energy.
84 WORK, ENERGY, AND VARIATIONAL CALCULUS 4.2 STRAIN ENERGY AND COMPLEMENTARY STRAIN ENERGY The first law of thermodynamics gives rise to the following energy equation (see Exercise 3.13) in material description (with no heat generation): p— = a : Vv-V q, (4.8) dt where e is the internal energy per unit mass, q the heat flux vector, a the Cauchy stress tensor, and v the velocity vector. For elastic bodies under isothermal conditions, the internal energy consists of only stored elastic strain energy, denoted pe = Uq and measured per unit volume. Tn this case (i.e.. elastic bodies under isothermal conditions), the energy equation takes the form ^ = ?:Vv (4.9) dt or (since v = du/dt) dU0 = a : V(</u). (4.10) Note that V(du) = d(Vu) = rf (i[Vu + (Vu)T] + |[Vu - (Vu)T]j = de + aco, where s is the linear strain tensor and go is the rotation tensor. Since the stress tensor is symmetric, the double-dot product o : co is zero, giving dUo = ex : de = 0ijd£U. (4.11) Integrating the above equation, we obtain the expression for the strain energy density of an elastic body: U0 = I ' crnden. (4.12) = / Wj df;ij- JO The existence of a scalar function Uq of strains such that the stresses are derivable from t/o is of special importance. Such stresses satisfy the energy equation, and consequently they are said to be conservative. We have from Eq. (4.11) the result — -ay. (4.13)
STRAIN ENERGY AND COMPLEMENTARY STRAIN ENERGY 85 Often one assumes the existence of Uq(sjj) such that Eq. (4.13) holds (even for nonlinear elastic bodies with large strains), and Eqs. (4.11) and (4.12) follow from there. Wc assume that the expression in Eq. (4,12) is valid for all elastic (linear or nonlinear) bodies with linear or nonlinear strain-displacement relations. When thermal effects are to be included, one assumes the existence of a free energy function ty, called Gibb\s free energy: p*(eUt T) = UQ(sih T) - f>Tn(eih T), (4.14) where T denotes temperature and ?? the specific entropy, such that hold. Equation (4.12) can also be arrived by considering the work done by internal forces (i.e., stresses) in moving through displacements. First wc consider axial deformation of a bar of area of cross section A. The free-body diagram of an element of length dx\ of the bar is shown in Fig. 4.4a. Note that the element is in static equilibrium, and we wish to determine the work done by the internal force associated with stress a^, where the superscript / indicates that it is the final value of the quantity. Suppose that the element is deformed slowly so that axial sU'ain varies from 0 to its final value e\l. At any instant during the strain variation from e\\ tof-:\\ + den, we assume that <j\j (due to en) is kept constant so that equilibrium is maintained. Then the work done by Ao\i in moving through the displacement ds\\dx[ is Ag\\ de\\dx{ — <7\\ d£\\(Adx\) = dUo(A dx\), where dUo denotes the work done per unit volume. Referring to the stress-strain diagram in Fig. 4,4b, dUo represents the elemental area under the stress-strain curve. The complementary area over the curve is given by dlly = £]\ d<T]\. The total area under the curve is obtained by integrating from zero to the final value of the strain (during which the stress changes according to its relation to the strain): tfo = / <r\\de\\, where the superscript / is omitted as the expression holds for any value of e. The quantity Uq is laiown as the strain energy density. The complementary strain energy density is given by Jo
86 WORK, ENERGY, AND VARIATIONAL CALCULUS U) A _ _ dxi dx\ — £\\ dx\ J l_ -Xy I I J L_ I H-4 dxi H X\ (a) (b) Figure 4.4 Computation of strain energy for the uniaxial case, (a) Work done by force a\ \A6x\. (b) Definitions of various energies. The internal work done by Ao\[ over the whole element during the entire deformation is dW[ = / an de\\{Adx\) = Uo(Adx\)t Jo and the total work done (or energy stored) in the entire body is obtained by integrating over the length of the bar: Wj = I AUodxi. -I Jo This is the internal energy stored in the body. U is called the strain energy and denoted by U = Wj. Similarly, the complementary strain energy is Wf = </* = [" AUfjdxi. Jo Next, we extend the above discussion to the three-dimensional case. Consider the rectangular parallelepiped element (taken from inside an elastic body) of sides dx\,dx2, and dx$ shown in Fig. 4.5a. Suppose that the element is subjected to a system of external forces that vary slowly until they reach their final values, so that the equilibrium is maintained at all times. The forces due to the normal components of stresses arc (negative of the same forces act on the opposite faces) <7ii dx2dx$, on dx$dx\, 0*33 dx[d.X2, and the forces due to the shear stresses on the six faces are ai2dxidx^ an dx^dxz, oi\ dxidx$, 023 dx$dxi, cru dxidxz, ayi dx^dx\.
STRAIN ENERGY AND COMPLEMENTARY STRAIN ENERGY 87 Figure 4.5 Computation of strain energy for the three-dimensional case. At any stage during the action of these forces, the faces of the parallelepiped will undergo displacements in the normal directions by the amounts d£\\dx\, ds2idx2, and dey$dx$, and distort by the amounts 2ds\2dx[, 2d£23dx2, and Ids^dx^, As examples, the deformations caused by normal force an dx.2d.x3 and shear force 0*21 dxidx3, each acting alone, are shown in Fig. 4.5b. The work done by individual forces can be summed to obtain the total work done by the simultaneous application of all of the forces, because, for example, an x\ -directed force does no work in the X2 or X3 directions. The work done, for instance, during the application of the force an dxzdxi is given by (force times displacement) (an dx2dxi)(d£]]dx\) = an deudV. Similarly, the work done by the shear force 02\ dx]dx$ is given by (a2 \dx\ dx$) (d£21 dx2) = a21 cI £21 d V. The internal work done by ail forces in varying slowly from zero to their final values is given by <T]]deu+ <r\2de\2-\ 1- / crttd£33)dV
88 WORK, ENERGY, AND VARIATIONAL CALCULUS where sum on repeated indices is implied but the limit corresponds to a fixed i and j. The expression U0= / CijdEr, (4.16) Jo is the strain energy per unit volume or simply the strain energy density. The complementary strain energy density, C/q , can be computed from J7*= P' ejjdcfi,. (4.17) Jo The total internal work done by internal forces is given by the integral of Uq over the volume of the body, and it is denoted by U: U= f U0dV, (4.18) Jv and it represents the mechanical energy stored in the body. It is called the strain energy of the body. The complementary strain energy U* of an elastic body is defined by [/*= f U£dV. (4.19) Analogous to Eq. (4.13), the strains may be derived from the complementary strain energy density function Example 4.2 Consider the pin-connected structure shown in Fig. 4.6. The members of the truss are made of an elastic material whose uniaxial stress-strain behavior in tension and compression are given by a ~ \ .— (4.20) E = Young's modulus Aj- cross-sectional area of the /ih member Figure 4.6 Pin-connected structure (truss).
STRAIN ENERGY AND COMPLEMENTARY STRAIN ENERGY 89 We wish to compute the strain energy density E/o> the complementary strain energy density U^, the strain energy £/, and the complementary strain energy U* of the structure. To compute Uq, we must find the strains in each member of the structure. On the other hand, U£ is computed using the stresses in each member. In both cases, the ultimate result will be expressed in terms of the applied (known) load P. Of course, because of the nonlinear stress-strain relation (4,20), Uo ^ Uq . However, £A)/ + ^0/ for the ith member should equal the product (no sum on /) 07s,-. In order to compute the stress and strain in each of the two members, we first compute the forces. From the free-body diagram of joint A, the axial forces in each member can be calculated as (F2 sin 30° = P and F\ + F2 cos 30° = 0) Fi = -V3P, F2 = 2P. Then the stresses in each member are calculated as (with rr = F/A) IP (4.21) V3P G\ = " , <*2 = —, A{ A2 (4.22) where A[ and A 2 denote the cross-sectional areas of members 1 and 2, respectively. The strains are calculated using the stress-strain relation (4.20) (using e = ± <j2/E2): e{ =■ 3P2 £2 = 4PZ ~A2E2' (4.23) Then the strain energy density of each member of the structure is Uqx A]E2 p f£i , , , 2E V2 2V3P3 = <Xide\=\ (-EV^dei=-—(-ei)3/2 = JO JO J The strain energy of the structure is U = f U0iclV+ f U02dV = UoiAiL\ + U02A2L2 (4.24) '/3V3P3Li\ (SP3L2 a a\e2 r A\E* (4.25) where L; is the length of the /th member.
90 WORK, ENERGY, AND VARIATIONAL CALCULUS The complementary strain energy density of each member of the structure is (4.26) 82 d<J2 The complementary strain energy of the structure is £/*^ f U*xdV+ f U*2dV JV] JV2 = USlA\L^+US1A2L2 - I r/3V3P3LA /8P3L2\1 3^ A^ r\A\E*)y (4.27) Example 4.3 (Eider-Bernoulli Beam Theory) Consider bending of a sti'aight beam according to the classical (Euler-Bernoulli) beam theory. We wish to compute the strain energy density, complementary strain energy density, strain energy, and complementary strain energy due to extension, bending, and transverse shear. For the strain energy density computation, the strain will be expressed in terms of the axial and transverse displacements, whereas for the complementary strain energy density computation, the stress will be expressed in terms of the axial and transverse forces, bending moment, and torque (valid only for circular cylindrical members). The three modes of deformation are separable for straight members with small strains. In particular, the following development treats bars as a subset of beams. We begin with a review of the basic equations of bars and beams. The Euler- Bernoulli hypothesis concerning the kinematics of bending deformation assumes that straight lines perpendicular to the beam axis before deformation remain (a) straight, (b) perpendicular to the tangent line to the beam axis, and (c) inextensi- ble after deformation. These assumptions lead to the following displacement field (see Fig. 4.7a): u = uq(x) —z—t-, dx v = 0, w = WoC*0. (4.28) where (w, v, w) are the displacements of a point (x, y, z) along the x, y, and z coordinates, respectively, and (wo, wo) are the displacements of the point (jc, 0,0). Under the assumption of smallness of strains and rotations, the only nonzero strain is erX = duo dx d2WQ l~d^' (4.29) Suppose that the beam is subjected to distributed axial force f(x) and transverse load q (x). Summing the forces and moments on an element of the beam (see Fig. 4.7b)
STRAIN ENERGY AND COMPLEMENTARY STRAIN ENERGY 91 q(x) M{x) t T t t t T j M(x)+AM(x) M{x) >N(x)+AN(x) N(x) q{x) a* Gxx VW V(jc)+4V(jr) (c) V(x) V(x)+AV(x) (b) Figure 4.7 Bending of beams, (a) Kinematics of deformation of an Euier-Bernoulli beam, (b) Equilibrium of a beam element, (c) Definitions (or internal equilibrium) of stress resultants. gives the following equilibrium equations: dN E'.-«= -£-/« dx ^ dV (4.30a) (4.30b) (4.30c) where N(x) is the axial force, M{x) the bending moment, and V(x) the shear force. These are known as the stress resultants and they can be defined in terms of the stresses <jxx and <jxz on a cross section as (see Fig. 4.7c) N(x)= [ crxxdA, M(x) Ja Ja -L rxxZ dA, V(x) = / axzdA. (4.31) Here A denotes the area of cross section. Note that for linear strain measures, Eq. (430a), which governs the axial deformation of bars (i.e., axially loaded members), is uncoupled from Eqs. (4.30b,c). Hence it can be solved independently of beam equations (4.30b,c).
92 WORK, ENERGY, AND VARIATIONAL CALCULUS The stress resultants (N, M) can be related back to the stress axx using the linear elastic constitutive relation for an isotropic material as (duo d2w0\ 0ju = Es"ssE{i7-l-d^)- (432) First, note that N(x) = f <rvv dA = E^f dA = EA^, (4.33a) J a dx J a dx Mix) = ^ axxZ dA = EJA{^- zfg?) z dA = -EI^ (4.33b) or duo _ N d2wo __ M Using the relations in Eq. (4.34) in Eq. (4.32), we obtain (4.34) A I where / is the moment of inertia about the axis of bending ()>-axis) and z is the transverse coordinate. Note that the x-axis is taken through the geometric centroid of the cross section so that jAzdA =0. The Euler-Bernoulli beam theory is inconsistent with the reality in the sense that the shear stress computed through the stress-strain relation (axz = 2Gexz) is zero, since exz = 0. Consequently, V(x) is zero as per the third equation of (4.31). However, in computing the complementary strain energy, we assume that the shear strain is derived from yxz = 2sxz = (<jxz/G) and shear stress is computed, as V is not zero in reality but given by Eq. (4.30c), using the relation (see a book on mechanics of materials for its derivation) VQ <r« = -r^. (4.36) lb Here Q(z) denotes the first moment of the hatched area (see Fig, 4.8a) about the centroidal axis y of the entire cross section, and b is the width (or thickness) of the cross section at z where the longitudinal shear stress (azx) acts. The hatched area is only a portion of the total area of cross section that lies below the surface on which the shear force acts; Q is the maximum at the centroid (i.e., z = 0), and axz is the maximum wherever Q/b is the maximum. A circular cylindrical member subjected to torque T about its longitudinal axis develops shear stress ax$ in its cross section (see Fig. 4.8b). It can be shown that the stress is related to the torque applied by the relation Ojrf(r) = y, (4.37)
STRAIN ENERGY AND COMPLEMENTARY STRAIN ENERGY 93 ■rfr ^Wa (a) (b) Figure 4-8 (a) Calculation of transverse shear stress on a cross section, (b) Torsion of a circular cylindrical member. where 7* is tho radial distance from the axis of the beam to the location where <jxq acts and ,/ is the polar moment of area of the entire cross section. For a solid circular member of diameter d, we have J = nd4/32. The stress is the maximum at the outer surface (i.e., r = d/2) of the member. Returning to the computation of various energies, we shall compute them for all components of stresses in a beam. For noncircular members, the part involving T should be omitted. The strain energy density of the beam subjected to axial and bending loads is given by Uq = / crXjc dexx = / Eexx dexx Jo Jo 2 xx 2 \ duo d2wo\ dx dx2 J (4.38) and the strain energy becomes (4.39) For pure extension we set wq = 0, and for pure bending we set uq = 0, to obtain the strain energy expression for bars and beams, respectively, The strain energy density due to shearing by torque T is given by rSxfi Uo = 2 / crxe dex$ - I 4Gex$ dsxo = 2GexB. Jo Jo The complementary strain energy density of a linear elastic beam is pox* . r<*xz r°xB Uo = / sxx d<rxx + 2 / exz doxz + 2 / ex$ dax Jo Jo Jo (4.40) o-: 1 2 2E 2G 2G x$ (4.41)
94 WORK, ENERGY, AND VARIATIONAL CALCULUS and the complementary strain energy becomes U*= [ U$dV Jv where fs I2b2 J a Q2(z)dA. (4.43) Example 4.4 Consider the frame structure shown in Fig. 4.9. Each member of the structure has the same cross-sectional area A, moment of inertia /, and Young's modulus E. The material of the frame is assumed to be linearly elastic. We wish to compute the complementary strain energy of the frame structure. The expressions for the axial and shear forces and bending moments for the two parts of the beam are MDB = Pv • x, Mb a = PV'b + Ph-x. Vba^Piu The complementary strain energy is U* = U^B + U^A [see Eq. (4.42)] with JO -^- 4- — (P xS1 4- fxPv IE A + 2EI { ° } + 2GA 2EA + 6EI + 2GA' FT" XX y \\p„b •T M = P.,x ¥ Jr i M=Pb+P,,x Figure 4.9 The frame structure of Example 4.4.
VIRTUAL WORK 95 u*ba = r Jo + ■ 1 2EA 2E7 (Pvb + Phx)2 + 2GA Jjc P.ffl + ■ 1 2EA 2EJ fa ^+PvPha2b+lp, V) + 2GA' 4.3 VIRTUAL WORK We shall use the term configuration to mean the simultaneous position of all material points of a body. From purely geometrical considerations, a given mechanical system can take many possible configurations. The set of configurations that satisfy the geometric constraints of the system is called the set of admissible configurations. Of all admissible configurations, only one of them corresponds to the equilibrium configuration under the applied loads; it is this configuration that also satisfies Newton's second law (i.e., equilibrium of forces and moments). The admissible configurations are restricted to a neighborhood of the true configuration so that they are obtained from infinitesimal variations of the true configuration (i.e., infinitesimal movement of the material points). During such variations, the geometric constraints of the system are not violated and ail applied forces are fixed at their actual equilibrium values. When a mechanical system experiences such variations in its equilibrium configuration, it is said to undergo virtual displacements. These displacements need not have any relationship with the actual displacements that might occur due to a change in the loads and/or boundary conditions. The displacements are called virtual because they are imagined to take place (i.e., hypothetical), with the actual loads acting at their fixed values. For example, consider a beam fixed at x = 0 and subjected to any arbitrary loading (e.g., distributed as well as point loads), as shown in Fig. 4.10. The possible geometric configurations that the beam can take under the loads may be expressed in terms of the transverse deflection wq(x) and axial displacement uq(x). The support conditions require that u>o(0) — wo, /dwp L- «0, «o(0) = «o> (4.44) «„(0) = 0 t»0(0) = 0 <(0) = 0 <?W z, w„ » 1Z T V * V * I T \m T T Vf > =*fc •Ax) Figure 4.10 A cantilever beam with a set of arbitrary loads.
96 WORK, ENERGY, AND VARIATIONAL CALCULUS where wq, Oq, and wo are constants (which arc zero for the beam shown in Fig. 4.10). These are called the geometric boundary conditions. Boundary conditions that involve specifying the forces applied on the beam are called force boundary conditions. The set of all functions w(x) and u(x) that satisfy the geometric boundary conditions (4.44) is the space of admissible configurations for this case. This space consists of pairs of elements {(u>/, w/)} of the form u)\ (x) = wq -f £o-v* + aiA'2, W2(x) = wo + Oqx -f a[X2 -f tf2*3, u\(x) = uq -f b[Xy uiW ~ «0 + b\x -f b2X2, — The pair (wo, «o) that also satisfies, in addition to the geometric boundary conditions, the equilibrium equations and force boundary conditions (which require the precise nature of the applied loads) of the problem is the equilibrium solution we seek. The virtual displacements, 8wq(x) and 8uq(x), must be necessarily of the form Swq = c\x , 8uq = d[X, 8wq = c\x 4- cix~, 8uq — d\x -f dix ,..., so that they satisfy the homogeneous form of the specified geometric boundary conditions: 8w0(0) = 0, (**p\ = o, 8uQ(0) = 0. (4.45) Thus, the virtual displacements at the boundary points at which the geometric conditions are specified, independent of the specified values, are necessarily zero. The work done by the actual forces through a virtual displacement of the actual configuration is called virtual work done by actual forces. If we denote the virtual displacement by Su, the virtual work done by a constant force F is given by SW = F-8u. The virtual work done by actual forces in moving through virtual displacements in a deformable body consists of two parts: virtual work done by internal forces, 8Wi, and virtual work done by external forces, 8We- These may be computed as discussed next. Consider a deformable body of volume V and closed surface area S. Suppose that the body is subjected to a body force f(x) per unit mass (or pf per unit volume) of the body, a specified surface force t(s) per unit area on a portion #2 of the boundary, and specified displacement u on the remaining portion Si of the boundary such that S = S{ U 1S2 and S\ D S2 = 0. For this case, the virtual displacement 8u(x) is any function, small in magnitude (to keep the system in equilibrium), that satisfies the requirement 5u = 0 on S].
VIRTUAL WORK 97 The virtual work done by actual forces f and t in moving through the virtual displacement <5u is 8WE = - ( f pt • 5u dV + f t • 5u dSJ . (4.46) The negative sign in front of the expression indicates that work is performed on the body. As a result of the application of the loads, the body develops internal forces in the form of stresses. These stresses also perform work when the body is given a virtual displacement. In this study, we shall be concerned mainly with only ideal systems. An ideal system is one in which no work is dissipated by friction. We assume that the virtual displacement <5u is applied slowly from zero to its final value. Associated with the virtual displacement is the virtual strain (for the linear case), 6s = £[V0u) + (V(<5u))T], (4.47) The internal virtual work stored in the body per unit volume, analogous to the calculation of the strain energy, is the virtual strain energy density /»8 s p Se.jj 8U0 = a : d(s7) - / crud(8£ij) = av/Se,-;, (448) Jo Jo irrespective of the constitutive behavior. The total internal virtual work stored in the body is denoted by $ W(, and it is equal to SWj = f 8U0dV= f aijSsjjdV. (4.49) Jv Jv Analogous to virtual displacements, we can also think of virtual forces. If a body is imagined to be subjected to a set of self-equilibrating force system <5F, the virtual work done by the virtual forces in moving through actual displacements u is given by SW* = ,5F-u, and it is called complementary virtual work. The complementary internal virtual work and complementary external virtual work for a deformable body can be expressed, following the ideas already presented, as 8W% = -(f p8f-udV+ f 8t-uds\ (4.50) 8U$ = eijSaij, (4.51) 8Wf = f '8U$ dV = [ sij 8aif dV. (4.52) Jv Jv Note that in selecting a virtual force system, one must make sure that they are in equilibrium among themselves.
98 WORK, ENERGY, AND VARIATIONAL CALCULUS Example 4.5 Consider the rigid link in Fig. 4.3 (Example 4.1). The initial, no-load position corresponds to 0 = 0, and the displacement to the equilibrium position is small (i.e., 0 is small). We wish to deteimine the virtual work done by external forces as well as internal forces when the link is given an infinitesimal rotation 50 from its equilibrium position. The virtual work done by actual external forces is given by SWE = -(P8v + M80) = -(PL + M)8G. The work done by internal force Fs in the extensional spring and internal moment Ms in the torsional spring is given by 8Wi = Fs8v + M5S0 = k\v8v + k2080 « (k]L2 + k2)086. Example 4.6 Consider the cantilever beam shown in Fig. 4.11a. Assume virtual displacements 8uq(x) and 8wq(x) such that 5wo(0) = 0, c18wq dx (0) = 0, *w0(0) = 0. «b(0) = o w0(0) = 0 w0'(u) = 0 z, % q(x) TTTTTTTTTTTTTTTr ■* «* L ^ •yw (a) Actual geometry and forces H(0) = 0 5<(0) = 0 k 5F » 5F (b) Virtual displacements and forces Figure 4.11 The cantilever beam discussed in Example 4.6.
VIRTUAL WORK 99 Then the virtual work done by the actual forces in moving through virtual displacement u)q(x) is given by 8WE = -\f fooSwo + f*uo)dx + FSwq(L) + P8u0(L) . (4.53a) The virtual work done by internal forces is SWr = [ 8U0 dV = / au8eij dV Jv Jv = / / 0\i8s\\ dAdx = / / oxx8sxx dAdx. (4.53b) Jo J A JO J A The virtual strains can be computed from the actual strains duo d2WQ dSuQ d28wo dx dx2 ' xx dx dx2 Then the virtual work is 8Wf = / crxx — z——5- dAdx Jo J A \ dx dx1 J [L{Ad8uo A,d28wp\ ^ where N — I axx dA, M —I axxz dA. The total virtual work done is 8W = 8Wi+8WR - F8w0(L) - P8u0(L). (4.54a) Note that no constitutive law is used in arriving at the result in Eq. (4.54a), but it is understood that N and M are known in terms of the kinematic variables wo and wolf we assume linear elastic behavior crxx = Eexx> then [see Eq. (4.34)] dx dxL
100 WORK, ENERGY, AND VARIATIONAL CALCULUS In this case the total virtual work done, which will be denoted by 511 to distinguish it from 5 W, is dependent on the constitutive law. We have sn fL (.jChiodSuo d2w0d28w0 \ - F8w0(L) - PSuq(L). (4.54b) Now suppose that the cantilever beam is subject to the virtual force system shown in Fig. 4.11b. The complementary virtual work done by external forces is 8W% = - \SFwo(L) + SPuq(L) - 8Fw0(0) + (8F - L) (^p) - 8Pu0(0) | = -[8Fwo(L) + 8Puo(L)\. The complementary virtual strain energy is computed assuming that the strains are of the form Then the complementary virtual internal energy is SWf = / (en8(7n +2en5an) dV Jv =CL K*«+w")8a"+^'^dv = f (e^8N + s^8M + 2s^8v)dx. The total complementary virtual work is SW* = f (ef>8N + s^SM + le^SV) dx - [SFw0(L) + SPu0(L)]. (4.55a) Jo Once again, if we use linear constitutive relations e(0) = £ (i) = in. OfiW = -« £A' "xx EI' xz GA' Bq. (4.55a) becomes • STl*= I (^N+~M+^8v)dx-[8FwQ(L)+t>Puo{L)l (4.55b) Jo \EA El GA J
VIRTUAL WORK 101 Example 4.7 Consider the frame structure of Fig. 4.9. We wish to compute the total complementary virtual work, 8W* = 8Wf + 8Wg. Assuming a linear elastic constitutive relation, the virtual complementary strain energy can be written as 8Wf = 8lT = fL\- M f V lEA EI GA dx. (4.56) where (N,Mt V) are the axial force, bending moment, and shear force due to actual internal forces, and (8N, <5M, 8V) are the virtual axial force, virtual moment, and virtual transverse force in the structure. The stress resultants (N, M, V) in the two segments of the structure can be expressed in terms of the applied external forces, Pv and Pfu as (see Fig. 4.9) Ndb = Ph> Nba = Pv, VDB = Pu, Mdb = Pv - xy MBa = Pv-b + Ph-x. Vba^Piu (4.57a) (4.57b) If wc apply self-equilibrating virtual forces 8PV and 8Ph at point D as shown in Fig. 4 J 2, the axial force and bending moment expressions due to the virtual forces 8PV and 8Ph for the two parts of the beam are 8NDB = Sft, 8NBA = 8PV> 8VDB = 8PV> 8VBA = 8Piu (4.58a) 8MDB = 8PV • a-, SMba = SPv • * + SPh • jc, (4.58b) which can also be obtained by taking the first variation of the quantities in Eq. (4.57b). However, in general, the virtual forces need not be the same as the actual forces. p W=8P, 4^ SM = SP„x f ■8P„ V \8PJ? 8P* 6Fh^%j5Pv +8Pfla 8V=8Ph^& f8M=dPvb + 8Phx 8PV *8N=8PV Figure 4.12 A frame structure with virtual forces.
102 WORK, ENERGY, AND VARIATIONAL CALCULUS The virtual complementary strain energies due to virtual forces 8PV and SP/7 are dx s^A=f [jxSPv+jj(P»b+Phx) (SPvb+SPhx)+WtSPh]dx \ Pva \ ( ,, ,a2M„ The total virtual complementary strain energy of the beam is the sum of 8U^B and8U*A: 8W? = 8U*DB+8U*BA. (4.60) The complementary virtual work done by external virtual forces is 5W| = -(v • 8PV + u • &Ph). (4.61) We cannot express the displacements in ternis of the forces unless we know the solution to the problem. 4,4 CALCULUS OF VARIATIONS 4.4.1 The Variational Operator The delta operator 5 used in conjunction with virtual quantities has special importance in variational methods. The operator is called the variational operator because it is used to denote a variation (or change) in a given quantity. In this section, we discuss certain operational properties of 8 and elements of variational calculus. Using these tools, we can study the energy and variational principles of general problems. Let u — u(x) be the true configuration (i.e., the one corresponding to equilibrium) of a given mechanical system, and suppose that u = u on boundary S\ of the total boundary S. Then an admissible configuration is of the form u-u+av ' (4.62) everywhere in the body, where v is an arbitrary function that satisfies the homogeneous geometric boundary condition of the system v = 0 on S\. (4.63)
CALCULUS OF VARIATIONS 103 Here av is a variation of the given configuration u. It should be understood thai the variations are small enough (i.e., a is small) not to disturb the equilibrium of the system, and the variation is consistent with the geometric constraint of the system. Equation (4.62) defines a set of varied configurations; an infinite number of configurations u can be generated for a fixed v by assigning values to a. All of these configurations satisfy the specified geometric boundary conditions on boundary Si, and therefore they constitute the set of admissible configurations. For any v, all configurations reduce to the actual one when a is zero. Therefore for any fixed x, av can be viewed as a change or variation in the actual configuration u. This variation is often denoted by Su: „/du\ /dv\ d(av) dSu ,A,A^ Su=avi s( — )=a( — ) = -^ = —, (4.64) \dx/ \dx/ ax dx and Su is called Xhe first variation of w. Next, consider a function of the dependent variable u and its derivative uf = du/dx: F = F(x,u9u'). (4.65) For fixed *, the change in F associated with a variation in u (and hence uf) is AF = F(„v, u + av, uf -f av') — F(x, w, uf) = F(x, u, it ) + —av + —av ou ow ^ (av)2 d2F 2(av)(av') 32F _{ dF dF , , = —av + —av' + 0(a2), (4.66) aw ou' where 0(a2) denotes terms of order a2 and higher. The first total variation of F(x,u, uf) is defined by 8F = a lim [a-»o a J (dF = a[—v \ou BF ; dF SF , OU Oil' = —Su 4- —&u'. (4.67a) du duf
104 WORK, ENERGY, AND VARIATIONAL CALCULUS L Alternatively, the first variation may be defined as [dF(u -f av. u! -f otvf) SF = or [ da dF dF , = -—ay + ~-ay du du' 3F dF = —Su + —Su'. (4.67b) du du' There is an analogy between the first variation of F and the total differential of F. The total differential of F is dF dF dF , dF = —dx + — du + — du'. (4.68) dx du div Since x is fixed during the variation of it to it -f Su, dx — 0, and the analogy between SF in Eq. (4.67) anddF in Eq. (4.68) becomes apparent. That is, the variational operator 8 acts like a differential operator with respect to the dependent variable. Indeed, the laws of variation of sums, products, ratios, powers, and so forth are completely analogous to the coiTesponding laws of differentiation; that is, the variational calculus resembles the differential calculus. For example, if F[ = F\(u) and Fi = Fi(ii), we have (4.69) If G = Giu, v, w) is a function of several dependent variables (and possibly their derivatives), the total variation is the sum of partial variations: 8G = 8ltG + SvG + 8wG, (4.70) where, for example, 8U denotes the partial variation with respect to u. The variational operator can be interchanged with differential and integral operators: „ (du\ dv d , x d yn x (1) 8[ — )=a— = — (av) = —(Su). \dx / dx dx dx Ora \ pa pa pa ' udx) =a v dx = / av dx = / Su dx. (4.71) o / Jo Jo Jo (1) S(Fi ± F2) = SFi (2) &{FXF2) *<%) (4) S(Fi)" = ±SF2. = 8F[F2 + Fl8F2. _ SFiF2 ■■ «(F,)»- - F\8F2 n ' ]SFi.
CALCULUS OF VARIATIONS 105 For example, the expressions in Eqs. (4.54b) and (4.55b) can be expressed as 2 ort J fL \EA /dito\2 EI {d2w0\ dx -Fwq(L)-Pu0(L)\, 8U" -[T( W + M2 + /,v- 2£M 2EI 2GA J dx - Fwq(L) - Puo(L) (4.72) (4.73) All of the above discussion can be extended to two dimensions and to functions F that depend on more than one dependent variable in two or more dimensions. Let F = F{x9 y, w, v, uX9 vX9 uy9 vy)9 where u — u(x, y) and v = v(x9y) arc dependent variables, and ux = du/dx, uy = duf'dy, and so on. The first variation of F is given by dF dF dF dF 3F 0 dF 0 SF = —&u + —«v + -—8ux + -r—Svx + t— *uy + — 5uy, 9w 3u 9wA- avx oily dvy ' and 8(F\(u9 v)F2(u, v)) = 8F](u9 v)F2 + SF2(u9 v)F\ Example 4.8 Consider the following functions of dependent variables: , , * / //, 2 fdu\2 (d2u\ du (a) f («, u , u ) = c\iC + c2 I — J + C3 I —3 I + c^ll~T^ (b) G(«, V, H , U ) = Ci« + C'2V + C3«U + C4 I —■ I + C5 I — ) + <?6 ——, \dx J \dx J dx dx where c,- denote functions of x only. The first variations arc dF a^7 3F 9F , dF „ (a) SF = -Su + —*«' + —-Su" au ow du" I\ du\ / du \ /dSu\ „ drii /d 8u\ dG dG , (b) 5lrG = — *« + v7 8u du duf = (2C1 W + C3V)bu + 2c4 — + Cfi —■ ) -— \ dx dx J \ dx
106 WORK, ENERGY, AND VARIATIONAL CALCULUS v „ dGc 3G. %G = —8v+—Sv/ av avf / dv du\ /dSv\ = (2c2v + c3u)8v + 2c5— + c6— — . \ dx ax) \ ax ) 4.4.2 Functionals In the study of variational formulations of continuum problems, wc encounter functions of dependent variables that are themselves functions of other parameters, such as position, time, and so on. In particular, integral expressions of dependent variables and their derivatives are of interest. Such integral expressions are termed functionals. A formal mathematical definition of a functional requires concepts from functional analysis. Here we attempt to present the formal definition without a detailed review of functional analysis, as it is outside the scope of the present study. A functional / is a mapping (or operator) from a vector space U into the real number field 9?. Thus, if u e U, then /(«) is a real number: I :U -+ 9i. (4.74) Note that /(•) is an operator and I (it) is a functional. For example, the integral expression I(u)= / \au(x)+ buf(x) +cuf(x)\dx, it = Jo du „ _ d2u dx ' dx2' qualifies as a functional for all integrable and square-integrable functions u(x) with their first and second derivatives. Note that the value of the functional / (u) depends on the choice of u. A functional is said to be linear if I (an + pv) = al(u) + pi(v) (4.75) for allaf, y3 e ))i and dependent variables u and v. A quadratic functional is one which satisfies the relation I(au) = a2I(u) (4.76) for all a € 9? and the dependent variable a. 4.4.3 The First Variation of a Functional The first variation of a functional of u (and its derivatives) can be calculated as follows. Let /(«) denote the integral defined in the interval (a, b): fh du I (it) = / F(jc, k, u')dx, u = —, (4.77) Ja dx
CALCULUS OF VARIATIONS 107 where F is a function, in general, of x, u9 and du/dx. When the limits a and b are independent of w, the first variation of the functional / (m) is fb rb rb /flf dF \ 8I(u)=8l F(x.u,uf)dx = / SFdx= I l—8u+—8u')dx. (4.78) Ja Ja Ja \ <*" 3" / When the limits of integration depend on the dependent variable, the Leibniz rule is used to calculate the first variation. Suppose that the boundary points a and b depend on w.Then 8T(u) = oc — I F{x, u + av, u + av1) dx <*<* Ja<u+av) tf=0 Using the Leibniz rule, we obtain fb 8I(u)= / 8F(x,u,u')dx + F(b,u(b)iu\b))8b-F(aMa),u\a))8a- (4.79) Ja Thus the variational operator can be used to compute the variation of any functional. Example 4.9 Wc wish to determine the first variation of the following functionals: ~L [n /sJu\2 h - 1 dx. (b) (a) We have , x r, x f \a (CIU\ b 2 r (a) m=L [2U) +r-ju_ I(u) = I I -Vw • Vtf — fit I dx — / qu ds 5/(w) = / a 1- bwou — /ow I dx Jo V "■* d* / (b) In this case, we have 8I(u) = / (Vm • V8u - f8u)dx- q8u ds Jo. Jr2 4.4.4 Fundamental Lemma of Variational Calculus The fundamental lemma of calculus of variations is useful in obtaining differential equations from integral statements. The lemma can be stated as follows: for arty integrable function G, if the statement f Ja G • 7] dx = 0 (4.80)
108 WORK, ENERGY, AND VARIATIONAL CALCULUS holds for any arbitrary continuous function n(x), for all x in (a, b), then it follows that G — 0 in (a, b). A mathematical proof of the lemma can be found in most books on variational calculus. A simple proof of the lemma follows. Since r\ is arbitrary, it can be replaced by G. We have L h G2 dx = 0. Since an integral of a positive function is positive, the above statement implies that G = 0. A more general statement of the fundamental lemma is as follows: Tf rj is arbitrary in a < x < b and n(a) is arbitrary, then the statement I b Gi] dx + B(a)i](a) = 0 (4.81a) implies that C = 0 in a <x < b and B(a) = 0, (4.8Jb) because ij(x) is independent of i](a). 4.4,5 Extremum of a Functional From elementary calculus it is known that a differentiate function /(*), a < x < b, has an extremum (i.e., minimum or maximum) at a point A'o in the interval (a, b) only if (necessary condition) dj_ dx The sufficient condition for maximum (or minimum) is that = 0. (4.82) .v=.r0 d2f A (_d2f dx2 <0 (or^><)). (4.83) Similarly, for a differential function f(x, y) in two dimensions, the necessary condition for an extremum at the point (xq, );o) is that its total differential be zero at U'o,J?o): df df df = ~- dx + -J-dy=0 at x = xq and y = y0. (4.84) dx dy If x and v are linearly independent (i.e., x and y are not constrained), Eq. (4.84) implies ' df df — = 0 and tt- = 0 (4.85) dx dy at x = xq and y = yo.
CALCULUS OF VARIATIONS 109 In the study of problems by variational principles and methods, we seek the extremum of integrals of functions of functions, F = F(x, u(x), uf(x))—or functional. Consider a simple, but typical, variational problem. Find a function u = u(x) such that u(a) = ua and u(h) = lib, and fb I(u)= I F(x,u(x),uf(x))dx Ja (4.86) is a minimum. In analyzing the problem we are not interested in all functions w, but only in those functions that satisfy the stated boundary (or end) conditions. The set of all such functions is called, for obvious reasons, the set of competing functions (or set of admissible functions). We shall denote the set by C. The problem is to seek an element it from C which renders / a minimum. If u e C, then (w + a v) eC for every v satisfying the conditions v(a) = v(b) — 0. The space of all such elements is called the space of admissible variations, as already mentioned. Figure 4.13 shows a typical competing function u(x) = a(x) + av(x) and a typical admissible variation v(x). Let I (it) be a differentiate functional in the sense that dl(u ~\-av, uf -f-ai/) da exists, and let C denote the space of competing functions. Then an element u € C is said to yield a relative minimum, (maximum) for I(u) in C if I(u)-I(u)>0 (<()). (4.87) If I (ft) assumes a relative minimum (maximum) at w e C relative to elements ii G C, then it follows from the definition of the space of admissible variations U and Eq. (4.87) that l(u+av)-I(u) >0 (<0) (4.88) uQc) + aK*) Figure 4.13 The variations of u(x).
110 WORK, ENERGY, AND VARIATIONAL CALCULUS for all v e Hy \\v\\ < e> and a e W. Since u is the minimizer, any other function u e C is of the form a = u H- av, and the actual minimizer is determined by setting a = 0. Once u(x) and v(x) are assigned, l{u) is a function of a alone, say /(a). Now a necessary condition for /(«) = /(a) to attain a minimum is that ^M = -£-[/(«+a W)]=0. (4.89) On the other hand, / (it) attains its minimum at w, that is, a ~ 0. These two conditions together imply (dI(<x)/da)\a=Q = 0, which is nothing but */(«) = 0. (4.90) Analogous to the second necessary condition for ordinary functions, the second necessary condition for a functional to assume a relative minimum (maximum) is that the second variation 82l(u) is greater (less) than zero. The second variation S2I(u) of a functional I (a) is given by d2 r/(«0 = y|_ for all v e C and a € d\. ^M^[^nu+aV\ (4.91) a=0 4.4.6 Euler Equations We now return to the problem of determining the minimum of a functional: rb l(u) = / F(x,u,it)dxi (4.92) Ja subject to the end conditions u(a) = uat u(b)^ab. (4.93) It is clear that any candidate for the minimizer of the functional should satisfy the end conditions in Eq. (4.93), and be sufficiently differentiate (twice in the present case, as we shall see shortly). The set of all such functions is the set of admissible functions or competing functions for the present case. Functions from the admissible set can be viewed as smooth (i.e., differentiable twice) functions passing through points (a, ua) and (by m/7), as shown in Fig. 4.13, Clearly, any element u eC (the set of competing functions) has the form u = u+avy (4.94) j where a is a small number and v is a sufficiently differentiable function that satisfies the homogeneous form of the end conditions (because u must satisfy the specified end conditions) in Eq, (4,93) u(fl) = u(fc) = 0f (4.95)
CALCULUS OF VARIATIONS 111 and u is the function that minimizes the functional in Eq. (4.92). The set of all functions v is the set of admissible variations, H. Now assuming that, for each admissible function u, F(x,u, uf) exists and is continuously differentiable with respect to its arguments, and / (w) takes one and only one real value, we seek the particular function u(x) that makes the integral a minimum. The necessary condition (4.89) for / to attain a minimum gives dl(u + av) da I = \-r / F(xyujtf)dx\ iev=o Ida Ja J^o rfdFdu dFdu'\\ , Cb (dF dF ,\ ,An^ = / ^T~ + rr:— dx = / — v + —v' dx, (4.96) Ja \ du da du' da J lor==0 Ja \ au aw J where it = u +av. Integrating the second term in the last equation by parts to transfer differentiation from vtou, we obtain n fb [dF d (dF\\ , (dF \ (4.97) The boundary term vanished because v is zero at x = a and x = b [see Eq. (4.95)]. The fact that v is arbitrary inside the internal (a, b) and yet the equation should hold implies, by the fundamental lemma of the calculus of variations, that the expression in the square brackets is zero identically: (£)- [ — I = 0 in a < x < b. (4.98) du dx l *"' ' Equation (4.98) is called the Eider equation of the functional in Eq. (4.92). Of all the admissible functions, the one that satisfies (i.e., the solution of) Eq. (4.98) is the tine minimizer of the functional /. Next, consider the problem of finding (w, v), defined on a two-dimensional region £2, such that the following functional is to be minimized: I(u, v) = I F(xt y,u9vt uXy vx, uy, vy) dxdy (4.99) where ux = dufdx, uy = du/dy, and so on. For the moment we assume that u and v are specified on the boundary r of Q. The vanishing of the first variation of / («, v) is written as 81 (u, v) = 8ltF(u, v) + 8vI(u, v) = 0. Here 8U and 8V denote (partial) variations with respect to u and i?, respectively. We have [see also Eq. (4.72)] f faFo dF dF dF dF0 dF 1 7 , 81 = I \ -—8u H 8ux H 8uy + —- Su + -—8vx + -—8vy \ dxdy, Jn [ du aux diiy dv dvx dvy ' J (4.100)
112 WORK, ENERGY, AND VARIATIONAL CALCULUS The next step in the development involves the use of integration by parts, or the gradient theorem on the second, third, fifth, and sixth terms in Eq. (4.100). Consider the second term. We have f dF 8Su , , f r 9 /9F \ 3 /3F\ 1 , , L ^i7dxdy=L b U;'v ~ Tx wJ 8u\dxdy r dF r a /3F\ = <A -r—Sunx ds- I — I -— ) 5« dxrfy. (4.101) /r 9k* Jq 9* \3«j/ Using a similar procedure on the other terms and collecting the coefficients of 5m and Sv separately, we obtain 0=[\\9*±(*L)-°-(°L]\8u JQ [I3u dx V dux) By \duyj] Law 9x\dvxJ dy\dvyjj J > f[/dF dF \ /dF dF \ 1 + % Lfe"*+ 5^"'j'"+ fe"*+ VJ 8v\ds- (4102) Since (u, v) are specified on r, 8u = 8v = 0 and the boundary expressions vanish. Then, since 8u and Sv are arbitrary and independent of each other in ft, the fundamental lemma yields the Euler equations 'dF 3 /3F\ 3 /3F\ 8u: -5--^- 3— )"ir(^—) = °» (4-103a) 9 m 8x \3ux/ oy \3iiyJ 3F 3 /3F\ 3 /3F\ *v- T" " T" ( TT" I " T" ( ^- I = 0- (4.103b) 9v 9x \9ujr/ 9y \3vyJ All of the foregoing discussion can be applied to the general case of a functional involving p dependent variables with /nth-order partial derivatives with respect to n independent variables. In this case there will be p Euler equations involving 2mth- order derivatives in n independent variables. 4.4.7 Natural and Essential Boundary Conditions First consider the problem of minimizing the functional in Eq. (4.92) subject to no end conditions (hence 8v ^ 0 at x = a and x = b). The necessary condition for / to attain a minimum yields, as before, Eq. (4.97). Now suppose that 3F/dul and v are selected such that t 3F —-v = 0 for x = a and x = b. (4.104) aw Then using the fundamental lemma of the calculus of variations, we obtain the Euler equation in Eq. (4.98).
CALCULUS OF VARIATIONS 113 Equation (4.104) is satisfied identically for any of the following combinations: (i) v(a) = 0, v(b) = 0. (ii) v(a) = 0. ^(« = 0. 3F (4.105) (iii) — (a)=0, v(b) = 0. dur (iv) ^(<i) = 0, j£(fr)=0. air air The requirement that v — 0 at an end point is equivalent to the requirement that u is specified (to be some value) at that point. The end conditions in (4.105) are classified into two types: essential boundary conditions, which require the variation of v and possibly its derivatives to vanish at the boundary, and natural boundary conditions, which require the specification of the coefficients of the variations of v and its derivatives. Thus we have Essential Boundary Conditions specify v = 0 or u = it on the boundary. (4.106a) Natural Boundary Conditions 8F —- = 0 on the boundary. (4.106b) dir In a given problem, only one of the four combinations given in Eq. (4.105) can be specified. Problems in which all of the boundary conditions are of essential type are called Dirichlet boundary-value problems, and those in which all of the boundary conditions are of natural type are called Neumann boundaiy-value problems. Mixed boundaiy-value problems arc those in which both essential and natural boundary conditions arc specified. Essential boundary conditions are also known as Dirichlet or geometric boundary conditions, and natural boundary conditions are known as Neumann or dynamic boundaiy conditions. As a general rule, the vanishing of the variation v (or Sit)—equivalently, the specification of u—on the boundary constitutes the essential boundary condition, and vanishing of the coefficient of the variation constitutes the natural boundary condition. This rule applies to any functional in one, two, and three dimensions, and integrands that are functions of one or more dependent variables and their derivatives of any order. Next consider the functional in Eq. (4.99), and suppose that («, v) are arbitrary on T for the moment. It is easy to identify the natural and essential boundary conditions of the problem from Eq. (4.102): in each of the pairings on boundary I\ specifying the first element (which contains no variations of the dependent variables) constitutes the natural boundary condition, and vanishing of the second element (or, equivalently,
114 WORK, ENERGY, AND VARIATIONAL CALCULUS specifying the quantity in front of the variational operator) constitutes the essential boundary condition. Thus we have either u = it (specified) so that Su = 0 on I\ (4,107a) v = v (specified) so that Sv = 0 on T, (4.107b) or dF dF —nx + —ny =0 on T, (4,108a) aux oily dF dF nK + ny = 0 onT. (4.108b) dvx dvy Equations (4J07a,b) represent the essentia) boundary conditions and Eqs. (4.108a,b) the natural boundary conditions. The pair of elements («, v) are called the primary variables and SF dF J „ dF dF Qx s —-nx + —ny and Qy = —nx + —ny 0UX dlly OVjc dVy are called the secondary variables. Thus, specification of the primary variables constitute essential boundary conditions and specification of the secondary variables constitute natural boundary conditions. In general, one element of the each pair (u, Qx) and (v, Qy) (but not both elements of the same pair) may be specified at any point of the boundary. Thus there are four possible combinations of natural and essential boundary conditions for the problem under discussion. Example 4.10 Consider an elastic bar of length L, modulus of elasticity E, and area of cross section A. Assume that it is fixed at the left end and subjected to distributed axial load f(x) and point load P at the right end. The total potential energy functional n = U + V of the bar is [see Eq. (4.39)1 „, x fL\EA (du\2 . " dx-Pu(L), (4.109) where it denotes the axial displacement of the bar. The first term in TT(w) represents the strain energy U stored in the bar, and the second and third terms denote the work done V on the bar by the distributed load / and point load P. We wish to determine the Eulcr equation of the bar by requiring that 11 (it) be a minimum subject to the geometric boundary condition w(0) = 0. As we shall see later, this, is known as the principle of minimum total potential energy. The first variation of 11 is given by (see Example 4,9) Sn(ii) = f (eA^^- - fSu) dx - P8u(L). J0 \ dx dx )
CALCULUS OF VARIATIONS 115 where Su is arbitrary in 0 < x < L and at x = L, but satisfies the condition Su(0) = 0. To use the fundamental lemma of variational calculus, we must relieve Sit of any differentiation. Integrating the first term by parts, we get du ~|L EA—Su\ -PSu(L) <** Jo \ dxJx=Q The last term is zero because 8u(0) = 0. Setting the coefficients of Su in (0, L) and £« at x = L to zero separately, we obtain the Euler equation and the natural (or force) boundary condition of the problem: Euler Equation -— (eA—\ - / = 0, 0 < jc < L. (4110a) Natural Boundary Condition EA—-P=Q at* = L. (4.110b) Thus, the solution u of Eqs. (4.110a,b) that satisfies w(0) = 0 is the minimizer of the energy functional 11(a) in Eq. (4.109), Equation (4.110a) can be obtained directly from Eqs. (4.98) by substituting . EA(du\2 , dF r <)F ^clu Example 4.11 The total potential energy of a linear elastic body in three dimensions, subjected to body force f (measured per unit volume) and surface traction t on portion S2 of the surface, is given by (summation on repeated indices is implied) n(u) = \ U<rueu - fiiii^j dV - J fa dS> (4.111a) where a, denote the displacement components. It is assumed that the body is subjected to specified displacements on the remaining portion S\ of the surface, and therefore the virtual displacements vanish there: ii\ = Hi and Su,- = 0 on S\, (4.111b)
116 WORK, ENERGY, AND VARIATIONAL CALCULUS The first term under the volume integral represents the strain energy density of the elastic body, the second term represents the work done by the body force f, and the surface integral denotes the work done by the specified traction t. For an isotropic body the stress-strain relations are given by ai} =2iieij+k8ijekkt (4.112) where \x and X are the Lame constants. Hence, Gij eij = 2/xeij eij + ken ekk, (4. J13a) where in Eq. (4.113a) and in the following discussion, sum on repeated indices is assumed. The linear strain-displacement relations of the linear theory are given by «// = 2{UiJ+Ujj)' (4.113b) where w/>7- = (duj/dxj). Substituting Eqs. (4.1 J3a,b) into Eq. (4.111a), we obtain n(u) = / j(uij + ujjXuij + itjj) + -UjjitM ~ fitti dV - / t;iti dS. (4.114) Now we wish to derive the Euler equations associated with the functional in Eq. (4. i 14) using the requirement <5I1 = 0 and S«/ = 0 on S\. Setting the first variation of II to zero, we obtain 0 = / [-(««/,; + SiijjHujj + ujj) + kSuijUM - fiSui^ dV - f tjiujdS, (4.115) wherein the product rule of variation is used and similar terms are combined. Using integration by parts, i / 8ujj(ttij +Uj,i)dV = - / Sui(Uij +Ujj)jdV Jv Jv H- <i Siti (uij -h «y,i)n/ dS9
CALCULUS OF VARIATIONS 117 where nj denotes the jth direction cosine of the unit normal to the surface, we obtain 0 = / [—(w/.y + Ujj)jSUf - -r-("/,j + Ujj)j8uj - *■«*,*/5"/ ~ fiS^i] dV + f "xfaij + Ujj)(nj8ui + tiiSuj) + XukjniSuj dS - j StiiU dS = / [-/*("/, / + " /,/)./ ~ XukM ~ fi] Siti dV Jv <b [ti(uij + ujj) + Xiikj^ij] njSui dS - / Sujti dS. (4.116) + In arriving at the last step, a change of dummy indices is made to combine terms. Recognizing that the expression inside the square brackets of the closed surface integral is nothing but <r;/ and <Jijiij = f,- by Cauchy's formula, we can write <p tiSuj dS= I tjSuj dS+ tjSiti dS = / f/5w/ dS. Js Jsi Js2 Js2 The integral over S\ is zero by virtue of Eq. (4.111 b). Hence, we have 0 = / [—AtC«/./ + «/,/),/ - A,«*.*/ - fi] Sui dV + I 8itj(tj - tj) dS. JV ... j^ Using the fundamental lemma of calculus of variations, we set the coefficients of 5m,- in V and Su, on S2 to zero separately and obtain fi(uijj + ujjj) + Xituxi + fi = ° in V, (4.117a) r,- — f>- = 0 onS2, (4.117b) for * = 1,2,3. Equations (4.117a) are the well-known Navier's equations of elasticity. 4.4.8 Minimization of Functionals with Equality Constraints Tn the preceding sections, we devoted our attention to the determination of a function that minimizes a given functional. The necessary condition leads to a differential equation (the Eulcr equation) and (natural) boundary conditions governing the function. It turns out that it is also a sufficient condition for linear problems, not requiring the evaluation of the second variation of the functional. In the present section, we consider ways to determine the minimum of quadratic functionals (i.e., functionals that involve quadratic terms of the dependent variables) with linear equality constraints.
118 WORK, ENERGY, AND VARIATIONAL CALCULUS Consider the problem of finding the minimum of a functional (assumed to be quadratic in u and v), Cb I(u,v)= / F(x,u,u'>v,v)dx (4.118) Jo subject to the constraint G(utu'tv,v') = Q. (4.119) Tn addition, u and v should satisfy, as before, certain end conditions (or essential boundary conditions) of the problem. In the constrained minimization problems, the admissible functions should not only satisfy the specified end conditions and be sufficiently continuous, but they should satisfy the constraint conditions. Furthermore, the admissible variations should be such that the constraint conditions are not violated. Here we consider two separate methods of including the constraints in the (modified) functional, Lagrange Multiplier Method The necessary condition for the minimum of I(u, v) in Eq. (4.118) is 81 = 0. We have fb /dF dF dF dF \ 0 = 81= (^Su + ^Su' + ^Sv+^-sAdx. (4.120) }a \du du( dv dv1 } Since u and v must satisfy the constraint condition (4.119), the variations 8u and <5u are related by dG dG , 3G 3G , 0 = 8G = ~^8u + —8u' + —8v + —8v'. (4.121) du dw dv ov' The Lagrange multiplier method consists of multiplying Eq. (4.121) with an arbitrary parameter X, integrating over the interval (a, b), and adding the result to Eq. (4.120). The multiplier X is called the Lagrange multiplier. We have -ns dF , 6F dF. Su + —8u' + — 8v + —Si/ duf dv dvf
CALCULUS OF VARIATIONS 119 The boundary terms vanish because 8v(a) = 8v(b) = 8u{a) = 8u(b) = 0. Since the variations 8u and 8v are not both independent, we cannot conclude that their coefficients are zero. Suppose the 8u is independent and 8v is related to 8u by Eq. (4.121). We choose X such that the coefficient of 8v is zero. Then by the fundamental lemma of variational calculus, it follows that (because 8u is arbitrary) the coefficient of 8u is also zero. Thus we have dlt a (F + XG)-^ IA_{F + XG)1 = o, (4.123a) (F + W)-4~ \tti(F + *G)1 = 0. (4.123b) ax \_dv/ J Equations (4.123a,b) and (4.119) furnish three equations for the determination of u, v, andX. It is clear from Eq. (4.122) that Lagrange's method can be viewed as one of determining u.v, and X by setting the first variation of the modified functional L(w, v, X) = /(«, v) -h / XG(uy u\ v, vf)dx Ja = f (F + XG)dx (4.124) Ja to zero. The Euler equations of the functional are precisely Eqs. (4.123a,b) and (4.119). Indeed, we have 0 = 8L= f Ja b 8(F + XG)dx = f (8F + 8XG+X8G)dx Ja fb[(3F dG\ „ (dF 9G\ „ . = / I — + *■ U« + I + X ) Su' }a L\a« duj \du' du'J (dF 8G\ c (dF 8G\ , , „, ~l J a IL^W dlt dx \duf duf JJ [dF dF d /dF 3G\1 ) [^+x^-rA^+k^)r+Gsxh (4125) + from which we obtain Eqs. (4.123a,b) and (4.119) by setting, respectively, the coefficients of 8u and 8v9 and 8X, to zero.
120 WORK, ENERGY, AND VARIATIONAL CALCULUS Penalty Function Method The penalty function method involves the reduction of conditional extremum problems to extremum problems without constraints by the introduction of a penalty function associated with the constraints. As applied to the problem in (4.118) and (4.119), the technique involves seeking the minimum of a modified functional obtained by adding a quadratic term associated with the constraint in (4.119) (i.e., the constraint is approximately satisfied in the least-squares sense): P(w, v) = /(if, v) + £ f [G(u, u\ u, v')]2dx, (4.126) where y is the penalty parameter (a preassigned positive parameter), Setting the first variation of P to zero, we obtain the Euler equations for the functional: dF d fdF\ [BG d /„dGY] „ ,,,^x li - TX (m) + * [G^ - Tx \Gm)\ = °' (4,27a) dF d /dF\ I" 3G d /„9G\1 „ ,Atnnuo ^-d-x\M)+Y[G^-^ (C^)J = °- (412?b) For successively large values of y, the solution of Eq. (4.127a,b) gets closer to the true solution. We note that in the penalty function method the constraint is satisfied only approximately, and that no additional unknowns are introduced into (he variational formulation. Further, in the penalty function method an approximation to the Lagrange multiplier can be computed from the solution (u(y), v(y)) of Eqs. (4.127a,b) by the formula [compare Eqs. (4.127a,b) with Eq. (4.123a,b)] \ = yG(u,u',v,v'). (4.128) Both the Lagrange multiplier method and the penalty function method can.be generalized for functional with many variables and constraints. Note that when the constraint equation is a relationship involving constants, the Lagrange multiplier is also a constant. For example, fb C[it - C2V = C3, / F(x, u(x), it (x))dx = C'i, Ja where ct (/' = 1, 2, 3) are constants, are constraint equations among numbers (note that a definite integral of a function is a number). Example 4.12 Given the problem of minimizing the functional 1 f'7 / = - / (xy-yx)dt, , (a) subject to the constraint /,= Ai2 + J2)l/2*, (b) Jf]
CALCULUS OF VARIATIONS 121 where t is a parameter, x — dxjdt, y = dy/dt, L is a constant, and x and y are dependent variables, we wish to determine the Euler equations using the Lagrange multiplier method. The constraint condition can be viewed as an algebraic one in the sense that the integral quantity is a number. Hence, the Lagrange multiplier is also a number. Thus, we have II = j / 2 &y ~ y*) dt + kfj y/x2 + y2dt - Lj — I F(x, y, A., x, y) dt, Jti where X is the Lagrange multiplier, independent of t. The two Euler equations, in addition to Eq. (b), are given by 8F d (9F\ n Sx: [ — ) = 0, dx dt\dx) dF d /dF\ n Sy: — =0, } By dt \ dy J (c) or Sx: Sy: 2y dt 2X It 1 Xx —y + -?= 2 s/x2 + y2 1 -x + Xy ^T. = 0, = 0. (d) (e) Example 4.13 Consider the problem of minimizing the functional subject to the constraint fL EI (d<j>\ dx — Fwq(L), G(u^0)sf^+0 = O. (4.129) (4.130) The functional represents the total potential energy of a cantilever beam subjected to uniformly distributed load q and point load F at its free end. The constraint in Eq. (4.130) represents the relation between the transverse deflection wo and rotation (j> of a transverse normal line. The essential boundary conditions for the problem are (see Fig. 4,10 for the geometry and coordinates) wo(O) = 0(O) = O. (4.13!)
122 WORK, ENERGY, AND VARIATIONAL CALCULUS In the Lagrange multiplier method, the modified functional for the problem is given by nL{wo94>,V = n(wo,4>)+ f xr^ + Adx, (4.132) where X(x) denotes the Lagrange multiplier. The Euler equations of the problem are obtained by setting the partial variations of YIl with respect to ti>o, 0, and X to zero: dX 8wYlL = 0: -f <? = 0. 0 <x <L. (4.133a) ax X-F = 0t at* = L. (4.133b) SaIIl = 0: -4" (EI^f) + ^ = 0, 0 < * < L. (4.134a) ax \ ax J E/^=0, at* = Z,. (4.134b) ax Sxnf = 0: ^+0 = 0, 0<;t<L. (4.135) ax The Lagrange multiplier X(x) in the present case turns out to be the shear force, V(x). In the penalty function method, the modified functional is given by np(wi<p) = n(w0,<i>) + ^f (^7 + ^) dx- <4"-136> The Euler equations are given by 8wYlp =0: -yj- (^p- + <p J + a = 0, 0 < x < L. (4.137a) y (— + 0 J - F = 0, at * = L. (4.137b) a#npS=0: ^±(eI^+Y{^+^)=09 0<x<L. (4.138a) EI—=0t aUc=L. (4.138b) a* t A comparison of Eq. (4.133b) with Eq. (4.137b) shows that the Lagrange multiplier is related to 0 and u)q by KX) = y(^ + 4>\ (4.139)
EXERCISES 123 For the choice of y = KSGA, Eq. (4.139) represents the shear force and shear- strain relation in the Timoshenko beam theory (see Exercise 4.24): X(x) ~ V(x) = KsGAyxz = K*GA U + -p J , (4.140) where Ks is the shear correction factor, G the shear modulus, and A the area of cross section of the beam. EXERCISES 4.1 Consider the double pendulum shown in Fig. E4.1 at some instant of time. Write the potential energy (only due to gravity) of the system assuming that the energy is zero when 0A = 02 = 0. Neglect the weight of the pendulum bars and assume that they are rigid. Figure E4.1 4.2 Consider the rigid-body assemblage shown in Fig. E4.2 at some instant of time. The bars are interconnected by hinges, and their relative rotations are resisted by torsional springs located at each hinge. Write the potential energy of the system assuming that the energy is zero when 0\ — 02 = 03 = 0. Include the weight W of the rigid bars and assume that the displacements are small and the torsional springs are linear. / H »► X " IV r Figure E4.2
124 WORK, ENERGY, AND VARIATIONAL CALCULUS 4.3 The rigid link shown in Fig. E4.3 is in equilibrium when pinned at one end and subjected to forces Fx and Fy and moment M as shown. Find the incremental work done, AW*, when the link is subjected to increments AFXi AFy, and AM. Express the answer only in terms of the increments AFA- and AFy. u,Fx Figure E4.3 4.4 Determine the strain energy and external work done for the bar with an end linear elastic spring shown in Fig. E4.4. E, A RiEid bar ■a frV^-w k Figure E4.4 4.5 Determine the strain energy and external work done for the beam with an end linear elastic spring shown in Fig. E4.5. Neglect the energy due to shear. Vi\ ^lllllll p: z, w0 E,I Figure E4.5 4.6 Determine the total complementary strain energy due to bending as well as shear for the beam shown in Fig. E4.6. 4.7 Repeat Exercise 4.5 when the spring behaves according to the relation Fs = IcWq, wq being the elongation of the spring. 4.8 Determine the complementary strain energy of the structure shown in Fig. E4.8. Include energies due to bending as well as torsion but neglect that due to transverse shear forces.
EXERCISES 125 A L BV ■ter Y L c —^KT i^ 2, w0 F0= 12 kN, L = 1 m, £ = 205GPa,v=0,3, A =1.25xl0"6m2, I =0.260 4x10*6m4 Figure E4.6 <7o (N/m) />(N) Figure E4.8 4.9 Consider the pin-connected structure shown in Fig. E4.9. Suppose that the material of all members obeys the following stress-strain relation: Ky/e% £ > 0, -Kyf^e, e < 0, where K is a constant. All members have the same cross-sectional area, A. Write the complementary strain energy densities of the members, and the total complementary strain energy of the structure. q For all members E»A (a and/? in ft.) Figure E4.9 4.10 Determine (a) the strain energy and (b) the complementary strain energy of the truss shown in Fig. E4.10. Assume clastic behavior of the form a = E^fe. Express your answer in terms of displacements in the former case and applied
126 WORK, ENERGY, AND VARIATIONAL CALCULUS loads in the latter case. Hint: You may have to use a displacement constraint to determine the reactions in the truss members. Figure E4.10 4.11 Identify admissible virtual displacements for the problem in Exercise 4.5. 4.12 Identify admissible virtual displacements for the problem in Exercise 4.8. 4.13 Identify admissible virtual forces for the problem in Exercise 4.4. 4.14 Identify admissible virtual forces for the problem in Exercise 4.6. 4.15 Identify admissible virtual forces for the problem in Exercise 4.9. 4.16 Write the total virtual work expression for the problem in Exercise 4.4. 4.17 Write the total virtual work expression for the problem in Exercise 4.5. 4.18 Write the total complementary virtual work expression for the problem in Exercise 4,3. 4.19 Write the total complementary virtual work expression for the problem in Exercise 4.4. 4.20 Write the total complementary virtual work expression for the problem in Exercise 4.5. 4.21 Write the total complementary virtual work expression for the problem in Exercise 4.8. 4.22 Write the total complementary virtual work expression for the problem in Exercise 4.9. 4.23 Two rigid links are pin-connected at point A and to a rigid support at point O as shown in Fig. E4.23. Compute the virtual work done by the forces due to the virtual changes 80\ and 8O2 from the equilibrium configuration shown in the figure. 4.24 A rigid bar supported by three linear elastic springs and subjected to load P and moment M occupies the equilibrium configuration shown in Fig. E4.24. Write the total complementary virtual work done, 8W*> and write the equilibrium equations that the virtual forces 8P, SM, and 8F3 (virtual force in spring £3) must satisfy.
EXERCISES 127 Figure £4.23 Figure E4.24 4.25 Write the virtual strain energy expression for the Euler-Bernoulli beam theory with the following nonlinear strain-displacement relation: duo 1 fdwo\ d2wo dx 2\dxJ dx2 4.26 The Tlmoshenko beam theory is based on the displacement field u(x9z) = Z((>(x), (a) w(x,z) = woOOi where wo is the transverse deflection of a point on the midplane (i.e., z = 0) of the plate and <p is the rotation of a transverse normal line about the )>-axis. (i) Compute linear strains using the displacement field (a). (ii) Write the expression for the total virtual work done by actual forces in moving through the virtual displacements (8wo, 50), assuming that the beam is loaded with a distributed transverse force q(x). Express the internal virtual work in terms of the stress resultants M = [ z<rxx dA, V = [ axz dA. (b) J A J A
128 WORK, ENERGY, AND VARIATIONAL CALCULUS 4.27 The classical theojy of circular plates is based on the displacement field uz(r,z) = woO"), (a) where wq is the transverse deflection of a point on the midplane (i.e., z = 0) of the plate, the /'-coordinate is taken radially outward from the center of the plate, the ^-coordinate along the thickness (or height) of the plate, and the ^-coordinate is taken along a circumference of the plate (see Fig. E4.27). In a general case where applied loads and geometric boundary conditions are not axisymmetric, the displacements (uri no, uz) along the coordinates (/*, 0, z) are functions of /\ 0, and ^-coordinates. Assume that the applied loads and boundary conditions are independent of the ^-coordinate (i.e., axisymmetric) so that the displacement uq is identically zero and («r, uz) are only functions of r andz. MrB Nre $■> JVrt Mro Figure E4.27 (i) Compute linear strains using the displacement field in (a). (ii) Write (he expression for the total virtual work done, S W = 8 Wj + 8 We, assuming that the plate is subjected to transversely distributed load a = q(r). Express the internal virtual work in terms of the moments (see Fig. E4.27) ' fh/2 Mrr = / < J-h/2 OyyZ dz, fh/2 M$e = I ooozdz, (b) J-h/2 where ary and cfqq are the radial and circumferential stresses, respectively.
EXERCISES 129 Calculate the first variation of the functional in Exercises 4.28-4.36. The following notation is used: u = du dx' it == d2u dx^' ttr S du uu = dxj All variables other than those listed in the argument of the functional are assumed to be functions of position. 4.28 7 (a) = /* Vl+(w')2 dx. 4.29 I {u) = /* uj\ +{u')2dx. 4.30 I(u) = /0l y/il+iuW/udx. 4.31 /(«) = ± /*{(u")2 + 2w V + (a')2 - 2m} Jjc. 4.32 n(«o.wo) Jo E4 djc 2 \ dx / Ei(d*m V " 2 V ^'2 / -f<?wo -Ptt0(L)-Af0—2(L). >d;c The boundary conditions are: «o(0) = 0, wo(0) = 0, and —— = 0. V dx J0 4.33 /(«, v) = 5 j^(u2x + 2^-Uv + 2wyvy + v2 + 2/w + 2$v) djtdy. 4.34 /(w,v)=/ I—v-v + v-Vw +Qwldx- / gw d>s\ 4.35 i(»i, «2* f) = / J -r(uij 4- Ujj)uij — Piiij \ dx\dx2 — I f\U\ dx\dx2 — f ffUf ds. Js2 Uj = 0 on the boundary Si and Si U 52 = S, the total boundary. xdxdy -f ^ /^ /c?^2 d;ef/;y — Jfi gw dxdy. w = 0 and dw/dn = 0 on the boundary of the domain £2. Express the next three problems in mathematical terms as one of minimizing appropriate quantities, subjected to certain end conditions, and possibly some constraints. 4.37 The Brachistochwne problem. Determine the curve y = y(x) connecting two points, A: (0,0) and B: (x/,, v^), in a vertical plane such that a material particle, sliding without friction under its own weight, travels from point A to point B along the curve in the shortest time (see Fig. E4.37).
130 WORK, ENERGY, AND VARIATIONAL CALCULUS Figure E4.37 4.38 Geodesic problem. Find the curve y = y(x) of minimum length joining two given points, A: (a, ya) and B: (b, yj?), in the *j>-plane (see Fig. E4.38). Bjo(b,yb) 4.39 Isoperimetric problem. Among all curves with a continuous derivative that join two given points, A: {a, ya) and B: {b, yt,), and have the given length L, find the one that encompasses the largest possible area. Determine (a) the Euler equations and (b) the natural boundary conditions associated with the functionals identified in Exercises 4.40-4.43. Specified essential boundary conditions are shown if there are any. 4.40 Functional in Exercise 4.31. 4.41 Functional in Exercise 4,32. 4.42 Functional in Exercise 4.33. 4.43 Functional in Exercise 4.35. 4.44 Functional in Exercise 4.36. 4.45 Obtain the Euler equations and the natural boundary conditions associated with the functional n(w<>) = which arises in connection with axisymmetric bending of polar orthotopic annular plates with inner radius r\ and outer radius ro. Assume that the deflection wq = 0 at r = ro, the outer radius of the plate.
EXERCISES 131 4.46 Derive the Euler equations and the natural and essential boundary conditions associated with the Lagrange multiplier functional of the following problem: Minimize the functional +d - v) (^ + ^)21 dxdy -J^wo dxdy subject to the constraints —— + 0, = 0, —- + 4>y = 0. ox By Use two different Lagrange multipliers, say Xx and Xy, with the two constraints. 4.47 Repeat Exercise 4.46 with the penalty functional of the problem (i.e., construct the penalty functional and derive the Euler equations). Use two different penalty parameters, say yx and yy9 with the two constraints. Note that the penalty functional represents the total potential energy of the first-order shear deformation plate theory (FSDT) discussed in Chapter 8 (for yx = yy = KsGh> where Ks is the shear correction coefficient, G the shear modulus, and h the thickness of the plate). 4.48 Consider the problem of maximizing the functional rb J (it) — I u(x)dx, u{a) = uQ, u{b) = m&, Ja subject to the constraint ™-CF® 2 dx-L — 0, Use (a) the Lagrange multiplier method and (b) the penalty method to introduce the constraint into /(w), and derive associated Euler equations. Note that the constraint is an integral constraint. 4.49 Consider the problem of minimizing the functional fb I (du\2 r(u) = / it J1 —I— I —— 1 dx, u(a) = ua, u(b) = lib, /( subject to the constraint /•» I /du\2 dx — L. ihm Use (a) the Lagrange multiplier method and (b) the penalty function method to introduce the constraint into the functional, and derive the Euler equations.
132 WORK, ENERGY, AND VARIATIONAL CALCULUS 4.50 Consider the problem of minimizing the functional EI fL 2 fL n(ii^), kxx) = — / kkk dx - / qwQ dx, 1 Jo Jo subject to the constraint d2w0 where wo is the tranverse deflection, q the distributed load, El is the flexural stiffness, and L is the length of the beam. Construct a functional using the Lagrange multiplier method and determine the Euler equations. REFERENCES 1. Charlton, T. M., Principles of Structural Analysis, Longmans (1969). 2. Charlton, T. M., Energy Principles in Theory of Structures, Oxford University Press, Oxford (1973). 3. Davies, G A. O., Virtual Work in Structural Analysis, John Wiley, Chichester, UK (1982). 4. Dym, C. L., and Shames, I. H., Solid Mechanics: A Variational Approach, McGraw-Hill, New York (1973). 5. Hildebrand, F. B., Methods of Applied Mathematics, 2nd edition, Prentice-Hall, Englewood CI iffs, NJ (1965). 6. McGuire, W., Gallagher, R. H., and Ziemian, R. D., Matrix Structural Analysis, 2nd edition, John Wiley, New York (2000). 7. Mikhlin, S. G., Variational Methods in Mathematical Physics, Pergamon Press (distributed by Macmillan), New York (1974). 8. Mikhlin, S. G, Mathematical Physics, An Advanced Course, North-Holland, Amsterdam (1970). 9. Oden, J. T., and Rippcrger, E. A., Mechanics of Elastic Structures, 2nd edition, Hemisphere/McGraw-Hill, New York (1981). 10. Reddy, J. N., An Introduction to the Finite Element Method, 2nd edition, McGraw-Hill, New York (1993). 11. Rcddy, J. N., and Rasmussen, M. L„ Advanced Engineering Analysis, John Wiley, New York (1982); reprinted by Krieger, Melbourne, FL (1990). 12. Reddy, J. N., Energy and Variational Methods in Applied Mechanics, John Wiley, New York (1984). 13. Reddy, J. N., Applied Functional Analysis and Variational Methods in Engineerings McGraw-Hill, New York (1986); reprinted by Krieger Publishing Co., Malabar, FLQ991). 14. Reddy, J. N., Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia, PA (1999). 15. Richards, T. H., Energy Methods in Stress Analysis, Ellis Horwood (distributed by John Wiley & Sons), Chichester, UK (1977).
5 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS 5.1 VIRTUAL WORK PRINCIPLES 5.1.1 Introduction In this chapter we study the virtual work and energy principles of solid mechanics. These include the principles of virtual displacements and virtual forces, the principle of minimum total potential energy, and the principle of maximum total complementary energy. These principles will be used to derive the unit-dummy-displacement and unit- dummy-load methods and Castigliano's theorems I and II of structural mechanics. The variational principles will be used to derive the equations of equilibrium or motion of deformable solids, including bars and beams. The use of energy methods (i.e., unit-dummy-displacement and unit-dummy-load methods and Castigliano's theorems) in the determination of point displacements and forces will be illustrated with bars, beams, trusses, and frames. Betti's and Maxwell's reciprocity theorems will also be discussed. Both direct variational methods (i.e., Ritz, Galerkin, and weighted- residual methods) and the finite element method that make use of variational methods will be discussed in subsequent chapters. 5.1.2 The Principle of Virtual Displacements Recall from Section 4.2 that virtual work is the work done on a particle or a deformable body by actual forces in moving through a hypothetical or virtual displacement that is consistent with the geometric constraints. The applied forces are kept constant during the virtual displacement. The principle of virtual displacements states that the virtual work done by actual forces is zero if and only if the body is in equilibrium. It is informative to consider the principle of virtual displacements for a particle in static equilibrium. Suppose that the particle is in equilibrium under the action of 133
134 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS n concurrent forces Fi, F2,..., F„. Now suppose that the particle is given an arbitrary virtual displacement 8u during which all forces along with their directions are fixed. The total virtual work done by all forces is given by 8W = Fi • Su + F2 • Su + • • • + Fn • <5u = (EF',)'*u- (5J) Note that the expression in parentheses is the vector sum of all forces acting on the particle. From vector mechanics we know that the sum is zero if the particle is in equilibrium, giving 8 W = 0. Conversely, if 8 W = 0 and 8u is arbitrary, it follows that YH! F/ = 0, i.e., the particle is in equilibrium. In other words, the particle is in equilibrium if and only if 8 W = 0 for any choice of 8u. The statement 8 W — 0 is the mathematical statement of the principle of virtual displacements for a particle. The discussion can be extended to rigid bodies because a rigid body is merely a collection of particles. Next, we consider a generalization of the principle of virtual displacements to deformable bodies. In deformable bodies, material points can move relative to one another and do internal work in addition to the work done by external forces. Thus we should consider the virtual work done by internal forces (i.e., stresses) as well as that done by external forces. Consider a continuous material occupying the volume V and in equilibrium under the action of body forces f and surface tractions t Suppose that over portion Si of the boundary displacements arc specified to be u. and on portion S2 tractions are specified to be t. The boundary portions S\ and S2 are disjoint (i.e., do not overlap), and their sum is the total boundary, S. Let u = (111,112,113) be the displacement vector corresponding to the equilibrium configuration of the body, and let ay/ and s;j be the associated stress and strain components, respectively. We make no assumption concerning the constitutive behavior of the material body. The set of admissible configurations is defined by sufficiently differentiable displacement fields that satisfy the geometric boundary conditions: u = u on S\. Of all such admissible configurations, the actual one corresponds to the equilibrium configuration with the prescribed loads. In order to determine the displacement field u corresponding to the equilibrium configuration, we let the body experience a virtual displacement 5u from the equilibrium configuration. The virtual displacements are arbitrary continuous functions except that they satisfy the homogeneous form of the specified geometric boundary conditions, i.e., £u = 0 on 3i. The principle of virtual work states that a continuous body is in equilibrium if and only if the virtual work of all forces, internal and external, acting on the body is zero in a virtual displacement: 8W = &WI+8WE=Ot (5.2) where 8Wj is the virtual work due to the internal forces and SWe is the virtual work due to the external forces.
VIRTUAL WORK PRINCIPLES 135 The principle of virtual work is independent of any constitutive law, The principle may be used to derive the equilibrium equations of deformable solids, as already illustrated by Examples 4.10 and 4.11. Here we consider additional examples. Example 5.1 (Euler-Bernoulli beam theory) In Example 4.3, we have discussed the kinematic hypothesis of Euler-Bernoulli beam theory, and used a displacement field consistent with the hypothesis (see also Fig. 4.7a). Here we use the principle of virtual displacements to derive the governing equations of the Euler-Bernoulli beam theory for the case of large deflections but small strains. The displacement field of a beam under the Euler-Bernoulli kinematic hypothesis is given by [see Eq. (4.28)] dwo u(xyy,z) = uo(x)-z—r-, i> = 0, w(x,ytz) = wo(x). (5.3) ax If we assume that the strains are small in the sense that the following terms are negligible compared to du/dx, /du\2 \dx) • du du „ ,v ««■ -r-r, (5-4) dx dz the only nonzero nonlinear strain is given by dx 2\ dx J dx2 We shall consider the bending and stretching of a beam on a linear clastic foundation with foundation modulus k, and subjected to a distributed longitudinal load f(x) and distributed transverse load q(x) at the top (see Fig. 4.7b). Then the external and internal virtual works due to the virtual displacements Suo and Swq are given by SWE = -\ I (fMSuQ + q(x)8w(x9—^JJ dx + / Fs8w(x^J dx\ = — / (f(x)8iiQ + q(x)Swo(x)) dx — I kwo(x)8wo(x) dx , (5.6) SWj — I I crxx8sxx dxdA Jo J A fL f /dSiiQ dwodSwo d28wo\ _ . . /c _. = / / <rxx —r- + -z : z—nr dxdA> <5-7) Jo J A \ dx dx dx dx1 } where L is the length and A the cross-sectional area of the beam. The foundation reaction force Fs is replaced with Fs — —kwo(x) using the linear elastic constitutive equation for the foundation.
136 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS The principle of virtual displacements requires that 8 W = SWj + SWe = 0, which gives 0= fL f a (d8uu dwpdSwp d28wo\dAdx Jo J a AA \ dx dx dx ' dx2 J - I [ fSuo + (q - kwo) 8wq\ dx Jo = fL\N(d8u° dwQdSwo\ Afrf2^w°l Jo L V dx dx dx J dx2 J - / [/too + (q— kwo) 8wq] dx Jo fL\ dN p rf /^lf\ d2Af 1 f = / ~—8uq - — -j—N Swq - —^-8w0 dx Jo L dx dx \ dx / dx£ J fL - I lf&UQ + (cj — kwo) 8wo\ dx Jo r o (dwo dM\ o ,</Sw0~|L where N and M are the stress resultants defined by N = f <rxx dA, M= f <txxz d.A. (5.9) J A J A Hence, the Euler equations arc dN Su0: - — = f(x). (5.10) dx d2M d /dw0 \ Sw* -^-TA-^N)+kw°-q=*' (5-n) in 0 < x < L. Note that 8uq, 8wo, and dSwo/dx appear in the boundary terms. Hence, wo. u?0> and dwo/dx are the primary variables of the theory, and their specification constitutes the essential boundary conditions. Since nothing is said about the primary variables being specified, S«o> 8wo, and dSwo/dx are arbitrary at x = 0 and x = L. Hence, the natural boundary conditions are: *=0. ^ + ^V=0, Af = 0. (5.12) dx dx
VIRTUAL WORK PRINCIPLES 137 5.1.3 Unit-Dummy-Displacement Method The principle of virtual work can also be used, in addition to deriving governing equations of equilibrium, to directly determine reaction forces and displacements in structural problems, in particular, truss and frame structures. If Ro is the force at point O in a structure, we can prescribe a virtual displacement <5uo at the point. The virtual strains due to the virtual displacement are determined from kinematic considerations. Then the principle of virtual work reads as Ro•Suo = f^ij&e?jdV, (5.13) where 07; are the actual stresses, and Ssf. are the virtual strains in the structure derived from the virtual displacement 5uo, consistent with the geometric constraints. Since <5uo is arbitrary, one can take Suq = c#, the unit vector along the force. If 8e9. are the strains due to the unit virtual displacement at point O, then = f *ijl Jv nO Rq= J o-ijSefjdV. (5.14) This procedure is called the imit-dummy-displacement method. The word "dummy" is used historically to mean "virtual." Equation (5.14) is also valid for the case in which Rq is replaced with a bending moment Mq\ -fv°,H M0= aijSsfj dV. (5.15) Equation (5.13) is more general than Eq. (5.14) in the sense that it is valid for an arbitrary virtual displacement vector (i.e., it need not be a unit vector). Equations (5.14) and (5.15) can be used to determine point loads and moments or deflections and rotations of structures. Tf we were to use the unit-dummy-displacement method for the determination of displacements, rotations, forces, and/or moments in a bar, truss, beam, or frame structure, it is necessary that the displacement field is represented in terms of displacements and/or rotations of discrete points in the structure. In other words, we must identify generalized coordinates that can be used to represent the deformation of the structure. Suppose that q\>q2,---,qn denote the generalized coordinates of a structure. Then the principle of virtual displacements can be stated as 0 = awfoi, 42,..., qn) = SWE + SWi (dWE (3W/\ Since 8WE sqi = F" (5-17)
138 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS we have where F/ is a generalized force causing the generalized displacement <?/. The generalized coordinate qi can be a displacement or a rotation, implying that F,- can be a force or a moment. For linear elastic beams under axial and bending deformation, we can write ,, _ /'' [*,*•>• (*S) + „ *»± (%»)]*. (5,8) where it is understood that the axial displacement wo and transverse displacement wq can be expressed in terms of the generalized coordinates <?/. Two examples of application of the unit-dummy-displacement method are presented next, Example 5.2 Consider the truss shown in Fig. 5.1a. We wish to determine the vertical displacement v and horizontal displacement u of point O using the unit- dummy-displacement method. We shall assume linear elastic behavior and small strains. Assume virtual (or dummy) displacements of 8vq downward and 8uq horizontal (see Fig. 5.1b) at point O so that <5uo = S«oeT 4 8vo&y. The load in this case is R = 0eY 4- Pey. Then we have R0 . Suo = 0 . Suq 4 P • 8vq = I (Xjj8eij dV Jv = f"1 ^H'^U dx + J"' A™o™Se™ dx, where A^ = A(2) = A, h[ = h, and hi = >/2/i. The actual stresses are computed in terms of the actual displacements uq and vq of point O from the deformed geometiy (a) (b) Figure 5.1 (a) Original truss, (b) Truss with virtual displacements.
VIRTUAL WORK PRINCIPLES 139 of each member of the truss as follows: *<{> = E^s™, <r<2) = EV>e™, £(,) = £(2) = E 4} = \ (V(A + «o)2 + w§ - ft) « y (5.19) £<2> = -^ (,/(/,+ ,,„)* 4-(*-v0)2 - V&) - ^. where the squares of wo/^ and vo/h were neglected in computing the strains (small strain assumption). Similarly, the virtual strains Ss[^ and Se\^ can be computed as (alternatively, the virtual strains can be computed by taking the variation of the actual strains): ,e{» = ^o, feW = ffSLL**. (5.20) Substituting Eqs. (5.19) and (5.20) in (5.18), we obtain 0 • Suo + P-8v0 = hAajf-^ + vW?^^ h ll 2h = Uv™ + ;4f A{rff}) *tt° + (~7f Acrn}) 5v°* (5'21) Since the variations 8uq and 6Do arc arbitrary, we can set them to (i) Suq = 1 and Svq = 0, and then to (ii) 5wo = 0 and Svo = 1, to obtain the two relations for the two displacements. Alternatively, collecting the coefficients of Su and Sv separately, we arrive at or />/? 2V2Ph Ph UQ vo = ^r~ + uo = —(1 + 2^/2). (5.22c) A£ AE AE Example 5.3 Consider a cantilever beam of length L and axial and bending stiffnesses EA and EI, respectively. We wish to determine the generalized deflections of the free end of the cantilever beam when it is subjected to transverse load Fo> bending moment Mo, and axial load Pq at its free end (x = L; see Fig. 5.2). We shall use the Euler-Bernoulli beam theory (see Example 5.1). First we must write the displacements uo and wo in terms of the generalized coordinates, which can be identified as the displacements of the free end, as discussed below. Then we apply the unit-dummy-displacement method to determine the generalized displacements in terms of the applied loads.
140 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS 4- *| M0 *■ x, uQ Figure 5.2 A cantilever beam with end loads. The axial displacement uq(x) is governed by the equation EA-—=- = 0 ->■ «o00 = ai + fl2*. (5.23) dx2 Using the only displacement boundary condition uq(0) = 0, we determine a\ — 0. Then aj can be used as the generalized coordinate; it is useful to express ai in terms of the axial displacement at the free end, ul- uq(L) = a^L or ci2 = uq{L)/L = ul/L. Thus, we have (a — ul) «oW = (^)Cl" Then the unit-dummy-displacement method gives P0 = / £i4-—— — </* = / fA7-dt = — ci, Jo dx dc\ \ dx ) Jo L L L or n\ P°L Similarly, we proceed to express wq(x) in terms of generalized coordinates q, which must be identified. The equation governing wq(x) is EI 4 = 0 -> wq(x) = «i -f- ^2^ 4- tf3*2 + a4Jt3. (5.24) Using the two Icnown boundary conditions ?/;o(0) = 0 and (dwo/dx)(0) = 0, we obtain ci[ = 0 and ^2 = 0. Next, we express a?> and a4 in terms of the displacement and rotation at the free end, x = L\ w0(L) = wL -> wL = a3L2+a4L3\ ( -^ j =9L-+eL = 2a3L + 3a4L2.
VIRTUAL WORK PRINCIPLES 141 Thus wl = C2 and Ol = t'3 can be used as the generalized coordinates. Solving the above equations for a^ and a^ in terms of wl and Ol, and substituting the result into Eq. (5.25), we obtain W0(X)==(4-4)C2+L("S + S)C3- Now using the unit-dummy-load method, we obtain p0 = [LEId2w° d (d2w°)dx Jo dx2 dc2 \ dx2 J Jo "* dx2 do, \ dx1 } -H(*-MM4*MH'B+M)" Upon carrying out the integration, we obtain 12EI 6EI 6EI 4EI Fo = —7jTc2 - ~JTC3> M° = —7T6"2 + ~T~C3y whose solution gives the deflection and rotation of the free end: ,__ F0L3 MQL2 fdw0\ F0L2 M0L The procedure discussed in the above example to express the displacements uo(x) and wq(x) in terms of unknown generalized displacement degrees of freedom holds for any bar and beam structure. When a distributed load q (x) ^ 0, one may convert it to a set of statically equivalent point loads F/ acting at the same points and directions as the generalized displacements cv of the beam. For bars and beams with constant EA and El but arbitrary load q(x), this procedure results in exact values of the generalized displacements q. The procedure can be generalized to bars and beams with arbitrary boundary conditions, geometric properties, and material properties. However, in such cases the solutions obtained are approximate. To further explain the idea, we express the displacement in terms of a linear combination of undetermined parameters a and known functions <p/ that satisfy the kinematic boundary conditions:* N mix) ~ J] cm(x). (5.25) ;=1
142 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS For example, the deflection of a simply supported beam under any combination of loads can be expressed as N inx <r~A . MIA, woO) ~ 2^c'sin_f-' /=1 where L is the length of the beam. This choice of functions satisfies the boundary conditions u>o(0) = wo(L) = 0. By substituting the expansion (5.25) into Eq. (5.18) with uq — 0, we obtain "L Ejd2w0d28wo ^ ._1 „u dx2 dx2 /v Jjc 1=1 -t(r«&&-)- 7-1 5 c/ which gives, because of the independent nature of c,, the necessary equations for the parameters c/: Then the continuous (approximate) solution can be obtained from Eq. (5.25). These ideas will be further discussed in Chapter 7 in connection with the Ritz method. 5.2 PRINCIPLE OF TOTAL POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM I 5.2.1 Principle of Minimum Total Potential Energy The principle of virtual work discussed in the previous section is applicable to any continuous body with arbitrary constitutive behavior (i.e., elastic or inelastic). A special case of the principle of virtual work that deals with elastic (linear as well as nonlinear) bodies is known as the principle of minimum total potential energy. Recall that for elastic bodies (in the absence of temperature variations) there exists a strain energy density function Uq such that [see Eq. (3.34)] Gjj = -—. (5.27)
PRINCIPLE OF TOTAL POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM I 143 The strain energy density Uq is a function of strains at a point and is assumed to be positive definite. The principle of virtual displacements. 8W = 0, can be expressed in terms of the strain energy density U$\ 0 = 8W = / <ju8sjj dQ - \ J f • 8u dQ + / t • 8u dS Ja Ssij ,J -/ 'S2 dQ + 8V 8U0 dQ + 8V = 8(U + V) = 811, (5.28) where or -8V = - [ f f • 8u dtl + f i . ju rfsl, V = -/f.udJ2+M.urf5L (5.29a) and U is the strain energy: U= f U0dtt. (5.29b) Note that we have used a constitutive equation (5.27) in arriving at Eq. (5.28). If the temperature changes are considered, the strain energy density Uo should be replaced by p^, where ^ is the free-energy functional defined in Eqs. (4.14) and (4.15). The sum V + U = FT is called the total potential energy of the elastic body, and the statement n =8(1) + V) =0 (5.30) is known as the principle of minimum total potential energy. The principle of virtual displacements as well as the principle of minimum total potential energy give, when applied to an elastic body, the equilibrium equations (3.13) as the Euler equations. The main difference between them is that principle of virtual displacements gives the equilibrium equations in terms of stresses (or stress resultants), whereas the principle of minimum total potential energy gives them in terms of the displacements, because in the latter a constitutive relation is assumed to replace the stresses in terms of the displacements. Example 5.4 Consider the bending of a beam according to the Euler-Bernoulli beam theory (see Example 5.1). We wish to construct the total potential energy functional and then determine the governing equation and boundary conditions.
144 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Z, M'o Figure 5.3 A beam with applied loads. Recall from Example 4.3 that the strain energy of a beam under the assumption of small strains and displacements for the linear elastic case (i.e., obeying Hooke's law) is given by Eq. (4.39): HfR£)2-'(s?y dx, (5.31a) where L is the length, A the cross-sectional area, / the second moment of area about the axis (y) of bending, and E is Young's modulus of the beam. Suppose that the beam is subjected to distributed axial force f(x) (measured per unit length), distributed transverse load q(x) (measured per unit length), horizontal point load P at x = L, transverse point load F at x = L, and bending moment Mq (counterclockwise) at x = L, as shown in Fig. 5.3. Then the potential of applied forces is given by =-[/"<■ (/ho + gw0) dx + Puq(L) + Fwo(L) + M0 (-SU (5.31b) Note that (—dwo/dx) is the rotation in the direction of the moment Mq. Hence, the total potential energy of the beam is n<*»»'=\ t h (^)2+E' (%?)' - 2(/u°+«,w) [' - I Puo(L) + Fwo(L) - Mo^ dx ,v=L. dx (5.32) Applying the principle of minimum total potential energy, SU = 0, and using the tools of variational calculus we obtain (see Example 4.6) o = sn= f (ea duo dSuo dx dx d2wod2Swo\ + El —r-z r^— I dx dx2 dx wq\ (fSuo + qSw0) dx + P8uq(L) + FSwq(L) - Mq dSwo I 1 dx \x=iJ
PRINCIPLE OF TOTAL POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM I + \ea-—8uq\ + \ei-1-t — r\EI^rr)SwQ L dx Jo L dx dx dx \ dxz / 145 lC(/5-° + qSw0)dx + PSuo(L) + FSivq(L) - M0 dSwo I dx \x= J- (5.33) An examination of the boundary terms resulting from integration by parts shows that the primary and secondary variables of the theory are Pri mary variables: wo - u)q , . dx ci • ui i?AduQ d (j7id2w^\ trrci2w(> Secondary variables: EA —-, — ( EI l, EI 2 . It is clear that the secondary variables are nothing but the axial force A^(x), shear force V(x) = dM/dx, and bending moment M(x). Only one element of each of the pairs (no, N), (wq, V) and (—dwo/dx> M) may be specified at a point. Recall that specification of a primary variable is an essential boundary condition and that of a secondary variable is a natural boundary condition. Returning to the expression in Eq. (5.33), first wc collect the coefficients of <5«o and 8wo in (0, L) together and set them to zero separately to obtain the Eulcr equations: 8u0: - 4" \EA~\ - f{x) =0, 0 < jc < L, (5.34) dx \ dx J Sw0: 4^(ET^¥)-Cl(x) = 0> 0<x<L. (5.35) dx2 \ dx2 J These equations are equivalent to those in Eqs. (5.10) and (5.11) (without the nonlinear and elastic foundation terms). Now considering all boundary terms in (5.33), we conclude that 8u0(Q) = 0, \ dx / v==l X d ( a V dx2 Jx==0 \ dx /A.=0 o i2WQ dx2 Suo(L) = 0, 5w0(0) = 0, |_ dx SwQ(L) = 0, =L \ dx Jx=L (5.36)
146 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS If any of the quantities 8uq> 8wq, and dSwo/dx arc zero at x = 0 or x = L, because of specified geometric boundary conditions there, the corresponding expressions vanish; the vanishing of the coefficients of Suo, 8wo, and dSwo/dx at points where the geometric boundary conditions arc not specified provides the natural boundary conditions. For example, suppose that the beam is clamped at jc = 0 and free at x = L. Then Siin(O) = 0, Swq(0) = 0, and d8wi)(0)/dx - 0, and the natural boundary conditions become L dx \X=L 5.2.2 Castigliano'sTheorem I Like the unil-dummy-displacement method, which can be used to determine the unknown point loads and displacements of structural systems composed of discrete structural members, Castigliano's theorem allows one to compute displacements or loads. Castigliano (1847-1884), an Italian mathematician and railroad engineer, was mainly concerned with linear elastic materials. The generalization of Castigliano's original Theorem I to the case in which displacements are nonlinear functions of external forces is attributed to Engesser. In the present study we consider Castigliano's theorems in a generalized form that are applicable to both linear and nonlinear elastic materials. Suppose that the displacement field (or geometric configuration) of a structure can be expressed in terms of the displacements (and possibly rotations) of a finite number of points x/ (j? = 1, 2,..., N) as u(x) = £></>,• (x), (5.38) f=i where u, are unknown displacement parameters, called generalized displacements, and (pi are known functions of position, called interpolation functions, with the property that <(>i is unity at the ith point (i.e., x = x/) and zero at all other points (x/, j jL /). Then it is possible to represent the strain energy U and potential V due to applied loads in terms of the generalized displacements u/. The principle of minimum total potential energy can be written as BU 3V 8U = -8V or — • Suf = -— • Su/ 3u/ ail/
PRINCIPLE OF TOTAL POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM I 147 where sum on repeated indices is implied. Since and 8u.j are arbitrary, it follows that (£-*)-.-' or 8U = F/. (539) Equation (5.39) is known as the first theorem of Castigliano. For an elastic body, it states that the rate of change of strain energy with respect to the displacement is equal to the load causing the displacement. When applied to a structure with point loads F-t (or moment A/;) moving through displacements «,- (or rotation 0,-), both having the same sense, the theorem states that diij or dU 30/ (5.40) It is clear from the derivation that Castigliano\s Theorem 1 is a special case of the principle of minimum total potential energy, and hence of the principle of virtual displacements. Consequently, the theorem is equivalent to the unit-dummy-displaccment method. Example 5.5 Let us revisit the truss problem of Example 5.2. The strain energy of the structure can be expressed in terms of the displacements it and v of point O (see Fig, 5.1). We have 1 2 r EA Then by Castigliano's Theorem I, we have 0= — = EA ou it \f2 it — v h +~2 2hT P^ — = EA[() dv \ V2u — v ~2 2/T which gives Ph EA' u = (1+2^2) Ph EA'
148 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Example 5.6 Consider a free-free straight beam of constant bending stiffness EI and subjected to point loads and moments, as shown in Fig. 5.4a, The equation governing the equilibrium of the beam according to the Euler-Bernoulli beam theory [see Examples 4.3, 5.1, and 5,3, and Eq. (5,35)] is The solution to this homogeneous fourth-order equation is given by wq(x) = tfi 4- ci2X -f a?>x H- a$x , (5.41) (5.42) where a\ (i — 1, 2, 3, 4) are constants of integration. We wish to express these constants in terms of the deflections and rotations at the two ends of an arbitrary beam of length L. Suppose that ( dwo\ \ dx Jx=s0 (5.43) «3 == wo(L) = a\ + aiL + a^L2 + atl?, «4 = f — I = — ti2 — la^L — 3a.4L2. \ dx )X=L Note from Eq. (5.43) that u\ and a3 are the values of the transverse deflection wo at x = 0 and x — L, respectively, and it2 and 114 are the rotations —dwo/dx, measured positive counterclockwise, at jc = 0 and x = L, respectively (see Fig, 5.4b), The four equations relate the four constants ax to the four displacements iij9 called generalized displacements, which serve as the generalized coordinates. These relations can be inverted to solve for «/ in terms of a\. Then substituting the result into Eq, (5.42) yields 4 wq(x) = <p\ (x)it[ + <p2(x)it2 + <P3(*)«3 + <P4(x)ii4 = £ p,-(*)w/, (5.44) i = \ F\ + Cl\ F3 + ^, «, «3 (a) Generalized forces (b) Generalized displacements Figure 5.4 (a) Beam with end forces and moments (or generalized forces), (b) Generalized displacements.
PRINCIPLE OF TOTAL POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM I 149 where <p\ A'\3 «-'-3(i)+j(i) V2(X) = -X 1 -2 (l)-(l)2] »w = (z)2(3-2i)' (5.45) The strain energy of the beam now can be expressed in terms of the generalized coordinates as "-¥f(£)'* 1 4 4 dx /=1 ;-1 where -rfjc. (5.46a) (5.46b) /o Jjc2 dfjc2 Note that Kjj is symmetric (£/_/ = /T;,). The work done by applied forces is given by *L 4 V = - ' / <y(x)u;oW dx + /"", ^/«/ 4 i = 1 where # = / q(x)(pi(x)dx Jo (5.47a) (5.47b) and F( are the generalized forces associated with the generalized displacements «/. Thus, F\ and F3 are the transverse forces at x = 0 and * = L, respectively, and Fi and F4 are the bending moments at x = 0 and x = L, respectively (see Fig. 5.4a). The transverse forces q\ and q$ and bending moments qi and 44 together are statically equivalent to the distributed load q(x) on the beam.
150 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS % TTTTTTTTTTTTTTX _ o T Figure 5.5 A beam fixed at x = 0 and supported by a spring at x = L Using Castigliano's Theorem I, we write dU diij 7=1 or, in matrix form, 2EI 6 -3L -6 -3L" -3L 2L2 3L L2 -6 3L 6 3L -3L 3L 2L2 \u\ "2 «3 W4^ ' = < [<7l </2 43 #4 > + < [Fll F2\ *3 lF4J (5.48a) (5.48b) As a specific example, consider a beam fixed at x = 0 and spring supported at x = L, and subjected to uniformly distributed load (see Fig. 5.5). We wish to determine the compression in the spring, i.e., determine v)q(L). For uniformly distributed load q == go, we have from Eq. (5.47b) Un l?2 |?3 l^. C/oL 12 f 6 ] -L| 6 1 t ] (5.49) The geometric boundary conditions require u \=U2 — 0. The force boundary conditions require F3 = —Fs = —kwo(L) = —kui and F4 = 0. Thus we have 2EI 6 -3/, -6 -3L" -3L 2L2 3L /,2 -6 3L 6 3L -3/, L2 3L 2L2 f o' 0 «3 «4 12 [ 6 ' 6 ■ L > + < f ^ 1 ^2 1 —/cm 3 j 0 J (5.50) or, setting up the equations for the unlaiown generalized displacements U3 and M4, we obtain L3 [3L 2L2J1«4) 12 \L\ + \ 0 J'
PRINCIPLES OF VIRTUAL FORCES AND COMPLEMENTARY POTENTIAL ENERGY 151 or 12EI 6£/1 L3 +k ~JT 6EI AEI L2 L J Solving for 113 and W4 by Cramer's rule, we obtain i"4j 12 \l] «3 = ^o(L) = qoL4 1 H4 = — %EJ (l + (kL3/3EI))9 gpL* (EI - (/cL3/24)) 6EI (EI + (/cL3/3)) ' Note that when k = 0, wc obtain the deflection 113 and rotation 114 at the free end of a cantilever beam under uniformly distributed load of intensity qo. 5.3 PRINCIPLES OF VIRTUAL FORCES AND COMPLEMENTARY POTENTIAL ENERGY Tn Section 5.1 we discussed the principle of virtual displacements. Naturally, the virtual work done by virtual forces in moving through actual displacements should have similar use. Here we formulate the principle of virtual forces. The basic idea is that each particle in the body undergoes actual displacements until the stresses developed satisfy the equations of equilibrium. Consider single-valued, differentiate variations of a stress field So and body forces St that satisfy the linear equilibrium equations both within the body and on its boundaries: V-5? + 5f = 0 inV, (5.51) n • Scr = St on S2. (5.52) We shall call such a stress field a statically admissible field of variation. These virtual stresses and forces, except for self-equilibrating, are completely arbitrary and independent of the true stresses and forces. The external complementary virtual work is defined by SWl = - [ u-8tdV- [ u-StdS, (5.53a) Jv J s{ where u is the displacement vector, and St and St satisfy Eqs. (5.51) and (5.52). The internal complementary virtual work is given by SWf= [ T:8*dV, (5.53b) Jv
152 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS where s are actual linear strains and 80 is the virtual stress field that satisfies Eqs. (5.51) and (5.52). The principle of complementary virtual work (or virtual forces) states that the strains and displacements in a defonnable body are compatible and. consistent with the constraints if and only if the total complementary virtual work is zero: 8Wf + 8Wl;=Q. (5.54) Tt can be shown that the principle of virtual forces gives the kinematic relations and geometric boundary conditions as the Eulcr equations. This is illustrated by considering the rectangular component form of 8 W* and 8 Wg for a dcformable solid: 0 = f eijScrij dV - f UiSfi dV - / fi/Sf/ dS JV ' Jv JS] = / M^V dV ~ / Uj{-8<Tijj) dV - I UfnjSaij dS Jv Jv Js\ = - / -ziuij + ujj) - Sfj 8aij dV + I («,- - Ui)nj8crij dS, (5.55) where sum on repeated indices is assumed. Because Scfjj is arbitrary, wc obtain the strain-displacement equations and the displacement boundary conditions as the Euler equations F<ij ~ -z(Mij + Ujj) = ° in V (5.56a) ui — Hi =0 on S[. (5.56b) The unit-dummy-load method is a special case of the complementary virtual work, and it can be used to determine displacements and forces in structures. The basic idea can be described by analogy with the unit-dummy-displacement method. If uq is the true displacement at point O in an elastic structure, we can prescribe a virtual force 8Rq at the point. The application of virtual force induces a system of virtual stresses Say that satisfy the equilibrium equations. Then from the principle of virtual forces (5.54), we have (8WE = -uqSRo): uoSRo - f SijSoZ dV. (5.57) Once again, one can take «5/?o = 1 and calculate corresponding virtual internal stresses 8ojj. Equation (5.57) represents the unit-dummy-load method. For the Eulcr-Bernoulli beams, Eq. (5.57) takes the form ' uqSPq + wqSFq + #o<5Mo + 4>o8Tq fL ( N M V T \
PRINCIPLES OF VIRTUAL FORCES AND COMPLEMENTARY POTENTIAL ENERGY 153 (a) (b) Figure 5.6 (a) Original truss, (b) Truss with virtual forces. where (8Pq, <$^o» SMq, 8Tq) are the generalized virtual forces at point O, and («o> wo> #o, 0o) are the corresponding generalized displacements. Example 5.7 Consider the frame structure shown in Fig. 5.6a. We wish to determine the vertical displacement v and horizontal displacement it of point O using the unit- dummy-load method. Assume virtual (or dummy) forces of SP downward and 8Q horizontal (see Fig. 5.6b). Then we have u-8Q + v-8P= f eijSajj dV = J2 f ' AU)e$8<rff dx, (5.59) where A^ — A(2) = A and h\ — h, h^ = sflh. The actual stresses are computed in terms of the actual forces in each member due to the applied load P. For static equilibrium of the structure, the member forces are found to be (5.60a) (5.60b) The virtual stresses are computed from virtual forces, which in turn are computed from the virtual loads 8P and 8Q on the structure. We obtain Fil) = p% ence F<2> = -V2P, *" - EA' (1) _ £ (2) _ _ p(2)_ ^P £" ~ EA ■ ~A~ <5F(1) = 8Q + 8P, SF(2) = -Jl8P, (5.61a) and 8a a). SQ + SP 11 *>$= *Jl8P (5.61b)
154 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Substituting Eqs. (5.60b) and (5.61b) in (5.59), we obtain 8Q + 8P P r- *Jl8P «JlP or, collecting the coefficients of 8Q and 8P separately, we obtain u = —, v = —- + 2V2—- = -—(I + 2V2). EA EA EA AEX } (5.62) Example 5.8 Consider the overhang simply supported beam shown in Fig. 5.7a. We wish to determine the rotation (or slope) at the left end of the beam using the unit-dummy-load method. Since no moment is applied at the left end where we wish to determine the rotation, we assume a virtual moment of 8Mo at the left end (clockwise). The virtual moment should be in self-equilibrium. A convenient set of self-equilibrating forces consist of a force of 8Mo/20 downward at the left end and another force of 5 Mo/20 upward at the other support (see Fig. 5.7b). Using the unit-dummy-load method, we obtain 0o 8M0 = ^t / Af(jc)*Af°(jc) dx, EI Jo (5.63) where M(x) is the actual bending moment and 8M°(x) is the virtual bending moment due to the virtual forces at any location x along the beam. We have (first we must compute the reaction forces at the supports using static equilibrium of forces): M(x) = 8MQ(x) = 3750* - 200*2, 3750* - 200*2 + 6250(x 20), 0<x< 20, 20 < x < 25, 8M0 - (^8Mo)x, 0, 0 < x < 20, 20 < x < 25. (5.64a) (5.64b) g=400 lb/ft. JV Ol T T T T T T T T"f ] x ^ 5% 20 ft *z, wQ 5 ft f §M0 20 20 ft *J"^ft (a) (b) Figure 5.7 (a) Given beam, (b) Virtual force system. 20
PRINCIPLES OF VIRTUAL FORCES AND COMPLEMENTARY POTENTIAL ENERGY 155 Using (5.64a,b) in (5.63) and evaluating the integral (the integral over the interval [20 to 25] is zero), we obtain 0o i />20 = "FT / (3750x ~ 200r2 ~ 187.5*2 + IOjc3) dx = - EI Jq 7x 105 6EI ' (5.65) Example 5.9 We wish to determine the horizontal and vertical displacements of point A of the frame structure shown in Fig. 5.8a. We use a dummy vertical load of SQ = 1 kN at point A to determine the vertical deflection. The self-equilibrating virtual force system is shown in Fig. 5.8b. We have SQ 1 f M • v = f —SMdx, fa 0 fa Fax r2.5a pna v = / tttO)dx + / -Trria + x) dx + / ~wr(^)dx Jo EI Jo EI J0 EI 5F«*3 + 5^=35^ = aoS83(m)> 6E1 6EI Similarly, we use a dummy horizontal load of SP = 1 kN at point A to determine the horizontal deflection and find 8P 1- M ~EI SM dx, -I "=r^<o),"+rf(o)'"+i Epa EI (—x) dx 3.125/W = -0.0313 (m). The negative sign for u indicates that the displacement is in the opposite direction to the assumed dummy load. Thus, point A moves downward and to the left. x w—m „ w i^OkN fl = 2m JE=200xJ0f,kN/m3 1= 200x10* mm4 CT |SQ=lkN 1 5F=lkN 2a 2.5a ZakN-m.^ Cx A*" 2a 2.5a lkN 2.5a 2.5a kN-m r--l kN (a) (b) Figure 5.8 (a) Given frame structure, (b) Virtual force system.
156 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS 5.4 PRINCIPLE OF COMPLEMENTARY POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM II Noting that the complementary potential energy due to virtual loads, and the complementary strain energy, are given by 8V* = 8W%, 8U* = 8W?) (5.66) we can arrive at the principle of maximum complementary potential energy: STl* = 8(U* + V*) = 0, (5.67) where n* denotes the total complementary potential energy. Analogous to the unit-dummy-load method, we can derive Castigliano's Theorem II from the principle of maximum complementary energy. We have STl* = 8U* + 8V* = Q -» 8U* = -8V*. Tf £/* and 7* can be expressed in terms of point loads Fj, then we have „. dU* ± dV* 8U* = —&Fi> *V* = -TrjrSFi = -Mfi. dFj aF( and (£-*) at/* SFi^O or ~—=«/t (5.68) dFj Equation (5.67) represents Castigliano's Theorem IL Equation (5.67) is valid for structures that are linearly elastic as well as nonlinearly elastic. When the material of the structure is linearly elastic, we have Uq = Uq and U = U* in value. However, U is always expressed in terms of displacements while U* is in terms of forces, and Castigliano's Theorem I is based on U while Theorem II is based on 17*. Castigliano's Theorem II, which is essentially the same as the unit-dummy-load method, is used to determine deflections and slopes under applied point forces and moments. However, when a structure does not have a load (or moment) at the point at which displacement (or slope) is required, the determination of the displacement (or slope) at that point requires the use of a fictitious (or dummy) load. For example, consider a beam that is subjected to uniformly distributed load <7o, and point loads F\, F2,..., etc. Suppose that we wish to determine the vertical deflection wo at a point at which there is no point load. We introduce a fictitious vertical load 7? at the point, and then write the complementary strain energy in tennis of tfo, R> and F\t F2,..., etc. Then we use Castigliano's Theorem II to determine the desired displacement at the point: at/* I dR l/?=0
PRINCIPLE OF COMPLEMENTARY POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM II 157 Example 5.10 Consider the frame structure of Example 5.7 (see Fig. 5.6a). We wish to determine the vertical displacement v and horizontal displacement u of point O using Castigliano's Theorem IL First, assume that there is a hypothetical load of Q applied in the horizontal direction at point O. This is necessary because, unless U* is a function of Q, wc cannot use Castigliano's Theorem TI to compute u. The complementary strain energy of the structure now becomes 2h'v'E) dV A 2E h 2EA *(^)\va {P2 + Q2 + 2PQ+2^/2P2). Then, by Castigliano's Theorem II, we have KvqJq^o (dU*\ Ph. „ ~ Ph ~EAy Example 5.11 Consider the pin-connected structure consisting of two bars and supporting a load P (see Fig. 5.9). We assume that the bars have the same cross-sectional area A, and are made of nonlinear elastic material whose stress-strain behavior is given by ic2 _ | Ey/e, e > 0, or e = < w a^ (7 < 0. Wc wish to determine (he vertical deflection using Castigliano's Theorem IT. For both members: E = Young's modulus A = cross-sectional area Figure 5.9 A two-member truss.
158 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS The forces in the horizontal and inclined members are given by Fx = -s/3P, F2 = 2P -* a(1) = The complementary strain energy of the structure is <s/3P a™ = IP _ /V3P3L 16^3P3L\ " \ E2A2 + 9E2A2 J _ 25V5 P3L 9 E2A2' The displacement v under the load P is given by dU* 25 P2L v = 3P J$E2A2% Example 5.12 Consider the indeterminate beam structure shown in Fig. 5.10. We wish to use Castigliano's Theorem II to determine the reaction force in the elastic support at the center of the beam. We assume linear elastic behavior of the beam as well as the support. Suppose R is the reaction in the center support. Since the center member is replaced with its reaction, we must consider the energy of the beam only. Due to symmetry, we can consider only half the beam, which receives half the reaction R/2. For the linear elastic case, neglecting the energy due to transverse shear forces, the complementary strain energy is given by (a) \ tit* nTHnunj 0 MR-%L) 1-^^LJ^) 5 {R r j Tz, w>o ° R Figure 5.10 The Indeterminate beam used in Example 5.12.
PRINCIPLE OF COMPLEMENTARY POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM II 159 where M(x) = - [- (qQL - R) x + q0x2] , 0 < x < -. (b) Using Castigliano\s Theorem II, we obtain du* « fL/2 M 8M , -iu0(0) = — = 2 / dx 3(R/2) J0 EI 3R 1 C1'!2 = ^y / [-(qoL - R)x + q0x ]x dx = _L (_2^L + — + qoL*\ 2EI \ 24 24 64 J = JL (Jj^l. + rl3\ 2EI \ 192 24 /" W The minus sign indicates that wq(0) is taken (by the sign convention adopted) opposite in sense to that of the reaction R. From the uniaxial deformation of the central support, it is clear that Hence, we have m(0) = -§^j-. (5.69) RLC = J_ (RL^ _ 5q0L4\ ECAC EI \ 48 384 /' { } from which wc obtain -1 R = SqpL* / Lc t L3 Y (; 3S4EI \ACEC 4SEIJ The center deflection is then given by where kc = (ECAC)/LC. Note that if the central support is rigid, i.e., Ec = oo, Eq. (a) gives the reaction R = 5qoL/S and deflection wq(0) = 0 at the midspan of a three- point supported beam. If the central support is not present, i.e., Ec — 0 (hence kc ~ 0), then R = 0 and the deflection becomes that of a simply supported beam under uniform load: Wo(0) = ^-. (e) 384£/ v '
160 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Finally, if the energy due to transverse shear force is included, the total complementary energy is L/2 y2 JfL/Z jufL rLJL yZ (f) where the factor fs is a geometric factor defined by Eq. (4.43). For rectangular cross sections it is equal to fs = 6/5. The shear force for the problem at hand is V{x) = 1 [- (<?0L -R) + 2q0x], 0 < x < |. Then Eq. (d) for the case with shear deformation becomes R _ 1 (RL3 5<70L4\ 1 (RL gpL2\ kc ~ EI \~W 384 ) + IQGA \ 4 8 ) ' (g) (h) or 5q0L4 R 384E/Ur + (- L" V1 \kc+4SEl) ' The deflection in Eq. (5.69) becomes -l W0(Q) = (*«*L + J^L) (i + *dL + JsLX , (5.7i) 0K J \3S4EI ^ SKsGAj \ 48 E/ AKSGAJ ' where Ks = \/fs is known as the shear correction factor. Clearly, shear deformation has the effect of increasing the deflection. Example 5.13 Castigliano's Theorem TI can be used to determine the reactions of an indeterminate beam more easily than using the statics and kinematics of the problem. As an example, consider the beam shown in Fig. 5.11. We wish to determine ihe reactions at the right support. Mi rtmr Rz Figure 5.11 The Indeterminate beam used in Example 5.13.
PRINCIPLE OF COMPLEMENTARY POTENTIAL ENERGY AND CASTIGLIANO'STHEOREM II 161 First we write the expression for the bending moment, *, x ** , / x\x x dM dM = 1, 0 < x < L. The complementary strain energy is given by "L M2 IE I U*= f Jo dx. Using Casligliano's Theorem II, we obtain BU* wQ(L) = dR2 -fL EI Jo M-—dx, o 3*2 dwo dx dU* 1 f1' dM , = = — / M dx. T_L dM2 EI Jo 8M2 Since the deflection and rotation are zero at x = L, we obtain •L r , ^ v „3 °-sl hft-(.T)l]"'-s(i,w'+5w'» Solving the equations, we arrive at R2 = —qoL, 20* M2 = -1«>L>. Example 5.14 Consider the structure shown in Fig. 5.12. We wish to determine the transverse deflections of points A and B, assuming linear elastic behavior with £,G, /, and / as the materia! and geometric parameters. The structure has bending and torsional deformations. First we note that the portion AB of the structure experiences only bending while the part BC experiences E, I} A, J = constant (m) P(N) Figure 5.12 The frame structure used in Example 5.14.
162 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS bending as well as torsion. Therefore, the strain energy of the structure involves computing the bending energies of the two parts and the torsional energy of part BC. The complementary strain energy (neglecting the energy due to shear) is given by For the structure at hand, we have (U* = U%B + U^c): 1 fb 1 Ca 1 fa The deflection at point A, by Castigliano's Theorem II, is given by A dU* 1 /Pb3 q0bA\ 1 /F1 Ixfl3 1 /„, qob2\ , wA = = — ( + -^— I + — (P + gob) — + \Pb + ^— ab. 0 dP EI\3%JEIK *u 3 GJ V 2 / (c) Clearly, the unit dummy-load-method would give exactly the same expression as above, except that the both sides of Eq. (c) would be multiplied by SP, In order to determine the vertical deflection at point B, we introduce a dummy vertical load SQ there (which is in equilibrium by the force SQ and bending moment aSQ at the fixed end). The portion BC of the bar is in a state of pure bending due to the dummy load SQ at B, The actual bending moment at any distance x from point B toward C is M(x) = ~(P + qob)x. The virtual moment is SMB(x) = -SQ - x. Note that the torque T = (Pb + qob2/2) docs not figure into the calculation because ST due to SQ is zero. Using the unit-dummy-load method, we write i WBSQ= [a—MSMBdx={P+J°b) (fax2dx)sQ Jo EI EI \Jq / = jp + «bW8Q (d)
PRINCIPLE OF COMPLEMENTARY POTENTIAL ENERGY AND CASTIGLIANO'S THEOREM II 163 or Wr, = (P + qob)a3 3EI (e) The answer can be easily verified by noting that the deflection at the free end of a cantilever beam of length L and end load Fq is FqL3/EI. Example 5.15 Consider the frame structure shown in Fig. 5.13. We wish to determine the reaction forces at support A using Castigliano's Theorem II. The material obeys Hooke's law, and EA and El are constant throughout. The expressions for the bending moment and axial force in each member can be expressed in terms of the reactions at point A. The sense of the local coordinate along the member length is indicated in the figures. Member AB: M|(x) = FAx- MA> Ni(x) = -NA, Member BC: Jf~a 9Mi = 0, = x, dMA dM\ ufA = 0, = -1, dMi = 0, SNa M2(x) - FAa - MA + NAx, 3N2 3M2 JFa~ = a. 8M2 Im~a = -1, dM2 ~3Na~ — x, N2(x) = FA Member CD: Po, dFA = 1, dN2 9Ma = 0, dN2 = 0. M3(x) = FA(a-x) + P0x - A/A - Mo + NAa, dM3 , 3M3 dMA dN3 3MA = -1, = 0, dNA a/v3 = a, N3(x)=NA, dNA dM3 JFa' 8N3 = a x, dFA = 0, 3NA = 1. Fa- -*i \A mffi '^ D| Figure 5.13 The frame structure used in Example 5.15.
164 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS The complementary strain energy of the structure (neglecting the energy due to shear) is given by f^k \2EI 2EAi \dx = U*(FA,NA,MA,Po,M0). Since the support is rigid (i.e., all generalized displacements are zero), we have dU* = 0, dU* MA " dFA Carrying out the integration, we obtain = 0, dU* ufA 0. 2MAa2 + NAa3 + °=MFAa3 EI V 3 0 = — (-2FAa2 + 3MAa - or, in matrix form, we have ,3 3MAa2 4NAa3 Poa3 + 2 ft*3 M0a2\ + 2NAa EA ' Mqcv 3NAa 6 2 2 A)"2 „ r Aa3 2a a 3EI ' EA EI 3 5a3 a -h in 3a2 ~2EI 3E1 EA _2^ El 3a2 2EI 2a2- El 3a_ ~El J ci J , NA FA MA P0a3 M0a2 El 2EI Poa3 Poa Mod2 ~6e7 + ^EA+ 2EI Poet2 M$a 2EI EI The coefficient matrix is known as the flexibility matrix. 5.5 BETTPS AND MAXWELL'S RECIPROCITY THEOREMS The principle of superposition is said to hold for a solid body if the displacements obtained under a given set of forces is equal to the sum of the individual displacements that would be obtained by applying the single forces separately. Consider, for example, the indeterminate beam shown in Fig. 5.14. The beam problem is equivalent to the two beam problems shown there. At point A the beam experiences the deflections Wq and Wy, due respectively to the distributed load qo and spring force Fs, Within the restrictions of the linear Euler-Bernoulli beam theory, the deflections are linear functions of the loads. Therefore, the principle of superposition is valid: A q , s CftL FSL3 SEI 3EI (5.72)
BETTI'S AND MAXWELL'S RECIPROCITY THEOREMS 165 % A t T T T T T T T ? T B EI = constant AjTT TTT Y T Tj^ ^ '* L t *~* £ *Z,K><{ Figure 5.14 Representation of an indeterminate beam as a combination of two determinate beams. Because the spring force Fs is equal to kvj$y we can calculate wfi from Eq. (5,72): ^6 = qoL* 8£/(l4-(/cL3/3£7))' (573) The procedure can be used to represent a complicated linear problem as a linear combination of simple linear problems. The principle of superposition is not valid for strain and potential energies because they are quadratic functions of displacements and/or forces. In other words, when a linear elastic body is subjected to more than one external force, the total work due to external forces is not equal to the sum of the works that are obtained by applying the single forces separately. However, there exist theorems that relate the work done by two different forces applied in different orders. The present section is devoted to the discussion of two reciprocity theorems for work done on structural systems. Consider a linear elastic solid which is in equilibrium under the action of two external forces, Fj and F2 (see Fig. 5.15). Since the order of application of the forces is arbitrary, we suppose that force Fj is applied first. Let W\ be the work produced by Fi. Then we apply force F2, which produces work W2. This work is the same as that produced by force F2, if it alone were acting on the body. When force F2 is applied, force Fi (which is already acting on the body) does additional work because its point of application is displaced due to the deformation caused by force F2. Let us denote this work by W12. Thus the total work done by the application of forces F\ and F2, Fi first and F2 next, is W = Wi + W2+ Wi2. (5.74a)
166 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Figure 5.15 Configurations of an elastic body due to the application of loads Fj and F2. — Undeformed configuration. --- Deformed configuration after the application of load F!. Deformed configuration after the application of load F2. Work W\2> which can be positive or negative, is zero if and only if the displacement of the point of application of force Fj produced by force F2 is zero or perpendicular to the direction of Fj. Now suppose that we change the order of application. Then the total work done is equal to W= W\ + W2 + W2j, (5.74b) where W21 is the work done by force ¥2 due to the application of force F\. The work done in both cases should be the same because at the end the elastic body is loaded by the same pair of external forces. Thus we have W = W or IV12 = W21. (5^75) Equation (5.75) is a mathematical statement of Betti's (1823-1892) reciprocity theorem. Applied to a three-dimensional elastic body, Eq. (5.75) takes the form / fiiii dV+ I tjiii dS= fit; dV+ £■«,■ dS, (5.76) JV JS2 Jv J S2 where Uj are the displacements produced by body forces f and surface forces t,-, and it} are the displacements produced by body forces f and surface forces f/. The left-hand side of Eq. (5.76), for example, denotes the work done by forces f) and U in moving through the displacements w/ produced by forces // and t\. We now state Betti's reciprocity theorem: if a linear elastic body is subjected to two different sets of forces, the work done by the first system of forces in moving through the displacements produced by the second system of forces is equal to the work ddne by the second system of forces in moving through the displacements produced by the first system of forces. Example 5,16 Consider a cantilever beam of length L subjected to a concentrated load Fo at the free end and to a uniformly distributed load of intensity #0 (see Fig. 5.16).
BETTI'S AND MAXWELL'S RECIPROCITY THEOREMS 167 ^0 A t JLT T T T T T T T | £/ = constant Figure 5.16 Configuration of a cantilever beam. The deflection equation due to the concentrated load alone is and the deflection equation due to the distributed load is The work done by the load Fo in moving through the displacement due to the application of the uniformly distributed load qo is Wl2 = F0wl(0) = SEI The work done by the uniformly distiibuted load q$ in moving through the displacement field due to the application of point load Fo is W2\ = f 7§r(Jc3 - 3L2* + 2L3)?0dx = Jo Ojfc/ which is in agreement with Wyi An important special case of Betti's reciprocity theorem is given by Maxwell's (1831-1879) reciprocity theorem. Maxwell's theorem was given in 1864, whereas Betti's theorem was given in 1872. Therefore, it may be considered that Betti generalized the work of Maxwell. We derive Maxwell's reciprocity theorem from Betti's reciprocity theorem. Consider a linear clastic solid subjected to force F1 of unit magnitude acting at point A, and force F2 of unit magnitude acting at a different point B of the body. Let uab be the displacement of point A in the direction of force F* produced by unit force F2, and uba be the displacement of point B in the direction of force F2 produced by unit force F1 (see Fig. 5.17). From Betti's theorem it follows that F1 • uAB = F2 • uba or uAB = uba* (5-77) Equation (5.77) is a statement of Maxwell's theorem. If ei and e2 denote the unit vectors along forces F1 and F2, respectively, Maxwell's theorem states that
168 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Figure 5.17 Configurations of the body discussed in the Maxwell theorem. W= 4,000 lb 4—\c 9ft 3ft (a) 4,000 lb 4,000 lb °bc B •^ wc\ 1 (b) *H> WCB\ (c) ^M Figure 5.18 The cantilever beam of Example 5.17. the displacement of point A in the ei-direction produced by a unit force acting at point B in the 62-direction is equal to the displacement of point B in the §2 direction produced by a unit force acting at point A in the ei -direction. We close this section with the following two examples that illustrate the usefulness of MaxwelFs theorem. Example 5,17 Consider a cantilever beam (E = 24x 106 psi, / = 120 in.4) of length 12 ft subjected to a point load of 4000 lb at the free end. We wish to find the deflection at a point 3 ft from the free end (see Fig. 5.18). By Maxwell's theorem, the displacement wjjc at point B (x — 3 ft) produced by the 4000 lb load at point C (x = 0) is equal to the deflection wcb at point C produced by applying the 4000 lb load at point B. Let wb and 0b denote the deflection and slope, respectively, at point B owing to load F = 4000 lb applied at point B. Then the deflection at point B (x = 3 ft) caused by load F0 = 4000 lb at point C (jc - 0)
EXERCISES 169 FB FA Fc FD Tff C +D . £1 T-p is +l/ aj n ?C T# , WAB WCB l^DB WB Figure 5.19 The simply supported beam of Example 5.18. is (wB = FL3/3EI and 6b = FL2/2EI): WBC = ™CB = WB + (3 X \2)0b _ 4000(9 x 12)3 (3 x 12)4000(9 x 12)2 ~ 3EI + 2EI 243 x 6000 x (12)3 24 x 106 x 120 = 0.8748 in. Example 5.18 Consider a simply supported beam subjected to a point load Fb at point B. The load produces deflections wab, wcw» and won at points A, C, and D, respectively (see Fig. 5.19a). We wish to find the deflection wb at point B produced by loads FA,FC, and FD (see Fig. 5.19b). A unit load acting at point B produces displacements wab/Fb, wcb/F#, and wdb/Fb at points A, C, andD, respectively. But, by Maxwell's theorem, wab/Fb* wcb/Fb, and wor/Fb are the displacements of point B caused by a unit load acting at points A, Cy and /), respectively. Therefore, the displacement at point B owing to forces Fa, Fc, and Fq is given by _ FAy)AB + Fcwcb + Fqwdb FB EXERCISES 5.1 Use the principle of virtual displacements to determine the governing equations and natural boundary conditions associated with the problem in Exercise 4.4. 5.2 Use the principle of virtual displacements to determine the governing equations and natural boundary conditions associated with the problem in Exercise 4.5. 5.3 Use the principle of virtual displacements to derive the equations governing the Euler-Bernoulli beam theory with geometric nonlinearity. Use a distributed transverse load of q = q(x) and assume that the displacement field is given by u{x, y, z) = mow - z— y-~r~» u = 0, w(x, y, z) = wo(x) dx dx
170 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS and the only nonzero strain is (nonlinear) t \ du° . dx 1 (dw0\2 c12wq d2wo dx2 -y dx2 Here uq and wo are the axial and transverse displacements of a point (x, y, 0) in the beam. Your answer should be expressed in terms of the area-integrated quantities [e.g., stress resultants of Eq. (5.9)]. 5.4 The Timoshenko beam theory. The Euler-Bernoulli beam theory is based on the assumption that a straight line transverse to the axis of the beam before deformation remains (i) straight, (ii) inextensible, and (iii) normal to the midplanc after deformation. In the Timoshenko beam theory, the first two assumptions are kept but the normality condition is relaxed by assuming that the rotation is independent of the slope (dwo/dx) of the beam. Using these assumptions, the displacement field of the beam can be expressed as u(x9 z) — woW + z<j)(x) v = 0 w = wo(*)» (a) 5.5 5.6 where (w, v, w) are the displacements of a point along the (x, y> z) coordinates, (wo. wq) are the displacements of a point on the midplane of an undeformed beam, and <j> is the rotation (about the y-axis) of a transverse normal line. Derive the equations of equilibrium of the Timoshenko beam theory using the principle of virtual displacements for the case of infinitesimal strains. Also derive the natural boundary conditions of the theory. Use the unit-dummy-displacement method to determine the displacement in the direction of the applied load P in the structure in Exercise 4.10. Use Castigliano's Theorem I to determine the axial forces in the wires of the structure shown in Fig. E5.6. Assume small deformation. Elastic kjj&^a^^ 1.5a 5.7 Use Castigliano's Theorem I to determine the unknown generalized displacements and forces of the beam shown in Fig. E5.7. Use the results of Example 5.7.
EXERCISES 171 P ,N >fr ^ fr y >fr ¥ y y y 1zy wQ Figure E5.7 5.8 Use the unit-dummy-load method to determine the horizontal and vertical displacements at point D of the structure in Exercise 4.9. 5.9 Use the unit-dummy-load method to determine the center deflection of a simply supported beam under uniformly distributed transverse load, q$. 5.10 Use the unit-dummy-load (moment) method to determine the rotation at x = L in Exercise 5.7. 5.11 Use the unit-dummy-load method to determine a relationship between the lip deflection we and deflection wc at the center of the beam shown in Fig. E5.11. Figure E5.11 5.12 Use the unit-dummy-load method to determine the deflection at point B of the beam shown in Fig. E5.12. Yu;n *D Figure E5.12 5.13 Use the unit-dummy-load method to determine the force in the cable of the cable-supported cantilever beam shown in Fig. E5.13.
172 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS ECJ^AC 5,14 Use the unit-dummy-load method to detemiine the rotation of joint B of the frame shown in Fig. E5.14. Consider energy due to bending only. F0=40k 9o q0= 0.5 kip/ft. a= 8 ft., b =12 ft. E= 29x10° psi. 1= 240 in/1 *4 A |L=16 ft. Figure E5.14 5.15 Use the unit-dummy-load method to determine the vertical deflection of point A of the structure in Fig. E5.15. Consider energies due lo both bending and torsion. P = 5kips Q= 10 kips a=6=5ft. #=30xl06psi. 7= 144 in/1 J= 288 in.4 Figure E5.15 5.16 5.17 Consider the frame structure shown in Fig. 4.9 of Example 4.4. Each member of the structure has the same geometric and material properties. Assume linear elastic behavior. Determine the horizontal and vertical displacements at the point of load application. Use the unit-dummy-load method. Consider the pin-connected structure shown in Fig. E4.9. Suppose that the material of all members obeys the following stress-strain relation: a = E^/s, s > 0, -E^ey s < 0,
EXERCISES 173 where E is a constant. All members have the same cross-sectional area, A. Determine the horizontal and vertical displacements at the point of load application using the unit-dummy-load method. 5.18 Solve the problem in Exercise 5.6 using Castigliano's Theorem IT. 5.19 Use Castigliano's Theorem II to determine the compressive force and displacements in the linear elastic spring (spring constant, k) supporting the free end of a cantilever beam under a triangular distributed load (see Fig. E5.19). % Z, WQ Figure E5.19 5.20 The thin curved beam shown in Fig. E5.20 has a radius of curvature Ry modulus E9 and moment of inertia / about the axis of bending. Use Castigliano's Theorem II to determine the vertical and horizontal deflections at the free end. Include only energy due to bending. x, u 5.21 5.22 The thin curved beam shown in Fig. E5.21 has a radius of curvature R, modulus E, and moment of inertia / about the axis of bending. Use Castigliano's Theorem II to determine the vertical and horizontal deflections at the free end. Include only energy due to bending. The thin curved beam shown in Fig. E5.22 has a radius of curvature R, modulus E, and moment of inertia / about the axis of bending. The radial distributed load (i.e., pressure) is constant, p per unit arc. Use Castigliano's Theorem II to determine the horizontal and vertical deflections, Include only energy due to bending.
174 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS Figure E5,21 p/unit arc Figure E5.22 5.23 Determine the horizontal displacement of the curved beam problem of Exercise 5.21 by accounting for the energy due to extension, bending, as well as transverse shear. 5.24 Use Castighano's Theorem II to determine the rotation at the free end of a spring-supported cantilever beam under uniformly distributed load (see Fig. 5.14). 5.25 Determine the transverse deflection of the free end of the cantilever beam shown in Fig. 5.16 using Castighano's Theorem II. Solve the problem (a) neglecting the energy due to shear force V(x)y and (b) including the shear energy [see Eq. (4.42); write the answer in terms of /TJ. 5.26 Determine the deflection at point C of the beam shown in Fig. E5.26 using Castighano's Theorem II. t Z,WQ Figure E5.26 5.27 Determine the slope at point A of the beam shown in Fig. E5.26 using Castighano's Theorem II. 5.28 Use Castighano's Theorem II to determine the vertical and horizontal deflections at point C of the statically determinate structure shown in Fig. E5.28.
EXERCISES 175 Assume that all connections are pin connections and that all members have the same area of cross section (A) and modulus (£"). 5.29 Suppose that the structure of Exercise 5.28 is made indeterminate by replacing the roller at point D with a connection identical to that at point A (see Fig. E5.29). Determine the reactions at points C and D. 1 A B C L^ A I $»/ >j i 1 \y p Figure E5.29 5.30 5.31 5.32 Prove Betti's theorem for a three-dimensional elastic solid. Consider a simply supported beam of length L subjected to a concentrated load Fo at the midspan and a bending moment Mq at the left end. Verify that Betti's theorem holds. Verify Maxwell's reciprocal relationship for the cantilever problem with two different loadings shown in Fig. E5.32. You must determine the required deflections and slopes to verify Maxwell's reciprocal relationship. ' A 71 -** x — a- 'A, M0 B| ztw0 Figure E5.32 5.33 5.34 Determine the deflection at the midspan of a cantilever beam subjected to uniformly distributed load qo throughout the span and a point load Fo at the free end (sec Fig. E5.33). Use Maxwell's theorem and superposition. A load P = 4000 lb, acting at a point A of a beam produces 0.25 in. at point B and 0.75 in, at point C of the beam. Find the deflection of point A produced by loads of 4500 lb and 2000 lb acting at points B and C, respectively.
176 ENERGY PRINCIPLES OF STRUCTURAL MECHANICS M H M T M T |. 1 EI = constant Figure E5.33 REFERENCES 1. Maxwell, J. C, "On the calculation of the equilibrium and the stiffness of frames," Phil. Mag. Ser. 4,27,294 (1864). 2. Engesser, F., "Ubcr statisch unbestimmte Trager bei beliebigen Formanderungsgcsetze und iiber den Satz von der Kleinsten Erganzungsarbcit," Zeitschrift des Architekten und Jngenieur-Vereins zu Hanover, 35 (1889). 3. Betti, E., "Teoria dellaJ clasticita," Nuavo Cimento, Serie 2, Tom Vll and VIK (1872). 4. Castigliano, A. C, "Intorno ai sistemi elastici," Dissertazione di Laurea presenlata all Comniissione Esaminatrice dclla Reale Scuola degli Ingegneri di Torino, Vinccnzo Bona, Turin, Italy (1873). 5. Lanczos, C, The Variational Principles of Mechanics, 4th edition, Dover, New York (1986). 6. Argyris, J. H., and Kelsey, S., Energy Theorems and Structural Analysis, Butlerworths (1954). 7. Davies, G. A. O., Virtual Work in Structural Analysis, John Wiley, Chichester, UK (1982). 8. Dym, C. L., and Shames, I. H., Solid Mechanics: A Variational Approach, McGraw-Hill, New York (1973). 9. McGuire, W., Gallagher, R. II., and Ziemian, R. D., Matrix Structural Analysis, 2nd edition, John Wiley, New York (2000). 10. Odcn, J. T., and Ripperger, E. A., Mechanics of Elastic Structures, 2nd edition, Hemisphere/McGraw-Hill, New York (1981). 11. Reddy, J. N., Energy and Variational Methods in Applied Mechanics, John Wiley, New York (1984). 12. Reddy, J. N.t Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia (1999), 13. Richards, T, H., Energy Methods in Stress Analysis, Ellis Horwood, (distributed by John Wiley & Sons), Chichester, UK (1977). 14. Shaw, F. S., Virtual Displacements and Analysis of Structures, Prenlice-Hall, Englewood Cliffs, NJ (1972). 15. Volterra, E., and Gaines, J. H., Advanced Strength of Materials, Prcnti,cc-Hall, Englewood Cliffs, NJ (1971). 16. Yang, Y. B., and Kuo, S. R., Theoiy and Analysis of Nonlinear Framed Structures, Prentice- Hall, Englewood Cliffs, NJ (1994).
6 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE 6.1 INTRODUCTION The principles of virtual work and their special cases discussed in Chapter 5 were limited to static equilibrium of solids. Hamilton's principle is a generalization of the principle of virtual displacements to dynamics of systems of particles, rigid bodies, or deformable solids. The principle assumes that the system under consideration is characterized by two energy functions, a kinetic energy K and a potential energy 1~L For discrete systems (i.e., systems with a finite number of degrees of freedom), these energies can be described in terms of a finite number of generalized coordinates and their derivatives with respect to time /. For continuous systems (i.e., systems that cannot be described by a finite number of generalized coordinates), the energies can be expressed in terms of the dependent variables (which are functions of position) of the problem. Hamilton's principle reduces to the principle of virtual displacements for systems that are in static equilibrium. The present chapter is devoted to the application of Hamilton's principle to dynamics of discrete systems as well as continuous systems. We begin the discussion with Hamilton's principle for particles. 6.2 HAMILTON'S PRINCIPLE FOR PARTICLES AND RIGID BODIES Consider a single particle of mass m moving under the influence of a force F = F(r). The path r (f) followed by the particle is related to the force F and mass m by Newton's second law of motion: F(r) = m^. (6.1) 177
178 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE A path that differs from the actual path is expressed as r-f- 5r, where 8r is the variation of the path for any fixed time t. We suppose that the actual path r and the varied path differ except at two distinct times t[ and t2, i.e., 8r(t\) = Srfa) = 0. Taking the scalar product of Eq. (6. J) with the variation <5r\ and integrating with respect to time between t\ and ?2, we obtain n-$-»] |m^4-F(r)|-*r</f =0. Integration by parts of the first term in Eq. (6.2) yields f'2 ( dr dSr r/v ,\J / dr \ = o. (6.2) (63) The last term in Eq. (6.3) vanishes because 8r(t\) = Srfo) = 0. Also, note that = SK, (6.4) dr d8r m — • —— = 8 dt dt m dr dr 2 dt dt where K is the kinetic energy of the particle: 1 dr dr 1 K = -m—- • — = -mv « v. 2 dt dt 2 Equation (6.3) now takes the form rr2 / (5AT+F-(5r)rfr = 0, (6.5) (6.6) which is known as the general form of Hamilton's principle for a single particle. Suppose that the force F is conservative (that is, the sum of the potential and kinetic energies is conserved) such that it can be replaced by the gradient of a potential: F = -grad V, (6.7) where V = V (r) is the potential energy due to the loads on the particle. Then Eq. (6.6) can be expressed in the form \\ (K - V) dt = 0, (6.8) because 'dV dV 8V grad V • <5r - -—8x.{ + ^—&x2 + -—Sx^ = 8V(x{,X2,X3). ax\ ax-2 °x3 The difference between the kinetic and potential energies is called the Lagrangian function: L = K - V. (6.9)
HAMILTON'S PRINCIPLE FOR PARTICLES AND RIGID BODIES 179 Note that for particles and rigid bodies considered here, the internal energy W> is assumed to be zero (and We = V). Equation (6.8) represents Hamilton's principle for the conservative motion of a particle. It states that the motion of a particle acted on by conservative forces between two arbitrary* instants of time t[ and ti is such that the line integral over the Lagrangian function is an extremum for the path motion. Stated in other words, of all possible paths that the particle could travel from its position at time t\ to its position at time t2, its actual path will be one for which the integral Ldt (6.10) n is an extremum (i.e., a minimum, maximum, or an inflection). If the path r can be expressed in terms of the generalized coordinates q\ (i = 1,2,3), the Lagrangian function can be written in terms of q; and their time derivatives: L = L(q\, q2, qi, q\, qi, qiY (6.11) Then from Section 4.4 the condition for the extremum of / results in the equation (note that 8qi = 0 at t\ and (2): 81 L(qu qi, 43, q\, q%> qs) dt = 0 When all q-t are linearly independent (i.e., no constraints among <://), the variations Sqi are independent for all ty except Bq\ = 0 at /[ and ft. Therefore, the coefficients of 8q\, 8q2, and 8q?> vanish separately: dq\ dt \dqij These equations are called the Lagrange equations of motion. In Section 4.4.6 these equations were also called the Euier equations. In the discussions to follow we shall refer to these equations as the Euler-Lagrangc equations. When the forces arc not conservative, we must deal with the general form of Hamilton's principle in Eq. (6.6), Recognizing that F • Sr represents the virtual work —8 Wk due to forces F, we can write Eq. (6.6) in the form / 8Kdt- \ 8WE dt = 0. (6.14) Jt\ Jtx
180 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE In this case, there exists no functional / that must be an extremum. If the virtual work can be expressed in terms of the generalized coordinates qj by 8WE = - (QiSqi + Q2Sq2 + fi3%), (6.15) where Q\ are the generalized forces, then we can write Eq. (6.14) as and the Euler-Lagrange equations for the nonconservative forces are given by 3K d_(BK^ dqi dt \dcii \ + Qi=0% / = 1,2,3. (6.17) Example 6.1 Consider the spring-mass system shown in Fig. 6.1. First, consider the case in which the mass is sliding on a smooth, frictionless surface (see Fig, 6. la). The kinetic and potential energies (measured from the unstretched position of the spring) are K = -(a-)2, V = -x2- fx. Hence the equation of motion is (L = K — V): dx dt \dx ) which yields mx+kx = /(/). Next, suppose that the surface on which the mass m slides is dry and offers resistance to motion (see Fig. 6.1b). For motion to begin, there must be a force acting on the body that overcomes the resistance to motion. The friction force is parallel to the surface, in the direction opposite to the motion, and proportional to the force normal to the surface. The force normal to the surface in this case is the weight W = mg7 and the constant of proportionality is known as the static coefficient of friction fix. a—jvw-Fh* i—wH k , ,, FS -mg (a) (b) Figure 6.1 A spring-mass system, with the mass sliding on a surface.
HAMILTON'S PRINCIPLE FOR PARTICLES AND RIGID BODIES 181 The value of jxs varies from 0 (when the surface is frictionless) to 1, depending on the surface finish and material. Once motion ensues, the value of the friction coefficient drops to fik < /x.v, known as the kinematic friction coefficient. The friction force remains constant in magnitude during the motion as long as the inertia force and the force in the spring are sufficiently large to overcome the static friction. Thus, in this case, the system involves a force that is nonconservativc. Denoting the magnitude of the friction force by Fci = mgixk, we can write the equation of motion as d fdK\ dt \dl) + Q = 0, Q = -kx - sign(i) Fd + /(0, where sign(i) denotes the sign of x (i.e., sign(i) is -hi when x is positive and — 1 if x is negative). The above equation of motion can be separated into two equations as m'x -h kx = —Fci + f(t) when x > 0, mx + kx = Fj -f f(t) when x < 0. Example 6.2 Consider the motion of a pendulum in a plane, ideally speaking, the pendulum i^ imagined to be a mass m (called the bob) attached at the end of a rigid massless rod of length / that pivots about a fixed point O, as shown in Fig. 6.2. Hence, q\ = / is fixed, q\ = 0, and 42 = 0 is the only independent generalized coordinate. The coordinate 0 is measured from the vertical position. The force F acting on the bob is the component of the gravitational force: F = mg(— sin0e<9 + cos0er). The component along er does no work (because / is a constant). The first component is derivable from the potential (VV = —F): V = - [-mgl(\ - cos0)| = mgl{\ - cos0). (6.18a) The expression signifies the potential energy of the pendulum bob at any instant of time with respect to the static equilibrium position 0=0. The gradient operator in Mm mgsmO mg Figure 6.2 Configuration of the pendulum.
182 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE the present case is given by (r = /): vt - a . *<> d 3r r 90 Thus the kinetic and potential energies are given by m ~2 K = ?(/fl)2, V = mgl{\ -cos<9), (6.18b) 8K = mrOSO, 8V = mgl sin 080. (6.18c) Therefore, the Lagrangian function L is a function of 0 and f). The Euler-Lagrange equation is given by 30 dt\d0) which yields -mgl sin (9 - 4- (m/2«) = 0 or 0 + - sin 0 = 0. (6.19) Equation (6.19) represents a second-order nonlinear differential equation governing 6. For small angular motions, Eq. (6.19) can be linearized by replacing sin 0 ~ 0: 0 + j0 = 0. (6.20) Now suppose that the mass experiences a resistance force F* proportional to its speed (i.e., the pendulum is suspended in a medium that offers resistance to motion). According to Stake's law, F* = -6ir palGeo, (6.21) where \x is the viscosity of the surrounding medium, a is the radius of the pendulum bob, and e# is the unit vector tangential to the circular path. The resistance of the massless rod supporting the bob is neglected. The force F* is not derivable from a potential function. Thus, we have one part of the force (i.e., gravitational force) conservative and the other (i.e., viscous force) nonconservative. Hence, we use Hamilton's principle expressed by Eq. (6.14) or Eq. (6.16) with 8WE = 8V -F* • (I80e0) = (mglsin0 + 671^112 6)80 = -Q80. Then the equation of motion is given by [K = K(9)]: ' -L (dK ~"di \~d0 ) + Q = Q or O + ^sm0 + ^^6^O. (6.22) / / m The coefficient c = 67tafj,/m is called the damping coefficient.
HAMILTON'S PRINCIPLE FOR A CONTINUUM 183 Hamilton's principle in Eqs. (6.8) and (6.14) can be easily extended to a system of N particles, hence to rigid bodies. We have 8 1 L(qj,qj)dt = 0 for conservative forces, (6.23) Jt{ W2 rh I 8K dt — / 8 We dt = 0 for nonconservative forces, (6,24) Jt\ J(\ where / = 1 1=1 and m.j and r,- denote the mass and path of the *th particle. Of course, Eqs. (6.23) and (6.24) apply to a rigid body, because a rigid body can be viewed as a set of particles with f\\Qd positions 17 relative to the mass center. 6.3 HAMILTON'S PRINCIPLE FOR A CONTINUUM Following essentially the same procedure as that used for discrete systems, we can develop Hamilton's principle for the dynamics of deformable bodies. The main difference between rigid bodies and deformable bodies is the presence of internal energy Wj for deformable bodies. Newton's second law of motion applied to deformable bodies expresses the global statement of the principle of conservation of linear momentum. However, it should be noted that Newton's second law of motion for continuous media is not sufficient to determine its motion u = u(x, t)\ the kinematic conditions and constitutive equations derived in Chapter 3 are needed to completely determine the motion. Newton's second law of motion for a continuous body can be written in general terms as F - ma = 0, (6.26) where m is the mass, a the acceleration vector, and F is the resultant of all forces acting on the body. The actual path u = u(x, t) followed by a material particle in position x in the body is varied, consistent with kinematic (essential) boundary conditions, to u -f <Su, where 8n is the admissible variation (or virtual displacement) of the path. We suppose that the varied path differs from the actual path except at initial and final times, t\ and ?2> respectively. Thus, an admissible variation <5u satisfies the conditions 8u = 0 on Si for all r, (6.27a) «u(x, t{) = Su(x, r2) = 0 for all x, (6.27b)
184 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE where S[ denotes the portion of the boundary of the body where the displacement vector u is specified. Note that the scalar product of Eq. (6.26) with <5u gives work done at point x, because F, a, and u are vector functions of position (whereas the work is a scalar). Integration of the product over the volume (and surface) of the body gives the total work done by all points. The work done on the body at time t by the resultant force in moving through the virtual displacement 5u is given by [ t-8udV + f i-8udS- f a : 8*Z dV9 (6.28) Jv Js2 Jv where f is the body force vector, t the specified surface traction vector, and a and e are the stress and strain tensors. The last term in Eq. (6.28) represents the virtual work of internal forces stored In the body. The strains 8 s are assumed to be compatible in the sense that the strain-displacement relations (3.20) or (3.23) are satisfied. The work done by the inertia force ma in moving through the virtual displacement <5u is given by where p is the mass density (which can be a function of position) of the medium, We have, analogous to Eq. (6.2) for discrete systems, the result f2 If p^--8udV - | f (t -Su-cr :8e)dV+ f t.8uds\\dt = Q or -f[/v4rl>+//'s"-?:8?>^i'H'''=0- (6.30) Tn arriving at the expression in Eq. (6.30), integration by parts is used on the first term; the integrated terms vanish because of the initial and final conditions in Eq. (6.27b), Equation (6.30) is known as the general form of Hamilton's principle for a continuous medium (conservative or not, and elastic or not). For an ideal elastic body, we recall from the previous discussions that the forces f and t are conservative: SV = -( f f -8udV+ f i-8uds\ , (6.31a) and that there exists a strain energy density function (Jo = Uo(£ij) such that crtf = ~. (6.31b)
HAMILTON'S PRINCIPLE FOR A CONTINUUM 185 Substituting Eqs. (6.31a,b) into Eq. (6.30), we obtain 8 [\k-(V + U)]dt = 0, (6.32) where K and U are the kinetic and strain energies: K = fylTt'ltdV- U = lVU°dV- (633) Equation (6.32) represents Hamilton's principle for an elastic body (linear or nonlinear). Recall that the sum of the strain energy and potential energy of external forces, U + V, is called the total potential energy, IT, of the body. For bodies involving no motion (i.e., forces are applied sufficiently slowly such that the motion is independent of time, and the inertia forces are negligible), Hamilton's principle (6,32) reduces to the principle of virtual displacements. Equation (6.32) may be viewed as the dynamics version of the principle of virtual displacements. The Eulcr-Lagrange equations associated with the Lagrangian, L = K — n, can be obtained from Eq. (6.32): L(u, Vu,u)dr 0 = 8 J = f2\f fp~J-diva-f J -8udV+ f (t-i)-8uds\dt, (6.34) where integration by parts, gradient theorems, and Eqs, (6.27a,b) were used in arriving at Eq. (6.34) from Eq. (6.32). Because 8\x is arbitrary for t, t[ < t < ^ and for x in V and also on 52, it follows that <r> P^-divcf-r-O inV, (635) t - t = 0 on S2. Equations (6.35) are the Euler-Lagrange equations for an elastic body. Using the results of this section and the previous sections, one can derive statements of Hamilton's principle corresponding to the principle of virtual forces (or complementary potential energy) and stationary principles. These are left as exercises for the reader. Example 6.3 Consider the axial motion of an clastic bar of length L, area of cross section A, modulus of elasticity E, mass density />, and subjected to distributed force / per unit length and an end load P. We wish to determine the equations of motion
186 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE for the bar. The kinetic and total potential energies of the system are n = / -crjjf-jjdV - / fudx - Pu(L) Jv 2 Jo = / —VxxGxx dx- I fu dx - Pu(L), Jo * Jo (6.36a) (6.36b) where u, <rJX, and exx are assumed to be functions of x and t only, and «(0,r) = 0 du S\x = dx (bar is fixed at x = 0), (strain-displacement relation). (6.37) Substituting for K and fl from Eqs. (6.36a,b) into Eq. (6.32), we obtain du 8u dt + pA—Su dt 0=i, \L r*ir-Mta)+"" -(Aaxx-P)\x=LSu(L)\dt, dx + PSu(L)\ dt dx Su dx — {Aaxx8u) f + P8u(L)\ dt o 8u dx (6.38) where Su(0, t) = 0 and 8u(xt t\) = Su(x912) = 0 are used to simplify the expression. The Euler-Lagrange equations are obtained by setting the coefficients of 8u in (0, L) and at x = L to zero separately: l(pA^\--jt.(Aoxx)-f = 0, Q<x<L, (6.39a) (A<7.„-)|t=/ -P = 0. (6.39b)
HAMILTON'S PRINCIPLE FOR A CONTINUUM 187 for all/,*i < t < t-2. For linear clastic materials, we have crxx = Esxx = E(du/dx), and Eqs. (6.39a,b) become _a^ St Kr)-s(Ms)-'-ft 0<"<L' <6'40a) K) - P = 0. (6.40b) x=L Now suppose that the bar also experiences a nonconservative (viscous damping) force proportional to the velocity: F* = -pljL, (6.41) where \x is the damping coefficient (a constant). Then the Euler-Lagrange equations from Eq. (6.30) are given by a Yt ( 3u\ a / a \ a« r ar) " ^ (£A W - f + " » = °' 0<X<L' (6-42a) |A£—) -P = 0. (6.42b) V 9xJx=L The next example illustrates the use of Hamilton's principle to derive the equations of motion associated with an assumed displacement field of a refined theory of straight beams. The displacement field is often arrived at by placing certain kinematic hypotheses on the system (see Reddy [7]). Example 6A (Third-order beam theory) Consider the displacement field n( SWq\ u(xt z, t) = uq(x, t) + z.(f>(x, t)-c\z \<j> + —— , V 9x ) (6.43) w(x,Zy t) = U)q(x, t)> where c\ = 4/(3/?2), «o is the axial displacement, wq the transverse displacement, and 0 the rotation of a point on the centroidal axis x of the beam. The displacement field is arrived at by (a) relaxing the Euler-Bernoulli hypotheses to let the straight lines normal to the beam axis before deformation become (cubic) curves with arbitrary slope at z = 0, and (b) requiring the transverse shear stress to vanish at the top and bottom of the beam. Thus, the only restiiction from the Euler-Bernoulli beam theory that is kept is w(x, z, f) = wq(x, t) (i.e., transverse deflection is independent of the thickness coordinate z). The displacement field (6.43) accommodates quadratic variation of transverse shear strain exz and shear stress crxz through the beam height, as can be seen from the strains computed next. Now suppose that the beam is subjected to a distributed transverse load of g(x, t) along the length of the beam. Since we are primarily interested in deriving the equations of motion and the nature of the boundary conditions of the beam that experiences
188 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE a displacement field of the form in Eq. (6.43), we will not consider specific geometric or force boundary conditions here. The procedure to obtain the equations of motion and boundary conditions involves the following steps: (i) compute the strains, (ii) compute the virtual energies required in Hamilton's principle, and (iii) use Hamilton's principle and derive the Euler-Lagrange equations of motion and identify the primary and secondary variables of the theory (which in turn help identify the nature of the boundary conditions). Although one can use the general nonlinear strain-displacement relations, here we restrict the development to small strains and displacements. The linear strains associated with the displacement field are where ^XX — *'XX + Z8XX -h Z S.vjt , Yxz — Yxz * <• Yxz > (0) _ ^0 (1) __ 90 (3) _ _,. (*± , 32«X) (6.44a) *xx " dx ' *xx ~ dx ,0) = _c. (** + ^M *«?=* + a , 3W<> (2) /a . dW°\ (6.44b) and C2 = 4/A2. Note that j/YZ = 2£vz is a quadratic function of z. Hence, axz = Gy.rz is also quadratic in z. From the dynamic version of the principle of virtual displacements (i.e., Hamilton's principle for deformable bodies) we have 0 = jT jf'' fA [a„ («,«? + a*™ + e3^) + <*« («y<? + Ay®)] rfArfxA -rr//{[----'(-^)] x «ii0 + z<$0 - ctz3 f 50 4- —^ J + w0Swo IdAdxdt fT fL — I I qSwo dxdt Jo Jo 0 JO = fj (nxxSs^ + Mxx8e™ + PxxSs^ + Qx8y^ + RxSY^)dxdt - I I I IquqSuq + \h4> ~ c\ ^(0+ -y- ) \W + <lSwo \ dxdt
HAMILTON'S PRINCIPLE FOR A CONTINUUM 189 -a:\{ 9NXX d2uo\x 'dx dt2 ■)■ + + UwQH d2w0- + /„>" " dt2 Sua + MxxS<p - c\Pxx dxdt dSwo dx + [Q-+c' (it - J'w+" *eS)] '-}„'* <"» where all the terms involving | • ]J vanish on account of the assumption that all variations and their derivatives are zero at f = 0 and t = T, and the new variables introduced in arriving at the last expression are defined as follows: Nxx Mxx Pxx -C\i\*""- {fcl-Cteh*- <6-46a) Mxx=Mxx-c}Pxx, Qx = Qx-c2Rx, ci = 3jj5. C2 = ^2 • (-6A6h^ J4 = U - c\h, K2 = h- 2ci U + c, 1$, r-h/2 f" It = / p(zY dz. (6.46c) J-h/2 Note that /,- are zero for odd values of i (i.e., 1\ = 1$ = 1$ = 0). Thus, the Euler-Lagrange equations are dNju 32«o BQX d2Pxx 8wq: -7— + c\ n , + q dx dx2 T a2uJ0 - /, a-V r B4w0 \ .. 3A?« ~ „ aV 93w0 "' ^x--Qx = K2WClJ4M^- (6.47) (6.48a) (6.48b)
190 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE The last two lines of Eq. (6.45) include the boundary terms, and they show that the primary variables of the theory are (those with the variational symbol) uq, wq, <t>, and dwo/dx. The corresponding secondary variables are the coefficients of 8ito> 5it»o, <50, and dSwo/dx: xx • (6.49) When c\ =0 in Eq. (6.43), it corresponds to the displacement field of the Timoshenko beam theory. Thus, the equations of motion of the Timoshenko beam theory can be obtained directly from Eqs. (6.47)-(6.48a,b) by setting c\ = t'2 = 0: 8NXX T 32it0 dQx dx BMXX dx dx + q -Qx = '°-9^ 92u>0 d2<p = /29^' (6.50) (6.51a) (6.51b) The primary and secondary variables of the Timoshenko beam theory are: (wo* ^o, <t>) anA(Nxx,Qx,Mxx). 6.4 HAMILTON'S PRINCIPLE FOR CONSTRAINED SYSTEMS The kinematic relations that restrict a motion are called constraints. When the relations are between generalized coordinates, and possible position and time, in the form G(/, q) = 0 for a discrete system, (6.52a) G(x, f, u, grad u) = 0 for a continuous system, (6.52b) they are called holonomic constraints. When the constraints are of inequality type or relate q\ to qi, they are called nonholonomic constraints. We consider only holonomic constraints here. As discussed in Section 4.4.8, constraints can be included into the Lagrangian function either by means of the Lagrange multipliers or by the penalty function method. To illustrate the ideas for a dynamical system, we consider the motion of a single particle whose motion is constrained by holonomic constraints: GiC. <7l, 42, 43) = 0, G2(f, qu qi, qi) = 0. (6.53)
HAMILTON'S PRINCIPLE FOR CONSTRAINED SYSTEMS 191 The variation of these equations yields 3GU 3Gi 3G\ -r—8qi + ^Sqi + -r^Sqs = 0, (6.54a) dq\ dq2 3q3 3G2 „ 3G2 3G2 —*</i + -r-^2 + T-^fcfe = 0. (6.54b) oq\ oq2 3q3 Since there arc two constraints and three degrees of freedom, only one of the variations 8q\, Sq2, and 8q?> is independent and the other two are related to the independent ones by Eq. (6.53). We multiply Eq. (6.54a) by the Lagrange multiplier A,i and Eq. (6.54b) by X2> integrate each equation over t\ to t2, and add the results to Eq.(6.16): Since the variations 8q\, 8q2i and 8q^ are not all independent, we cannot use the usual argument to set the coefficients of 8qi (i = 1, 2, 3) to zero. However, the Lagrange multipliers k{ and X2 are arbitrary. Therefore, we choose k± and X2 such that the coefficients of two of the variations out of 8q\, 8q2i and 8q$ vanish. Since the remaining variation is linearly independent, its coefficient should be zero. Thus we obtain dK d /3K\ SGl 3G2 for / = 1, 2, 3. Equations (6.56) together with (6.53) provide five equations for the five unknowns (21, ?2,43. M» X2). In the penalty function method, Hamilton's principle in Eq. (6.14) is modified to read f 2 8KP dt- f2 8WEdt + 8 f2 {^G\ + ^G|) dt = 0, (6.57) where y\ and y2 are the penalty parameters, and Kp is the kinetic energy of the constrained system. Since the constraint conditions (6.53) are now included, in the least-squares sense, we suppose that the variations 8q\, 8q2, and 8qi are independent of each other. Performing the variation indicated in Eq. (6.57), we obtain (after the usual argument) the Euler-Lagrange equations dKp d (3KP\ 3G] 3G2 P ' P X + Qi+Y\G{-^ + y2G2~=0. (6.58) li dt \ 3q\ ) 3cjj dt \ 8qj ) 3q\ 3qi A comparison of Eq. (6.58) with Eq. (6.56) (Kp = K) shows that the Lagrange multipliers X\ and k2 can be computed from *i = yiGi, k2 = y2G2, (6.59)
192 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE where G\ and Gj are given by Eq. (6.53) for the generalized coordinates q\ computed from Eq. (6.58); G\ and G2 thus computed are not identically zero because q\ computed from Eq. (6.58) are different from the true values, the error being inversely proportional to the penalty parameters y\ and 3/2. Example 6.5 To illustrate the ideas presented in the preceding paragraphs, we reconsider the damped motion of a pendulum of Example 6.2. The generalized coordinates are q\ = r, C12 = 0, (6.60a) and the constraint is Gx(qi) = qi-l = Q. (6.60b) The work done by external forces in moving through virtual displacement <5q is given by SWe = —F • <5q = — [mg(- sin 0e# 4- cos0er) - 6nfiarGeo] • (Sq\er + r&q2^e) = —[ro£ cos0<fyi - r(mg sin6 + 6njjLar0)8q2] = —(Qi&qi + Qibqi) • (6.61) The ldnetic energy is given by K = \ [(rf + (lef] = I (q\ + l2ql) = \m{W)\ (6.62) For the Lagrange multiplier method, the Euler-Lagrange equations arc obtained by substituting Eq. (6.62) and Q\ from (6.61) into Eq. (6.56): —mr + mgcos0 -f X\ =0, (6.63a) -ml26 - r(mg sin 0 + 6niiar9) = 0. (6.63b) In view of the constraint condition (6.60b), we have r = 0, and Eqs. (6.63a,b) become \{ = -mg cos0, 0 + -^^0 + f sin 0 = 0. (6.63c) m I The Lagrange multiplier X\ can be inteipreted as the force exerted on the pendulum bob by the massless rod. It is the force necessary to oppose gravity and maintain the motion of the bob in a circular arc. In the penalty function method, the Euler-Lagrange equations arc obtained by substituting Eqs. (6.61) and (6,62) into Eq. (6.58): mg cos 0 + y\ (r - /) = 0, (6.64a) -ml20 - r (mg sin 0 + 67tfiar6) = 0. (6.64b)
HAMILTON'S PRINCIPLE FOR CONSTRAINED SYSTEMS 193 A comparison of the first equation with the first equation in (6.63) shows that the approximate Lagrange multiplier is given by Myi) = Yi(r -0 = -nigcos0. The error in the constraint is given by mg cos 0 ■/ = ■ which goes to zero as y\ goes to infinity. Y\ The ideas discussed above for particles and rigid bodies can be readily extended to deformable bodies by accounting for the internal energy due to deformation. In particular, Eqs. (6.56) and (6.58) take the following forms: Lagrange multiplier method ,c 3 fdLc\ 3 /dLc\ „ 3Gi ' dG2 „ „„, 7 dt\dcjij dx\dqitX/ dqi dqt Penalty method ■H-"57"^"-5-r^+ Qi + wGi-r1 + K2G2—^ = 0. (6.66) 3qf dt \ dqi J dx \dq^T J dqi dqi where q^x = (dqi/'dx), and Lc = Kc — (Uc -f Vc). The quantities KC9 UCi and Vc arc the kinetic energy, strain energy, and the potential energy due to conservative loads, respectively, of the constrained system. The next example illustrates the use of Eqs. (6.65) and (6.66). Example 6.6 The Lagrange function associated with the dynamics of an Euler- Bernoulli beam is given by L = K — {U + V), where dAdx 2" K= fL tie.(dJ^_ ^i\ +1Cdw°\ Jo JA[2\dt zdxdt) +2 V at ) = fL [pA /auo\2 pl_ /a2w0\2 pA /dwp\ Jo [ 2 \ at J 2 \3xat) 2 \ at ) - fL \EA (3uA2 EL (d2w°\ ~J0 [2 [dxj'+T\d^~) V = - f ' [f(x, Jo dx, (6.67a) 8x ) ' 2 >t)uQ + q(x,t)wo\dx. (6.67b) (6.67c)
194 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE Here, «o denotes the axial displacement and wq the transverse displacement, which are functions of x and /, and / and q are the axial and transverse distributed loads. In arriving at the expressions for K and £/, we have used the fact that the x-axis coincides with the geometric centroidal axis, fA z dA = 0. Note that there is no constraint among the two dependent variables c/[ — «o an<i q2 = wo, and L = L(uq, wo, Mq, Wq\ «o, ^0, Wq). Hence, the Euler-Lagrange equations of the Euler-Bernoulli beam theory are given by Swq: 8 wo: dL duo SL dwo _a_ Ft \duoJ dx \du0J a (sl\ _^_(^\ °2 ('aL\_0 dt\dwo) + dx2\dwz) + dxdt\dw'0)~ ' Note that ^— = /. ttt- = Mk0. —- = -EAuQ, Olio "Uq "wn dL = <7> dwo Hence we have —- = pAwo, dwo dL OWn &--"* 8t V 9t J dx\ dx J J 3 ( 3w0\ 32 fe,91m\ d2 ( J2wo\ Now suppose that we introduce the function <£(x, t) such that ox (6.68a) (6.68b) (6.69) Now the problem can be looked at as one of minimizing L subject to the constraint in Eq. (6.69). We wish to determine the equations of motion of the beam using (a) the Lagrange multiplier method and (b) the penalty method. First we write out the expressions K, £/, and V for the constrained problem. We have {V = Vc): v fL \pA fduo\2 ^ pi (dct>\2 pA (dwo\7 K^J0 [t\it) +tU; +tUtJ _ fL\EA fdu0\2 EI fd<p\2 dx, (6.70a) (6.70b) and LC = KC- (Uc + Vc).
HAMILTON'S PRINCIPLE FOR CONSTRAINED SYSTEMS 195 (a) Lagrange multiplier method: We have L, = Lc+f k(f + ^f\dx. (6.7la) Hence, the Euler-Lagrange equations are "°- du0 Sx \du'Q) dt \duo) dw0 dx\dw'0J dt \dwoJ dcp dx \dfyj dt \d4>o/ >ki IT-0- We have 5wo: q-yx-jt {pA-dT) = °- ^ a* \ 3x/ 8f \ dt) a^o (b) Penalty method: In this case we have Lp — Lc ~f- UX*+S)'*- Note that Lp = Lp(uq, wq, <j>, u'q, WQy 0\ iiQy wo> < Hence, the Euler-Lagrange equations are *» '+h("%)-h{>*%)-°-
196 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE Swq: q — 6.5 RAYLEIGH'S METHOD For natural vibration of conservative systems, the Lagrangian function consists of only the strain energy and kinetic energy, L = K — U. For an ideal mechanical system (i.e., without thermal and other dissipative effects) the kinetic energy is the maximum when the strain energy is zero, and vice versa. Thus, in the absence of other energies, the principle of conservation of energy requires that K-max — Umax* (6.72) Equation (6.72) merely implies that the maximum kinetic energy has the same value as the maximum strain energy. However, at any given instant, the two energies are not equal to each other but their sum is constant. For example, for the spring-mass system shown in Fig. 6.3, we have (6.73) "-max — 2™{*max' ' "max — 2 ^max' * If the system is in a state of natural vibration, then x — A sin cot, xwax = A, where A is the amplitude and co is the frequency of natural vibration. Therefore, x = Ao> cos <ott xmax = Aco. Then ±mA2a)2 = A*A2 -> a)2 = ~. z z m Thus, given the stiffness of the spring and mass attached to it, we can calculate the natural frequency. Fig u re 6.3 Asp ring-mass system.
RAYLEIGH'S METHOD 197 Extending the discussion to natural vibration of a continuous system, say to an elastic bar, we have where p is the mass density (measured per unit volume), E the Young's modulus, and A is the area of cross section of the bar. Fot natural vibration, we have u(x, t) = u„(x) sin(<w„r + 0/7), ii(x, t) = un(x)o)cos(a)nt 4- 0„). Hence, pAu^co* cos2 ((L>nt + 9n) dx, 2 fL . 2 , fL T,*(dun\ o Jn EA{dun/dx)2dx o)l / pAul dx = EA —- dx or col = ^ JL^IZ—i , ^0 Jo \dx LLoAuldx Equating the maximum values of K and U (Kmax = Umax), we obtain rL it PAlll dx (6.75) The right-hand side of Eq. (6.75) is known as the Rayleigh quotient (see [9]) applied to a bar. With an appropriate choice of the eigenfunctions uu (x), the Rayleigh quotient allows us to compute the eigenvalue ajfr A suitable candidate for u„ (x) is the function that is sufficiently differentiable as required in Eq. (6,75) and satisfies the geometric boundary conditions of the problem, As we shall see in Chapter 7, the Rayleigh quotient is a special case of the Ritz approximation. The Rayleigh quotient is good for the estimation of fundamental frequencies. Example 6.7 We wish to use the Rayleigh quotient to estimate the fundamental frequency of vibration of a cantilever beam. For this case, the Rayleigh quotient takes the form 2 f0LEI(dh»„/dx2)2dx " = fl a 2 J ' (6J6) JO PAwndx where w„(x) is the eigenmode associated with the transverse motion of the beam. A suitable candidate for the eigenfunction wi is provided by the deflection of the beam under its own weight. For a cantilever beam, the deflection is given by w0(x) = pA(x4 - 4a-3I + 6x2L2).
198 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE Therefore, we choose wn(x) = x4 — 4x3L + 6x2L2> and compute the Rayleigh quotient ^El[\2{x-L)2fdx $ pA(x4 - 4x*L + 6x2L2)2dx EI 144 x 63 pAL4 728 <wi = The exact fundamental frequencies can be obtained by solving the transcendental equation cos X cosh X + 1 — 0, The first root of this equation is \\ = 1.875, giving co\ = (3.5\6/L2)<s/EI/pA. Thus, the solution estimated by the Rayleigh quotient is very accurate. A higher-order frequency can be obtained only if we can find an appropriate candidate for the corresponding eigenfunction. For example, in the case of a simply supported beam, the eigenfunction associated with the nth mode is wn(x) = sinOiTix/L), which can be used in the Rayleigh quotient to determine cou — (nn/L)2. This turns out to be the exact value of the frequency because of the selection of the exact eigenfunction for the simply supported beam. EXERCISES 6.1-6.3 Derive the Lagrangian function for the linear spring, mass, and linear dash- pot systems shown in Figs. E6.1-E6.3. Assume that motion starts from rest and that the contact surfaces are free of friction. The constitutive equations for the spring and dashpot are given by equations of the form Force = k x displacement, for springs {k — spring constant), and Force = rj x velocity, for dashpots (rj = dashpot constant). Figure E6.1 Rigid bar Figure E6.2 -o—^Ffr) Figure E6.3
EXERCISES 199 6.4 Derive the equations of motion for the double pendulum of Exercise 4.1, 6.5 Derive die equations of motion for the rigid-body assemblage of Exercise 4.2. 6.6 Consider a pendulum of mass mi with a flexible suspension, as shown in Fig. E6.6. The hinge of the pendulum is in a block of mass mi, which can move up and down between the frictionless guides. The block is connected by a linear spring (of spring constant k) to an immovable support. The coordinate x is measured from the position of the block in which the system remains stationary. Derive the Euler-Lagrange equations of motion for the system. Unstretched length of the spring Figure E6.6 6,7 Consider a block of mass mi sliding on another block of mass m\, which in turn slides on a horizontal surface, as shown in Fig. E6.7. Using u \ and itj as coordinates, obtain the equations of motion. Assume that all surfaces are frictionless. «zA \m2\s Ah"1' wmm. —>u B Figure E6.7 6.8 Figure E6.8 represents a double pendulum that is suspended from a block which moves horizontally with a prescribed motion, x = f{t). Derive the Euler-Lagrange equations of motion of the pendulum when it oscillates in the (*» y)-pl'dne under the action of gravity and prescribed motion. 6.9 Two masses m\ and mi are attached to the ends of an inextensible cord that is suspended over a frictionless stationary pulley, as shown in Fig. E6.9. Find the equations of motion of the system. 6.10 Repeat Exercise 6.9 for the case in which a monkey of mass m$ is climbing up the cord above mass m\ with a speed vq relative to mass m\ (see Fig. E6.10). 6.11 Determine the motion of all masses in the suspended double-pulley problem represented in Fig. E6.11.
200 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE *=M Figure E6.8 "#■ Figure E6.9 tHtl J- 0 m2 ®'™i Figure E6.10 4m 2m Figure E6.11 6.12 Derive the equations of motion of the system shown in Fig. E6.12. Assume that the mass moment of inertia of the link about its mass center is J — mtf, where Q is the radius of gyration. 6.13 Consider a cantilever beam supporting a lumped mass m at its end (/ is the mass moment of inertia), as shown in Fig. E6.13. Derive the equations of motion and natural boundary conditions for the problem (see Example 6.6).
EXERCISES 201 *-x Figure E6.13 o 771, J 6.14 Derive the linear equations of motion of the Timoshenko beam theoiy, starting with the displacement field u(x9 z, t) = «ot*i /) + z<f>(x, t)9 v = 0, w(x, z,0 = wo(x9t). Assume that the beam is subjected to distributed axial load f(x9 t) and transverse load q(x9 t)9 and that the a*-axis coincides with the geometric centroidal axis. Compare your results with Eqs. (6.50M6.51a,b). 6.15 Express the equations of motion of the Timoshenko beam theoiy, Eqs. (6.5 la,b), in terms of the generalized displacements (wo, 4>)- Assume linear clastic behavior. In particular, show that h[<H%+*). + /o 92wp di2 = ?. 6.16 Rewrite the equations of motion of the Timoshenko beam theory from Exercise 6.15 solely in terms of the transverse deflection wq. 6.17 Consider the Euler-Bernoulli beam theory, whose displacement field is given by dwo U(X, Z, 0 = Mot*, 0- Z-r—, ax v — 0, w(x9 z, t) = wq(x9 t). Assume that the beam has two types of viscous (velocity-dependent) damping: (1) viscous resistance to transverse displacement of the beam, and (2) a viscous resistance to straining of the beam material. If the resistance to transverse
202 DYNAMICAL SYSTEMS: HAMILTON'S PRINCIPLE velocity is denoted by c(x), the corresponding damping force is given by #£>(*» 0 = c(x)wq. ^ the resistance to strain velocity is cSy the damping stress is a^x = csexx. Derive the equations of motion of the beam with both types of damping. 6.18 Derive the equations of motion of the Euler-Bernoulli beam theory (sec Example 6.6) when the following nonlinear strain £xx is used: "* ~ dx 2 V dx ) d2W0 6*19 Suppose that the displacements («i, «2, "3) along the three coordinate axes x 1 = x, X2 = v, and X3 = z of a point (x ty,z) in the beam can be expressed as r dw LClaJ Mi (JC, Z, 0 = m(a\ 0 + Z\Ci—+ C20(A% 0 M2(Jf,z,0 = 0, (a) «3(jc,z,0 = iu(a:,0i where (cj, 6*2) are constants, («i, W2, "3) denote the total displacements along the x = jci , y — X2> and z — A'3 coordinates, respectively, of a point (xy j>, z) in the beam, u(x, t) is the displacement of a point on the midplane, w(x,t)is the transverse deflection of a point on the midplane, and <f> (x, t) is the rotation of a transverse normal about the v-axis. Assume that the beam is subjected to distributed transverse load q(x, t) and axial load f(x, f), both measured per unit length. (1) Compute the linear strains, using the strain-displacement relations. (2) Use the dynamics version of the principle of virtual displacements and the following definitions of stress resultants to (i) derive the governing equations of motion (i.e., the Euier-Lagrange equations) and (ii) identify the primary and secondary variables of the theory: Nx = / axxdA% Qx- / <rxzdAt Mx = / z-<rxxdA J A J A J A (b) Here Nx denotes the axial force, Qx is the transverse (shear) force, and Mx is the bending moment. Your answer should be in terms of the stress resultants defined in Eq. (b). 6.20 Derive the equations of motion of the Timoshenko bearn theory when the strains exx and exz are given by duo , i f9wo\2 d<j>
REFERENCES 203 6.21 Consider a rectangular membrane of dimensions a by b, and mass density p, and subjected to distributed transverse load f(x, /). The membrane is fixed at all points of its boundary. If the strain energy stored in the membrane is given by »-rn[(£),+®' clxdy, where 7b is the tension in the membrane and u is the transverse displacement, determine the equation of motion. 6.22 Use the Rayleigh quotient to estimate the fundamental natural frequency of a simply supported beam. Use the static deflection of the beam under uniform load as the eigenfunction. 6.23 Use the Rayleigh quotient to estimate the fundamental natural frequency of a fixed-hinged beam. Use wi(x) = x2(2x2 - 5xL + 3L2). 6.24 Use the Rayleigh quotient to estimate the fundamental natural frequency of a fixed-fixed beam. Use w\(x) = x2(L -x)2. REFERENCES 1. Lanczos, C, The Variational Principles of Mechanicsy 4th edition, Dover, New York (1986). 2. Hildebrand, F. B„ Methods of Applied Mathematics, 2nd edition, Prentice-Hall, Englewood Cliffs, NJ (1965). 3. Meirovitch, L., Analytical Methods in Vibrations, Macmillan, New York (1967). 4. Petrov, I. P., Variational Methods in Optimal Control Theory, translated from the 1965 Russian edition, Academic Press, New York (1968). 5. Reddy, J. N., and Rasmussen, M. L., Advanced Engineering Analysis, John Wiley, New York (1982); reprinted by Krieger, Melbourne, FL (1990). 6. Rcddy, J. N., Energy and Variational Methods in Applied Mechanics, John Wiley, New York (1984). 7. Reddy, J. N.} Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia (1999). 8. Rayleigh, Lord (John William Strutt), Theory of Sound, 2nd edition, Vol. I, Macmillan, London (1894); reprinted by Dover, New York (1945). 9. Clough, R. W., and Pcnzien, 1.; Dynamics of Structures, McGraw-Hill, New York (1975). 10. Weaver, Jr. W., Timoshenko, S. P., and Young, D. H., Vibration Problems in Engineering, 5th edition, John Wiley, New York (1990).
7 DIRECT VARIATIONAL METHODS 7.1 INTRODUCTION In Chapters 5 and 6 we saw how energy principles can be used to obtain governing equations, associated boundary conditions, and, in certain simple cases, solutions for displacements and forces at selective points of a structure. However, the energy methods considered in Chapters 5 and 6 cannot be used, in general, to determine continuous solutions to complex problems. The present chapter deals with approximate methods that employ the variational statements (i.e., either variational principles or weak formulations) to determine continuous solutions of problems of mechanics. Recall that the energy principles contain, in a single statement, the governing equation(s) and the natural boundary condition(s) of the problem. The energy principles involved setting the first variation of an appropriate functional with respect to the dependent variables to zero. The procedures of the calculus of variations were then used to obtain the governing (Euler-Lagrange) equations of the problem. In contrast, the methods described in this chapter seek a solution in terms of adjustable parameters that are determined by substituting the assumed solution into the functional and finding its extremum or stationary value with respect to the parameters. Such solution methods are called direct methods, because the approximate solutions arc obtained directly by using the same variational principle that was used to derive the governing equations. The assumed solutions in the variational methods are in the form of a finite linear combination of undetermined parameters with appropriately chosen functions. This amounts to representing a continuous function by a finite linear combination of functions. Since the solution of a continuum problem in general cannot be represented by a finite set of functions, error is introduced into the solution. Therefore, the solution obtained is an approximation of the true solution for the equations describing a 204
CONCEPTS FROM FUNCTIONAL ANALYSIS 205 physical problem. As the number of linearly independent terms in the assumed solution is increased, the error in the approximation will be reduced, and the assumed solution converges to the desired solution. The equations governing a physical problem themselves are approximate. The approximations are introduced via several sources, including the geometry, the representation of specified loads and displacements, and the material behavior. In the present study, our primary interest is to deteimine accurate approximate solutions to appropriate analytical descriptions of physical problems. The variational methods of approximation described here include the classical methods of Ritz, Galerkin, and Petrov-Galerkin (weighted residuals). Examples of applications of these methods are drawn from the problems of bars, beams, torsion, and membranes [1-23]. Applications of these methods to circular and rectangular plates are considered in Chapter 8. We begin with some mathematical preliminaries. 7.2 CONCEPTS FROM FUNCTIONAL ANALYSIS 7.2.1 General Introduction Before we discuss the variational methods of approximation, it is useful to equip ourselves with certain mathematical concepts. These include the vector spaces, norm, inner product, linear independence, orthogonality, and linear and bilinear forms of functions. Since the objective of the present study is to learn about variational methods, we limit our discussion here only to concepts that are pertinent in the context. The reader already familiar with these concepts may browse through the section to gain familiarity with the notation. Others who do not wish to burden themselves with the formalism of functional analysis may skip this section; it would not prevent them from gaining an understanding of the main ideas of variational methods. A set X is any well-defined collection of things, which are called members or elements of X In the present study we are concerned with collections of numbers, sequences, functions, and functions of functions. Examples of sets are provided below: 1. The set ))i of all real numbers. 2. The set C[0, L\ of all real-valued continuous functions f(x) defined on the closed interval 0 < x < L. 3. The collection of all closed intervals, /,• = [*/, x/+i ], on the real line. The following notation, very standard in mathematics, is adopted here: C means "a subset of (£ means "not a subset of € means "an element of £ means "not a clement of (7.1)
206 DIRECT VARIATIONAL METHODS V means "for all" 3 means "there exists" 3 means "such that". One way of defining a set S is to specify two pieces of information: (1) assume that each element of S is an element of a universal set (i.e., a well-known set), say X, and (2) list the properties that elements of the universal set must satisfy in order to be in S. For example, let X be the set of all sequences of complex numbers x = {x\, JC2, *3, • • •} and S be all elements of X possessing the property £|*„| < oo. We shall use the following notation Ioo x eX: ^2\xn\ < oo /7 = 1 (7.2) which is read "5 is the set of all elements of X such that (the colon stands for 'such thatJ)i:^il^l<cx>." A set A c Sft is said to be bounded from above if there exists a real number fi such that a < \x for all a € A. The real number \x is said to be an upper bound of the set A, Similarly, a set A is said to be bounded from below if there exists a real number y such that a > y for all a e A. The real number y is said to be a lower bound of the set A, If a set A is bounded from above and from below, we say that A is bounded. An upper (lower) bound M(m) for A is said to be the maximum (minimum) of A C ))i if M G A (m e A). It should be noted that even a bounded set need not have a maximum or a minimum. Every nonempty set of real numbers bounded from above has a "least upper bound," and every nonempty set of real numbers bounded from below has a "greatest lower bound." The least upper bound of a set A is denoted by sup A ("supremum of A"), and the greatest lower bound of A is denoted by inf A ("infimum of A"). 7.2.2 Linear Vector Spaces As wc have seen in Chapter 2, the term vector is used often to imply a physical vector that has "magnitude and direction" and obeys certain rules of vector addition and scalar multiplication. These ideas can be extended to functions, which are also called vectors, provided that the rules of vector addition and scalaj: multiplication are defined. While the definition of a vector "from a linear vector space" does not require the vector to have a magnitude, in nearly all cases of practical interest the vector is endowed with a magnitude, called the norm. In such cases the vector is said to belong to a normed vector space. We begin with a formal definition of an abstract vector space.
CONCEPTS FROM FUNCTIONAL ANALYSIS 207 A collection of vectors, «, v, u;, ... is called a real linear vector space V over the real number field SR if the following rules of vector addition and scalar multiplication of a vector are satisfied by the elements of the vector space. Vector Addition To every pair of vectors u and v there corresponds a unique vector it 4- v e V, called the sum of u and v, with the following properties: (la) u + v = v + u (commutative); (lb) (it -h v) + w = u + (v + w) (associative); (lc) there exists a unique vector, O, independent of u such that u + £> = u for every it e V (existence of an identity element); (Id) to every u there exists a unique vector, — u (that depends on w), such that u ■+- (—u) = ® for every it e V (existence of the additive inverse element). (7.3) Scalar Multiplication To every vector u and every real number a e d\ there corresponds a unique vector au e V, called the product of u and a, such that the following properties hold: (2a) a (fill) = (ap)u (associative); (2b) (a + p)u = aw + /3 m (distributive w.r.t the scalar addition); (2c) a(u + v) = an + av (distributive w.r.t. the vector addition); (2d) l-« = «. 1. (7.4) Note that in order to prove that a set of vectors qualifies as a vector space, one must define the identity and inverse elements and prove the "closure property" u + v € V and ait e V for all w, v e V and a e sJi. A subset S of a vector space V is called a subspace of V, denoted 5 C V, if 5 itself is a vector space with respect to vector addition and scalar multiplication defined over V. Example 7.1 1. The set of ordered n-tuples (x\, x%, A3,..., jc„) of real numbers x\, A2,..., jc„ is called the Cartesian space, denoted Sft". A typical element of SB" is denoted x = (a*i, *2, *3, •..»*n). The Cartesian space is a linear vector space with respect to the usual rules of addition and scalar multiplication: Vector addition: x + y = (*i + )'i, *2 + J'2,..., xn + y„) Vx, y <= JR". Scalar multiplication: ax = (qtaj , ax2>.... aA„) Vx € $\" and a e 31. The identity element is 0 = (0,0,0,...) (n zeros) and the inverse element is the negative of the vector.
208 DIRECT VARIATIONAL METHODS 2. Let V be the set of all polynomials in x with real coefficients. A typical element of V is of the form p(x) = an -f a\x -f a-ix2 H , where ao, a\,... are real numbers. Then V is a linear vector space with respect to the usual rules of addition and scalar multiplication. Also, the set Vn of polynomials of degree less than or equal to degree n is also a linear vector space, as can be verified (by the closure property). Moreover, Vn C P. However, the set of polynomials of degree equal to n is not a vector space as the closure property is violated. For example, consider the set of all cubic polynomials. The sum of p\ (x) = 1 - 2x + 3x2 + 6x3 and p2 = -3 + 5x 4- 2x2 - 6x3 is not a cubic polynomial. 3. Let C"[a, fe], where « > 0 is an integer, denote the set of all real-valued functions u(x) defined on the interval a < x < b such that u is continuous, and the derivatives dku/dxk of order k < n exist and are continuous on |a, b\. It can be shown that C" [a, b] is a linear vector space with respect to the usual rules of vector addition and scalar multiplication. 4. The set So = \u: u(x) e C2L0, L], -^— fflW^J + c(x)u = 0, 0 < x < L\ is a vector space with respect to the usual addition and scalar multiplication. However, the set S=\u:u(x)eC2\Q,Ll-— (a(x)'-^J+c{x)u = f(x), 0<x<L\ is not a linear vector space (why?). 5. Consider the transverse motion of a cable of length L, fixed at its ends (sec Fig. 7.1). Let C[0, L] denote the set of all real-valued, continuous functions a(x, t) defined on the closed interval 0 < x < L for any time t. The transverse deflection h(-, 0 (i.e., configuration) of the cable at any time t can be viewed as an element of C[0, /,]. However, not every element of C[(), L] is a possible u(x,t^ Figure 7.1 Transverse motion of a cable fixed at both ends.
CONCEPTS FROM FUNCTIONAL ANALYSIS 209 configuration of the cable because all possible configurations must pass through the points x = 0 and x = L; i.e., the boundary conditions w(0, t) = 0 and w(L, r) = 0 must be satisfied. Let S be the subset of C[0, L] made up of ail real-valued, continuous functions u{x, t) suchthat«(0, 0 = 0andw(L, /) = 0: S = (u: u(x) e C[0, LJ, «((), r) = 0, «(L, /) = 0}. Then S is a subspace of C[0, L], and all possible configurations (i.e., deflections) are contained in this space. Let U and V be each a linear vector space. An ordered pair is a pair of elements u € U and v G V where one of the elements is designated as the first member of the pair and the other is designated as the second. We denote ordered pairs by (w, v) with the obvious order. Then 17 x V is called a product space W with elements w = («, v), it € U and v e V: W = {w: w = (w, v), wet/, lie V}, which is also a linear vector space with respect to the following definitions of vector addition and scalar multiplication of a vector in the product space V x V: (Mi, V\) + (W2, V2) = ("1 + «2, "I + V2), (7.5a) a(w, u) = (aw, av), a e d\ (7.5b) for («i, ui), («2» ^2) 6 W = U x V with Hi, W2 € £/ and ui, t>2 € V7. Consider an open-bounded domain Q C St3. Note that Q is a set of points x = C*i, X2, xj). A real-valued function u(x) is said to be square-integrable in the domain Q if the integrals (in the Lebesgue sense) / w(x)dx, /* |«(x)|2dx (7.6) exist and are finite. The space of square-integrable functions u defined over a domain £2 is called the L2 space: L2(0)= |i#(x): /" |«(x)|2dxf. (7.7) There is a corresponding space 7,00 (£2), which consists of all real-valued functions w(x) defined in the domain Q such that there exists an N with the property that l«(x)| < N.
210 DIRECT VARIATIONAL METHODS Linear Independence Recall the concepts of coplanar and collinear vectors in Euclidean space from Chapter 2. These concepts can be generalized to function spaces. An expression of the form n a\it\ +<X2ii2 H = 2_2a'u' ^>%) i=\ for all functions ut{x) and scalars a; e SW (real number field) is called a linear combination of h/. The equation £? a/Wf = 0 is called a linear relation among the functions w/. A set of n functions, ii[, U2> •. •, un, is said to be linearly dependent if a set of n numbers, ot[, o^, ...,«„, not all of which are zero, can be found such that the following linear relation holds: n £(*,«,=<>. (7.9) If there does not exist at least one nonzero number among at such that the above relation is satisfied, the vectors are said to be linearly independent Example 7.2 1. Consider the following set of polynomials, {/?,•}, with /?](*) = 1 +*, P2(x) = 1+A'2, p3(x) = I +X+X3, Consider the linear relation Ctipi +OL2P2 +<*3/>3 =0 for a/ € tft. Since the above relation must hold for all x, it follows that the coefficients of powers of x must be zero separately. Collecting the coefficients of various powers of jc and setting them to zero, we obtain «i+c*2-h<*3 = 0, ai +«3=0, «2 = 0, a-} = 0. The solution to these equations is trivial (i.e., all a,- = 0); hence, the set (Pi. £>2, #3} is linearly independent. 2. If p^ is replaced by 774 = 2 -f x + x2, we see that the linear relation <x\Pi + «2i?2 + &4P4 = 0 requires that #1 + #2 + 2<*4 = 0, 0:1+0:4=0, a2 + Qf4=0. An infinite number of solutions to the above set of equations exists. For example, CK4 =1, <X\ =&2 = —1
CONCEPTS FROM FUNCTIONAL ANALYSIS 211 is a solution. Hence, the set {/?[, p2> pa] is linearly dependent. Indeed, p4 can be expressed as a linear combination of p\ and pi\ P4 =aipi +a2/?2, ofi =a2 = 1. 7.2.3 Normed and Inner Product Spaces Norm The concepts of distance between two points and length of a physical vector can be generalized to abstract vectors, i.e., vectors that are functions. Let V be a linear vector space over the real number field SW. We shall use the notation || • || to denote the norm of real-valued functions w(x), xefic SR3. Then, associated with every vector u € V\ there exists a real number || ■ || e dt, called the norm, that satisfies certain rules, as discussed below. Thus the norm is the operation, || ■ || : V -* 5ft. (1) Nonnegative: (a) ||«|| >0 for alia. (b) \\u\\ = 0onlyifw =0. (7JO) (2) Homogeneous:||aw|| = \ot\ \\u\\. (3) Triangle inequality:||« + v\\ < \\u\\ + ||i>||. If ||a|| satisfies (ia), (2), and (3), it is called a seminorm, and is denoted by \u\. A linear vector space endowed with a norm is called a normed vector space. A linear subspace S of a normed vector space V is a linear subspace equipped with the norm of V. A norm || • || can be used to define a notion of distance between vectors, called natural metric: d(u, v) = ||u - v\\ for u, v € V. (7.11) Examples of norms will be given shortly. For I < p < oo, we define the Lebesgue spaces [see Eq. (7.7)]: L/,(Q) = {ii:||ii||/i<oo}, (7.12) where \\u\\Lpm = \\u\\P = \f |u(x)|' dx\ ' < oo. (7.13) For p — oo we set llwllz^Q) s Hulloo = sup{|M(x)|: x e Q]. (7.14) This is called the "sup-norm."
212 DIRECT VARIATIONAL METHODS Two norms || • || i and || ■ H2 on a normed vector space V are said to be equivalent if there exist positive numbers c\ and C2, independent of u e V, such that the following double inequality holds: Cilluili < \\u\\2 <C2\\uh. (7.15) A normed space V is called complete if every Cauchy sequence {uj} of elements of V has a limit u € V. For a normed vector space, a Cauchy sequence is one such that j It -* 0 as 7, /c 00, || Uj - Uk and completeness means that ||u — Uj\\ -> 0 as j -» 00. A normed vector space which is complete in its natural metric is called a Banach space. A linear subspace of a Banach space is itself a Banach space if and only if the subspace is complete. Example 7.3 1. The n-dimensional Euclidean space St'1 is a Banach space with respect to the Euclidean norm: X ss N X>.2 (7.16) 2. The space C[0, 1] of real-valued continuous functions f(x) defined on the closed interval [0, 1] with the sup-norm (7.14) is a Banach space. It is a linear vector space with respect to the vector addition and scalar multiplication defined as (/ + *)(*) = fix) + *(*), («/)(*) = «/(*), a s 91. Further, it is complete with respect to the sup-norm in (7.14): Il/Hoos max|/(x)|. 3. Sobolev space, W"up(Q). Let Cm (Q) denote the set of all real-valued functions with m continuous derivatives defined in Q G ffi3, and let C°°(Q) denote the set of infinitely differcntiable continuous functions. We define on Cm(Q) the norm, called the Sobolev norm, Wuup [ £ \Dau(x -iVp )\"dx (7.17) ' \ct\<m
CONCEPTS FROM FUNCTIONAL ANALYSIS 213 for 1 < p < oo and for all u e Cm(&). In Eq. (7.17), a denotes an rc-tuple of integers: n (7.18) ai«i For m = 1, n = 2, and 1 < p < oo, we have fa = (ai, 0C2), a\, <*2 = 0, 1], and ll"lli,P = {( [l« Ugcsh2 L lp + Sx p + Aw dxdy i/p (7.19) The space Cm (Q) is not complete with respect to the Sobolev norm || « ||,„iP. The completion of Cm (Q) with respect to the norm |[ * \\mtp is called the Sobolev space of order (m, /?), denoted by Ww-p(£l). The completion of C(£2) is the L2(Q) space. Hence the Sobolev space is a Banach space. Of course, the Lebesque space LP(Q) is a special case of the Sobolev space Wm,p form = 0, and L2(£2) is a special case of LP{Q) for p = 2, with the norms defined in (7.13). Tf U and V are each normed vector spaces, we can define a norm on the product space U x V in one of the following ways: (1) ||(»,i;)|| = ||ii||y + ||i;||v. (2) ||(tt, v)\\ = (Nlf; + llO^, p > I. (7.20) (3) ||(ji.v)||= max(||ii||c/f||i;||v). Then U x V is a normed vector space with respect to any one of the above norms. Inner Product Analogous to the scalar product of physical vectors, the inner product of a pair of vectors it and v from an abstract vector space V is defined to be a real number, denoted (wf v)y (i.e., (♦, -)v : V x V -+ W], which satisfies the following rules for every u\, 112, u,v e V and a e ?i\: (1) Symmetry: (w, v)y = (i>, w)y. (2a) Homogeneous: (ecu, v)y =a(uy v)y. (7.21) (2b) Additive: (u\ + «2» u)y = («i > i>)k + ("2» u)y. (3) Positive-definite: («, u)y > 0 for all w -fi 0. One can define a number of inner products and associated natural norms for pairs of functions that, along with their derivatives, are square-integrable. In particular,
214 DIRECT VARIATIONAL METHODS the Sobolev space Wm>2(Q) = Hm (£2), which is also known as the Hilbert space of order m, is endowed with the inner product (u,v)m= [ ][] D*u(x)D*i;(x)dx, (7.22) ° |a|<w for all w, v € Jfw(Q). Note that for m = 0, we have //0(J2) = L2(ti). Some special cases of Eq. (7.22) are given by (w, v)o = I uvdxdy, ||w||o = >/(u> M)o. (7.23a) • x f / 9"9u 3«3iA 7 , „ n n r~ „™^ (m,w)i=/ wu+—^^^^H*"-^ Mil =v("iw)i» (7.23b) Jq V 9x dx dy dy) f ( du dv du dv d2u d2v Jn \ dx dx dy dy dxdy ax ay d2ud2v 32ud2v\ y , „ „ n — ,„„„ . + ^T^ + l^TT Ja'^', ll«lb = V(«7")2. (7.23c) 3xz dx1 dy1 dy1) A linear vector space on which an inner product can be defined is called an inner product space. A linear subspace S of an inner product space V is a subspace with the inner product of V. Note that the square root of the inner product of a vector with itself satisfies the axioms of a norm. Consequently, one can associate with every inner product in vector space V a norm (7.24) The norm thus obtained is called the norm induced by the inner product. Since we can associate with each inner product a norm, every inner product space is also a normed vector space. It should be obvious to the reader that the converse does not hold in general. Orthogonality Two vectors w, v e V are said to be orthogonal if («,u)k=0, (7.25) where (•, -)v denotes an inner product in V. Note that the concept of orthogonality is a generalization of the familiar notion of perpendicularity of one vector to another in Euclidean space. A set of mutually orthogonal vectors is called an orthogonal set. A sequence of functions {<pj} in L2(^) is called orthonormal \i 'l>iU = J> ' (7.26) 0, ifi £j. (<Pii<Pj)v = Sij = Here <$/; denotes the Kronecker delta. It can be shown that every orthonormal system is linearly independent.
CONCEPTS FROM FUNCTIONAL ANALYSIS 215 If two vectors u and v of an inner product space V are orthogonal, then the Pythagorean theorem holds even in function spaces: II" + wily = (« + Vy u + v)v = (m, u)y + 2(«, w)y + (v, v)y = ||«||y + ||v||^. A set of functions {fy} is said to be complete in L2 (£2) if every piecewise continuous function / can be approximated in ft by the sum X^=i cj4>j m sucn a wav that can be made as small as we wish (by increasing n). This property of the coordinate functions is the key to the proof of the convergence of the Ritz and Galerkin approximations. A complete (in its natural metric) inner product space is called a Hilbert space. We mention without proof the fact that every inner product space (hence a nornied space) has a completion. The following lemma, referred to as the fundamental lemma of variational calculus, plays an important role in variational theory. Lemma 7.1 Let V be an inner product space. If (w, v)y=0 for all f e V, then u = 0. Proof: Since (a, v)v = 0 for all v, it must also hold for v = w. Then («, u)y = 0 implies that u = 0. 7.2.4 Transformations, and Linear and Bilinear Forms A transformation T from a linear vector space U into another linear vector space V (both vector spaces are defined on the same field of scalars) is a correspondence that assigns to each element u e U a unique clement v = Tit e V. We use the terms transformation, mapping, and operator interchangeably, and the transformation is expressed as T : U —> V. A transformation T : U -+ V, where U and V vector spaces that have the same scalar field, is said to be linear if 1, T(au) = aT(u)t for all it e U, a e 9? (homogeneous); 2. T(u\ + u2) = T(u\) + T{u2)y for all «,, «2 € £/, (additive). (7,28) Otherwise it is said to be a nonlinear transformation. Transformations that map vectors (functions) into real numbers are of special interest in the present study. Such transformations arc called functionals. A linear transformation I : V -* ffl that maps a linear vector space V into the real number field d\ is called a linear functional.
216 DIRECT VARIATIONAL METHODS Similarly, a linear transformation that maps pairs of vectors («, v) e V x V into real number field flt, or B(>, •) : V x V -► £fl, is called a bilinear form. Examples of linear and bilinear forms are provided by fb fh (du dv \ Ku) = / fudx, B{u, v) = I —— 4- uv I dx. Ja Ja \dxdx J A bilinear foim is said to be symmetric if it is symmetric in its arguments: B(u,v) = B(vyu). (7.29) 7.2.5 Minimum of a Quadratic Functional Consider an operator equation of the form Au = /, (7.30) where A is a certain operator (often a differential operator), A : VA -> H, and / e H is a given function. Here T>,\ denotes a set of elements from a Hilbert space H. The denseness of Va in H is often assumed, but we will not discuss this topic in the present study. The differential equations ~^(EA^)=:J{X)' EA>0' 0<x<L, (7.31) d2 / d2u\ wAEl**rm' EI>0' 0<X<L' (7.32) are special cases of the operator equation (7.30) with A~ dx\tAdx)' A dx2\ "'dx2)' respectively. In these cases, H — £2(0, L). The set Va for Eq. (7.31) consists of functions from C2(0, L) and for (7.32) functions from C4(0, L). An operator A : Va -> H is called symmetric (or self-adjoint) if (Am, v)h = ("> Av)h (7.33) holds for all uyv e Va, where (•, -)/f is the inner product in H. An operator A is called strictly positive in D4 if it is symmetric in Va and if (Aw, 1/)// > 0 holds for all u e £>/t and « ^ 0, (7.34a) (Au, «)// = 0 if and only if u e VA and « = 0. (7.34b)
CONCEPTS FROM FUNCTIONAL ANALYSIS 217 A quadratic functional Q'.H —^ 9} is one that is quadratic in its arguments, Q(au) = a2Q(u) for a e 9L Every bilinear form £(•, ■) can be used to generate a quadratic form Q by setting Q(tt) = B{u, u), u e H. (7.35) The following results are of fundamental importance for the present study. Theorem 7.1 If A is a strictly positive operator in T>a, then Au — f in H has at most one solution u e T>a in H. Proof: Suppose that there exist two solutions U{, u-i € T>a- Then Au\ = f and A112 = / -> A(wi — 1*2) = 0 m & and (A(wi — H2), "1 — «2)tf =0 -> M] — W2 = 0 or M| = «2 € Z^, which was to be proved. Theorem 7.2 Let A be a positive operator in T>a, f e #• Eet Eq. (7.30) have a solution «o e X>/i. Then the quadratic functional I(u)^{(Au,u)H -(/,«)// (136) assumes its minimal value in T>a for the clement «o> i«e., I (it) > /(mo), and /(«) = I (no) only for w = wq. Conversely, if /"(«) assumes its minimal value, among all u e T>a, for the element «0, then uo is the solution of Eq. (7.30) (i.e., Auq = f). Proof: First note that I(u) is defined for all it e Va- Let uq be the solution of Eq. (7.30). Then / = Auq. Substituting for / into Eq. (7.36), we obtain for u e X^*: I(u) = j(Am, u)h - (Awo, u)h = £ [(Aw, w)ff - (Aw0,11)// - («, Auo)h] = j [(Ait, u)H - (Auo, u)n - (Aw, uo)h) = \ [(An, u)H - (Auo, 11)11 - (An, uo)h + (Auo, uo)h - (Awn. «o)#] = \ [(A(u - w0), it - wo)// - (4«o. «o)f/J 1 (7.37)
218 DIRECT VARIATIONAL METHODS where the linearity and symmetry of A, as well as the symmetry of the bilinear form, are used in arriving at the last step. From Eq. (7.37) it follows that /(«o) = -%(Auo> «o)/y- (7.38) Next, we use the strictly positive property of A to conclude that I{u)>I(uo) for ue Va, and I(u) = I(uq) if and only if w = wo in V&. Consequently, if the equation A wo — f is satisfied, then the functional /(«) assumes its minimal value in V& precisely for the element u = mo- Now suppose that / (w) assumes its minimal value in Va for the element wo. This implies that / (wo + av)>I (wo) for ae%veT>A. (7.39) Using again the symmetry of A and the symmetry of the inner product, one obtains /(uo+au) = \{A(un + av)9uo + av)n - (/, wo + cnOw = \ [(Ait0 + olAv, wo + av)n - 2(/, w0)// - 2a(/, v)H~\ = j[(Alio, «o)ff + <*(A v, «o)ff + «(Aw0, u)// + a2(Av, v)H - 2(/, wo)// - 2a(/, u)ff] = £[(A«0, «o)ff + 2a(A«0, v)ff + oc2(Av, v)H -2(/,«o)ff-2a(/,v)//]. (7.40) Since wo € Z>a and f e H are fixed elements, it is obvious that for arbitrarily fixed v e Va the function /(wo + <xv) is a quadratic function in the variable a. From Eq. (7.39) it follows that the function has a local minimum at a = 0, which implies that its first derivative is equal to zero at a = 0 [or, equivalently, the first variation of / is zero; see Eq. (4.89)]: — /(wo + au) =0, Id* Ja=o or by Eq. (7.40) that (Aw0, v)H ~ (/, v)H = 0 or (Aw0 - /, v)H = 0. Since v e H is arbitrary, by Lemma 7.1 it follows that A wo — / = K) in H. Example 7.4 Consider the differential equation ~Tx (a{x)^) = /(J°* a(x) > a ° < x < L' (7,41)
CONCEPTS FROM FUNCTIONAL ANALYSIS 219 subjected to the boundary conditions «(0) = 0, u{L) = 0, (7.42) which arise in connection with the transverse deflection of cables. Here u(x) denotes the deflection of a cable of original length L, tension a = a(x), and subjected to distributed transverse load f{x) (see Fig. 7.2). The boundary conditions in Eq. (7.42) indicate that the cable is fixed at x — 0 and x = L. Let us choose H = 1/2(0, L), and define Va as the linear set of functions that are continuous with their derivatives up to the second order inclusive in the interval [0, L] and satisfy the end conditions in (7.42). Define the operator A on Va by i (/ ( \u = I a dx \ dx J (7.43) We now set out to prove that A is strictly positive on T>a> First, we note that A is symmetric in Va'- For every u e T>a and v e Va, we have (^w)H=r[^(flS)],,rfx L dx J0 Jo dx \ dx) fL < Jvdu / = / a{x)-r--^dx Jo dx dx (7.44) L dx J0 Jq I dx \ dx) (7.45) where we have used the fact that w(0) = «(L) = v(0) = i>(L) = 0. Thus, A is symmetric on Va. From Eq, (7.44), it follows that fL (du\2 {Au, u)h = I a(x) I — ) dx for all it e Va- Jo \dx) (7.46) Figure 7.2 Transverse deflection of a cable fixed at its ends.
220 DIRECT VARIATIONAL METHODS Due to the fact that ^(jc) > 0 in [0, L], it follows from (7.46) that (An, u)h > 0 for every u e VA\. Moreover, if (An, u)H — 0, it follows that du — = 0 in [0, L]. dx This in turn implies that u(x) = c, constant in [0? L], Since «(0) = 0, it follows that c = 0 or u(x) = 0 in fO, L]. This proves that ,4 is strictly positive in P^. The quadratic functional associated with Eqs. (7.41) and (7.42) is given by 1 1 fL (du\2 fL n(u) = -(Au,u)H~(f\u)H = -j ci(x)l—\ dx-J fudx, (7.47) which represents the total potential energy of the cable. The first term is the elastic strain energy and the second term is the potential energy of the external load /(*). Let uq(x) be the solution of Eqs. (7.41) and (7.42). Then / = Auq and we have n(tt) = -(Auy u)h - (Alio, u)h =\ ta{x) (s)dx - r \~i {a{x)C^)]"dx = o / "W T~ dx- I a(x)-—— dx 2 J0 \dxj Jo dx dx (7.48) Since a(x)(uf - u'0)2 > 0, it is clear from Eq. (7.48) that U(u) is minimal in T>a if and only if uf = HflinX^.Thus, if wo e T>a is the solution of Eqs. (7.41) and (7.42), then the functional n (it) assumes its minimum for uo G Va< Conversely, let uq e T>a be the element minimizing the functional Ti(u) in (7.47). Let v e T>a be an arbitrary element from T>a and let a be an arbitrary real number. Then the minimum of Tl(u) implies that (u = mq + <xv) 0= — n0in + au) da t*=0 d_ da \[a{x)^Tx) dx-hfudx\a=a
THE RITZ METHOD 221 fL duodv fL = / a(x)—-—dx- / fvdx Jo dx dx J0 Since this result must hold for every v, it follows that -£(^)-/=o taff=^«- The above arguments are equivalent to the principle of minimum total potential energy discussed in Chapter 5. Recall that the operator A : T>a -> H is symmetric: (Au, v)h = («, At;)// for «, v sVa and positive definite on T*a, i.e., there exists a constant C > 0 so that (Aw, «)/? > C||«]|2 holds for every u e VA. (7.49) Hence, wc can define a new inner product (u, v)a on T>a as follows: (m, u)a = (Aw, v)H for all «, u e V\. (7.50) It can be easily verified that (w, i>)a satisfies the axioms (l)-(3) in Eq. (7.21) of an inner product. The linear set VA with the inner product (7.50) constitutes a linear vector space, called the energy space, and denoted by Ha • The norm and natural metric follow from the definition in (7.50): llwIIa = ("> ")*. d(u, v) = \\u - v\\A. (7.51) The energy space HA can be shown to be complete with respect to the metric defined in (7.51), and hence it is a HHbert space. Moreover, it can be shown that the functional n (u) can be extended to all elements of Ha, that the functional assumes its minimum at «o s Ha, and that the element uq is uniquely determined by the element / E H. These proofs are beyond the scope of the present study, and interested readers may consult Refs. [4], and [12-15]. 7.3 THE RITZ METHOD 7.3.1 Introduction As discussed in Chapters 5 and 6, the principles of virtual displacements and forces as applied to continuous systems can be used to determine the governing equations and natural boundary conditions of the problem. The energy methods (e.g., unit-dummy- force and unit-dummy-displacement methods and Castigliano's Theorems T and II)
222 DIRECT VARIATIONAL METHODS derived from these principles were used to determine deflections and forces at selected points. Here we consider a powerful method of determining approximate solutions to the governing equations of a problem by directly using the variational statements (i.e., virtual work principles, the principle of total potential energy, or the principle of complementary energy). The method bypasses the derivation of the Euler equations and goes directly from a variational statement of the problem to the solution of the Euler equations. One such direct method was proposed by German engineer W, Ritz (1878-1909). 7.3.2 Description of the Method Consider the linear operator equation Au = f in £2, (7.52) where A is a strictly positive operator from Ha into H, and f € H. The solution wo of Eq. (7.52) is the element wo e Ha that minimizes the quadratic functional /(«) = Uu,u)A -*(«), (7.53) where /(•) denotes a linear functional. In structural mechanics problems, the functional I (it) represents the total potential energy and S J (it) = 0 yields Eq. (7.52) as the Euler equation. We seek an approximation Un(x) of wo00, f°r a fixec* and preselected N, in the form N U0(X) R* UN(X) = Yl C'0/(*) + *>(*)> <7-54) where (j>i(x) are the elements of a base in Ha, and c,- are as yet unknown real constants. These constants are determined by the condition that /(t/jv) is the minimum. Since {</>/} is a base in Ha, the solution wo € Ha can be approximated to arbitrary accuracy by a suitable linear combination of its elements. Therefore, it can be expected that the approximate solution Un , with constants determined by minimizing the functional /(E///), will differ sufficiently slightly in Ha from the actual solution wo if N is selected sufficiently large. This process of determining Un is known as the Ritz method. To fully illustrate the basic idea of the Ritz method described above, we consider the axial deformation of a nonuniform bar with an end spring. The governing equation is -4~ (eA(x)^-) = /(*), 0 < x < L, (7.55a) ax \ ax J ' subjected to the boundary conditions = P, (7.55b) x=L w(0)=0, f du\ EA(x)— + ku(x) , ax J
THE RITZ METHOD 223 where E denotes Young's modulus, A = A(x) the area of cross section, L the length, k the spring constant, f(x) the distributed axial load, and P the axial load at x = L (see Fig. 7.3). The space Ha in this case is the completion of the set T>a of functions that are continuous with their derivatives up to the second order in [0, L]. As discussed earlier, the problem is equivalent to minimizing the total potential energy functional II: „, x rL\EA (du\2 ^ dx + - lu(L)]2 - Pu(L). (7.56) The necessary condition for the minimum of II is 0 = STl = B(Su, it) - l(Su) or B(Su, u) = I(8u), (7.57a) where £(•, •) is the bilinear form and /(♦) is the linear functional. Equation (7.57a) is known as the variational problem associated with Eq. (7.52). The inner product in Ha is defined by (u,v)A = B(u,v). (7.57b) For the specific case of Yl(u) in Eq. (7.56), the bilinear and linear forms are fL du dv fL B(u,v)= I EA-—-dx+ku(L)v(L), /(«)=/ fudx + Pu(L). (7.58) Jo dx dx J0 The Euler equations andnatural boundary conditions associated with the minimization of U(u) in Eq. (7.56) are d ( du\ —T- EA~T I - / = 0 in 0 < x < L, (7.59a) dx \ dx J EA~ + ku-P=0 at x = L. (7.59b) dx The essential boundary condition of the problem is provided by the geometric constraint «(0) = 0. (7.59c) The exact solution u to the problem is one that satisfies Eq. (7.59a) at every x e (0, L) and the boundary conditions in Eqs. (7.59b,c). Thus, the solution of Eqs. (7.59a-c) E,A f{x) [EEHwwh ^x Figure 7.3 Axial deformation of a nonuniform bar with an end spring.
224 DIRECT VARIATIONAL METHODS is equivalent to minimizing U(u) over a set of functions that satisfy the condition in (7.59c). In the Ritz method, we seek an approximate solution Un (which may be exact if we choose the right kind of approximate solution) to the problem as a finite linear combination of the form (7.54). The reason for selecting the particular form of the approximate solution will be apparent in the sequel. If we select (po and fa such that Un satisfies the specified essential boundary condition, «(0) = 0, and substitute Un into the total potential energy functional n in Eq. (7.56), we obtain n as a function of the parameters c\, C2,..., c^ (after carrying out the indicated integration with respect to x)\ TT = n(ci,c2,...,Civ). Then c,- are determined (or adjusted) such that <5FT = 0; in other words, we minimize n with respect to c,-9 i = 1, 2,..., N: 0 = «n = — Sci + —8c2 + • • • + —ScN = V -r-tcf. dc\ dc2 9cyv .t ocj Since the set {c{} is linearly independent, it follows that — =0 fori = 1,2, ...,W (7.60a) dej or [A]{c] = {b}. (7.60b) Equations (7.60a,b) represents a set of N linear equations among c\, d,..., c#, whose solution together with Eq. (7.54) yields the approximate solution Un(x). This completes the description of the Ritz method. Since the natural boundary conditions of the problem are included in the variational statement, we require the approximate solution Un to satisfy only the essential boundary conditions. In order for Un to satisfy the essential boundary conditions for any c;, it is convenient to choose the approximation in the form (7.54) and require (f>o(x) to satisfy the specified essential boundary conditions. For instance, if u(x) is specified to be u at x = 0, we require <J)q(x) be such that 0o(O) = w. Then N N UN(0) = £>0,(O) -I- 0o(O) = X>*/(0) + u. (7.61a) Since £//v(0) = u, it follows that N Y^CffriO) = 0 ->- 0/(0) = 0 for all i = 1, 2,..., N. (7.61b) i=\
THE RITZ METHOD 225 Thus, <t>i(x) must satisfy the homogeneous form of specified essential boundary conditions. Equation (7.54) can be viewed as a representation of u in a component form; the parameters c,- are the components (or coordinates) and {<£/} are the coordinate functions. Another interpretation of Eq. (7.54) is provided by the finite Fourier series, in which c/ are known as the Fourier coefficients. 7.3.3 Properties of Approximation Functions The approximation functions 0o and 4>i should be such that the substitution of Eq. (7.54) into SU or its equivalent results in N linearly independent equations for the parameters cj (j = 1, 2,..., Af) so that the system has a solution. To ensure that the algebraic equations resulting from the Ritz approximation have a solution, and the approximate solution U^{x) converges to the true solution u(x) of the problem as the value of N is increased, </>/(/= 1, 2,..., N) and <f>o must satisfy certain requirements. Before we list the requirements, it is informative to discuss the concepts of completeness of a set of functions and convergence of a sequence of approximations. To make the ideas presented simple to understand, mathematical rigor is sacrificed. Convergence A sequence [Un] of functions is said to converge to u if for each e > 0 there is a number M > 0, depending on €, such that || UN (x) - u (x) || < € for all N > M, where || * || denotes a norm of the enclosed quantity and u is called the limit of the sequence. In the above statement, U^ represents the approximate solution and it the true solution. The statement implies that the Af-parameter solution E/# can be made as close to u as we wish, say within e, by choosing N to be greater than M = JV/(e), provided that the approximate solution is convergent. While there is no formula to determine M, a series of trials will help determine the value of N for which the approximate solution Un is within the tolerance. Completeness The concept of convergence of a sequence involves a limit of the sequence. If the limit is not a part of the sequence {Vn}^=\ > then there is no hope of attaining convergence. For example, if the true solution to a certain problem is of the form u(x) = ax2 + bx3 + ex5, where a, b, and c arc constants, then the sequence of approximations £/l = ClA'3, U2 = CIA'3 + C2A-4, . . . , UN = QJC3 + C2X* + • • • + CNXN+2 will not converge to the true solution because the sequence does not contain the x2 term. The sequence is said to be incomplete. As a rule, in selecting an approximate solution one should include all terms up to the highest-order term. If a certain term is not a part of the true solution, like the x4 term, its coefficient will turn out to be zero by the time all terms of the true solution are included in the approximation.
226 DIRECT VARIATIONAL METHODS We now list the requirements of a convergent Ritz approximation (7.54): 1. </)q must satisfy the specified essential boundary conditions. When the specified essential boundary conditions are homogeneous, then (po(x) = 0. 2. <pj e Ha (i = 1, 2,..., N) must satisfy the following three conditions: (a) be continuous, as required by the variational statement being used; (b) satisfy the homogeneous form of the specified essential boundary conditions; and (c) the set {<£,•} must be linearly independent and complete. (7.62) 7.3.4 Rifz Equations for the Parameters Returning to the problem of determining the approximate solution U^(x) of the bar problem described by Eqs. (7,59a,b), we substitute Eq. (7.54) into the total potential energy functional II: x fL\ EA (4^ d<f> n{cl,C2.....CN) = ]Q l—^cj- £♦£-/!>*+*> U=1 dx + o \J2ci<j>j(L)+<j>o(L) KJ=i I -plf^Cj4>j(L)+4Ki(L)y Now differentiating n with respect to c\ (i = 1,2,..., N), we obtain N linearly independent equations for the unknowns, ci, C2,..., c^. We ^ave dci Jo cl(j>i d<f>j d<po '*% 2"^+ 7 = 1 dx f<t>\ dx + /c0,(L) ^Cj<j>i(L) +*o(L) - P0i(L), (Eq. I) KJ=* 9c2 Jo N dx I f^ dx dx ' N dx + *^2(L) [ ^2cj<f>j(L) +<Pa(L) ] - P<h(L), (Eq. 2)
THE RITZ METHOD 227 3c,- Jo d(f>i U=i d<t>j d<f)Q dx dx dx + k<t>i(L) ( £c,tf>,(L) +MD I - P<t>i(L), (Eq. 0 o-r-f "%[£'>%+%)-'** N dx + k<j>N{L) ( J2cJ<l>j(L) + ML) ] - P4>n{L). (Eq. tf) Note that Eq. i is the *"th equation of the set of N equations. The ith equation can be written in the short form N 0 = Y2"ijCj ~bi> where the coefficients a\j and b\ are defined by (7.63a) Jo dx dx bi=- f EA^^dx-kMLWiiL) Jo dx dx + f f4>n Jo ;dx + P<t>i(L). The N equations can be written in matrix form as [A]M = {b}. (7.63b) (7.64) Equations (7.64) are called the equations for the Ritz parameters C{. Once c\ (i — 1, 2,..., N) are determined from Eq. (7.64), the approximate solution of the problem is given by Eq. (7.54). This displacement can be used to evaluate strains and stresses: du ^ y^\ d<j>i d<f>o dx ~ ' dx dx
228 DIRECT VARIATIONAL METHODS Equations (7.63a,b) can also be arrived at by substituting the Ritz approximation (7.54) and its variation iV i=l in the variational statement 8YI = 0, instead of substituting (7.54) in Ft and then taking the variation with respect to c,-. This results in N / = ! +/<0,(L) \^2cj4>j(L)+4>o(L) J - P0/(L) I . Since 8c; arc arbitrary, we obtain the result in Eq. (7.63a). Next we discuss the task of selecting the approximation functions 0o and 0/. The properties listed in Eq. (7.62) provide guidelines for selecting the coordinate functions 0o(a*) and 0/(x); they do not, however, give any formulae for generating the functions. Thus, apart from the guidelines, the selection of the coordinate functions is largely arbitrary, As a general rule, the coordinate functions 0,- should be selected from an admissible set [i.e., those meeting the conditions in Eq. (7.62)], from the lowest order to a desirable order, without missing any intermediate terms (i.e., the completeness property). Also, 0o should be any lowest order (including zero) that satisfied the specified essential boundary conditions of the problem; it has no continuity (differentiability) requirement. For the problem at hand, 0o = 0 since the specified essential boundary condition is homogeneous. Next, we find 0i (a) such that <p\ (0) = 0 and differentiate at least once with respect to x because n involves the first derivatives of it ^ U^. If an algebraic polynomial is to be selected, the lowest-order polynomial that has a nonzero first derivative is 01 (x) — a + bx, where a and b are constants to be determined. The condition 0i (0) = 0 gives a = 0. Since b is arbitrary, we take it to be unity (any nonzero constant will be absorbed into C[). When N > 1, property 2(c) in Eq. (7.62) requires that 0/, i > 1, should be selected such that the set {0/ }£L, is linearly independent and makes the set complete. In the present case, this is done by choosing 02 to be x2. Clearly, <jfi{x) = x2 meets the conditions 02(0) = 0, linearly independent of 0i (x) = x (i.e., 02 is not a constant multiple of 0i), and the set [x, x2} is complete (i.e., no other admissible term up to quadratic is omitted). In other words, in selecting coordinate functions of a given degree, one should not omit any lower-order terms that arc admissible. Otherwise the approximate solution will never converge to the exact solution, no matter how many f Jo EA dx d</>j d(/>o dx dx ■f<tn dx
THE R!TZ METHOD 229 terms are used in the Ritz approximation, as the exact solution may have those lower- order terms that were omitted in the approximate solution. Note that faW = x -f x2 is also an admissible function that meets all requirements. Then Ul(x) = C\<t>\ + C2$2 = C\X + C2X2, With C[ = C\ + C2, C'2 = C2> which is equivalent to U%(X) =C\(t>\ +C2<j)2. Thus, one may select (pi = xl,i — 1, 2,..., N. If trigonometric functions are to be selected, one may be tempted to select (p\ = sm(nx/L), which satisfies the condition <p\ (0) = 0. However, this choice also gives U\(L) = 0 since 01 (L) = 0. A better choice would be to select (p\(x) = sin(jtx/2L)y or for N > 1, select 0/ = sin[(2i - 1 )nx/2L\\ this choice will yield a good solution. For the choice of algebraic polynomials, the JV-parameter Ritz approximation for the bar problem is N UN (x) = £ a^ (*), 0, (x) = x!, (7.66) and the coefficients a// and b\ for EA = a0 (2 - j-} , / = /0 (a constant), P = P0. (7.67) are given by = CO'] j (l-~) x'+J~2 dx + Jfc(L)'+> = flvf +nVl ^L?+J~l + ^^> <7'68a) (1 + J - l)(i + j) Jo 1 + 1 bi = / f<Pi dx + P<Pi(L) = ^-(D'H"1 + pQ(Q(. (7.68b) For one-term approximation (N = 1 and k — 0), we have 3 1 9 Cl — n—T I 2^L + pL\ = — a\\ 9ciqL \6 J 3a0
230 DIRECT VARIATIONAL METHODS and foL + 2Po U\(x) = r X. 3ao For N — 2 and k = 0, we have 3 4 9 5 3 ail = 2a°L* an = a21 = ^a°L~> a22 ~ oaoZ/'' ft, = ~/0L2 + P0^ *2 = 3/oi3 + PqL2. The Ritz equations can be written in matrix form as 6 whose solution by Cramer's rule is Hence the two-parameter Ritz solution is U rx) = W1±^Pox + 3(/>o-/oL)T2 13^o 13ao^ The exact solution of Eqs. (7.55a,b) with w(0) = 0, *: - 0, J?A = ao[2 - (*/£)], and / = fo is u(x) = x + log(l - —) (7.69a) Table 7.1 contains a comparison of the Ritz coefficients c,- for Af = 1, 2,..., 8 with the exact coefficients in Eq. (7.69b) for L = 10 ft, a0 = J 80 x 106 lb, f0 = 0, and Po = 10 kip. Clearly the Ritz coefficients c,- converge to the exact ones as AT goes from 1 to 8. 7.3.5 General Features of the Method Some general features of the Ritz method are listed below: 1. Ifthe approximate functions </>/(*) are selected to satisfy Eq. (7.62), theassumed approximation Un(x) normally converges to the actual solution u(x) with an increase in the number of parameters (i.e., AT -> oo). A mathematical proof of such an assertion is not given here, but interested readers may consult the references at the end of the chapter.
THE RITZ METHOD 231 Table 7.1 The Ritz coefficients*1 for the axiai deformation of an isotropic elastic bar subjected to axial force n c\ c2 c3 cA c5 c6 c7 c8 I 2 3 4 5 6 7 8 Exact "*=* 37.037 25.641 28.219 27.691 27.794 27.775 27.778 27.778 27.778 f x 105+'\ 12.821 4.409 7.788 6.701 7.009 6.929 6.948 6.944 2. For increasing values of N, the previously computed coefficients of the algebraic equations (7.60b) remain unchanged (provided the previously selected coordinate functions are not changed), and one must add newly computed coefficients to the system of equations. 3. If the set of approximation functions {0;} chosen is an orthogonal set in the sense B((pi, <}>j) = «(//)<$/,• (no sum on i and y), then one need not invert the system of equations, and the solution is obtained as c,- = bi/an. 4. The Ritz method applies to all problems, linear or nonlinear, as long as the variational problem B(8u, u) = l(8u) (7,70) is equivalent to the governing equation and natural boundary conditions. In general, B(8u, u) is linear in 8u but may be nonlinear in it (and thus £(•,«) may not be symmetric), and l(8u) is a linear functional. 5. If the variational problem used in the Ritz approximation is such that its bilinear form is symmetric (in u and 8u)s the resulting algebraic equations are also symmetric and, therefore, only elements above or below the main diagonal of the coefficient matrix need to be computed. 6. If 811 (or B{8u,u)) is nonlinear in w., the resulting algebraic equations [A({c})]{c} = {b} will also be nonlinear in the parameters c/. To solve such nonlinear equations, a variety of numerical methods are available (e.g., the Newton-Raphson method). Generally, there is more than one solution to the set of nonlinear equations. 7. Since the strains are computed from the approximate displacements, the strains and stresses are generally less accurate than the displacements, 8. The governing equation and natural boundary conditions of the problem are satisfied only in the variational sense, and not in the differential equation sense. 4.879 0.000 3.389 1.904 2.453 2.272 2.315 3.029 -1.040 2.012 0.320 1.094 0.868 1.664 -1.142 1.447 -0.287 0.347 0.952 -0.980 1.136 0.145 0.560 -0.769 0.062 0.336 0.027
232 DIRECT VARIATIONAL METHODS Therefore, the displacements obtained from the Ritz approximation generally do not satisfy the equations of equilibrium pointwise. 9. Since a continuous system is approximated by a finite number of coordinates (or degrees of freedom), the approximate system is less flexible than the aclual system (recall that 11(h) < Tl(U) for any U that is not exact). Consequently, the displacements obtained by the Ritz method using the principle of minimum total potential energy converge to the exact displacement from below: U\ < U2 < • • • < Un <£/,„•••< w, for m > 77, (7.71) where U^ denotes the N-parameter Ritz approximation of u obtained from the principle of minimum total potential energy. The displacements obtained from the Ritz approximations based on the total complementary energy principle provide an upper bound for the exact solution. 10. Although the discussion of the Ritz method in this section thus far is confined to a linear solid mechanics problem, the method can be employed for any equation that admits a variational formulation (in the sense discussed in Comment 4), as will be illustrated through several examples shortly. However, the bounds mentioned above do not hold unless the variational problem is based on a minimum variational principle. 7.3.6 Examples Tn this section we illustrate the application of the Ritz method to a variety of problems. These include static, eigenvalue, and transient problems. As will be shown in the sequel, the Ritz method can also be applied to problems that either do not admit a quadratic functional or where one knows only the governing equations of the problem. In the latter case, a way to develop the so-called weak form is discussed in Section 7.4. Example 7.5 Consider a uniform cross-section bar of length L, with the left end fixed and the right end connected to a rigid support via a linear elastic spring (with spring constant /c), as shown in Fig. 7.4. We wish to determine the first two natural axial frequencies of the bar using the Ritz method. The kinetic energy K and the strain energy U associated with the axial motion of the member are given by fL pA /Su\2 , fL EA (du\2 k 9 K = L ~ {*) dx< U = L- (al) dx + 2[U(L'r)] • a72) E,A h :hww\tI »~x Figure 7.4 Natural vibrations of a bar with an end spring.
THE RITZ METHOD 233 Substituting for K and (J from Eq. (7.72), and V = 0 in Hamilton's principle, we obtain [8u(x, t\) = Su(x912) = 0 and 5w(0, t) = ()]: 0 = f2 8(K-U)dt = / / (-My7Sw~EA^—^J^-fciiCL,^^"^,^ U^ (7.74) We seek the periodic motion of the form i«(;c, 0 - boW^, / = v73!, (7.75) where o> is the frequency of natural vibration, and u$ (x) is the amplitude. Substituting Eq. (7.75) into Eq. (7.74), we obtain 0 = / ( pA(d2uq8uq - EAC-4^£~~P^ | dx - ku0(L)8u0(L), (7.76) Jo V «* «* / where (Jco)2 = ~a)2, and j^e2i0)tdl, being nonzero, is factored out. Wc use Eq. (7.76) to determine the values of co. Note that the Rayleigh quotient for the problem at hand is given by a? = f0 EA(diio/dx)(d8uo/dx)dx + kuo(L)8uo(L) j0 pAitQ8uQdx The Euler equation and natural boundary condition associated with Eq. (7.76) are -4-{£A^ J -pAcD2iiQ = Q, 0<x<L, (7.77a) dx \ dx J dim EA—^+ kitQ = 0 at x = L. (7.77b) dx The essential boundary condition is wo(0) = 0. Anondimcnsionalization of the variables is used for simplicity: x _ «o kh , co2pL2 ,„„™ x = t u = t- a = JX> x = Hh (7-78)
234 DIRECT VARIATIONAL METHODS Then Eq. (7.76) becomes 0=/ (ku8u-^^)dx-au(l)8u(l). (7.79) Jo \ dx dx ) The bar over the nondimensional variables will be omitted in the interest of brevity. Further, in the following discussions, we shall assume that a = 1, Substituting an /^-parameter Ritz approximation (obviously, we have <po = 0): N u(x) « Un(x) = ]Pc;0/(*) j=1 into Eq. (7,79), we obtain N °=E 1 = 1 N E '{^^(I^^^^^]1 Because of the independent nature of <5c/, we obtain \8cj. c,\ and in matrix form where ([A]-k[M])[c} = [0}9 (7.80a) (7,80b) aU =J ^^7dx+ c^(l)<^(l), my - J <^, dx. (7.80c) Equation (7.80b) represents a matrix eigenvalue problem, and we obtain N eigenvalues, A,,-, / = 1, 2,..., Af. An analytical method for finding eigenvalues and eigenvectors was discussed in Chapter 2 (see Section 2.3.4). For the problem at hand, the approximation functions can be taken as Substituting <pt — x* into Eq. (7.80c), we obtain (7.8 la) ma = / <pi<pjdx = ——r Jo i + J ^d^d^ J Jq dx dx + r d* + fc(l)0y(l) = U i+j-1 + 1. (7.81b)
THE RITZ METHOD 235 Since we wish to determine two eigenvalues, we take N = 2 and obtain 77211 = 5, m\2 = |, m22 = 3, an - 2, a\2 = 2, a22 = 5, and the matrix eigenvalue problem (7.80b) becomes (B IH Dlsl-I!)- For a nontrivial solution (i.e., c\ -fi 0, c2 / 0), the determinant of the coefficient matrix in Eq. (7.82) is set to zero: 9 A 9 X * 3 L 4 9 _ A 2 A ^435 = 0 or 15A2-640A-f 2400 = 0. The quadratic equation has two roots: aa*a* •> 00 ^. 2.038 /iF 6.206 [e AA =4.1545, k2 = 38.512 -► ^ = /—, ^2 = J —. L Y p L y /o (7.83) The eigenvectors are given by U^=c^x + c^x2t where cj° and cj* are calculated from Eq. (7.82) for A = A,,-, / = 1,2 (see Example 2.8). The exact values of X arc the roots of the transcendental equation A + tanA = 0, (7.84) whose first two roots are (co2 = X): 2.02875 [e 4.91318 [E °*l = -irfa (0(a = —L-p- (7-85) Note that the first approximate frequency is closer to the exact than the second. If one selects <j>o and 0/ to satisfy the natural boundary condition also, the degree of polynomials will inevitably go up. For example, the lowest-order function that satisfies the homogeneous form (we still have </>o = 0) of the natural boundary
236 DIRECT VARIATIONAL METHODS condition w'(l) + w(l) = 0 is 4>x = ?>x-2x2. (7.86) The one-parameter solution with the choice of (j)\ in Eq. (7.84) gives Xi = 50/12 = 4.1667, which is no better than the two-parameter solution computed using <j>{ = x and fa — A'2. Of course, solution ci$[ would yield a more accurate value for Ai than the solution c\4>\. Although c\4>\ and C[(p\ + C2<p2 are of the same degree (polynomials), the latter gives better accuracy for X\ because the number of parameters is greater, which provides greater freedom to adjust the parameters. Example 7.6 Consider a uniform cross-section bar of length L with the left end fixed and the right end connected to a rigid support via a linear elastic spring with spring constant k, Suppose that the bar is subjected to a body force /(*,/) (see Fig. 7.4). We wish to determine the transient response of the bar under the assumption that the motion starts from rest, i.e., the initial conditions of the problem are w(x,0) = 0, m(jc,0)=0. (7.87) The kinetic and strain energies associated with the axial motion of the bar are given in Eq. (7.70). The potential energy due to f(x> /) is V = - f J fudx. Jt] Jo Then Eq. (7.73) becomes 0 = f2\ f (-pA^-8u - EA^yt + /*") dx ~ k"(L> t)8u{L, t) \dt. (7.88) The Euler-Lagrangc equations associated with Eq. (7.88) are ±(EA*L\_±( Ate\ Qt 0<X<L. f>0t (789a) dx \ dx J dt \ dt / EA — + ku =0, at x = L\ t > 0. (7,89b) dx When one is interested in determining the time-dependent solution u(xy /) under applied load f(x, /), the Ritz solution is sought in the form u(x, t) « 5^cy(O0; W, <Pi - x\ (7.90) 7=1
THE RITZ METHOD 237 where Cj arc now time-dependent parameters to be determined for all times t > 0. Substituting Eq. (7.90) into (7.88), we obtain °=-£ 7 = 1 / 4>ifd> Jo c ~dt2 r ^-^ ^) ^+(r ^^^+^(,)^(i)) ^ = - E (m'J C-^T + a'JcJ ) + h>. ^7-91a) where xt u, /, and t are nondimensionalized as * = !• l7 = obzW /=i ' = wb' (7-9,b) /o being a constant, and m,7 = / 4>i<l>jdx> Jo Jo c/x dx bi = I <t>if(x,t)dx. Jo For N = 1 and / = /o (or / = 1), we have d2c\ , \d2c\ ^ 1 Wll¥+flllCl=t| °r 3^ + 2Cl==2' The solution to the second-order differential equation is ci(t) = A sin V6t + B cos V6* + \, Hence the one-parameter Ritz solution is given by U[ (a% /) = (A sin V6r + B cos \/6* + \)x, where A and B are constants to be determined using the initial conditions. For zero initial conditions «(jt, 0) = 0, w(x,0) = 0, (7.92) we can determine the constants as B = — 1 /4 and A = 0. The solution becomes f/l (jt, /) = |(l - COS y/6t)x.
238 DIRECT VARIATIONAL METHODS j^^^^^fr^fr^frj/ 1 lz,w0 Figure 7.5 A simply supported beam under uniform (oad. For N > 2, the resulting system of differential equations in time, Eq. (7.91a), may be solved for c{(t) using the Laplace transform method or a numerical method, the latter being more practical. In general, the initial conditions (7.92) cannot be satisfied exactly, requiring approximation. Further, the numerical solution may be obtained only for discrete values of time. For example, at time ts = sAt (i.e., the total time interval is divided into a finite number of time steps of size At), wc would have "n (x , is) = ^2 CJ &)0j (*)■ 7=1 For more details of the ideas discussed here, see Example 7.17. (7.93) Example 7 J Consider a simply supported beam of length L. We wish to find the transverse deflection of the beam under uniformly distributed transverse load qo (see Fig. 7.5) using the Euler-Bernoulli beam theory. The principle of virtual displacements for the problem becomes °=iL(EI^d-^-Sw^)dx- (7-94a) The essential boundary conditions are w0(0) = wq(L) = 0. (7.94b) We choose a two-parameter approximation of the form wo ~ W2C*) = cj0i + c2(j)2 + 00, 8W2 = Sc\4>\ + Sc2(f>2, (7.95a) where 4o = 0, 0i = x(L - jc), <fe = x2(L - *). (7.95b) Substituting Eq. (7.95a) into Eq. (7.94a), we obtain 0 = / [El(8ci4>f[ + 8c24%){ci<l>i + C20£) - (^101 +&C2<h)]qodx i/O
THE RITZ METHOD 239 where 2/2 = X) 12aUcj ~bi J5^ or 1=1 \j=\ fL d2<t>id2<j>j 0 = ^/aiJCj -bu 7=1 Jo qo&dx- For the particular choice of <j)\ in Eq. (7.95b), we have EIL 4 2L "I f c, 2L 4L2 •](: 40^ c2 J 12 I L and the solution of these equations yields the result qoL2 C[ = 24ET so that the two-parameter Ritz solution becomes qoL* C2 = 0, W2(x) = 24E/ (f"S) The exact solution of the problem is given by q0L4 (x ^ . . qoL4 (x A'3 x4\ The maximum deflections according to the exact and Ritz solutions are 37 q0L4 Wq (§> 5 ?0L4=0.0I302?0L4 384 EI EI ' W2 ©- (7.96) (7.97) 2352 £7 = 0.01042 EI Thus the two-parameter Ritz approximation is about 20% in error. The three-parameter Ritz approximation with fa = x3(L — x) yields EIL ' 4 21 2L2 "» 2L 4L2 4L3 2L2 4L3 4.8L4 qoL3 12 2 L 0.6L2 (7.98) and we obtain r t 2 ^L Cl=c2L = -c3L = _ The three-parameter Ritz solution coincides with the exact solution in Eq. (7.97).
240 DIRECT VARIATIONAL METHODS Remark 1 If the beam is subjected to a point load Fq at x = L/2, instead of a uniform load throughout the span of the beam (see Fig. 7.6), the exact solution will be in two parts: wq(x) = FpL3 48 £7 FQL3 4SEI H-4)- L 0<*<-, L - < x < h. 2 ~ ~ (7.99) Then it is clear that we must seek the Ritz solution also in two parts. Suppose that we use the virtual work statement (or <$n = 0) Jo L d23wod2w0 EI—— z-^-dx — Fqowq (§) dx2 dx2 with the three-parameter approximation Wj(x) = C]x(L - x) + t'2X2(L - x) + C3X3(L — jc), (7.100) (7.101) we obtain the same coefficient matrix as in Eq. (7.94), because the bilinear form did not change but the linear form changed, and the right-hand side is given by FoL2 L ~2 L 4 The solution for the Ritz coefficients gives F0L F0 c\ }2ET' Q = — 12EI C'3 = - 5*0 64E/L* The Ritz solution does not coincide with the exact solution. In particular, the maximum deflection predicted by the three-parameter Ritz approximation (7.101) is W3 (-) = - \2j 38 F0L3 FQL3 384 EI 54.S6E1' Fu Ll2 * L/2 >| z,w0 Figure 7.6 A simply supported beam under center point load.
THE RITZ METHOD 241 whereas the exact value from Eq. (7.99) is ©- MTl=i§7- (7-102) The reason for the Ritz solution based on the variational problem (7.100) not being exact even for a point load is that (7.100) does not account for the discontinuity in the load. Note that the exact shear force, Q = EI(d^\vo/dx3)y is discontinuous at jc = L/2, but the approximate one is continuous. Thus, we must modify the variational problem, as discussed in Example 7.8. Remark 2 One can also use trigonometric polynomials in place of algebraic polynomials for the approximation functions (ft. For instance, the deflection of a simply supported beam subjected to continuous distributed load q(x) can be represented by . nx . 3nx ^nx wo ^ C[ sin — + c2 sin — 1 h cN sm(2N - 1) —. (7.103) The functions 0/ = sin(2/ — 1)(ttx/L) are linearly independent, and are complete if all lower functions up to sin(2N - l)(nx/L) are included. When the load is sinusoidal, . x . mnx q (x) = c/0 sin (for fixed m), (7.104) we obtain the exact solution (c\ = ci = • • • = c,„_j = cm+\ = • • • = cyv = 0): , x . mnx c/qL4 w0(x) = cm sin ——, cm = 4 4. (7.105) This solution cannot be represented using a finite set of algebraic functions (<£,}, although the Ritz solution with a finite number of terms may be very close to the exact when evaluated at a point. On the other hand, if the load is representable by an algebraic polynomial (e.g., q (x) is a constant, linear, or higher-order function of*), then the Ritz solution (7.103) will not coincide with the exact solution (7.97) for any finite value of N, because the sine-series representation of such a load is an infinite series. For example, when q = c/o, a constant, then , . x-^ 16</n . inx q(x)=qo= V —~sin —-. (7.106) /=ft. m L However, the Ritz solution (7.103) converges rapidly, giving an accurate solution, especially away from the ends, for a finite value of N. Thus, in general, a judicious choice of approximation functions <pj based on the source term q(x) will not only make the computational effort minimal, but also gives an accurate solution.
242 DIRECT VARIATIONAL METHODS Example 7.8 Here we consider a beam with discontinuous loading. As an example, consider the beam shown in Fig. 7.6. We divide the beam into as many parts as there are regions with continuous loading. Consider the ith part, located between x — Xj and x = Xi+i. Within each part of the beam, the differential equation for wq(x) is d2 /-rd2w\ , , Xj < X < Xj 4, \ , (7.107) where q(x) is any distributed load in the part. We isolate the ith part and set up a free- body diagram depicting the internal forces {Q\ , Q^) and moments (Af, , M% ), as shown in Fig. 7.7. The total potential energy for the ah part is f*>« I EI fd2w0\2 1 . ..</) / dwo\ M, f(- (7.108) where the left end is labeled as 1 and the right end as 2. To use the Ritz method, we must select approximation functions based on specified boundary conditions. If we assume for the moment that all of the boundary conditions are of natural type, then the Ritz approximation over the zth part may be assumed in the form (see Example 5.6): Wyix) — C\ + C2X + C?>X2 + C4.X3 _,„('>, „<o «<'*> ,(0 (')/ ..<') .(')/ <p\"(x)w™ + wwey + ifl'Ww? + fl'MOZ 3<0 (7.109) M=0' Q '±± pM = 0 M\ If" V Qf2' Q?1 Qf+ef2,=F0 Figure 7.7 Boundary, continuity, and equilibrium conditions for the beam of Example 7.8.
THE RITZ METHOD 243 where (Jc = x — x; and L, = jc/+i — Xj) «.;>- (7.110) Substitution of Eq. (7.109) into 8Ylj the Ritz equations for the i th part: 0 and carrying out the integration yields 2E,I, where 6 -3Li -6 -3L," -3L, 2L? 3L, /.? -6 3/,/ 6 3L/ -3L/ 3L,< 2Lf X'^ < w(r ^r. ► r= ^ K}] «2W ^ Ufl ► + « [ fii'} 1 < Gf l"H , (7.111a) qf= I*"" q(i){x)<pf(x)dx. (7.111b) Here the superscript or subscript i on variables indicates that they belong to the ith part. Note that the modulus of elasticity and moment of inertia arc assumed to be constant within each part, while the load is arbitrary but continuous in each part. For uniformly distributed load q^(x) — c/q on part /, we have (7.111c) „<'>1 1\ tf iT „(') «4 J , _ tf*" 12 ! 6 -L 6 1 U Equations (7.111a) relate the four generalized displacements (w\ , 0j i(0 m(0 n(f) Oh o2 C/1 , Wn, , Or 2 > to the four generalized forces (Q\\ M\\ Q%\ M$}). Obviously, four of the eight variables should be known in order to solve the four equations. If there arc n parts, there will be a total of An equations in 8/1 variables. Hence the remaining An variables should be eliminated through known conditions (e.g., boundary conditions, continuity conditions, and equilibrium conditions). To illustrate the ideas, consider a simply supported beam with a point load Fq at the center. Obviously, the beam needs to be divided into two parts, 0 < x < L/2 and L/2 < x < L. The Ritz equations for the two parts are given below (E\ = Ej = E, 7X = I2 = /, L| = L2 = L/2, and q(x) = 0).
244 DIRECT VARIATIONAL METHODS Parti: )6EI Part 2: 16£7 6 -1.5L -6 -1.5L -1.5L 0.5Z,2 1.5L 0.25L -6 1.5L 6 1.5L -1.5L 0.25L 1.5L 0.5L2 6 -1.5L -6 -1.5L -1.5/. 0.5L2 1.5L 0.25L -6 1.5L 6 1.5L -1.5L 0.25L 1.5L 0.5L2 N)_ ^i(1) U20 Ui>_ h<2)] 0™ Lf Uf] re» 1 M,(l> fiH .M^J ffiPl <H fl2W Uf>J (7.112a) (7.112b) There are a total of eight equations in 16 variables. However, some of the variables are duplicative and others are related. In particular, we have the following eight conditions (see Fig. 7.7): Boundary Conditions w U) = 0, w. (2) = 0, M1(,)=0, M<2)=0. (7.113a) Continuity Conditions w. (i) w (2) ft<l) = 3 (2) (7.113b) Equilibrium Conditions Q2[) + Q? = F6, M^[) + M[2) = 0. (7.113c) Imposition of conditions in Eqs. (7.113a,b) is straightforward. However, to impose the equilibrium conditions (7.113c), we must add the third equation in (7.112a) to the first equation in (7.112b) and then, using the first condition of Eq. (7.113c) (conditions in Eqs. (7.113a,b) are also used], we obtain \6ET where w(2l) = W2, 6>,(,) - (1.5L@i + 12W2 - 1.5L©3) = F0, (7.114a) ©i, 6{2V) - ©2, and 02<2) = ©3. Similarly, adding the fourth equation in (7.112a) to the second equation in (7.112b) and then using the second condition in (7.113c), we obtain —— (0.25L©| + L2W2 + 1.5L©3) = 0. (7.114b)
GENERAL BOUNDARY-VALUE PROBLEMS 245 Thus, we have a total of six equations: first two equations of Eq. (7.112a), the last two equations of Eq. (7.112b), and two equations in Eqs. (7.114a,b) in six unknowns Q\\ ©1, W2, ©2> ®3> and <2;> . Solving the middle four equations for (©1, W2, ©2, ©3), we obtain 01 =- FpL2 16E7' W2 = FpL3 48E7' ©2 = 0.0, ©3 = FpL2 16E/' (7.115) and Q] ) and Q,\ ' can be computed to be —0.5Fo from the first equation of (7.112a) and the last equation of (7.112b). The rotation ©2 at the center of the beam is correctly predicted to be zero (why?). The four-parameter Ritz solution [but with the same polynomial degree as in the two-parameter solution (7.95a,b)] becomes w4(x) = FqL3 48£7 <$^&vmm-i$ 0 < X < <-*</„ (7.116) where x = x — L/2. Simplification of Eq. (7.116) gives the expression in Eq. (7.99); thus, the Ritz solution matches with the exact solution. The procedure described in this example closely resembles that of the finite element method, which will be discussed in detail in Chapter 9, The procedure is valid for all beam problems irrespective of the nature of the loading and boundary conditions. In addition, the procedure gives exact values of the deflection wq(x) and rotation —(divo/dx) Sit the end points of each part for all loads and boundary conditions, provided that the flexural rigidity EI is constant within each part. 7.4 GENERAL BOUNDARY-VALUE PROBLEMS All the examples presented in Section 7.3.3 utilized the minimum total potential energy principle or Hamilton's principle. For problems outside the field of solid and structural mechanics, the construction of an analog of the minimum total potential energy principle is needed to derive the Ritz equations. Here the procedure for constructing weak forms from differential equations and use of these statements in the Ritz approximation are studied. 7.4.1 Variational Formulations The steps involved in the weak- formulation of differential equations arc described with the aid of three model equations: (1) a second-order equation in one dimension, (2) a fourth-order equation in one dimension, and (3) a second-order equation in two dimensions. These equations are quite general and they arise in a number of fields in engineering and applied sciences. Most of the ideas in developing a weak form are
246 DIRECT VARIATIONAL METHODS presented in connection with Model Equation 1. Model Equation 2 is used to illustrate the weak form development for a higher-order equation, and Model Equation 3 is used to extend the ideas to two dimensions. Model Equation 1 Consider the problem of finding the function u(x) that satisfies the differential equation -— (ci(x)—\ + c(x)u ~f = 0 for 0 < x < L, (7.117a) dx \ dx ) and the boundary conditions ( du\ «(()) = «o, \cll~') = Py (7.117b) x=L where a = a(x)t c — c(x), f = /(*), wo, and P are the data (i.e., known quantities) of the problem. Equation (7.117a) arises in connection with the analytical description of many physical processes. For example, conduction and convection heat transfer in a plane wall or fin (ID heat transfer), flow through channels and pipes, transverse deflection of cables, axial deformation of bars, and many others. Table 7.2 contains a list of several field problems described by (7.117a) when c(x) = 0. The mathematical structure common to different fields is brought out in this table. Thus, if we can develop a numerical procedure by which Eq. (7.117a) can be solved for all physically possible boundary conditions, the procedure can be used to solve all field problems listed in Table 7.2, as well as many others. This fact provides us with the motivation to use Eq. (7.2) as a model second-order equation in one dimension. The motivation for the development of the weak form and a step-by-step procedure for the weak form development are discussed next. In the Ritz method, we seek an approximate solution to Eq, (7.117a) in the form N UnM = J^cj4>i(x) + <Pq(x)9 (7.118) 7 = 1 and determine the unknown parameters Cj such that Eqs. (7.117a) and (7.117b) are satisfied by the //-parameter approximate solution U^. For example, suppose that L — 1, a = x, c = 1, wo = 1, / = 0, and P = 0; then we could take N = 2 and write the approximate solution of (7.117a) in the form {<f>\ = x2 - 2x, <f>2 =*3 ~ 3a\0o — 1): uf*UN= a(x2 - 2x) + c2(x3 - 3x) + 1, which satisfies the boundary conditions (7.117b) of the problem for any values of c\ and C2. Then the constants c.[ and C2 must be determined such that the approximate solution C/yv satisfies Eq. (7.117a): d ( du\ rt , x -— [x— )+u = Q (a) dx \ dx/
GENERAL BOUNDARY-VALUE PROBLEMS 247 Table 7.2 Some examples of second-order equation (7.117a) in one dimension: : / for 0 < x < L d ( du\ EBC: tf(0) = i/0; NBC: (a — \ ch xJx=L Field 1. Cables 2. Bars 3, Heat transfer 4. Pipe flows 5. Viscous flows 6, Seepage 7. Electrostatics Primary Variable it Transverse deflection Longitudinal displacement Temperature Hydrostatic pressure Velocity Fluid head Electrical potential Coefficient* a T EA k 7tD4 128 At M e 6 Source Term / Distributed vertical force Distributed axial force Internal heat generation Flow source Pressure gradient Fluid flux Charge density Secondary Variable P Axial load Axial load Heat flux Heat Flow rate Stress Flow Electric flux '<*E ~ Young's modulus; A = area of cross section; D = diameter of the pipe; k — thermal conductivity; li = viscosity; T — tension; e = permeability; e = dielectric constant. in some sense. If we require Up? to satisfy the above equation in the exact sense, we obtain dU,\i d2Ufj o 0 + UN = -2c{(x - 1) - 3C2(x2 - I) - 2c\x - 6c2x2 dx — x- dx2 + a {xl - 2x) + 6'2<y - 3*) + 1 = 0. 3 Since this expression must be zero at all jc, the coefficients of the various powers of x must be zero: 1 + 2ci + 3c2 = 0, -(6c i +3c2) = 0l a - 9c2 = 0, C2 = 0. The above relations are inconsistent; hence there is no solution to the equations. If we were able to find a unique solution to these equations, then Un = c\ (x2 — 2x) + C2(x3 — 3x) -h 1 is the exact solution of the problem. Tt is not always possible for
248 DIRECT VARIATIONAL METHODS arbitrary data of the problem to find the exact solution. An alternative is that we may require the approximate solution Um to satisfy the differential equation (a) in a weighted-integral sense, / w{x)Rdx = 0, (b) Jo where R is the residual (i.e., error) in the differential equation, dUN d2UN dx dx2- and w(x) is a function, called the weight function, which is introduced to provide as many independent relations among c/ as there are unknown parameters cj (j = 1,2,..., N). This is accomplished by selecting N independent functions for w(x). For example, if we take the two choices w(x) = 1 and v)(x) = x9 which are linearly independent, we obtain 0 = / 1 • R dx = (1 + 2ci + 3c2) + £(-6cj - 3c2) + £(n - 9c2) + |c2, Jo 0 = / x • Ad* = £(1 + 2ci + 3q) + ^(-6c, - 3c2) + \{c\ - 9c2) + £c2, Jo or 2„ , 5^ | 3„ , 31, 1 ici + ±C2 = l, fa +%C2=$. (C) These equations provide two linearly independent relations for C[ and c2 that can be solved to obtain unique solution, c\ = 222/23 and c2 — —100/23, and the approximate solution U2(x) becomes U2(x) = c](x2-2x)+c2(x3~3x) + l = ?§(x2-2x)-l§(x3-3x)+\. (d) Thus, integral statements of the type in (b) provide means for obtaining as many algebraic equations as there are unknown coefficients in the approximation. There are several variational methods, in addition to the Ritz method, in which approximate solutions of the type u ^ Yl cj(l)j + 00 arc sought, and the coefficients cj arc determined, as shown above, using an integral statement. These methods differ from each other in the choice of the weight function w(x) and the integral statement used, which in turn dictates the choice of the approximation functions <f>j and 0o- For the moment we deal with the Ritz method of approximation. As discussed above, the necessary and sufficient number of'algebraic relations among the cj's can be obtained by recasting the differential equation (7.117) in a weighted-integral form: o=fwwh£(H£)+<"-']*• ail9)
GENERAL BOUNDARY-VALUE PROBLEMS 249 where w(x) denotes the weight function, which for the moment is arbitrary. For u(x) ss Un (x) and each independent choice of w(x), we obtain an independent algebraic equation relating all cy. A total of N independent equations are required to solve for the N parameters c/. When a weighted-integral statement like Eq. (7.119) is used to obtain the N equations among c/, the method is known as the weighted-residual method, and it will be discussed in Section 7.5. Note that the use of Eq. (7.119) precludes that Un(x) satisfies all specified boundary conditions and is differentiable as many times as required in the differential equation. To weaken the continuity (i.e., differentiability) required of U^(x) and therefore of <pj(x)9 we trade the differentiation in (7.119) from u to w such that both u and w are differentiated equally—once each in the present case. The resulting integral form is termed the weak form of (7.117). This form is not only equivalent to (7.117) but it also contains the natural boundary conditions of the problem, and therefore /7yv(A') need not satisfy the natural boundary conditions. The three-step procedure of constructing the weak form of (7.117) is discussed next. The first step is to multiply the governing differential equation with a weight function w and integrate over domain (0, L), giving Eq. (7.119). The second step is to trade differentiation from u to t/;, using integration by parts. This is achieved as follows. Consider the identity —w d / du\l _ d / du\ dw du dx \ dx)\ dx\ dx) dx dx' which is simply the product rule of differentiation applied to the product of two functions, a(du/dx) and w. Integrating this identity over the domain, we obtain _ fL \jL( —\] /.-_ fL jL( —\i. fL —— Jo \_dx \ dx)\ " Jo dx \ dx) Jo dx dx f duV' fL dwdu , = - wa~- + / a-——dx. (7.120b) L dx J0 Jo dx dx Substituting (7.120b) into (7.119), we arrive at the result fL ( dwdu \ , T duV' 0=1 a—--— + cum - wf dx -\w- a— . (7.121) J0 \ dx dx ' / L dx]0 The third and last step is to identify the primary and secondary variables of the variational (or weak) form. This requires us to classify the boundary conditions of each differenti al equation into essential (or geometric) and natural (or force) boundary conditions. The classification i-s made uniquely by examining the boundary term appearing in the second step of the weak form development, namely, Eq. (7. J 21): T duV'
250 DIRECT VARIATIONAL METHODS = P. (7.122a) As a rule, the coefficient of the weight function in the boundary expression is called the secondary variable, and its specification constitutes the natural or Neumann boundary condition. The dependent unknown in the same form as the weight function in the boundary expression is termed the primary variable, and its specification constitutes the essential or Dirichlet boundary condition. For the mode) equation at hand, the primary and secondary variables are da _ u and a— = Q. dx Next, we denote the secondary variables at the end points by some symbols: -«-(■£) L °-('£) With the notation in (7.122a), the variational form becomes 0= f (a—^+cwu-wf)dx-w(0)Qo-w(L)QL. (7.122b) Jo \ dx dx J Now is the time to discuss the conditions on the weight function w(x). Clearly, it should be differentiable at least once (like u(x) is). The Ritz method uses the weak form, with w = <£/ to obtain the /th equation of the set of N relations among c/'s. Thus, we require w(x) to satisfy the homogeneous form of specified essential boundary conditions (i.e., to belong to the set of admissible variations). In the present case, w(x) should be once differentiable and vanish at x = 0. Hence, the final weak form is 0= / (a — ^+cwu-wf)dx-w(L)P. (7.123) Jo \ dx dx J This completes the three-step procedure of constructing the weak form. The weak form in (7.123) contains two types of expressions: those containing both w and u; and those containing only w. We group the former type into a single expression, called the bilinear form: f1' ( dw du \ B(w, u)= f a—— + cwu J dx- (7«124a) We denote all terms containing only w (but not u) by l(w), called the linear form: l(w)= f wfdx + w(L)P. (7.124b) Jo The statement (7.123) can now be expressed as one of finding u from the set of admissible functions (i.e., differentiable at least once and satisfying the essential boundary conditions) such that the variational problem B(w,u) = l(w) (7.125)
GENERAL BOUNDARY-VALUE PROBLEMS 251 is satisfied for all w from the set of admissible variations. As seen before, the bilinear form results directly in the coefficient matrix, while the linear form gives rise to the right-hand-side column vector of the Ritz equations. Since £(-, •) is symmetric, the functional associated with the variational problem (7.125) is given by Eq. (7,53) (which represents the total potential energy in the case of bars or cables): J(u) = -B{uyu)-l{u) J0 2 \dx )+r' dx -f Jo wfdx-w(L)P. (7.126) The weak form development up to Eq. (7.125) is valid even for the case in which the coefficients a and c are functions of the dependent variable w, making the problem nonlinear. In that case, B(iv, u) will be no longer a bilinear form and the functional J(u) may not exist. However, to use the Ritz method one needs only the variational statement (7.125), which always exists. Model Equation 2 Here we consider a fourth-order differential equation of the form dx2 («*£) -j- c(x)u = /, 0 < x < L, (7.127) We will not be concerned with any specific boundary conditions, as the weak form development naturally leads to the classification of the variables into primary and secondary type, and their specification constitutes, respectively, the essential and the natural type. This equation arises, for example, in connection with the bending of straight beams, where u{x) denotes the transverse deflection, b(x) = E(x)I(x) the bending rigidity, c{x) = k the foundation modulus (if any), and f(x) the distributed transverse load. The first two steps in the development of the weak form of the equation are summarized below. Note that in this case we must transfer two derivatives to the weight function so that both w and u are required to have the same order of continuity (or differentiability). Step 1: °=lLMU%h CM - f dx. (7.128) Step 2: n fL[ dw d /tA\ "I , f d (,d2u\
252 DIRECT VARIATIONAL METHODS d2W d2U fL/,d2wd2u \ f d (,d2u\\L [dw ,d2ulK It is clear from the boundary expressions that the secondary variables are -'2-^ d2u A (1 fji\ — (7.129) (7.130a) In the case of beams they represent the shear force and bending moment in the beam. The primary variables are [w -> u and (dw/dx) -> (du/dx)] w, dx (7.130b) Step 3: Denoting the shear forces and bending moments at the two ends of the beam as (proper signs arc inserted to make all of the Q's and M's have the negative sign in the weak form; this also happens to be the correct definition of the bending moments and shear forces on the left and right ends of the beam): [dx V d*2/Ir=o ~ °' L ^21y=o (7.131) Then Eq. (7.129) takes the final form « fL (, d2u> d2u J\ , 0=1 [ b—^r —=• + cwu - wf]dx — Jo \ dx2 dx2 ' / ML. w(0)Qo-w(L)QL (-S)h-(-S) ML. (7.132) The bilinear form, linear form, and functional for the problem arc w- (d2wd2u , , f' (,d2vi)d2u \ , (w, u) = I b-r-r-rs + cwu I «•*•"> Jo \ dx2 dx2 } = j wf dx + w{0)Q0 + w(L)QL + (—£■) . fL\b (d2u\\c 2 " ( du ■u(0)Qo-u(L)QL-l — /(»;) M0 + (_dw\ \ dx) M,„ M0 (-S) ML. (7.133) In the case of beam bending, F(u) is nothing but the total potential energy T7(w).
GENERAL BOUNDARY-VALUE PROBLEMS 253 The weight function w(x) is required to be twice differentiable and to vanish at the points where it and to du/dx are specified. Equation (7.132) can be specialized to any beam with specific boundary conditions and loading. Model Equation 3 Lastly, we consider the problem of determining the solution w(a% v) to the partial differential equation 9 / du\ 3 / du\ r . n ,„, ^ -alK)-^K)+fl0W = / m* al34) in a two-dimensional domain R. Here ao, ci\, a%, and / are known functions of position (x, y) in /?. The function u is required to satisfy, in addition to the differential equation (7.134), certain boundary conditions on the boundary S of R. The variational formulation to be presented tells us the precise form of the essential and natural boundary conditions of the equation. Equation (7.134) arises in many fields of engineering, including in 2D heat transfer, stream function or velocity potential formulation of inviscid flows, transverse deflections of a membrane, and torsion of a cylindrical member. The three-step procedure applied to Eq. (7.134) results in the following equations: Step J: °-fA-£ (fli£)_ h (a%)+aou ~f]dxdy- (7J35) Step 2: f ( 3w du dw du \ , f 0 = / I «i — — + a2 — — + aowu - wf ) dxdy JR\ ox dx dy dy J 1 ( dlf $u \ -p Un—n* + ai-r-ny 1 ds9 (7.136) where we used integration by parts [or the Green-Gauss theorem, Eq. (2.89)] to transfer differentiation to w so that both u and w have the same order derivatives. The boundary term shows that u is the primary variable while Bu du ct\—nx +a2—> ax dy is the secondary variable. Step 3: The last step in the procedure is to impose the specified boundary conditions. Suppose that u is specified on portion S\ and the natural boundary condition is
254 D! RECT VARIATIONAL METHODS specified on the remaining portion S? of the boundary: u = it on S\, ai —nx -h ci2 —nv = 8 on $2. (7.137) dx ay ' Then w is arbitrary on S2 and equal to zero on S\. Consequently, Eq. (7.136) simplifies to f ( dw du dw du \ , , f A , 0 = / ai ^T" ^ + fl2 — — + flow« - w/ <^J' - / w# rfJ. (7.138) JR \ dx dx dy dy } Js2 The bilinear form, linear form, and functionals are f ( dw du dw du \ B(w.u) = / ( a\ {"^2~^ hciQWii \dxdy, (7.139a) JR \ dx dx dy dy ) l(w)= J wfdxdy+ J wgds, (7.139b) Jr Js2 r x 1 /• r /9"\2 (du\2 dxcly - f ufdxdy - f gu ds. (7.139c) JR Js2 Once again, 1 (u) represents the total potential energy in the case of a membrane problem. 7.4.2 Ritz Approximations The Ritz method can be applied directly to the weak forms (7,124), (7.132), and (7.138)—or into the variational problem B(w,u) = l(w), (7.140) Here we consider the case in which £(•, -) is a bilinear form. Substituting N u**Un = Y1CJ^J + 0o> w = <t>t, (7.141) 7 = 1 into Eq. (7.140) to obtain the /th equation N J^aijcj = bh 7 = 1,2 //, ' (7.142a) where afj - B(fc, 0y-), ft, = /(&) - * «>o, 0,-). (7.142b)
GENERAL BOUNDARY-VALUE PROBLEMS 255 The specific expressions of A,y and b\ of each model equation can be written using the respective bilinear and linear forms. For example, for the third model equation, we have Clij = JR{a] ^^+a2^^7 + ao<pi<Pi)dxdy" f ( 3<j>i d(j)o Scpi d<po \ f h = / I 4>if - «i-—z "2 — - «o^/0o dxdy + / fcg dS. Jr\ 3* ox dy dy ) Js2 (7.143) Once fa and <po are selected, subject to the conditions stated in Section 7.2, the coefficients of matrix [A] and column vector [b] can be computed, and the linear algebraic equations in Eq. (7.142a) can be solved for the Ritz coefficients. We consider specific examples next. Example 7.9 We wish to solve the partial differential equation (Laplace's equation) fd2u d2u\ subject to the boundary conditions (see Fig. 7.8): u = 0 on x = 0, 1 and y = 0, u = sin 7i x on y = 1. (7.144b) The weak form of the equation can be obtained as a special case from Eq. (7.138) by setting / = 0, ci[ =ci2= 1, ao = 0, and ^2 = 0 (S\ = S): Jo Jo \ Sx dx dy dy) Figure 7.8 Domain and boundary conditions of the problem in Example 7.9.
256 DIRECT VARIATIONAL METHODS The Ritz equations are given by Eq. (7.142a) with a,j and bj given by J k h \Bx dx dy dy) Jo Jo \ &x dx $y <>y / (7.146) Next we select <j>$ and 0/ for the problem. The function 0o is required to satisfy the boundary conditions in Eq. (7.144b) because all of them are of the essential type. The following choice for 0o meets the conditions: 00 = )> sin jtx. (7 J 47a) The functions 0;(/ = 1,2,..., n) arc required to satisfy the homogenous form of Eq. (7.144b) and to be linearly independent. We choose 01 = sin7TA-sin7T3>, 02 = sin 7zx sin 27T v, 03 = sin lixx sin jt v, etc. (7.147b) From Eq. (7.146) we obtain «,-[> b, = 2' 0. if* =j> if i = 1, if i ^ 1. (i.e., [A] is diagonal), (7.147c) Hence, the Ritz solution becomes (c[ = —1/jt and all other c,- = 0): Ui(x, y) = ci0i(a% ;0 + 4>o(x, y) 1 y sin 7TX sin nx sm ny = sin n x 7X f y — — sinTr^j . The exact solution of Eq. (7.144a) is given by sin nx sinh ny u(x,y) = sinh7r (7.148a) (7.148b) Example 7.10 Consider the following pair of coupled differential equations, which arise in connection with the Timoshcnko beam theory (see Exercise 6.15): dt2 ' 2„ (7.149) (7.150)
GENERAL BOUNDARY-VALUE PROBLEMS 257 where S is the shear stiffness (S = KSGA\ Ks is the shear correction coefficient, G the shear modulus, and A the area of cross section), D the bending stiffness, wo the transverse deflection, <px the rotation, q the distributed transverse load, and Iq and I2 are mass inertias. Assume that Z), S, Iq, and I2 are constants. The "specified" boundary conditions arc of the form (as will be clear from the third step of the weak form) -("EL-*- Kh-»- "•»"> -*(^7+*.)t 0=0>. s(^+*,)ti-fi2. <7.151b) Wc wish to derive the Ritz equations of the problem. First we develop the weak form of the equations using the three-step procedure. Multiply the first equation with weight function V[ and the second one with weight function v2 and integrate over the length of the beam to obtain = ]0 \-sii{-9r++*)+viq-Im-wr + Qm(.0) + Q2v\(L), (7.152a) + M\ v2(0) + M2v2(L). (7.152b) Note that integration by parts was used such that the expression wo,.v +0* is preserved, as it enters the boundary term representing the shear force. Such considerations can
258 DIRECT VARIATIONAL METHODS only be used by knowing the mechanics of the problem at hand. Also, note that the pair of weight functions (v], vi) (from a product space of admissible variations) satisfy the homogeneous form of specified essential boundary conditions on the pair (wo> <£*•) (with the correspondence V[ ~ wq and V2 ^ <t>x)- Writing the bilinear form for the problem is a little involved; one may treat u — (wo, <t>.\) and v = (v\, 1*2) as vectors (from a vector space) and then write the bilinear form B(v, it). Wc assume Ritz approximations of the form M N (7.153) and derive the ordinary differential equations involving the time derivatives of d\ and cv. From weak forms (7.152a,b) it is clear that all of the specified boundary conditions are of the natural type. Hence ^oto = 0 and Oo(x) — 0. Next wc substitute the approximations into the weak forms for wo(x) and <f>jc(x)9 and v\(x) = irt{x) and V2(x) = 0j(x), and obtain -s"-£ It t£*+p'e'j+*'*' - ** \t "S*> Jo d2dj IT2 dx y-i y=i y=i Jo de, /A«w, t M l^dfj -^Ig^l-^IE^+E^ ;=i 7=1 ,/=i ' ai (7.154) (7.155) where A. Jo dx dx Jo d* } = [ Iofr^jdx, Fl=\ t^dx + Qiiri(Q) + Q2fi(L), Jo Jo
WEIGHTED-RESIDUAL METHODS 259 °>i = [s,"d-td*- "w-fC"^* **')*•• (7'156) Mfj = f hOiOj dx, Ff = M[9i(0) + M20i (L). In matrix form, we can write this as [A] \B] [C] \D] 1 f{rf}l [IM1] [0] 1 Ud}\_ UF])\ JlWr [[0] |M2jJi^)-l{F2})' (7.157a) or [K]{A} + [M]{A) = {Fl (7.157b) In closing this section, we make a couple of additional comments on the Ritz method. In developing the weak form, one should bear in mind that the boundary terms obtained from the integration by parts should be physically meaningful. The variational form used for the Ritz method does not have to be a quadratic functional, but it should be a form that includes the natural boundary conditions of the problem. Therefore, the Ritz method can be applied even to nonlinear problems. Of course, the resulting simultaneous algebraic equations are nonlinear and there can be more than one solution to the equations (see Example 7.20). The selection of approximation functions becomes increasingly difficult with the dimension and shape of the domain. 7.5 WEIGHTED-RESIDUAL METHODS 7.5.1 Introduction As discussed earlier, weighted-residual methods are those in which we seek approximate solutions using a weighted-integral statement of the equation(s). To fix the ideas, consider a boundary-value problem described by the operator equation A{u) = f in ft, (7.158a) subjected to boundary conditions Bi(u) = ii on Fi, B2(u) = g on r2, (7.158b) where A is a linear or nonlinear differential operator, u is the dependent variable, / is a given force term in the domain ft, B\ and B2 are boundary operators associated with essential and natural boundary conditions of the operator A, and it and g are specified values on the portions Fi and-F2 of the boundary F of the domain. An example of Eqs. (7.159a) and (7.159b) is given by d ( du\ du A(w) =—— [a— , Bi(u) = u, B2(u)=a — , dx \ dx } dx
260 DIRECT VARIATIONAL METHODS V\ is the point x — 0, and F2 is the point x = L, The weighted-integral statement of Eq. (7.159a) is 0 = /" w(x) [A(it) - f) dx. (7.159) We seek a solution Un that satisfies the specified boundary conditions (7.158b) such that the above equation is also satisfied. A complete discussion of the procedure will be given shortly. The weak form of the operator equation (7.158a) can be constructed whenever the operator A permits the use of integration by parts [or gradient/divergence theorems in Eqs. (2.89a,b)] to transfer part of the differentiation from the dependent variable u to the weight function w and incorporate the natural boundary conditions of the problem. In general, the weak form can be constructed if the operator A is expressible as a product of two operators: A = T*(aT) (7.160a) where operator T* is called the adjoint of T and is related to T by / T(u)v rfx = / uT*(v) dx + <p Ci(w)C2(v) dS (7.160b) Jo. J a Jr for all u and v. Here £2 is domain with boundary F. The boundary operators C\ and C2 depend on the operator A. For example, when A = —(d/dx)\a(du/dx)]y operators 7\ 7*, Ci, and C2 are given by dx dx If A = —V2 = —V ■ V, then we have [u must be a scalar and v a vector function of position in Eq. (7.160b)]: T = V (grad), 7* = -V • (div), d = 1, C2 = n- . The weak form of Eq. (7.158a,b), with A given by Eq. (7.160a), can be derived as follows: w\A(u)-f]dx = f w[T*(aT(u))-f]dx = J \T(w)(aT(u)) - wf]dx - £> B{(w)B2(u) dS,
WEIGHTED-RESIDUAL METHODS 261 where C\ = B\ and B2 = C2(aT(u)). Since B\ (w) = 0 on portion Ti (where B[(u) is specified), and £200 = g on T2, with Ti -f Ta = T\ wc have 0= f {aT{w)T{u)-wf}dx- f Bi(w)g dS, which is the weak form we set to derive. Returning to the weighted-residual methods, we seek an approximate solution of u in the form "(as in the Ritz method) n UN{x) = Y,Cj<l>j{x) + <h{x), (7.161) 7=1 where the parameters c/ are determined by requiring that the residual in the governing equation due to the approximation Kn = aIJ2 Cj4>j + 0o I - / ^ 0 (7.162) be orthogonal to a set of TV linearly independent weight functions ^(x), which in general are different from the approximation functions <£,•: / innN(xt {c}, {0}, f)dx = 0, (1 = 1,2,..., N). (7.163) Jn Equation (7.163) is the same as that obtained by substituting approximation (7.161) into the weighted-residual statement (7.159). Tt provides N linearly independent equations for the determination of the parameters c/. If A is a nonlinear operator, the resulting algebraic equations will be nonlinear. The approximation functions (0o, 0r) and weight functions V/ in a weighted- residual method must satisfy the following conditions: 1. (f>j(j = 1, 2,..., N) should satisfy three conditions: (a) Each 0, is continuous as required in the weighted-residual statement; i.e., 4>j should be such that Un yields a nonzero value of A(Un). (b) Each 4>j satisfies the homogeneous form, of all specified (i.e., essential as well as natural) boundary conditions. (c) The set [<pj} is linearly independent and complete. 2. 0o has the main purpose of satisfying all specified boundary conditions associated with the equation. It is necessarily zero when the specified boundary conditions are homogeneous. 3. The set {1/0} should be linearly independent. (7.164)
262 DIRECT VARIATIONAL METHODS There are two main differences between the approximation functions used in the Ritz method and those used in weighted-residual methods: 1. Continuity. The approximation functions used in the weighted-residual methods arc required to have the same differentiability as in the differential equation, whereas those used in the Ritz method must be differentiable as required by the weak form. 2. Boundary conditions. The approximation functions used in the weighted- residual method must satisfy the homogeneous form of both geometric and force boundary conditions, whereas those used in the Ritz method must satisfy the homogeneous form of only the essential boundary conditions, since the natural boundary conditions are already included in the weak form. Both of these differences require 0/ to be of a higher order than those used in the Ritz method. Various special cases of the weighted-residual method differ from each other due to the choice of the weight function, ^r,-. The most commonly used weight functions are: The Petrov-Galerkm method: ■*/// ^ 4>i • Galerkin *s method: if, = (pi. (7.165) Least-squares method: yffj = A (<£/). Collocation method: \jfi = 5(x — x,). Here 8 (•) denotes the Dirac delta function. Although the least-squares method is listed as a special case of the weighted-residual method here, it is based on the concept of minimizing an integral statement. In general, the least-squares method is not a special case of the weighted-residual method. These remarks will be discussed in more detail below. In addition to the methods listed above, there are other variational methods (methods in which the unknown parameters c,- are adjusted such that the governing equations are satisfied in a certain sense). These include the subdomain method and Trefftz method. These methods will also be discussed briefly in this chapter. 7.5.2 Galerkin's Method The Galerkin method is a special case of the Petrov-Galerkin method in which the approximation functions and the weighted functions are the same (<fr = ^). Hence, the Galerkin integral is given by / 4>i1lN(x, {c}, {<£}, f)dx = 0, (1 = 1,2,..., N). (7.166) If the Galerkin method is used for second-order or higher-order equations, it would involve the use of higher-order coordinate functions and the solution of nonsymmetric equations.
WEIGHTED-RESIDUAL METHODS 263 The Ritz and Galerkin methods yield the same set of algebraic equations for the following two cases: 1. The specified boundary conditions of the problem are all of the essential type, and therefore the requirements on fy and <po in both methods are the same. 2. The problem has both essential and natural boundary conditions, but the coordinate functions used in the Galerkin method are also used in the Ritz method. 7.5.3 Least-Squares Method The least-squares method is based on the idea of minimizing the integral of the square of the residual: minimize I{c\, cj,..., cjv) = / 7£^(x, {c}, {<£}, /) dx. (7.167a) We obtain (from SI = 0) °= f ?^(x> M> i4>l f) **> (7.167b) which, when A is a linear operator, becomes 0 = I A(4>i)KN(x, {c}, {<£}, /) dx. (7.167c) Clearly, Eq. (7J 67c) is a special case of Eq. (7.163) with fi = M4>i)- The least- squares method is more suitable for first-order equations. For eigenvalue problems and time-dependent problems it is not suitable, as shown below. On the other hand, the least-squares method is the only other method, in addition to the Ritz method, that is based on the minimization of a functional. The least-squares method also results in a positive-definite coefficient matrix. 7.5.4 Collocation Method In the collocation method we require the residual to vanish at a selected number of points x' in the domain: KaKx'", {c}, {0}, /) = 0, (i -1,2,... ,.A0, (7.168a) which can be written, with the help of the Dirac delta function, as / «(x - x')fttf(x, M, {<£}, /) dx = 0, (i = l,2 N). (7.168b)
264 DIRECT VARIATIONAL METHODS Thus, the collocation method is a special case of the weighted-residual method (7.163) with Y>7(x) = <$(x — x'). In the collocation method, one must choose as many collocation points as there are undetermined parameters. In general, these points should be distributed uniformly in the domain. Otherwise, ill-conditioned equations among cj may result. 7.5.5 Eigenvalue and Time-Dependent Problems Tt should be noted that if the problem at hand is an eigenvalue problem or a time- dependent problem, the operator equation (7.158a) takes the following alternative forms: Eigenvalue Problem A(u)-XC(u) = 0. (7.169) Time-Dependent Problem Mu) + A(u) = f(x,t). (7.170) In Eq. (7.169), X is the eigenvalue, which is to be determined along with the eigenvector u(x), and A and C are spatial differential operators. An example of the equation is provided by the buckling of a beam column d2u d2 ( dlu\ ¥ dzu ^ where u denotes the lateral deflection and P is the axial compressive load. The problem involves determining (he value of P and mode shape u(x) such that the governing equation and certain end conditions of the beam are satisfied. The minimum value of P is called the critical buckling load. Comparing the above equation with Eq. (7.169), we see that A' d2 /Td2u\ d2u\ , „ ^, , d2u Tn Eq. (7.170), A is a spatial differential operator and At is a temporal differential operator. Examples of Eq. (7.170) are provided by the equations of heat transfer in a plane wall and axial motion of a bar, respectively: du d / du\ pCv^-^{k^) = f(x't)'
WEIGHTED-RESIDUAL METHODS 265 In the first equation, u denotes temperature, p the density, cv specific heat at constant volume, k the conductivity, and / is internal heat generation. Clearly, we have du d ( du At{u) = pcv — , A{u) - —— I teat ax \ ax In the second equation, u denotes the axial displacement, p the density, E Young's modulus, Aq area of cross section, and / body force per unit length. In this case, we have d2lt ,, x 3 / au\ dtz dx \ dx J Application of the weighted-residual method to Eqs. (7.169) and (7.170) follows the same idea, i.e., Eq. (7.163) holds. However, two comments are in order. 1. In the case of time-dependent problems, the integral in Eq. (7.163) is still over the spatial domain and the weight function \j/ is a function of the spatial coordinate only. Thus, Eq. (7.163) leads to a set of ordinary differential equations in time among cj(t). These need to be further approximated using lime-approximation schemes. 2. In the case of the least-squares method, the question arises as to what should be the weight function if/j. Let us examine the least-squares method first for the linear eigenvalue problem (i.e., A and C are linear operators). Wc have 0 = 8 f ll2N Jo, dx, TlN = A(UN) - XC(UN) £2 [4(0/) - XCifr)} KN dx, (i = l,2 N) = 2]C {/ [A^ " XCWM [Mtj) - AC(0,)] dx\ cj. Clearly, the eigenvalue problem becomes quadratic in X, which is not desirable from a computational viewpoint. In the case of time-dependent problems, we have 0 = 8 [ n2N dx, KN = A(UN) + At(UN) =/ N 3 M<f>i) + J2</>k—At(ck) TlNdx, (i = 1,2, ...,N). which is also complicated. Thus, the least-squares method leads to complicated systems of equations for eigenvalue or time-dependent problems. An alternative is to use V// = A{<t>i) in all cases. This avoids the problems seen above.
266 DIRECT VARIATIONAL METHODS It is possible to develop the so-called space-time approximations, i.e., to use the variational methods, treating time as an additional coordinate. It is found that such approaches are complicated, and they are not used in practice. Therefore, we will not consider the space-time approximations in the present study. 7-5.6 Equations for Undetermined Parameters Here we develop the discrete equations for the equilibrium, eigenvalue, and time- dependent problems under the assumption that A is a linear operator, as is the case with most problems considered in this study. The equations arc valid for all special cases of the weighted-residual method except for the least-squares method. They are also valid for the least-squares method if one accepts the use of A (<£/) for t/t/? which is certainly true for equilibrium problems. For eigenvalue problems we assume that all boundary conditions are homogeneous and therefore </>q = 0. Equilibrium Problems We have (N \ N J2cj4>j+4>q\ =YtcJA(4>j) + A{<M, (7.171) and Eq. (7.163) for the equilibrium equation (7.158a) becomes JT \f ifiA(<t>j)dx\ cj - j fiU - A(0o)]dx = 0 or N J2 ctijcj - bi = 0 (i = 1,2,..., tf), [A]{c] = {&}, (7.172a) where ai} = [ i,iA((Pj)dx, bi = f ft 1/ - A(<M}dx. (7.172b) Note that ay is not symmetric in general, even when fy = 4>\ (Galerkin's method). It is symmetric in the least-squares method because rj/j = A{<j)j). Also, if (a) A is an operator of the form in Eq. (7.160a), (b) ^ = 4>i, and (c) <j>{ satisfy the homogeneous form of specified essential and natural boundary conditions, it can be shown that au = / r(0/)a71(0/)dxt */ - I <t>ifdx+ f frgdS. . (7.172c) Jo, Jn Jr2
WEIGHTED-RESIDUAL METHODS 267 Eigenvalue Problems For linear eigenvalue problem of the form in Eq. (7.169), we have £ U filA^j) - XCWtfdxJ cj = 0 7=1 or N J2(aU-k8ij)cj=0V = h29...,N), ([A]^k[G]){c} = [0}9 (7.173a) 7 = 1 where ay are defined by Eq. (7.172b) and gii= [ ir;C(<t>j)dx. (7.173b) Note that using Eqs. (7.173a,b) for the least-squares method amounts to using ^y = A(0/); a\j is symmetric but gjj is unsymmetric. Time-Dependent Problems For linear time-dependent problem of the form in Eq, (7.170), we have J2 if MM<t>j)cj +4>jMcj)\dx\ - / fi\f - A(<t>0)]dx = 0 or N X>,/c, + myA,(c;)) = 0 (i = 1, 2, ..., N)9 UJ{c} + X\M]{At(c)} = {0}, 7-1 (7.174a) where a,j are defined by Eq. (7.172b) and muAt(cj)= / ilrtfjAticfidx. (7J74b) Again recall that using Eqs. (7.174a,b) for the least-squares method amounts to using fi = AM). 7.5.7 Examples A number of examples arc considered here to illustrate the use of various weighted- residual methods. Equilibrium, eigenvalue, and time-dependent problems are considered.
268 DIRECT VARIATIONAL METHODS Example 7.11 Consider the eigenvalue problem of Example 7.5 in a nondimen- sional form [see Eqs. (7.77a,b)]: d2u Xu = 0, u(0) = 0, du }-« = 0 at jc = 1. dx In the weighted-residual method, 0; must satisfy not only the condition 0/ (0) = 0 but also the condition 0;'(1) + 0/(1) =0. The lowest-order function that satisfies the two conditions is .2 <t>\(x) = 3x-2xz. (7.175) The one-parameter Galerkin's solution for the natural frequency (see Example 7.5) can be computed using ° = Ci/ol0,(^ + ^1)^ M V 3 '5 / 10 4\ (a) which gives (for nonzero c\) X = 50/12 = 4.167. If the same function is used for 0j in the one-parameter Ritz solution, we obtain, as discussed in Example 7.5, the same result as in the one-parameter Galerkin solution. If we use the one-parameter collocation method with the collocation point at x = 0.5, we obtain [0i(O.5) = 1.0 and (d2<p\/dx2) = -4.0]: O = ci0i(().5) dx2 ) (f± + A0i(O.5) v=0.5 or (-4 + X) a = 0, which gives X = 4. The one-parameter least-squares approximation with }f/\ = A(</>\) gives which gives X = 4.8. If we use t/q = A((f>\) — A0i, we obtain = (*A2-fX+16)c,, whose roots are A.1,2 = f ± gV'iiS ->- Ai = 7.6825, A2 = 0.6508. (b) (c) Neither root is closer to the exact value of 4.116. This indicates that the least-squares method with i//y = A(0/) is perhaps more suitable than t/t; = A(0,) — XC((f>j).
WEIGHTED-RESIDUAL METHODS 269 Let us consider a two-parameter weighted-residual solution to the problem U2(x) = ci0i(jc) + c2ct>2(x)y (d) where 0i (a) is given by Eq. (7.175). To determine 02(•*), we begin with a polynomial that is one degree higher than that used for <j>\: 02 (*) = a + bx + ex2 + dx3y and obtain #2(0) = 0 -> a = 0; ^2(D+02(1) = 0 -► 2fe+3c+4rf = 0ord = -\b~ \c. We can arbitrarily pick the values of b and c, except that they are not both equal to zero (for obvious reasons). Thus we have an infinite number of possibilities. If we pick b = 0 and c — 4, we have d = — 3, and 02 becomes 02(X) = a + bx + ex2 + dx3 = Ax2 - 3a-3. (7.176a) On the other hand, if wc choose b = 1 and c = 2, we have d — -2, and 02 becomes 02(X) =a + bx + ex2 + dx3 = * + 2a2 - 2x3 = <pi(x). (7.176b) The set {0i, 02} is equivalent to the set {0i, (p2}4 Note that U2(x) = ci0i (x) + c2<j)2(x) = c\ Ox - 2x2) + c2(4x2 - 3a-3) = 3cix + {-ley + 4c2)x2 - 3c2a3, = ci (3a - 2a2) + c2(* + 2x2 - 2a3) = (3c\ + c2) x + (~2d[ + 2c2) a2 - 2c2x3. Comparing the two relations, we can show that C{ = C\ — 0.5C2, ^2 = 1.5C2. Hence, cither set will yield the same final solution for U2(x) or A. Using 0i and 02 [from Eq. (7.176a)], wc compute the residual of the approximation as „ d2u2 - d2<t>x d24>2 .,._,_ . , R = ~~d^ ~ W2 = ~Cllx? ~C27^ -A(CI* +C2(h)
270 DIRECT VARIATIONAL METHODS For the Galerkin method, we set the integral of the weighted residual to zero and obtain fl f{ r d2<pi d2<f>2 1 0 = / <p\ (x)R dx = I fa (x) -C] —y - Q —^ ~ * (c\ 01 + C2<f>2) dx = K[\c\ 4- ^i2c2 - X (Afuci + W12C2), r1 r1 r ^/20i d2<j)2 1 o = / 02C*)/?tf* = / <p2(x) -ci -^-2- - c'2_^r " A fr^* + C2^ dx = ^21 Cl + /^22t'2 - A. (M21C1 + M22C2) • In matrix form, we have [*]{c}-X[M|{c} =={<)}, where f[ d2(b; fl First, for the choice of functions in Eqs. (7.176a) and (7.176b), we have **L__4. ^ _ 8 - IS,, dx1 dxL Evaluating the integrals, we obtain Ku=-j <t>i-^dx = J (3x-2x2)(4)dx = —, Kn = ~ [ 01 %rr dx = / Q* " 2x2)(-8 + J8jc) rfx = J, Jo ^ Jo 3 if21 = - / fa^rrdx = f (4x2 - 3x3)(4)dx = I Jo «xz Jo 3 ^22 = - / 02 ^ dx - / (4x2 - 3x3)(-8 + 18x)dx - ?§, Jo d*z Jo 15 M\\= I fa fa dx = / (3* - 2x2)(3x - 2x2) dx = -, Jo Jo 5 Mi2= / fafadx = / (3x-2*2)(4x2-3x3)dx = - = Af2i, Jo Jo 5 M22 = f 0202 dx = f (4x2 - 3x3)(4x2 - 3x3) dx = ^, Jo Jo 35
WEIGHTED-RESIDUAL METHODS 271 and /MT50 35] _ ^ [28 2i]\fcil Jo) v15L35 38J 35L21 i7J/i^r (o)* For nontrivial solution, c\ ^ 0 and ci ^ 0, we set the determinant of the coefficient matrix to zero to obtain the characteristic polynomial 675 which gives 1332 315 0 or 525- 148X4-5^ = 0, (f) X{ =4.121, A2 = 25.479. (g) Clearly, the value of X\ has improved over that computed using the one-parameter approximation. The exact value of the second eigenvalue is 24.139. Tf we were to use the collocation method, we may select x = 1 /3 and x = 2/3 as the collocation points, among other choices. We leave this as an exercise to the reader. Example 7.12 Next we consider the transient response of the problem discussed in Example 7.6. The governing equations are d2ii d2u _ w(0,f) = 0, du 'dx + it = 0 at x = L for all t > 0, with zero initial conditions. We use the one-parameter approximation u{x,t) ci(t)(p\(x) with <f)\(x) defined inEq. (7.175). For the Galerkin method, we obtain whose (exact) solution is (V50/I2 ^ 2.0412): 4 d2d 10 5 °r 5^ + T'1 = 6' (a) t'i(f) = A sin 2Mt + B cos2.04r + \. For zero initial conditions, u(x, 0) = ii(x, 0) = 0 [or c\ (0) = 0 and c\ (0) = 0], the total solution becomes {A — 0 and B = — ] /4): ui(x, t) = 4(1 - cos2.040(3x - 2x2). (b) For the onc-paramctcr collocation method with the collocation point at x = 0.5, we obtain 0 = 01(0.5) -Cl {*?) 0.5 + ^.(0.5) -1 or d2ci + 4ci = 1, (c)
272 DIRECT VARIATIONAL METHODS so that d(t) = A sin2/ + B cos 2/+ £, and the one-parameter collocation solution becomes Ul(x%t) = \(\- cos2/) (3* -2a2) . (d) Example 7.13 Consider the simply supported beam problem of Example 7.7. Since all specified boundary conditions are homogeneous, again we have 0o = 0. In the Galerkin method, 4>i must satisfy the homogeneous form of all specified boundary conditions (wq = Mxx = 0 at x = 0. L)\ fc=0, ^T=°- W For the choice of algebraic polynomials, wc assume a five-parameter polynomial because there are four conditions in Eq. (a): <f>\ (,r) = a + tot + c*2 + <t/a3 + ejr4. Using the boundary conditions, we find that a = c = 0, M, + <:/L3 + eL4 = 0, ML + I2e£2 = 0. Thus we have b = (?L3 and d = —2eL. The function <p\ is given by (taking eL4 = 1) A' / X2 X3\ = l{1-21? + 1?)' « = L{l-2T>+i?)- (b) Substituting the one-parameter Galerkin approximation W\ = ci<£i into the residual, ^ ^d4^ 24£/ Since the residual is already a constant (which implies thai the solution is exact), there is no need to integrate the weighted residual over the domain. By setting K to zero we obtain c\ = qoL4/(24EI), and the solution becomes ™-££[(iH(i)3+(i)4]- which coincides with the exact solution. Note that the solution obtained is independent of the particular weighted residual method. It can be shown that a one-parameter Ritz solution with <j>\ given by Eq, (b) also yields the same exact solution (d). Note that the solution to this problem is symmetric about x = L/2. Hence, we can use a half-beam model to solve the problem. In using a half beam, we must address the boundary conditions at x = L/2. The boundary condiliohs at this point
WEIGHTED-RESIDUAL METHODS 273 are that the slope is zero, (dwo/dx) = 0, and shear force is zero. Hence, for this case, the approximation functions 0/ in the weighted-residual method must satisfy the conditions d2<bj L clcbi d?d>; al* = O:0f=O, ~=0, and at.v = -: — = 0, ~rr=0. dxl 2 ax ax5 Obviously, the function (f>\ {x) of Eq. (b) satisfies the above conditions. Another choice of (pi is provided by . (2i - \)7VX 4>i(x) = sin . Example 7.14 Another weighted-residual method that has not been introduced formally in this study is the subdomain method. In the subdomain method, we divide the domain of the problem into as many subdomains as there are undetermined parameters, q, and then on each subdomain we require the integral of the residual IZn to be zero: / KN{x, [c], {<£}, /) dx = 0, (/ = 1, 2,..., N)t (7.177a) where Qj is the zth subdomain. The method can be viewed as a piecewise application of the weighted residual method with & = 1: *i = l' ****' (7.177b) 0, otherwise. Obviously, in this method, negative errors can cancel positive errors to give zero net error (unless 1Z^ is a positive function in £2,), although the sum of the absolute values of the errors may be very large. The finite volume method that is popular in fluid dynamics is based on ideas similar to the subdomain method. As an example, we consider the beam problem of Example 7.13 (see Fig. 7.5). Using the half-beam model with the following two-parameter approximation: Vf2\x) = c\ sin — -f C2 sin ——, (a) i-i l* we determine approximate solutions using both the collocation method and the subdomain method. Collocation Method Using collocation points at x = L/4 and x — L/2, we obtain ,-wT /n\4 . * Px\4 . 3jrl hI[C]\l) Sln4+C2(TJ ^tJ"90^0' „J /7r\4 . n /3tt\4 . 3tt 1 EI[C]\I) *m2+C2\T) ^tJ"90 = 0'
274 DIRECT VARIATIONAL METHODS which yield "®Tf 8-tf]IS)-{!}■ or Cl ~ 2?r4 £/ ' °2 162jt4 El The two-parameter collocation solution becomes H,*) = i^(195'55si"T+0114si°5r)- <b) The maximum deflection, W2(L/2) = 1.205(q0L4/EIjt4) = (q0L4/SQ.SlEI), is 5% in error compared to the exact value (qoL4/76.$EI). Subdomain Method We consider two subdomains, Q1 = (0, L/4) and Q.2 = (£/4, L/2)? in the first half of the beam. For the same choice of coordinate functions as in the collocation method, we obtain Jo [C] EI \t) sin t+C2£/ It Jsm—" *°Jdx = 0j A/4 f*7 It; s111t+^7vtJ ™—-*> dx = 0, or »(!)4i(-i)«*«'(*)4£0+£)- 4 ' 4 ' The solution of these equations is V2+lq0L4 V2-1 ?o£4 Cl = , vr ,^77"^ C"2 = 4V5tt3 e/ ' " 108V2tt3 £/ ' and the solution W2 becomes Q()L4 ( 7VX 37tx\ W2(x) = ^L 65.184 sin — + 0.414sin ). (c) 10$V2EIti3 V L L ) The center deflection obtained in the subdomain method, W2(L/2) = (<yo£4/ 73,12£7), is —5% in error compared to the exact value.
WEIGHTED-RESIDUAL METHODS 275 *-x Figure 7.9 A triangular membrane. Example 7.15 Consider the equation A{u) = V2u + Xu = 0 in ft, u = 0 on T, (a) where ft is the triangular domain shown in Fig. 7.9 and T is its boundary Equation (a) describes a nondimensional form of the equation governing the natural vibration of a triangular membrane of side a, mass density p, and tension T (X = pa2co2/T, co the frequency of vibration). We wish to determine the fundamental frequency (i.e., determine X) of vibration by using a one-parameter Galerldn approximation of the problem. The Galerkin method is based on the weighted-integral statement °=H (b) The function <pi(x, y) must vanish on the boundary I\ Thus, we have <j>i(x,0) = 0, 4>i(0, ;y) = 0, $\(x,y) ~ 0 on the line x + y - I = 0. (c) Hence the choice for 4>\ (x, y) is <t>i (*, y) = (x - 0)(y - 0)(x + y-l). Hence we have 920, , d2<Pi dx2 + dy2 = 2(x + y). (d) 0 -"II e/0 JO 1 r\-y \2(x + y) + A,ry(jt + y - 1)\xy(x + y — I) dxdy x = /q /o 3' 2(* + 3'Ky (,y + y - 1) </ay/)> = (e)
276 DIRECT VARIATIONAL METHODS Example 7.16 Consider the differential equation (Poisson's equation) u = 0 on the boundary (7.178) in a square region. The origin of the coordinate system is taken at the center of the domain, as shown in Fig. 7.10. The equation arises, among others, in connection with (1) the transverse deflection of a membrane fixed on all sides and subjected to uniform pressure fo, (2) the study of the torsion of a square cross-section prismatic bar with /o = 2G0, where G denotes the shear modulus and 0 is the angle of twist per unit length, and (3) conduction heat transfer in a square region with internal heat generation of /n unit area. The function u denotes the deflection u in the case of a membrane, the Prandtl stress function <I> in the case of the torsion problem, and the temperature T in the case of conduction heat transfer. The quadratic functional associated with Eq. (7.178) is given by '«-m[(£),+®'-*: dxdy, (7.179) which represents the total potential energy functional for the membrane problem and the total complementary energy functional for the torsion problem. For the heat conduction problem, /(«) does not have a physical meaning. The functional /(«) can be derived by the weak form procedure outlined in Section 7.4.1. The boundary condition in Eq. (7.178) is of the essential type. The natural boundary condition for the equation involves the specification of du/dn, which is not the specified boundary condition of the problem, unless a quadrant is used. For the Galerkin method, we assume (0n = 0) (7.180) H = 0 7°j \y i j i 2a '_ = 0 * 2a Figure 7.10 The square domain of Example 7.16.
WEIGHTED-RESIDUAL METHODS 277 where 4>i = (a2 - x2){a2 - y\ <p2 = (x2 + y2)<j>{..... (7.181) Clearly, <pj are twice difterentiable with respect to x and y and satisfy the boundary conditions. FromEq. (7.172b), wc have /a pa pa pa / <t>jA(<f>j)dxd)\ bj = I / fafodxdy. -a J—a J—a J —a For N = 1, we have 256 _ 16./b 45 Cl "i? and the one-parameter solution is given by For the two-parameter approximation, we have a 256 45 1024^2 525 az 1024^2 11264^4 525 " 4725 " J :;i=v f 16 9 I 45" whose solution yields ci = I295/o C2 = 525/o 4432a2 * 8864a4* The two-parameter Galerkin solution of Eq. (7.178) is given by U2(x, y) = ^[2590 + 525(i2 + f)](] - i2)(l - y2), (7.182) (7.183a) (a) (b) (7.183b) where x = x/a and y = y/a. For the Ritz method, the same approximation functions as used in the Galcrkin method must be used (why?), and we obtain /a pa I fofcdxdy. -a J—a (7 J 84) Calculations will show that a,j and b\ are the same as those in Eq. (a). Therefore, the Ritz and Galerkin solutions coincide.
278 DIRECT VARIATIONAL METHODS The exact solution to Eq. (7.178) can be obtained using the separation of variables method, and it is given by x irfe n L cosh(«7r/2) J 2a (7.185) The exact solution at the center of the region is w0(0,0) = 0.2942/ofl2. whereas the two-parameter Galerkin/Ritz solution is 0.2922 foa2, which is only 0.68% in error. Examples 7.17 and 7.18 presented below illustrate the use of variational methods for the solution of problems in two dimensions and time-dependent problems. The approximation is selected such that the parameters c/ are functions of one coordinate (say, time) and fa are functions of the remaining (say, spatial) coordinates, The procedure used in Examples 7. J 7 and 7.18 to obtain ordinary differential equations from partial differential equation(s) is termed the semidiscretization method or Kantorovich method. The Levy method of solution to be discussed in Chapter 8 is also similar to these methods, and all of them are based on the separation of variables concept. Example 7.17 Consider the following nondimensionalized partial differential equation (such as the one arising in transient heat transfer); i7-^ = 0' 0<A'<2< u(0j) = u(2,/) = 0 for/ > 0, u(xA))= 1.0 forO <x <2. Owing to the symmetry about x = 1, we solve the following equivalent problem: Bit dht n —- —=0, 0<jc < 1, dt 3x2 du w(0,f) = 0, —(Lr) = 0 forr>0, dx m(a;i0) = 1.0 for0<jt<l. ' (7.186a) (7.186b) (7.187) Let us consider the following form of a one-parameter Galerkin approximation: Ui(x, t) = a(t)4>i(a-) = c, (0(2* - x2). (7.188)
WEIGHTED-RESIDUAL METHODS 279 The function <pi(x), which is a function of x only, satisfies the boundary conditions in Eq, (7.186b). We should determine a{t) such that the initial condition (7.187) is satisfied. This requires that ci(0) = l. (7.189) Substituting Eq. (7.188) into Eq. (7.186a) and setting up the Galerkin integral, we obtain A'2)-c,(-2) (2x — x2) dx 8 dc\ 4 = i5ir + 3cl- (7J90) The solution of Eq. (7.190) is given by Cl(0 = Ae~<s/2)t% ci (0) = 1 -> A = 1, and the one-parameter Galerkin solution becomes U\(x,t) = e~2-5t(2x -x2). (7.191) The exact solution of Eqs. (7.186)-(7.187) is given by B(l>0 = 2f;£^iiEi=£, *„ = ^±^. (7.192) ,,=o A" z For a two-parameter Ritz approximation, wc seek the solution in the form ii(at, t) « U2(x, t) = ci(O0i (x) + c2(r)<feto, (7.193) where fa satisfy only the essential boundary condition, fa (0) = 0. Thus the functions 4>\(x)=x, <h(x)=x2 (7.194) are admissible. We use the weak form (for the Ritz approximation) „ fl L (Bu 32u\ , /•' / au dfadu\ ,
280 DIRECT VARIATIONAL METHODS Substituting Eqs. (7.193) and (7.194) into Eq. (7.195) and carrying out the integration, we obtain 1 dc\ 1 clc2 3Tr + 4iF+cl+C2=0> (7]96a) r£ + v£+c' + ^ = 0' (7-196b) 4 dt 5 dt 3 which must be solved subject to the initial condition in Eq. (7.187). It is clear that there is no way in which the initial condition will be satisfied exactly by the selected approximation (7.193). Therefore, we satisfy the initial condition in the Galerkin- intcgral sense: f \u(x, 0) - l]<f>j dx = 0, i = 1, 2, (7.197a) Jo which gives ^i(O) + £c2(0) = £, ±c,(0) + \c2(0) = f (7.197b) Now we use the Laplace transform method to solve Eqs. (7.196a,b) and (7.197b). Let c/ (s) denote the Laplace transform of c\ (/): d <!)e-st dt. (7.198) -OO The Laplace transform of the derivative of c/ (/) is given by Ltel=^(40-c/(O). (7.199) Using Eq. (7,199), Eqs. (7,196a,b) can be transformed to fa + l)ci + fa + \)c2 = $c\ (0) + ±c2(0), (7.200a) fa + l)ci + fa + f )c2 = £c, (0) + ±c2(0). (7.200b) The right-hand side can be substituted from Eq, (7.197b); and the resulting equations can be solved for C[ (,v) and C2(s): = ^_ K^ + ?)-|(^ + 1) _ 12.V + 240 Cl(5) " (., + 4)(.,+ ))_a, + l)2 " 3,*+104,+ 240' (7-20l3) r-r,._ s^ + D-^ + O _ iav + 120 C2W - (J,+ $)<*, + !)-(*, + !)* - -^+,04,V240- (7'201b) The inverse transform of Eqs. (7.20la,b) can be obtained by using the identity L-1 s 4- eh I a\ — a „t ct\ — b ht -^ = — e~at + — e~bt. (7.202) Xs+a)(s + b)\ h-a a-b K }
WEIGHTED-RESIDUAL METHODS 281 Hence we have a(t) = 1.640&T32'1807' + 2.3592e~2Amy (7.203a) c2(t) = -(2.265<r321807' + 1.06fe_2-486r). (7.203b) Equations (7.193) and (7.203a,b) together give the two-parameter Ritz solution. The two variational solutions in Eqs. (7.191) and (7.193) are compared, for various values of (a\ f), with the series solution (7.192) in Table 7.3. The two-parameter Ritz Tabie 7.3 Comparison of the variational solutions with the series solution of a parabolic equation with initial condition, u(x,0) = 1 [see Eqs. (7,186)-(7.187)] X 0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0 0,2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0 Series Solution Eq. (7.192) 0.4727 0.7938 0.9418 0.9880 0.9965 0.2135 0.4052 0.5560 0.6520 0.6848 0.1145 0.2177 0.2996 0.3522 0.3703 0.0617 0.1174 0.1616 0.1900 0.1997 0.0333 0.0633 0.0872 0.1025 0.1077 0.0097 0.0184 0.0254 0.0298 0.0313 Variational Solution Eq. (7.191) 0.3177 0.5648 0.7413 0.8472 0.8825 0.1927 0.3426 0.4496 0.5139 0.5353 0.1031 0.1834 0.2407 0.2751 0.2865 0.0552 0.0981 0.1288 0.1472 0.1534 0.0296 0.0525 0.0690 0.0788 0.0821 0.0085 0.0151 0.0198 0.0226 0.0235 Eq. (7.193) 0.4265 0.7413 0.9443 1.0357 1.0154 0.2306 0.4152 0.5538 0.6466 0.6934 0.1238 0.2230 0.2975 0.3473 0.3725 0.0665 0.1198 0.1600 0.1866 0.2001 0.0357 0.0643 0.0858 0.1002 0.1075 0.0103 0.0186 0.0248 0.0289 0.0310
282 DIRECT VARIATIONAL METHODS solution is more accurate than the one-parameter Galerkin solution and agrees more closely with the series solution for large values of time. Example 7.18 Consider the membrane/torsion problem considered in Example 7.16. Here we seek a one-parameter approximation of the form Vi (x, y) = ci (x)0i 00 = a(x)(y2 - a2). (7.204a) Since u = 0 on jc = ±a and on y = ±a, it follows that we must determine c\(x) such that it satisfies the conditions ci(-a) = ci(a)=0. (7.204b) Substituting Eq. (7.204a) into the Galerkin integral (w & U\), °-££K^+tfH**<* <"»* wc obtain ~~ff/fl[(>2-fl2)^+2ci+^](y2^fl2)^ldx- (7,2o5b) Performing the integration with respect to y and dividing thi'oughout by the coefficient of d2c\/dx2> we obtain -£(£-5*"-3)*- An examination of the integrand in (7.206a) shows that C{ (x) is noUa. periodic function. Hence, the integral vanishes only if the integrand is identically zero: £-2P«-8-*
WEIGHTED-RESIDUAL METHODS 283 This completes the application of the Galerkin method. We can solve the ordinary differential equation (7.206) either exactly or by an approximate method, such as the Ritz or Galcrkin method. We consider the exact solution of Eq. (7.206) subjected to the boundary conditions in Eq. (7.204b). The exact solution of Eq. (7.206) is given by a (x) = A cosh kx -h B sinh lex — —, 5 2a2' (7.207) Using the conditions in Eq. (7.204b), the constants of integration, A and B, arc evaluated to be A = /o 2 cosh ka The solution in Eq. (7.204a) becomes .2 / tl2v B=0. «*<••'>-^-K) [-££]■ A two-parameter approximation of the form U2(x, y) = (v2 - a2)[Cl(x) + c2(x)y2] (7.209) gives the differential equations « ,d2c\ * <*2 15*' dx2 8 105 8 4d2c2 /4 4 2 \ 2^ + W5a^-{3C^T5aC2) = 3f0 4d2a 8 6d2c2 (4a2 44 4 \ 2 2 5 <£x2 315 g?jc2 \ 15 105 / 15 (7.210) with the boundary conditions c\{-a) = C] (a) = c2(-a) = c2(a) = 0. (7.211) The simultaneous differential equations in Eq. (7.210) can be solved as follows: let D = d/dx, D2 = d2/dx2, and so on. Then we can write Eq. (7.210) in the operator form -D2-4 15 " 3 8"! n2 _ 4^ 8^ n2 _ 44flf -105 ^ 15 315^ 105. 8fl4 n2 4«2 105^ " 15 C2l = 15/0|a2 c\ (7.212) Using Cramer's rule, but keeping in mind that D's operate on the quantities in front of them, wc obtain L(Cl) = 2 4o2 ±f0a2 m&- I/O Sal 105 J 315 J 44a4 105
284 DIRECT VARIATIONAL METHODS / 8 44 A 2 , / 8 4 , 4a2\ 2 , , = (315° * - 105° ) 3/0 - (iSfl °" - 1?) 15** ' 384 T575 ybfl4, £(C2) = 8a2 r>2 4 15 U 3 8 /t4n2 4« i/o 105 " /y 15 15^°" = 0, (7.213a) (7.213b) where £(■) is the determinant of the operator matrix in Eq. (7.212), 256«8 L = 33075 The general solutions of Eqs. (7.213a,b) are ("->+3y (7.214) ci(x) = —— + A\ cosh/ci* + A2 sinhfci* 4- A3 cosh fo* + A4 sinhk2X, C2(x) — B\ cosh/qx + ^2sinhfcijr + Z?3COsh&2-x +fi4sinh/c2^, (7.215) where A,- and Bj(i — 1, 2, 3, 4) are constants to be determined, and k\ and k\ are the roots of the quadratic equation 28, 63 n 1* Substituting Eq. (7.215) into the first equation in (7.213a), we obtain the relationships k2 =14 -VI33, *? = 14 + Vl33. (7.216) (7.217) Use of the boundary conditions in Eq. (7.211) gives A2 = A4 — 52 = #4 = 0, and Ai cosh/qc/ + A3 cosh/^tf = 0.5 ft, (7.218) #1 cosh/qtf 4- £3 coshfotf = 0, which can be solved along with Eq. (7.217) for A\, B\, A3, and #3, and we have ci (*)= 0.5/o+0.516/0 cosh/ci.v c2U) =-0.1138/0 cosh/q a cosh /ciA' ■0.0156/0 cosh k2X cosh/qtf + 0.11387b cosh /c2« * cosh £2* cosh /c2fl (7.219)
WEIGHTED-RESIDUAL METHODS 285 Table 7.4 Comparison of the solution u(x,0) of Eq. (7.178) obtained by the Galerkin/Ritz and semidiscretization methods with the series solution21 Ritz-Galerkin Solution** Semidiscretization Eqs.(7.183a,b) Eg. (7.209) x/a (y = 0) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.0 Series Solution Eq. (7.185) 0.29445 0.29194 0.28437 0.27158 0.25328 0.22909 0.19854 0.16104 0.11591 0.06239 -0.00038 0.67500 One Parameter 0.31250 0.30937 0.30000 0.28437 0.26250 0.23437 0.20000 0.15937 0.11250 0.05937 0.00000 Shear stress, 0.62500 Two Parameter 0.29219 0.28986 0.28278 0.27075 0.25340 0.23025 0.20065 0.16382 0.11884 0.06463 0.00000 <ryz{at0) 0.70284 One Parameter 0.30261 0.30014 0.29266 0.27999 0.26180 0.23765 0.20693 0.16886 0.12250 0.06667 -0.00447 0.72636 Two Parameter 0.29473 0.29222 0.28462 0.27177 0.25339 0.22909 0.19839 0.16072 0.11546 0.06203 0.00000 0.66416 a« = u/tfotf2; <ryz = ayz/fw- ^The Ritz and Galerkin solutions are the same for the same coordinate functions. Equations (7.219) and (7.208a) together define the two-parameter solution of the torsion problem. A comparison of the solutions obtained using the Galerkin method and the semi- discretization method (with one and two parameters) with the series solution of the torsion problem is presented in Table 7.4. The table also includes the shearing stress, ayz — —(du/dx), at x = a and y = 0. It is clear from the results that the two- parameter solutions arc more accurate than the one-parameter solutions. Example 7,19 (The Trefftz method) In all the variational methods discussed in this chapter, the coordinate functions were selected such that they satisfied the boundary conditions of the problem, and the unknown parameters were determined using a variational procedure, such as the minimization of a quadratic functional or setting the weighted residual to zero. An alternative approach that is complementary to the above approach is to select the approximation functions to satisfy the governing differential equation and determine the unknown parameters such that the boundary conditions are satisfied in a variational/intcgral sense. This method is known as the Trefftz method, and its application is illustrated via the torsion problem of Example 7.16. The torsion of cylindrical members can be formulated alternatively in terms of the conjugate function ^, which is related to the Prandtl stress function <S>: * = * - —■ (x2 + y2) in Q, (7.220)
286 DIRECT VARIATIONAL METHODS where /o = 2G0, as in Example 7.16. Consequently, ^ is governed by the Laplace equation VlV = 0 in Q = {(Jt% y): - fl < a-, }> < a}, subjected to the boundary condition (7.221) y = £l(x2 + y2) on T. (7.222) Note that a nonhomogeneous equation (i.e., Poisson's equation) with homogeneous boundary conditions is transformed to a homogeneous equation (i.e., Laplace's equation) with nonhomogeneous boundary conditions. For the square domain of Example 7.16, we wish to determine an approximate solution of Eqs. (7.221) and (7.222) using Hie Trefftz method. We select an Af-parameter approximation of the form (7.223) where </>/ are selected such that ^n satisfies the governing equation V2^ = 0 (i.e., 4>j should be harmonic functions). The parameters cj are determined by the requirement that the boundary condition (7.222) is satisfied in an integral sense: °=/r[*»- -i -j(x +y ) /o, dn ds E^-f^+y2) dn ds ^^ijcj-bi, (7.224) where tiJ ^ £d-tVJ Ui' U,=%4- --dn As a specific example, we choose a one-parameter approximation with '-4>j ds, b> = <p t (x2 + y2) ir>ds- (7.225) <j>x = x4 - 6x2y2 + /, (7.226)
(7.227) WEIGHTED-RESIDUAL METHODS 287 which satisfies the equation V20j = 0. We have r,, = 4 f (4a3 - I2ax2)(a4 - 6a2x2 + x4) dx = ^a\ J—a 5 J h=-f fo(-12x2a + 4a3)(x2 + a2)dx = ~a6f0, J—a 3U and the solution is given by *'= "life <*4 - 6x2y2+A (?-228> The Prandtl stress function <f> becomes * = * - ^(x2 + y2) = —~£{*A - 6*V + /) - ^(x2 + y2). (7.229) The maximum shear stress ayz is a^(a, 0) = - ( —-i ) = 7/0fl = 0.833/0«. V dx /(flt0) 6 The maximum shear stress is about 23.4% greater than the exact solution. The Trefftz method has limited use, because it can only be utilized for Dirichlet- type boundary-value problems (i.e., boundary-value problems in which the function is specified on the boundary). Furthermore, it is not easy to find approximation functions that satisfy the governing differential equations. Some of the hybrid finite element models are based on ideas similar to the Trefftz method. Example 7.20 Consider the nonlinear differential equation d / du \ dx \ dx) subject to the boundary conditions / du\ hi— )+!=(), 0 < x < 1, (7.230a) ■*>-* (£) = 0. (7.230b) v=0 We wish to determine a one-parameter Ritz solution to the problem. The weak form of Eq. (7.230a) is given by Jo L dx dx + w • 1 dx. The boundary term is zero because w(l) = 0 and (du/dx)(0) = 0. Let u m JJ\ — c\(j>\ + 0o, with 0o = V2, 0i = 1
288 DIRECT VARIATIONAL METHODS Substituting into the weak form, wc obtain 0=1 {[ci(l -x) + V2] (-l)(-c,) + (1 -x)J dx = c\ [ci • £ 4- V2] + 5 or c\ + 2V2c-! +1=0, (ci)i.2 = -V5± VT=T = -V2± 1. Thus, there are two approximate solutions to the nonlinear problem. Wc must choose one value of q using some meaningful criterion. We shall take the value of C[ that yields the smallest integral of the residual in the differential equation: Clearly, the smaller (in absolute value) root gives the smaller value of the residual. Hence, we choose (c\)[ = — \fl + 1. The solution becomes V\ = (1 - V2)(l - x) + V2 = 1 + (a/2 - 1)jc. The exact solution is w(jc) = \/l -f x2. At ;t = 0 the approximate solution matches with the exact. 7.6 SUMMARY In this chapter, the Ritz and weighted-residual (e.g., Galerkin, least-squares, and collocation) methods were presented and their application to simple problems was illustrated. The Ritz method makes use of the weak form provided by the principle of virtual displacements, the principles of minimum total potential energy, or the one developed from the governing equations of the problem as discussed in Sections 7.4.J and 7.5.1. The key feature of the weak form is that it includes the governing equation(s) as well as the natural boundary condition(s) of the problem. Hence, the Ritz approximation is not required to satisfy the natural boundary conditions. All weighted-residual methods, except the least-squares method, are based on an weighted-integral statement of the governing equation(s), whereas the least-squares method is based on the minimization of the square of the governing equation(s). Thus, the integral statements used in all weighted-residual methods do not include any boundary conditions as a part of the statements. Hence, the approximations chosen for the weighted-residual methods are required to satisfy all (both natural and essential) specified boundary conditions of the problem. Consequently, the approximation functions are of higher order than those used in the Ritz method. The Ritz as well as the weighted-residual methods may be used for linear and nonlinear differential equations. However, in the case of the least-squares method, there arc some limitations, as
EXERCISES 289 discussed in Section 7.5.5. Although the least-squares method can be interpreted as a special case of the weighted-residual methods for linear static problems, it is based on an integral statement whose Euler equations, if derived, are not the same as the governing equations. Thus, the least-squares method is quite different from the other weighted-residual methods. Overall, the Ritz method is the most efficient method, especially for solid and structural mechanics problems. The single most difficult step in using the variational methods presented in this chapter is the selection of the approximation functions. The requirements (7.62) and (7,164) on the approximation functions merely provide the guidelines for their selection. The selection of the approximation functions becomes even more difficult for problems with irregular domains (i.e., noncircular and nonrectangular) or discontinuous data (loading as well as geometry). Further, the generation of coefficient matrices for the resulting algebraic equations cannot be automated for a class of problems that differ from each other only in the geometry of the domain, boundary conditions, or loading. These limitations of the classical variational methods can be overcome by representing a given domain as a collection of geometrically simple subdomains for which we can systematically generate the approximation functions (see Example 7,8). One such technique, namely, the finite element method, is discussed in Chapter 9. The finite element method is based on ideas similar to the classical variational methods, especially in developing the system of algebraic equations for the unknown coefficients, but the method views a given domain as a collection of conveniently chosen subdomains that allow a systematic generation of the approximation functions. EXERCISES 7.1 Let V be the vector space of all polynomials with real coefficients. Determine which of the following subsets of V are subspaces: (a) Sx=lp(x): p(\) = 0). (b) S2 = {p(x): degree of p(x) = 3}. (c) 53 = {p(x): degree of p(x) < 3}. (d) 54 = {p(x): constant term is zero}. 7.2 Determine which of the following sets of vectors in St3 are linearly independent over the real number field 3"i: (a) {(-U,0),(-l,l,l),(-2,-l,l), (1,1,1)}. (b) {(1,0,0), (1,1,0), (1,1,1)}. (c) {(1, I, 1),(I,2,3), (2,-1,1)}. 7.3 Determine if the following sets of functions are linearly independent. (a) {sin(77jrx/L)}fI=1, 0 <x <L. (b) {x"(l-x)}l={), 0<x<). (c) {1 + x + x2, 1 + 2x + 3jc2, 2 - 3jc + jc3}, 0 < x < 1.
DIRECT VARIATIONAL METHODS 7.4 Compute the L2 norm and the sup-norm of the following functions in the interval indicated: (a) u(x) = sin nx — x, on 0 < x < 1. (b) u{x) ~xl/3, onO <x < 1. (c) u(x) = cos nx + 2x — 1, on 0 < x < 1. (d) «(*) = Vl -hx2, on 0 < A' < 1. , x , x f 100 sin 100* a;, 0<x < 0.01 (e) h(a) = < (0, 0.01 <x < 1. 7.5 Prove the following relations in a real inner product space: (a) Parallelogram law: \\u + v\\2 + \\u - v\\2 = 2(\\u\\2 + \\vf). (b) (utv) = $[\\u + v\\2-\\u-v\\2]. (c) IN|-NII<II«-*I|. 7.6 Let u = u(a) and v = \(x) be two vector functions of a:. Which one of the following products qualifies as an inner product? (a) (u,v) =/0Lu-vJa'. (b) (u,v)=ft'u(L-x)-v(x)dx. 7.7 Compute the inner product of the following pairs of functions on the interval indicated. Use the Z,2-inner product and the Z/1-inner product. (a) 11 = A' — x2, v = sin nx, on 0 < x < 1. (b) u = (1 + jc), v = 3x2 - 1, on -1 < jc < 1. (c) u = sin7TA, v = cos7TA% on 0 < x < 1. (d) w = sin nx, v ~ a + bx -\- ex2, on 0 < x < 1. (e) w = sin jta- sin 717, v = (1 — x2 — )>2), on 0 < jc, )> < 1. (f) it = (jc2 - a2)(y2 - b2), 0<x< a and 0<y<b. 7.8 Find the distance between the following pair of functions using the inner product in Eq. (7.23a): it = x3 - 3x + 2 and v = (x - l)2, 0 < x < 1. 7.9 Check whether the following pair of functions are orthogonal in the Z,2(0,1)- space: u(x) = 2 + 3x2 — x and v(x) = 3 + 3a' — 5a2. 7.10 Check whether the pair of functions in Exercise 7.9 arc orthogonal in the //■(O, l)-space. 7.11 Determine the constants a and b such that w(x) = a+bx +(3a2 is orthogonal in Z,2(0,1) to both u(x) and v(x) of Exercise 7.9. 7.12 Determine C if the indicated pairs of vectors are orthogonal in d\4: (a) (1,C-1,2, l + C),(2C,4,C,i). (b) (2, 1 - C, 4, 3C), (C, 1 + C, 3C> -7).
EXERCISES 291 7.13 Determine a vector in the space V of polynomials of degree 2 such that the vector is orthogonal to the polynomials pi = 1 + x — 2x2 and p2 = — 2+ 4jc+*2inL2(l,0). 7.14 Which of the following operators qualify as linear operators? <*>*»--£(■£)■ (b) r(w) = V2w + 1. (c) I(u) = f£K(x)^dx-u(0). 7.15 If 71) and 72 are linear operators defined on SR3 by T] (x) — (0, *i — X3, A'2 + X3), 72(x) = (x2 — X3, 2xi — X2, *i), determine: (a) 7i + T2. (b) TiT2. (c) T27i. 7.16 Find the matrix representing the linear transformation from dl4 into ffl3 with respect to the standard bases of $\4 and $H3: T(X\, X2, X3, X4) = (X\ + 2X2, 2X2 + X3, A'3 + X]). 7.17 Identify the linear and bilinear functionals: (a) I(u9 v) = fQ Vw • Vvdxdy. fa / tf2« d2v \ 7 J 8 Determine which of the following operators represent bilinear forms: (a) B: W2 x m2 -> % B(xy y) = (Xi + y\)2 - (xx - )>i)2. (b) B: m2 x m2 -> S», B(x, y) = *D'2 - x2y\. 7.19 Consider the functional B(u,v) = k\\u + v\\2-\\u-v\\2) on an inner product space. Show that: (a) #(*> •) is bilinear in u and u. (b) For fixed u, fu(v) = B(uy v) is a (bounded) linear functional. (c) B (h, v) satisfies the axioms of an inner product. 7.20 If / (it) is a quadratic functional, show that (a) its first variation is a bilinear functional of u and Sw, and (b) its first variation is a linear functional of 8u. Give admissible approximation functions, either algebraic or Uigonometric, for a two-parameter Ritz approximation of problems in Exercises 7.21-7.27. Assume
292 DIRECT VARIATIONAL METHODS that the total potential energy principle or weak form is used to construct the Ritz approximation. 7.21 A cable suspended between points A : (0, 0) and B : (L, h) and subjected to uniformly distributed transverse load of intensity fa. 7.22 A cantilever beam subjected to uniformly distributed load of intensity qo. 7.23 The symmetric half of the simply supported beam problem considered in Example 7.7 (0 < x < L/2). 7.24 A beam clamped at the left end and simply supported at the right end, and subjected to point load Fo at x = L/2. 7.25 A simply supported beam with a spring support at x = L/2, and subjected to uniformly distributed load of intensity qo. 7.26 A square elastic membrane fixed on all its sides and subjected to a uniformly distributed load of intensity /o. 7.27 A quadrant model (because of the biaxial symmetry) of the membrane problem of Exercise 7.26. 7.28-7.34 Find the two-parameter Ritz approximation using algebraic polynomials of the problems in Exercises 7.21-7.27 and compare the solutions with the exact solutions when possible. 7.35 Find the first two natural frequencies of a cantilever beam. Take EI = (a -f bx)~x where a and b are constants. 7.36 Find a two-parameter Ritz approximation of the transverse deflection of a simply supported beam on an elastic foundation that is subjected to uniformly distributed load. Use (a) algebraic polynomials and (b) trigonometric polynomials. 7.37 Derive the matrix equations corresponding to the ^-parameter Ritz approximation WN = cxx2 + c2x3 + • ■ ■ + cNxN+] of a cantilever beam with a uniformly distributed load, qo- Compute a,j and bj in explicit form in terms of i, j. L, EI, and </o- 7.38 Use a two-parameter Ritz approximation with trigonometric functions to determine the critical buckling load P of a simply supported beam. 7.39 Consider the budding of a uniform beam according to the Timoshenko beam theory. The total potential energy functional for the problem can be written as "<-*- uLH$)'+'(%+*)'-<(%)' dx, where wq(x) is the transverse deflection, <px is the rotation, D is the flexural stiffness, S is the shear stiffness, and N is the axial compressive load. We wish to determine the critical buckling load Ncr of a simply supported beam using
EXERCISES 293 7.40 7.41 the Rilz method. Assume a one-parameter approximation of wo and </>x and determine the critical budding load. Determine the N-parameter Ritz solution for the transient response of a simply supported beam under step loading q(xj) — cjqH (t — fo) > where H(() denotes the Hcaviside step function. Use trigonometric functions for 0,- (jc). Show that the two-parameter Ritz solution for the transient response of the bar considered in Example 7.6 yields the equations B Ph aizi-ii (a) Use the Laplace transform method to determine the solution of these equations. 7.42 Derive the weak forms of the following nonlinear equations: 4 (£»-)- dNx, dx d2Mxx dx2 /<*). = q(x), 0 < x < L, 0 < x < L, (a) (b) where Nxx = EA tlllQ 1 lx~+2 m' Mxx = -El crwp lb?* 7.43 7.44 7.45 Consider a uniform beam fixed at one end and supported by an elastic spring (spring constant k) in the vertical direction. Assume that the beam is loaded by a uniformly distributed load #o- Determine the one-parameter Ritz solution using algebraic functions. Consider the problem of finding the fundamental frequency of a circular membrane of radius a, fixed at its edge. The governing equation for axisymmelric vibration is \_d_ / du\ r dr \ drJ Xu = 0, ()</•< a, where X is the frequency parameter and a is the deflection of the membrane, (a) Construct the weak form, (b) use one-parameter Ritz approximation to determine X, and (c) use two-parameter Ritz approximation to determine A. Select trigonometric functions for <f>i(r). Consider the problem of finding the solution of the equation dlu ~d? + u+ x = 0, 0 < x < 1; w(0) = i/(l) = 0.
294 DIRECT VARIATIONAL METHODS (a) Develop the weak form, (b) assume TV-parameter Ritz approximation of the form UN(x) =x(\ - x)(c[ + c2x H h cNxN~[) and obtain the Ritz equations for the unknown coefficients, and (c) determine the two-parameter solution and compare it with the exact solution sin jc sinl 7.46 Consider the Bessel equation 2d2u . dti —^ +x— + (* - 1)« = 0, 1 < x < 2 dxz ax subject to the boundary conditions h(1) = 1 andw(2) = 2. Assume u — w+x and reduce the equation to dx \ dw \ x2 - 1 2 JC I W = AT, 1 < JC < 2, djt / x subject to the boundary conditions w(l) = w;(2) = 0. Determine a one- parameter Ritz approximation of the problem and compare it with the analytical solution u(x) = 3.6Q12J[(x) +0.75i95Yl(x)i where 7i and Y[ are Bessel functions of the first and second kind, respectively. 7.47 Derive the weak form and obtain a one-parameter Ritz solution of the problem /32u d2u\ i . — —t- -{ r- =1 in a unit square, \3x2 dy2J w(i,y) = «U, l) = o, 3jc 3v The origin of the coordinate system is taken at the lower left corner of the unit square. 7.48 Determine the two-parameter Petrov-Galerkin solution (with \j/\ = 1 and \fr2 = x) of the following differential equation and compare, if possible, with the exact solution: ~ = ——, 0 < x < J; M(0) - u'(l) = 0. dx2- I + x
EXERCISES 295 7.49 Determine the two-parameter Petrov-Galerkin solution (with ^ = 1 and \fs2 — x) of the following differential equation: d r o du~\ , (1-hr2) — +u=sin7r*+3*-l, 0<;t<l; m(0)=h'(1)=0. ax L a* J 7.50 Determine the one-parameter Galerkin solution of the equation x\ d2WQm d2 [/ x\ 5? (2 + l) dx2 x + kwo = ^oy- that governs a cantilever beam on an elastic foundation and subjected to lineai'ly varying load (from zero at the free end to go at the fixed end). Take k = L = 1 and #0 = 3, and use algebraic polynomials. 7.51 Determine a two-parameter Petrov-Galerkin solution (with Vi — i and y/2 = xy) of the problem . -V2« =0 inO < (x, y) < 1, u = sin nx on y = 0, u = 0 on all other sides. 7.52 Solve the nonlinear differential equation / du\ w(l) = V2, n'(0)=0. f t „ _ + 1 = 0, 0 < x < 1, ax Use a one-parameter Petrov-Galerkin approximation with (a) 1//1 ~ 1 and (b) i/f\ = x. Compare the results with the exact solution, u(x) = Vl -\-x2. See Example 7.20. 7.53 Find the first two eigenvalues associated with the differential equation d2u — —^ = Xlt, 0 < X < 1, w(0) =0, w(l) + «7(l) = 0. Use the least-squares method. Use the operator definition A = —(d2/dx2) to avoid increasing the degree of the characteristic polynomial for A. 7.54 Solve the Poisson equation —V2w = /o in a unit square, u =0 on the boundary.
296 DIRECT VARIATIONAL METHODS using the following N-parameter Galerkin approximation: N Ut\ = Y^ cij sin inx sin jny. 7.55 Solve the nonlinear equation in Exercise 7.52 by the Galerkin method. 7.56 Find a two-parameter Galcrkin solution of a clamped (at both ends) beam under uniformly distributed load. 7.57 Solve the equation in Exercise 7.49 using the least-squares method. 7.58 Solve the problem of Exercise 7.52 using the least-squares method. 7.59 Solve the equation in Exercise 7.53 using the least-squares method. 7.60 Solve the equation in Exercise 7.54 using the least-squares method. 7.61 Consider a cantilever beam of variable flcxural rigidity, EI = ao\2 — (x/L)2], and carrying a distributed load, q = q$[\ — (x/L)]. Find a three-parameter solution using the collocation method. 7.62 Repeat Exercise 7.52 using the collocation method. 7.63 Solve the differential equation in Exercise 7.53 by the collocation method. 7.64 Solve the problem in Exercise 7.54 using the one-point collocation method. 7.65 Consider the Laplace equation -V2u =0, 0 < x < 1, 0 < y < oo, u(0,y) = u(L,y) = 0 for y > 0, u(x, 0) = x(\ - jc), u(x, co) = 0, 0 < x < 1. Assuming an approximation of the form u(x,y) = ci(y)x(l -x), find the differential equation for c\ (y) and solve it exactly. 7.66 Use the semidiscretization method to find a two-parameter approximation of the form nx / x 3ttx u(x, y) = c\ (y) cos — + c2(y) cos —-, 2a 2a i to determine an approximate solution of the torsion problem in Example 7.16. 7.67 Use the semidiscretization method to find a two-parameter approximation of the form U)(xt t) = 6*1 (/)(! — C0S27TA-),
REFERENCES 297 to determine an approximate solution of the equation 34w d2w a*+W = 0' 0<x<1' t>0' w = — = 0 at x = 0, 1 and / > 0, dx w = sinnx — nx(\ — x), — = 0 at t = 0, 0 < x < 1. Use the Galerkin method to satisfy the initial conditions. 7.68 Consider the problem of finding the eigenvalues associated with the equation V2u + Xu = 0 in Q; u = 0 on I\ where Q is a rectangle, —a < x < a and —b < y < b, and V is its boundary (see Example 7.18). Assuming approximation of the form V\=ci{x)4>\{y) = cx{x){y2-b2\ determine the first eigenvalue X\. 7.69 Repeat Exercise 7.68 with a two-parameter approximation of the form U2 = ci{x)4>\ (y) + c2(x)<p2(y) = \c{(x) + c2(x)y2\(y2 - b2). 7.70 Solve the problem in Example 7.19 using a two-parameter approximation of the form (which is more complete than the one-parameter approximation used in Example 7.19, although it has no effect on the derivative of the solution) *2= c\ +C2(x4 -6x2y2 + y4). REFERENCES 1. Odcn, J. T., and Redely, J. NM Variational Methods in Theoretical Mechanics, 2nd edition, Springer-Verlag, Berlin (1982). 2. Oden, J. T., and Ripperger, R. A., Mechanics of Elastic Structures, 2nd edition, Hemisphere, New York (1981). 3. Reddy, J. N., Energy and Variational Methods in Applied Mechanics, John Wiley, New York (1984). 4. Reddy, J. N., Applied Functional Analysis and Variational Methods in Engineering, McGraw-Hill, New York, 1986; reprinted by Krieger, Melbourne, FL (1992). 5. Reddy, J. N., An Introduction to the Finite Element Method, 2nd edition, McGraw-Hill, New York (1993). 6. Reddy, J. N., Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia (1999). 7. Reddy, J. N., and Rasmussen, M. L., Advanced Engineering Analysis, John Wiley, New York (1982); reprinted by Krieger, Melbourne, FL (1991).
298 DIRECT VARIATIONAL METHODS 8. Washizu, K., Variational Methods in Elasticity and Plasticity, 3rd edition, Pergamon Press, New York (1982). 9. Timoshenko, S. P., and Woinowsky-Krieger, S., Theory of Plates and Shells, McGraw-Hill, Singapore (1970). 10. Lanczos, C, The Variational Principles of Mechanics, University of Toronto Press, Toronto (1964). 11. Langhaar, H. L., Energy Methods in Applied Mechanics, John Wiley, New York (1962). 12. Mikhlin, S. G, Variational Methods in Mathematical Physics, translated from the 1957 Russian edition by T. Boddington, Macmillan, New York (1964). 13. Mikhlin, S. G, The Problem of the Minimum of a Quadratic Functional, translated from the 1952 Russian edition by A. Feinstein, Holdcn-Day, San Francisco (1965). 14. Mikhlin, S. G, An Advanced Course of Mathematical Physics, Elsevier, New York (1970). 15. Rektorys, K., Variational Methods in Mathematics, Science and Engineering, 2nd edition, Reidel, Boston (1980). 16. Shaw, F. S,, Virtual Displacements and Analysis of Structures, Prentice-Hal J, Englewood Cliffs, NJ (1972). 17. Dym, C. L., and Shames, I. H., Solid Mechanics: A Variational Approach, McGraw-Hill, New York (1973). 18. Charlton, T M(> Energy Principles in Theory of Structures, Oxford University Press, Oxford (1973). 19. Richards, T. H., Energy Methods in Stress Analysis, Ellis Horwood (distributed by John Wiley & Sons), Chichester, UK (1977). 20. Davics, G A. O., Virtual Work in Structural Analysis, John Wiley, Chichester, UK (1982). 21. McGuire, W., Gallagher, R. R, and Ziemian, R. D„ Matrix Structural Analysis, 2nd edition, John Wiley, New York (2000). 22. Kantorovich, L, V., "Some remarks on Ritz's method" (in Russian), Tr. Vyssh. Voen, Morst lnzh. Strait Uchil., No. 3 (1941). 23. Kantorovitch, L. V., and Krylo v, V. I., Approximate Methods of Higher Analysis, translated by C. D. Benster, Noordhoff, Groningen (1958). 24. Galerkin, B. G, "Series-solutions of some cases of equilibrium of elastic beams and plates" (in Russian), Vest. Inzh. Teck., 1, 897-903 (1915). 25. Galerkin, B. G, "Bcrechung der frei gelagerten elliplschenPlatteauf Biegung," Z Agnew. Math, Mech., 3(2), 113-117 (1923). 26. Rayleigh, Lord (John William Strutt), "Some general theorems relating to vibrations,** Proc. Lond, Math. Soc, 4, 357-368 (1873). 27. Rayleigh, Lord (John William Strutt), Theory of Sound, 1st edition (1877); 2nd revised edition, Dover, New York (1945). 28. Ritz, W, "Uber eine ncue Methode zur Losung gewisser Variationsprobleme der mathernatischen Physik," Journal fur reine unci angewandte Mathematik, 135, 1-61 (1908). 29. Trefftz, E., "Zur Theoric der Slabilitat des elastischen Gleichgewichts," Z Angew. Math. Mech., 13, 160-165(1923). 30. Trefftz, E., "Ein Gegenstuck zum Ritzschen Verfahrcn," Proc. 2nd Int. Congr. Appl Mech., E. Meissner (cd.), Zurich, pp. 131-137 (1926).
8 THEORY AND ANALYSIS OF PLATES 8,1 INTRODUCTION 8.1.1 General Comments In this chapter we study bending, budding, and natural vibration of plate structures. A plate is a flat structural element with planform dimensions much larger than its thickness and is subjected to loads that cause bending deformation in addition to stretching (see Fig, 8.1). Street manhole covers, table tops, side panels and roofs of buildings and transportation vehicles, glass window panels, turbine disks, bulkheads, and tank bottoms provide familiar examples of plate structures. A shell is initially a curved structural element with thickness much smaller than the other dimensions. Like a plate, a shell is subjected to loads that cause stretching and bending deformations. Examples of shell structures are provided by pressure vessels, pipes, curved panels of a variety of structures including automobiles and aerospace vehicles, tires, and roof domes and sheds. In most cases, the thickness of plate and shell structures is about one-tenth or less of the smallest in-plane (i.e., within the plane) dimension. When the thickness is one-twentieth of an in-plane dimension or less, they are termed thin; otherwise they are said to be thick. Because of the smallness of the thickness dimension, it is often not necessary to model plate and shell structures using 3D elasticity equations. Simple 2D theories that arc based on some kinematic assumptions can be developed to study deformation, stresses, natural frequencies, and global buckling loads of plate and shell structures. In the present study, the governing equations of the 2D theories are derived using the principle of virtual displacements for circular plates in cylindrical coordinates and for noncircular plates in rectangular Cartesian coordinates. Bending, buckling, and vibration solutions are obtained for particular cases of 299
300 THEORY AND ANALYSIS OF PLATES Figure 8.1 Geometry of a plate with curved boundary. geometry, boundary conditions, and loads using the energy and variational methods developed in the earlier chapters. The two-dimensional theories (in terms of the planform coordinates x and y) of plates and shells arc developed from an assumed displacement or stress field. They are often taken as a finite linear combination of unknown functions and the thickness coordinate z. For example, the zth displacement component u\ is expanded as N 7=0 where (x, y) are the coordinates in the plane of the plate, z the thickness coordinate, t the time, and </y; are functions that describe the deformation and are to be determined. The equations governing <p, (j = 0, 1, 2,..., AT) are determined by the principle of virtual displacements: 0= / (8U + 8V-8K)<H, (8.2) Jo where 8U, 8V, and 8K denote the virtual strain energy, virtual potential energy due to external applied forces, and virtual kinetic energy, respectively. These quantities are determined in terms of actual stresses and virtual strains, which depend on the assumed displacement expansion (8.1). The integration over the domain of the plate structure is represented as the tensor product of the integration over the midplane (mid surface) of the plate (shell) and integration over the thickness, This is possibly due to the explicit nature of the assumed displacement field in the thickness coordinate. Thus, we can write r rh/2 p \ 0)tfV= / / Orftafe, (8.3) J Vol. J-h/2 Jn0 where h denotes the total thickness of the plate and Qq denotes the undeformed midplane of the plate, which is assumed to coincide with the jty-plane. Since all
INTRODUCTION 301 undetermined variables (p) are explicit functions of the thickness coordinate z, the integration over plate thickness is carried out explicitly, reducing the problem to a two-dimensional one. Consequently, the Euier-Lagrangc equations resulting from Eq. (8,2) consist of differential equations involving the dependent variables (pi (x, v, t) and thickness-averaged stress resultants, Rj}- < , fh/2 *S'7)= / (z)m<xudz. (84) J J-h/2 The stress resultants R;" can be written in terms of #y with the hel p of the assumed constitutive equations (stress-strain relations) and strain-displacement relations. More complete development of this procedure is forthcoming in this chapter. The same approach is used when n.( denote stress components, except that the basis of the derivation of the governing equations is the principle of virtual forces. In the present book, the stress-based theories will receive very little attention. 8.1.2 An Overview of Plate/Shell Theories There exist a number of plate and shell theories, and they differ from each other in two principal ways: (1) the choice of the assumed field (i.e., displacement or stress), and (2) the choice of terms in the expansion [i.e., number of terms and powers of the thickness coordinate z retained in Eq. (8.1)]. The choice of the field is often restricted to either displacements or stresses, although a mixed approach is possible. The number of terms retained in the displacement or stress expansions is often limited, at most, to cubic in thiclcness coordinate in the interest of keeping the theory simple. Thus, a plate or shell theory can be developed for any combination of the field variables and number of terms in the expansion of the variables. The number of theories further multiply if different order expansions are used for different components of a field variable and types of nonlinear terms included in the strain-displacement relations. A general but brief review of various theories of plates is presented before closing this section. The ideas arc also applicable to shells. No review of plate/shell theories will be complete, as there are thousands of papers dealing with one aspect or the other of the many combinations mentioned above. Kirchhoff Plate Theory The simplest plate theory is known as the Kirchhoffplate theory or simply the classical plate theory (CPT), which is an extension of the Euler- Bernoulli beam theory (see Example 4.3) to two dimensions. The theory is based on the following assumptions, known as the Kirchhoff hypotheses [11: 1. Straight lines perpendicular to the midplane (i.e., transverse normals) before deformation remain straight after deformation. 2. The transverse normals do not experience elongation (i.e., they are inextensible).
302 THEORY AND ANALYSIS OF PLATES 3. The transverse normals rotate such that they remain perpendicular to the mid surface after deformation. These assumptions allow us to describe the plate deformation in terms of certain displacement quantities. Assumption (1) requires that the displacement field be a linear function of the thickness coordinate z: u(x9 y, z, r) = «o + zF\ (x, y, t)9 v(x, y, z, t) = do + zF2(x, y, 0, (8.5) u>(*, y9 z, 0 = wq + zF3(a\ y, /). where (wo, vo>wq, F\, F2, F$) are functions to be determined such that the remaining two assumptions of the Kirchhoff hypothesis are satisfied. The inextensibility assumption (2) requires that -— = 0 -> F3 = 0 for all jc, y9 and t. dz Thus, w is independent of z, i.e., w = wq(x9 y, t). Assumption (3) requires that (8.6) dlt _ dw _ 3wq dz dx dx dv dw _ du)Q dz dy dy Hence, the displacement field (8.5) takes the forni 3 wo u(x9 y, z, t) = w0(a% y91) - z-—, ax v(x9 y, z, 0 = vo(x, y91) - z——, (8.7) dy w(x9y9z,t) = wo(x9y9t). Thus, the Kirchhoff plate deformation is completely determined by the functions (wo, vo, u>n), which denote the displacements of a point on the midplane along the three coordinate directions. Note that the displacement field (8.7) will result, as will be seen shortly, in the neglect of all transverse strains, i.e., szz = sxz = syz = 0. Displacement-Based Shear Deformation Theories The simplest displacement-based plate theory that accounts for transverse shear strains is one based on the displacement field [2-81 ' u(x9 y9 z) = "o + z<f>x(x* y), v(xy y9 z) = vq + z<f>y(x9 y)9 (8.8) w(x9y9z) = wo(x,y)9
INTRODUCTION 303 where <f>x and — <py denote rotations about the y and x axes, respectively: In this theory, the normality restriction is removed by allowing for independent and arbitrary rotations {(f>x, <f>y) of a transverse normal. The theory is known in the literature as the Mindlin plate theory, although the use of the displacement field (8.8) and associated plate theory were developed much earlier by Bassett |2], Hildebrand et al. [31, and Hencky [4J. Mindlin [5J extended the theory developed by Hencky [4] to the vibration of crystal plates. Hence, it would be incorrect to attribute the theory to Mindlin alone. The theory is now being referred to as the first-order shear deformation plate theory (FSDT); see Reddy [6-8]. Note that the classical plate theory (CPT) is also a first-order theory in the sense that the first-order terms in the thickness coordinate are included; however, it is not a shear deformation theory. As will be seen later in this chapter, the transverse shear strains in FSDT are represented as constants through the plate thickness. Therefore, the transverse shear stresses are also constant through the plate thickness, whereas the stress equilibrium equations predict them to be quadratic. To make the shear forces (Ca-i Qy) computed in the FSDT to be equal to those obtained using the transverse shear stresses from the stress-equilibrium equations, shear correction factors were introduced. In both CPT and FSDT, the plane-stress state assumption is used and the plane-stress reduced form of the constitutive law is used. In both theories, the inex- tensibility (i.e., ezz = 0) and/or straightness of transverse normals can be removed. Such extensions lead to second- and higher-order theories of plates. Second- and higher-order plate theories1 use higher-order polynomials [i.e., n > \ in Eq. (8.1)] in the expansion of the displacement components through the thickness of the plate. The higher-order theories introduce additional unknowns that are often difficult to interpret in physical terms. A third-order plate theory with transverse inextensibility is based on the displacement field [9-16]: «(*, v, z, t) = uq(x, v, t) + f{z)<px(xy y9 0, v(x, v, z, t) = vq(x, y, t) + f(z)(py(x, yy f), (8.10) w(a\ y9z,t) = wo(x,ytt), where f(z) is a cubic function of z. There are number of people who have used a displacement field of the form (8.10), but they differ in actual form because of the choice of variables; consequently, the resulting governing equations have different looks. However, it can be shown that all third-order theories are special cases The order referred here is to the degree n of the thickness coordinate in the displacement expansion and not to the order of the governing differential equations.
304 THEORY AND ANALYSIS OF PLATES of that derived by Reddy [9-11] for laminated plates and shells using the displacement field in Eq. (8.10) and the principle of virtual displacements. The displacement field (8.10) accommodates quadratic variation of transverse shear strains (and hence stresses) and vanishing of transverse shear stresses on the top and bottom of a plate. Thus, there is no need to use shear correction factors in a third-order plate theory. The third-order plate theory of Lcvinson [16J is based on the same displacement field as in Eq. (8.10), but he used the equilibrium equations of the FSDT, i.e., he did not use the principle of virtual displacements to derive the equilibrium equations. The Levinson theory results in much simpler equations and does not involve the higher- order stress resultants. However, the theory leads to an unsymmetric stiffness matrix even for linear problems. Stress-Based Theories The plate theories based on the expansion of the stress field are due to Reissner [17-19] and Kromm [20,21], and the book by Pane [22] contains chapters devoted to these theories and their extensions. In the Reissner plate theory, the distribution of the stress components through the plate thickness, for the static case, is assumed to be of the form 0xx = z- 12MYA- A3 ' MM Oyy = Z~ yy <?X7 = 3&v 2/2 m G?7 = /?3 <JyZ = <*xy = Z- \2MXy 2h -(!) A3 '■ 2" -'(?)♦(?)' (8.11a) (8.11b) (8.11c) where (MXXi Myy,Mxy) are the bending moments per unit length, (Qx, Qy) the shear forces per unit length, q is the distributed transverse load per unit area, and h is the plate thickness. The stress field in Eq. (8.11a) is the same as that of the classical plate theory (to be shown shortly); the transverse shear stress field in Eqs. (8.11b,c) is the same as that obtained from 3D stress-equilibrium equations after using the in-plane stress field from (8.11a). Thus, the in-plane stresses are linear functions of z, the transverse shear stresses are quadratic functions of z, and transverse normal stress is cubic in z. Obviously, these stress components satisfy the stress-equilibrium equations of 3D elasticity, Eqs. (3.13). The transverse displacement w of Reissner's theory is a function of (x,y,z), and this complicates its determination. To make the theory tractable, Reissner introduced the thickness- integrated transverse displacement (a "mean deflection with respect to the plate thickness"): 3 fh/2 I /2z\ dz. (8.12)
CLASSICAL PLATE THEORY 305 The refined theory of Kromm [20,211 is based on a more general stress distribution, especially for the transverse shear stress components across the thickness of the plate. For a complete description of theory and its governing equations, the reader may consult the book by Pane |_22~|. The stress-based theories have not received as much attention as the displacement- based plate theories. This might be due to the fact that the stress-based theories are relatively more complicated, and inconsistencies between the actual (i.e., consistent with the assumed stress distributions) and adopted displacements may exist. In this chapter we consider only small strain and small displacement bending deformation of plates; i.e., the stretching deformation is omitted in the derivation of the governing equations, as the equations governing stretching deformation are not coupled to bending deformation. Therefore, in the absence of loads that cause stretching, the extensionai displacement field (uq, uo) is identically zero. We begin with an assumed displacement field, compute strains, and then use Hamilton's principle (or the dynamic version of the principle of virtual displacements) to derive the governing equations of motion. 8.2 CLASSICAL PLATE THEORY 8.2.1 Governing Equations of Circular Plates Consider a circular plate on a linear elastic foundation, with foundation modulus /c. Let r denote the radial coordinate outward from the center of the plate, z the thickness coordinate along the height of the plate, and 0 the coordinate along a circumference Figure 8.2 A circular plate with stress resultants.
306 THEORY AND ANALYSIS OF PLATES of the plate, as shown in Fig. 8.2. In a genera) case where applied loads and boundary conditions are not axisymmetric, the displacements (ur, uq, uz) along the coordinates (r, 0, z) are functions of r, 0, and z coordinates. We begin with the following displacement field implied by the Kirchhoff assumptions: t a +\ dw° or uz(r>e,zit) = wo(r%0tt), where wq is the transverse displacement at point (x, y) in the plate. The linear strain components referred to the cylindrical coordinate system are given by (see Exercise 3.34) dllr Ur 1 dltg or r r 30 8uz dz 1 (\ diiy due u0\ /Q ... ^ = 2(7^ + ^7-yj' (8-14) 1 (duo 1 #wA 1 (Bur duz\ For the choice of the displacement field in Eq, (8.13), the only nonzero strains are err = z*#\ s#e - tew* 2er0 = zyr00\ (8.15) where s(l) = d2w0 ' 3r2 ' - \ dr (1) i (dm ^ i d2m\ ,Q1~ ™ r \drd0 r 30 J' (1) _ 2 /d2?jJo 1 dwp Next we use Hamilton's principle Ju (8U + 8V-8K)dt=0, (8.17)
CLASSICAL PLATE THEORY 307 to derive the Euler-Lagrange equations. The virtual strain energy W is given by rh/2 su f f = / / (O'rr^rr + ^99^66 + VrO&Yrd) dz rdrdO JQq J-h/2 f r I* d2sw° iM fssm ^ id28m\ 2 . /92«w<> L'dSwoX —M,.q I r \ drdO r d0 } rdrdO, (8.18) where h is the plate thickness and (Mrr> Mro, Mqq) are the moments per unit length: rh/2 rh/2 M> orrz dz, Mob = / aeez dz, -A/2 J-h/2 J-h/2 Mi rh/2 ro = / <yrez dz. J-l (8.19) The virtual total potential energy due to applied loads is calculated as follows. Suppose that q = q(r, 0)is the distributed transverse load, Fs is the reaction force of a linear elastic foundation, and (JV,T, N&0, N}$) are the compressive and shear forces per unit length applied in the midplane of the plate. Wc have 8V = — I (4 + Fs)8wq r drdO -i [*- Jn0 L Swq ddwo a 1 3wq 1 SSwq dr dr r 30 r 30 + rdrdO, (8.20) where the foundation force is given by Fs = —kwo, and k is the modulus of the elastic foundation. The virtual kinetic energy SK is 2dwodSwo z? Bwq 1 38wq . ft. 1 z ~ - 1—- -^ ^tt + woou>o r dr dr dr r dO r 30 Jn0 J-I f [ /Bwo 38w0 , 1 dm dSw0\ . 1 = L r U7^ + ^i^ir)+Wu*JrdrM* (8*21) rh/2 SK=I p too «>-/*/2 where /o is the principal mass inertia, and I2 is the rotatory inertia: h= / z2pdz=P—, -h/2 J-h/2 1* and the superposed dot on wq denotes differentiation with respect to t. fh/2 h = I pdz = ph, J-) (8.22)
308 THEORY AND ANALYSIS OF PLATES Next, substitute the expressions for 8U, 8V, and 8K into Eq. (8.17), and carry out integration by parts with respect to /\ 0, and / to relieve 8wq of any derivatives. Note that 8wo(ry 0, t\) = 0 and 8wq(i\ 0, (2) = 0. The mixed derivative term 328wo/3r30 in Eq. (8.18) requires special attention, instead of integrating by parts with respect to r and then with respect to 0, or vice versa, it should be split into two terms so that the integration by parts is carried out in a symmetric way: d28wo 3r30 -drcW f MrB( d2Sw0 drdO d28w0\ , , + lm7)drde- We obtain -jTjU-H^- dMoe 1 d2Mee &M, dr r dO2 drdd \e 2dMr$-] 9 r 36 J + + r [ dr V dr J 1 a2wp r* 96>2 m I< (rMrrnr + Mrene) 0 r t °r h~ ~„n + '0^0 J 8wqi' drd0dt d8u)Q 1 / xjr xjr x 38wq + - (rMrtnr + Moon0) ■ a/. + [Tr<rM») + SMrQ 30 Moo nr + ISMpo 3M} r 30 +~di- — + -Myo no r /3wq 1 0u)o \ ] „ f „, . +/2 I —«f + -^^fno 1 [ Swo r drdOdt, no (8.23) where (nr, no) are the direction cosines of unit normal n = nrer + /i^e^ to the boundaiy T of the domain (midplane) Qq of the plate. Setting the coefficient of 8wq to zero inside Qq, we obtain the Eulcr-Lagrange equation: 1 ra2 + dMee \32Med d2Mr8 2dMre~\ dr r B92 drdd r 39 J (8.24)
CLASSICAL PLATE THEORY 309 The natural boundary conditions follow from the boundary expressions in Eq. (8.23). Assuming that Swq, dSwo/dr, and dSwo/30 are arbitrary, which means that wq, dwuf'dr, and dwo/dO are not specified on the boundary, we obtain 8wq: SSwq dr SSwo rV,ny + Vgne = 0, rMrrnr + Mre tig = 0, rMyeiiy + M^ng = 0, where Vr and Vq are the effective shear forces: 2 BO K = Gr - - I rNrr— + Ntf— + ^ —, r \ 9r 30 / Or 1 /a ] du)o a 9«>(A r 1 fir n , 3M0B , ,, re) + -TT- + M#-0 (8.25a) (8.25b) (8.25c) (8.26a) (8.26b) (8.26c) (8.26d) The equation of motion, Eq. (8.24), can be expressed, in view of the definitions (8.26c,d), as -H£K2)+^SK('-SK('-S)]- (8.27) For the axisymmetric case (i.e., when the materia) properties, load, and boundary conditions are independent of 0), Eq. (8.27) becomes ~T- (rfir) + too + -— (rJVrr—H ) + - -72 — r—^- + 70u*) r Or r dr \ dr ) r [ Or \ dr / (8.28) for & < r < a> where a and fr are the outer and inner radii, respectively, of an annular plate. For the solid circular plate,-we set b = 0. This completes the derivation of the governing equations of a circular plate. The equations are valid for any suitable constitutive behavior of the plate material. In order to express the governing equations in terms of the displacement wq9 the bending moments in Eq. (8.19) must be expressed in terms of the displacement wq via
310 THEORY AND ANALYSIS OF PLATES stress-strain relations and strain-displacement relations. For an isotropic linear elastic material, assuming that the elastic stiffnesses are independent of the temperature, the stress-strain relations are I see Eqs. (3.45) and (3.4 l)J: arr E 1-v2 1 v 0 v 1 0 0 0 ^ srr — a AT \ eee-a AT 2t>0 (8.29) where AT(t\ 0, z) is the temperature rise from a stress-free (or undeformed) state, a is the coefficient of thermal expansion, E is Young's modulus, and v is Poisson's ratio. Then we find that M; rh/2 Eh3 12(1-v2) Eh3 f 'rr = / Vrrt dz = ./-A/2 rh/2 7-A/2 12d - V2) rh/2 Mro = I Or$zdz = J-h/2 (8.30) G/z3 12 v(1) where MT is the thermal moment: M' Ea Ch'2 Ea f ~ 1 - v }_ ATzdz. ■A/2 (8.31) Now, the stress resultants can be expressed in terms of the deflection wo using the strain-displacement relations (8.16) as Mr, ^-(l-v)/)-^—--—J. (8.32a) (8.32b) (8.32c) The bending stiffness D is given by D = Eh2 12(1 -v2)' (8.33)
CLASSICAL PLATE THEORY 311 The equation of equilibrium (8.27) can now be written in terms of the deflection wo as ri a / a \ i a2 *| rl a / dw0\ i a2u>ol , ir a / 3w0\ , l a2w0 + /QW0 ri a / dMT\ i 32Mrl (8.34) (8.35a) r2 802 For the axisymmetric case, Eq. (8.34) simplifies to 7a7 |ra7 [^ (rirjJ| +kw°+ 7Tr Kirj For the static case, Eq. (8.35a) becomes (8.35b) The standard boundary conditions for axisymmetric bending of solid circular plates and annular plates (see Fig. 8.3) are listed in Table 8.1. h T TTmmTTTTTTTMTTTTTT - A T JUIUJ. fm"T * a 2, W>o(>') z, u>o(r) Figure 8.3 Axisymmetric circular and annular plates.
312 THEORY AND ANALYSIS OF PLATES Table 8.1 Typical boundary conditions for solid circular and annular plates Plate Type/Edge Free Hinged Clamped r = 0 (for all cases) r = a r=b dr = 0 Qr Mrr Qr Mrr Qr Mrr = Qa = Ma = Qb = Mh = Qa = Ma Circular Plate 2n r Qr = w0 = Q Mrr = Ma Annular Plate wo = 0 Mrr = M^ WQ = 0 Mrr = Ma -Go WQ = 0 dr w0 = 0 dr wo = 0 dr Qa* Qb, Mat and M/, arc distributed edge forces and moments and Q$ is a concentrated force. 8.2.2 Analysis of Circular Plates Analytical Solutions: Bending Here, we develop analytical solutions of isotropic circular plates for axi symmetric boundary conditions and loads (sec Reddy [8| for additional details). For axisymrnetric circular plates with fc=0, MT = 0, and Nn. = 0, Eq. (8.35b) simplifies to 7htt[,hffl]br[-'>hffl+>*>. For the static case, this further simplifies to D d f d p d ( divpX r dr \ dr \_r dr \ dr J = <!> = q. (8.36a) (8.36b) and the stresses, bending moments, and shear force are related to the deflection by Ez (d1 Gyy = - : Ariup vdwp\ v*j \ dr2 r dr J ' Ez ( d2wo 1 dwo\ M, (d2 wo vdwo\ rr = -V\ -T^r + --J- . V dr dr J ( d2wp 1 dwo\ -\d_ (rdmY\ jdr \ dr J]' Qr = -D (8.37a) (8.37b) (8.38a) (8.38b) (8.38c)
CLASSICAL PLATE THEORY 313 The general solution of Eq. (8.36b) can be obtained by successive integrations [see Eq. (8.28)]: ~rQ(r) = j rq{r)dr + cu d ( dw0\ f 1 f Ddwo ^ ff(r) + r y _ ^ + r + 1^ dr 4 2 r 2 2 r r Dwo(r) = F(r) + — (log r - i) C] + —c2 + c3 log r + c4, (8.39) (8.40) (8.41) (8.42) where F(r)= J - jr J - f rg(r)drdrdrdr9 F'(r) = - fr f - f rq{r)drdrdr, (8.43) and Cj (i = 1,2, 3,4) are constants of integration that will be evaluated using boundary conditions. Note that d2wo // , 1 , 1 1 drz 4 2 vz (8.44) For solid circular plates, the requirement that the slope be zero due to symmetry at r = 0 yields c.3 = 0 |scc Eq. (8.41)1. Further, if there is no point load at r = 0, the shear force is zero there, giving c\ = 0. Example 8.1 Consider a clamped circular plate under distributed load. For this case, c\ = c'3 = 0. The boundary conditions associated with the clamped outer edge are wo = 0 anddwo/dr = 0 at r — <7 (seeFig. 8.4). These conditions yield c2 = — 2F (a) c4 = ~F(a) + F (a)a Qo Qo * hTTTTIMTTTTTfj _^ L -* . a H *2, Wiy(r) h-* a friz, iVo(r) (a) (b) Figure 8.4 A clamped circular plate under (a) a uniformly distributed load and (b) a point load.
314 THEORY AND ANALYSIS OF PLATES For uniformly distributed load qo> wc have qua2 qoa4 Hence the deflection becomes ,4/ ,2X2 64 D and the maximum deflection occurs at the center of the plate (r — 0): ~4 m(r) _££(!_£.), (8.45) ^;»^v = 7777. (8.46) Expressions for the bending moments and stresses become 2 r 2 *i M,T(r) = *2fL I (1 + v) - (3 + v)^|, (8.47a) 2 r -2 ~i JfM(r) = ^fL I (1 + v) - (1 + 3v)£j I, (8.47b) *r(r.z) = ^[(l + v)-0 + v)^], (8.47c) o»«(r, z) = ^^ [d + v) - (1 + 3v)^J . (8.47d) The maximum (in magnitude) values of the bending moments are found to be at the fixed edge: *„<*) =-«£ Moo(a) = -V-*f. (8.48) Hence, the maximum stress is given by orr[a. --)= JJ3- = — (j) • (8-49) For a solid circular plate with a point load £>o at the center, we require 2n (rQr)o = -Qq. From Eq. (8.39) we obtain Qo The boundary conditions of the clamped edge at r = a give 2°^i in Goa2 c2 = -— (21oga - 1), c4 - —— - 4?T 167T
CLASSICAL PLATE THEORY 315 Hence, the solution becomes Go"2 T r2 r2 /r\l u,0('-) = T6^[,-^ + 2?lo8(«)j' (8-50) Mrr(r) = -^ [l + (1 + v) log Q] , (8.51a) Mooir) = ~ [v + (1 + v)log Q] . (8.51b) The maximum deflection occurs at r = 0 and is given by uw = -g^. (8.52) Example 8.2 Consider a simply supported annular plate under uniformly distributed load of intensity c/o- The boundary conditions are At r = b: Mrr=0, (rfir) = 0, (8.53a) AXr =cr. wo - 0, Mrr = 0. - (8.53b) These conditions give c\ = —qob2/2 and fl + 3v\q0b2 (3 + v\q0a2 anb2. q0b4 . a (3 + v\q0b2a2 /l+v\ ?oa2/?4 , a qoa4 f3 + v\q0a2{a2-b2) /3 + v\q0b2a2l C4 = -^r + (jT^J—32—+{i=z)-i6-]x*a + 8^) W + {—v) lk^) l°ga ^ The deflection and bending moments become A (3+")[|-D2]-^<3+")[i+0I WO 04 i + 4(1 + v)j82k [l - (£)2j +4(1 + v)^2logQ[ , (8.55a)
316 THEORY AND ANALYSIS OF PLATES M09 = ^r-\{3 + V) 16 + ,-02]+^[(5v-1) + (3 + v)G)2] 4(l + v)^[l + Q2]j> (8.55b) where a, =(3 + v)(l -p2)-4{l + v)P2K, a2 = (3 + v)+4(I +v)k, K = -^5 log ft /» = -. For a simply supported solid circular plate under uniform load, we set b = 0 or P = 0 (hence, ci = c3 = 0) in Eqs. (8.54)-(8.55a,b) and obtain U)q = Mqb [v«/ 1+vU + 1 + vJ' .^[0 + ,).(1+3.)(^. 64D M,T = ^(3 + JO (8.56) (8.57a) (8.57b) The maximum deflection, bending moments, and stresses occur at r — 0, and they are W,: ax ~ (fr^) 64/)* Mmax = (3 + v) 16 ' ^m«A- = 3(3 + V) <7Qfl 8/i2 * (8.58) Finally, for a simply supported solid circular plate with a central point load Qq, we have c\ = Qo/2n and C3 = 0. Using the boundary conditions wo(a) = 0 and Mrr (a) = 0 gives C2 = _g[Iogfl + ^|^)]1 C4=(|±J) go«2 ltor The solution for any r ^ 0 is given by ,2 7T Vfl/ <7/t04) = - Cty0 /z3 3zGo (T)-^ [(! + .)*. ©-(.-.)]. (8.59) (8.60a) (8.60b)
Mrr(r) = God + v) An te«G)' Go M^(r) = ~ [(1 + M)log Q - (1 r- v)] CLASSICAL PLATE THEORY 317 (8.61a) (8.61b) The maximum deflection is given by Wmax = WQ(0) = Qoa2 16tvD m- (8.62) The stresses and bending moments cannot be calculated at r = 0 due to the logarithmic singularity. The maximum finite stresses produced by load Qo on a very small circular area of radius rc can be calculated using the so-called equivalent radius re in Eqs. (8.60a,b) and (8.61a,b) (see Roark and Young [23]); re = yi.6/f + A2 - 0.675/7 when rc > 1.7ft. when rc < 1.7/?, '« ™ rc (8.63a) (8.63b) Analytical Solutions: Buckling Here we consider exact buckling solutions of circular plates. Consider the shear force-deflection relation (8.38c): Qr = -D 1 d / d2w0\ _ J. r dr \ dr2 J r"A d.U)Q 2 dr (8.64a) We write the equation in ternis of 0 = (dw$fdr)y which represents the angle between the central axis of the plate and the normal to the deflected surface at any point (see Fig. 8.5), as °>--°W%)-$. (8.64b) For a circular plate under the action of uniform radial compressive force Nrr = No per unit length, the shear force at any point is given by [an integration of Eq. (8.28), for the static case with q = k = 0, gives the result, because Qr = 0 at r = 0 for axisymrnctric mode]: * dwn Qr = Nrr~ = Nrr4>. dr (8.64c) From Eqs. (8.64b,c) we obtain -°[;U'%)-?*]-** or dr ( d<t>\ + £H-a (8.65)
318 THEORY AND ANALYSIS OF PLATES Figure 8.5 Buckling of a circular plate under uniform radial compressive load. Equation (8.65) can be recast in an alternative form by invoking the transformation r = rat a2 = ~p (8.66) and the alternative form of Eq. (8.65) is (with n = 1) 'i('-|)+'r"2-"2»*=0' <8-CT) which is recognized as the Bessel differential equation of order n. The general solution of Eq. (8.67) is <Kr) = AMr) + BYtt<r)% (8.68) where Jn is the Bessel function of the first kind of order ra, Yn is the Bessel function of the second kind of order n, and A and B are constants to be determined using the boundary conditions. In the case of buckling, we do not actually find these constants but determine the stability criterion. Example 8.3 Clamped Plate For a clamped plate, the boundary conditions are (dwo/dr is zero at r = 0, a) 0(0) = 0, 4>(a) = 0. ' (8.69) Using the general solution (8.68) for an isotropic plate, wc obtain AJ} (0) + BYi(0) = 0, AJX(oca) + BY\ (aa) = 0, a2 - -~.
CLASSICAL PLATE THEORY 319 Since J\ (0) = 0 and Y[ (0) is unbounded, we must have B = 0. The fact that B = 0 reduces the second equation (since A ^ 0 for a nontrivial solution) to J\ (aa) = 0, (8.70) which is the stability criterion. The smallest root of Eq. (8.70) is aa = 3.8317. Thus we have Ncr= 14,682—. (8.71) Simply Supported Plate For a simply supported plate, the boundary conditions are (dwo/dr = 0 at r = 0 and Mrr = 0 at r = a) *(0) = 0, [^7 + v7^1 =0- <8-72a) The second boundary condition can be written in terms of r as |~p + v^l =0. (8.72b) ldr r lT=aa Using the general solution (8.68), we obtain A/i(0) + BYi(0) = 0, AJ[{aa) + BY[(aa) + — [AJ\(aa) + BY{(aa)\ = 0. aa The first equation gives 5 = 0, and the second equation, in view of B — 0 and the identity dJ„ I -JL = jn_{(r)--Jn(^ (8.73) dr r gives aaJ^ipta) - (1 - v)/i(aa) = 0, (8.74) which is the stability condition for the simply supported plate. For v = 0.3, the smallest root of the transcendental equation (8.74) is aa = 2.05. Hence the buckling load becomes #cr =4.198-^. (8.75) a1 Simply Supported Plate with Rotational Restraint For a simply supported plate with rotational restraint (see Fig. 8.6), the boundary conditions are 0(0) = 0, D r-^ + i;0 +kRa4>(a) = 0J (8.76)
320 THEORY AND ANALYSIS OF PLATES Torsional spring, kR i a »4* a *■ *z, w(r) Figure 8.6 Buckling of a rotationally restrained circular plate under uniform radial compressive load. Table 8.2 Buckling load factors N = Ncr(a2P) tor rotationally restrained circular plates /*-> N -► 0 4.198 0.1 4.449 0.5 5.369 1 6.353 5 10.462 10 12.173 100 14.392 00 14.682 where /cj? denotes the rotational spring constant. We obtain aaJfaa) - (I - v - P)J\(aa) = 0, 0 = —, (8.77) as the stability condition. When /if = 0, we obtain Eq. (8.74), and when (1 = oo, we obtain Eq. (8.70) as special cases. Table 8.2 contains numerical values of the buckling load factor N = Ncr(a2/D) for various values of the parameter, fi (for v = 0.3). Variational Solutions: Bending Next, we consider the Ritz and Galerkin solutions for axisymmetric bending of circular plates, possibly on an elastic foundation. Toward this end, we first write the variational statement (i.e., the weak form) needed for the Ritz method. From the virtual work statement in Eq. (8.17), with (8U, SV) from Eqs. (8.18) and (8.20), 8K = 0, and omitting terms involving derivatives with respect to 0, we have 0 = -2tt wo 1 cISwq fa ( d28nxt „ --U „ / Mrr J "+Moe ^-kwQ8wo+q&WQ-Nrr—rL—r^ )rdr Jb \ "r r dr dr dr J £ dwodSwo\ ~d^ + 27t\aQa8m(a)-bQbSwo(b)-aMa(^^j +bMh(^^\ 1
CLASSICAL PLATE THEORY 321 ,2n r\D\(^+^)fp^+(v^+i^Yd^] Jb l l\ d* r dr J dr1 \ dr1 r dr J r dr J . # * * 1C7 dxvodSwo] dr dr J + 2n [aQaSwo(a)—bQbSwo(b)] + 2n /dSwo\ , . .. fd8wQ\ (8.78a) where a and b denote the outer and inner radii of an annular plate, k is the spring constant of the linear elastic foundation, D the bending stiffness, q the distributed load, Qa and Qj7 the intensities of effective line loads (including in-plane compressive force) at the outer and inner edges, respectively: Qa = Q ~ Nrr dwo 17 Qb = \Q-Nrr^r] , (8.78b) L dr \r=b and Ma and M[} the distributed edge moments at the outer and inner edges, respectively. When b — 0 (for a solid circular plate), we have M]? = 0 and 2nbQi7 = £>o> go being the applied point load at the center of the plate. Equation (8.78a) is the weak form used in the Ritz method. The weak form (8,78a) should be modified if the edge r = a of the plate is elastically restrained (see Fig. 8.7): (rQr)r=a + kEw0(a) = 0, (-rM,),=a + kR (^\ - 0, (8.79) where kx and J<r are spring constants associated with extensional and rotational springs, respectively. One can simulate the simply supported boundary condition (/c£ = oo and kR = 0), the clamped boundary condition (fc# = oo and I<r = oo), Rotational spring, kR ^%« la f a ^^ Extensional spring, kPi -5- j ~3*"* *z, u>(r) Figure 8.7 An elastically restrained annular plate.
322 THEORY AND ANALYSIS OF PLATES and free edge condition (k^ = 0 and Izr = 0). Of course, similar expressions can be written for the edge at r — a. The weak form (8.78a) takes the form a n fa\,Jfd2wo , vdw0\d28wQ , / d2w0 , I dw0\ \ dSw0l + kwoSwo — q$wo - Wrr—p- —— irdr A7- ar J + 27T [*^5iyo(a)wo(«) — bQb8wo(b)] + 27r|**(^^| +bMb( +*'^)J- (8.80) Assume an TV-parameter solution of the form (assuming that all specified geometric boundary conditions are homogeneous so that we can take 0oO') = 0): 7 = 1 (8.81) The approximation functions <\>\ for the Ritz method must be continuous as required by the weak form, linearly independent, and satisfy the homogeneous form of specified essential boundary conditions. Substituting (8.81) into Eq. (8.80), we obtain or [AIM = {b}t (8.82a) where Clfj—lTtD [ d2<f>id2<t>j v/d^d2^ d24>jd4>j\ 1 d(j)id<j)j dr2 dr2 r \ dr dr2 dr2 dr f r2 dr dr rdr (8.82b) fa fa * d<t>i d<bi +2/C7T / <k4>j rdr-ln \ Nrr-TLJ7L rdr, Jb Jb dr dr bi=2n j qb rdr-2jt\aQa<t>i(a)-bQbMb)-a +bMb\jf) (8.82c) Similar expressions can be obtained using the variational statement (8.80). Once 0/ are selected, a.;j and b\ can be computed by evaluating the integrals in Eqs. (8.82b,c), and Eq. (8.82a) can be solved for C{ (I = 1,2,..., AT). Then, the solution is given by Eq.(8.81).
CLASSICAL PLATE THEORY 323 For a weighted-residual method, we use the weighted-integral statement Id ( .y dW0\ ' r dr, (8.83) where 1/// (r) is the weight function that takes different forms depending on the method used. Substituting the N-parameter approximation of the form N wo(r) « WN(r) = ^cjtjir) + 0o(r) 7-1 (8.84) into Eq. (8.83), we obtain Eq. (8.82a) with atj ~2nD (8.85a) (8.85b) The approximation functions fy must be continuous as required by the weighted- integral statement (8.83), linearly independent, and satisfy the homogeneous form of all specified boundary conditions, while <£o satisfy all specified boundary conditions. Example 8.4 As a specific example, we consider a simply supported solid (b = 0) circular plate under uniformly distributed load of intensity qo, as shown in Fig. 8.8. Note that Nrr = 0 here. ?o » I ' ' ' r t t t T T T T 1 TJ-f fr a * z, w0(r) Figure 8.8 A simply supported solid circular plate under uniformly distributed load.
324 THEORY AND ANALYSIS OF PLATES Recall that for this problem the essential boundary conditions are dr w0(a) = 0, and the natural boundary conditions are given by Mrr = 0 at r = ay Qr = - r ' d — (rMrr)-M90 dr = 0 at r = 0. The natural boundary conditions will have no bearing on the selection of 0/ in the Ritz method. Each 0,- must satisfy the homogeneous form of the geometric boundary conditions ^(0)=0, 0/(a)=O. dr For the choice of algebraic polynomials, we assume 0i(r) —a\ + a2r + «3'*2, and determine, using the above conditions, that 0:2 — 0 and <x\ + a$a2 — 0, Thus, we have (for the choice of ot\ = 1) 01 = 1--. or The procedure can be used lo obtain a linearly independent and complete set of functions: ..3 01 = 1 02 = 1- *--© rw+i (8.86) Substituting for 0/ from the above approximation into Eqs. (7.82b,c), and noting that k = 0, we obtain ■2(^ + 1) (8.87) For N = 3 we have the algebraic equations D ^2 4(1+y) 6(1+ v) 8(1+ v) 6(1+ v) 9(1.25 + v) 12(1.4+u) 8(1+ v) 12(1.4+ v) 16(f+v) J C3 60 15 18 20 (8.88)
CLASSICAL PLATE THEORY 325 The one-, two- and three-parameter Ritz solutions are .4 / „2> *.(*•) = Cm 160(1 + ^0-?)' W2(° = 80D\T+Tj I1 ~ ?J - 30^ V ~ ?j ' w,. <?o«4/6 + 2v\/ r2\ ^4/, >A\ = fof^ \5 + v _ 2 (3 + v\ (L\2 + (L\ 64D [l + v \) +v)\a) W (8.89) The three-parameter solution (8.89) coincides with the exact solution in Eq. (8.56): Example 8.5 Consider axisymmetric bending of a clamped (at r = a) solid circular plate under uniformly distributed transverse load of intensity 40 (acting downward), and with Nrr = 0. The boundary conditions are wo(a) = 0, — = 0 at r = 0, a, (8.90a) The three boundary conditions in Eq. (8.90a) are of the essential type while that in (8.90b) is of a natural type. Approximation Functions First, we discuss the selection of the approximation functions. Obviously, </>$ is zero since all specified boundary conditions are homogeneous. The choice of <\>\ depends on the method we use. We derive them for both the Ritz and weighted-residual methods. In the Ritz method the approximation functions are required to satisfy the homogeneous form of only the geometric boundary conditions. Since there arc three essential boundary conditions, we begin with a four-parameter polynomial, 01 (r) = cq + c\r + cir2 + c3r3, and determine the constants (one of them is arbitrary but nonzero) such that the three conditions 0,'(O) = O, 4>i(a) = 4>'i(a) = 0
326 THEORY AND ANALYSIS OF PLATES are satisfied. We obtain (co = a3C3/2, c\ = 0, ci = —3ac3/2): *i=l~3G)2+2©3- (8-91a) Similarly, we pick the five-term polynomial for <j>2 and determine the constants (two of them are arbitrary but the coefficient of the highest-order term, C4, should never be taken as zero): 02 (>') = co + cir + c2r2 + c$r* + C4/-4. We obtain Co = \a: 6'3 + C\a , C\ — 0, C2 — — 2aC'3 ~ 2fl2C4, For simplicity, we take C3 = 0 and obtain <fe -©T- 1- - . (8-9lb) In the weighted-residual method, the approximation functions are required to satisfy the homogeneous form of all specified boundary conditions. Since there are four specified boundary conditions, we begin with a five-parameter polynomial, 4>\ (r) = co 4- cir + c2r2 + c3r3 + c4r4, and determine the constants such that the four conditions «<">-* *«-««-* {i;\;z(r*)]h-0 are satisfied. We obtain (co = aACA, c\ = 0, c2 = — 2a2C4, c$ = 0): 0)(r) = [l-Q2l . (8.92a) Next, we pick the six-term polynomial for <p2 and determine the constants (two of them are arbitrary but the coefficient of the highest-order term, cs, should not be taken as zero): <p2(r) = co 4- c\r + c2r2 + c$r3 + C4/'4 + c5r5. We obtain Co = a ca -f ^a cs, c\ = 0, c2 = —2a C4 — §« C5, C3 = 0, and *«-[i-0T-4«+H3-i©,+©']*
CLASSICAL PLATE THEORY 327 Since the first part is already represented by <j>\, wc set C4 = 0 and obtain fc<r) = 3-5(£)2 + ©5- (8'92b) Solutions For the two-parameter Ritz approximation, wc use the approximation functions 0i and <fe from Eqs. (8.91a,b) and compute aij and b, from (8.82b,c) (note that k, Qcl, Qb, Ma, and Mb are zero): I^l—40--). ^r = -4[i-2(-)l. r dr a1 \ a/ drl a2 I \a/\ 64D Thus, the two-parameter Ritz solution coincides with the exact solution in Eq. (8.45): *4 uv y 64£> [-©T- Note that the one-parameter solution is given by {c\ — bi/a\\): The maximum deflection obtained with the one-parameter Ritz method is in 10% error. In Galerkin's method, we first compute the residual using the functions in Eqs, (8.92a,b). We have (<t>0 = 0): dW2 d<j)\ d<f>2 dr dr dr d2W2 dr2 -:©K-)V;K-)+©> 7 d2W2 V2W2 = —-± + dr2 12 **-£<*"^7^-£■ + £(>•
328 THEORY AND ANALYSIS OF PLATES Also, note that d _ </ f </ ri d ( dWiXW r 64/r\ 225/r\2 1 Hence the coefficients «/y and b-t of Eqs. (8.92a,b) have the values 64D 120D 176/) flII="&5- *I2 = ^"' *2, = ~^' 225/> , (?o«2 , H^o«2 and the solution of the Galerkin equations gives C, = 64D' ca = a Therefore, the two-parameter solution is the same as the one-parameter solution, which coincides with the exact solution: 64D [ W n2 ana ' i / /* \ u>o(r) The next example illustrates the use of Betti/Maxwell reciprocity theorem in obtaining the deflections of circular plates under asymmetric loading. Example 8.6 Consider a circular plate of radius a with an axisymmetric boundary condition, and subjected to an asymmetric loading of the type (sec Fig. 8.9): q(r,9) = qo + q\-cosGt (8.93) a where qo represents the uniform part of the load for which the solution can be determined for various axisymmetric boundary conditions. In particular, the deflection of a clamped circular plate under a point load at the center is given by Eq. (8.50) and that of a simply supported circular plate under a point load at the center is given by Eq. (8.59). Here we wish to use the Betti/Maxwell reciprocity theorem to determine the center deflection of a clamped plate under asymmetric distributed load. By Maxwell's theorem, the work done by a point load (Qo = 1) at the center of the plate due to the deflection (at the center) u;c caused by the distributed load q(f% 6) is equal to the work done by the distributed load q{i\ 0) ,in moving through the displacement wo(r) caused by the point load at the center. Hence, the center deflection of a clamped circular plate under asymmetric load (8.93) is -ifefM-SO-**:) '■'"'"> = il£ (894)
CLASSICAL PLATE THEORY 329 q{r, 0) =qo + qi — cos0 qo-q simply supported z, w0(r) T " f | J ,T T M M M, 'S//S/ TTTTTTTTTT 1 X ^ a 1 z> w0(r) Figure 8.9 A circular plate subjected to an asymmetric loading. Variational Solutions: Vibration Determination of the natural frequencies of circular plates with various boundary conditions by analytical means leads to complicated equations involving Bessel functions (see [8,24-28]), and the solutions are difficult to obtain, requiring the use of approximate methods of solution. Here we use the Ritz method to determine the natural frequencies. The equation of motion of an isotropic plate is given by (see Reddy [8]) d~wo DV2V2w0 + kwo + I0^f - /2^(V2W) = 0, (8.95a) where /u = ph and h = ph3/12 are the principal and rotatory inertias, respectively, and the Laplace operator V2 is defined in the polar coordinates as *2 rdr V BrJ r2302 (8.95b) For free harmonic motion (i.e., natural vibration), the deflection can be expressed as w0(f\ 0, t) = w(r, 0) coscot, (8.96) where co is the circular frequency of vibration (radians per unit time), and w is a function of only r and 0. Substituting Eq. (8.96) into Eq. (8.95a), we obtain DVlVzw + kw - Iqco2w + l2(olVlw = 0, .2vy2, (8.97)
330 THEORY AND ANALYSIS OF PLATES The weak form of Eq. (8.97) is given by d2w d28w Jo, I v /dwd28w 9Swd2w + + - v /dwd28 1 d2w 328w ld28w<Pw\ rF ar2 + 7"ae2~ 8r2 J 3r 3r2 1 32*t /13u; 1 d2w\ n d8w_ 1 d28w\ ° V "37 + r5 30V V "^ + 'r2~Mr) /[ d2w 1 3uA /1 a2an; 1 d8w\ 2d - v)d {-^-^-^) {--^ " ;* i*/ V^r^7 + P23^"30"/JJr f 3w 3<5w 1 3w d8wy Assume an N-parameter Ritz approximation of the form N wfa 0) « WN(rt 0) = ]£ cjMr> 0)' 7 = 1 Substituting Eq. (8.99) into Eq. (8.98), we obtain N circle. 0 = E(-S)+^)--V)^J where a\: , a\, , and my are defined as (8.98) (8.99) (8.100) a'j ~ J a I dr2 Br2 \r Br r2 B92 ) \r Br r2 B92 J vhfrBHj d<j>jd2<Pi id2(/>id24>j l d2<i>j d2<f>j\ r \ Br Br2 Br Br2 r BO2 Br2 r B02 Br2 J 1 }\rdrdO r2B9)\rBrB0 r2 B6 a?) =k I <pi<pj rdrdB, rdrdB. rdrdS, (8.101a) (8.101b) (8.101c)
CLASSICAL PLATE THEORY 331 In matrix notation, Eq. (8.100) has the form of an eigenvalue problem: ([A(1) + A(2)] - a)2[M]){c} = {0}. (8.102) For a nontrivial solution, the coefficient matrix in Eq. (8.102) should be singular, i.e., |[A(1)] + [A{2)] - w2\M]\ = 0. (8.103) The above formulation is valid also for annular plates. Example 8.7 As an example, consider free vibration of a clamped solid circular plate. For a one-parameter (N — 1) Ritz solution, let (j>{ (r, 0) = f\ (/•) cos nO (8.104) and compute crn , au , and ma as \ r drd0t x cos2?20 +2(J - v) C-fl - -7/1) sin2nO a™ = k I / J\ J\ cos2 nO r drd0, Jo Jo mu=f f r(/o/l/l + /i7i)cos2n0 + /2^-/i/isin2/ie (8.105a) (8.105b) rdrdO. (8.105c) The conditions on the approximation functions arc ft (r, 0) = ^P- = 0 at r = a, and — = 0 at r = 0, (8.106) or dr which translate into the conditions f[(a) = f[{a) — //(0) = 0. Clearly, (he choice <r\2 /r\3 /lW.,-,(i)' + 2Q- (8.107) satisfies the conditions. Since f,_ 6r 6r2 6 12r it is clear that the integrals (for n > 0) /•a 1 pa 1 / -/i/r^. / -3/1/1* Jo r Jo >*3
332 THEORY AND ANALYSIS OF PLATES required in aKn} of Eq. (8.105a) do not exist because of the logarithmic singularity. Thus /i (r) defined in Eq. (8.107) is not admissible for n > 0. The next function that satisfies the boundary conditions f\{a) = f[{a) = f{(0) = 0 is /,«-[.-Of-.-2 ©' + ©\ (8..08, which is also not admissible for n > 0. The next admissible function is ™-:K)'H-2C)I+c-)'- <-») which will not present any problem in evaluating the integrals for n > 0. Here we seek the fundamental frequency corresponding to the axisymmetric mode, n = 0. For this case, we can use the function in (8.107) and obtain a<X = 2nD f° (V^ + ^f[f[ + ^flf'l) r dr = 2n^ r f\f\ Jo Ink r dr = 2n - 3kal 35 m u = lit jT (7b/i/i + hfifi) rdr = 2n Q^/o + ^/2) Hence 0) (9D 3ka2\ [3a2 T 3 1 ! D ^0 105 + (fa4/D) i+7(/2//oa2), (8.110) Cleai'ly, rotatory inertia has the effect of reducing the frequency of vibration while the elastic foundation modulus increases it. For a clamped isotropic plate without elastic foundation (i.e., k = 0), the frequency parameter becomes {h = Ioh2/12)\ X2 = a)a2J— = \0241 -1 1/2 1 + 0.583- h2 (8.111) For very thin plates, say h/a — 0.01, the effect of rotatory inertia is negligible. Even for h/a = 0.1, the effect is less than 1%. Note that the one-parameter Ritz solution differs from the analytical solution a> = 10.216 listed in Table 5.5.1 (for m = n — 0) of Reddy [8] by less than 0.5%!
CLASSICAL PLATE THEORY 333 For n = 1, the function in Eq. (8.109) may be used.. We obtain -,(') _ 7ZD N-G-^HH- (2) na2, fa2 1 \ «n =-60-*.- »«=jr(eo/o+6/2J- The frequency parameter is given by i2 2 MO -1 1/2 480 + 60D .1 + 10/2 a2/o (8,112) For a clamped isotropic plate without elastic foundation, the frequency for the case m = 0, n — 1, becomes (h = fall2/12): k2 = (oa2Jy= 21.909 1 + 5tf 6a2 J 1/2 (8.113) When rotatory inertia is neglected, the frequency predicted by Eq. (8.113) differs from the analytical solution co = 21.26 listed in Table 5.5.1 (for m = 0, n = 1) of Reddy J.8J by only 3%. For other values of n, one must select functions that are admissible (i.e., allow evaluation of the integrals). Generally, higher values of n require higher-order functions //(/*)• Variational Solutions: Buckling Next, we consider buckling of circular plates under uniform compression N in the middle plane of the plate (see Fig. 8.7). The governing equation is DV2V2w + NV2w = 0. (8.114) The weak form of this equation is a slight modification of that given in Eq. (8.98): 0: 2w 328w _ v /dw d2Sw dSw d2w 1 d2w d28w f f d2wd28w v /dwd28 0-G r-2 3r2 i d2Sw d2w 7 de2 dr2 + ffr dr2 + r 36>2 dr2 'dw 1 d2w )0 3Sw 1 d28ws + + N„/l'92w 1 dw\ (\ 2(1 ~V)D (7^-^w)(7 dr ' r2 SO2 d28w 1 d8w\ drdO r 2 86 *d8w dw . , f/i -N — \rdrd0. dr 3r (8.115a)
334 THEORY AND ANALYSIS OF PLATES The /^-parameter Ritz approximation (8.99) results in {[A]-N[G]){c} = [Oh (8.115b) where cnj are the same as aj^ in Eq. (8.101a) (with k — 0) and gu arc given by Example 8.8 As an example, consider a clamped circular plate under uniform compression. Since we are interested in the minimum buckling load, which occurs in the axisymmetric mode (n = 0), we can use the one-parameter approximation (see Example 8.6) W\ (r) = c\ f\ (r), where fx (r) is defined in Eq. (8.107). We obtain 9D -3 „ * 157) a2 5 a1 which differs from Ihe exact solution in (8.71), 14.682(2>/a2), by 2.18%. 8.2.3 Governing Equations In Rectangular Coordinates Here we consider the classical theory of plates in rectangular Cartesian coordinates. We wish to develop the total Lagrangian functional and to use Hamilton's principle to derive the equations of motion and natural boundary conditions for linear bending of a plate according to the Kirchhoff hypotheses. Stretching deformation is not considered as it can be uncoupled from bending deformation for the linear problem. Let us denote, as before, the undeformed midplane of the plate with the symbol £2o> and let it coincide with the xy-planc of the coordinate system with the z coordinate along the thickness of the plate. The total domain of the plate is symbolically expressed as the tensor product Q = £2o x (—h/2, h/2). The boundary of the total domain £2 consists of the top surface St(z = h/2), bottom surface Sj?(z = —h/2), and the edge f = r x (—h/2, h/2). In general, f is a curved surface, with outward normal ii = nxex 4- nyey, where nx and nY are the direction cosines of the unit normal (see Fig. 8.10). We begin with the following displacement field [cf. Eq. (8.7)], which is a direct result of the Kirchhoff hypotheses in the pure bending case (see Fig. 8.11): Swq diVQ l u(x, y, z> t) = -z-—-, v(x, y, z, t) = -z-r—, w(x, y, z, t) = uuo(*» y> 0» ox ay (8.116) where («, v, w;) denote the displacements of a material point in the (x, ys z)coordinate directions at timer.
CLASSICAL PLATE THEORY 335 Figure 8.11 Kinematics of deformation of the classical plate theory in rectangular Cartesian coordinates. Assuming small strains and displacements, the linear strains can be computed using Eq. (3.23). For the displacement field in Eq. (8.116), the linear strains are _ d2wp _ d2w0 _ d2w0 *** ~ ~Z1*9 Bxy " "'W 6yy ~ ~ZT^' exz = 0, eyz = 0, ezz = 0. (8.117) Note that the transverse strains (exz, syzt £zz) are identically zero in the classical plate theory. Hamilton's principle can be used in two different forms: one way is to construct the Lagrangian functional L = (U + V — K) and use Hamilton's principle (the dynamic version of the principle of "minimum" potential energy): 0 = S Ldt=S (U + V-K)dtt (8.118a) Jt[ Jt] where (U9 V, K) are the total strain energy, potential energy due to applied loads, and kinetic energy, respectively. The other way is to construct only the virtual Lagrangian SL = SU -h <5 V — 8K and use Hamilton's principle in the form (the dynamic version
336 THEORY AND ANALYSIS OF PLATES of the principle of virtual displacements): 0= [2&Ldt= f2(SU +SV-SK)dt. (8.118b) The first way makes use of the constitutive equations to write U and the assumption that the forces are conservative to write V, much the same way as in (he principle of minimum total potential energy. The second way is general and does not require the use of constitutive relations or the assumption that the forces are derivable from a potential. Here we use the latter, but also give the expressions for (U, V, K) for completeness. The virtual strain energy SU is given by 8U = I I ((fxx&Gxx + cTyySeyy + 2aXySsxy) dzclxcly JQq J-h/2 f ( 328wq c)28wp ^A/f 328w0\ j (8.119) where (Mxx, MYyy Mxy) are the moments per unit length (see Fig. 8.12): r/i/2 -A/2 Mxx My, M xy -f Jyy 7*y\ zdz. (8.120) To write U, we assume linear elastic behavior of the plate material and write [see Eqs. (3.40) and (3.41)] Jyy 7*y J 1-v2 J v 0 v 1 0 1 - v 0 0 Sxx £v yy [2Sxy\ (8.121) Figure 8.12 Definitions of moments and shear forces.
CLASSICAL PLATE THEORY 337 for an isotropic material and < <yxx) Oyy \ = 0Xy) [fill 2l2 0 for an orthotopic material, where On Ei 1 - 1>12V21 266 = G\2, 012 = E2 VU = V(2 —■ 2l2 0 222 0 0 266 Vi2^2 1 - Vl2V2l' t_yy (8.122a) l2s.v3J 222 = £2 ) — V12V21 ' (8.122b) Then substituting Eq. (8.117) into Eq. (8.122a), and the result into Eq. (8.120), gives Mxx = -(d\ M, d2w0 dx2 d2w0 + Dl2 ryy = -(D^ + D" cHwo dy 92 9. »2W0\ dx2 j ' (8.123a) Mxy = -2D66- d2wo dxdy' where A3 121 Dij = ^Qi.h h = plate thickness, (8.123b) For an isotropic plate we have (£j = £2 = E9 v\2 = 1^21 = v, and G12 = G = £/[2(l + v)]): DH=Z)22 = £> = £/*3 12(1 -v2Y The strain energy of an orthotropic plate is Dn = vD, 2D66 = (l-v)D. (8.124) = 2 / / [°u & + D22^' + 2Z),2^ve^ + 4D<* *? J dzdxdy dxdy (8.125)
338 THEORY AND ANALYSIS OF PLATES The virtual work done by the applied distributed load q(x, y) on the top surface, z = h/2, the applied transverse edge force V,u the applied normal edge moment M,m on boundary V2 (a portion of the total boundary T of Qo), and the applied in-plane compressive and shear forces (Nxx, Nyy, Nxy) is -H/J*-~+^ Swq dSwo 17 + / <:/(a-, y)Swo rf*<ty + / ( V»Swo - M„„ —-^ J ds J (8.126) The potential energy due to applied loads is - [ q(x,y)wQdxdy+f \Vnwo - Mnn^j ds\ . (8.127) dxdy + The virtual kinetic energy is given by /■A/2 r rh/2 8K — I I p (uSu + vSv + wSw) dz dxdy JQii J-h/2 JaoJ-h/2 LV a* / V 9* / V <W V 3)' / + wo^wo qWqSwq + h dzdxdy Wq dWQ 38wq +-si-sr)]dxdy' (8.128) where p is the mass density and the superposed dot on a variable indicates time derivative, tio = dwo/dt, and (/o, .fc) are the mass moments of inertia: The kinetic energy is K = {f yo^^ + h m1+m' dxdy. (8.130)
CLASSICAL PLATE THEORY 339 Now we have all the elements in place to apply Hamilton's principle (8.118b). We have r (%w - Iqw08wo + h I — 32Swq ' 3y2 + 2M. 328w0\ 3xdv / T . 0 . r /3wq3Swo 3wq38wo v ) f- 3wo 38wo , a dwo 38wq y \ dx dy dy dx )\ J / / qSwodxdy- I ivnSwo-Mlin——jds /fi Una Jr2 \ 9re Integrating terms by parts to relieve 8wq of any differentiation, we obtain dt, (8.131a) 0 -f|jL[-pS 92M,v.v , 92MW 92M0,\ , /92w0 , 92u;0\ # /- dtun s. 3wq\ 3 (* 3wq - 3wq\ +TX (""17 + N-Hi) + Yy (""17 + N»-W) ~ * 8wq dxdy I f dSwo , ., 38w0 38wo 38w0 % [M»n'-9T + M»'n>'-dy- + """'ST + M^^x~ 8x-n>+ By 3Mxy \ \ dx 3y h (i7'Ij+17"'j+ r-vi7+^17 J "* ' (""17 + "w'iv) "*] 5u,°ds ~ /, (^"5u,° ~ ^" 9<5u>o On ds > rff, (8.131b) where all tenns evaluated at t = t\ and t = t2 are zero (by assumption) and not included in the above statement. The Euler-Lagrange equation is clearly /92M,v,v „ 92M,VJ, 92MWA 9x2 + 2- + + 9y9;t 9y2 /
340 THEORY AND ANALYSIS OF PLATES Inspection of the boundary terms in (8.131b) indicates that w;o> 9wo/3x, and dwo/dy are the primary variables, and (a,+ *£)., + (fe +*£).,. md Mxxnx + MXyfiy, Mxynx + Myyny, (8.133a) are the secondary variables of the theory. Here Qx and Qy are defined as V dx dyJ (8.133b) rt _ 3MV.V 8Mvy c-rs=~ar+"ar* _<WV>. 9Afw, In order to cast the boundary conditions on an edge arbitrarily oriented in the xy- plane, we convert the derivatives with respect to x and y to those in terms of the normal and tangential coordinates (n, s). If the unit outward normal vector ii is oriented at an angle 0 counterclockwise from the positive jc-axis, then its direction cosines are nx = cos 0 and ny = sin 0, Hence the transformation between the coordinate system (77, s) and (x, y) is given by ev = cos0 e„ — sin 0 es = nx en — ;?v e^, « (8.134) ey = sin 0 e„ + cos 0 es = ny e„ 4- nx es. Hence the normal and tangential derivatives (dwo/dn, Bwo/ds) are related to the derivatives (dwo/Bx, dwo/dy) by ^— = nx~- ny-—, -— = ny-— + «*-r-. (8.135) dx dn ds dy dn ds Using Eq. (8.135), we can rewrite the boundary expressions in (8.131b) as 38wq °=f\ (v*nx + Vyny)8wo - [Mxxn] + 2Mxynxny + Myyn]){- - [(Myy - Mxx)nxny + Mxy(n2x - n2yj\ ^ J ds ( (* * 38wo\ - / I V„8wq - Mn„ -^- I <fa 3/2
CLASSICAL PLATE THEORY 341 dSwo = j> \\Vxnx + Vyiiy + ^- - Va\ 8w0 + (Mfm - Mfm) (9/2 ds. (8.136) Here, Swq = 0 and (dSwo/dn) = 0 on F\ where wq and (dwo/dn) are specified, and i 3wq Mw„ = Mv.vrtJ + 2Mxynxny + Myyii2y, (8.137a) (8.137b) K, = e.-(^Y— + *,„_ ) + /2—f (8.138a) A#,w = (M;7 - Mxx) nxny + M^. - n*). (8.138b) Note that the expression involving BSwo/ds was integrated by parts to give 3Mns Mtts—-ds = r ds Jr ds 8wq ds - [MnsSwo\r , (8.139) and the term [Mns8wo]r is set to zero since the end points of a closed smooth curve coincide. For plates with corners (i.e., for polygonal plates), concentrated forces of magnitude Fc = -2Af„, (8.140) will be produced at the corners. The factor of 2 appears because Mns from two sides of the corner arc added there. Finally, the natural boundary conditions are Vxnx + Vyny + —^-Vn=Q% as Mtt. Mnn =0 on To. (8.141) The first of the two boundary conditions is known as the Kirchhqfffree-edge condition. Thus, the effective shear force is Vn = Vxnx + Vyny + ^. (8.142) Equation (8.132) can be expressed in terms of the transverse deflection with the help of moment-deflection relations (8.123a). We have d*4 dxzdyz dy4 T* {""IT + N"-W) + 5? (""IT + N»! WQ dy (8.143a)
342 THEORY AND ANALYSIS OF PLATES for orthotropic plates, and a /* 3wq ^ dwo\ , .. T (s2m d2m\ /01,„.x + Yy (N"l* + N» VJ = * - l0W° + h \J* + IF) ' <8J43b) for isotropic plates, where D = Eh3/[\2(l - v2)], 8.2.4 Navier Solutions of Rectangular Plates In this section we briefly discuss Navier solutions of isotropic rectangular plates for bending, natural vibration, buckling, and transient response. Navier solutions can be developed for simply supported rectangular plates in the following cases: (1) bending solutions under arbitrary transverse load q(x, y); (2) natural frequencies; (3) buckling loads under in-plane biaxial compressive loads Nyy = yNxx — yNo; and (4) spatial part of the solution for the transient case. Keeping the scope of this book in mind, only a brief discussion of analytical solutions is presented. For additional information, the reader may consult the books on plates [8,22,24-29], especially those by Reddy [8] and Timoshenko and Woinowsky-Krieger [24]. The simply supported boundary conditions for a rectangular plate, for the choice of coordinate system shown in Fig. 8.13, are u>o(0, y, r) = 0, wo(a9 y, 0 = 0, u>o(*, 0, t) = 0, w0(x, 6, t) = 0, (8.144a) Mxx(0, y, 0 = 0, Mxx(a, y* 0 - 0, Myy(x, 0, /) = 0, Myy(xtbt 0=0. In Navier's method the displacement wq is expanded in double sine series with unknown coefficients: OO OQ wq(x> y, 0 = Yl X W,,m(/) sin a,,,x sin £„)'>' (8.145a) 77=1 m=1 where a,n = , A, = -p. (8.145b) a b
CLASSICAL PLATE THEORY 343 WO=Myy = 0 Wq=Mxx-Q . t l WQ=Myy = 0 7 j fe» b Figure 8.13 Geometry and coordinate system for a rectangular plate. Here Wmn denote the coefficients (that depend on time t) to be determined such that Eq. (8.143b) is satisfied everywhere in the domain of the plate for all t > 0. The choice of expansion (8.145a) is dictated by the fact that the double sine series satisfies the simply supported boundary conditions in (8.144a5b). Substitution of the expansion (8.145a) into the governing equation (8.143a) or (8.143b) shows that the mechanical load q may also be expanded in double sine series: <?(*, y, 0 = E E *»"i(0 sin«n»*sin P*y> 4 fb fa Qmn (0 = -r / / <? (■*, y, 0 sin am x sin p„ y dxdy. ab Jo Jo (8.146a) (8.146b) The coefficients qmn for various types of loads can be calculated using Eq. (8.146b) (see Table 8.3 for typical loads; see Reddy [7,8]). Substituting Eqs. (8.145a) and (8.146a) into Eq. (8.143b), we obtain 0 = E E (DK + 2a™ti + fi)W™ ' ("I + Y&)N* - qmn + [k + h{*l + A?)] Wllin } sin amx sin #,y (8.147) for any (jc, y) and f > 0. Hence.it follows that [lo + h{cti,+fi)] d2W,m dfi + \p{a2m + /3n2)2 - (a2 + yfi)No] Wmn = qma, (8.148)
344 THEORY AND ANALYSIS OF PLATES Table 8.3 Coefficients in the double sine series expansion of loads in Navier's method Loading Coefficients, qmn Uniform load 16</o Qmn — ~~y nAmn (m, n = 1, 3, 5,...) Hydrostatic load (m = 1,3,5,...) (n = 1,2,3,...) Point load [i.e., Qq at (jt0, y0)] 4Q0 . mnxQ . /77r^o ^mii = ~r sln sin ~T"~ (/w,/i= 1,2,3,...) Line load q(x,y) = qo&(y-y6) 8tfo . niryo Qmn — sin - ixbm ~" b (m = 1,3,5,...) (/? = 1,2,3,...) for every pair of m and n and /; > 0. Equation (8. J 48) can be specialized to static bending, natural vibration, or budding of simply supported rectangular plates. For the transient case, the ordinary differential equation in time, Eq. (8.148), must be solved subject to initial conditions on the deflection wq and velocity wq. We discuss these cases next for isotropic plates. Bending For the static bending response under applied transverse load q(x, y) and in-plane biaxial compressive forces (No, yNo)> the coefficients Wmn (which are constants) in the displacement expansion (8.145a) can be obtained from Eq. (8.148)
CLASSICAL PLATE THEORY 345 by setting time derivative terms to zero: Wm„ = ? ^ = (8.149a) [D(«2+^)2-(«2+^2)JVo] and the static solution becomes 00 OO woC*, 3;) = ^2 ^ ^"^ sin a'»A' sin &'3'' (8.149b) /7=1 m=l Note that the compressive force No has the effect of increasing the deflection (or reduce the deflection when the in-plane forces are tensile). When the in-plane compressive loads are zero, we have wot*, )0 = 77-4 T] V . 2 fw" 2.2 sin awx sin £„y. (8.150b) D7T4 *—' ^—' (mfv2 + nz) where 5 is the plate aspect ratio, s = b/a, and the solution is j 4 00 00 n—[ 777= I The bending moments can be calculated from (s = b/a): 00 00 2 Mxx = D ^ X! IT (//?2,y2 + y/l2) ^"'" sin a'"x sin 0'1?' 71=1 77?= 1 00 00 2 M)7 = D J2 Yl ^i H*2*2 + r2) w»»> sin a'»x sin P*y> <8-] 5 ]) 77 = 1 wi = l 2 00 00 Mxy = -(1 - v)-gTD ^ ^ mnWm„ cosamx cos/3„y. 7?=1 777—1 The shear forces (2a and 2}> can be computed using Qx = -^- + -~- ^ 2^L^ Sxx Wm" cosa'»x sin ^z3;> (8.152a) ^} 7? = i 777=1 3MA,Y 3MYV dx By Qy = ~^~ + -~- = 2222 SyyW>»» sino^cos A,?, (8.152b) 7f=l 711 — 1 where S^ = D{al + vamp*)% Syy = D(val + a,„p*). (8.152c)
346 THEORY AND ANALYSIS OF PLATES The effective shear forces (i.e., reaction forces) Vx and Vy along the simply supported edges x = a and y = b, respectively, can be calculated using Vx(a9y)=Qx + dMxy _ 3MXX dx + 2- dM xy 'dy = J2 S(_')W 5« Wm»sin pny> n=] m=l Vy(X, b) = gj, + BM-. xy dx 23Mxy f dMyy dx + ■ dy (8.153a) = ^^2 ^~^" ^ W"m si*1"'"*- (8.153b) where 5A,V = D[al + (2 - u)of/w^], 5^ = Z>[^ + (2 - v)a* ft,]. (8.153c) Thus, the distribution of the reaction forces along the edges follows a sinusoidal form. Similar expressions hold for Vx(0t y) and Vy(xt 0). In addition to the reactions in Eqs. (8.153a,b) along the edges, the plate experiences concentrated forces at the corners of a rectangular plate due to the twisting moment Mxy per unit length (which has the dimensions of a force). The concentrated force at the corner x = a a.nd y = b is given by Fc = -2Mxy = 2(1 - v)D dxdy -*-)»EE(7)(f)H)-'. (8.154) «=1 ;n=l For the pure bending case considered here, the stresses in a simply supported rectangular plate are given by Oyy a*y J Ez (1 - v2) 1 v V 1 0 0 0 " 0 1-v 2 - [ d2w0 dx2 92i/j0 dy2 2**L n=\m=\ 3x3y Rxx &inamx sin ft, y Ryy sinamjt sinft?y [— Rxy cosamx cos ft, y J (8.155a)
CLASSICAL PLATE THEORY 347 where n2E R™ - b2(i _ ,2) n2E R = mns -ir- -. } h2(\ + v) (mrs2 + wr), (ymV+n2), (8.155b) The maximum stresses occur at (x, yt z) = (a/2, b/2y ±h/2). In the classical plate theory, the transverse stresses (axz, ayt, <rzz) are identically zero when computed from the constitutive equations because the transverse shear strains are zero. However, they can be computed using the 3D stress equilibrium equations for any — h/2 < z < h/2: an = -[Z (d-^L+d-^L)dz + C2{x,y), (8.156) J-h/2 \ <>x (>y ) J_/7/2 V 9x dy ) where the stresses crxx, <rxy9 and <rvv are known from Eq. (8.155a) and C, are functions to be determined using the conditions crxz(x, yt —h/2) = ayz(x9 y, —h/2) = <?zz(x> y> -h/2) = 0. We obtain C, = 0 and — 1 0>z 0yz h2 8 h2\ 8 p cos - mnx ._ nny sm ■ a EL*; n = l /«=1 oo oo v^ v^ „vz . m7r* nny (8.157a) (8.157b) "48 -3 1 + (d: „ . mnx , nny „jn sin sm n=l m=l b ' (8.157c) where S£n = Sl3Wmn, S& = 523 Wmn, S«, = 533 W,„„, (8.158a)
348 THEORY AND ANALYSIS OF PLATES with Sjj defined by 523 = (TT^) (^ + *« &)• (8-158b) Note that <ta.2 and cr^ are zero and azz = q at the top surface of the plate (z = h/2). The transverse shear stress axz is the maximum at (x, y, z) = (0, 6/2, 0), o^ is the maximum at (x, y, z) = (a/2, 0, 0), and the transverse normal stress azz is the maximum at (x, y9 z) = (a/2, Z?/2, h/2), Example 8.9 For an isotropic rectangular plate subjected to a uniformly distributed load, the deflection is given by n=i,3,... «i=l.3f... The maximum of wq and Mxx are given by (.? = h/a) l6qob4 « ~ (_1)((»»+»)/2)-l wo(*,y) = —-T- > > — 2 2 2 , sin sin—-. (8.159) D7T" ^^ *—' mn{misi+niy a b «W =<«o r.r = n 6 >^ >, —, 22± 2x2. (8.160a) x ' h=1,3,... m=l,3,... (Mxx)max = Mxx(^ = ^ ± ± (-,)(0»+»)/2)-. \ ' m — I ^ in- I 1 ??=:!,3.... ro=L3,... (m2J2 + W2)2 X / 22x 2^2' (8.160b) mn(mzsz + rcz)z For a square plate (£> = a) wc have lfifloo4 v^ ^ (-D(('"+")/2)-' „..„,, n=1,3,... m=l,3,... I6q0b2 i?=l,3,... wi=1,3,.. A one-term solution is given by (v = 0.3) (j*„w = ^£ E E (-o^^1 (,?;2trX- <8161b> uw = -^s-=0.00416—, (MxxW = 0.05338</0a2, (ff«)m« = 0.3203^-.
CLASSICAL PLATE THEORY 349 The deflection is about 2.4% in error compared to the solution obtained with m, n = 1, 3,..., 9. Thus, the series in Eq. (8.161a) converges rapidly. The expressions for bending moments do not converge as rapidly. For m9n = 1, 3,..., 29, we obtain 4 wmax = 0.004062^-, (8, J 62a) 2 (MV,Y Wv = 0.04789 q0a2, (<rXJ[)max = 0.2816^-. (8.162b) For an isotropic rectangular plate under a point load Qo at (xq> j>o)> the deflection is given by . mnxQ , nnyo , , 4Go&2* ^ ^ sin—— *« — . mnx . nny wo(x, y) = _ , ) } —t r ^ =-^— sin sin——. Dn4 *-? 1-i (m2s2 + n2)2 a b ;i=l,3,... w~l,3,... (8.163) The center deflection when the load is applied at the center is given by W{) (a b\ 4Q0b2s ~ ~ 1 /i=l,3,... m=l,3,... In the case of a square plate, the center deflection becomes 4<2o"z v^ v* l ^ Ti -Ti (m2+n2y »=l,3.... m-1,3,... The first term of the series yields (v = 0,3) Onct2 Onct2 «W = 0.01027^- = 0.1121^3-, {Mxx)max = O.LWQoa, (a,,W = 0.7903^. Taking the first four terms (i.e., m,n= 1,3) of the series, we obtain wmax = 0.01121^— = 0.1225^, D Eh5 Qoa (Mxx)max = 0.1990o«, (ff«W = 1.194 h2 ' The deflection is about 3.4% in error compared to the solution 0.0116 (Qq<i2/D) obtained using m, n = 1,3,..., 19.
350 THEORY AND ANALYSIS OF PLATES Table 8.4 Transverse deflections and stresses in Isotropic (v = 0.3) square and rectangular plates subjected to various types of loads Load SL UL(19)a HL(I9) PL (29) SL UL(19) HL(19) PL (29) w 0.0280 0.0444 0.0222 0.1266 0.0908 0.1336 0.0668 0.1845 <rXx 0.1976 0.2873 0.1436 2.4350 dyy Square Plates 0 0.1976 0.2873 0.1436 2.4350 °*y = b/a = 0.1064 0.1946 0.0775 0.3658 Rectangular Plates (s = b/a 0.5088 0.7130 0.3565 2.3523 0.2024 0.2433 0.1217 1.8828 0.1149 0.2830 0.0579 0.0566 *XZ 1) 0.2387 0.4909 0.2455 1.0010 = 3) 0.4297 0.7221 0.3610 0,9032 °yz 0.2387 0.4909 0.1353 1.0010 0.1432 0.5110 0.0636 0.2257 aThe number in parentheses denotes the maximum values of ih and n used to evaluate the scries. Table 8.4 contains the nondimensionalized maximum transverse deflections and stresses of square and rectangular plates under various types of loads. The transverse deflection and stresses are nondimensionalized as follows: /Eh3\ ( h2 \ w = u^O, 0) -J— ; 5XX = crxx(a/2, b/2, h/2) -5— ) , \crgoJ \^0/ dyy = (Xyyia/2, b/2, h/2) (-£- J; axy = axy{a, b, -h/2) (-^— J , *xz = o>z(0, b/2, 0) (—\ ayz = *yZ(a/2% 0, 0) (—) . (8. \aqoJ \aqoJ 166) The loads considered are sinusoidal (SL), uniform (UL), hydrostatic (HL), or central point load (PL). Since convergence for derivatives of a function is slower than the function itself, the convergence is slower for stresses, which are calculated using the derivatives of the deflection. In the case of the point load, convergence for stresses will not be reached due to the stress singularity at the center of the plate. Natural Vibration For natural vibration, the deflection is assumed to be periodic in time: CO OO «*(*, * 0 = £ £ <»sin ^sin njv e''a","' (8-167a) m=\ n=l or Wmn(t) = W«ei"»«>ty (8.167b)
CLASSICAL PLATE THEORY 351 where / = V—1 and o),mi is the frequency of natural vibration associated with the (m, n)th mode shape (W®n is the amplitude of vibration): . mnx njvy sin sin—-1-. a b (8,168) Substituting (8.167b) into Eq. (8.148), we obtain, for nonzero W^„, the result *>{*l + ft?)2 - (a* + Ytf)No - a£fl[/o + K + ft?)/2] = 0, (8.169) where am = inn/a and fin — rnt/b. Solving Eq. (8.169) for the natural frequency, we obtain 2 * COmii = — /o*2 ,2\2 {m2s2 + n2) - No(m2s2 + yn2) where 70 = /o + /a /mn\2 {nn\i~\ (v)+(t)J j = -, (8.170) (8.171) For different values of m and /*, there corresponds a unique frequency co,7U1 and mode shape given by Eq. (8.168). The in-plane compressive force as well as the rotatory (or rotary) inertia h have the effect of reducing the magnitude of the frequency of vibration and their relative effect depends on m and /z, and the plate thickness-to-side ratio h/b. For most plates with h/b < 0.1, the rotary inertia may be neglected. The smallest value of coJHn is called the fundamental frequency. When the rotatory inertia h is neglected, the frequency of a rectangular isotropic plate without in-plane forces No reduces to (/q = ph) D = ia*h:(ms +»)> _b a' b2]j ph For a square isotropic plate without rotary inertia, the frequency is (8.172) to, a1 V ph v ' (8.173) The fundamental frequency of a square isotropic plate is 2;r2 aA (8.174) Table 8.5 contains nondimensionalized frequencies comn of isotropic (v = 0.25) plates for (m, n) = (1, 1), (1,2), and (2,1). The effect of rotary inertia is negligible for plates with h/b < 0.1.
352 THEORY AND ANALYSIS OF PLATES plates' b/a 0.5 1.0 1.5 2.0 2.5 3.0 1 fanm - w/o 1.250 2.000 3.250 5.000 7.250 i 0.000 <*>mn(hr W|| 0.01 1.250 2.000 3,250 4.999 7.248 9.996 tx*>JphlOl 0.1 1.249 1.998 3.246 4.990 7.228 9.959 w/o 4.250 5.000 6.250 8.000 10.250 13.000 o>\2 0.01 4.249 4.999 6.248 7.997 10.246 12.993 0.1 4.243 4.990 6.234 7.974 10.207 12.931 w/o 2.000 5.000 10.000 17.000 26.000 37.000 0)2\ 0.01 2.000 4,999 9.996 16.988 25.972 36.944 0.1 1.998 4.990 9.959 16.882 25.726 36.450 aw/o = without rotary inertia; the second and third columns contain frequencies when the rotary inertia is included for h/b — 0.01 and 0.1, respectively. yN„ mmmmm 1 YiV0 Figure 8.14 Biaxial compression of a rectangular plate (Nxx = N$ and N§y = y NQ). Buckling Analysis For buckling of rectangular plates under in-plane biaxial compressive loads (see Fig. 8.14), Eq. (8.148) reduces, after omitting the time derivative and load terms, to [d(o& + fif - (o& + yP$)N0] Wm„ = 0, which gives, for nonzero Wmn, the result ,2x2 /Vo(m, ri) = nlD(sJ'ml + nl) b2 s2m2 -f yn2 (8.175) where s is the plate aspect ratio s = (b/a). For each choice of m and n there corresponds a unique value of Nq. The critical buckling load is the smallest of No(m, n). For a given plate this value is dictated by a particular combination of the values of m and n, the value of y, plate geometry, and material properties. Next we present critical buckling loads for various cases.
CLASSICAL PLATE THEORY 353 Biaxial Compression of a Plate For a rectangular isotropic plate under biaxial compression (y = 1), the buckling load is No(m, n) = -^r{m2s2 + n2), (8.176) where s is the plate aspect ratio s = bfa. Clearly, the critical buckling load occurs at m. = n — i and it is equal to Ncr = (l+S2) 7tZD s = (8.177a) and for a square plate it reduces to Ncr - 2n2D (8.177b) Biaxial Loading of a Plate When the edges x = 0, a of a square plate are subjected to compressive load Nxx — No and the edges y = 0, b are subjected to tensile load Nyy = —yNo, Eq. (8.175) becomes n2D No(m9n) = —=- (m2+/z2)2 m2 — yn2 (8.178) when yn2 < ;?z2. The minimum buckling load occurs for n = 1: \ bz / 77izi'z — y In theory, the minimum of No(m, 1) occurs when m2s2 = 1 + 2y. For a square plate with y = 0.5, we find 8tt2Z> tfo(lJ) = —2-. W0(2, 1)= 7.1429 7T2D = Mr. (8.179b) Uniaxial Compression of a Rectangular Plate When a rectangular plate is subjected to uniform compressive load Wo on edges x = 0 and * = a, i.e., when y — 0, the buckling load is given by „ , , n2a2D (m2 n2\ N0(m, n) = 5— -r- + -5- , Ar , 1N n2D ( I a2\2 N0(m, 1) = —5- (« + —7T • (8.180) (8.181)
354 THEORY AND ANALYSIS OF PLATES Table 8.6 Effect of plate aspect ratio on the nondimensionalized buckling loads N of simply supported (SSSS) rectangular plates under uniform axial compression (y — 0) and biaxial compression (y = 1) a/b 0.5 1.0 1.5 N 6.250 4.000 4.340<2- Y i)a = 0 a/b 2.0 2.5 3.0 N 4.000(2'1} 4.134'3'1) 4.000(3U) a/b 0.5 1.0 1.5 Y = N 5.000 2.000 1.444 1 a/b 2.0 2.5 3.0 N 1.250 1.160 1.111 a Denotes mode numbers (wi, /?) at which the critical buckling load occurred; (/??, n) = (1.1) for ail other cases. For a given aspect ratio, two different modes, m \ and 777.2, will have the same buckling load when +Jm.\m2 = ajb. In particular, the point of intersection of curves m and m + 1 occurs for aspect ratios % = 72, VS, -s/12, V20,..., Vm2 + m. Thus, there is a mode change at these aspect ratios from m half-waves to m + 1 half-waves. Putting m = 1 in Eq. (8.181), we find n2D (a b\2 *—-?-(»+:)- (8-I82) For a square plate we obtain 4tt2D Table 8.6 shows the effect of plate avspect ratio and modulus ratio (orthotropy) on the critical buckling loads N = Ncrb2/(7T2D) of rectangular isotropic (v = 0.3) plates under uniform axial compression (y = 0) and biaxial compression (y = 1). In all cases, the critical buckling mode is (m, n) = (1, 1), except as indicated. Transient Analysis The determination of the solution wq(x, y, t) of Eq. (8.143a) or (8.143b) for all times t > 0 under an applied load q(x, v, /) and known initial conditions mix, J>» 0) = do(xt y), —— (*, y, 0) = vq(x, y) for all k and v, (8.184) at where do and do are the initial displacement and velocity, respectively, is termed the transient response. The transient response of simply supported rectangular plates can be determined by assuming a solution of the form in Eq. (8.145a) and determining
CLASSICAL PLATE THEORY 355 the solution Wmn (/) of the ordinary differential equation in (8.148). It is necessary to expand the nonzero initial displacement and velocity field in double sine series: 00 OO do(x, y) = ^2 X! Dmn sina'nXsin P»y> (8.185a) OO 00 vq(x, )')=^]^ Vmn sin amx sin pny, (8.185b) H=l 771=1 where am = mx/a, pn = «7r/6, and D„m and Vm„ are given by /),„„ = — / / d(jt,;y)sinam;tsin/5,?)'d*<tf;y. (8.186a) flfr Jo Jo 4 r'7 /•* Vmn = — / / v(*, )0 sin a,„x sin /3„>* djedy. (8.186b) «frJo Jo Equation (8.148) has the general form KmnWm„(t) + M>™—^T = Imnif), (8.187a) where Knm = D(al + pff - NQ(al + ypj), Mmn = 70 + 72(<4 + A?)- (».187b) The second-order differential equation (8.187a) can be solved either exactly or numerically. The numerical solutions are often based on the finite difference methods. To solve Eq. (8.187a) exactly, we first write it in the form ^ + (tF) w™ = T^WO - *„<0. (8.188) dtl \MmnJ Mmn The solution of Eq. (8.188) is given by W,„„(0 = Cx^f + C2ek* + W*f (0. (8.189) where C\ and C2 are constants to be determined using the initial conditions, W&n(t) is the particular solution W£/i(0=/ , NW v r-7—-— ^nif(t)dr, (8.190a) J n(T)r2(T)-ri(r)r2(T) with n(0 = £*ir and r2(0 — eXl\ and A,i and A2 are the roots of the equation A2 + ^=0; Jl, =-»/*, ^2 = «>,i = vCTf/t =/£!!!. (8.190b)
356 THEORY AND ANALYSIS OF PLATES The solution becomes Wmn(t) = Acqs fit + B sin fJLt + W,Jn(0. (8.191a) WLif) = jf- (^ f e-^q^it) dx - e~illt J* e^xqmn{x) dz\ . (8.191b) Once the load distribution is known, the solution can be determined from Eq. (8.189). For a step loading, qm„(t) = q%nH(t\ where H(t) denotes the Heaviside step function, Eq. (8.191a) takes the form W„w(t) = Amn cos fit + Bmn sin fit + —$°m. (8.192a) Using the initial conditions (8.184), we obtain Amn = Dmn - ^-q°n. Bma = ^2L. (8.192b) Thus the final solution is given by w0(x, y, t) = Y Y) Dmn cos pa + — sin (it + ^-(I - cos fit) x sin amx sin #,>*• (8.193) The coefficients qfnn were given in Table 8.3 for various types of load distributions. The same holds for Dmn and VmfJ. The exact solution of the differential equation (8.188) can also be obtained using the Laplace transform method. 8.2.5 Levy Solutions of Rectangular Plates The deflection of a rectangular plate with two opposite edges (x = 0, a) simply supported and the other two edges (—b/2 < y < b/2) having arbitrary boundary conditions (see Fig. 8.15) can be represented in the form (sec the semidiscretization method of Examples 7.17-7.19): OO u>o(x, y) = J2 W»*W sma,,,*, (8.194) i which satisfies the simply supported boundary conditions on edges x = 0 andx = a. Substituting the expansion (8.194) into the equilibrium equation of the plate, < £+>&+&-
CLASSICAL PLATE THEORY 357 simply supported u>o=Mxx = 0 A 6/2 T i b/2 T Figure 8.15 Geometry and coordinate system for a rectangular plate with sides x = o, a simply supported. I II If If L T a 1 1 1 i »- t» we obtain an ordinary differential equation for Wm(y): D(a*,Wm-2al-jv2 l2Wm d4Wni dy2 4- cly4 J = q»i, where qm is defined by 2 /"' 3m00 = - I Q(x. y) sm or,,,a: d*, « Jo mjr Otm = (8.196) (8.197) Tabic 8.7 contains the coefficients q,„ for various types of loads. The solution of Eq. (8.196) may be determined exactly or numerically. Here we discuss its exact solution. One may also solve it using the Ritz method (see Rcddy [8]). The solution Wm consists of the homogeneous solution W^ and particular solution Wm. The homogeneous solution of Eq. (8.196) is W*(y) = (Am + Bmy) cosh amy + (Cm + D,„y) sinha„,y. The particular solution W,f2 is given by the solution of «mWm 2a,,, dy2 + ^ - D. *m rvm Assuming solution of Eq. (8.199) in the form oo WZ(y) = Y,WmnS™P„y, fin = n=\ expanding the load q,„ also in the same form: oo nn T' (8.198) (8.199) (8.200) (8.201)
358 THEORY AND ANALYSIS OF PLATES Table 8.7 Coefficients in the single trigonometric series expansion of loads in the Levy method Loading Coefficients, qn (x) Uniform load <? = Qn ■■ (n~- m nn = 1,3,5,...) Hydrostatic load q(x qn ■- (n = Poii Qo v y = 2qo( \y1+l nn = 1,2,3,...) it load at (xq, )>o) 22o., v . nnyo qn = -—S(x - x0) sin —— D O (« = 1,2,3,...) Line load q(xty)=qQd(y-yo) 2^o . nnyo qn b b (n = I,2,3,...) and substituting both into Eq. (8.199), we obtain u>mn (8.202) The particular solution becomes W;n{y)^ >-r— sin—-, »=i (8.203)
CLASSICAL PLATE THEORY 359 and the complete solution becomes 00 u>o(*. v) = ^ [(Aw + Bmy) ooshamy + (Cm + Dmy) sinha,„y + W,£] smamx, (8.204) i»=l where W,£ is given by Eq. (8.203). The constants Am, B,fli Cw, and Dm must be determined for the particular set of boundary conditions on edges y = ±b/2. Example 8.10 Consider a rectangular plate simply supported on all four edges and subjected to distributed bending moments along edges y = ±b/2 (see Fig. 8.16): J to = }^ Mm sm <*,„*, am , w=1 with the coefficients Mm loiown from Mn _2 fa a Jo f(x)sinamx dx. (8.205) (8.206) We wish to determine the deflections and bending moments throughout the plate using the Levy solution procedure. The general solution (8.204) for rectangular plates with simply supported edges x = 0, a is valid here except that the distributed load is zero: q = 0 (hence, the particular solution is zero). The boundary conditions on edges y = ±fe/2 are wo = 0, Myy = f(x). (8.207a) Since wq = 0 on y — ±b/2, the condition Myy = / on y = ±b/2 reduces to -*>$? =/w. Sy2 (8.207b) f(x)=Myy(x,~b/2) bj 2 FS^S^^S=S^^ i!mm 1 f(x)=Myy(x,b/2) Figure 8.16 A simply supported rectangular plate with distributed moments along edges y=±b/2.
360 THEORY AND ANALYSIS OF PLATES Due to the symmetry of the problem about the x axis, woix, y) must be an even function of y. This implies that Bm — Cm = 0, and Eq. (8.204) becomes oo wq(x, y) = ^^ (Am cosh a,„y + Dmy sinhamy) sina,„x. (8.208) Using the boundary conditions (8.207a)? we obtain a b n i u - n ClM'» - b ,o iAm Am = - « Dm tanh am, /)„, = - ——, am = am -. (8.209) 2 2m7rL>coshar„, 2 The deflection becomes u>o(*, 3?) = ^—pr T2 M>"~- rV- ( t tanh<5,„ cosha,„)> - >' sinha,7,)> J. 2^D ^ mcosha,,, \2 / (8.210) The distribution of bending moments for the problem can be computed using the definitions in Eq. (8.123a). Nondimensional deflections and bending moments of simply supported rectangular plates with uniformly distributed moments of intensity / = A/o (or M,„ = 4Mo/mn9 m = I, 3, 5,...) are presented in Table 8.8 for various aspect ratios. The results were computed using m — l, 3...., 9. The nondimensionalizations used are as follows: b D - l For - < I: w = Wo(fl/2,0)—-x, M = Af(a/2,0) —, « y M; (8.2H) For - > l: w = wQ(a/2,0)- 7, M = Af (a/2, 0)—-. a Mqci1 Mq Example 8.11 Consider a rectangular plate with edges x = 0, a simply supported and edges y = ±b/2 clamped, and subjected to distributed transverse load (see Fig. 8.17). The clamped boundary conditions on edges y — ±b/2 are u/o = 0, ^=0. (8.212) dy Table 8.8 Nondimensionalized center deflections and bending moments of simply supported plates with applied edge moment along y= ±b/2 Variable b/a -> 0.5 0.75 1.0 1.5 2.0 3.0 w 0.0965 0.0620 0.0368 0.0280 0.0174 0.0055 Mxx 0.3873 0.4240 0.3938 0.2635 0.1530 0.0446 Myy 0.7701 0.4764 0.2562 0.0465 -0.0103 -0.0148
CLASSICAL PLATE THEORY 361 '///////////////////////, b/T b/T V//////////////////////, Figure 8.17 Geometry and coordinate system for a rectangular plate with sides x = 0, a simply supported and y = ±b/2 clamped. The general solution is given by CO / OO \ wu(*> y) = ^2 ( Am cosh<xmy + Dmy sinhar,„.y + ^ Wm„ sin pny J sine*,,,*, (8.213) m=l «=1 and its derivative is CO ~T"r = 2^ ( a'"A»' sinhamy + A» smhawy + Dmamy coshct„ m = \ oo . + X! A' ^""r C0S #^ ) Sin a'«*' Boundary conditions inEq. (8,212) give Am = - - A« tanh am - Qm, An = " ^ r^ i 2 2m7r D cosh am (8.214) fim = qm cosh <5„ §« =£(-1)" H^wm 11=1 M,„ = am -f (1 — am tanham) tanh <£„,, _ mn rin b a b 2 (8.215) where Wm„ is defined in Eq. (8.202): Qmn Dte+ftf and <5fom are defined in Table 8.3 for various types of loads. Table 8.9 contains nondimensionalized deflections and bending moments, w = w0( r)xi02, Af = M( =■ ) x 10, (8.216)
362 THEORY AND ANALYSIS OF PLATES Table 8.9 Maximum nondimensional deflections and bending moments in isotropic (v - 0.3) rectangular plates with edges x = 0, a simply supported and edges y = ±b/2 clamped, and subjected to distributed loads Variable w(a/2, 0) Mxx {a/2,0) Myy (Cl/2,Q) -Mxx(a/2,b/2) -Myy(af2,h/2) w(a/2, 0) Afc, (a/2,0) Myy (a/2,0) -Mxx(a/2,b/2) -Myy(a/2,b/2) w(3a/4t0) Mxx(3a/4J?/2) MyyQafab/2) -Mxx(a/2,3b/4) -Myy(a/2,3b/4) 1=3 1.1681 1.1442 0.4214 0.3740 1.2467 0.5841 0.5721 0.2107 0.1870 0.6233 0.4368 0.5085 0.1803 0.1635 0.5449 1=2 0.8445 0.8693 0.4738 0.3574 1.1915 t = « Uniform i 0.1917 0.2445 0.3326 0.2097 0.6990 a __ 3 T?~ 2 \oad, q = tfo 0.2476 0.1794 0.4067 0.2470 0.8233 l=* 0.2612 0.1441 0.4214 0.2535 0.8451 Hydrostatic load, q = qo(x/a) 0.4222 0.4346 0.2369 0.1787 0.5957 0.3218 0.4097 0.1997 0.1576 0.5254 0.0959 0.1222 0.1663 0.1048 0.3495 0.0841 0.1640 0.1562 0.1042 0.3475 0.1238 0.0897 0.2033 0.1235 0.4117 0.1313 0.1665 0.2276 0.1443 0.4810 0.1306 0.0721 0.2107 0.1267 0.4225 0.1614 0.1509 0.2705 0.1670 0.5566 b ~3 0.2619 0.1302 0.4201 0.2525 0.8417 0.1309 0.0651 0.2100 0.1263 0.4209 0.1884 0.1153 0.3055 0.1842 0.6140 w -"»(i^)*'°2' *=M(^)*10' <8-2") for various aspect ratios. The first set of nondimensionalizations is used for b > a (the first three columns of values) and the second set is used for a > b (the last three columns of values). The values are obtained using the first five terms of the series (m = lf3 9). 8.2.6 Variational Solutions: Bending Here we discuss application of the Riiz method to the bending of plates with various boundary conditions. The virtual work statement for an orthotropic rectangular plate is Jq l 32wq d2Swp /32wq32Sw0 d2w0d28w0\ 11 dx2 dx2 + i2\dy2 dx2 + dx2 By2 ) _,Ark d2WQd25w0 , ^ d2w0 d28wp '1 + 4D66^7^7-QmQ^ + #22~^r-i^> qSwo dxdy i( 'dxdy dxdy dy2 dy2 -Mnn —^ + V„8w0 ds, (8.218) an /
CLASSICAL PLATE THEORY 363 where Mnn and V„ arc the applied edge moment and effective shear force, respectively, on the boundary T of the domain Q (the midplane of the plate). We seek an N-parameter Ritz solution in the form and substitute it into Eq. (8.218) to obtain [R][c] = [Fl (8.219) (8.220) where R u r\ d2<t>i d2<t>j (a2^ d24>f ^ d2<t>i a24>i dxdy, (8.221a) Fi" Lq<t>i dxdy+£ (^" ^"n it)ds' (8*22lb) For rectangular plates, it is convenient to express the Ritz approximation the form in N N N N W(x, y) * WN(x, y) = J2 HCV 4>ij(x, V) = £ Y,c,jXfWYj(y), (8.222) ; = I ; = 1 i=\ ./=1 where </>rj(*, )0 is expressed as a tensor product of the one-dimensional functions Xi and Yj of x and }>, respectively. Substituting Eq. (8.222) into Eq. (8.218) (8wo>n = 5it>o = 0 or Mnn = 0, and yrt = 0 on T), we obtain D„(xd2Yid2XpY +d2XiYX d2YA if*pw idxdy cij ~ L LqXpYq dxdy + d2Yi d2Ya = R '{pqWJ)cU Fpi> (8.223)
364 THEORY AND ANALYSIS OF PLATES where pb pa R(pq)(ij) = / / Jo Jo d2X, d2Xp dX;dYjdXpdYq ^l^^-d^^+^-dJ-dy-^Tly- + . d2Yj d2Y„ + D22Xi-^lx, —* dy2 p dy2 ply pa Fpq = qXpYqdxdy. Jo Jo rb pa «+ dx2 dxdy, IjA»7y^) (8.224a) (8.224b) One choice for X; and Yj is to use algebraic polynomials. A second choice is provided by the characteristic equations of beams. Both sets of functions are given here for typical boundary conditions (see Fig. 8.18). The notation CFSF, for example, means that edgt x = 0 is clamped (C), edge x = a is free (F), edge y = 0 is simply supported (S), and edge y = b is free. In the case of characteristic polynomials, the roots A,/ arc to be determined by solving a transcendental equation. The first few values of kja are given in Table 8.10, which is the same as Table 4.5.1 of Reddy[8]. CFSF CCCC 'L— x I SSSS ' I Figure 8.18 Rectangular plates with various boundary conditions.
CLASSICAL PLATE THEORY 365 Table 8.10 Values of the constants and eigenvalues for natural vibration of plate strips with various boundary conditions (x£ s *>n'o/D = (<Wa)4)« The classical plate theory without rotary inertia is used End Conditions at Characteristic Equation x = 0 and x = a and Values of e„ = Xna • Hinged-Hinged (S-S) sin en = 0, en = nn © Fixed-Fixed (C-C) cos en cosh e„ - 1 = 0 e„ =4.730,7.853,... • Fixed-Free (C-F) cos en cosh ew -f- 1 = 0 en = 1.875,4.694,... • Free-Free (F-F) cos en cosh en — I = 0 ew =4.730,7.853,... © Hinged-Fixed (S-C) tan en = tanh en e„ =3.927,7.069,... o Hinged-Free (S-F) tan en = tanh en e„ = 3.927, 7.069,... CCSS Plates Algebraic Polynomials w-(D'-gr (8.225) Characteristic Polynomials X-,(x) = sin A/a' — sinh XjX + a,- (cosh A/a — cos A,-*), . /27T^ (8.226a) ^00 = sin—, where A,- are the roots of the characteristic equation cos Xici cosh A/a - 1 = 0, (8.226b) and a{ are defined by sinh Xjci — sin Xjci ,„ ~~^ . oti - ——^ J-. (8.226c) cosh A,-a — cos A/a CFSS P/ates Algebraic Polynomials
366 THEORY AND ANALYSIS OF PLATES Characteristic Polynomials Xj (x) = sin XjX — sinh A/x + a,- (cosh XjX — cos Xjx) , . uny (8.228a) y/00 = sin—, sinh A;a -f sin A/a _ „^M v cos A/a cosh A/a -h 1 = 0, a* = —— . (8.228b) cosh A/a + cos Xi a FFSS Plates Algebraic Polynomials (n>1) Characteristic Polynomials Xj(x) = sin X(x + sinh A/ jc — a?/ (cosh A/x 4- cosA/x), . n*)> (8.230a) 7/00 = sin—, sinh A/a — sin A/a ~™i_s cos A/a cosh A/a —1=0, a,- = — —. (8.230b) cosh Xj a — cos A/a SCSS Plates Algebraic Polynomials Characteristic Polynomials Xj (x) = sinh A/a sin A/,v + sin A/a sinh A/ x, . 72TO> (8.232a) F/00 = sin——, b tan A/a - canh A/a = 0. (8.232b) SFSS Plates i Algebraic Polynomials *«-©'■ w(!)'-(ir- <8-»
CLASSICAL PLATE THEORY 367 Characteristic Polynomials Xt (x) = sinh Xxa sin Xjx — sin Xja sinh a,x, v , , . nny (8.234a) 7/00 = sin —, tanX,-fl - tanh A,-a = 0. (8.234b) CFFF Plates (N>1) Algebraic Polynomials XiM ^ ©'+1' Yjiy) = (l)J~]' (8-235) Characteristic Polynomials Xi(x) = sin a,x — sinh a,x 4- a,-(cosh a/x — cos A/x), (8.236a) 7/O0 = sin jjbjy 4- sinh /x^ - #(cosh /^y + cos fijy), cos Xja cosh Xja H- 1 = 0, cos iXjb cosh iijb — 1=0, (8.236b) sinh X{a + sin A./a ^ sinh u/b — sin usb <**• = —T7—7 r-. ^7 = —T^T ^T' (8-236c) cosh A/ a -h cos a/ a cosh /x; & — cos /x / /? CFSFP/ates Algebraic Polynomials Xi(x) = (~)l+i, Yj(y) = (£)' . (8.237) Characteristic Polynomials Xf (x) = sin XiX — sinh A,x + (Xj (cosh a;x — cos A/A') , (8,238a) *0' 00 = sinn l*>j b sin /xy ;y — sin /xy- fc sinh jtx; y, cos A,a coshXtci + 1=0, tan/x/fc - tanh/xyi = 0, (8.238b) sinh a,-a 4- sin A;a a/ = —rr r—■ (8.238c) COShA;tf + COS A/fl CCCC P/ates Algebraic Polynomials
368 THEORY AND ANALYSIS OF PLATES Characteristic Polynomials Xj (x) = sin XiX — sinh X,x + a,- (cosh Xjx — cos A;x), (8.240a) Yj(y) = sinA/j> — sinh A ;)> 4- a /(cosh Xj y — cosAj)>), cos A,- a cosh A/a - 1 = 0, (8.240b) sinh A/a — sin kjct cosh A/a — cos X\a , oti = —— -*- = . ' , . ' ■ (8.240c) cosh XjCi — cos Xta sinh X\a + sin A/a The plate types presented above indicate how one can construct the approximation functions for any combination of fixed, hinged, and free boundary conditions on the four edges of a rectangular plate. The difficult task is to evaluate the integrals of these functions as required in Eq. (8.22 la,b). One may use a symbolic manipulator, such as Mathematica or Maple, to evaluate the integrals. Of course, one may also use trigonometric functions, which have the orthogonality property, for certain boundary conditions. In general, the Ritz method for general rectangular plates with arbitrary boundary conditions is algebraically more complicated than a numerical method, such as the finite element method. Example 8.12 As an example of the application of the Ritz method to rectangular plates with arbitrary boundary conditions, we consider a CCCC plate subjected to distributed transverse load q(x, y). The approximation functions of Eqs. (8.239a,b) or (8.240a-c) may be used in Eq. (8.222) (see Fig. 8.18 for the coordinate system). First we consider the algebraic functions in Eqs. (8.239a,b) with N = 1 and q — q$. Equation (8.223) for this case takes the form 0 = or (^){^D«+4D66{lL){-iL) r 7 4 7 1 49 UDu + aW (°12 + 2D6(,) + ¥D22\ Cn = J'10' The one-parameter Ritz solution becomes qoa — (?) i-(§)T[HS)T 7Dn + 4(D12 + 2D66)*2 + 7D22*4' where s = a/b denotes the plate aspect ratio. The maximum deflection occurs at x = a/2 and y = b/2: 0.00342<?Qfl4 D„ + 0.5714(2?i2 + 2D6t)s2 + D22s4'
CLASSICAL PLATE THEORY 369 Next we use the characteristic functions in Eqs. (8.240a) for N = 1 (X[ = A .13/a andai = 1.0178): v , x . 4J3x . i- 4J3x . ™™/ t 4.73* 4.73. X| (,v) = sm smh f- 1.0178 ( cosh cos r9 /? b \ b b } Evaluating the integrals, we obtain '537 U816 324.829 „ x 537.181a 1 ^ n + gfr (pi2 + 2Arc) + —^T—^22 cH = 0.715c/o^. The maximum deflection is given by (X\ (a/2) = 7i(fe/2) = 1.6164): ).00348<r/0tf4 iyr /a b\ 0, \2' 2J ~ Z)n +0.6047( 6047(Z),2+2D66)^ + 2922^4, (b) For an isotropic square plate, the maximum deflection in Eq. (a) becomes Wn(a/2J?/2) = QMm(qoa*/D), whereas Eq. (b) gives W\\{a/Xb/2) = Q.Q0133{qoa4/D). The "exact" solution (see Timoshenko and Woinowski-Kricger [241, P. 202) is 0,001260y0a4/Z)). Example 8.13 Consider a uniformly loaded, isotropic, rectangular plate (a x b) that has as its edges x = 0 and x = a simply supported, the edge y = 0 clamped, and the edge y = b free (i.e., a SSCF plate; see Fig. 8.19). The geometric boundary conditions of the problem are u;n(0, y) = wo(a, y) = u>o(*, 0) = —- = 0. <*.0) We wish to determine a one-parameter Ritz solution to the problem. We choose an approximation of the form u>o(x, y) ^ W\ (x, y) = C[(f>i (x9 y), 0i = ('-) sin —, \b/ a (a) (b) A t> * '///////////////////////, 1 i 1 J 1 1 1 1 1 1 V II Figure 8.19 A rectangular plate with sides x = 0, a simply supported, side y — 0 clamped, and side y = 6 free.
370 THEORY AND ANALYSIS OF PLATES which satisfies the geometric boundary conditions of the problem. Substituting Eq. (b) into the total potential energy expression [see Eqs. (8.125) and (8.127)], rb rar/*2.„.\2 /*2_\2 92^32 »«-!jU" (£)♦(£)♦ 2v- Wq ♦«-(Sf dx2 dy2 b ra dxdy Jo Jo qowo dxdy, we obtain ^(l) (*) sin V+U^C0STJ -ci / 90 7 sin-—dxdy Jo Jo ^b/ a +2(1-w) ra rb dxdy , 4(1-V)ffz 2 + ,9,a r dy 8(1- -3— / cos —ax / y ay}—cii?o / sin — dx / -r~dy b* Jo a J0 J Jo a Jo b a2b4 Dc\a [^ i7t4b 47V2 4q-v)7z2 U3 + 5a4 ' 3a2b a2b 2abc\qo 3n ' (c) Using the principle of minimum total potential energy, 511 = (dU/dci)Sc] = 0, we obtain c\ — 20q0b4 s = b/a, Dn [60 + 3jt V + 2forV(2 - 3v)]' and the solution is given by 20qob4 /y\2 . nx W] , . Mqptr /yy = nx_ y D7i[60 + 3k4.<:4 +207T2s2(2-3v)]\b) Mn a ' The maximum deflection is Wmax = Wl(£,b) = cu (d) (e) which, for a square plate with v = 0.3, is equal to 0.0111%(qoa*/D).
CLASSICAL PLATE THEORY 371 Example 8.14 Here we solve the clamped plate problem using the semidiscrctc approximation method of Chapter 7 (sec Example 7.19). Consider an isotropic, clamped rectangular plate of dimensions 2a x 2b, and subjected to uniformly distributed load #o« We seek a one-parameter solution of the form 2n2 Wi Or, v) - cx (*)*, (v) - C] (*)(/ - b") (a) For convenience, the origin of the coordinate system is taken at the center of the plate. The approximation W\{x, y) must satisfy the boundary conditions (see Fig. 8.20) jy,=0, dWi/dx^O atjc = ±fl, W] = 0, 9 Wi /dy = 0 at )> = ±ft. (b) (c) The boundary conditions in (c) are satisfied (because of the choice of <j)\) for any x. To satisfy the boundary conditions in (b) for any y, the function c\ (x) is required to satisfy the conditions dc\ c\ = 0, —- =0 atx = ±a. ax The Galerkin integral of Eq. (8.195) over the domain (—h, b) is <t>\ (y) dy, (d) fb r /d4w{ „ s4Wi d4w{\ °-D£[c .2x2, (r-*) <ty- After performing the integration, we obtain or 5 (256b4 d4ci 512b2 d2d 384 ^ 16qo\ _ 7 ^ 315 dx4 105 ~dx* + 1^C1 "~ 15 DJ ~~ rf4ci _ d^ci_ 63 _ 21 tfo (e) I y.w;»wA>Mw;/>M'. Figure 8.20 A rectangular plate with sides x = ±a and y=±b clamped.
372 THEORY AND ANALYSIS OF PLATES where £ = x/b. The homogeneous solution of Eq. (e) can be calculated using the roots of the characteristic equation associated with Eq. (e): d4 - 6d2 + — = 0 -* (rf2),f2 = 3 ± y/9 - (63/2) = 3 ± Z4.7434, where / = v —* • To determine the roots d\ through d<\, we use the polar form (</2)i,2 = x ± /}> = r(cos0 ± 1 sin 0) = re*'"*, where x = 3, y = 4.7434, r = yjx2 + y2 = 5.6125, and 0 = tan"1 (.)'/•*) = 57.69°. Therefore, we have (tf)i_4 = ±Vre±ie/2 = ±Vr (cos ^ ± 1 sin ^ J = ±(2.075 ±H. 143), di =2.075 + /1.143, rf2 = 2.075-i 1.143, d3 = -(2.075 + il.143), d4 = -(2.075 - 11.143). Therefore, the homogeneous solution is given by = A1 cosh a£ cos /?£ 4- A2 cosh a£ sin /:*§ + A3 sinh a% sin ^^ + A4 sinh af cos £§, (f) where or = 2.075, £ — 1.143, and £ = ,r//?. The particular solution is given by C\p 63 \16DJ 24D The general solution of Eq. (e) becomes cj(§) = cift + ci/>- (g) Because of the symmetry of the expected solution about the y-axis, it follows that A2 = A4 = 0. The remaining constants A \ and A3 can be determined using boundary conditions in (d) on c\ at x = d=a. We obtain A1 cosh k\ cos /C2 4- A3 sinh k\ sin ki = — 24D' Ai(— sinh/q cos/c2 cosh/q sin/c2 I \b b } (a 0 \ -f A3 — cosh/ci sin &2 + r sinh A: 1 cos ^2 I = 0, • \b b J
CLASSICAL PLATE THEORY 373 where k\ = aajb and /q = fia/b. Solving these equations, we obtain A _ fj± ( go \ A _ /f2 / go \ Al^MV24Z)>'5 A3~ m\24d)' where a = 2.075, f$ = 1.143, and /zo = P sinh /cj cosh k\ + a sin fe cos /<2, /xi = -(acosh/cj sin/c2 -f/3sinh/q cosfo), (h) fjL2 = a sinh &j cos /c2 — P cosh /ci sin /c2. Thus, the one-parameter solution becomes 1 — cosh -—cos 1 smh — sin —- + 1 (y - /r) . \fi0 b b no b b J J v ' (i) W^y)=UD The maximum deflection occurs at the center of the plate; for a square plate of dimensions 2a x la, the maximum deflection is given by (k\ -> a = 2.075, fe —► 0= 1.143): Wmax = 1*(0, 0) = ?£ (l + ^ = 0.4977^ = 0.0207-^4 24£> V Mo 7 24D Z) For a square plate of dimension a x a, this reduces to (divide the above result with 16): 4 WW =0^)0129^-, which is closer to the "exact" solution 0.00126(go<34/#) than that obtained in Example 8.12. Example 8.15 Much of the discussion presented in Example 8.14 is also valid for a rectangular plate with edges y = ±b clamped and x = ±a simply supported (see Fig. 8.21) and subjected to a uniformly distributed load. The general solution in Eq. (f) is valid with the following boundary conditions on c\: 'J2c . (a) """"-(i^L.-1 These conditions give (a = 2.075 and p = 1.143): A i cosh k\ cos &2 + A3 sinh k\ sin ki — — IAD'
374 THEORY AND ANALYSIS OF PLATES '{'"ft'?""'"""""' V/////////A'//////////A WMSMW Figure 8.21 A rectangular plate with sides x=±a simply supported and sides y=±b clamped. A[ sinh k[ sin k2 — A$ cosh k\ cos ki = The solution of these equations is given by J70_ 247) \ 2*P ) AX=- qo cosh k\ cos /c2 -f #o sinh fci sin fo 24D cosh2 /ci cos2 fe + sinh2 k\ sin2 /c2' A3 = qo «o cosh k\ cos /c2 + sinh 7q sin ^ 24D cosh2 k\ cos2 /c2 + sinh2 k\ sin2 fc2' (b) where cio = [01 — a2) /2ap. For a square plate, the maximum deflection is given by ^0, = ^, (0,0) = 0.0248 D Example 8.16 The Galerldn solution of a simply supported rectangular plate using trigonometric functions coincides with the Navier solution. To show this, we begin with the approximation [see Eq. (8.145a)] N N IVQ 2_^(/)sin sin i = \ 7=1 . ittx . jjty which satisfies all of the boundary conditions in Eq. (8.144a,b). Substituting Wpj into the (scmidiscrete) Galerldn integral: [82WN d2WNY\ V 3jc2 + dy2 )\ q + IqWn - h ( „.,, + A , ) j sin — sm —— dxdy
CLASSICAL PLATE THEORY 375 for m, n = 1, 2,..., N, we obtain rb pa ft* ra o= / Jo Jo N N d2c D(af + ^2cij + [l0 + h(af + ^)]^ . jtt^: . jny I . m.7rx . nny f , x sm sin — a } sm sm —-— axay. a /? I a b Since we can write N N b ' q(x, y, t) = 2^ 2^ <?U W sm sm " and due to the orthogonality of the trigonometric functions / sin — Jo a we have «/? ?rx . mux , — sm ax = -(P\2c~„ + \[n±T {.' ■,{„2 - fori = m, fori ^ m, 2x1 ^£»m flf?2 ' Qmn i The expression in the curl brackets is identical to Eq. (8.148) with cmn = Wmn and Nq = 0. Hence, for N ->■ oo the Galerkin solution is the same as the Navier solution. The use of the same approximation functions in the Ritz method will also produce the same result. Example 8.17 Consider an isotropic equilateral triangular plate (see Fig. 8,22) subjected to uniform transverse load of intensity q$. We wish to determine the deflection using the Ritz method. Since the essential boundary conditions require wq = 0 on Figure 8.22 An equilateral triangular plate with simply supported edges.
376 THEORY AND ANALYSIS OF PLATES the edges, the approximation functions should vanish on the edges. A natural choice for the first approximation function is the product of the equations of the edges themselves. The equations of the edges, with respect to the coordinate system shown, are i + _ = 0 on BC. a 3 1 x y y/3a a 1 x y 2 3a/3 2 = 0 on AC, = 0 on AB. (a) >/3 a a 3s/3 Hence a one-parameter Ritz approximation is of the form „,„. „, _ „ (i +1) (-Li +1 - ^) (-Li -1 - -^ = C{<f>[(x,y), ci = First note that 801 2x 3y2 3x2 Bx a2 a3 a* 320i 2 6x B2<t>\ 3x2 a2 a}' By2 3C1. B4>\ _ 2y dy a2 2 6x a2 a3 * 6xy B2<j>\ BxBy (b) 6y Substituting the above expressions into Eq. (8.221 a,b), we obtain fDu = D22 = D, Da = vD, and 2/^66 = (1 - v)D]: R"'DfT d2<f>i 32<pi /a20i d2<j>i B2<t>i a20i a*2 Bx2 + -<s + • ±on 920i 920! a2^j O20! + 2(1 - v)-^-^--^-^- + 3y2 9*2 Bx2 By2 J dxdy dxdy dxdy By2 By2 (4 36 24 \ / 2 6 36 \ =D U/0°+^/2°" jIi°)+2vD wTw+^/10" ^hv 36 / 1 36 T 12 \ + 2(1 - v)D-sIm + D -jToo + -g fto + -5 Ao ,
CLASSICAL PLATE THEORY 377 /4A0 /20 h% 3/i2 , /3o\ where him = / xmyn clxdy, Im = A0, /10 = Aqx, /ni = A0y, * = 5 J2Xh y= 3 Jl*> /u = T2 (t2xw+9*>)' (°) '» = £ (i:*? + 9^2) • '« = £ (E jf + 9j*) . (x;, y,) are the coordinates of the three vertices of a triangle, and Aq is the area of a triangle (Aq = a2/^/3). For the triangle in Fig. 8.22, we have 2a a a X{ = -^-» X2 =X3 = --, yi = 0, y2 = -3'3 = -p, 3 3 V3 * = )" = 0, I io = Iqi = /n = 0, /20 = /o2 = -j^-, /30 = l35-' 7,2 = —135"- Substituting these values into the expressions for /^n and F\, we obtain 16/Mo tfo^o - <W4 Thus the one-parameter Ritz solution becomes ">*■>>-£[*-&-&->&&+&]■ «> For an //-parameter Ritz approximation, one may use 4>i(x, v) = p{i-\){x, y)0</-l)O, y)t i = 2, 3, ..., N9 where /?; (,r, y) is a polynomial of degree /: * y * y x2 y2 P\(xt y) = - 4- -, P2U, y) = -- + -t + ^o . - • ■ • a a da az az
378 THEORY AND ANALYSIS OF PLATES The exact deflection is given by [which can be verified by substitution into the governing equation (8.195)] (e) Example 8.18 Consider a clamped, isotropic elliptic plate with major and minor axes 2a and 2b, respectively (see Fig. 8.23). We wish to obtain a one-parameter Galerkin solution for the case in which the plate is subjected to uniformly distributed load qo- The boundary conditions are u>0 ■■ 0, = 0 or = 0, = 0 . dn \ dx dy J (8.241) As before, we use the equation for the boundary of the ellipse but square it to satisfy the deflection as well as slope boundary conditions: x2 y2^2 Wi(x,y)-c\(p\(x,y) = ci [ 1 - -^ - p- (8.242) We have W}_ = _& ( J? _ ?\ W = _*l ( _ x2 _ f\ dx a2\ a2 b2)' dy b2 \ a2 b2)' 8 84<pi d«4>l 24 dx* ~ a4 aVi Bx2dy2 a2b2' 8y4 24 b*' Clearly, the boundary conditions in Eq. (8.241) are satisfied by the choice. Substituting into the expression K (d4W 4Wi d4W) d4Wi + 2„ ' + 'dx20y2 dy4 ) -qo, Figure 8.23 A clamped elliptic plate.
CLASSICAL PLATE THEORY 379 wc obtain /24 8 24\ n = clD{^ + 2^ + ¥)-qo' (8"243) which is independent of a: and v. In this case, there is no need to evaluate the weighted integral of the residual //> ra / K4>\(x>y)dxdyt -b J-a as it yields only a constant multiple of the expression in Eq. (8.243). Thus we have c = 4o 1 8D [(3/a4) + (2/a2b2) + (3/b4)] ' and the Galerkin solution, which coincides with the exact solution, is given by When a = by the solution reduces to that of a circular plate (x2 + v2 = a2). 8.2.7 Variational Solutions: Vibration This section is dedicated to the study of natural vibrations of rectangular plates by the Ritz method. Equation (8.143), without the applied loads q and (Nxx, Nyy, Nxy), reduces to fd4w0 d*w0 d4w0\ V dx* + dx2dy2 + 3/ ) d2w0 ( dAWQ d4w0 \ where h = Poh9 l2 = ^' (S245b) For natural vibration, the solution is assumed to be periodic: u>o(*, v,f) = w(x9y)ei0)t9
380 THEORY AND ANALYSIS OF PLATES where / = \/^T and co is the frequency of natural vibration associated with mode shape Wy and Eq. (8.245a) takes the form ) ,r /d2w d2w\ fd4w d4w d4w \dx* = 0. (8.246) We wish to find values of co such that Eq. (8.246) has a nontrivial solution w (mode shape). The weak form of Eq. (8.246) is cbrut [d2wd2Sw /d2wd28w d2wd28w\ + 2(1 -v) 32w 328w 32wd2> + dxdy dxdy dy wd28wl J~d^\ -co2 IowSw + h ( dw 88w dw B8w dx 3x dy dy )]) dxdy. (8.247) Note that the weak form is based on the principle of minimum total potential energy [and not Eq. (8.246) directly]. Using an N-parameter Ritz solution of the form iV W (X, V) « ^Cy0;(x, )>) 7=1 in Eq. (8.247), we obtain the eigenvalue problem ([R]~co2[B])lc} = {0l where a2yk. a2 RU = f D 9^a^ (d2Jpid2<Pj &4>,a24>j 'dx2 Bx2 V \ dy2 dx2 Bx2 dy2 12(1 v)3^ ^ I 82fr*2»/ k 'dxdy dxdy dy2 By2 dxdy, B'j-]a['°*>*j + h\j;-to +1717/ dxdy. (8.248) (8.249a) (8.249b) (8.249c) Equation (8.249a) represents a standard eigenvalue problem, which must be solved numerically.
CLASSICAL PLATE THEORY 381 As discussed earlier [see Eq. (8.222)], it is convenient to express the Ritz approximation of rectangular plates in the form N N w(*. y) « Wmn(x, j) = JJ cu Xi(x)Yj(y)f (8.250) where the functions X; and Yj for various boundary conditions were given in Eqs. (8.225)-(8.240). These functions satisfy only the geometric boundary conditions of the problem. Substituting Eq. (8.250) into Eq. (8.247), we obtain Eq. (8.249a) in which the coefficients R{ij){ki) of [R\ and coefficients £(//)(*/) of [#l are defined by Jo Jo d2X, d2Xp dx2 J dx2 Yg+2(l-v) dXi dYj dXp dYq +JXl€$.**f.T. + ™TlX-- dx dy dx dy 2y- d2Ya\ dy2 dx2 q dx2 'j^p dy2 +Xi d2Y: d2Ya -X dy2 " dy2 dxdy, ^-rrh^^££^**«)] (8.251a) dxdy. (8.251b) The size of the matrix [R] is N2 x N2, and Eq. (8.249a) can be used to calculate the first N2 frequencies of the infinite set. Example 8.19 For an isotropic (v = 0.25) plate simply supported on all sides, wc may select the following approximation functions: <hj = X,(x)Yj(y)9 fx\'+l j;y+1 For square plates with N = 2, the matrices [R] and [B] are given by (a) (b) [R) = [B] = 0.4346 0.2173 0.2173 0.1086 0.2173 0.2314 0.1086 0.1157 0.2173 0.1086 0.2314 0.1157 0.1086 0.1157 0.1157 0.0993 1.1111 0.5556 0.5556 0.2778 0.5556 0.3175 0.2778 0.1587 0.5556 0.2778 0.3175 0.1587 0.2778 0.1587 0.1587 0.0907 x 10 ,-i (c) x 10' l-5
382 THEORY AND ANALYSIS OF PLATES where the following notation is used to store the coefficient ^11 = #iiji> ^12 = fin,12, ^13 = #11,21, ^21 = #12,11, #22 = #12,12, #23 = #12,21, #31 = #21,11, #32 ~ #21,12, #33 = #21,21, #41 = #22,11, #42 = #22,12, #43 = #22,21, The first four frequencies (<omn = comna2V'ph/'D) are o)U = 20.973, coi2 =. 58.992, w2[ = 59.007, The exact frequencies from Eq. (8.173) are a>n = 19.739, co[2 = cb2\ = 49.348, cb22 = 88.826. i?14 = #n,22, #24 = #12,22, #34 = #21,22, #44 = #22,22* £22 = 92.529. (d) (e) For larger values of TV and M, the Ritz method will yield increasingly accurate and more frequencies (see Licw et al. [29]). Example 8.20 For a rectangular plate clamped on all sides, we may select the following approximation functions: <t>U = Xi(x)Yi(y), w-(r'-2(5r+sr- (a) (b) (c) For isotropic (v = 0.25) square plates with N = 2, the matrices [R] and [#] are given by [#] = [*] = 0.2902 0.1451 0.1451 0.0725 0.1451 0.1007 0.0725 0.0503 0.1451 0.0725 0.1007 0.0503 0.0725 0.0503 0.0503 0.0335 0.2520 0.1260 0.1260 0.0630 0,1260 0.0687 0.0630 0.0344 0.1260 0.0630 0.0687 0.0344 0.0630 0.0344 0.0344 0.0187 x 10" (d) x 10 -7 The first four nondimensionalized frequencies (comn = comna2*/ph/D) are cou = 36.000, a>n = 74.296, oni = 74.297, ti22 = 108.592, and the approximate frequencies (see Leissa [28]) are o)U = 35.999, tin = 73.405, w2\ = 73.405, cb22 = 108.237.
CLASSICAL PLATE THEORY 383 Table 8.11 Effect of the plate aspect ratio on the nondimensionalized frequencies of clamped isotropic (v = 0.25) rectangular plates m 1 2 n 1 2 1 2 0.25 22.890 24.196 63.466 64.943 0.5 24.647 31.867 65.234 71.941 a>mn u Vph 1D t°r values of 0.667 1.0 27.047 36.000 41,899 74.296 69.297 74.297 80,394 108.592 a/b 1.5 60.856 94.273 151.419 180.887 2.0 98.590 127.466 260.938 287.766 Table 8.11 contains the first four natural frequencies of clamped rectangular plates. The results are obtained with N = 2 fcf. Tables 4.28 and 4.29 on pages 63 and 64, respectively, of Leissa [28]; no mention is made of the Poisson's ratio used]. The mode shape is of the form w(x, y) = Xi (ci YX + c2Y2) + X2 (c3Yx + cAY2) . (c) The vectors {c}1 for the four modes in the case of a/b = 0.25, for example, are given by f10 1 1 0.0 | 0.0 [ 0.0 •■ M2=' f -°-5 ] 1.0 0.0 0.0 ', (c}3=- ^ -0.50 ' 0.08 1.00 0.16 >. (c}4=< 0.25 1 -0.50 -0.50 1.00 J The vectors {c}{ for the four modes in the case of a/b = 1 are given by f 1.0] 0.0 0.0 1 I o.o ► , (c}2=< [ -°-51 1.0 0.0 0.0 •. W3=' [ -0.5 | 0.0 1.0 ! 0.0 '. {€}*=■ ' 0.25 ] -0.50 1 -0.50 1 i 10° J Example 8.21 Here we consider natural vibrations of rectangular plates with sides x = 0, a and v = 0 clamped and side y — b simply supported (CCCS). For this case, the approximation functions are given by The first four natural frequencies of rectangular plates are presented in Table 8.12 for various aspect ratios. Additional examples of the natural vibration of rectangular plates with other boundary conditions can be found in Leissa [28], Liew et al. [29], and Reddy [8].
384 THEORY AND ANALYSIS OF PLATES Table 8.12 Effect of the plate aspect ratio on the nondimensionalized frequencies of isotropic (v = 0.25) rectangular plates (CCCS) Mode (m,n) 0J) (1,2) (2J) (2,2) 0.25 22.840 24.654 63.411 65.444 0.5 24.215 34.421 64.957 74.254 comna2*Jph/D for values of, 0.667 1.0 25.913 31.849 46.951 72.025 66.663 86.497 84.914 120.087 a/b 1.5 48.217 86.045 179.342 208.448 2.0 73.573 108.452 310.663 337.474 8.2.8 Variational Solutions: Buckling Rectangular Plates Simply Supported Along Two Opposite Sides and Compressed in the Direction Perpendicular to Those Sides In Section 8.2.6 we discussed bending solutions of rectangular plates simply supported along two sides and having arbitrary boundary conditions on the other two sides. We used the Levy method of solution. Here we consider the buckling of rectangular plates simply supported along edges x = 0, a and subjected to uniform compression along the same edges (see Fig. 8.24). From Eq. (8.143b), by setting the time derivative and load terms to zero, and taking Nsx = No, Nyy = 0, and Nxy = 0, we obtain In the Levy method, we assume a solution of the form u>o(*, y) = ]P W,„(y) sina,„x, am = . (8.253) That is, the plate buckles into m sinusoidal half waves. The assumed solution satisfies the simply supported conditions fd2WQ . B2vjq\ «* = <>, il(„ = -Df-^ + v-jjl=() at*=0,fl. Substitution of (8.253) in (8.252) yields the equation d4Wnt 2d2Wm ( 4 <4 w\ n /q>)*a\ ~^T ~ 2am-^-j- + [am - — Noj Wm = 0, (8.254) which must be solved using the boundary conditions on edges y = 0, b. Here we wish to solve Eq. (8.254) using the Ritz method.
CLASSICAL PLATE THEORY 385 Simply supported No H iV0 Figure 8.24 Buckling of rectangular plates simply supported along two opposite edges (x = o, a) and subjected to uniform compressive forces on the same edges; other edges may have any combination of boundary conditions. The weak form of Eq. (8.254) is *b[d2Wmd28Wm f[ dry1 dy2 ♦[(3 + 2«^^ + W - Noal)Wm8Wm~\ cly m 0 2 dWm \ u vym d2Wm dSWm dy2 dy (8.255) where Nq = (Nq/D). For a simply supported edge [W = 0 and (d2W/dy2) = 0] or a clamped edge LW — 0 and (dW/dy) = 0], the boundary terms in Eq. (8.255) vanish identically. However, for a free edge the vanishing of bending moment and effective shear force requires d2Wm -va£ 1*^=0, ^Wm 2dWm (2- v)<x,r^— = 0. dy2 -m' - - dy3 - —i" rfy Hence, on a free edge, the boundary expression of Eq. (8.255) can be simplified to (d3Wm 2dWm\ d2Wmd&W„ -,/? dy2 dy = -v< Wm dSWm , dWm_w dy dy so that the weak form for SSSS, SSCC, and SSCS plates becomes rh[d2W,nd2SWm -rt „ ? dW dSWm , a - 9W„ «„, dy* dy* >^lfy-dT + tt-N»Z>)W->iW» + : dy, (8.256a)
386 THEORY AND ANALYSIS OF PLATES and for SSSF and SSCF plates it is given by fb [d2W,„ dzSW,„ . 0dWmdSW. fb \d2Wm d2SWm 2 dWm dSWm 4 - 2 -j 0=Jo h^^/?~ + ^-dT~dT + ('" ~ "^"^-Jdy 1^ (8.256b) o Next we assume an N-parameter Ritz approximation of the form N W,,,O0~X>?V/O'), (8.257) where <pj are the approximation functions [see Eqs. (8.225)-(8.240); replace x with y and a with b]. Substituting the approximation (8.257) into the weak form (8.256b), we obtain where ([*]-JV0[5]) {*} = {<)}, (8.258a) -alDvLj^ + ^p-w] , (8.258b) L dy dy J0 Bjj = o£ / <p;<pj dy, am = . (8.258c) Jo a The boundary term in Rfj is nonzero only for SSSF and SSCF plates. Equation (8.258a) has a nontrivial solution, a ^ 0, only if the determinant of the coefficient matrix is zero: \[R]-N0[B]\±0. (8.259) Equation (8.259) yields an Nth-order polynomial in Nq that depends on m. Hence, for any given aspect ratio a/b, the critical buckling load is the smallest value of No for all in. Table 8.13 contains numerical values of nondimensional critical buckling loads Ncr = Ncrb2/(7t2D) obtained using the algebraic approximation functions given in Eqs. (8.225)-(8.240) for various boundary conditions. The accuracy of the buckling loads predicted by the three-parameter Ritz approximation is very good and in agreement with the analytical solutions. Note that for certain aspect ratios there is a change in the buckling mode.
CLASSICAL PLATE THEORY 387 Table 8.13 Nondimensionalized buckling loads N of rectangular isotropic (v = 0.25) plates under uniform compression NXx = Wo (the Ritz solutions) a/b 0.5 1.0 1.5 2.0 N 1 3 E 1 3 E 1 3 E 1 3 E SSSS 6.334 6.250 6.250 4258 4.001 4.000 4.497a 4.341 4.340 4.258 4.001 4.000 sscc 7.725 7.693 7.691 8.606 8.605 8.604 7.120a 7.116 7.116 8.606 8.605 8.604 SSCS 7.915 6.860 6.853 8.149 5.741 5.740 7.040a 5.432 5.431 8.149 5.741 5.740 SSCF 4.896 4.526 4.518 2.050 1.699 1.698 1.751 1.339 1.339 1.915 1.386 1.386 SSSF 4.456 4.436 4.404 1.456 1.445 1.434 0.900 0.894 0.888 0.706 0.702 0.698 SSFF 4.000 3.958 — 1.000 0.972 — 0.988 0.426 — 0.250 0.238 — "Denotes change to the next higher mode; E denotes the "exact" solution obtained in Section 8.2. Formulation for Rectangular Plates with Arbitrary Boundary Conditions In this section we consider buckling of orthotropic rectangular plates using the Ritz method. Buckling problems of plates with combined bending and compression or under pure in-plane shear stress do not permit analytical solutions; therefore, it is useful to consider the Ritz method to solve such problems. The statement of the principle of virtual displacements or the minimum total potential energy for isotropic plates subjected to in-plane edge forces Nxx, Nyy, and Nxy can be expressed as pb pa 0=/ / Jo Jo ra2 D 82w d28w + •<$ S2w d28w 82w B2Sw 32w d28w d2wd28w dxdy dxdy By2 dy2 * dw d8w - diu 88w dx2 dy2 } ) dx dx w d8w dx + dy dy dw d8w dx dy )) dxdy. (8.260) In general, the applied edge forces are functions of position and they are independent of each other. Let Nxx = No, Nyy = Y\ No, Nxy = y2N0, (8.261) where No is a constant and y\ and y2 are possibly functions of position.
388 THEORY AND ANALYSIS OF PLATES Using an N-parameter Ritz solution of the form W in Eq. (8,260), we obtain 7-1 ([/?] - N0[B}) [c] = {0}, (8.262a) where pb pa U=D / Jo Jo d2<pi d2<pj / d2(pi d2<pj d2(pi B2<pj + 2(1-v) d2(pi B2<pj d2<pi d2<pj dxdy, Bxdydxdy dy2 By2 13 Jo Jo [^ dx Yi By By Y2\Bx By By Sx J (8.262b) dxdy. (8.262c) As discussed earlier, it is convenient to express the Ritz approximation of rectangular plates in the form M N mix, y) » Wmn(x, y) = ^ Y1CU %i WY]()>)• (8,263) The functions X\ and Yj for various boundary conditions arc given in Eqs. (8.225)- (8.240). Substituting Eq. (8.263) into Eq. (8.260), we obtain Eq. (8.262a) with the following definitions of the coefficients: pb pa %« = 0 / / Jo Jo ~d2X, d2Xk YY,yyd2Yj d2Y, dx2~dl2~YjYl + XiXkly2l^ + V\X' dx> dy2Y'+ dx2XkY'dy> + 2(1 -v) dXj dXk dY^ dx dx dy -— \dxdy9 9 dy J (8.264a) dYi r r [dx, dxk dY; d] + n (^yB. + X^^-Yi)] dxdy. (8.264b) \ dx dy dx dy ) \
CLASSICAL PLATE THEORY 389 Example 8.22 Consider a simply supported plate. For the choice of algebraic functions and for M = N = 1, we obtain n( "2b 2 2a \ (a) B"<" = ¥o + ^ + y2X°- (b) Note that the budding load under in-plane shear cannot be determined with a one- parameter approximation. The buckling load under uniaxial compression along the x-axis is given by setting y\ = 0: ., ^/I2b 20 I2a\ and the critical buckling load under biaxial compression is given by (y\ = 1): , /6/> 10 6a\ For square isotropic plates, the above expressions become D Dn2 . D Dn2 N?,. = 44-^ = 4.458^-, N*r = 22^ = 2.229^. a2 a2 a2 a1 The exact values for the two cases are 4Dtc2/ci2 and 2D7c2/a2, respectively. The results are in about — 11.5% error, and the values do not change for N — M — 2. For a square isotropic plate clamped on all sides and subjected to uniaxial compression (N"r) or biaxial compression (N%r)9 the buckling loads obtained using the Ritz method with algebraic polynomials given in Eqs. (8.239a,b) are (for N = M — 1 or 2): A^ = , 0.943^, <= 5.471^. Example 8.23 Let us consider a simply supported rectangular plate (see Fig. 8.25) with distributed in-plane forces applied in the middle plane of the plate on sides x = 0, a. The distribution of the applied forces is assumed to be ffXx = -Non = ~M) (l - co|) , (a) where No is the magnitude of the compressive force at y — 0 and cq is a parameter that defines the relative bending and compression. For example, co = 0 corresponds
390 THEORY AND ANALYSIS OF PLATES No(l-co) r MH3 JVo(l-Co) Figure 8.25 Buckling of simply supported plates under combined bending and compression. to the case of uniformly distributed compressive force, as discussed in Section 8.2, and for cq = 2 we obtain the case of pure bending. All other values give a combination of bending and compression or tension, We seek the deflection of the buckled plate, which is simply supported on all sides, in the form of a double sine series N M w(x, y) *tf WMN = 2-^2^ Cmn S!n Sm ~T~ n=\ m=\ b (b) that satisfies the geometric as well as the force boundary conditions of the problem. Substituting Eq. (b) for w and . pizx . qny Sw = sin sm ™— a b into Eq. (8.260) with Nyy - Nxy = 0, we obtain o-°EE-{(v)2(f)2+(T)2(f)2 «[(t)2(t)2+(v)2(t)'])" ;j=l m—1 N M M /?=1 m=l 9=1 where /i and h are nonzero only when p = m and q = n: -ff Jq Jo t/0 ./O . mux . nny . pnx . ^;ry f , a/? sin sm sm sm —— dxdy = —, a b a b 4 mnx rnty pnx qny ab cos cos cos cos dxdy — —, a b a b 4
CLASSICAL PLATE THEORY 391 and Inq is defined as '"-L[^-<) mux pitx . nity . qny cos cos sin —— sin —^- dxdy, a a b b which can be computed with the help of the following identities: h -L . nny . any t y sin —— sin —— ay o b b = — when n = a 4 = 0 when n ^ q and n ± q is an even number b2 2nq n2 (n2 — q2)2 Thus Inq = 0 if p ^m and when n ^ q and n ± <? is an odd number. /?<? -id-?*) when p =m. For any /7z and n, Eq. (c) becomes °-[(7)4+(t)4+2(t)2(v)2} *("/ A) 9 r™ ^L 8 ^^ 2/z#c, 7110 77,2 ~f fa2 ~~ ^^ (d) where the summation is taken over all numbers q such that n ± q is an odd number. Taking ?rc = 1 in Eq. (d), we obtain D(j2+n2) ci/T = M, s2b2 7ZA / co\ 8co ^ nqc\q CUY 2)+ x2^(n2-q2)2 (e) where s denotes the aspect ratio s = bja. A nontrivial solution, i.e., for nonzero cj/, the determinant of the linear equations in (e) must be zero if the plate buckles. For the one-parameter approximation (N = M = 1), we obtain No- 7T2D M2 tf + n*) 2x2 1 l-O.Sco1 (f) which gives a satisfactory result only for small values of en, i.e., in cases where the bending stresses are small compared with the uniform compressive stress. Higher- order approximations yield sufficiently accurate results for the case of pure bending. Table 8.14 contains critical buckling loads N = No(b2/7i2D) obtained using
392 THEORY AND ANALYSIS OF PLATES Table 8.14 Nondimensionalized critical buckling loads A/of simply supported rectangular isotropic plates under combined bending and compression NXy — W0(1 - cQy/b) (the Ritz solutions) co a/b 0.4 0.5 0.6 2/3 0.75 0.8 0.9 1.0 1.5 2 4/3 1 4/5 2/3 29.1 18.7 15.1 13.3 10.8 25.6 — — — — 24.1 12.9 9.7 8.3 7.1 23.9 — — — — 24.1 11.5 8.4 7.1 6.1 24.4 11.2 8.1 6.9 6.0 25.6 — — — __ 25.6 11.0 7.8 6.6 5.8 24.1 11.5 8.4 7.1 6.1 Ng Figure 8.26 Buckling of rectangular plates under the action of shearing stresses. the two-parameter approximation, except for Co — 2, where the three-parameter approximation was used (see [30]). Example 8.24 When the plate is simply supported on all its edges and subjected to uniformly distributed in-plane shear force Nxy = N®y (see Fig. 8.26), the Navier or Levy solution procedure cannot be used to determine the critical budding load. Hence we will seek an approximate solution hy a variational method. Let us seek the solution in the form N M W, y) % WMn = > > cmn sin sin -— ii a b (a) The approximate solution satisfies the geometric (w = 0) as well as the natural Wya- = 0 on sides x = 0, a and MYY = 0 on sides y — 0, b) boundary conditions of the problem. Therefore, both the Ritz and Galerkin methods yield the same solutions. The Galerkin solution is obtained by substituting Eq. (a) in the weighted-residual statement •-rrKS" 94 w dx28y2 d4w\ By*) 2< 32 w dx'dy <Ppq dxdy. (b)
CLASSICAL PLATE THEORY 393 We obtain ()=- Dab (VHz-nf)1+&)']■ pq N M -<EE(")(")m nq wwm (c) m = l n=\ where fa mrtx . piix T /2a\ p 1 , 0 hup = / cos sin dx = I -=■ } 7-5 ~r for pA 96 mr fh rnty . qny (2b\ lttq=]oco* — sin — <h>=[-2) (q2-n2) for q2^n2% and the integral lmp is zero when p = m or p ± m is an even number, and Inq is zero when </ = n or 4 ± n is an even number. The set of inn homogeneous equations (c) define an eigenvalue problem N M /J a2 (A(»'"),(pcf) ~ ^xyO(nw)(pq))cmn = 0 (d) ([A]-<,[G]){c} = {()}, where #fc AHiu,M = DSmpS„^ [dial + 2alHappnpq + fl^*] 1 {jinn.pq — 2>&mPn Imp'nqt 0im = , pn = —r~. a 0 Equation (d) has a nonirivial solution (i.e., cmn ^0) when the deteiminant of the coefficient matrix is zero. Note that [A] is a diagonal matrix while [G ] is a nonpositive- definite matrix; hence the solution of (d) requires an eigenvalue routine that is suitable for nonpositivc-definite matrices, it is found that the solution of (d) converges very slowly with increasing values of M and N. We note that [G] does not exist for M = N=l. For M = N = % we find from Eq. (d) the result (Gnj2 = Gu?2i = 0) (\A" ° 1_a/0 T 0 G,2~|\ I en \_\0\ \[ 0 A22J vvLG2i 0 J;|c22]^ lOJ * (e) where the coefficients are given by 7T4Z> o 0 Gn = Giiji = G22 = ^22,22 = 0, A22 s= ^22,22 = I6A11J1, G12 = Gn,22 = G2t = G22jn = 32
394 THEORY AND ANALYSIS OF PLATES Table 8.15 Nondimensionalized critical buckling loads N of rectangular isotropic (v = 0.25) plates under uniform shear NXy = N$y a/b N 1.0 9.34 1.2 8.0 1.4 7.3 1.5 7.1 1.6 7.0 1.8 6.8 2.0 6.6 2.5 6.1 3 5.9 4 5.7 and s = bja is the plate aspect ratio. Setting the determinant of the coefficient matrix in Eq. (e) to zero, we obtain N0 _ , 9P7T4 2 2 The two signs indicate that the value of the critical budding load docs not depend on sign. Timoshenko and Gere [30] obtained the following equation for short isotropic plates (a/b < 2) using a five-term (en, C22, C|3 = £31, C33 and C42) approximation: 81(1+j2)4 1 + 81 81 /I +s2\2 81_ / 1+j2 V 625 + 25 \9 + s2) +25 V1+9.V2/ where _ b a' A = - 32abN$y For a square plate, the above equation yields the critical budding load (N^y)cr — 9A(7t2D/b2)i whereas the value obtained with a larger (than 5) number of equations give 9.34 in place of 9.4. Table 8.15 contains the critical budding loads TV — (N®y)cr(b2/7t2D) obtained using a large number of parameters (see [30]). Example 8.25 Here we consider the budding of a clamped rectangular plate under in-plane shear. The principle of the minimum total potential energy for this case is pb pa Jo Jo d2w d28w dx2 dx2 (d2l 2w d2Sw d2w d28w" + 2 dx2 dx2 dy2 d2w d28\D d2w d28w ( ~ V)~dxJy"dxdy + dy2 dy2 _<r dSwdw dwdSw" + D \ dx dy dx dy dxdy. (a)
CLASSICAL PLATE THEORY We assume a Ritz approximation of the form N M w 7 = 1 i = \ and we obtain N M [ rb r« ° = D£E / f d2Xj d2Xf + 2 dx2 J dx2 " dXi dYj dXp dYq Ya+Xt d2Yj v d% dy2 Xp dy2 dx dy dx dy D \ dx J p dy dy dx qJ Using the two-parameter approximation w(x9y)^CiiXi(x)Yx(y)+C22X2(x)72(y)9 with Xi(x) = sin sinh h 1.0178 ( cosh cos - dxdy Cii X2&) = sin sinh a 7.853* 73* \ r, rt,™ / i 7-853* 7.853jc + 0.9992 I cosh cos a \ a a Y[{y) = sin 4-^l - sinh ^ + 1.0178 (cosh AJ3y 7.853v 72(v) = sin — sinh b 7.853y / 4.73v 4.73 v\ I cosh cos —-— I, V b b ) „™„ ( , 7-8533' 7,%53y + 0.9992 cosh —-^- - cos ——^ wc obtain where -( Au -N°yAi2 -N°yAl2 An 537.181 324.829 537 * a2b2 + a|" l°r c C22 7.181 \ Aw = 23.107 A22 = 0 /3791.532 4227.255 3791.532N l~? + a2b2 b*
396 THEORY AND ANALYSIS OF PLATES For a nontrivial solution, the determinant of the coefficient matrix in Eq. (d) should be zero, A[[A22 — AuAniN^)2 = 0. Solving for the budding load N®y, we obtain A)2 The ± sign indicates that the shear buckling load may be either positive or negative. For an isotropic square plate, we have a = b, and the shear buckling load becomes #£=±1764, (e) ; a1 whereas the "exact" critical budding load is D a The two-term Ritz solution (e) is over 21 % in error. < = ±I45fi- (f) This completes the discussion on the application of the Ritz method to the buckling of rectangular plates. 8.3 SHEAR DEFORMATION PLATE THEORY 8.3.1 Governing Equations of Circular Plates We begin with the following displacement field: ue(r90tz) = z<h(r>0), (8.265) uz(r,6,z) = u>o(r,0), where wq is the transverse displacement and {<pr, <f>g) are rotations of a transverse normal line about the (r, 0) coordinates. The quantities (wq, </>,., </)#) are called the generalized displacements. For thin plates, i.e., when the plate in-plane characteristic dimension to thickness ratio is on the order of 50 or greater, the rotation functions 0r and 4>q will approach the respective slopes of the transverse deflection: Bwq 1 dwo The linear strain components referred to the cylindrical coordinate system are given by p _ -p(l) pao _ 7p(l> 1 dWO r, 2e?.e = <l>9 H tt- , 2ezr r ad 2srg=zy%\ = *, + —. (8.266a) (8.266b)
SHEAR DEFORMATION PLATE THEORY 397 where i a^ *a ^ (8.266c) Since we are interested in deriving the equations of motion and the form of the boundary conditions, we assume that the plate is subjected to transverse load q. Using the principle of virtual displacements, SW = 0, we can write 0—1 I (<Trr&£rr + ^B$SS00 + <Jr9&YrB + Grz&Yrz + Otf&YzB) dz r drdO JQ{) J-h/2 — / q8u)Q r drdO + fir (stf>, + ~ J + Q0 (gfa + ^ ^^) - qSwo] r drdO, (8.267a) where ("also see Eq. (8.19)] /•A/2 /-A/2 fio = /Cv / ^ dz, Gr = Ks / o>r rfz. (8.267b) J-h/2 J-h/2 The Euler equations are Sfa: -~ \^ (rAirr) + ^ - M*J + fir = 0, (8.268) 1 f d dMaa 1 Ho: — I — (rMre) + -zr + M>e \ + Qo = 0, (8.269) r ]_dr d0 J (8.270) The natural boundary conditions are [in addition to Eqs. (8.25a-c)] 8<t>o rMrrnr + Mrono=Of (8.271) rMrotir + Meene = 0, (8.272) rfir«r + G*/i*=0. (8.273) The bending moments and shear forces are related to the displacements (u/o,0r,0e)by (8.274a)
398 THEORY AND ANALYSIS OF PLATES [ dtby 1 / 13^\" J^ = (l-v)Z>^— +r— -*j, g* = Ks f <roz dz = KsGh Ue + \^f) ■ (8.274b) (8.274c) (8.274d) (8.274c) where Ks denotes the shear correction factor, D = Eh3/[12(1 — p2)], E Young's modulus, G the shear modulus, v Poisson's ratio, and h the thickness of the plate. 8.3.2 Governing Equations in Rectangular Coordinates Wc begin with the displacement field for the general case that includes in-plane deformation and time dependency: u(x% )', Zy t) = uq(x, y, t) + z4>x(x> )>> 0. v(x, yt z, t) = vo(x, y, t) + z<t>y(x, y, /), w(x,y>z,t) = wo(x>ytt). (8.275) As before, (mo, uoi u>o) denote the displacements of a point on the plane z = 0, and <f>x and (f>y arc the rotations of a transverse normal about the y- and A'-axcs, respectively (see Fig. 8.27). The linear strains associated with the displacement field (8.275) are or [e} = [e°} + z{elh (8.276a) [ Sxx Eyy \Yyz Yxz [y*y. ► = A axx 8° *yy Yyz fxz y° L rxy J - +Z « \sl 1 byy 0 | 0 k) {*°} = duo ~~dx dvp dy dW° JLA, ~r~+ vy 3y dx dllQ dvo 17 + "^ {e>} = dc^ dx d<l>y dy 0 0 d(/)X d<t>y dy dx (8.276b)
SHEAR DEFORMATION PLATE THEORY 399 r^4 Figure 8.27 Undeformed and deformed geometries of an edge of a plate under the assumptions of the first-order shear deformation plate theory (FSDT). The governing equations of the first-order plate theory can be derived using the dynamic version of the principle of virtual displacements (or Hamilton's principle of Chapter 6): 0 = L {L [Nxx8s*x + Mxx*e**+ NyySe>>+ Myy8£>>+ Ns>*Y*y + Mxy8Yx\ + QxSy% + Qy8y$z + *«*>*">() - q&W - /rj (AqSuo + VO&VQ + WqSwq) - h (4>x84>x + 4>y$4>y) * Swq SSwq - dwo 38wo Nxx~ ~ 'Vvv Sx dx dy dy - / (Nnn&Un + N„s8us + Mnn8<f>n + M„s8<f>s + Qn8wo)ds dt9 (8.277) where k denotes the modulus of the elastic foundation, and the inertias (/o, h) were defined in Eqs. (8.129), (Af„„, Mns) in Eq. (8.138a,b), the transverse shear forces (fix. Qy) are defined by Or * 7-/I/2 l°"yz J dz. (8.278)
400 THEORY AND ANALYSIS OF PLATES and in-plane forces (NXXi Nxy, Nyy) are defined by fh/2 rh/2 rh/2 VXX = / <*XX dZ, Nyy = / (XXy ClZ, Nyy = / CTyy dZ. (8.279) J-h/2 J-h/2 J-h/2 Nx Here (Nxx, Nyy, Nxy) denote the applied in-planc compressive and shear forces on the edges and (Nnn, Nnx) are defined in a manner similar to the moments in Eq. (8,138a,b), The Eulcr-Lagrange equations are 8u0: Svq: dNxx dNxy ox dy dNXy i dNyy dx + 32uq d2vo dy = '°V (8.280) (8.281) 8QX SQy d /* dm t a 3wo\ d2WQ S<f>x: %: -T,\!i»l? + N»17r'*-5!t' 9M„ + dhhy_ _ = Jd2<l>x dx dy ' c)t2 dMXy dMyy 824>y -JX- + -dy--Qy h~aW The natural boundary conditions are M,„, ■ - N„„ = 0, Nns - N„s = 0, Q„ - Q„ = 0, - M„„ = 0, Mns - M„, = 0, (8.282) (8.283) (8.284) (8.285) where Qn 55 QXHX + Qy/ly. (8.286) The stress resultants in an isotropic plate are related to the generalized displacements by Nxx Nyy NXy ^A I v 0 v 1 0 o o l~v 2 J dx dvp By duo dvo il)7 + ~a7J (8.287)
SHEAR DEFORMATION PLATE THEORY 401 Mxx Myy Mxy Myy \ = D I v 0 v 1 0 1 -v 0 0 2 J a^. dx dy dy dx (8.288) Qy Qx -*.»[J ','] 3w0 dy dwQ dx + <Av (8.289) where the extensional stiffness A and bending stiffness D are defined as: Eh Eh3 A = 1-v2' D = 12(1 -v2)' In the present study we are interested in the bending deformation of plates, and it is governed by Eqs. (8.282)-(8.284). These equations can be expressed in terms of the generalized displacements (wo, <t>x, <£>>) by means of Eqs. (8.288) and (8.289): S2w0 a /. du)0 - 'dwo\ 9 /. au>0 , - 3w0\ to (""17 + N»'l^) + Ty (""17 + N»ly-) ' (8.290) D &4>x a2^, (l-v) /a2&. a2^, 3jc a^2 9y3xy 2^ Z) a^ax 2 ya*3y ax2 y vax9y a^2 -w»(£+*)-*&- (8.291) (8.292) 8.3.3 Exact Solutions of Axisymmetric Circular Plates For axisymmetric bending, the equilibrium equations (8.268) and (8.270) become _l\d_ 7 [17 (rMrr) - Mob + Q, = 0, —r(rQr)-q=0, r dr (8.293) (8.294)
402 THEORY AND ANALYSIS OF PLATES where [see Eqs. (8.274a-e)] Mrr = D(^ + v^\t (8.295) Moe = D (v^- + ^\ , (8.296) Qr^KsGhUr + ^\ (8.297) andD = £/T3/[12(l-v2)]. Integration of Eq. (8.294) gives = - / rq(r) dr rQr = - / rq(r) dr + c\. (8.298) Use of Eqs. (8.295), (8.296), and (8.298) in Eq. (8.293) and integration leads to the result DMr) = -F(r) + J(2logr - l)^i + ~c2 + -<*, (8.299) where F(r) = f - f r f - f rq(r) drdrdrdr. (8.300) Finally, from Eqs. (8.297M8.299), we arrive at _ dwo KsGh r ' >' ^ r 11 KsGh —- = - -^— -F + -(2 logr - 1)^ + -c2 + -c3 ar D L 4 2 r J -~ I rq(r)dr + ~, (8.301) KGh f r2 r2 1 K,GA w0(r) - - -^- -F(r) + — (logr - l)ci + — c2 + c3 logr - I ~ I rq(r) dr + a log r + c4. (8.302) The constants of integration, ci, c2, c3, and C4, are determined using the boundary conditions. The derivative of <py is given by D^L = _F" + r-(2\ogr + l)c, + ic2 - t^c3. (8303) dr 4 2 2rz For solid circular plates, the condition that the rotation <f>r be finite at /• =0 requires c3 = 0 [from Eq. (8.299)], In addition, if the plate is not subjected to a point load at
SHEAR DEFORMATION PLATE THEORY 403 r = 0, the shear force must be zero there. This implies that [from Eq, (8.298)] C[ — 0. Thus, for solid circular plates without a point load at the center, we must have c\ = 0, c3 = 0. (8,304) If a solid circular plate is subjected to a point load Qq at the center, we have 2x(rQr) = -fio at r = 0 -» c\ = -^. (8.305) Obviously, conditions (8.304) and (8.305) are not meaningful for annular plates. Example 8.26 Consider a clamped, solid circular plate of radius a. For this case, C3 = 0. The boundary conditions associated with the clamped edge r = a are wq(o) = 0, (pr(a) =0. Uniform Load For uniform load of intensity c/o, we have c\ = 0 and qua2 qua2 KsGh qoa4 C2-—^> C4 = —; h 8 4 D 64 Hence the deflection and rotation become |4 / J2\2 „„„2 ^W = ^-(l--|. (8.307) Note that the deflection in (8,306) has two parts, one due to bending u^{r) and the other due to shear Wq(v): The bending deflection is the same as that predicted by the classical plate theory [see Eq. (8.45)]. Thus, the effect of including transverse shear strain in the formulation is to increase the deflection by w^(r). In other words, the classical plate theory underpredicts deflection compared to the first-order shear deformation theory. We also note that the rotation 0r, in the present case, is equal to Mr) = --7^-. (8.309) dr That is, the rotation of a transverse normal is not affected by the shear deformation. Also, note that (dw^/dr) is not zero at r = a.
404 THEORY AND ANALYSIS OF PLATES The maximum deflection of the plate is _ q0a4 qua2 _ q0a4 Wma* ' 64D + AlTGh - 64D / 8 h2 \ 1 + ^ . (8.310) For v = 0.3, Ks = 5/6, and h/a = 0.01 (a thin plate), the difference between the maximum deflections predicted by the classical and first-order shear deformation theories is 4.57 x 10~4, which is negligible. For h/a = 0.1 (a moderately thick plate) it is 4.57 x 10~2, or 4.57%. Thus, the effect of shear deformation on the deflection is greater for thick plates. The expressions for the bending moments and stresses are the same as those in Eqs. (8.47a-d) (why so?). The shear force Qr and shear stress crrz are given by (2,(0 = K.Gk (* + ^) = K,Gkd-§. = -f (8.311a) ,ri(r) = G(* + ^)=-^. (8.311b) Point Load For point load <2o at the center, we have The boundary conditions of the clamped edge give Co,-, „ KsGhQoa2 Q0 c2 = - (21ogfl - 1), c4 = —-j^ + ^rlogfl. Hence the solution becomes , . Qoa2 \6tcD l-^ + 2^1og(-)l a1 a1 \af \ Uo j r (8.312) (8.3)3) 0r(r) = -7-=-log-. As before, the deflection predicted by the first-order shear deformation plate theory has two parts: one due to bending and the other due to shear. The bending part is the same as that predicted by the classical plate theory, and <j>r (r) = —(dw^/dr). However, the shear part w^(r) is singular at r = 0, making it difficult to determine the maximum deflection. . Example 8.27 Here we consider a simply supported solid circular plate under uniformly distributed load of intensity qo. The boundary conditions arc at r = 0: (r Qf) = 0; at r = a: w0 = 0, Mrr = 0. (8.314)
SHEAR DEFORMATION PLATE THEORY 405 The constants of integration are calculated to be (ci = cj — 0) _ q0a2 (3 + v) _ q0a2 KsGh. q0a4 (5 + v) C2~ 8 (1 + y)' C4~ 4 + Z) 64 (1 + v)' The deflection and rotation become \6D a |_(1 + y) azJ Note that the deflection is affected by shear deformation [the second term in Eq. (8.315)]; however, for this problem the rotation is not affected by transverse shear deformation. The maximum deflection is *~--«"-(?£) ^£s- <o,7a> and the maximum rotation is equal to the negative of the slope at r = a: 3 4>max = Ma) = *f"_ v (8317b) 8Z)(1 -h v) 8.3.4 Exact Solutions of Rectangular Plates As in the case of the classical plate theory, analytical solutions of the first-order shear deformation plate theory can be developed for simply supported rectangular plates using Navier's method and for rectangular plates with parallel edges simply supported by Levy's (semidiscretization) method. Here we limit our discussion to the Navier method of solution for the pure bending case (i.e., omit stretching deformation). The Levy method of analysis for the first-order shear deformation plate theory is more involved than the classical plate theory (as there are three equations to be solved even for pure bending case), and details can be found in the books by Reddy [7,8]. The simply supported boundary conditions for the first-order shear deformation plate theory (FSDT) can be expressed as (Fig. 8.28): wq(x, 0, r) = 0, wot*, b, t) = 0, u;o(0, y, t) = 0, wo(a9 y, 0 = 0, (8.318a) &(*, 0, f) = 0, 4>x(x, b, t) = 0, 0,(0, y, t) = 0, 4>y(a, y, t) = 0, (8.318b) Myy{X, 0, t) = 0, Myy(X, k, *) = 0, UXX (0, )>, t) = 0, MXX{Cl, J, t) = 0. (8.318c)
406 THEORY AND ANALYSIS OF PLATES Myy = 0 lo0 = $y s 0 fw0=<fcy«0 | Myy = 0 Figure 8.28 The simply supported boundary conditions of the first-order shear deformation theory. The boundary conditions in Eqs. (8.318a-c) are satisfied by the following expansions: OO OO mix, y, 0 = 2^}^ W"m(t) sm sln i • (8.319a) 77=1 777=1 OO OO & WTTA' . 717T}? 0.v(x, yy 0 = ^2^2 Xm„(t) cos sin ;jr=l 777 = i oo oo b ' #y(*, >', 0 = > > ^inn (0 Sin COS -—, (8.319b) (8.319c) 77 = 1 «? = 1 where a and b denote the dimensions of the rectangular plate. The mechanical load is also expanded in double sine series: v—\ x~^ , s . mux , nny q(x9 y> 0 = > > qmn (f) sin sin -—, ^—' z—' a b /7=1 7/7=1 where pa rb Hmnit) T / I 4 C*> )?' 0 sin Sln ~r~ dxdy. ab Jo Jo ah (8.320a) (8.320b) Substitution of Eqs. (8.319a-c) into Eqs. (8.290)-(8.292), with Nxy = 0, yields the following equations for the coefficients (W,„fii Xmn, Ymn): S\\ — A'll J12 *13 S\2 S'22 *23 S\2 523 ^33 ^ W/NH X-mn *mn ► + mil 0 0 0 m22 0 0 0 '"33 w™ X?)tn i 'mn 1 ► = H Qmn 0 0 (8.321)
SHEAR DEFORMATION PLATE THEORY 407 where sjj and m,-y are defined by (for orthotropic plates) jii = Ks(A5SafH + A44A?) + *. *U = ^v.v<4 + Nyytf, sn = KxAss<*m> *13 = ^^44^1, ^22 = (/>ll«w + #660* + ^55). *23 = (D,2 + D(s)ampn9 S33 = (#660& + />22j8^ + ^44), 77111 = /b» "»22 = ^2, OT33 = ^2, (8.322) and (*,„ — mit/a and /J„ = 7i7r//3. Bending Analysis The static solution can be obtained from Eq. (8,321) by setting the time derivative terms to zero: (8.323) Solution of Eq. (8.323) for each m, n = 1, 2,... gives (Wfnn, Xmn, Ym„)9 which can then be used to compute the solution (ivq, 4>x, </>y) from Eqs. (8.319a-c). We obtain su *12 S\3 S12 ^22 *23 *13 S23 ^33 _ ' wmn' X-mH mwi t = 1 &/TM 0 i o J bo b\ Y,„„ = ^-Qm„, (8.324a) where *m« = *?l l ^0 + ^12^1 + A-13&21 ^0 = ^22^33 — $23*23. /?1 = $23*13 - $12*33, bz = $12*23 ~ $22$13- The bending moments are given by 00 00 Mxx = -D ^ ^ (am*™ + v#, F„m) sin aw* sin ft,?, oo oo Myy = -DYllL, (vamXmn + /37mH) sin a,„x sin /J,,)7. (8.324b) (8.325) n=.\ m=l oo oo + <*m Ymn) COS OfOTX COS ft,)', n—\ m—\ and the shear forces can be computed from oo oo Qx = -^GA ]T ^ («;,iWffl/l + X,„„) cosafljj: sin^y, oo oo Qy = -KsGhY^Yl (^» W'"» + Ym>^ sin a'"* cos P«y' M = l #« = ! (8.326)
408 THEORY AND ANALYSIS OF PLATES The stresses can be computed using the constitutive equations: Ez <T, ayy (1 - v2) (amXnw + vpnYmn) sine*,,,* sin finy (vajnXmn + p„ Ymn) sin atnx sin pny - (pnXmn + am Ymn) cosa,„x cos p„y \<7yz\ _ Gj y^ y^ I (Ymn+PnWm„)smamxcosP„y i °xz J £1 ,£i 1(X'"" + a'n W,,w ^ °°s a"f * sin ^" }' J (8.327) (8.328) The transverse shear stresses can also be computed using 3D equilibrium equations in terms of stresses. They are given by &X7 = —- h2 <r« = -T 'yz -(f)2 (I)3' x $ina,nxsml$nyt (TuXmtt + T]2Ymn) co&amx smpny9 (TnX„m + TnYmn) sin amx cos pny9 (TllXmn + T32Ymn) -3 ■♦■?) (8.329a) where r" = (i^2) («i + VA'J ri2 = 2(1 - v) &m Pn j Tfe = ^^ (l^a* + tf) . (8329b) ?31 = 732 = (1 - V2) E (1 ~ V2) (0&+««£»). Example 8.28 Numerical results for the deflection and stresses of simply supported, isotropic, rectangular plates are presented here. The following nondimensionaliza- tions are used: W = tt>o forUL, w = W°{lk)X 102 forSL,
SHEAR DEFORMATION PLATE THEORY 409 Table 8.16 Effect of the transverse shear deformation on deflections and stresses In isotropic (v = 0.25) square (a = b) plates subjected to distributed loads". Load SL UL(J9)C b/h 10 20 50 100 CPT 10 20 50 100 CPT w 0.2702 0.2600 0.2572 0.2568 0.2566 0.4259 0.4111 0.4070 0.4060 0.4062 <*xx 0.1900 0.1900 0.1900 0.1900 0.1900 0.2762 0.2762 0.2762 0.2762 0.2762 dyy 0.1900 0.1900 0.1900 0.1900 0.1900 0.2762 0.2762 0.2762 0.2762 0.2762 <rxy 0.1140 0.1140 0.1140 0.1140 0.1140 0.2085 0.2085 0.2085 0.2085 0.2085 &XZ 0.1910 0.2387b 0.1910 0.1910 0.1910 0.2387h 0.3927 0.4909b 0.3927 0.3927 0.3927 0.4909b *yz 0.1910 0.2387b 0.1910 0.1910 0.1910 0.2387b 0.3927 0.4909b 0.3927 0.3927 0.3927 0.4909b aSee Fig. 8.28 for llic plate geometry and coordinate system. b Stresses computed using the stress equilibrium equations (8.329a,b); they are the same for aJJ ratios of*//1. cThe numbers in parentheses denotes the maximum values of m = n used to evaluate the scries. *" = °xx Gfe)' dyy = "» Gfe)' *" = Gxy Gfe)' *" = ffjr8(4)' ^ = ^(4)- (8-330) Table 8.16 contains the maximum nondimensionalized deflections (w) and stresses of simply supported square plates under sinusoidally distributed load (SL) and uniformly distributed load (UL), and for different side-to-thickness ratios. The stresses were evaluated at the locations indicated below: . fa b h\ . fa b h\ _ / , h\ a**{rr2)' ""{rri)' ^^J* The transverse shear stresses are calculated using the constitutive equations as well as equilibrium equations. They are the maximum at the locations indicated below: / * h\ fa h\ Of course, the constitutively derived transverse stresses arc independent of the ^-coordinate. The nondimensionalized quantities in the classical plate theory are independent of the side-to-thickness ratio. The influence of transverse shear deformation is to increase the transverse deflection. The difference between the deflections predicted by the first-order shear deformation theory and classical plate theory decreases with the increase in the ratio a/h (see Fig. 8.29).
410 THEORY AND ANALYSIS OF PLATES O.Ofr I '3 flection 0.08- 0.07t 0.06t 0.05 1 0.04" 0.03- 0.02-^ 0.01- 71—\—I—S—l—l—l—i—l—'—l—I—*—l—I—i—'—'— y~ .".I Uniform load Sinusoidal load isotropic orthotropic 0.00-pn-! i | i r-n | c r t r | i t I tjtt\ iji n i| i jh ;ii hji I i ! prm 0 5 10 15 20 25 30 35 40 45 50 afh Figure 8.29 Nondimensionalized center transverse deflection (w) versus side-to-thickness ratio (a/h) for simply supported square plates. Natural Vibration For natural vibration, we n&t the mechanical load to zero and seek a periodic solution to Eq. (8.321) in the form Wmn\t) — "mn^ ' Xmn{t) — &mne » Ymn(t) = ¥%,/»', (8.331) and obtain the following 3x3 system of eigenvalue problem: *u S\2 S13 S\2 $21 J'23 svi «?23 *33_ -CD2\ mn 0 0 0 WZ22 0 0 0 WI33 \\ 1)1 TT mn 1 y0 Y° ► ^^ < [0] 0 (8.332) where sjj and my are defined in Eq. (8.322). Setting the determinant of the coefficient matrix in Eq. (8.332) to zero yields a cubic equation for o)2. If the rotatory inertia h is omitted (i.e., mn = WI33 = 0), the frequency equation can be solved for o?\ (D 1 / ^13*23 - *12*33 -W23 - J13J22 \ /0 ,™. = \s\\ s\2 - s[3 - (8.333) Wli V S22S33 — S23523 ^22^33-^23^23 / Example 8.29 Consider a simply supported rectangular plate. We wish to determine the natural frequencies using the first-order shear deformation plate
SHEAR DEFORMATION PLATE THEORY 411 Table 8.17 Effect of the shear deformation, rotatory inertia, and shear correction coefficient on nondimensionalized natural frequencies of simply supported plates (a> = (o{^jh)y/JjE\ v = 0.3, a/h = 10) cpTa w/oRI CPT with RI Ks FSDT w/oRI FSDT withRI 5.973 14.933 5.925 14.635 23.893 23.144 1.0 5/6 2/3 1.0 5/6 2/3 1.0 5/6 2/3 5.838 5.812 5.773 14.127 13.980 13.769 21.922 21.582 21.103 5.794 5.769 5.732 13.899 13.764 13.568 21.424 21.121 20.688 aw/o RI means without rotary inertia. Table 8.18 Effect of shear deformation, material orthotropy, and rotatory inertia on dimensionless fundamental frequencies of simply supported square plates [£ = viaZ/ftJp/E, K= 5/6, v = 0,25] Theory FSDT CPT a/h-+ w-RIa w/o-RI (5.885)b 5 5.232 5.349 5.700 10 5.694 5.736 5.837 20 5.835 5.847 5.873 25 5.853 5.860 5.877 50 5.877 5.879 5.883 100 5.883 5.883 5.885 aw-RI - with rotatory inertia; w/o-RI = without rotatory inertia. b Value in parentheses is the frequency without rotatory inertia. theory. Using Eq. (8.333), we can compute the frequencies. Table 8.17 contains the first four natural frequencies of square isotropic plates. The rotary inertia (RI) has the effect of decreasing the frequencies. Table 8.18 contains fundamental natural frequencies of square plates for various values of side-to-thickness ratio a/b. Buckling Analysis For buckling analysis, we assume that the only applied loads are the in-plane compressive forces N yy Nxx = Nq, Nyv = YN0, y = -^> Nxx (8.334) and all other loads are zero (and k = 0), From Eq. (8.231) we have *13 ^23 ^'33 1 i \wmn] X-mn *mn ► = < [0] 0 oj (8.335)
412 THEORY AND ANALYSIS OF PLATES For a nontrivial solution, the determinant of the coefficient matrix in Eq. (8.335) must be zero. This gives the following expression for the buckling load: NQ \*l + Yfi) (8.336) KfAwAss Ks A$s KsAa4 co = Dua*n + 2 (Dl2 + 2Z>66) a* ft? + D72& cx = c2c3 - (c4)2 > 0, c2 = Dnal + DeeP^ c3 = Art<4 + #22/^, c4 = (Dn 4- AaO awAi. (8337) The expression in (8.336) is of the form i +/q with k\ < hj which implies 1 > -—, 1 + /C2 indicating that transverse shear deformation has the effect of reducing the buckling load. Table 8.19 contains the critical buckling loads N = Ncrb2/(n2D22) as a function of the plate aspectratio a/b, side-to-thickness ratio b/ h, and modulus ratio E\jEi for uniaxial (y = 0) and biaxial (y = 1) compression. The classical plate theory (CPT) results are also included for comparison. The effect of transverse shear deformation is significant for lower aspect ratios, thick plates, and larger modular ratios. For thin plates, irrespective of the aspect ratio and modular ratio, the budding loads predicted by the shear deformation plate theory arc very close to those of the classical plate theory. 8.3.5 Relationships Between Bending Solutions of Classical and Shear Deformation Theories From the analytical solutions of axisymmetric bending of circular plates, it is clear that there exists a relationship between the solutions of the two theories, namely, the classical (CPT) and first-order shear deformation (FSDT) theories. Indeed, Reddy and Wang [31 ] developed such relationships using the similarity of the equations of the two theories and the load equivalence (also see Wang et al. |32]), We present a brief discussion of the relationships for the case of bending. First, we consider axisymmetric bending of circular plates. The governing equations of the two theories are summarized below. The superscripts on variables refer to the theory (C for CPT and F for FSDT). CPT li(rQ') = ^ (8-338)
SHEAR DEFORMATION PLATE THEORY 413 Table 8.19 Buckling loads N of simply supported plates under in-plane uniform uniaxial (y = o) and biaxial (y = 1) compression (Ks=5/6) 1/ a y s 0 0.5 1.0 1.5 3.0 1 0.5 1.0 1.5 3.0 h b 10 20 100 CPT 10 20 100 CPT 10 20 100 CPT 10 20 100 CPT 10 20 100 CPT 10 20 100 CPT 10 20 100 CPT 10 20 100 CPT aDenotes mode numbers (m, £ = ' 5.523 6.051 6.242 6.250 3.800 3.948 3.998 4.000 4.045(2,1 )a 4.262(2J) 4,337^1> 4340(2,0 3.800^ ^ 3.9480.0 3 998(3,l) 4.000(3>l) 4.418 4.841 4.993 5.000 1.900 1.974 1.999 2.000 1.391 1.431 1.444 1.444 1.079 1.103 1.111 1.111 ^l -3 11.583 13.779 14.669 14,708 5.901 6.309 6.452 6.458 5.664 5.942 6.037 6.042 5.664<2-1) 5.942<2'1> 6.037^J) 6.042^ l> 9.405 11.070 11.737 11.767 3.015 3.173 3.227 3.229 1.788 1.841 1.858 1.859 1.151 1.172 1.179 1.179 £ = "> 23.781 35.615 42.398 42.737 11.205 12.832 13.460 13.488 8.354 8.959 9.173 9.182 8.354<2'U 8.959<2']> 9.173^ 9.182<2-1) 15.191<^3> 21.565^-3> 25.241 <!'3) 25.427(1'3> 5.662 6.433 6.731 6.744 2.614 2.769 2.823 2.825 1.227 1.251 1.259 1.260 , 7?) al which the critical buckling load occurred; (m, n) — (1, %=* 34.701 68.798 100.750 102.750 19.252 25.412 28.357 28.495 13.166 15.077 15.823 15.856 13.1660'1) 14.052 14.264 14.273 17.773(U> 30.073(M) 40.157<1'4> 40.784(i'4) 7.518<!'2> 9.308(1-2) 10.156<1'2> 10.196(1>2> 4.093 4.651 4.869 4.879 1.375 1.414 1.426 1.427 1) for all other where r /d2w£ vdu)n\ (8,339a) (8.339b)
414 THEORY AND ANALYSIS OF PLATES Ko / d2w% ldtv£\ FSDT where MrFr = D (^- + j<j>n , (8.341a) M&> = D (v*£- + 1^ , (8.341b) Q? = KsGhL? + ??f\. (8.341c) Next, we introduce the moment sum 1 + v Using Eqs. (8.338)-(8.342), we can show that M = -^———. (8.342) MF = D(^+l-(j>lSj = D~(r4>r)> (8.344) 1 d ( dMc\ 1 d ( dMF\ -vTA'-Trr-"' -rTr\r-dr)=-«> ***> and dMc d { wr* , - ~ r"dT = 57(rM^) " *& = '■fi?. (8.346) dMF d , , Fx F F r~rf7~ = 2^rJI/")" M°°= r2r • (8-347) From Eqs. (8.338) and (8.340a), because the load q is the same, it follows that rQ?=rQf + Ci, ' (8.348) and from Eqs. (8.346)-(8.348) we have dM1' dMc r r r—^T = r~JP + Ci -»■ M ~ M + ci lo§'' + c2, (8.349)
SHEAR DEFORMATION PLATE THEORY 415 where C\ and C2 are constants of integration. Next, from Eqs. (8.343), (8.344), and (8.349), we have du£ C\r Cir C3 ^^-^+4^(2l0g'"-i) + ^ + ^- (8350) FinaUy, from Eqs. (8.341c), (8.346), (8.347), (8.349), and (8.350), we obtain -**+jk{&+c-f)- (8-3S1) dw^ dr and noting that Q% — dMc/dt\ we have F c, Mc .Cir2 , , Ci , C2r2 C3logr , C4 (8.352) The four constants of integration are determined using the boundary conditions. The boundary conditions for various cases are given below. Free Edge rQr = rQf = 0, rMf.r = rAf£ = 0. (8.353) Simply Supported Edge vug = w$ = 0, rAf£ = rAf£ = 0. (8.354) Clamped Edge du) wf = wf = 0, <£,F = —^- = 0. (8.355) So/M Circular Plate at r = 0 (i.e. at the plate center): ;/ du;£ A? = —SL =0, Ci = 0. (8.356) dr Example 8.30 Consider a circular plate of radius a and subjected to a linearly varying axisymmetric load q — qo(] — r/a) (set q\ = 0 in Fig, 8.30). For simply supported as well as clamped (at r = a) circular plates, the boundary conditions (8.354) and (8.355) give DMZ d = C2 = C3 = 0 and C4 = —^f, (8.357) where M% is the moment sum at the simply supported or clamped edge (r = a) of the classical plate theory, and it is given as follows.
416 THEORY AND ANALYSIS OF PLATES Ql-QQ I h T J T T 1 T T T T T T T T T T T T J T° ^Q^& ' z, wo(r) Figure 8.30 Simply supported solid circular plate under linearly varying axisymmetric load. Simply Supported Edge Mc = M&(a) = DQ-v)dw<~ 0 1 + v a Clamped Edge dr DO -v)d2w9 dr2 d2w\ M«=-°dl, Hence, the relationships (8.350) and (8.352) become Mr) = - dr ' c F c MH')-M^ W° <r) = "'« + KsGh (8.358a) (8.358b) (8.359a) (8.359b) The classical plate theory solution for such plates is given by (see Reddy [8]) C = 0 14400D 3(183 +43v) 10(71 +29v) /r\2 l + v + 255 + v W Q4-«(;)'] (8.360a) for a simply supported plate, and .4 r 0 14400D 129 - 290 (-)2 + 225 (-)* - 64 (~)5~| (8.360b) for a clamped plate. Substitution of Eqs. (8.360a,b) into Eqs. (8.358a), and the result into Eq. (8,359b), yields the deflection of the simply supported plate according to the first-order plate theory: f c , cloa £s ['-»(;)'+*(:)']• «*»
SHEAR DEFORMATION PLATE THEORY 417 The maximum deflection is (occurs at r = 0) i = «,!, +3^. (*•**) Next, we consider the relationships between the bending solutions of the classical plate theory and the first-order shear defomiation plate theory for polygonal plates (i.e., plates with multiple straight edges). The equations of equilibrium of plates according to the classical [Eq. (8.195)] and the first-order [Eqs. (8.290)-(8.292)] plate theories can be expressed in terms of the deflection wq and the moment sum (or Marcus moment) M. of each theory, r A/fv + Af£ w MFX + MFV MC^_^ ^ MF = _J^ JJt (g363) as 14- v 1 + v Mc V2MC = -q, V2w% = -1—, (8.364a,b) D F 9 . F 9 / F MF \ MF (8.365a,b) where the superscripts C and F refer to quantities of the classical and first-order plate theories, respectively, D is the flexural rigidity, and v Poisson's ratio. The moment sum in each theory is related to the generalized displacements of the theory by the relations MF = D(d-^+d-p.). (8.366b) From Eqs. (8.364a) and (8.365a), since q is the same in both equations, it follows that MF = Mc + DV2d>, (8.367) where 4> is a function such that it satisfies the biharmonic equation V44> = 0. (8.368) Using this result in Eqs. (8.364b) and (8.365b), we arrive at the relationship wF = < + P— + * - 4> (8.369a) = ^ + 6W^0V2WoC + *-*' (8369b)
418 THEORY AND ANALYSIS OF PLATES where ^ is a harmonic function that satisfies the Laplace equation V2* = 0. (8.370) Note that the relationship (8.369b) is valid for all plates with arbitrary boundary conditions and transverse load. One must determine <S> and ^ from Eqs. (8.368) and (8.370), respectively, subject to the boundary conditions of the plate. It is difficult to determine these functions for arbitrary geometries and boundary conditions. In cases where wfi = w$ on the boundary of the polygonal plates and Mc is either zero or equal to a constant M*c (which can be zero) over the boundary, ty — <f> simply takes on the value of -M*c/{KsGh). However, if Mc varies over the boundary, the functions * and 4> must be determined separately. Restricting our analysis to the former case, Eq. (8.369a) can be written as w£(x,y) = w£(x%y) + Mc -M*c KsGh (8.371) Equation (8.371) can be used to establish relationships between deflection gradients, bending moments, twisting moment, and shear forces of the classical and first-order shear deformation plate theories, as given below: By dx H 8y + + MFXX = Mg + = <,+ = <v + MF — Mc + KsGh: KsGh' D(l- (8.372a) (8.372b) -Myy+KsGh 2KsGh D KsGh D(l-v) 8 KsGh Jx D(l - v) 2KsGh D [s<fc'-«>+4(e'-^ £<0?-flf)-£<a,'-flf)] (8.373a) - £>(1- v) 9 .F „Cs = M„cv + -±——L — (Qr - Qcy), yy M^ = M^y + KsGh dy £>(l- 2KsGh ?[i(of-flf) + i(ef-flf>]. (8.373b) (8.373c) The above relationships are exact if wfi = wfi at the boundaries and the Marcus moments in both theories at the boundary are equal to the same constant.
SHEAR DEFORMATION PLATE THEORY 419 In the case of simply supported polygonal plates, it can be shown that the Marcus moment (or moment sum) in the classical plate theory is zero along with the deflection: Wq = Mc = 0 along the straight simply supported edges. (8.374) In the first-order plate theory, the simply supported boundary condition involves specifying, in addition to the deflection and normal bending moment, the tangential rotation on the edge. This boundary condition is known as the simple support of the "hard" type: u£=0, Af£=0, & = 0, (8.375) where n is the direction normal to the simply supported edge and s is the direction tangential to the edge. Owing to these conditions, 3<ps/ds = 0 and the Marcus moment MF is thus equal to zero. The boundary conditions of the FSDT for the simply supported plate are therefore Wq = MF = 0 along the straight simply supported edges. (8.376) Since the Marcus moments at the boundaries of plates with any polygonal shape and simply supported edges are equal to zero, Eq. (8.371) applies to such plates with X*c=0.Wehave /;• r Mc < = <+ksTh' (8'377) Equation (8.373) is an important relationship between the deflections Wq and wJq of a simply supported polygonal plate. That is, if deflection of a simply supported plate using the classical plate theory is available, we can immediately determine the deflection according to the first-order theory for the same problem from Eq. (8.377), thus bypassing the task of solving the equations of the first-order shear deformation plate theory. Using the same reasoning, one may readily deduce that Eq. (8.371) holds for simply supported polygonal plates under a constant distributed moment M*K along their edges. For additional results on this topic, the reader may consult the book by Wang et al. [32] and references therein. Example 8.31 Consider the simply supported equilateral triangular plate of Example 8.17. Equation (e) of that example contains the classical plate theory analytical solution for the deflection of the plate under uniform load qo: w° = %5 [* ~3f* ~ <? + ^ + ^](l- *2 - f)' (8-378) where x = x/a and y = y/a (see Fig. 8.22). In view of Eq. (8.378), the Marcus moment is given by Mc = -DV2w% = ^- \x3 - 3xy2 - (x2 - f) + ^1. (8.379)
420 THEORY AND ANALYSIS OF PLATES Therefore, the deflection wfi according to the first-order plate theory of the same problem is given by Eq. (8.377) as F qo<r -lf-{x2+f) + = 27 4 -2 r:2 9~A —y D 16 K,Ga2 (8.380) Example 8.32 Consider a simply supported rectangular plate of side lengths a x b, as shown in Fig. 8.13 for the classical plate theory and in Fig. 8.28 for the first-order plate theory. The plate is subjected to a distributed load q(x> y). The deflection of this problem according to the classical plate theory is given by Eq. (8.150b): wt . (m2 «2y a . niTtX . 727TV sm sin * ' (8.381a) where n(r vi sin Sin - a b In view of Eq. (8.366a), the Marcus moment is given by 4 fh f° . mnx . nny , f ab Jo Jo " u (8.381b) Qmn ~ (m2 n2\ a , mnx . nny sin sin—-. (8.382) Using Eq. (8.377), the deflection w£ according to the first-order plate theory applied to the simply supported rectangular plate under distributed load is given by 00 OO < = ££ . Q« i . (m2 n'\ . mnx . nny sin sm —-— a b (8.383a) where Qnw — tfmn i + 2/m2 n2\ 6KSG (8.383b)
SHEAR DEFORMATION PLATE THEORY 421 This solution can be verified to be the same as that in Eq. (8.319a) with Wnw defined in Eq. (8324a). In particular, for a sinusoidally distributed load, q(x» y) = <7o sin — sin —. (8384) a b The deflection becomes (qmn = qo) F qob* |\ , 7r2a+s2) W0 " ^4(1 + ^2)2 | ' "*" f>rnah2 L 6KsGb2 jtx . 7iy sin — sm-p-, (8385) a b where s is the plate aspect ratio, s = bja. 8.3.6 Variational Solutions of Circular and Rectangular Plates In this section we consider variational formulation of axisymrnelric bending of circular plates and plates in rectangular Cartesian coordinates using the first-order shear deformation plate theory. The variational fomiulation of the shear deformation plate theory mirrors that of the classical plate theory. Note that one may use either the Ritz method or the Galerkin method to solve the problem. First, we consider axisymmctric circular plates. If we wish to use the Ritz method, we must construct the weak form. The weak form based on the virtual work statement (8.267a) is given by Jo L V dr r ) dr \ dr r } r + K,Gh (V + ^) (s<f>r + ^) -0Sm]rdr. (8.386) One must add additional terms corresponding to any applied loads and moments to the above expression. Next we seek Ritz approximation of the form M N w0(r)«J]fl/Vr/(r), <M'0 ~^bjvjir), (8387) ' = ! 7=1 where we assumed that all specified essential boundary conditions are homogeneous. Substituting Eq. (8387) into the virtual work statement (8386) and collecting the coefficients of 8aj and 5/?/, we obtain -gM££H"'+sM^-4- [ qf-,rdr, (8.388)
422 THEORY AND ANALYSIS OF PLATES M 0 = KsGh £ IT n^j- r dr] aj + £ UxGh J* <pk<pj r dr for k = 1,2,..., N. In matrix form, these equations can be written as \[A] [B]}Ua}\_UF}\ [[B]r [C]\\{b}l-\{0}\ (8.390) where An = K,Gh r *h*h. ,dr, Bn = K,Gh [ ^-9jrdr, Jo dr dr J0 dr + KsGh I <pk(pj rdr, Jo r Fj = j q^i r dr. Jo (8.391) Example 8.33 Consider a simply supported solid circular plate. We wish to seek the Ritz approximation (8.387) with M = 2 and N — 1, and Vr,(r) = l--f a in{r) = 1 --r, 0>i(r) = which meet the essential boundary conditions of the problem: wq(ci) = 0 and 4>r(0) = 0. Evaluating the integrals in Eq. (8.391), we obtain (A = D/KsGh) KsGh 12 whose solution is 6 8 -4a 8 12 -6a -4a -6a 3a2 + 12(1 4- v)A 12 ai =0, a2 = 24 2 r + ■ (1 + v)D K; —1 *i 2^0a3 3(1 + v)D The Ritz solution becomes „4 j W2(r) = *i(r) = ^oa 24D (1 + 2^0a3 v) V W 6JT,GA V W ' 3(l + v)£> v)£> W (8.392) (8.393)
SHEAR DEFORMATION PLATE THEORY 423 A close examination of the problem indicates that the deflection is symmetric and the rotation is antisymmetric about r = 0. Hence, wo(r) is a symmetric function and <j)r is an antisymmetric function of r. This also explains why a\ = 0 (only when M > 1). This understanding helps in selecting proper functions and reducing the effort. Thus, the solution in Eq. (8.393) amounts to using one-parameter (M = 1 and N = \) approximations with r2 r ir\(r) = 1 =■, <pi(r) = -. a1 a If a two-parameter approximation for wq and a two-parameter approximation for <pr are used, with the following choice of approximation functions: ,2 „4 „ 3 in (r) = 1 - ^ ir2(r) = 1 - ^r, <pi (r) = -, <p2(r) = ^, (8.394) the exact solution (see Example 8.27) would be obtained. Next, we consider the variational formulation of plates in rectangular Cartesian coordinates. While no numerical examples are included, the formulation should prove useful in obtaining Ritz solutions of plates with arbitrary boundary conditions. We begin with the weak form [which is the same as the virtual work statement (8.277) specialized to the static, pure bending case]: 0 = /" [MxxSexxx + MyySslyy + MxySyx[y + QxiYxz + Qy^Yyz ~ cl8wo] dxdy - <b (Mn„84>n + Mns&4>s + QnSw0) dsy Jqo \ \9x dy J dx \ dx 8y J By 2 \ 8y dx J V dy dx J - qSwo I dxdy - * {Mnn8<t>n + Mnx8(ps + Q„Sw0) ds. (8.395) Assuming Ritz approximation of the form N m<* W^(x9y) = ^al-^I-(*f y), i=\
424 THEORY AND ANALYSIS OF PLATES M 4>x *<$>™(x,y) = J^bi<pj]\x>y)y (8396) M 4>*™*y(x,y) = J2cWi(x>fl> 1=1 and substituting into the weak form (8.395), we obtain the following Ritz equations: (8.397a) r [A] [*] [cii [B\T [D] [E] LiC|T [Ef [G]J \[a)} {b} \{c}\ y = < \{F1} {F2} {IF3} where B,j = / KsGh ^-<p{P dxdy, Jsin ox } C,j = f KsGh d-^-yf dxdy, „<» *JX) OS*'."' ^^ + Lzl^l^uKsGh(pj^? Dn= f \d( Jn0 [_ \ d* d* 2 By 9* dy 2 5)? 3jc + ■ i dxdy, Jq0 |_ \ $y h} 2 3x dx J Pi = / cifi dxdy + f Qnfi ds, Jqq Jr F? = £ (Mxxnx + Mxyny) <p\[) ds, F? = d> (Mxynx + Myyny) <p<2) ds. + KsGh <p™<pf dxdy, (8.397b) dxdy, This completes the variational formulation of the first-order shear deformation plate theory in rectangular Cartesian coordinates. Note that the geometric boundary conditions involve specifying only (wo, <f>,x> <j>y)>
EXERCISES 425 EXERCISES 8,1 Obtain the Euler equations and the natural boundary conditions associated with the functional tt/ ^ n r(„ /d2u>o\2 2Di26lw0d2wo fldw0\ rdi\ which arises in connection with axisymmctric bending of polar orthotopic annular plates with inner radius b and outer radius a. Assume that the deflection wq = 0 at r = a, the outer radius of the plate. 8.2 Use the virtual work statement in Eq. (8.23) for the pure bending case to show that the total potential energy functional for the bending of circular plates is given by d /* [/a2** idw0 i B2w0\2 + - / ku)Q rdrdO — I qwQ r drdO. 8.3 Derive the Euler equation and natural boundary conditions of the functional in Exercise 8.2. 8.4 Show that the exact deflection of a simply supported circular plate subjected to applied bending moment Mrr = Ma at r = a is 2 21 rdrdO .. MQa2 (x r2\ 8.5 Determine the deflection and bending moments of a circular plate under uniformly distributed transverse load c/o when the edge r = a is elastically built-in. The boundary conditions are wo(a) = 0, Mrr = p —- at r = a, dr where £ denotes rotational stiffness constant.
THEORY AND ANALYSIS OF PLATES 8.6 Show that the maximum deflection and bending moment of a simply supported circular plate under linearly varying load q = qo(l — r/a) are 4800D 4 /183 + 43v\ 1# 2/71 + 29v\ id v~n^r)' Mmax=qoa l~72o-J • 8.7 Show that the maximum deflection and bending moment of a simply supported circular plate under linearly varying load q = q\ (r/a) are w„ qxa4 /6 + v\ 2/4 + A 8.8 Show that the deflection of a clamped circular plate under the load CI = <lo(r2/a2) is given by -<"=!£H02+G)6]- 8.9 Show that the expression for the deflection of a clamped (at the outer edge) circular plate under linearly varying load q = qo(l — r/a) is q0a4 { r2 r4 r5\ voir) = « J" 129 - 290^- + 225-j - 64-^ ). 14400D V a2 a4 a5 J 8.10 Show that the expressions for the deflection and bending moments of a clamped (at the outer edge) circular plate under linearly varying load q = q\ (r/a) are 450D \ a1 a>} 45 L a 2 r 3 "1 M&0(r) = ^ (1 + v) - (1 +4v)^ 45 L tt 8.11 Solve the problem in Exercise 8.7 by one-parameter Ritz approximation. 8.12 Solve the problem in Exercise 8.10 by one-parameter Ritz approximation. 8.13 Determine the fundamental frequency of a simply supported circular plate using a one-parameter Ritz approximation. 8.14 Determine the fundamental frequency of a clamped circular plate using a two- parameter Ritz approximation. Determine the eigenvector. 8.15 Determine the deflection at the center of a simply supported circular plate under asymmetric loading given in Eq. (8.93) using the reciprocity theorem, 8.16 Determine the center deflection of a simply supported circular plate under hydrostatic loading q(r) — qo(i — r/a) using the reciprocity theorem.
EXERCISES 427 8.17 8.18 8.19 8.20 8.21 Repeat Exercise 8.16 for a clamped circular plate. Determine the center deflection of a clamped circular plate subjected to a point load <2o at a distance b from the center (and for some 0) using the reciprocity theorem. Repeat Exercise 8.18 for a simply supported circular plate. The total potential energy functional for the axisymmetric bending of a circular plate on an elastic foundation and subjected to concentrated load go at the center is given by 1 dwo d2WQ T +D d dw0\2 2 rdr - Qqwq(Q), where v)q is the transverse deflection, a is the radius, D the flexural rigidity of the plate, and k is the foundation modulus. Determine the two-parameter Ritz solution in the form u?o(r) = C[ -f- C2f2. Consider an isotropic, simply supported rectangular plate subjected to hydrostatic load that varies linearly along the y-axis, as shown in Fig. E8.21 sec also Table 8,3). Using the Navier solution method, determine the expressions for deflection wo(xt y) and bending moment Mxx. What are the maximum values of these quantities for a square plate? Figure E8.21 8.22 Consider an isotropic rectangular plate, clamped at v = 0, simply supported at x = 0, a and y = b (see Fig. E8.22), and subjected to uniformly distributed load qo. Using the Levy method of solution, determine the expression for deflection wo(x, y). What are the maximum values of the deflection and bending moment for a square plate? 8.23 Consider a square isotropic plate, clamped at y — ±a/2, simply supported at x — 0, a (see Fig. 8.17), and subjected to sinusoidal load q{x,y) = </o sin—. a
428 THEORY AND ANALYSIS OF PLATES '////////////// tr* Figure E8.22 Use the Levy method of solution and determine the expression for deflection wq(x, y), What are the maximum values of the deflection and bending moment? Give the algebraic functions X; and Yj required in the Ritz approximation (8.222) for the following rectangular plates: 8.24 Plate with edges x = 0, a and y = 0 clamped and edge y = b simply supported* 8.25 Plate with edge x — 0 clamped and the remaining edges free. 8.26 Plate with edges x — 0, a and y = 0 clamped and edge y = b free. 8.27 Plate with edges x = 0 and y = 0 clamped and edges x = a and y = b free. 8.28 Determine a one-parameter Ritz solution of an isotropic rectangular plate with edges x = 0, a clamped and edges y — 0, b simply supported, and subjected to a uniformly distributed transverse load qo> Use the approximation W(x , y) = en I 1 - cos J sin — (a) 8.29 8.30 to determine the maximum deflection of a square plate. Determine a one-parameter Ritz solution of an isotropic rectangular plate with edges x = 0, a clamped and edges y = 0, b simply supported, and subjected to a center point load go- Use the approximation in Eq, (a) of Exercise 8.28, Determine the maximum deflection of a square plate. Determine a one-parameter Ritz solution of an isotropic rectangular plate with all edges clamped and subjected to uniform load qo. Use the approximation W(xt y) = en (l - cos ^\ (l - cos ^A , (a) where the origin of the coordinate system is taken at the top left corner of the plate with the y-axis down (i.e., 0 < x < a and 0 < y < b). Determine the maximum deflection of a square plate. 8.31 Repeat Exercise 8.30 for a point load go at the center of the plate.
EXERCISES 429 8.32 Repeat Exercise 8.30 for an orthotopic plate, and determine the maximum deflection forthecase D\\ = D0> D22 = 0.5Do, and Dn + £>66 = 1.248D0. 8.33 Determine the deflection surface wq(x, y) of the simply supported, isotropic, equilateral triangular plate of Fig. 8.22 when the plate is subjected to a point load Qq at the centroid (a% y) = (0, 0). Use the one-parameter Ritz approximation of Example 8.17. 8.34 Consider the equilateral triangular plate of Fig. 8.22. Suppose that all edges are simply supported and loaded by a uniform moment Mnn = Mq along its edges. Use the one-parameter Ritz solution of Example 8.17 and determine F[. The exact solution is , , M0a2 wJo(*,;y) = AD Hint: Evaluate the line integral s-© -©-©©+©- (a) «-jN-E+*£)*- where T denotes the boundary of the triangle and (nAy ny) are the direction cosines of the line segment. 8.35 Consider a clamped, isotropic elliptic plate with major and minor axes 2a and lb, respectively. Obtain a one-parameter Galerkin solution for the case in which the plate is subjected to distributed load q = qo(x/a). 8.36 Determine a one-parameter Ritz solution of a simply supported, isotropic elliptic plate under uniformly distributed load. Hint: Use (a) ^O0=(l-^), which satisfies the geometric boundary condition wq = 0. 8.37 Show that substitution of Eqs. (8.295) and (8.296) for Mrr and Moo into Eq. (8.293) yields the result
430 THEORY AND ANALYSIS OF PLATES 8.38 Show that the one-parameter Ritz approximation (8.263) with Xj and redefined by Eq. (a) of Example 8.22 yields the following critical buckling load for a simply supported plate under combined bending and compression: /\2b 20 12a \ 1 ^-D(^ + ^ + ^Ja^o^)- 8.39 Show that the two-parameter Ritz approximation in Eq. (b) of Example 8.23 yields the equations [seeEq. (e) of Example 8.23] C" 0C09^DCl2 / a2\ a2 (l + ^j -"o^(l-0.5co) / a2\2 a1 (1+Vj -^o^(l-0.5co) „ 16«2 -NoC09^DCU + and show that the associated characteristic equation is AX2 - BX -h C = 0, where A = No(a2/7Z2D) and -<—2-(^)2' c\i = 0, (a) (b) (c) 8.40 Eliminate </>x and <pY from Eqs. (8.290)-(8.292) and express them in terms of wo alone. REFERENCES 1. Kirehhoff, G, Vorlesungen iiber mathematische Physik, Vol. 1, B. G. Tcubncr, Leipzig (1876). 2. Bassett, A. B., "On the extension and flexure of cylindrical and spherical thin elastic shells/' Phil Trans. Royal Soc. (London) Sen A, 181(6), 433-480 (1890). 3. Hildebrand, F. B., Reissner, E., and Thomas, G B., "Notes on the foundations of the theory of small displacements of orthotropic shells," NACATN-1833, Washington, DC (1949).
REFERENCES 431 4. Hencky, H., "Uber die Beriicksichtigung der Schubverzerrung in ebenen Platten," Ing. Arch., 16, 72-76 (1947), 5. Mindlin, R. D., "Influence of rotatory inertia and shear on flexural motions of isotropic, elastic plates," 7. AppL Meek, 18, 31-38 (1951). 6. Reddy, J. N., Energy and Variational Methods in Applied Mechanics, John Wiley, New York (1984). 7. Reddy, J. N., Mechanics of Laminated Composite Plates: Theory and Analysis, CRC Press, Boca Raton, FL(1997). 8. Reddy, J. N., Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia (1999). 9. Reddy, J. N., "A simple higher-order theory for laminated composite plates," J. AppL Meek, 51, 745-752 (1984). 10. Reddy, J. N., "A general non-linear third-order theory of plates with moderate thickness," Int. J. Non-Linear Meek, 25(6), 677-686 (1990). 11. Reddy, J. N., "A small strain and moderate rotation theory of laminated anisotropic plates," /. AppL Meek, 54, 623-626 (1987). 12. Vlasov, B. R, "Ob uravneniyakh tcovii isgiba plastinok [On the Equations of the Theory of Bending of Flatesl," Izv. Akd. Nauk SSR, OTN, 4, 102-109 (1958). 13. Jemielita, G, "Techniczna tcoria plyt srednieej grubbosci [Technical Theory of Plates with Moderate Thickness]", Rozprawy Insynierskie (Engineering Transactions), Polska Akademia Nauk, 23(3), 483^199 (1975). 14. Schmidt, R., "A refined nonlinear theory for plates with transverse shear deformation," J, hid Math. Soc., 27(1), 23-38 (1977). 15. Krishna Murty, A. V., "Higher order theory for vibration of thick plates," AIAA J., 15( 12), 1823-1824(1977). 16. Levinson, M., "An accurate, simple theory of the statics and dynamics of elastic plates," Mech. Res. Commun,, 7(6), 343-350 (1980). 17. Reissner, E., "On the theory of bending of elastic plates," /. Math. Phys., 23, 184-191 (1944). 18. Reissner, E., "The effect of transverse shear deformation on the bending of elastic plates," J. AppL Meek, 12, 69-77 (1945). 19. Reissner, E., "Reflections on the theory of elastic plates," AppL Meek Rev., 38(11), 1453-1464(1985). 20. Kromm, A., "Verallgerneinerte Theorie der Plattenstatik," IngMrck, 21, 266 (1953). 21. Kromm, A., "Uber die Randqucrkrafte bei gestutzten Platten," Z angew. Math, Meek, 35,231(1955). 22. Pane, V, Theories of Elastic Plates, Noordhoff, Leyden (1975). 23. Roark, J. R., and Young, W. C, Formulas for Stress and Strain, McGraw-Hill, New York (1975). 24. Timoshenko, S. P., and Woinowsky-Krieger, S., Theory of Plates and Shells, McGraw-Hill, Singapore (1970). 25. Ugural, A. C, Stresses in Plates and Shells, 2nd edition, McGraw-Hill, New York (1999). 26. Szilard, R., Theory and Analysis of Plates: Classical and Numerical Methods, Prentice- Hall, Englewood Cliffs, NJ (1974). 27. McFarland, D., Smith, B. L., and Bernhart, W. D., Analysis of Plates, Spartan Books, Philadelphia (1972).
432 THEORY AND ANALYSIS OF PLATES 28. Leissa, A. W., Vibration of Plates, NASA, Washington, DC (1969). 29. Liew, K. M., Wang, C ML, Xiang, Y, and Kitipornchai, S„ Vibration ofMindlin Plates: Programming the p-Version Ritz Method, Elsevier, Oxford (1998). 30. Timoshenko, S. R, and Gere, J. M., Theoiy of Elastic Stability, 2nd edition, McGraw-Hill, New York (1961). 31. Reddy, J. N., and Wang, C. M., "Relationships between classical and shear deformation theories of axisymmetric circular plates", AlAA /., 35(12), 1862-1868 (1997). 32. Wang, C. M., Reddy, J. N., and Lcc, K. H., Shear Deformable Beams and Plates: Relationships with Classical Solutions, Elsevier, Oxford (2000).
9 THE FINITE ELEMENT METHOD 9.1 INTRODUCTION As was pointed out in Chapter 7, the traditional variational methods are ineffective in solving problems that are geometrically complex, have discontinuous loads, or involve discontinuous material or geometric properties. In such cases, the selection of the coordinate functions is a formidable task. Even in cases where the coordinate functions are available, the computation of associated coefficient matrices cannot be automated for a fixed class problems (e.g., bars, beams, or plates) because the coordinate functions arc not always algebraic polynomials and they depend on the boundary conditions of the specific problem. Consequently, each time the essential boundary conditions are changed for the same differential equation (i.e., the class of problems is the same), the approximation functions are also changed and the coefficient matrices have to be recalculated. This process is not readily adaptable to computer programming. The finite element method is a procedure that uses the philosophy of the traditional variational methods to derive the equations relating undetermined coefficients. However, the method differs in two ways from the traditional variational methods in generating the equations of the problem. First, the approximation functions are often algebraic polynomials that are developed using ideas from the interpolation theory; second, the approximation functions are developed for subdomains into which a given domain is divided. The subdomains, called finite elements, are geometrically simple shapes that permit a systematic construction of the approximation functions over the element. The division of the whole domain into finite elements not only simplifies the task of generating the approximation functions, but also allows representation of the solution over individual elements. Thus, geometric and/or material discontinuities can be naturally included. Further, since the approximation functions are 433
434 THE FINITE ELEMENT METHOD algebraic polynomials, the computation of the coefficient matrices of the approximation can be automated on a computer. As will be shown shortly, the construction of the approximation functions is systematic, and the process is independent of the boundary conditions and data of the problem. In short, the finite element method is a pieccwise application of classical variational methods. The undetermined parameters often, but not always, represent the values of the dependent variables at a finite number of preselected points, whose number and location dictate the degree and form of the approximation functions used. The method is modular and therefore well suited for electronic computation and the development of general-purpose computer programs. The present study is intended to expose the reader to some of the basic concepts of the finite element method. Therefore, the material presented is introductory in nature, and confined to the basic finite elements of one- and two-dimensional problems of solid mechanics, although the developments presented for these problems apply to all field problems that are governed by the same differential equations (but may have been arrived at using different physical principles). The reader may consult references on the finite element method listed at the end of the chapter for further details (see, for example, [1-5]). The method is described and illustrated here for (1) one-dimensional second-order equations governing bars, (2) one-dimensional fourth-order equations governing Euler-Bernoulli beam theory, (3) pairs of coupled one-dimensional equations governing the Timoshenko beam theory, (4) two-dimensional fourth-order equations governing the classical plate theory, and (5) sets of coupled two-dimensional equations governing the first-order plate theory. While the general procedure is the same for all five classes of problems, the details differ from each other and therefore it is best to consider the five classes of problems separately. The equations governing these problems closely resemble the equations governing a variety of field problems, e.g., heat transfer, ground water flow, electrostatics, and so on, and thus the developments are readily applicable to those problems as well. 9.2 FINITE ELEMENT ANALYSIS OF BARS 9,2.1 Governing Equation From earlier discussions, the governing equation of a bar is ~Tx iaJx) + CU " fM* ° < X < L' {9A) where u(x) is the axial displacement, a{x) = E{x)A{x) the axial stiffness (E is Young's modulus and A is the area of cross section), c = c(x) the. surface resistance (zero in most problems), and / = f(x) is the body force per unit length. We are interested in determining an approximate solution that satisfies Eq. (9.1) and appropriate boundary conditions in some sense, say in an integral sense as in the variational methods. In developing the finite element model, we wish to consider ail possible boundary conditions that a bar can be subjected to.
FINITE ELEMENT ANALYSIS OF BARS 435 1 m -V * = 0 a2= a2{x) <73=constant -V-A i k ■ ^ JC = A 3 = L Figure 9.1 A stepped composite bar with an end spring. In order to account for all possible problem cases, we allow the data, a, cy and /, to be continuous or discontinuous in the domain (0, L). For example, a tapered, stepped composite shaft has discontinuities in its area of cross section as well as material properties at a finite number of points along the length (see Fig. 9.1): a(x) = a\(x), 02(*)> 0 < x < X{y X[ < X < X2, (9.2) a„(x), xn~\ <x <x„ = L. Also, we must allow for various combinations of the two possible boundary conditions that we derived for bar problems in Chapters 5-7: it = specified or a— + ku = specified, dx (9.3) where k is the spring constant. Thus, in a general case, we must seek approximate solution to Eq. (9.1) in each of the subintervals of (0, L) in which the equation has continuous coefficients (i.e., a, t\ and / arc constant or continuous functions of x). A step-by-step procedure for the finite element analysis of Eq. (9.1) for arbitrary variation of a, c, and /, and arbitrary boundary conditions, is given below. 9.2.2 Representation of the Domain by Finite Elements In the finite clement method, the given domain (0, L) is divided into a number of subdomains or intervals, called finite elements. This is necessitated by one or both of the following reasons: (1) It is easier to represent the solution u(x)> irrespective of the degree of its variation with x, by a polynomial of a desired degree over each element than to use a single polynomial to approximate u(x) over the entire domain. For example, a linear polynomial over each element may be used to adequately represent a cubic or higher-degree variation of it with x (sec Fig, 9.2). Obviously, the greater the number of elements, the smaller is the error between the true solution and the piecewise linear solution. (2) The actual solution is defined piecewise because of the geometric and/or material discontinuities.
436 THE FINITE ELEMENT METHOD u(p) A Figure 9.2 Piecewise linear approximation of the solution u that is cubic. eth element Global node numbers Element node numbers X* <*••:* 12- n » ^ :•*,**/*' 4-.* * o <>* a«<j> tf A] 7*« Jill 12 3-- rt-1 n -^ /te fr- (a) Finite element mesh (b) A typical finite element Figure 9.3 Subdivision of the domain into finite elements, and a typical finite element. The collection of elements is called \he finite element mesh (see Fig. 9.3). The number of divisions is analogous to th& number of parameters in the traditional variational methods. Therefore, the larger the number of elements, the more accurate the solution will be. The minimum number of subdivisions is equal to the number of subdivisions created by the discontinuities in the data (i.e., loads, material properties, and geometry) of the problem. An element or interval in one-dimensional problems is a line of finite length. The elements in a finite element mesh are connected to neighboring elements at a finite number of points, called global nodes (see Fig. 9.3). The word "global" refers to the whole problem as opposed to an clement. The end points of individual elements are called element nodes, and they match with a pair of global nodes. The relationship between the element nodes and global nodes will be discussed below. 9.2.3 Approximation over an Element Consider an arbitrary element, Q* — (xat xt,)> located between points x = xa = x\ and xh ^ x% in the domain. We isolate the element and study a variational approximation of Eq. (9.1) using any one of the variational methods presented in Chapter 7. In the classical variational methods, we seek approximation of the form N u(x) %5Ic,-<fcU) (9,4) »=i
FINITE ELEMENT ANALYSIS OF BARS 437 over the whole domain and determine relations among a such that Eq. (9.1) is satisfied in the weak form or weighted-integral sense over the entire domain. In the finite element method we seek a solution of the form u(x) « ue(x) = J2 <#(*>, with #(*) = xh (9.5) /=i over an element, and determine relations among a by satisfying Eq. (9.1) in a variational (or weak form) sense over each element. Since the elements are connected to each other at the nodes, the solutions from various elements connected at a node must have the same value at that node. In order to enforce this continuity (or uniqueness) of solution, it is convenient to express the solution over each element in terrns of its values at n nodes of the element: u€{x) = £c?0?(*) = X>f>fto. (9.6) /=! r=l where uet is the value of ue{x) at the zth node (i.e., ue[xi) = «p. Hence, by definition, the functions rfrf satisfy the interpolation property Vrf (JcJ) = *y. (9.7) That is, 1//7 is unity at its own node (i.e., the ith node) and zero at all other nodes. If there are n nodes in the element, a polynomial with n terms is required to fit the solution at the n points. Thus, each ij/f is a (n — .1) degree polynomial. For example, when n = 2, we have ue(x) = c* + c\x. Using the definition ue(xf) — w?, we obtain W^IJCi ) == tti ^ Ci -J- C-jXx , WglX'j I —— lln ^ C'-i -f- C<}X<2* Solving for (cf, c|) in terms of («j, «£), we °btaM1 ,1 HI 1 1 u,a lA 1 c2 = 4 Substituting for c^ and c^ into Eq. (9.8), we obtain „tW = 4+^ = <^#I + (^l). XX —X (9.8) = (f^)": + (^)"|E|"ww' <9-9)
438 THE FINITE ELEMENT METHOD where *"-(!^)- *<»-{&■ <9,0) Clearly, irf(x) satisfy the interpolation property in Eq. (9.7) (see Fig. 9.4). Thus the nodal values u€. take the place of the undetermined parameters while -ff take the role of the approximation functions in the traditional variational methods. Alternatively, we can determine yjrf(x) using the interpolation property (9.7). For example, for linear interpolation, the functions have the properties Since V'f and t/t| must vanish at x\ and x\, respectively, we may write them as V\ M = A(x- xe2), Irfa) = B(x- x\), where the constants A and B are determined so that ifrf and y/| are unity at jcJ and xf, respectively: rx (jef) = 1 = >l(jcf — *!), *|(**) = ! = *(*! " *')• The above equations give A = -l/he and B — \/he, where he denotes the length of the element, he = x\ — x\ = xeb — xfr Hence the approximation functions irf(x) are given by ViW = r2M = he ' which ai*e the same as those listed in Eq. (9,10). The interpolation functions \frf are defined over the clement Qe, and they are zero outside the element. Hence they are said to have compact support. The interpolation in which only the function alone is interpolated—and not its derivatives—is known as the Lagrange interpolation, and the corresponding functions are termed the Lagrange interpolation functions. It is convenient to express the interpolation Figure 9.4 Linear interpolation functions.
FINITE ELEMENT ANALYSIS OF BARS 439 functions in terms of a coordinate system fixed in the element. For example, we may take the origin of the element coordinate, x, at node i such that x=x+x\. (9.11) Then the interpolation functions (9.10) can be expressed in terms of x as ^(i)=l-~, *!<*) = f. (9.12) On the other hand, if f is the element coordinate such that £ = -1 when x = 0 and £ = 1 when* = he, i.e., * = y(l+*), (9.13) the ^ (x) of Eq. (9.12) take the form itf(f) = I (1 - £) , ^|(f) = 1 (1 + f), M < f < 1. (9,14) The coordinate £ is known as the normalized or natural coordinate. For values ;i > 2 we must identify additional points or element nodes in the interior of the clement. These points, in principle, can be any distinctly different points of the domain. For example, when n = 3, the third node can be placed anywhere other than at the ends. The optimal location is the one that is equidistant from both end nodes, i.e., the midpoint of the element. When n = 4, the one-third points are selected for the two interior nodes. The (n — l)-degree Lagrange interpolation functions (for an element with n nodes) arc defined, for arbitrary location of the nodes, by As specific examples, the quadratic and cubic Lagrange interpolation functions are given below. Quadratic (n= 3) „ X ~*** X i X """" X'x o ~~ ■'*' i X ™~ X^ r\ „e ve ,,e ^c ' Y2 _.e „e ve ,.e' ^3 *Y"c ... . V Y* "V* T* . V* "V* _ V V V ~Y*^ "V** (9.16a) tbic (n = 4) Xi Xn Xt Ao Ai Xa . A ~~ Ai A ~" Ao A ' ' X.\ ft = -—> ~ -, A Ai A X-y X Xn ' " -a | A«} At An A4 ~ X\ A^ A%, JC4 " - X€ x4 -A -A -A (9.16b)
440 THE FINITE ELEMENT METHOD For equally spaced nodal points, these functions have the following simpler form, especially when expressed in terms of the local coordinate x (see Fig. 9.5): Quadratic Functions «w-4s('-c)- ««-£(-!;)■ Cubic Functions (9.17) (9.18) «»-(-£)0-£)(-c)- «<*-»£(-£) K) • ««-s(-S)(i-S)- Note that the interpolation functions Vf to (i = 1, 2y..., «) depend only on the geometry and position of the nodes in element Q? = (xa>Xb) = {xeL, xf}); they do not depend on the solution or the boundary conditions of the problem. The functions rfrf satisfy the interpolation property (9.7). In addition, wc note that *f W + ^1 to + " ' + tn to - 1 or J2ff(x) = 1, (9.19) i=i 3 tl 2. , x=x i + & i- /j ■ i -i (a) ^. (b) Figure 9.5 Quadratic interpolation functions.
FINITE ELEMENT ANALYSIS OF BARS 441 ui —b*»;::>- » •».,«•• ..*..:— 1 2 3 — n-\ n (a) Element nodal forces .P. &*» K-> 12 3- n-\ ?i (b) Element nodal forces Figure 9.6 (a) A typical finite element with n nodes and nodal displacements, (b) A typical finite element with n nodes and nodal forces, which is sometimes referred to as "the partition of unity." This property is the direct result of including the constant term in the approximation (9.8). For example, if the solution is constant, say ce, over the entire element, then u\ = u€2 — • • • = uen = cc\ and we have de = Ytu^f(x) = c£J2iff(x) -> l=£>fO). i=i j=i /=! 9.2.4 Weak Form Having established a systematic way of deriving the approximation functions needed for the Ritz solution over an element, we now turn our attention to developing the weak form of the governing equation (9.1) over the domain Qe — {xa, xi>) = (x*t xf}) of a typical element. A typical element with n nodes and nodal displacements is shown in Fig. 9.6a, while Fig. 9.6b contains the n-node element with forces (i.e., a free-body diagram of the element). The forces and displacements at the end nodes are defined by / du «(*f) = «f, — pe «(*«) = "«' aTx) (9.20) The nodal forces P2, P%,..., P^ arc externally applied (i.e., known) forces, if any. The variational statement for the bar element in Fig, 9.6 is provided by the principle of minimum total potential energy, 8Tle — 0: -tr EeAe (du\ c€ 2 /« dx ~ P{ it £ — P2 U2 Peue = [Ab (EeAe^^ + ceSu u - fSu) dx - V P^uek> (9.21) where S is the variational symbol, and the subscript e on the variables indicates that the variables are defined in element Qe.
442 THE FINITE ELEMENT METHOD Since the model equation (9, l) also arises in fields other than solid and structural mechanics, it is informative to discuss the procedure by which we can obtain the weak form (9.21) from Eq. (9J) directly. As discussed in Chapter 7, we use the three-step procedure to construct the weak form. We have o=Cw[-T^£)+c-fh fx»[ dwdu 1 , T / du\\Xb = I a——-—|- cwu — wf \dx — \w ■ [ a-r- I L L dxdx J L V dx)\Xa rx»\ dwdu i ^4r / du\yi = I a-:—;—hcwu - wf Idx - > \w • { a — I 4 I dxdx Jj ^ I dx)\xl =£ [a^fx+cwu -wf]dx - g u;(^)p*' k (9.22a) where x* — xa,xen — xt>, and (-&-* clu dx = Pt (*=2 n-l)t (•£).-* (9.22b) and [-]p denotes the jump in the enclosed quantity at point P. Equation (9.22b) is the same as Eq. (9.21) with the weight function w replaced by 8u and a = EA, 9.2.5 Finite Element Equations We seek approximation of u(x) in the form u(x) % u€(x) = J^ *)Vj (•*)■ w = 5»(A") ** X! KV'f (*)> (9-23) where ty€. are the approximation functions derived earlier; they can be linear (n = 2), quadratic (n = 3), or higher (/? > 3). Substitution of Eq. (9.23) into Eq. (9.21) yields fXb Jx0 °<A-(p»it) (pm^(p<>"*) it*,* -/(£>?*] /? dx-J^PfSu'; ;=1
FINITE ELEMENT ANALYSIS OF BARS 443 -|>? (£[£' {^-T^^hh-iy*- *-« =X>f /=! Y,KM-f?-P?\ U=i J (9.24) Since 8u\, Su^ .. -, Suen arc arbitrary and the above equation must hold for all / = 1,2,.,,, ji, we obtain where 7=1 ;=i (9.25a) ht dtie dM \ Ee(x)Ae{x)^r~^ + ceirffej dx, ft = r M*wfdx = [ e M*wfmd** (9-25b) dx dx The coefficient matrix [Ke\ is called the stiffness matrix, and {Fe} = {ff) + [Pf] is the force vector. Equation (9.25a) is often referred to as the finite element model of the differential equation (9,1), and it provides n linear algebraic equations relating n nodal values ue.y (j = i, 2,,,,, n), The coefficient matrix [Ke], which is symmetric, and the source vector {fe} can be evaluated for a given element type (i.e., linear, quadratic, etc.) and element data (ae, ce> and fe). For element-wise constant values of ae, ce, and fe, the coefficients Kfj and ff can easily be evaluated. For linear and quadratic elements, these matrices are presented below. Linear Element For a typical linear element of length he = Xb — xQ, we have [K€] = ~ he r i [-i -r i_ 6 '2 1] J. 2 J in = /?*« f i i ae — EeAe. (9.26a) (9.26b) lfa = ae-x and c = cet the coefficient matrix L^*] for a linear element can be evaluated as ^-nmu -\ + ■ efle 2 1 1 2 (9.27)
444 THE FINITE ELEMENT METHOD where (A'pA'f) are global coordinates of node 1 and node 2 of the element £2* =(xa,xb) = (jcf, jc|). The reader should verify this. Note that [Ke] in Eq. (9.27) is the same as that in Eq, (9.26a) with ae replaced by the average value -*avg = ~(*i +4W For example, in the study of bars with linearly varying cross section A but constant modulus of elasticity £, we have a{x) = EA(x) -•(■ A\ + ¥lzA-x where x is the local or element coordinate with origin at node 1, and A\ and A% are areas of cross section at nodes 1 and 2, respectively. Then the element stiffness matrix is the same as that of a constant cross-section bar with the cross-sectional area being the average of the two ends, AaVg = (A\ + Ap/2. When a(x) — ae = constant, and f(x) = fe = constant, and ce = 0, the finite element equations corresponding to the linear element reduce to Oe_ he ' 1 -1 -1 1 I - fehe fl 1 + (9.28a) or it* _rA = _ O he fehe + Pf, ■fr+fc-fa+x- (9.28b) Xb - xa = x\ - x\, we have 3he 7 -8 -8 16 1 -8 + ■ ement Cc*lc tt* = (Xa,Xb) "4 2-1" 2 16 2 -12 4 — {«£j , '^3)1 "C ~~ (9.29a) {/*} = /eA«f (9.29b) For arbitrary variation of the data ae> cey and fey numerical integration may be used to evaluate the coefficients Kf- and // (see Reddy ]4]),
FINITE ELEMENT ANALYSIS OF BARS 445 9.2.6 Assembly (or Connectivity) of Elements The finite element equations (9.25a) can be specialized to each one of the elements in the mesh by assigning the values of xa, xy, ae, ce, etc. Because each of the elements in the mesh is comiected to its neighboring elements at the global nodes, and the displacement is continuous from one element to the next, one can relate the nodal values of displacements at the interelemcnt connecting nodes. To this end, let Uj denote the value of the displacement u(x) at the 7th global node. Then we have the following correspondence between Ui and ue- (see Fig. 9.7a) of a linear element mesh: u\=Uu u\ = it] ~ £/2, ■ u2 = "l+1 - ^H-l. ' ' ■ > U2 = UN + U (930) where N is the total number of linear elements connected in series. Equation (930) relates the global displacements to local displacements and enforces interelement continuity of the displacements (see Fig. 9.7a). The assembly of clement equations is based on the satisfaction of the principle of minimum total potential energy for the whole system: *+l ATT or dUi / = 1,2....,W+1, (9.31) Exact solution, u($) Finite element solution, u.^) Pf. ' 1 eth element 2 (e + l)st element 3 (a) he+x *• Global node numbers 2 P| J>«-1 e+l 2^2 (b) Figure 9.7 Connectivity of elements in one dimension, (a) Element-wise linear approximation of the displacement and interelement continuity of the displacements, (b) Balance of element forces at common nodes.
446 THE FINITE ELEMENT METHOD where n is the sum of Ue for e = 1,2,..., N: N 2 n = Enc = EE»HiE^-^ e=i e=\ i = l y=i = h E [«TWj«f + *?2«2 " 2^) + "2(^21"J + *22«2 ~ 2*2)] e=l = |[«|(*:>,«{ + ^^ -2F/) + uftKlu\ + A^k* _2Ff) + ... + 4{Kitu\ + K2\u\ - 2F2') +"2(*!l«J + KiA-Wl) + • ■ 0 = $[£/,(X^l/i + /r,'2£/2 - 2F\) + U2{K2nU2 + Kf2U3 - 2F,2) + ... + U2(K2\U{ + K\2U2 - 2F2') + U3{KlxU2 + KlzUi - 2F22) + ■•■], (9.32) where, in arriving at the last line, the correspondence in Eq. (9.30) is used to replace the element nodal displacements by the global nodal displacements. Using Eq. (9.32) in Eq. (9.31), we obtain KluUX + \{K\2 + K2\)U2 + ■ • ■ - Fi = 0, F2 - F\ = °. +2(^12 + ^21)^3 + ■ s(*?2 + ^21)^2 + (4> + K3n)U3 + ...-F2-F?=0, (9.33a) F»=0, 2(^12 + ^21)^ + *&</*+! - F« = 0. Using the symmetry of the element stiffness matrices, Ke.. = Kfj, and the fact that [Fe] = [fe] + {Pe}, we can write the above equations in the matrix form U2 I/3 *l\ ^12 0 0 0 k\2 Kii + Xu KYl 0 0 0 ^12 ^22 + ^11 K22 0 k\2 -•+JCJ? A12 0 0 0 K\2 K22 0N
FINITE ELEMENT ANALYSIS OF BARS 447 f rt 1 /2'+/l2 fN ■ + < Pi + pf y P^-' + if pN it is clear from Eq. (9.33b) that the diagonal elements of stiffness matrices of elements Qe and Qe+i add up at the common global node / = e + 1, and the global stiffness coefficient Kjj is zero if global nodes / and J do not belong to the same element. Thus the resulting global stiffness matrix is not only symmetric, but is also banded, i.e., all entries beyond a diagonal parallel to the main diagonal, below and above, are zero. This is a feature of all finite element equations, irrespective of the differential equation being solved, and is a result of the piecewise definition of the coordinate functions. Keeping in mind the general pattern of the assembled stiffness matrix and force column, one can routinely assemble the element matrices for any number of elements. The assembly procedure for a general case is based on the following two requirements: 1. Continuity of the primary variable(s) at the i nterelement boundary, as expressed by Eq. (9.30). 2. Balance of secondary variables; i.e., the secondary variables from the elements connected at a global node should add up to the value of the externally applied secondary variable at the node. The second condition for the mesh of two linear elements shown in Fig. 9.7b requires where JVj-l is the value of externally applied force at node e + 1 (see Fig. 9,7b). These conditions require the addition of the second equation of element Qe to the first equation of element Qe^] so that we can replace P£ + P^1 with Pe+[. This reduces 2N equations to N + 1 equations, where N is the number of linear elements connected in series, as shown in Fig. 9.7a, In general, if the ith node of element £2e is connected to the jth node of element &, the balance of secondary variables requires P? + pf = FK, (9.35) where K is the global node number of the i th node of element Qe, which is the same as the jth node of element Qf.
448 THE FINITE ELEMENT METHOD 9.2.7 Imposition of the Boundary Conditions Equations (9.33b) contain more variables than the number of equations, prior to applying the boundary conditions of the problem. In general, for the model problem in Eq. (9.1), we must iaiow the value of either the primary variable u or the secondary variable a(du/dx) at each node, including the boundary nodes. In all bar problems, we have either a displacement u or a force a(du/dx) specified at a point. If the displacement is specified at a global node, the corresponding nodal displacement should be replaced by the specified value. To impose a specified force, recall that Ff consists of contributions due to the disUibuted force f(x) and internal force a(du/dx) at the nodes [see Eq. (9.25a)], Ff = ff + Pf, where Pf and ff are the components defined in Eqs. (9.22b) and (9.25b), respectively. At any node where the displacement is unknown, the externally applied point force at the node should be equal to the sum of internal forces from all elements connected at the node. To illusttate this point further, consider a bar fixed at the left end, mo = 0, and subjected to distributed force / throughout the length of the bar and a point load Pq at the right end, The fact that no point loads are specified at the intermediate nodes implies that P* + p.2 = 0, ft2 + /T=0 P' "l + /f = 0. (9.36) Then Eq. (9.33b) becomes k\2 + k\y k\2 0 0 0 0 = < 1 f? + ff f2N~[ + f? L f2N + ?o t > - « 0 0 k\2uq 0 0 ... » 0 " 0 K)2 A22J ' u2 (9.37) which consists of N equations for the AT unknowns, Uj {I — 2,..., N + 1 )♦ Since the right-hand column is completely known, one can use any standard solution procedure to solve the linear algebraic equations.
FINITE ELEMENT ANALYSIS OF BARS 449 9.2.8 Calculation of Reactions and Derivatives of Solution; Postprocessing The displacement at any point can be computed using the equation ue{x) = mi (a-), 0<,y < hu h[ <x< h2, (9.38) un(x), htf^\ <x<h where ue(x) is the finite element solution of (9.23) in element £le. For example, if w(*n), h[ < xq < h2i is desired, for a linear element we have u2(xQ) = u^irl2\xo) + u^ir^ixo) = U2is{2)(x0) + t/atffW (9.39) Similarly, the derivatives of ue(x) can be computed from due ~dx~ £? dx I>5— 7=1 dx " dV- 7=1 AT 0 < x < h\, /*i < x < h2i /l/y„l < X < hflj. (9,40) It should be noted that the nodal values of the derivative computed from the two elements sharing that node do not coincide, irrespective of the order of the elements, because the continuity of the derivative is not imposed in the procedure. Therefore, one must further process these values to assign a single value at the node (e.g., the weighted average of the values of the derivative from all elements connected at the node). The unknown reaction forces at any global node can be computed after the nodal displacements Uj are computed. They can be computed from either (a) the definition (9.22b) or (a) the element equations (9.25a). For the problem at hand, the first equation of element i can be used to determine the reaction force at node 1 (U[ = uq): Alternatively, P/ can be computed using the definition (9.41a) ''-(-"£L
450 THE FINITE ELEMENT METHOD The two values for Pt! from Eqs. (9.41a) and (9.41b) are the same only when the distributed force f(x) is zero. The difference reduces as the number of elements is increased (i.e., ff gets smaller in magnitude as the number of elements is increased). This completes the description of the basic steps in the finite element analysis of the model problem. The derivation of the coordinate functions is not a usual step in the finite element formulation because these functions are already derived and are available in textbooks. One should only select appropriate coordinate functions for the problem at hand. The derivation is included here to illustrate the procedure. Although we used the terminology of solid mechanics in the development of the finite element equations, all the concepts are applicable to any physical problem described by Eq. (9.1). In other words, if a computer program is developed to solve Eq. (9.1) for any boundary conditions, then the program can be used to analyze not only bars but all other physical problems described by the equation (e.g., transverse deflection of cables, heat transfer in fins, flow through pipes, etc.; see Table 8.2). For additional details, the reader may consult the textbook by Reddy [4]. In the following example, we illustrate the steps involved in the finite element analysis of a composite bar. Example 9.1 Consider the composite bar consisting of a tapered steel bar fastened to an aluminum rod of uniform cross section, and subjected to loads as shown in Fig. 9.8. We wish to determine the deformation in the bar using the linear elements. The governing equations are given by d (' x du\ -—\EsAs—\ = 0, 0<* </?i, dx \ ax J dx \ dx ) (9.42a) i < x < h\ +Ii2 = L, Stee1[a, = #,4,(*)] Aluminum <a2= EaAa) ^Element® Element® Global node -^— ^ 1 \ 2 \ 3 number (c) Po §; -^ 2|1V-2P0 ^ ' Pi I /t,= 961n.,A2= 120 in. (b) ® Element nod number p\ _\ pi n A ^...,,.. ,.,„, „. „. Pi + P? = 2P(1 Pt-Po (d) Figure 9.8 Axial deformation of a composite member, (a) Geometry and loading, (b) Finite element representation, (o) Continuity of displacements, (d) Balance of forces.
FINITE ELEMENT ANALYSIS OF BARS 451 where the subscript s refers to sleel and a to aluminum. We need not worry about the boundary conditions until step 4 of the analysis. We take the following values of the data: Es = 30 x 106 psi, As = (ci + C2X)2, Ea = 107 psi, , 2 (9.42b) 4*= I in.2, /2i=96in„ £, = 216 in., P0 = 10,0001b. Wc will follow the steps in the finite element analysis of the problem. In actual practice, all the steps can be carried on a digital computer. 1. Finite Element Mesh. From the discontinuity in the material property, area of cross section, and loading at x — ft i > we are required to divide the domain Q = (0, L) into at least two elements: Q} — (0, h i) and fi2 — (ft [, L). For the mesh of two linear elements, there will be three global nodes; if quadratic elements are used, then there will be five global nodes. 2. Element Equations, For each element, whether linear or quadratic, the element stiffness matrix and force vector are given by Eq. (9.25b), with a€ = Ee(a\ + a\x) , ce - 0, fe = 0. We now compute \Ke] using the linear interpolation functions. The element stiffness matrix and force vector can be computed using either the local coordinate x or the global coordinate x. If the local coordinate is used to evaluate the coefficients, then the transformation from x to x (x = x + xe) should be used to convert all functions of x to functions of x. We have A|2 = A21 - -iin, A-22 " *ll> or
452 THE FINITE ELEMENT METHOD where [we have used the algebraic equalities, a2 — b2 = (a - b) (a + b) and a3, - b3 — (a-b)(a2 + b2 + ab)] Ae = (af) -h -(flf) (*«+i + *e + *<?*«+!) + ^2(^+1 + *e)- (943b) More specifically, we have {a[ = 1.5, a^ = —0.5/96, a^ = 1, a2 — 0, h[ — 96 in., A2= 120 in.): (9.44) 3, Assembly 0/ Element Equations. For the two-element mesh we have from Eq. (9.33b) the assembled set of equations [*'] = ^ " ">' m-is"10'! r 1 [-1 1 -1 -1" -i" 1 |. n- {F2} = hi r*f *?2 0 ^21 ^22 +^11 *12 0 4, ^22J t/2 U3 >a> pf+p, d(2) (2) For the problem at hand, these have the explicit form r 4J5 _4J5 n 96 96 u 10' 4.75 4.75 , _1_ 1 96 96 -1" 120 120 L _L 120 120 J 0 ^1 U2 t/3 ,0) »(') .(2) (2) 4. Imposition of Boundary Conditions. From Fig. 9.8, we have «(0) = Ui = 0, P2(1) + P,(2) = 2P0, P2(2) = P0. Using the boundary conditions, Eq, (9.46) can be written as 104 49.479 -49.479 0.000 -49.479 57.812 -8.333 0.000 -8.333 8.333 i [°1 u2\ = < u*\ r P(2)) 1 2Po . ^o J 5. Solution of Equations, From Eq. (9.47), we have 10' [ -3 812 333 -8.333 8.333 _ {2P01 _ f20,0001 -|p0J "iio.oooj (9.45) (9.46) (9.47) or U2 = 0.06063 in., U3 = 0.18063 in. (9.48)
RNITE ELEMENT ANALYSIS OF BARS 453 6. Postprocessing of Results. The reaction force at the left end of the bar is given by [either by definition (9.41b) or from the first equation of (9.46)]: p[x) = 104(49.479C/l -49.479£/2) = -30,0001b. The negative sign indicates that Pj ^ is acting opposite to the convention used in Fig. 9.8, and therefore the reaction is acting away from the end (i.e., tensile force). The magnitude of p[ * is consistent with the static equilibrium of the forces: Pl(1)+2P0+Po-0 or >(D ^" = -3^0 = -30,000 lb. The axial displacement at any point x along the bar is given by ue(x) = u\l)f\]) +u{2[)f^) =: 0.06063^/96, 0<x< 96, IufV{2) + 42)V'22) = -0.03537 + 0.001*, , 96 < x < 216, (9.49a) and its first derivative is given by due _ f 0.06063/96, 0 < x < 96, dx "" J 0.001, 96 < x < 216. (9' ^ } The exact solution of Eq. (9.42a) subject to the boundary conditions I£(0)=0, is given by (4) -(4) = 2P0, u(x) = 0.I28*/(288-*), 10.001(jc-32), ch^_ (36.864/(288-;c)2, dx "10.001, 0 < x < 96, 96 <* <216, 0 < x < 96, 96 <* <216. (9.50a) (9.50b) In particular, the exact solution at nodes 2 and 3 is given by u(96) = 0.064 in., w(216) = 0.1840 in. Thus the two-element solution is about 1.8% off from the maximum displacement. Next consider a two-element mesh of quadratic elements. The element matrix and force vector for element 1 are [K]] = 104 142.19 -159.37 17.18 -159.37 266.67 -107.29 17,18 -107.29 90.10 n= P(»i 0 d(0
454 THE FINITE ELEMENT METHOD Table 9.1 Comparison of the finite element solutions with the exact solution of the bar problem in Example 9.1 x (in.) Linear Quadratic Exact l + la 2+1 3 + 2 4 + 2 6 + 2 1 + 1 2 + 1 16 24 32 48 64 72 80 96 156 216 0.00753 0.01600 0.01600 0.02560 0.03657 0.04267 0.04923 0.06400 0.12392 0.18400 0.00752 — — 0.01161 0.01164 — 0.02532 0.01593 0.03638 0.02553 — 0.04253 0.06063 0.06309 0.06359 0.18063 0.18309 0.18359 0.06377 0.12377 0.18377 0.01598 0.02557 0.03652 0.04916 0.06390 0.12390 0.18390 0.02572 0.02560 — 0.04268 0.06392 0.12392 0.18392 0.06399 0.12399 0.18399 *m + n means m elements in the interval (0, 96) and n elements in the interval (90, 216); all elements in each interval arc of the same size. The assembled stiffness matrix is of order 5x5, and it is of the form [K] = r*i\ Kb K3l 0 _ 0 Ku *22 ^32 0 0 K\3 ^23 *33 + *U K21 K2 A31 0 0 K2 A12 ^22 ^32 0 I 0 K2 A13 K2 *23 ^33J After imposing boundary conditions {U{ - 0, P\ + P\ = 20,000, P2 = 10,000), and solving the resulting 4x4 equations, we obtain U2 = 0.02572 in., U3 = 0.06392 in., U4 = 0.12392in„ U5 = 0.18392in. Obviously, the two-element solution obtained using the quadratic element is very accurate. A comparison of the finite clement solution obtained by various meshes of linear and quadratic elements with the exact solution is presented in Table 9.1. 9.3 FINITE ELEMENT ANALYSIS OFTHE EULER-BERNOULLI BEAM THEORY 9.3*1 Governing Equation Here we consider the finite element formulation of the fourth-order equation governing the bending of elastic beams according to the Euier-Bernoulli beam theory
FINITE ELEMENT ANALYSIS OF THE EULER-BERNOULLI BEAM THEORY 455 q(X) w0,z TT T T T M T TTTTTT TTT' (a) Xb- -M e-th element - xtt~ J J WsPM 8 1 N JV+1 1 2 e (b) Primary variables (c) Secondary variables Figure 9.9 Finite element representation of a beam, (a) Geometry and loading on a straight beam, (b) Finite element mesh, (c) Beam element with primary (displacement) and secondary (force) degrees of freedom. (see Examples 4.3 and 5.1). The equation governing the transverse deflection is given by[seeEq. (5.35)] dx2 V £4?)-«=°- 0 < x < L, (9,51) where wq denotes the transverse deflection, EI the bending stiffness, and q the distributed transverse load (see Fig. 9.9a). A finite element mesh of the beam is shown in Fig. 9.9b. The basic steps in the formulation are the same as those discussed for bars, but the specific mathematical details differ from those of second-order equations. Since the basic terminology of the finite element method has already been introduced in the preceding section, we go straight to the approximation over an element (see Example 5,6). 9.3.2 Weak Form over an Element A finite element for a beam is again a free-body diagram of a portion, Qe = (xa, xt,), of a typical beam, as shown in Fig. 9.9c, with nodal displacements (primary variables)
456 THE FINITE ELEMENT METHOD and forces (secondary variables), As discussed earlier, the choice of the nodal variables is dictated by the variational formulation. From the total potential energy principle for beams, we know that the essential boundary conditions involve the specification of wq and dwo/dx: du}0, , e , v dm. -—(xa) = -ir2, w0(xb) = «3, —— ax ax (9.52) Wo(xa) = U\, ~]—(xa) = ^'2> U>o(*&) = "3, ~T~^X^ ~ ~U4 and the natural boundary conditions involve the specification of bending moments and shear forces [see Eq. (4.33b) for the moment-deflection relation and Eq, (4,30c) for the moment-shear force relation; sec also Fig. 9.9c for the sign convention]; [=("£)L-^ ("£)L--* <953b) The derivation of the weak form of Eq, (9.51) was discussed in Chapter 7, Following the same procedure, we obtain the following weak form over an element: -ew*)-as (-£)„• <9'54a> where v is the weight function. The total potential energy functional for the beam element is given by -jcN n*(wo) = / I ^^r ( ~£r) ~ Woq* I clx - Q\u\ - g'tt* -' fi|«f - G5«5t (9,54b) where the sum X)/=i 2f "f represents the work done by the forces £^' and Q\ and moments 2^ anc^ Q\ m moving and rotating the ends of the element through displacement it\ and W3, and slopes ue2 and u\. The weak form (9.54a) is the same as the statement of the principle of miniirium total potential energy, 8Ue — 0, with v = 8wo* 9.3,3 Derivation of the Approximation Functions To derive the coordinate functions for the Ritz approximation of the function wo, we must first select a polynomial that is continuous and complete up to the degree required. It is clear from the functional Yle that the function we select should be twice- differentiable to yield nonzero strain energy, and it should be cubic to yield nonzero shear forces Q\ and Qe2. Another reason for selecting wq to be a cubic polynomial is
FINITE ELEMENT ANALYSIS OF THE EULER-BERNOULU BEAM THEORY 457 that we need four parameters in the polynomial in order to satisfy all four essential boundary conditions in Eq, (9.52). Thus, a minimum degree polynomial is cubic (see also, Example 5.6): wo(x) % «o + &\x + aix2 + <*3X3> (9,55) where x denotes the local coordinate. Next, select the parameters ao, &{> #2, and #3 such that u>o satisfies the essential boundary conditions, Eq. (9.52): it\ = wj0(0) = ofo, ax (9.56) W3 = wo(he) = ao + a\he + ajhl 4- 0:3/1;!, «4 = —TT^e) = -«i - 2a2/7e - 3a3/^, where h.e—x^-xa> Solving Eq. (9.56) for ao, «i> «2> and as in terms of uUi = 1, 2, 3,4), and substituting the result into Eq. (9.55), we obtain (9.57) where the approximation functions <f>^(x), x=x—xe, for the beam element are given by *fm-i-3(£)2+2(i)3, <t>%{x) = -x he) \ (9.58) The approximate functions <pf are called the Hermite cubic interpolation functions. We note the following properties of (p?, i = 1, 2, 3, 4 (see Fig. 9.10): 0far-l t^) = ^' ^2a (*fi) = 0> Yl *2or-1 = °> a=l (9.59) **L- 2ff-l dx (xp) = 0, -^C*/J) = **/>. (a, 0 = 1, 2),
458 THE FINITE ELEMENT METHOD Figure 9.10 Hermite cubic interpolation functions for the beam element. where x\ — 0 and X2 = he are the local coordinates of nodes 1 and 2 of clement Qe = (xe, xe+\), and xe is the global coordinate of global node e. It should be noted that by including the slopes u^ and ueA as nodal quantities, we will be able to enforce the slope continuity at the interelement boundaries during the assembly process. Of course, the total potential energy formulation of the governing equation yields slopes as part of the essential boundary conditions. All dependent variables and their derivatives that enter the specification of the essential boundary conditions of a problem always end up as the nodal variables, and they take the role of the undetermined coefficients of the Ritz method. 9.3.4 Finite Element Model To derive the Ritz equations corresponding to the functional in Eq, (9.54), we substitute Eq. (9.57) for we into Ue, and set its derivatives with respect to each ue., j = 1,2, 3,4, to zero: Ue"(«|, «2, u\, ttj) = / Xh=Xe+\ Xc E«5 d2<p<j ^e^e 2 X^'1 dx2 E»7tfW \i=\ dx -Hq>i 4 4 irr*i(r'E,,M^X -iED* i=l J=\ dx2 dx2 ' I ■' - X>? (£**'*# ** + #) , = 1 ^e 4/4 98 5 E E«f^-2^ > ;=i y=i (9.60a)
FINITE ELEMENT ANALYSIS OF THE EULER-BERNOULLI BEAM THEORY 459 3TT 3«f 7 = 1 (9.60b) where In Eq. (9.60c) it is understood that 0? are expressed in terms of the global coordinate jc. In the element coordinate i, Eq. (9.60c) takes the simple form, Kfj - j EeIe-^^-dx, Ff = j qe<j>f dx + Ql (9.60d) where Eey fe, and qe are the transformed functions, Ee = Ee(x), Ie = Ie(x)i qe = q(x). For the case in which Ee> Ie, and q€ are constant over the element Q€, the stiffness matrix [Ke] and force vector {Fe} are given by \K£} = - - 6 -3Ae -6 _-3Ae -3ft, 3he hi -6 3fte 6 3fte -3Ae"| /i2 3he 2/t2 J 1 ' 12 f 6 ] -he 6 I /*e 1 • + ' fen Ifij] (9.6 h . 1) where the label e on the variables is omitted in the interest of brevity. One can show that the first column of the force vector represents the "statically equivalent" forces and moments at the nodes due to uniformly distributed load q§ over the element. This element gives, for element-wise constant values of EI, the exact values of wq and dwo/dx at the nodes for any load q(x)> and thus it is called a superconvergent element. 9.3.5 Assembly of Element Equations The assembly procedure for the beam clement is analogous to that described for the bar element. The difference is that in the beam element there are two unknowns, called degrees of freedom, per node. Consequently, at every global node thaL is shared by two elements, the elements in the last two rows and columns of /th element overlap with the elements in the first two rows and columns of (/ + l)th element. For a
460 THE FINITE ELEMENT METHOD Element node Global node numbers numbers , * N © (b) Figure 9.11 Assembly of two beam finite elements, (a) Continuity of generalized displacements, (b) Balance of generalized displacements. three-element case, the assembled force vector and stiffness matrix are shown below (see Fig. 9.11): {*"} = + e3- (3) (3) (9,62a) [*H = *!, K. K K 41 0 0 0 0 K[2 Kn K32 K42 0 0 0 0 *13 K22> ^33 + ^11 ^43 + ^21 *31 *4*1 0 0 *!4 *24 ^3l4+^l22 44*~ 22 *32 ** 0 0 0 0 *?3 *l> 4j + *11 at43+a:21 *31 ^41 0 0 A14 ^24 ^34 + ^12 ^44 + ^22 ^2 *& 0 0 0 0 Kn K23 ^33 ^43 0 0 0 0 4t 34 a: ^44. (9.62b)
FINITE ELEMENT ANALYSIS OF THE EULER-BERNOULLI BEAM THEORY 461 9.3,6 Imposition of Boundary Conditions The boundary conditions on the generalized displacements can be imposed in the manner described earlier for bars. The boundary (or equilibrium) conditions on forces are imposed by modifying the second column of the assembled force vector [see Eq. (9.62a)]. If, for example, the force at the 7th global node is specified to be Fq> and the moment at Kth global node Ls specified to be Mo, then the following assembled coefficients get modified: Qi~l + Q\ - fi). Q*~[ + Qi= Mo. (9.63) If no force or moment i s specified at a nodal point that is unconstrained, it is understood that the force and moment are specified to be zero there. The solution and postcomputation of the displacements and their derivatives at various points of the beam follow along the same lines as that described for bar problems. We now consider a specific example to illustrate the steps involved in the finite element analysis of beams. Example 9.2 Consider the indeterminate beam shown in Fig. 9.12. The beam is made of steel (E — 30 x 106 psi) and the cross-sectional dimensions are 2 x 3 in. (7=4.5 in,4). We are interested in finding the deflection wo and its derivatives using the Euler-Bernoulli beam finite element. Because of the discontinuity in loading, the beam should be divided into three elements: Q] - (0, 16), £22 = (16, 36), and Q3 = (36,48); the element lengths are; h\ = 16 in., /i2 — 20 in., and A3 = 12 in. <?"fc) ql7\x) 20 lb/In. 30 lb/in, w, z 4 ± TTTTTTTTTTTTTTi W I^MI 16 in. 20 in. 12 in. R> = 500 lbs. \ © J © \®\ t/1=o u3 u5=o u7 Ql Qi + QNOQl + Q? Ql=F0 QV £ Figure 9.12 Finite element modeling of an indeterminate beam.
462 THE FINITE ELEMENT METHOD The element equations for all three elements arc readily available from Eq. (9.61). The only exception is that the load vector for element 1, which has linearly varying distributed load, must be computed. The load variation is given by 70)M -(*>-£)■ qi2)(x) = 20. /<3>(A-) = 0, and their contribution to the element nodes can be computed using Eq. (9.60d). For element 1, we have q} = j (30-^x\4>}(x)dx (» = 1,2,3,4). Wc have {«'} = 216.00 -554.67 184.00 512.00 k2) = 200.00 -666,77 200.00 666.67 {<H - {0}. (9.64) The element stiffness matrices can be computed from Eq. (9.61) by substituting appropriate values of/*, £\ and /. For example* for element 1 we have (h[ = 16 in., EI = 135 x 106 lb.-in.2): [K[] = 106 0.3955 -3.1641 -3,1641 33.7500 -0.3955 3.1641 -3.1641 16.8750 -0.3955 -3,1641 3.1641 16.8750 0.3955 3,1641. 3.1641 33.750 The assembled matrix for this mesh is of the same form as given in Eq. (9.62b). The boundary conditions for the problem are tuo(0) = 0 -► U[ = 0, ax 0, uj0(36) = 0 -> Us = 0, 63 + 61=°' 64 + 62 = °. 64 + 62=0, fi*=500, Ql = Q. (9,65) Note that Q\t Q\, and 63 -f 6 j are the reactions that are not known, and are to be calculated in the postcomputation. Since the specified essential boundary conditions are homogeneous, one can delete the rows and columns corresponding to the specified displacements (he,, delete rows and columns 1,2, and 5) and solve the remaining five equations for U3, U4, £/$, Ujy and U%: t/3 = -0.000322 in. J77^ 0.00515 in,, U4 = 0.0000593 rad, Uz = -0.000518 rad. t/6 =-0.0002513 rad, (9.66) As stated before, these nodal values are exact for any number of elements.
FINITE ELEMENT MODELS OF THE TIMOSHENKO BEAM THEORY 463 The exact solution to the problem can be obtained by integrating the moment- deflection expression (d2wo/dx2) — —M/EI [see Eq. (4.33b)J, in which the bending moment can be readily obtained in terms of the reactions at the supports and applied loads. The deflection in the three intervals of the beam is given by (large numbers were rounded to whole numbers): wq(x) = L(26S.543x2-— *3 + :^J_ A 0<jr<16, 11 \ 15 4 192 ) ~ ~ 1 / 16384 5120 ccnnnn2 m 3 5 A „ + _^^x +55.2097^2 + -—x2 + 7x4 y 16 <x <36, <A \ 3 3 15 6 / ^^6554368-506066*+I2000x2-^*3), 36<jc<48. E J El It can be verified that the nodal values in Eq. (9.66) are exact, A comparison of the finite element results for deflections, slopes, and bending moments (calculated in the postcomputation) at points other than the nodes arc compared with the exact values in Table 9.2 for three different meshes. The values of deflections, slopes, and bending moments were computed using the finite element solution (9.57) and its derivatives. 9.4 FINITE ELEMENT MODELS OF THE TIMOSHENKO BEAM THEORY 9.4.1 Governing Equations The displacement field of the Timoshenko beam theory (sec Problems 4.26, 5.4, and 6.15, and Examples 6.4 and 7.10; see also Rcddy [5-7]), for the pure bending case, is u{x, z) — zcp(x), v = 0, w(x, z) — woO), (9.67) where wo is the transverse deflection and <p the rotation of a transverse normal line about the v-axis. The strains and stresses of the Timoshenko beam theory arc d<f> . . dw0 _ _ e.Yx = z— = ZKxx* YxZ = <P + —r-» <*xx = E£xx* <Txz = Gyxz- ax dx (9.68) The balance of internal moments and transverse forces give the relations M(x) = f zoxx dA = EI^f Q(x) = KS f axz dA = GAKS U + ^ J , (9.69) and the equilibrium of moments and trans verse forces give G-^0. f = -qM. (9.70) dx dx
464 THE FINITE ELEMENT METHOD Table 9.2 Comparison of the finite element solution with the exact solution of the beam problem considered in Example 9-2 x 2.0 6.0 12.0 21.0 31.0 42.0 /Va 3 5 6 3 5 6 3 5 6 3 5 6 3 5 6 3 5 6 U)Q X FEM -0.8570 4.1405 5.2319 -18.4510 8.3936 9.4751 -138.2300 -122.5900 -122.5900 -674.3600 -643.4900 -643.4900 -812,9300 -782.0700 -782.0700 2174.9000 2174.9000 2174.9000 106 Exact 5.3739 9.6048 -120.8100 -639.6300 -778.1900 2174.8000 (-dw0/dx) FEM 1.1553 -3,3386 -4.1558 8.8348 4.4199 5.2323 33.7760 39.6630 39.6630 71.5620 62.3030 62.3030 -83.7970 -74.5370 -74.5370 -451,3600 -451.3600 - -451.3600 x 106 Exact -4.1552 5,2328 39.6720 62.3020 -74.5390 451.3600 ~{d2wQ/d. FEM -1.0250 0.4663 0.4638 -2.8147 -4.3456 -4.3431 -5.4992 -5.4794 -5.4794 3.5507 3.5507 3.5507 27.5210 27.5210 27.5210 22.2220 22.2220 22.2220 x1) x 106 Exact 0.3219 -4.4727 -5.9238 2.9335 26.9040 22.2220 a3-Elements h\ = 16, h2 = 20. A3 = 12. 5-Elements: h\ == 8> h2 = 8, A3 = 10, A4 = 10, A5 = 12. 6-Elements: A { = 4, h2 - 4, A3 = 8, A4 = 10, A5 = 10, A6 =12. Here, q (x) is the distributed transverse load, E Young's modulus, G the shear modulus, A the area of cross section, / the moment of inertia, and Ks the shear correction factor. Using (9.69) in (9,70), we obtain the following equilibrium equations: ax \ ax J \ ax J dx GAi = q(x). (9.72) In the following, a number of displacement finite element models of these equations are presented for element-wise constant properties, E = Ee> G = Ge, A — Ae, and/ = Je, 9.4.2 Displacement Finite Element Models The displacement finite clement model of the Timoshenko beam theory is constructed using the principle of minimum total potential energy, or equivalently, using the
FINITE ELEMENT MODELS OF THE TJMOSHENKO BEAM THEORY 465 weak form, 0 = jT [e.U^ + OAK, (* + ^) (* + ^) - *(,)*«*] «/, - V/5w0fe) - VbeSw0(xb) - Mea84>(xa) - MebS<t>(xb) (9.73) and -Q{xa) = G.A.K, (*£+*) Mea = -M{xu) = - [^4^1 vf} a Gfo) = [g€a,js:, (^ + <*-) = M(xb) = fa/,^] • L ^Jx=.v* (9.74) M; Suppose that too and $ are approximated as (9.75) where (WJ, <J>p are the nodal values of (wq, 4>) and (V^j, £>p are the associated interpolation functions. Substitution of (9.75) for w<$ and <f>, and Su>o = iff and 50 = p? into Eq. (9.73), yields the finite element model [Kn] [Kl2]l i{W}\ _ UF1} 12iT r]r22-|J I ($} f — 1/p2 [[Ki2]J [K2 if2}]' (9.76) where KsGeAe-^L<pefdx> fXb I dy* d<pe{ \ *i=r ^dx+w<*«>+wc**). Jx„ /f = M>f(*a) + M^f(^). (9.77)
466 THE FINITE ELEMENT METHOD Ql = Ml Q'4 = K (a) Generalized displacements (b) Generalized forces Figure 9.13 Displacement finite element for the Timoshenko beam theory. 9.4,3 Reduced Integration Element (RIE) For a linear interpolation of wq and <f> (i.e., m — n = 2; see Fig. 9.13) and exact evaluation of (he integrals of Eq. (9.77), Eq. (9.76) takes the form GeAf>Ks 6he ' 6 -6 -3he -3he' —6 6 3he 3he -3he 3he 2h]k ti*% -3he 3he h2e$ 2hjk_l < \wr \W2 M ' =' \q*\ 0 0 . ► + \ \K) Wb\ (9.78) or, rearranging the equations, we obtain 6 — 3he —6 —3he 2h^k 3he (2EeIe\ \ mo*5 / -6 -3K 3he h2A 6 3he -3he- h2eH 3he 2h*\\ "Wex M i- = -i Kl 0 0. y + < \Q\\ \Q\\ IfijJ (9.79) where Q\ = V/, Qe2 = W«, <2^ = V£, fij = M|, and qf=r"ffqdx 0 = 1,2), J.v„ Mo = i2A, $ = 1 - 6A, A = £«/« GeAeKsh2' A = 1+3A. (9.80) (9.81) Tn the thin beam limit, the effect of shear deformation must vanish, i.e., A -*■ 0. Then the first two equations of (9.79) imply the following relation among (Wf,W|,^,4>p: m+p^H. (9.82) which is equivalent to the (Kirchhoff) constraint <j> + (dwo/dx) = 0 (or shear strain yxz — 0). The last two equations of (9.79), in view of (9.82), yield the constraint he = 0. (9.83)
FINITE ELEMENT MODELS OF THE T1MOSHENKO BEAM THEORY 467 This is equivalent to d<j)/dx = 0, which is an incorrect condition to be satisfied as it forces the curvature and hence the bending energy to zero. Thus, the finite element equations in (9,79), in an effort to satisfy the constraints (9.82) and (9.83), will yield the trivial solution W* = W% = $ \ — $2 — 0- This is known in the finite element literature as shear locking. The Kirchhoff condition (9.82) suggests that wq and <p be interpolated such that dwo/dx is a polynomial of the same order as <p. If wq is approximated using a linear polynomial (a minimum requirement), then <p should be a constant. Since the minimum continuity requirement on <j> is also linear, it follows that wq be approximated using a quadratic polynomial. This is termed a consistent interpolation. Alternatively, both wo and <p appearing in the bending energy may be approximated with linear polynomials while <p appearing in the shear energy is treated as a constant. This can be achieved by using numerical integration with one Gauss point to evaluate the shear stiffness. This procedure is known in the literature as the reduced integration of the shear stiffness. It amounts to evaluating the second term of Kfj- in Eq, (9.77) with one-point integration as opposed to two-point integration required to exactly evaluate the integral. This results in the element equations 6 -3he -6 -3ft, —3he —6 —3he A2(1.5+6A) 3he h2(\.5-6A) 3he 6 3he h2e(\.5-6A) 3he h2(1.5+6A) 'Wf <J>f W| .*!. K 0 «5 0 + QV (9.84) This element is designated as the reduced integration element (RIE). In the thin (or slender) beam limit, these element equations reduce to only one constraint, namely, the Kirchhoff condition in Eq. (9.82). While the element does not lock, it does not yield exact displacements at the nodes. However, with a sufficient number of elements in the mesh of a problem, one does get a very accurate solution. 9.4.4 Consistent Interpolation Element (CIE) As suggested earlier, if we use a quadratic approximation of u>o and linear approximation of <S>, Eq. (9.76) yields the following 5x5 system of equations: GAKS 6Ju 14 -16 2 — 5he —he -16 2 5A, -he 32 -16 4he -4he -16 14 K 5A* Ahe he 2h2eX *% -Ah 5he h]S 1h\Y < W\' \w? n of m. , = , [tfl i% ?! 0 0 '+< 'v;| Qec\ n\ K\ Ml , (9.85) where (V/, V£, M^, M£) are the generalized forces defined in Eq. (9.74), W* and Qec are the deflection and applied external load, respectively, at the center node of the
468 THE FINITE ELEMENT METHOD % = M[ ef-v; <?.• qi=v£ {a) Generalized displacements (b) Generalized forces Figure 9.14 Consistent finite element model of the Timoshenko beam theory. quadratic element, and ?f=/ V'f qdx (i — 1, 2, c). (9.86) This element is designated as the consistent interpolation element (CIE) (see Fig. 9,14). The center degree of freedom W* may be condensed out to reduce the system (9,85) to a 4 x 4 system of equations. The second equation of (9.85) can be used to express W* in terms of Wf, W|, 4>*, 4>2> </*> anc* Qc: w; 6h 32G( ^^m--m -» Substituting for W* from Eq. (9.87) into the remaining equations of (9.85) [i.e., eliminating W* from Eq. (9.85)], we obtain (2EeI€\ \ 1^1 ) 6 — 3he —6 — 3he -3hc h*(\.5 + 6A) 3he A2(1.5-6A) -6 -3he 3he 6 fc*(1.5 -6A) 3fctf 3/t* ft*(l.5 + 6A) 1*1+ 2#' I -&•*« 11{ + W -neh • + ■ [Q'l ia^i (9,88) where qec = #* + Qec. For simplicity, but without the loss of generality, we will assume that Qec — 0 so that q* = qecr Note that the element has the same stiffness matrix as the reduced integration element but a different load vector. The load vector is equivalent to that of the Euler-Bernoulli beam clement [see Eq. (9.61)]. In fact, for uniform load q> the load vector in Eq. (9.88) is identical to that of the Euier-Bemoulli beam element. Note that the elimination qfWc is not possible for the dynamic case because of the presence of Wc.
FINITE ELEMENT MODELS OF THE TIMOSHENKO BEAM THEORY 469 9.4.5 Superconvergent Element (SCE) The next choice of consistent intcipolation is to use the Lagrange cubic polynomials for wq and quadratic polynomials for <f>. This will lead to a 7 x 7 system of equations. The displacement degrees of freedom associated with the interior nodes can again be condensed out, for the static case, to obtain a 4 x 4 system of equations. Here we will not consider this approach. Instead, we consider the Hermite cubic intcipolation of w\] and a related quadratic approximation of 0 (see Reddy [5,7]). The 4x4 system of equations thus obtained yield exact values of the nodal displacements for the Timoshenko beam theory, as in the case of Euler-Bemoulli beam element. Such elements are called superconvergent elements. The exact solution of Eqs. (9.71) and (9.72) for the homogeneous case (analogous to what was done in Example 5.3 in connection with the Euler-Bernoulli beam theory) is «*<*> = ~±j (alX-+a2^+a,x + a^ + ^- («„), (9, 89) x2 E14>{x)=a\—+ aix + a3, (9.90) where ai through 04 are the constants of integration. Note that the constants ai, #2, and a-*, appearing in Eq. (9.90) are the same as those in Eq. (9.89). Thus, Eqs. (9.89) and (9.90) suggest that one may use cuhic approximation of wq and an interdependent quadratic approximation of 4>. The resulting finite element is termed the interdependent interpolation element (HE). The constants a, appearing in Eqs. (9.89) and (9.90) can be expressed in terms of the nodal values of w<$ and <j> (see Example 5.5). Then Eqs. (9.89) and (9.90) take the form Af = Wf, Ae2 = Of, A| = W|, AJ = t>e2, (9.92) where \jrf and <pf are the approximation functions i tfrf = -[/i-12A»j-<3-2i,)u2], *1 = --[(1-»J)2»> + 6A(1-ij)u], V'! = -[(3-2ij)^ + 12A>j]i (9.93)
470 THE FINITE ELEMENT METHOD he\L 6 J hejJi <pl=-(3r12-2r1 + 12Ar}). (9.94) Here, i] is the nondimensional local coordinate x xc( (9.95) (9.96a) At = I + 12A. Substitution of Eq. (9.91) into Eq. (9.73) yields the finite element model [Ke]{tf} = [qe} + {Q% where qei = / irfq{x)dx9 J.xa and Q* = Vf, <2£ = Mf\ 63 = V2*' and g* = Mf. Equation (9.96a) has the explicit form [see Eq. (9.81) for the definitions of § and X]: (2EeTe\ 6 — 3he —6 —3he -3he 2h2ek 3he h*% —6 3he 6 3he -3h€ ti*$ 3he 2h2eX \Wf] Wi\ y = i fflf] «2 kl . + , en IgsI Q5I gsJ (9,97) As stated earlier, this element leads to the exact nodal values for any distribution of the transverse load q(x)> provided that the bending stiffness EI and shear stiffness KSGA are element-wise constant. In the thin beam limit, Eq, (9.97) reduces to the Euler-Bernoulli beam equations (9.61), and it experiences no shear locking. Example 9.3 Consider a simply supported beam. The following data are used: E = 10° v = 0.25, K = 40= 1. Two different beam length-to-height ratios, L/H = 10 and 100, aTe considered. Table 9.3 shows a comparison of the finite element solutions obtained with one, two, and four elements in half beam with the exact beam solutions for two different types of loads, namely, uniform load and sinusoidal load. Clearly, more than two elements
FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 471 Table 9.3 Comparison of the finite element solutions with the exact maximum deflection and rotation of a simply supported isotropic beam (A/= number of elements used in half beam) wq x 102 -0 x 103 Element N-l N = 2 N = 4 N = \ N =2 N =4 Uniform had (L/ H = 10) R1H 0.09750 0.14438 0.15609 0.37500 CIE 0.12875 0.15219 0.15805 0.50000 IIEa 0.16000 0.16000 0.16000 0.50000 EBEa 0.15265 0.15265 0.15265 0.50000 Uniform load{L/H = 100) R1E 0.09379 0.14066 0.15238 0.37500 CIE 0.12504 0.14847 0.15433 0.50000 HEa 0.15629 0.15629 0,15629 0.50000 EBEa 0.15265 0.15265 0.15265 0.50000 Sinusoidal load {L/ H = 100) RIE 0.07639 0.11079 0.12007 0.30543 CIE 0.09679 0.11682 0.12163 0.38705 IIEa 0.12322 0.12322 0.12322 0.38702 EBEa 0.12319 0.12319 0.12319 0.38702 "Exact values compared to the respective beam theories. of CIE and RIE are required to obtain acceptable solutions. On the other hand, ITE yields exact nodal values with one element. 9.5 FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 9.5.1 introduction Here we develop displacement finite clement models of the equations governing the motion of plates according to the classical plate theory (CPT). While many of the basic ideas of the finite element method from one-dimensional beam problems carry over to two-dimensional plate problems, finite element models of plates are considerably more complicated due to the fact that two-dimensional problems are described by partial differential equations over geometrically complex regions. The boundary T of a two-dimensional domain £1 is, in general, a curve. Therefore, the two-dimensional finite elements must be simple geometric shapes that can be used to approximate the geometry of a given two-dimensional domain as well as the solution over it. Consequently, the finite element solution will have errors due to the approximation of the solution as well as the geometry of the domain. In two dimensions, there is more than one geometric shape that can be used as a finite element (see Fig. 9.15). The interpolation functions depend not only on the 0.46875 0.50000 0.50000 0.50000 0.46875 0.50000 0.50000 0.50000 0.36702 0.38702 0.38702 0.38702 0.49219 0.50000 0.50000 0.50000 0.49219 0.50000 0.50000 0.50000 0.38204 0.38702 0.38702 0.38702
472 THE FINITE ELEMENT METHOD Triangular element Quadrilateral element Figure 9.15 Finite element discretization of a two-dimensional domain using triangular and quadrilateral elements. number of nodes in the element, but also on the shape of the element. The shape of the element must be such that its geometry is uniquely defined by a set of points, which serve as the element nodes in the development of the interpolation functions. A triangular element is the simplest geometric shape, followed by a rectangle. 9.5.2 General Formulation Here we consider the general case of the dynamic response of plates subjected to in-plane compressive and shear forces. The virtual work statement of the classical plate theoiy over typical finite element Qe is given by [from Eq. (8.131a), except that the time derivative terms arc integrated by parts in time]: A] Jae L Mr. - 2——MXy Z-T-My 38wo ( - dm ~ dwo\ BSwq (« dw0 * 8wq\ ~ "aT [N"TT + ""If) - IT \N"1T + N»-&) + TqSwqwo + h — — + -t ;r- ™ SwM dxdydt \ 3x dx dy dy } J +C i (~SwoVn+i!?Tx+*wT')dsdu i9M) where Tx = Mxxnx + MXytiy, Ty = Mxfnx + Myyny, (9.99a) *-(!£^-»»S-»«£-*£S)«'
FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 473 Here (nx, ny) denote the direction cosines of the unit normal on the element boundary T*, (M.v.v> Mxyy Myy) are defined by Eq. (8.120), and (Nxx, Nxy, Nyy) are the in-plane compressive and shear forces. It is clear from the variational statement (9.98) that the finite elements based on the classical plate theory require continuity of the transverse deflection and its normal derivative across element boundaries. Also, to satisfy the constant displacement (rigid-body mode) and constant strain requirements, the polynomial expansion for wq should be a complete quadratic. Let us assume finite clement approximation of the form n wQ{x, y, 0 = J2 A/(0?y(*. 30, (9.100) where Ay are the values of wq and its derivatives at the nodes, and <pe. are the interpolation functions, the specific form of which will depend on the geometry of the element and the nodal degrees of freedom interpolated. Substituting approximations (9.100) for wq and <pf for the virtual displacement 8wq into Eq. (9.98), we obtain the /th equation of the finite element model: D(^-G?>y + ««A5] = /7f (9.101) 7=1 where i>j = 1, 2,..., n. The coefficients of the stiffness matrix Kf. = K"?-, mass matrix M*. = Me>^ geometric stiffness (or stability) matrix G*. = G^, and force vectors Ff and Ff1 are defined as follows; Kfj = f D\T^X + v(7V™»' + T,fxx) + 2(1 - v)T*?xy + T?™]dxdy, M?j = fjhSf? + h{Sff + Sff)]dxdy, 7 = f 4<Pi dxdy + j> (vn<p* - Tx ^ - Ty M\ ds> (9.102a) F, where and £, ?/, f, and \x take on the symbols x and y. In matrix notation, Eq. (9.101) can be expressed as (\Ke\ - \Ge}) {Ae} + [Me]{Ae] = {Fe}. (9.103)
474 THE FINITE ELEMENT METHOD This completes the finite element model development of the classical laminate theory. The finite element model in Eq. (9.103) is called a displacement finite element model because it is based on equations of motion expressed in terms of the displacements, and the generalized displacements are the primary nodal degrees of freedom. 9.5.3 Conforming and Nonconforming Plate Elements There exists a large body of literature on tiiangular and rectangular plate-bending finite elements of isotropic or orthotopic plates based on the classical plate theory (see, e.g., [8-16]). There are two kinds of pi ate-bending elements of the classical plate theory, A conforming element is one in which the interelement continuity of wq, 6X = dwo/dx, and 0y = dwo/dy (or dw^jdn) are satisfied, and a nonconforming element is one in which the continuity of the normal slope, Swo/dn, is not satisfied. A nonconforming rectangular element has wo, 0Y, and 0y as the nodal variables (see Fig. 9.16a). The normal slope variation is cubic along an edge, whereas there are only two values of Bwo/Sn available on the edge, Therefore, the cubic polynomial for the normal derivative of wo is not the same on the edge common to two elements. The interpolation functions for this element can be expressed compactly as <pf = gii (i = 1, 4,7,10); <p* = gl2 (i = 2, 5, 8, 11), P? = g,-3(i = 3,6,9,12), Si I = *0 + &)(1 + >7o)(2 + £o + no ~ ¥ - rj1), S/2 = ^0-l)(l+>7o)(l+So)2, £/3 = ^/(W ~ 1)0 + W + W)2, £ = (x - xc)/2, n = (y- yc)/h> lb = ttt> (9.104a) (9.104b) *70 = W» where 2a and 2b are the sides of the rectangle, and (xCi yc) are the global coordinates of the center of the rectangle. u ^x dy dxdy (a) Nonconforming element (b) Conforming element Figure 9.16 (a) Nonconforming and (b) conforming rectangular elements.
FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 475 A conforming rectangular element has wo, Swq/3x, Owo/rty, and 32wo/3xdy as the nodal variables. The interpolation functions for this element (see Fig. 9.16b) are < - gn 0' = 1, 5,9,13); <pf = m (i = 2y 6, 1.0, 14), tf = gn (i = 3, 7,11,15); < - gi4 (I = 4, 8, 12,16), g/i = t^« + £/)2(?o - 2)0? + w)2(»to - 2), */2 = ^/« + l/)2(l " So)0? + »?/)2ft0 - 2), S/3 = JtMS + ft)2«0 - 2)fo + 1?,)2(1 - 170), 5/4 - T^ftW« + fr)2(l - &)0? + *?/)2(l " %). The conforming element has four degrees of freedom per node, while the nonconforming element has three degrees of freedom per node. For the conforming rectangular element the total number of bending nodal degrees of freedom per element is 16, and for the nonconforming clement the total number is 12. 9.5.4 Fully Discref ized Finite Element Models Static Bending In the case of static bending under applied mechanical loads, Eq. (9.103) reduces to (\Ke]-[Ge]){Ae} = iF% (9.106) where it is understood that all time-derivative terms are zero. Once the nodal values of generalized displacements (wq9 dwo/dx, dwo/dy) have been obtained by solving the assembled equations of a problem, the strains are evaluated in each element by differentiating the displacements. The strains and stresses are most accurate if they are computed at the center of the element [17,18]. Buckling In the case of buckling under applied in-plane compressive (Nxx, Nyy) and shear Nxy edge loads, Eq. (9.103) reduces to ([K€]-k[Gel)[*e) = {Fl (9.107) where A = Nxx/N°x = Nyy/N°, = Nxy/N%, (9.108a) Glj = f \N°xS?f + N°y(Sf? + Sff) + N°yS>?] dxdy, Fk = l(vn4-Tx^~Ty^)ds. (9.108b) (9.105a) (9.105b)
476 THE FINITE ELEMENT METHOD Natural Vibration In the case of natural vibration, the response of the plate is assumed to be periodic. Equation (9.103) becomes ([/H - o?[Me]){Ae] = [Fel (9.109) where co denotes the frequency of natural vibration. Transient Response In the case of transient response, Eq. (9.103) must be integrated with respect to time t to determine the nodal values Ay (0 as functions of time. Here we consider the Newmark family of time-integration schemes [1,3,19,201 that is used widely in structural dynamics. In the Newmark method, the first and second time derivatives are approximated as {A*W, = {A*}, +*i{A'}.? + ^{A'Wi, (9.11 Oa) {A<Wi - 03 ({A*Wi ~ {A*U - ci4[Ae}s - a5{'£'}Si where the superposed dot denotes differentiation with respect to time, 2 _, (I- Y) ci[ = (1 —ot)dts, ci2 = adts, a$ = 0 , 04 = dts 03, CI5 — —, Y(dts)2 y (9.110b) dt is the time increment, dts = ^4.1 - ts, and {-}s, for example, denotes the value of the enclosed vector at time ts. The parameters a and y in the Newmark scheme are selected such that the scheme is either stable or conditionally stable; i.e., the error introduced through the time approximation (9.110a) does not grow unboundedly with time. All schemes for which y > a > 1/2 are unconditionally stable. Schemes for which y < a and a > 0.5 are conditionally stable, and the stability condition is dt < dtcy - |><W (a - y) ]~]/\ (9.111) where comax denotes the maximum frequency associated with the finite element equations (9.109) for the same mesh. The Newmark family contains the following widely used schemes: a = \, y = \, the constant average acceleration method (stable). a = \, y = ^, the linear acceleration method (conditionally stable). a = \, y = 0, the centered difference method (conditionally stable). a = \, y = 2, the backward difference method (stable). (9.112) Using Eq. (9.110a), Eq. (9.103) can be expressed as [K<][we}s+{ = {Fe)> (9J13a)
FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 477 where [K*] = (IX'Wi - [GeU0+a3\Me]s+u {F*} = [Fe}s+i + [M€Ui {a3{Aeh +a4{A*h +a5{fr)s), (9.113b) Note that for the centered difference scheme (y — 0), it is necessary to use an alternative form of Eq. (9.113a). Equation (9.113a) represents a system of algebraic equations among the (discrete) values of [Ae(t)} at timer = f.y+i in terms of known values at time/ — /^. At the first time step (i.e., s = 0), the values {A*}0 = {A*(0)} and {A*}0 = {A*(0)J arc known from the initial conditions of the problem, and Eq. (9,101) is used to determine {A£'}o at f = 0: [Aeh = [Merl {{Fe} - (\K*] - [Ge\) (Ae}0}. (9.114) The conforming (C) and nonconforming (NC) rectangular finite elements discussed herein arc used to analyze plates for bending and natural vibration response, and the results are presented in the next two examples. We shall use the notation "m x n mesh" to indicate that m elements along the *-axis and n elements along the y-axis are used. In the case of the conforming element, \t is necessary that the cross-derivative 32wo/dxdy also be set to zero at the center of the plate when a quarter-plate model is used, otherwise the results will be less accurate. The stresses in the finite element analysis were computed at the center of the elements. Example 9.4 Consider a rectangular isotropic (v = 0.25) plate with simply supported edges, and subjected to uniformly distributed load of intensity q$. Due to the symmetry, a quadrant of the plate may be used as the computational domain. Full integration (i.e., 4x4 Gauss rule) was used to evaluate the stiffness coefficients in both elements, although the nonconforming element gave the same results for 3x3 Gauss rule. The geometric boundary conditions of the computational domain (see the shaded quadrant in Fig. 9.17) are dwo dwn —- = 0 at x = 0; —- = 0 at y = 0, (9.115a) dx dy w0 = ~^0 at a- = -; wo = ~^ = 0 aty = -. (9.1I5b) ax 2 ay 2 In addition, 32wo/Bx3y = 0 was used at x = y = 0 for the conforming element. The boundary conditions (for CPT) along the symmetry lines are shown in Fig. 9.17. Table 9.4 shows a comparison of finite clement solutions with the analytical solutions developed in Chapter 8. The series solutions were evaluated using m,« = ], 3,..,, 19. The exact maximum deflection occurs at x = y = 0, maximum stresses axx and cryy occur at (0,0, h/2), and the maximum shear stress axy occurs at {a/2, b/2, —h/2). The locations of the maximum normal stresses are (a/8, b/S),
478 THE FINITE ELEMENT METHOD b 2 b 2 «<.-g*o n «^°=0 ^■©•-o ":^ -r-^iJ^Computational domain Symmetry conditions: |£°=Oats = 0;j^0=oat,y = 0 Figure 9.17 Symmetry boundary conditions for rectangular plates. Table 9.4 A comparison of the maximum transverse deflections and stressesa of simply supported (SSSS) square plates under uniform load Variable w &XX (Tyy &xy Nonconforming 2x2 4.8571 0.2405 0.2405 0.1713 4x4 4.6425 0.2673 0.2673 0.1964 8x8 4.5883 0.2740 0.2740 0.2050 2x2 4.7388 0.2259 0.2259 0.1669 Conforming 4x4 4.5952 0.2637 0.2637 0.1935 8x8 4.5734 0.2732 0.2732 0.2040 Analytical Solution 4.5701 0.2762 0.2762 0.2085 *w = 102lu>QE2h3/(qoa4)l a = ah2/(q0«2). Table 9.5 Maximum transverse deflections and stresses" of clamped (CCCC), isotropic (> ~ 0.25) square plates under uniform load Variable wb &xx dyy *xy 2x2 1.5731 0.0987 0.0987 0.0497 Nonconforming 4x4 1.4653 0.1238 0.1238 0.0222 8x8 1.4342 0.1301 0.1301 0.0067 2x2 1.7245 0.1115 0.1115 0.0700 Conforming 4x4 1.5327 0.1247 0.1247 0.0297 8x8 1.4539 0.1305 0.1305 0.0086 au; = lQ2[w0E2h*/(qoa4)\t d = crh2/{q{)a2). b The analytical solution (see Example 8.12) is 1.4231. (a/16, 6/16), and (fl/32T 6/32) for uniform meshes 2 x 2, 4 x 4, and 8 x 8, respectively, while those of axy are (3tf/8> 36/8), (7a/\6, 76/16), and (15a/32f156/32) for the three meshes. Similar results are presented in Table 9.5 for a clamped square plate. Reduced integration (i.e., 3x3 Gauss rule) was used to evaluate the stiffness coefficients in both elements. The locations of the normal and shear stresses reported for the three
FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 479 Table 9.6 A comparison of the fundamental frequencies1 of simply supported (SSSS) and clamped (CCCC) rectangular plates a/b 0.5 1.0 1,5 2.0 2x2 4.8301 1.9736 1.4144 i.2074 SSSS Plates 4x4 4.9752 1.9959 1.4395 1,2439 Exact 4.9003 1.9838 1.4359 1.2436 2x2 7,8213 3.3103 2.3435 1,9551 CCCC Plates 4x4 9.1185 3.5203 2.5861 2.2781 Rilza 9.9891 3.6476 2.7404 2.4973 *<b^co(b2/n2Wph/D22. b Rilz — Ritz solution discussed in Chapter 8. meshes are: a-^2x2: (a/8,A/8); 4x4: (a/16,6/16): 8x8: (a/32,A/32); r^2x2: (3«/8,3*/8); 4x4: {la/\6Jh/\b)\ 8x8: (15a/32,15A/32). These stresses are not necessarily the maximum ones in the plate. For example, for an 8 x 8 mesh the maximum normal stress in the plate is found to be 0.2316 at (0.46875a, 0.03125*, -A/2) and the maximum shear stress is 0.0225 at (0.28125a, 0.09375/?, -A/2) for the nonconforming element. Both conforming and nonconforming elements show good convergence. Since the stresses in the finite element analysis are computed at locations different from the analytical solutions, they are expected to be different. A mesh refinement not only improves the accuracy of the solution, but the Gauss point locations also get closer to the true locations of the maximum values. Example 9,5 Next consider natural vibrations of isotropic rectangular plates. Table 9.6 contains fundamental frequencies for isotropic (v = 0.25) simply supported and clamped plates. The results were obtained using the conforming plate-bending element. Only a quadrant was modeled and rotary inertia was included (b/h — 0.1). Because a quadrant is used to model the problem, only symmetric vibration modes can be calculated. 9,6 FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 9.6.1 Governing Equations and Weak Forms Here we develop the finite element models of the equations governing the first-order shear deformation plate theory (FDST). We consider the linear equations of motion of FSDTfrom Eqs. (8.282)-(8.284) that are expressed in terms of the stress resultants (MXXi Myy% Mxy, Qx, Qy) defined in Eqs. (8.288) and (8.289). These equations are
480 THE FINfTE ELEMENT METHOD expressed in terms of the generalized displacements Oo? <px, <t>y) m Ec!s> (8.290)- (8.292). The weak forms of Eqs. (8.282)-(8.284) are given by w0 dSwo S-wq Qx 4- — Qy - Sw0q + IoSwo—pr- '38wq ~BJc °-(\7 • (""17 + ""17 J + U7" {"»-* + "»V J J *"" -0 (2.v«v + Qyny) Swq ds, (9.116a) (9.116b) 0 = /, ( 17^ + ^TMyy + s<pyQy + Wvll/dxdy ~~ &.TyS<py ds' C9.116c) where ?i = Afr.r^ + Af,^, 7> = MA7nA. 4- Myyny. We note from the boundary terms in Eqs. (9.116a-c) that (wq, </>x, 4>y) are the primary variables and (Qn> TXt Ty) are the secondary variables, where Qn = Qxnx + Qyny. (9.117b) Unlike in the classical plate theory, the rotations (<px, <py) are independent of wq. Note also that no derivatives of wq are in the list of the primary variables and therefore the finite element to be developed is called a C° element. 9.6.2 Finite Element Model The dependent variables (wq, <px,(j>y) can ail be approximated using the Lagrange interpolation functions, and, in principle, wq and (<pXi <f>y) can be approximated with differing degrees of functions. Let n mix, y, t) = ^2wj(t)fej(x, y), (9.118a) 7=1
FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 481 fr7 ? * : clL± -7-* "••:.':7'.-. ■ =.- 5' • Tiff 6* M 4 >■ 8-■-'.' V-'.- :■ : 7 ': •:'©:'-.:':-; ■v': 9 ;;■■;.:. 3 6$ (a) Bilinear element (b) 8-node quadratic element (c) 9-nodc quadratic element Figure 9.18 Linear and quadratic rectangular elements. 4>Ax>y,0 = yEfS}{t)<pej(x9y), /7 (9.118b) (9.118c) where (\jfe., <pej) are Lagrange interpolation functions, and (wj , £j, 5"?) are the nodal values of (u^o, 4>x>4>y\ respectively. One can use linear, quadratic, or higher-order interpolation functions for each of the variables. Here we use rectangular elements. The linear and quadratic Lagrange interpolation functions of rectangular elements (see Fig. 9.18) are given below in terms of the element coordinates (|, ??), called the natural coordinates. The subscripts of these functions ^\ correspond to the node numbers shown in Fig. 9 J 8. For a derivation of these functions, see Reddy [4,32]. Bilinear Interpolation Functions (9.119) [*fl n n \fi\ i 4 r(i-|)(i-„)| (1 + 1X1-;?) (l+|)(i + )?)| 1(1-§)(!+>?) J Nine-Node Quadratic Interpolation Functions n n n (l - $)(1 -17)<-£ - ij - l) + (l - l2)d - n2) (l + $)<i -17)<§ - n - i) + (i - £2)(i - >?2) (1 +|)(1 + »?)(£ + >? - 1) + (1 -£2)(1 - V2) (l - £)(i + ,?)(-£ + J? - i) + (i - t2)(i - n2) 2(1 - |2)(1 - )?) - (1 - £2)(1 - ,j2) 2(l+|)(l->72)-(J.-?2)(l-,?2) 2(l-?2)(l+)7)-(l-^2)(l-J?2) 2(l-£)(l-}72)-(i-£2)(l-)72) 4(l-f2)(l-?/2) (9.120)
482 THE FINITE ELEMENT METHOD (9.121) Eight-Node Quadratic Interpolation Functions (]-£)(J-)7)(-£-??-l)l (l + £)(l-ij)(!-!?-l) (l+f)(l + »7)« + i/-l) (1-1)0 +»?)(-? + »j-l) 2(l-£2)(l-i,) 2(l+£)(l->72) 2(l-f2)(l+»j) 2(l-£)(l-„2) Substituting Eqs. (9.118a-c) for (u;o, <f>x, 4>y) into Eqs. (9.116a~c), we obtain the following semidiscrete finite element model of the FSDT: [Kn] [K12] {Kn]~l [Kl2]T [K22] [K73] [Kn]T [K23f [K*] \*v ki m M Hi \n Wi m\ i \G] [0] [0]"| [0] [0] [0] [0] [0] [0]J \ / [{«/}] }{Sl)\ {S2> [Mu] [0] [0] [0] [M22] [0] [0] 1 [0] [M33J < f{«»e] {S1} {S2} \{F1} {F2} {F3} {0} (9.122a) or [Ke]{Ae} + [Me]{Ke} = [Fe] The coefficients of the submatrices [KaP] and \MaP] for (a, y3 = 1,2,3) by the expressions (9.122b) and vectors [Fa} are defined K = / KSA55^—L + KsA^^—L dxdy, J Jo? \ dx dx dy dy J 42= / KsA55^p-<pej dxdy, K}f = / KsAjp-<pejdxdy, J Jw ox J J Jo* dy J dxdy, (9.123)
FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 483 (a) Bilinear FSDT element (b) Quadratic FSDT element Figure 9.19 C° rectangular plate-bending finite elements for the first-order shear deformation plate theory. A*}} = f Mf** dxdy, Mff = f htftf dxdy = Af* J or Jw Ff] = f tjirfdxdy+f Qnffds, Ff=i Tx<pfds> Ff = <L Tytfds. Here Djj are the bending stiffnesses and A44. and A55 are shear stiffnesses of an orthotopic plate: A>6 = Erf3 12(1 » V12U21)' G]2h3 D22 = E2h3 12 12(1- V[2v2\) ^44 = G23h, A55 = G13A. D[2 = v\2D22l (9.124) When the bilinear interpolation functions are used with tyf = <p? for all generalized displacements (wo, 4>x><PyX the element stiffness matrices are of the order 12 x 12; for the nine-node quadratic element they are 27 x 27 (see Fig. 9.19). The C° plate-bending element of Eq. (9.122b) is among the simplest available in the literature. It is known, when lower-order (quadratic or less) equal interpolation of the transverse deflection and rotations is used, that the elements experience shear locking (similar to the Timoshenko beam element). A commonly used technique to alleviate the problem of shear loclcing is to under-integrate the transverse shear parts of stiffness coefficients (i.e., all coefficients in Kf? that contain A44 and A55). To better understand what is reduced integration in the present context, we visit briefly the concept of numerical integration using the Gauss-Legendre quadrature. For alternative finite element models of the FSDT, the reader may consult the papers [21-26] listed at the end of the chapter. 9.6.3 Numerical Integration Numerical integration schemes, such as the Gauss-Legendre numerical integration scheme, require the integral to be evaluated on a specific domain or with respect to a
484 THE FINITE ELEMENT METHOD • •frv;-- -« Qe, Physical element Master element Figure 9,20 Transformation between the domain on which an integral is defined (actual element domain) and the domain on which the integral is evaluated. specific coordinate system. Gauss quadrature, for example, requires the integral to be expressed over a square region £2 of dimension 2x2 and for the coordinate system (£, ?/) be such that —!<($, r\) < 1. The coordinates (£, rj) are called normalized or natural coordinates. Thus, the transformation between (xt y) and {£, r/) of a given integral expression defined over an element W to one on the domain Q facilitates the use of Gauss-Legendrc quadrature to evaluate integrals. The element Q is called a master element (see Reddy [4], chap. 9). The transformation between Qe and Q is accomplished by a coordinate transformation of the form (see Fig. 9.20) (9.125) j=i while a typical dependent variable u{x, y) is approximated by 7=1 7=1 (9.126) where ifr* denote the interpolation functions of the master element Q and ij/j are interpolation functions of a typical element Qe over which u is approximated. The trans formation (9.125) maps a point (x, y) in a typical element Sle of the mesh to a point (£, rj) in the master element £2, and vice versa if the Jacobian of the transformation is positive-definite. The positive-definite requirement of the Jacobian dictates admissible geometries of elements in a mesh (see Reddy [4], pp. 421-448). The interpolation functions xj/j used for the approximation of the dependent variable are, in general, different from tM used in the approximation of the geometry. Depending on the relative degree of approximations used for the geometiy and
FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 485 the dependent variable(s), the finite element formulations are classified into three categories: (a) Superparametric formulation (m > n). The polynomial degree of approximation used for the geometry is of higher order than that used for the dependent variable. (b) Isoparametric foundation (m = n)t Equal degree of approximation is used for both geometry and dependent variables. (c) Subparametric formulation (m < n)> Higher-order approximation of the dependent variable is used. For example, the nonconforming and conforming finite elements are based on higher- order interpolation (e.g., cubic or higher order) of the deflection wq while the geometry is represented by linear or quadratic interpolation functions. Hence the formulation falls into the subparametric category. The first-order shear deformation plate-bending elements presented here are based on the isoparametric formulation. Next, consider the evaluation of the integral /*= [ Fe(x,y)dxdy. (9.127) The integrand Fe (x, y) in the present case consists of the elastic coefficients Df. and A(jj and the interpolation functions t//f and their derivatives. Hence, if D]-} and A*. are constant within the element Qe, Fe{x, y) is a polynomial of x and y of a certain degree, depending on the degree of iff. When a Gauss rule is used, the integral in (9.127) is expressed in terms of the natural coordinates (£, 77) as -/ Fe(Xiy)dxdy Fe(x($,ri),y($,ri))\J<\dSdii F($>ri)d$dTi, (9.128) where \Je\ is the determinant of the matrix of coordinate transformation, called the Jacobian: [■/'] = dx dy dx dy di] di] J (9.129) For straight-sided elements | Je | is a constant, and the polynomial degree of F€ (§, t}) is the same as that of Fe („v, y).
486 THE FINITE ELEMENT METHOD "-ff--| .«-ff ff, ' *-ff L^. A-- i I i T\=0 [f----$— --ff- Figure 9.21 Gauss points in a master rectangular element. The numerical evaluation of Ie by the Gauss quadrature amounts to evaluating the integrand at a number of points in &e, called Gauss points, multiplying the function values with so-called Gauss weights, and adding them up: (9.130) where (f/, r]j) are the coordinates of the (7, J) Gauss point in Qe, Wj and Wj are the corresponding Gauss weights, and (M, N) are the number of Gauss points in each coordinate direction (see Fig. 9.21). If F(£, r}) is a polynomial of degree p in £ and degree q in r}7 then the choice M = > + l" W = tf+1 (9.131) gives the exact value of the integral Ie. Here [•] denotes the nearest integer equal to or greater than the enclosed value. For example, for the C° elements based on isoparametric formulation, the interpolation functions are of the same degree in both £ and r]. Hence, for the evaluation of the stiffness coefficients, we have M = N\ and p = q are at most equal to twice the degree r of the polynomials (of £ or i\) in V'/Cf»*i)> This can be seen by examining the coefficients Kf? (a — i, 2, 3), which contain the products yfrfi/rj. Thus we have M T2r + 1 "1 r = order of the element. (9.132) Note that r — 1 for the linear element and r — 1 for the quadratic element. The Gauss rule M x N means that M Gauss points in the §-coordinate direction and N Gauss points in the ^-coordinate direction. If N x N is the full Gauss rule, then (N — 1) x (N - 1) is the reduced Gauss rule. Equation (9.122b) can be simplified for static bending, buckling, natural vibration, and transient analysis, as described for the classical plate element in Section 9.5.4. The simplifications are obvious and are therefore not repeated here. Numerical results for bending, buckling, and natural vibration will be discussed next.
FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 487 Example 9.6 This example is designed to illustrate the effect of full and reduced integration Gauss rules used to evaluate the stiffness coefficients on the deflections and stresses. We consider a simply supported, isotropic square plate under the (sinusoidal) load 7VX 7iy q{x>y) = qocos — cos —, (9.133) a a where the origin of the coordinate system is taken at the center of the plate, -a/2 < x < a/2, -a/2 < y < a/2, and -A/2 < z < h/2. Because of the biaxial symmetry of the solution, the quadrant 0 < x < a/2 and 0 < y < a/2 is used as the computational domain. We use the notation nh for n x n uniform mesh of linear rectangular elements, nQS for n x n uniform mesh of eight-node quadratic elements, and «Q9 for n x n uniform mesh of nine-node quadratic elements in a quarter plate. Three different Gauss rules are used to evaluate the stiffness coefficients: 1. Full integration (F). All stiffness coefficients [K01^] of Eq. (9.123) are evaluated using the integration rule N x N that would yield their exact values. 2. Selective (mixed) integration (S). All terms of the stiffness coefficients except those containing A44 and 455 are evaluated using the full integration rule, and the terms involving A44 and A55 are evaluated using the reduced integration rule. 3. Reduced integration (R). All stiffness coefficients [KaP] of Eq. (9.123) are evaluated using the reduced integration rule (N — J) x (N — 1). Various stresses in the finite clement analysis are evaluated at the reduced Gauss points as indicated below: VxAa, A, -I (TyylA, A, ^V crxy(Bt B, --V axz(B,A) atz = ±^, cryz(A,B) atz = 0, where the values of (A, B) are given in Table 9.7. Table 9.7 The Gauss point locations at which the stresses are computed in the finite element analysis Coord. Exact 2L 4L 8L 2Q8/2Q9 4Q8/4Q9 A 0.0 0.125a 0.0625« 0.03125a 0.05283a 0.02642a B 0.5a 0.375a 0.4375a 0.46875a 0.44717a 0.47358a
488 THE FINITE ELEMENT METHOD The maximum deflection and stresses obtained for the problem are presented in Table 9.8. The following nondimensionalizations are used: w = iVQ(Ay A) x 10 D crqo h\ h2 9 ' azq0 *»-a»{A'A4)^' &xz = -o>z(5, A) , aqo &xx ^Vxx{A,A, -\ Gyz = -ayz{AtB) . aq0 (9 J 35) An examination of the numerical results presented in Table 9.8 shows that the FSDT finite element with equal interpolation of ail generalized displacements does not experience shear locking for thick plates even when the full integration rule is used. Shear locking is evident when the element is used to model thin plates {a/h > 100) with the full integration rule (P). Also, higher-order elements show less sensitivity for locking but with slower convergence. The element behaves uniformly well for thin and thick plates when the reduced (R) or selectively reduced integration (S) rule is used. It is clear that the effect of shear deformation is to increase the deflections. It should be noted that the nondimensionalization in Eq. (9.210) is such that the CPT solution is independent of the side-to-thickness ratio. Higher-order elements or refined meshes of lower-order elements experience relatively less locking, but sometimes at the expense of rate of convergence. With the suggested Gauss rule, highly distorted elements tend to have slower rates of convergence but they give reasonably accurate results. Plates composed of multiple layers are called laminated plates {see Reddy [27] and Fig. 9.22). A laminated plate composed of multiple orthotropic layers, with the material axes (x\, X2) of each layer oriented at either 0° or 90° to the plate axes (x, y\ has the following bending moment-deflection relationships: Mxx Du Dn 0 Dn D22 0 0 " 0 D<%_ \sL *!,- k (9.136) where el are the bending strains, defined by Eq. (8.117) for the CPT andEq. (8.276b) for the FSDT The laminate bending stiffnesses Djj are defined by n Dv-iEc/W+i-^- (9.137a) k=l Here N denotes tht total number of layers in the laminate, zk is the z-coordinate of the bottom of the /cth layer (measured in the positive ^-direction), and Q\^ are the
09 Q) CO CL E CD 14 ■* sr s « e?t 18 ■— o ona (0/9 « JB C CO 2 c I E T3 CO c — O TS C © <1> fc £& c a. o = — w s* •^ ct £ S 1? « c o T*lte a> <a o <u o zi o c 3= <^ W » c oo o 43 ^ to to |2 T3 Ol c ■D CO "co o to 3 r 'w o 1 T> a> n 3 a) n «b ib ^ •^ •b :< ib •si i> P 3 o ^ <3 in i/5 Tf vo oo in "nj- ^ en cN m so !> Tj- C- cn 1— m vn 00 00 on Tt m cN r-- i> in m os os OO CO OO 00 00 OOO OOO OOO OOO OO m in "d- sooo r- ^r r* o^m cn r- -3- -*t cn in id ui cn«»««rr) 1-- v> m o\ o^ in *n \D so so o in oc 00 ocoooo 0000 odd odd odd odd do os in tn r- cn 00 cn t^-oo soocoo so o c— ,000c o o os oooo 000 poo oj —1 —« odd odd odd O so vO cn CN 000 cn cn 00000 0 r- — 00 0. cn — cN d d 0 r- —1 0C o\ m —1 <N d d §§ d d »n ^f m O m -d- m —H d Os O 8 d sO xr in O so Tf V^ —" d in t> 00 0 d -3- cn so O ^d- m •so «—< d 0 Os 00 O d r- sO m 0 r- sO in »— d m m 8 d oc m so O 00 m NO *-H d vn 00 Os O d Os •n sO O Os \n sc —• d 0 Os s d m 00 0 cn ««H 00 •—H d CN -3- 8 d ^t 00 0 2 oc —* d 00 0 0 d r- cn oc 0 i> cn 00 r—t O cn r^ O d so Tt 00 0 so ^ 00 T— O cn 00 2 0 ■^r in 00 O ^r in 00 •—H O SO 0 d 0 m 00 0 O m oc —1 d SO O ^2 d ON 0O O Os *-m 00 —H d cn CN d SO Os 00 0 SO Os oc —^ d CN m d 0 r- —■ 00 as cn -H CN d d O r- H OC os m —■ CN 0 d 0 0 "* ^r d d ■^toccn cNmoc o — 00 Tj-moo o ^ cn n 00 x r-^r^- Tfsosc -vf r-* r- o ,— ~-« ^ ^ _ _ _ _- odd odd odd -^foocn cn in 00 so •— 00 -^rmoo o ti- n cnoooo r*- •* "* Ttscso "d-r-r- O —I —« -H-H—• 1-. 1 odd odd odd cn —1 •«- — o in osOcn oor-cN sccn-h r-r-00 c^fvnso ^ i> r*- h o \C -^CNCN CNCNCN CNCNCN odd odd odd CN oc — 0 IN 0O T—< O CN 00 SO CN 55 00 '-* 0 Tf ^ 00 WM 0 cn O r- CN 00 i-H O ptoe if— O cn O r- CN r^ 00 "—• 0 _H r- 00 •»H O r- a\ sO CN SO OC 00 —< 0 so 00 00 T—< 0 CN O r- CN 3 t: ■C^ 1 p^, Anal 006 1—1 0 0 0 Os •—• 0 CN O r-- CN O O 0^ —' O O O Os. 1-^ O sO sO in CN ^ cs ^ 1 ement &i •— m 0 0 0 m 0 O 0 SO CN O O oc cn m 00 -t tj- **" T—' O O 00 cn m 00 ^r ^t- •^— 1—1 0 0 >— CN cn r*- ^- ^f CN CN cninoo o h oc ^h^^ r* so CNxhcn h- 00 oc cn -tf- ^- i>oo cnsoso or--r- r-0000 oooo odd odd odd do cninoc o^^oo — -^ ^t -—so cN^cn r-oooo cn "«t •* r-00 cnsosc or-r- r-0000 oooo odd odd tzi ^S (z> <z> d> odd do cn cn <— -it, in cn 00 00 o cn rt »-h m m —< m m CN CN CN O CN CN odd odd **fr Os Os CN oc OS sO so sO so -^t- ir< m m m CN CN CN CN CN d cS <5 c5 c5 CN CN CN 1 Ph ftS Cr? OS OS OS 33 ^ ^- -rf "^t PU P^ W 1 j l Os Cs Os CN CN CN co a? 00 313 CN cN CN IlHpfCO tlH^c/} ^^^ CO ^^_,^. "* TJ- Tf cNcNCN 00 to 80 o Os OS d d §§ OS Os 00 SC VO so m m cn cn d d Ph
490 THE FINITE ELEMENT METHOD Figure 9.22 The layer numbering and coordinate system used for a typrca! laminated plate. elastic stiffnesses of the kth layer referred to the plate coordinates (x, y): 1 - v\2vk2l' Olk)- v21cl o, m 0, ^66 = G 12» m= U44 -^23* Ek k k ' V12V2l 6{k) - Gk (9.137b) The engineering constants E\, E\, G\v Gk2V and v\2 referred to the principal material axes of the /cth layer. Note that in laminated plates the principal material axis A-3 is always taken along the plate axis z that is transverse to the plane of the laminae. Example 9.7 Consider a square sandwich plate subjected to sinusoidally distiibuted transverse loading. The sandwich plate is treated as a three-layer plate with different thickness of the face sheets (layers 1 and 3) and the core (layer 2). The face sheets are assumed to be orthotopic, with the principal material coordinates (jq, *2, x-$) coinciding with the plate axes (x, y, z), and the material properties are given by Ei = 25£2, G23 = 0.2£2, E2 = 10*psi, vl2 = 0.25. r\2 Gi3=().5£2, (9.138a) The core material is assumed to be transversely isotropic and is characterized by the following material properties: Ex ^£2=106psi, G13 = G23 = 0.06 x 106psi, vn = 0.25,
FINITE ELEMENT MODELS OF THE FIRST-ORDER SHEAR DEFORMATION PLATE THEORY 491 —- =0.(116 x M6mi. K„ = 5 2(1 + Vi2) G\2 = 0/1 , .. x = 0.016 x i06psi, ** = f • (SU38b) Each face sheet is assumed to be one-tenth of the total thickness of the sandwich plate. The finite clement results obtained with a uniform 4x4 mesh of eight-node quadratic elements with reduced integration (4Q8-R) are compared with the exact solutions based on FSDT and elasticity solution (ELS) of Pagano [28] in Table 9.9. The deflection is nondimensionalized as w = w0(0.0) x 102 ( ^—). (9.139a) The stresses are nondimensionalized as inEq. (9.135), and their locations with respect to a coordinate system whose origin is at the center of Che plate are as follows: *,(o,o.£), ^(0,0^), ^(|.|.-|)- (9.139b) The results of Table 9.9 indicate that the effect of shear deformation on deflections is significant in sandwich plates even at large values of a/h (i.e., the thin plate range). The equilibrium-derived transverse shear stresses [see, for example, Eqs. (8.l44a-c) for the CLPT] are very close to those predicted by the elasticity theory for aj h > 10, while those computed from constitutive equations are considerably Table 9.9 Comparison of nondimensionalized maximum deflections and stresses in simply supported sandwich plates subjected to sinusoidal load (hf = h^ — 0.1 h, h2 = 0.8/7, K3 = 5/6) a/h 10 20 100 Source ELSa Exactb FEM ELS Exact FEM ELS Exact FEM CLPT w — 1.5604 1.5603 — 1.0524 1.0523 — 0.8852 0.8851 0.8782 V.xx 1.153 1.0457 1.0384 1.110 1.0831 1.0755 1.098 1.0964 1.0887 1.0970 5yy 0.1104 0.0798 0.0792 0.0700 0.0612 0.0608 0.0550 0.0546 0.0542 0.0543 &xy 0.0717 0.0552 0.0548 0.0511 0.0466 0.0462 0.0437 0.0435 0.0432 0.0433 &XZ 0.300 0.1374 (0.3134) 0.1365 0.317 0.1409 (0.3213) 0.1399 0.324 0.1422 (0.3242) 0.1412 (0.3243) ayz 0.0527 0.0293 (0.0408)c 0.0278 0.0361 0.0234 (0.0325) 0.0233 0.0297 0.0213 (0.0296) 0.0161 (0.0295) a3D elasticity solution from [40j. hBased on the first-order shear deformation plate theoiy (FSDT). cValues computed from equilibrium equations.
492 THE FINITE ELEMENT METHOD 0.0 0.1 0.2 0.3 0.4 0.000 0.025 0.050 0.075 0.100 0.125 Stress, gxz (0,b/2,z) Stress, oyz {a/2, 0,z) Figure 9.23 Distribution of shear stresses through the thickness of a simply supported sandwich plate under sinusoidal load, (a) aXz> (b) cryz- underestimated for small sidc-to-thickness ratios. The transverse shear stress component ayz is significantly overestimated by CLPT. Figures 9.23a and 9.23b show the variation of the transverse shear stresses through the thickness of the sandwich plates for side-to-thickness ratios a/h = 2,10, and 100. The same sandwich plate discussed above is analyzed for simply supported and clamped boundary conditions when a uniformly distributed load is used. Once again a quarter-plate model is used with 4x4 mesh of quadratic FSDT elements and 8x8 mesh of CPT conforming cubic elements. The results are presented in Table 9.10* The effect of shear deformation on the deflections is even more significant in clamped plates than in simply supported plates. Example 9.8 Here we consider natural vibration of a simply supported laminated plate. The plate is made of four orthotopic layers. The principal material axes (^i, X2) of the first and fourth layers coincide with the plate axes while those of the middle two layers are oriented at 90° to the plate axes (x, y). Such a laminate is designated as (0/90/90/0) or simply (0/90)iV> where the latter notation implies that the laminate is a symmetric (about the midplanc) laminate. The material properties of each layer (or lamina) are taken to be E[ = 4QE2, G[2 = C13 = 0.6E2> G23 = 0.5E2, v\2 = 0.25. (9.140) A 2 x 2 mesh in a quarter plate is used to obtain the results (rotary inertia included). Effects of side-to-thickness ratio, integration, and type of element on the nondi- mensionalized fundamental frequency w = co(a2/h)yfp/E2 can be seen from the results presented in Table 9.11. From the results obtained, it is clear that both the reduced integration (R) and selective integration (S) rules give good results for a wide range of side-to-thickness ratios.
EXERCISES 493 Table 9.10 Nondimensionallzed maximum deflections and stresses in square sandwich plates with (a) simply supported and (b) clamped boundary conditions (hi = h3 = 0,1 h, h2 = 0.8/7, Ks = 5/6) a/h Source w x 10w axx dyy 5Xy &xz &yz (a) Simply supported plate under uniformly distributed load 2.3370 1.5430 0.0883 0.1136 0.2396 0.0991 1.3671 1.5964 0.0526 0.0916 0.2433 0.0881 1.3359 1.5978 0.0514 0.0906 0.2394 0.0880 1.3296 1.5830 0.0509 0.0906 — — (b) Clamped plate under uniformly distributed had 1.2654 0.5018 0.0550 0.0120 0.2318 0,1445 0,3111 0.5356 0.0108 0.0039 0.2406 0.1160 0.2785 0.5347 0.0094 0.0030 0.2400 0.1148 0.2951 0.5401 0.0145 0.0605 — — 518 x 8 mesh of conforming cubic elements with full integration for stiffness coefficient evaluation and one-point Gauss rule for stresses. bThc 4Q9-S clement gives the same results as 4Q9-R. Table 9.11 Effects of side-to-thickness ratio, integration, and type of element on the nondimensionallzed fundamental frequency w of simply supported square laminates (0/90/90/0) 10 50 100 CLPT 10 50 100 CLPT 4Q8-R 4Q8-R 4Q8-R 8CC-Fa 4Q9-Rb 4Q9-R 4Q9-R 8CC-Fa a/h 2 10 100 Serendipity Element F 5.502 15.174 19.171 R 5.503 15.179 18.841 S 5.501 15.159 18.808 Lagrange Element F 5.502 15.182 19.225 R 5.503 15.193 18.883 S 5.501 15.172 18.933 Exact 5.500 15.143 18.836 d dx du dx EXERCISES 9.1 Derive the finite element model of the equation = /, 0<jc<1, for the boundary conditions of the type T du 1 u(0) = w, (1+*)— +u\ = q, L dx Jv=l and solve the problem for the data / = 0, u — 1, and q = 0. Use two linear finite elements. 9.2 Use the following procedure to derive the cubic interpolation functions for a line clement with equally spaced (four) nodes. Set up the local coordinate x at
494 THE FINITE ELEMENT METHOD the left-end node (i.e., node 1), and then use the interpolation property of ^7, $i(xj) = Su> 1,7 = 1.2,3,4, to write ty as a polynomial that vanishes at all xj, xj ^ x,\ \fri(x) = c\{x - x2){x - xi)(x - x4), lfa(*) - c2(x - x\)(x - X3)(x - m), and determine C[, c2, etc., such that V'l t*;i) = 1, V^fe) == U etc. Use the minimum possible number of linear finite elements to analyze the axially loaded structures of Exercises 9.3-9.9, 9.3 Find the stresses and compressions in each section of the composite member shown in Fig. E9.3. Use Es = 30 x 106 psi, EQ = 107 psi, Eb = 15 x 106 psi, and the minimum number of linear elements. Aluminum (Aa=6 in.2) Figure E9.3 9.4 A solid circular brass cylinder (Eb — 15 x I06psi, ds — 0.25 in.) is encased in a hollow circular steel shell {Es — 30 x 106psi, cls = 0.21 in*)« A load of P = 1,3301b compresses the assembly as shown in Fig, E9.4. Determine (a) the compression and (b) the compressive forces and stresses in the steel shell and brass cylinder. Use the minimum number of linear finite elements. Assume that the Poisson effect is negligible. 9.5 A rectangular steel bar (Es = 30 x 106 psi) of length 24 in, has a slot in the middle half of its length, as shown in Fig. E9.5. Determine the displacement of the ends due to the axial loads P — 2,000 lb. Use the minimum number of linear elements. 9.6 The two members in Fig. E9.6 are fastened together and to rigid walls. If the members arc stress-free before they are loaded, what will be the stresses and deformations in each after the two 50,0001b. loads are applied? Use Es = 30 x I06 psi and E(l — 107 psi; the aluminum rod is 2 in. in diameter and the steel rod is 1.5 in. in diameter.
EXERCISES 495 :1,3301b P K 6 in. 6 in. 6 in. 6 in. Figure E9.5 Aluminum " 20 in. Figure E9.6 9.7 The aluminum and steel pipes shown in Fig. E9J are fastened to rigid supports at ends A and B and to a rigid plate C at their junction. Determine the displacement of point C and stresses in the aluminum and steel pipes, Use Lhc minimum number of linear finite elements. P:=50kN Steel (E, =20 0 GPa, A3=60 mm2) p 30cm Aluminum (E(t = 70 GPa,Aft=60 0 mm) Figure E9.7
496 THE FINITE ELEMENT METHOD 9.8 A steel bar ABC is pin-supported at its upper end A to an immovable wall and loaded by a force F\ at its lower end C, as shown in Fig. E9.8. A rigid horizontal beam BDE is pinned to the vertical bar at B, supported at point D, and carries a load F2 at end E. Determine the displacements «# and uc at points B and C. 20 in. E=30xW*psi 30 in. L 30 in. *\* 25iri7- D II, • A2=0.25 in2. F2=6,000 lb f F. -2.000 lb Figure E9.S 9.9 Repeat Exercise 9.8 when point C is spring-supported (see Fig. E9,9). -& D E 13 w Fx =2,000 lb F2 =6,000 lb k -1,000 lb/in. Figure E9.9 9.10-9.17 Use the minimum possible number of Euler-Bernoulli beam elements to analyze (Le.> determine deflections and slopes at the nodes of) the beam structures shown in Figs. E9.10-E9.17,
EXERCISES 497 Steel members (Es = 30x l()c psj) 200 lb/in. / 4inL i 4 in. t • 1 2 in:—H*" *\ Figure E9.10 600 lb/ft 1,000 1b ¥ftttfftf " •8 ft. 4 ft—M £7 = 6 x 108lbtn.2 Figure E9.12 200 N/m 1 kN Steel members (£,.-200GPa) 1.5 hxdm. |TWOTMW 15 in.dia. ,fH|HjH V.3 cm :di&; 5kN-m 2 cm dia= -w«- 12 cm 12 cm Figure E9.11 "►I HHHmmm z,w0 Figure E9.13 ^0=1,000 lb k 1 ft g0= 500 lb/ft. >j^: 5 ft^ 5 ft*** 5 ft ^5 ft^ #7~ constant Figure E9.14 S0= 500 N/m^ F0=1,000 N ^mm ^iir i p ,^.^ T z,wn 5 in ^L 5 m J^ 5 m #7= constant Figure E9.15 <?0= 1,000 lb/ft Fn= 1,000 lb H 8 ft *\* 6 ft H 7£7= constant z,w0 Figure E9.16 9.18 Construct the finite element model, based on the weak form, of the following differential equation, which arises in connection with the buckling of columns: irAE1i^)+N^^ 0<X<L'
498 THE FINITE ELEMENT METHOD U, U3 Figure E9.17 where N is the buckling load to be determined and EI is the flexural rigidity, 9.19 Evaluate the finite element matrices of Exercise 9.18 using the Hermite cubic polynomials. Take EI constant and make use of the results already available in the text. 9.20 Find the critical buckling load of a simply supported beam using the least number of Euler-Bernoulli beam elements. This problem requires the solution of an eigenvalue problem. 9.21 The axi symmetric bending of a thin, uniform thickness, circular plate can be described by a one-dimensional differential equation. For isotropic material the differential equation is given by [sec Eqs. (8.36) and (8.80)] 1 d V ( d*w d2w\ ldw~\ r/ N ^ where r is the radial coordinate and D = Eh3/[12(1 — v2)\ is the bending stiffness. Derive the variational formulation and associated finite element model of the equation over a typical element. 9.22 The differential equations governing axisymmctric bending of circular plates according to the first-order shear deformation plate theory are where As$, Du, ^I2» and D22 are plate material stiffnesses, <j) is the rotation, wo is the transverse deflection, and q is the transverse load. Develop (a) the weak form of the equations over an clement and (b) the finite element model of the equations.
REFERENCES 499 9.23 Consider the fourth-order equation governing the Euler-Bernoulli beam theory. Suppose that a two-node element is employed, with three primary variables at each node (tuo, 0, and /c), where 0 — dwo/dx and k = d2w$/dx2. Show that the associated Hermite interpolation functions are given by A'3 V4 X5 ( X2 X3 X4\ 2 3 \ 3 4 5 r x^\ , txJ x* sx X (. ~.X ~X* X" \ , .„■*" *~X~* ,X* 4n = Y{1-\+3^~^l ^=-10^+15^-6^, / x2 x3 x4\ x2 (x x2 x3\ where x is the element coordinate with the origin at node 1. 9.24 Consider the weak form (9.54a) of the Euler-BernouUi beam element. Use a three-node element with two degrees of freedom (wq,&), where 6 = —dwo/dx. Derive the Hermite interpolation functions for the clement. Compute the element stiffness matrix and force vector. 9.25 Give the variational formulation and associated finite clement model of the equation 3 / du du\ d ( du du\ . ~^~ C^T"+C^— -7-U-2i™+C22—)+co« = /, in ft. (a) dx \ dx dy / dy \ dx By/ The coefficients c;j (i> j — 1, 2) and co and the source term / are given functions of position (x, y) in the domain Q. Equation (a) arises in the study of a number of engineering problems, including torsion of constant cross-section members, conduction heat transfer, irrotational flow in a viscid fluid, transverse deflection of a membrane, and so on. REFERENCES 1, Bathe, K. J., Finite Element Procedures, Prentice-Hall, Englcwood Cliffs, NJ (1996). 2. Burnett, D. S., Finite Element Analysis, Addison-Wesley, Reading, MA (1987). 3. Hughes, T. J. R., The Finite Element Method? Linear Static and Dynamic Finite Element Analysis, Prentice-Hall, Englewood Cliffs, NJ (1987). 4, Reddy, J, N,, An Introduction to the Finite Element Method, 2nd edition, McGraw-Hill, New York (1993), 5* Rcddy, J. N., "On locking-free shear deformable beam finite elements," Comput Methods AppL Meek Engng,, 149, 113-132 (1997). 6, Rcddy, J. N., "On the dynamic behavior of the Timoshcnko beam finite elements," Sadhana (Journal of the Indian Academy of Sciences), 24, 175-198 (1999). 7. Reddy, J. N., "On the derivation of the supcrconvergentTimoshenko beam finite clement," Int. J. Comput. Civil Struct. Engng., 1(2), 71-84 (2000).
500 THE RNITE ELEMENT METHOD 8, Zienkiewicz, O. C, and Cheung, Y. 1C, "The finite element method for analysis of elastic isotropic and orthotopic slabs," Proc. Inst. Civ. Engrs., London, 28, 471-488 (1964). 9. Bazeley, G. R, Cheung, Y K., Irons, B. JVi, and Zienkiewicz, O. CM "Triangular elements in bending: conforming and non-conforming solutions/' Proc. Cortf Matrix Methods in Structural Mechanics, Air Force Institute of Technology, WPAFB, Dayton, Ohio, AFFDL- TR-66-80, 547-576 (1965), 10. Bogner, R KM Fox, R. L,, and Schmidt, L. A., Jr., "The generation of interelement- compatible stiffness and mass matrices by the use of interpolation formulas/* Proc. Conf. Matrix Methods in Structural Mechanics, Air Force Institute of Technology, WPAFB, Dayton, Ohio, AFFDL-TR-66-80, 397-443 (1965). 11. Clough R. W., and Tocher, J. L,, "Finite element stiffness matrices for analysis of plate bending," Pwc. Conf. Matrix Methods in Structural Mechanics, Air Force Institute of Technology, WPAFB, Dayton, Ohio, AFFDL-TR-66-80, 515-546 (1965). 12. Herrmann, L. R. "Finite element bending analysis for plates,*' Proc, ASCE, J. En grig. Meek Div., 93(8), 13-26 (1967). 13. Fraeijis de Veubeke, B., "A conforming finite element for plate bending," Int. J. Solids Slruct.,4(\), 95-108 (1968). 14. Irons, B. M., "A conforming quartic triangular element for plate bending,'* Int. J. Numerical Methods Engng., 1, 29-45 (1969). 15. Melosh, R. J., "Basis of derivation of matrices for the direct stiffness method," AIAA /., 1, 1631-1637(1963). 16. Hrabok, M. M., andHrudcy, T ML, "Areview and catalog of plate bending finite elements," Comput. Struct, 19(3), 479-495 (1984). 17. Barlow, J., "Optimal stress location in finite element models,'* Int. J. Numerical Methods Engng., 10, 243-251 (1976). 18. Barlow, J., "More on optimal stress points: reduced integration element distortions and error estimation," Int. J. Numerical Methods Engng, 28,1486-1.504 (1989). 19. Newmark, N. M., "A method for computation of structural dynamics," ./. Engng. Mech., 85,67-94(1959). 20. Nickell, R. E., "On the stability of approximation operators in problems of structural dynamics," Int. J. Solids Struct., 7, 301-319 (1971), 21. Hughes, T. J. R., Cohen, ML, and Haroun, M„ "Reduced and selective integration techniques in the finite element analysis of plates," Nucl. Engng. Design, 46, 203-222 (1981), 22. Belytschko, T., Tsay, C. S., and Liu, W K„ "Stabilization matrix for the bilinear Mindlin plate clement," Comput. Methods Appl. Meek Engng., 29, 313-327 (1981), 23. Bathe, K. J., and Dvorkin, E. NL, "A four-node plate bending clement based on Mindlin/Reissner plate theory and mixed interpolation," Int. J. Numerical Methods Engng., 21, 367-383 (1985). 24. Huang, H, C, and Hinton, E., "A nine-node Lagrangian plate element wilh enhanced shear interpolation," Engng. Comput, 1,369-379 (1984). 25. Reddy, J. N., "Simple finite elements with relaxed continuity for non-linear analysis of plates," Proc. 3rd Int. Conf. on Finite Element Methods, University of New South Wales, Sydney, 2-6 July (1979).
REFERENCES 501 26. Reddy, J. N., "A penalty plate-bending element for the analysis of laminated anisotropic composite plates," Int. J. Numerical Methods Engng., 15(8), 1187-1206 (1980), 27. Reddy, J. N.> Median ics of Laminated Composite Plates, Theory and A nalysis, CRC Press, Boca Raton, FL (1997). 28. Pagano, N. J., "Exact solutions for rectangular bidirectional composites and sandwich plates," J. Composite Mater., 4, 20-24 (1970).
10 MIXED VARIATIONAL FORMULATIONS 10.1 INTRODUCTION 10.1.1 General Comments Tn this chapter we will cover mixed variational principles and their applications. We define a "mixed" (or sometimes called "hybrid") variational formulation as one where secondary variables of the conventional formulation are also treated as dependent variables along with the primary variables. Most often these formulations are developed with the objective of determining the secondary variables, which are often quantities of practical interest, directly rather than from postcompulations. In this chapter we will present a brief exposure of mixed variational principles of elasticity, including Reissner's variational principle and the Hellinger-Reissner variational principles |_1»2]. Use of variational methods as well as the finite clement method to determine solutions based on the mixed variational formulations will also be discussed. 10.1.2 Mixed Variational Principles To illustrate the basic ideas in the development of mixed variational principles, we shall consider the simple case of axial deformation of a bar. From equilibrium considerations (see Example 4.3) we have -— = f (*), 0 < x < Lt N = f axx dA, (10.1) dx ' J a where L is the length of the bar, / is the body force along the axis i, and aA-x is the axial stress in the bar. The axial force N is related to the axial displacement u through 502
INTRODUCTION 503 the kinematic and constitutive relations du *™ = T, °xx=Egxx% (10.2) dx by [substitute Eqs. (10.2) into the definition of TV in Eq. (10.1)] du du N N = EAT~ or ___=o. (10.3) dx dx EA Substituting Eq. (10.3) for N in Eq, (10.1), we arrive at d ( du\ -— (EAfa) = /(*>■ 0 < * < L. (10,4) The conventional (or displacement) variational formulation is based on Eq. (10.4). Consequently, the displacement it is the primary dependent variable of the formulation, and TV is postcomputed from the known displacement using Eq. (10.3). The variational (or weak) formulation of Eq. (10.4) is obtained by either following the steps presented in Section 7.4.1 (Model Equation 1) or using the principle of minimum total potential energy. The total potential energy functional is given by [see Eq. (4.109)] [L\EA (du\2 dx-WPU (10.5) where WPL stands for work done by applied point loads. Equation (10.4) is the Euler cquation of this functional. Tn a mixed formulation, we treat Eqs. (10.1) and (10.3) as a pair of coupled equations in two independent variables u and TV. The mixed variational formulation of this pair can be developed using the same steps as discussed in Section 7.4.1 (or see Examples 7.10 and 7.11). We have fL ( dN \ o=L ""hfe-'r*' (,o'6a) °-jf «(£-£)<•• (i06b) where w\ and w% are the weight functions, which may be interpreted as follows. If w\ ' f dx were to represent-the work done, w\ must be an axial displacement. Similarly, W2 • (dujdx)dx represents work done if w2 is an axial force. Thus, w\ — Su and W2 ~ <$TV, Since TV appears without differential in Eq. (10.6b), in the interest of lowering the order of approximation used for TV, we carry out the integration by parts of the first term in Eq. (10.6a). One reason is that we do not wish to make u a secondary
504 MIXED VARIATIONAL FORMULATIONS variable by carrying out integration by pails of the first term in Eq. (10.6b). There are other reasons that will be apparent below. We obtain ° = foW*{-%-f)dX Now it is clear that, since w\ ~ w, w is the primai'y vaiiable and TV is the secondary variable. Note that N is also an independent variable of the fonnulation but it is not a primary variable. Equations (10.7a;b) can be written, since W\ = Su and it>2 = <$ N are linearly independent, as 0 = / ( ^W - Suf + SAf^ - -^rSW ) dx - f«wW|J Jo \ dx dx EA } =6C(%N-^-fi)dx-[suN],»- (,a8) The boundary term can be simplified for a specific problem. Whenever it is specified at x = 0 or x = L, <5w is zero there. If « is not specified at a boundary point, then N is known (either as zero or some specified value) at that point. Suppose that it is zero at x — 0 (essential boundary condition) and u is not specified at x = L, but N is specified to be P at a* = L. Then we have 0 = 5 ' /'L {du N2 \ I {TxN-2EA-fVdX-U(L)P, (10.9) This is the mixed formulation associated with Eqs, (10,1) and (.10,3). Clearly, the functional nm(M|JV) = jf ^N-J^-fiSdx-u{L)P (10,10) yields Eqs. (10.1) and (10.3) as its Euler equations along with the natural boundary condition N-P=0 atx = L. (10.11) In otlier words, SUfn — 0 is equivalent to Eqs. (10.1), (10.3), and (10.11). Note that n,„ does not attain a minimum or maximum with respect to u and N. It may be minimum with respect to u but maximum with respect to N. Thus the mixed variational
INTRODUCTION 505 principle can be stated as: find (u, N) such that o = mm = suiim + sNnm, or sunm = o and ** nm = o. (10.12) 10.1.3 Extremum and Stationary Behavior of Functional It can be shown that U(u) of Eq, (10,5) attains a minimum when u is the solution of Eq. (10.4) and satisfies the boundary conditions M(0) = H0=0, (10.13) To establish this, suppose that u{x) = u(x) + av(x) is an arbitrary element from the set of admissible functions (i.e., it satisfies the specified geometric boundary conditions and is differentiable as required by the functional Fl), where a is the solution of Eq. (.10,4), a is a real number, and v is an element from the space of admissible variations (i.e., v satisfies the homogeneous form of the specified geometric boundary conditions and is differentiable as required by n). Then we have dx — u(L)P dx -u(L)P ™>-f [¥(£)-" «7/¥(£)2- r rL ( du dv \ , „, x 2 fL EA /clv\2 , Jo L dx V dx) + < + < v dx + a EA—-P dx = Tl(tt) + a2 I — [ — | dx > n(«' JQ 2 \dxj v{L) (10.14) where in arriving at the last step we have used the fact that u satisfies Eqs. (10.4) and (10.13). It is clear from Eq. (10.13) that 11(5) > U(u) for any a ^ 0, and U(tt) — Tl(u) if and only if a = 0 (i.e., u = u). Hence n(«) attains its minimum only at u = u, and for any other, admissible function u we have 11(5) > U(tt). This is essentially the statement of the principle of minimum total potential energy [see Eqs. (4.87M4.90) and (5.30)].
506 MIXED VARIATIONAL FORMULATIONS Next, we establish the stationarity of I1w(m, AT). Let w = u+av and// = N+fiQ, where a and /? are real numbers, and (u, N) satisfy Eqs. (10.1) and (10.3) inside the domain, Eq. (10.11) at x = L, and u = 0 at x = 0, Then 2EA N2 CL (du N2 \ + <x\[ \^-N-fv\dx-v{L)p\ P2 fL Q2 fL dv 2 J0 IE A * Jo dx = nm(u,N)~^ £-dx+ap —Qdx> (10.15b) The third and fourth lines of Eq. (10.15a) vanish by virtue of Eqs. (10J), (10.3), and (10.11). It is clear that Tlm(u, N) decreases from its value at equilibrium when u is kept at its actual value u (i.e., a — 0) and TV is changed to N — N + PQ; that is, (10.15a) Ilm(ufN)< Um(u9N). (10.16a) The value of the functional is unchanged when N is kept at its actual value (i.e., p = 0) and u is changed to u = u + av. In fact, if the end x = L is connected to an elastic spring, one can show that nm(u,N)>Tlm(u,N). (10.16b) Thus, IIW exhibits a stationary behavior and not an extremum character. 10.2 STATIONARY VARIATIONAL PRINCIPLES 10.2.1 The Minimum Total Potential Energy Principle The principles of minimum total potential energy and maximum total complementary energy are both externum principles. Recall from Eq. (4.87) that a functional II (u) is a minimum at the displacement vector u if and only if n(u) > n(u) and the equality holds only if n = u. for all u (10.17)
STATIONARY VARIATIONAL PRINCIPLES 507 To show that the total potential energy of a linear elasticity body is the minimum at its equilibrium configuration, consider the total potential energy functional [of an anisotropic, linear elastic body under the assumption of infinitesimal strains; cf. Eq. (4.114)1: n(u) = f (-Cijk(ek}eu - fan) dV - J fa ds, (10.18a) where u is the displacement vector, f is the body force vector, t is the specified traction vector on boundary 52 of the volume V occupied by the body, Cjjm are the components of the fourth-order elasticity tensor, and e/y are the components of the infinitesimal strain tensor: *y = ^("/,j+";.*). (10.18b) In Eqs, (10.18a,b), summation on repeated indices is implied, and a comma followed by a subscript i denotes differentiation with respect to *,-. The geometric boundary condition on u is u-=u on Si, (10.19) where S\ + Sj = S, the total boundary of V. Let u be the true displacement field and u be an arbitrary but admissible displacement field. Then u is of the form u — u-hav, where v is a sufficiently differentiate function that satisfies the homogeneous form of the essential boundary condition v = © on Si. From Eq, (10.18a), we have n(u + ay) = / -Cjjki (eid + ctgu) (eij + agu) - f; (uj + at//) dV - f ti(u; +av{)ds, 'Si where 8ij = 2^./ + ^)- Collecting the terms, we obtain (because Cjjkt — Q/,7) n(u) = n(u) + a \ I (-fiVi + Cijkjekigij + ~Cijk}gugkl) dV - fa ds (10.20)
508 MIXED VARIATIONAL FORMULATIONS Using the equilibrium equations, we can write - / fjVidV= / (TijjVf dV = / CijkieujVidV = - / CijkieklVijdV + f CijkiekiViiijds = ~ I CiJk,ekigijdV+ [ fads* (10.21) Jv Js2 where the condition v; = 0 on S\ is used in arriving at the last step. Substituting Eq. (10.21) into Eq. (10.20), wc obtain a2 f n(u) = n(u) + — J QjMigijgki dv. (io.22) In view of the nonncgative nature of the second term on the right-hand side of Eq. (10.22), it follows that n(u)> n(u), (10.23) and Il(u) = n(u) only if the quadratic expression \Cijkigijgki is zero. Due to the positive-definiteness of the strain energy density, the quadratic expression is zero only if Vj — 0, which in turn implies u\ — ii\, Thus, Eq. (10.23) implies that of all admissible displacement fields the body can assume, the true one is that which makes the total potential energy a minimum. Therefore, the total potential energy principle is a minimum principle. Similar arguments can be made for the total complementary energy. It should be noted that the consideration of whether a functional is minimum or maximum does not enter the calculation of the Euler equations. Such a consideration may be used in determining approximate solutions. 10.2.2 The Hellinger-Reissner Variational Principle A stationai'y principle is one in which the functional attains neither a minimum nor a maximum in its arguments. In fact, a functional can attain a maximum with respect to one set of variables (while the others are fixed) and a minimum with respect to another set of variables involved in the functional, as discussed earlier. An example of such functional is provided by the functional based on the Lagrange multiplier method (see Section 4.4.8). For example, the functional in Eq. (4.132) attains a minimum with respect to (woy <f>) and a maximum with respect to X. Stationary principles, also called mixed variational principles, are of special importance in the analysis of structures. Here wc consider the Hellinger-Reissner variational principle [3-6] for an elastic body. The stationai'y functional is constructed from the equations of an elastic continuum by treating all the dependent variables as independent of each other, The stationary conditions of the principle are the strain-displacement equations,
STATIONARY VARIATIONAL PRINCIPLES 509 stress—strain equations, stress-equilibrium equations, and both natural and essential boundary conditions—in short, all of the governing equations of elasticity (see Odcn and Reddy [ 1]). Here we consider the principle for an elastic body undergoing large displacements (static case). Recall the total potential energy functional FI from Eq. (4J14): n(u) = / -(itij 4- ujj)(Uij + ujj) + -u}jukj( - flu,- dV - / tiW ds, (10.24) Js2 where \i and A. denote the Lame constants. In writing the total potential energy functional n, it was assumed that the displacements and strains are related by a kinematic relationship (10.18b) and that the displacement field satisfies the specified boundary conditions on Si [see Eq. (10.19)1, The principle of minimum total potential energy gives (see Example 4.11) the equilibrium equations (4.117a) and traction boundary conditions (4.117b) as the Euler equations: ix(tiijj + iijjj) + Xukjki + fi = 0 in V, (10.25a) */ = [jj, (tijj + "/,;) + htik.k&ij] rtj = f} on 52. (10.25b) Here rij denote the components of the unit normal vector on the boundary S of the volume V occupied by the elastic body. Now suppose that we wish to include the strain-displacement relations (3.20) £ij = 2(w/.j + Ujj + umjumj) (10.26) and displacement boundary conditions (10.19) Ui - hi (i = 1, 2, 3) on Si (10.27) into the variational statement by treating them as constraints. Toward this end, we first rewrite the total potential energy functional II of Eq. (10.24) in terms of the nonlinear strains [i.e., revert to Eq. (10.18a)]: n(K/, fjj) = / (Uo(s{i) - fnu)dV - f tun ds% (10.28) Jv Js2 where Uq is the strain energy density function. Next, let Xjj = A/,- (i, 7 = 1,2,3) and /X}(i = 1, 2, 3) he the Lagrange multipliers associated with the strain-displacement relations (10.26) and the displacement boundary conditions (10.27), respectively.
510 MIXED VARIATIONAL FORMULATIONS Using the Lagrange multiplier method, we introduce the new functional: XlHw(Mi,Sfj,kij, fXi) = II(w/, Sij) - / fii (iii - &i) ds + / XU [l(uU + UJJ + umJumj) - Eij] dV J V = / {[I (UU + UJJ + UmjUmj) ~ £ij\ XU + ^ofoy) - fjui }dV - I fii (it; - it;) ds - I fat ds (10.29) where 5} -f 52 — 5 is the total boundary, and uf and i\ are the specified displacement and traction components, respectively, on Si and S2. We now wish to check if the Euler equations of the new functional n hw are indeed die governing equations of elasticity. Setting the first variation of Ylnw to zero, 0 = snHW s 8unHW + 8£nHW + skn„w + sfJinHW, (10.30) and noting that the variations in u, s, k and \jl are arbitrary, it follows that Suriw = o, 8anHw = o, hnHw = o, 8flnHW = o. (10.31) Caixying out the variations with respect to the dependent variables, we obtain SUhw = / [{faij +$uj,i +&umiiuntj+umt;8umj)kjj - fiSui]dV j v — / fifSuf ds — I ijSui ds -h / I —detjXjj -f 8eij ) dV J Si Js2 Jv \ fey / + / [k{uUi + uiJ + UmjUm, /) - Sif]8X.ij dV - I &flj(Uj - Ui)ds = / Suij(Si„, -f Uij„)kmJ 4- f j-^ - kij 1 8eij - fjSuj dV + I [l(uU + UJJ + umjUmj) ~ Sij]8Xij dV JV — I iXi&iiids— f ij8uids— I 8fj,i(u; — u{)ds. (10.32) JS) JS2 JSi Using integration by parts, the first term in Eq. (1032) can be expressed as / 8ujj(8im + ui>m)kmj dV = - I kmj(8;m + «,>) Mi dV Jv Jv ,J + / njkmJ(Sjm + Ui,m)8uj ds. (10.33)
STATIONARY VARIATIONAL PRINCIPLES 511 Substituting Eq, (10.33) into Eq. (10.32), and setting the coefficients of 8w, <$£,*/, and 8Xjj in V, Siii on S] and 52, and 8 m on S\ to zero, we obtain Su(: frmjtfim + Uitm)]j + fi = 0 in V, (J0.34) S£ij: i^ _ Ai7 =0 in V, (10.35) 5A/;-: ±(m,j + itjj + umjumj) - 8ij =0 in V, (10.36) <5w/: njXmj (8im + u/ilrt) - /*/ = 0 on 5i, (10.37) Sw/: njkmj(8im + m/,/m) - ft = 0 on 52, (10.38) 5/i/: «/ - Ui = 0 on 5i. (10,39) Equations (10.35) and (10.37) can be interpreted as the definitions of the Lagrange multipliers Xfj and m, respectively. Clearly, they have the meaning of stress and traction components, respectively. Thus the functional Tlfiw gives the equilibrium equations (10.34), constitutive equations (10.35), strain-displacement relations (10.36), traction boundary conditions (10.38), and displacement boundary conditions (10,39) as the Euler equations. Consequently, finding the stationary values of YIhw is stated as a variational principle, and it is termed the Heiiinger-Reissner variational principle for an elastic body. Example 10.1 Consider the following equations of the Euler-Bernoulli theory of beams: Kinematics k = — ^1 w(0) = Wq (10.40) dx1 dw d v Constitutive equation M - EIk, - —(0) = #o (10,41) Kinetics (equilibrium) d2w M= EIk, d2M __ dx2 w(0) = wq dw - — (O)=0O dx M(L) = M dM dx (10.42) where wo and $o denote specified displacement and rotation, respectively. The Heiiinger-Reissner type variational principle for the Euler-Bernoulli beams is given by fL /d2w \ nffwOi;,*,*,^!,/^) = nOO-f- / I-j-j+zrJXrfjc + Mi N(0) -wo] ->2[( +'wi(S)L+* (10.43) where T\(k) is the total potential energy functional expressed in terms of the curvature k . The functional n accounts for the moment and shear force boundary conditions
512 MIXED VARIATIONAL FORMULATIONS because they arc the natural boundary conditions for FL We have ( dw\\ + fll[w(0)- W()] + M2 (10.44) The Euler equations of functional Tluw are 8k: 8w\ SX: dSw dx 8w: Sfx\: 8fi2: ElK+k: = 0, d2k dx1 d2w dx1 -X(0)+/t2 = 0, ^ \lx~) + Mi=0, 0 xv(0) — wo — 0, -r-(0) + fo = 0. inO < jc < L, inO < x < L, inO < a- < L, k(L) + ML = 0, (I) + Vl = 0, (10.45) (10.46) (10.47) (10.48) (10.49) (10.50) (10.51) From a comparison of Eqs. (I0.40M10-42) with (10.45M10.49), we note that the Lagrange multipliers X, jA{, and fX2 are given by X{X) — ~M(x)y /JL[ - (dM\ ix2 = -M(0). (10.52) Example 10,2 Here we develop a mixed formulation of the Timoshenko beam theory. In this formulation the displacements (wo, 4>) and strains (/c_™, yxz)are treated as the dependent variables. The variational statement associated with this mixed formulation is given by the stationarity of the following functional (see Oden and ReddylUp. 116): U?" = f [EI (S - 5*") "" + GAK< (l7 + * ~ 5y") "< - qW« - Vxwo(0) - V2U>o(L) - Afi0(O) - M24>(L)9 dx (10.53)
STATIONARY VARIATIONAL PRINCIPLES 513 where V{ = [-GAKsyxz]0, V2 = [GAKsyxz]L, MX = I-EIkxx]o, M2 = \Elicxx]L. (10.54) The Euler-Lagrange equations associated with the functional in Eq. (10.53) are wo: - -T- [GAKsyxz] - q = 0, (10.55) ax 8$: - — [EIkkx] + GAK,Yxz=Q, (10.56) ax dd> Stcxx: -f-Kxx=0, (10.57) ax 8yxz: 0^ + 4>)-Yxz=0- (10.58) Note that the first two equations represent the equilibrium equations (9.72) and (9.71) expressed in terms of the curvature kxx and shear strain yxz, while the last two equations are the kinematic relations [see Eq. (9.68)]. 10.2.3 The Reissner Variational Principle A special case of the Hellinger-Reissner variational principle is the Reissner principle, which is obtained by eliminating strains eij by assuming that the strains are related to the stresses a-tj = X// by the constitutive relations 8U( £ji = 3 9cfU £, U$ = <rjjSij-UQ. (10.59) where Uq is the complementary strain energy density. Then wc obtain frornEq. (10.29) n«(tt/, <Jij) = / [\{iti,i + ujj + umjumj)aij - U$(cTij) - fiU(]dV Jv - I t((iii - Cij)ds - / t{Uids. (10,60) JS] Js2 The Euler equations are given by Eqs. (10,34), (10.37)-(i0.39), and Scrij\ i(uij + uJti + umjitmj) - —^ =0 in V. (10.61)
514 MIXED VARIATIONAL FORMULATIONS 10.3 VARIATIONAL SOLUTIONS BASED ON MIXED FORMULATIONS The advantage of slationary principles lies in incorporating all of the governing equations into a single functional. Such functionais form the basis of a variety of new finite element models that often yield better accuracies for stresses than the displacement finite element models (which are based on the total potential energy functional). The variational methods presented in Chapter 7 can also be used to determine solutions of problems based on mixed formulations. The variational methods require a variational statement to determine the undetermined parameters of the approximation, and the variational statement can be anything that is equivalent to the governing equations. Thus the Ritz solution, for example, of an Euler-Bernoulli beam can be determined using the principle of minimum potential energy or a mixed variational principle/statement. The conditions on the approximation functions are dictated by the variational principle. If all governing equations as well as the boundary conditions are included in the variational statement, then the approximation functions are required to satisfy only the continuity requirements. Here we illustrate these ideas through an example. Example 10.3 Consider the bending of a beam using the Euler-Bemoulli beam theory. The governing equations are [see Eqs, (4.30b,c) and (4.34)] d2wo M -^-jj=°< (1062a) d2M —£2=1- 00.62b) In Examples 7.7 and 7.8 we have solved the Euler-Bernoulli beam problem using the principle of minimum total potential energy, whose Euler equation is the fourth-order equation governing the deflection: which is obtained by combining Eqs. (10.62a,b) above. Here we wish to consider a mixed formulation in which both wq(x) and M(x) are treated as dependent variables with independent approximations. That is, we solve Eqs. (10.62a,b) simultaneously (without eliminating M). For comparison purposes, we shall also solve the same problem using the Ritz method based on the principle of minimum total potential energy. As a specific example problem, we consider a cantilever beam under uniformly distributed load, <jo- The boundary conditions are wo(0) - 0} /dwo \~dx~ = 0, x=0 M(L)=0, V^^(^ = 0. (10.64) x=L
VARIATIONAL SOLUTIONS BASED ON MIXED FORMULATIONS 515 In the Ritz approximation based on the minimum total potential energy principle we use the two-parameter approximation W2(x) = c\x2 + C2*3 (which satisfies only the geometric boundary conditions) and obtain the solution The bending moment obtained from this Ritz solution is M2(,) = _£/^ = _^!(5_6^)) M2(0) = -^. (10.66) The exact solutions for wq(x) and M(x) are given by *,« = -«£ (,-£)*, «(0) Thus the two-parameter Ritz solution based on the total potential energy principle (or displacement model) yields exact maximum deflection, but maximum bending moment is in 16.67% error, Let us consider the pair in Eqs, (10.62a,b) and construct their weak forms. Using the three-step procedure, we obtain after two steps 0 = i ~r-r -viqvXdx-Ux'-j-] , (10.68) J0 \ dx dx J [ dx J0 where (ui, 1^2) are the weight functions. To physically interpret the weight functions, we examine the expressions in the integrands. The most obvious one to interpret is the product v\qodx. In order for it to have the units of "work done," v\ should have the meaning of displacement. Thus, v\ ~ wo* Since (M/EI) dx has the units of rotation, i>2 must have the units of a moment, giving vi ~ M. This information is needed in developing the weak forms (and later in applying the Ritz method, where we must replace v\ and V2 with the proper approximation functions). It is clear that V[ must be zero at points where u>n is specified, and V2 should be zero where M is specified. It is also clear from the boundary expressions in Eqs. (10.68) and (10.69) that wq and M are the primary 8E7* q*L2 (10.67)
516 MIXED VARIATIONAL FORMULATIONS variables, while dwo/dx (rotation) and dM/dx (shear force) are the secondary variables. Thus, the nature of the boundary conditions (i.e., classification into essential and natural) depends on the set of equations used. Returning to the wealc form development, we use the specified boundary conditions in Eq. (10.64) to simplify the boundary terms [(dwo/dx)(0) = 0, (dM/dx)(L) = 0, v\ (0) = 0, and V2(L) = 0 maice all boundary terms vanish] and obtain the final weak forms; Jo \ dx dx ) Jo \dx dx EI J (10.70) (10.71) Let us consider the two-parameter Ritz approximations of wq and M\ W2M - a\<p\ (x) + tf2<feM = a\x + C12X2, M2(x) = btinix) + hfaW = fci(L - x) + b2(L - x)2, (10.72) which satisfy the specified essential boundary conditions [w(0) = 0 and M(L) = 0]. Tn general, there is a relationship between the number of parameters chosen for wq(x) and M(x) to insure the invertibility of the resulting algebraic equations for the unknown parameters (af, b{). Substituting Eq, (10.72) for wq and M» and V[ = 4>i and ^2 == ^f into Eqs. (10.70) and (10,71), we obtain Jo Jo dfa I ^ 777 2^> dx dfti 2 - 4>m dx =0, d4j ^[X-ft-k*^* J=l 2 W=1 i = 1,2, dx = 0, In matrix form, we have 2 2 2 J2KJiaJ-T,Gubi = 0> /=1 y=l (10.73) i = 1, 2. (10.74) (10.75a) where Kij = i H7-dTdx' Gij = ^JJo **'dx> fi = h «* dx (10.75b)
MIXED FINITE ELEMENT MODELS OF BEAMS 517 Evaluating the integrals for the choice of 0; and fa from Eq. (10.72), we obtain [*]■ # \L\ [G] = 1 EI [L3 3 L4 . 4 L4] 4 L5 5 _ {/} = c/oL2 I2L • Solving the first equation in (10.75a) for /?\s yields b\ = 0, Substituting for b's in the second equation of (10.75a), and solving for a's„ we obtain b2 = -*>. 2 ai = 1 flo^3 20 EI oi 3 c/qL2 40 £/ The two-parameter Ritz solutions based on the mixed formulation are W2(. } 40E/ \ L L2/' 2 V L/ ' W2(L) = <?oL4 Af2(*) = - M2(0) = - 8£7 ' (10.76) Note that maximum values of both deflection and moment coincide with the exact values, and that the expression for the bending moment coincides with the exact solution. Although the maximum deflection predicted by the mixed Ritz approximation coincides with the exact solution, the deflection obtained violates the slope boundary condition at x — 0 (which is a natural boundary condition in the present case). Thus, the mixed variational formulation gives more accurate solutions for the force variable because they are approximated independently; however, the accuracy of the displacements is somewhat degraded. It can be improved by increasing the number of parameters (note that the polynomial degree used for wq is only 2 in the mixed formulation, while it is 3 in the displacement formulation). In cases where the analysis is carried out to support design based on stresses, it is desirable to have accurate values of the force variables. Table 10.1 contains a comparison of pointwise deflections, slope, bending moment, and shear force obtained by the two approaches with the exact solution. 10.4 MIXED FINITE ELEMENT MODELS OF BEAMS 10 A1 The Euler-Bernoulli Beam Theory Governing Equations and Weak Forms Here, we develop a mixed finite element model of the equations governing the Euler-Bernouili beam theory. Consider the following equations of the Euler-Bcrnoulli beam theory [see Eqs. (10.62a,b)
S 3 £ o o CO > O C 1.1 (0 to Z o L_ < MUL O LL _J < Z o ATI VARI D LU X 2 00 u> cj> £ 8 CO 7i to ■o •o (D N 75 a_ O C <D Ul r ■a a> seal t o C 0 CO TO a £ o O r- © ^ 0) .a TO H c 2 o a is C £ o o r* " m = O is TO 0 3 ^ (O C * £ c ~ o a '•5 c IS 2| Q. > Q.T3 TO 0 *1 DC = w "5 <u £ * il £iT TO M TO* H 4= CO £ E a o s P5 a o q=l a> Q > S x: Oviomomomomo ooN^oocJoo^riNOO in •* m cs -h r- ooooo ooooooooooo I I I I I I I I I I I r-r-t-r-c^r-c-r-r-cocn Tfc^mc^c^i— ^oooo o o © o o d o o o d d I I I I I I I I I I t 0"no<no<nomo«no OO<N"d-00CN00-*J-<NOO vi-Trcn<Nr-i,—.ooooo I I I I I [ I I I J I oc*-oocooooi>or-o ooooooooo — —i oddoooddo'dd oor--»nor~-omr-oo oc^r-M5oo—«r-enooom ooo^cSTtmt^avOcN OOOOOOOOOr^-i Ocor-^^cn-stcor-coo CMooocorr^in-HflO»ri ooo-Hc^Tj-tnr-a^oiN OOOOOOOOO— —< dddddododdd o J* ooddddddddd ^dddddddddd o*no*no^Omo«no »nVDooo\T— c^^'nr^ooo 0000^— ^^-h—'i— <N ododdddddoo I I I i I I I I I I I OtNovir^ooorJc^mr- CO\ncN\0iOQ0\ncS\0 gcnr-ocNTt-vo^or-r-vo OO—I^^T-HT-Hr^r-^-1 ddddodooooo 0(NninhoooMn<nh O <n i— QjO>0\£lN^\D\D OOO-H^—«-H-Hr-.^-^-| dddddododdd O^H<Ncr>'3"|nNOr-*oocJ\0 ooodddddddd II I CO 5i k*\ ^ cm ^3 ^5 0) £ a * bl •3 '-4- +-» O G ia E *■> is «* (D 1-4 > a> * fe aj O S .2 "B & c a> ° ca o ^ J'-fc S IS § & CO ^ 3§ c ■I42 OH g § 3 S CO us s-8 E > o GO J±h c^ rp ^ ^ . if & s O 43 i-> C3
MIXED FINITE ELEMENT MODELS OF BEAMS 519 We have [see Eqs. (10.68) and (10,69)] (10.78a) fx*> (dv\ dM \ - , - Jx0 \dx dx J dv2dw0 M\ V2 dx - V2(xa)®\ + V2(Xb)®2, dx dx EI (10.78b) where (i>i, U2) are the weight functions that have the inteipretation of (see Example 10,3) virtual deflection 8wq and virtual moment <5M, respectively, and *~(£L- " _ (dM\ 2~\dx )x=Xb ' 01 = (**\ , e2 = Lip) . \ dx )x=Xa \ dx JXSSXb (10.79) Weak Form Finite Element Model The weak forms (10.78) suggest that both wo and M may be approximated using the Lagrange interpolation. Suppose that wq and M are approximated as wq(x) to "^ Wffc (*), M(x) to ^ Mifi (*), 1 = 1 1 = 1 (10.80) where ($,-, \j/j) are the Lagrange interpolation functions of different degree used for wo and My respectively. Substituting Eq. (10.80) for wo and M, and V[ = <pj and v2 = i/j into Eqs. (10.78a,b), we obtain % E-^l-* Ja- d;c /</ m rfjc-g; (/ = 1,2, ...,w), (10,81a) diff} f ^ rf0y (i = lt2t...,n). 1 e7 iME*^ d*-(-l)/+l0f (10,81b) or X] KfjM] = ft + Of, £ K^w) - J^ G^M) = (-ly+^f, (10.82a) 7=1 ;=i where *5-jC*^- ^r^- ^/>- (10.82b)
520 MIXED VARIATIONAL FORMULATIONS In matrix notation, we have [Ke]{Me] = (/'} + {Q% [Kef{we] - [Ge]{Me] = {<=><% (10.83a) or where [0] [K*] [Ke]T -[Ge] ife} + {Qe) i+l&e 0f = (-l)'+1©; (! 0.83b) (10.83c) The pair of matrix equations (10.83a) in nodal values of the displacements and moments may be reduced to a single matrix equation in [we] and {(H)e). Solving the second equation in (10.83a) for [Me], wc obtain {Me} = [GTl([Kef{we} - {B*}). Substituting the result into the first equation in (10.83a), we obtain [JH[GT,[JH"IV} - [Ke][Ge]-l{®e) = {f) + {Qe}. We can now write Eqs. (10.84a,b) in a single matrix equation as '[Ke][Gerl[Ke]T -[Ke][Ge]~l -[Gerl\Kef [Ge}~} lwe) if) + {Qe) -{M<} (10.84a) (10.84b) (10.85) For the choice of linear interpolation of both wq and M, we obtain [Ke\ = - He 1 -1 -1 I [Ge] = 2 i 1 2 [Ge] e->~\ 2E(.ie r he [- 2 -1 1 2 6Eele \Gerl\Ke]J = in = \f! l/f hi :-i -a- [Ke][Ger][Ke]T = \2Eele hi 1 -1 -1 1 [GT,l*'lT = ([JH[<?T1)T. (10.86) Hence, Eq. (10.85) becomes 6 —3he —6 — 3he 2E„L I -3he 2h2e 3he hi —6 3he 6 3he _-3he hi 3he 2h2e j f/fl 0 ft 0 + 21 21 2! [Oil (10.87)
MIXED FINITE ELEMENT MODELS OF BEAMS 521 Interestingly, the stiffness matrix of the mixed finite element model with linear interpolation of both wq and M is the same as that of the displacement finite element model derived in Section 9.3 using the C1 (Hermite cubic) interpolation. However, the load vector differs in the sense that the mixed model does not contain contributions of distributed load q{x) to the nodal moments. Weighted-Residual Finite Element Models It is possible to construct a mixed finite element model of Eqs. (10,77a,b) using a weighted-residual formulation. As will be seen shortly, this model requires higher-order approximations of both wo and M because they must satisfy both essential and natural boundary conditions. This leads to a complicated set of finite element equations that are not practical. Here we present the main ideas behind the development of the model. The weighted-residual statement of Eqs. (10.77a,b) is j2„ fXf> ( d2w0 M\ , o=l »{-■*?-v**- (ia89) where (v\, 1*2) are the weight functions. A close examination of the above statements indicate that v\^ M and i>2 ~ wq. Using approximations of the form WW * Y, A'^0) W> M^ * J2 Ay^2)(*), (10.90) 1=1 i=i we obtain the following Galcrkin (i.e., v\ ~ <pj and V2 ~ <p\ ) finite element model: [[«"] [£>£jj({Ae)| {{0}}' (WM) where A? For Hermite cubic inteipolation of both wq and Af, we have <p\{) — (p^2). The coefficient matrix in Eq. (10.91) is not symmetric.
522 MIXED VARIATIONAL FORMULATIONS The least-squares finite element model of the pair (10.77a,b) is based on the variational statement [*» I" fdhvo M\2 (d2M \ „ fx»\ (cfiwo M\/d2Sw0 SM\ (d2M \d28M~\, =2L r\^+^){-d^+'^)+P2{^+vii^\dx' or . fx"d28w0 (d2w0 M\ « rV(To,./^2™0 M\ EIp2d2SM /d2M \ dx. (10.93) (10.94) where (pi, P2) are weights used to make the entire statement have the same units. Substituting the approximation (10.90) into Eqs. (10.93) and (10.94), wc obtain [B«\T [De]\\{A'}\ ({0} (10.95) where , {«> (Elp-td2^ cl\f m (2)\ e p^f ^ Af ** d2^ d29) en a!*2 ■</;c. (10.96) Obviously, p\ and p2 should be selected such that EI(p2/pi) has the dimension ofl/L2. We close this section with the comment that finite element models other than the displacement finite element model are not commonly used in practice. In fact, finite element models based on the Timoshenko beam theory that accounts for transverse shear strain yxz may be used to analyze beams irrespective of whether shear deformation is significant. The next section is devoted to the discussion of Timoshenko beam finite elements.
MIXED FINITE ELEMENT MODELS OF BEAMS 523 10.4.2 TheTimoshenko Beam Theory Governing Equations The governing equations of the Timoshenko beam theory for pure bending arc given by (see Eqs. (9.71) and (9.72); also see Reddy [J7-19]) Here q{x) denotes the disti'ibuted transverse load, E Young's modulus, G the shear modulus, A the area of cross section, / the moment of inertia, and Ks the shear correction factor. Displacement finite clement models of Eqs. (10.97) and (10.98) were presented in Section 9,4.2. Here we shall discuss a number of mixed finite element models of these equations or their equivalent. Genera! Finite Element Model The mixed finite element model to be discussed here is based on the stationary functional (10.53) in which the displacements (u;o, <p) and strains (kxx, yxz) are treated as the dependent variables (see Example 10.2). The Euler equations of this functional are given in Eqs. (10.55)^(10.58). The weak forms associated with Eqs. (10.55)-(10.58) over a typical element Qe — (*„,**) are I lGAKs—JLyxz _ 8woq\ dx _ Va&w(xa) - VbSrvo(xb) = 0, (10.99) / b (eIC—^kxx + GAK,8^yxzj dx - Ma84>(pca) - Mb&4>(xh) = 0, (10.100) jXh EISkxx (^ - kxx) dx = 0, (10.101) jXt GAKs&Yxz (^ + 4> ~ Yxz} dx = 0, (10.102) (10.103) where V{ = [-GAKsYxzlc*, V2 = [GAKsYxzU> MX = I -EItcxx]Xa , M2 = [EIicxx]Xb. Let the variables (tun, <!>, kxx, yxz) be approximated as m n p q w^j2+?)wh ^J2^?^er <^E^3)/q' y*z~Y,*Vtp ;=i y=i j=i y=i (10.104)
524 MIXED VARIATIONAL FORMULATIONS where (W*t $y, JCcj, Tp are the nodal values of (wq, <p, icXXi yxz) and if^(x) (a — 1, 2, 3,4) are the associated interpolation functions whose choice is yet to be made. Substituting (10.104) into (10.99)-(9.102), we obtain the following finite element model: 101 fO] [01 \Aef [0] [0] [Be]T [Ce]T where r0| [A«] ' \.Be] [C*\ ~[De] [0] [0] -[Ge] • '{We}' {Te} ► = < {F<y [0) (0) fl»J • + ' '{V«}1 (0) (0} J (10.105a) xbGAK^r^ ax J Jx0 Jxa <% = f Jxa GAK^ff dx, v\ — Yct » Ve = V? Bf, D?.= f"El^tyfdx, J.Xa Jx> -,(!> Qlry'dx, Aff = A£, M* = M£. (10.105b) A couple of observations are in order concerning the finite element model in Eq. (10.105a), We note that \Ae] is a vector [Ae] when yX7 is approximated as a constant, Fq. In addition, the first equation of (10.105a) has the form GeAeKs r-f 0 0 [ i >r0=. fFf] .n .K. \ +' fv'l 0 0 . Ill J (10.106) when ido is interpolated using quadratic or higher-order polynomials. The nonzero entries correspond to the deflection degrees of freedom at node 1 and node m. For linear interpolation of u;o, wehavem = 2andEq. (10.106) is correct. However, when m > 2, Eq. (10.106) implies that Ff = 0 for i — 2,..., m — 1, which, in general, is not true. Thus, either the distributed load is zero or it is converted to generalized point forces at the end nodes through the Hermite cubic polynomials <pj of Eq. (9.58). In the latter case, the force components can be added to V£ and Vf and the moment components to M\ and M\ at nodes 1 and m, respectively. ASD-LLCC Element For linear (L) interpolation of (wo9<f>) and constant (C) representation of (kxx, yxz), an(i f°r constant values of EI and GAKS> the element
MIXED FINITE ELEMENT MODELS OF BEAMS 525 equations become (in = n = 2 and p = q — 1) GeAeKs "H:a+ Vf| E'ele ^+g^t|!)'-J";) Km 2 {-i i} $« \<$>% - />«r§ = o, - /^/cg = o. (10.107) (10.108) (10.109) (10.110) Solving Eqs. (10.110) and (10.109) for /Co and T0 and substituting into Eqs. (10.107) and (10.108) (i.e., condensing out /Co and To), we arrive at the following 4x4 system of equations: GAK, 4 —2he _4 -2he -2he -' A2fl + 4A) 2h 2he 4 A*(1-4A) 2/t -2A " (1 - 4A) 2he (1+4A) i W^ *i W| .*2- ► — < Kj 0 ?! ;0; , 4- . Gil GSI G5 GSJ 4/z, (10.111) where A = Eele/GeAeKsh2e, These arc exactly the same equations obtained in the displacement formulation with the linear interpolation of wq and 4> and using one- point Gauss quadrature to evaluate the shear stiffnesses. The element was refeired to as the reduced integi'ation element (RTE). Thus the assumed strain-displacement formulation eliminates the need for reduced integration concepts. ASD-QLCC Element A quadratic interpolation of iuo, linear interpolation of <f>, and constant representation ofKxx and yxz will also yield the same stiffness matrix as in Eq. (10. II1), However, as noted earlier, the distributed load q must be calculated using Hermite cubic polynomials (pi (i — 1, 2, 3,4): •» =/ ,('0 q(x)<pi(x)dx, (10.112) and the force components of q\ } and q\ must be added to V/ and Vg, and the moment components q^ and q\ to M\ and M|, respectively: f 7f-' Mf \n l*ff. r k — \ V2' + *,W A/|+^' (A) (10.113)
526 MIXED VARIATIONAL FORMULATIONS Thus the load vector is the same as that of the Euler-Bernoulli beam equations. This element was referred to as the consistent interpolation element (CIE). ASD-HQLC Element Suppose that the distributed load is represented using Eq. (10.112). A Lagrange cubic interpolation of wo, quadratic interpolation of <p, linear interpolation of kxx, and constant representation of yxz yields the equations GeAeKsl 1iJro=u{L (10.114) ^>p. 'p. k*ele -5 -1 [-5 4 L i -1- -4 5 J \K\\ GeAeKshe CI 4}r0=- I'J \ Me] 0 Ml 4 -4 6 PS {-1 1) ♦S" 4 1} *? 4>< *?}-Aerg = o, (10.115) (10.116) (10.117) where the end nodes of the element arc designated as "1" and "2," and the middle node as "c," and the interior nodal degrees of freedom associated with w are omitted as they do not contribute to the equations. Solving Eq. (10.116) for {K} andEq. (10.117) for To, and substituting the result into Eqs. (10.114) and (10.115), we obtain (10.118) *!1 (10.119) where Se = GeAeKs and A = (EeIe/Se). The second equation of (10.119) can be used to eliminate <Pec from Eqs. (10.118) and (10.119). Thus, we obtain -6 I I Wf I I-3A. 6 Wwfrl 3/», "-3A, -3A. 3A 3/? 1I3M- -3/t 3/). 2/^A tit 2h2eX M)- (10.120) (10.121)
MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 527 Superposing Eqs. (10.120) and (10.121), we obtain 2EeIe ixh] 6 -3he -6 _-3he -3he -6 2h2X 3he 3/ze 6 h2£ 3he -3he~ h2ei; 3he 2h2X_ ■ wf' $>\ Wf w2\ 1 f Ve 1 ! \K \n 1 U| (10.122) The stiffness matrix is the same as that of the superconvergent clement derived in (9.97); however, the load vector is different. It is the same when either the applied load q is element-wise uniform or the load vector is computed using Eq. (9.96b). It should be noted that the degree of the polynomial interpolation used for iuo does not enter the equations presented in all models discussed in this section. However, the load representation implies that wq be interpolated with Hermite cubic polynomials <pf of Eq. (9.58) or <pf of Eq. (9.94), 10.5 MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 10.5.1 Preliminary Comments As mentioned in the previous section, a conforming plate finite element based on the displacement formulation of the classical plate theory is algebraically complex and computationally demanding. A quintic polynomial would satisfy the continuity requirements at element interfaces, and results in 21 degrees of freedom per element. The continuity requirements can be relaxed by reformulating the fourth- order equation (8.143a) as a set of second-order equations in terms of the deflection and bending moments. Such a formulation is called a mixed formulation. Here we discuss two such models for the static bending of orthotopic plates (see Reddy and Tsay [20,21]), There exist many other papers dealing with mixed and hybrid formulations based on modified total potential energy functional, modified complementary energy functional, and variants of the Hellinger-Reissner and Hu-Washizu principles. 10.5.2 Mixed Model I Governing Equations Consider the following set of equations [cf. Eqs. (8.123a) and 8.132)]: d2Mx- ^ 2d2Mxy | dx2 dxdy Mxx = -( 32M ^ 82wq n (10.123)
528 MIXED VARIATIONAL FORMULATIONS *" = -Hsr + D^)' (,0-,24) S2Wq Mxy = -2D66—-, dxdy where Djj are the bending stiffnesses of an orthotopic plate [see Eq. (8.123b)]: Du - Qu — 9 (ij - 1,2,6), (10.125a) and the elastic stiffnesses Q{j are defined in terms of the principal moduli (E\, Z^X shear modulus Gj2, and Poisson's ratio v\2 as Q\\—-. > Gl2 = - , Ql2 l-V12V2l 1-V12V21 I — P12V21 ,ln104-Uv (10.125b) £?66 — Gi2, V2| = ^12 — - Substitution of Eqs. (10.124) into Eq. (10.123) leads to the fourth-order equation in Eq. (8.143a), whose finite element models were discussed in Section 9.5. Here we formulate a mixed finite element model of Eqs. (10.123) and (10.124), which involve second-order equations among the dependent variables (wn> MXXt Myyt Mxy). First we express Lhc three equations in (10.124) in the alternative form 32wc\ , - _ -^ = -(D22M,.,-D12M„,), «2 -j^- = - {bnMxx - DnM„) , (10.126) t dx'dy where 3 wq t Dij = ^T. Do = DX1D22 - D2n. (10.127) Weak Forms The weak forms ofEqs.(10.123)and(10.126)ovcra typical element £2e can be derived using the inverse procedure described in Section 7.4.1. We have A t \Mwq (BMXX , BMxy\ , dSwo (BMyy dMxy\ s ~\ , , f \(BMXX SMKV\ /'dMxv dMYV\ 1
MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 529 dxdy 0 = / ^r~r^ " SM^(°22MXX - D[2Myy) Jne I ox dx f S\V{) - <2> 5MLl—n, ds, f [dWQdSMyy ,- ' (10.129) dxdy — (t 8Myy ——n v as, By (10.130) It is clear that the primary variables of the formulation are WQ9 MXX, Myy, MXy, and the secondary variables are (dMxx dMxy\ (dMxy 3Myy\ (10.131) (10.132a) dwp dx -nx* dwp dy -fly, 3wq dwQ dx -nv + By (10.132b) Note that the natural boundary conditions involving the bending moments in the displacement formulations become the essential boundary conditions in the mixed formulation. The functional associated wilh the weak forms (10.128)—(10.131) is given by [see Eq. (7.53)| /f (A) = f\j (~D2iM2xx 4- 2bnMxxMyy - DuM*. D66 xy) i {Mxxexnx + Myyeyny + Mxy(0xny + 0ynx) + Q{lw0]ds» (10.133a) where 0X = dwp 3X ' 0y = ^ > dy A - (W[)9MXX,Myy,MXy). (10.133b) The condition SJ[ = 0 gives Eqs. (10.123) and (10,126) as the Euler equations.
530 MIXED VARIATIONAL FORMULATIONS Finite Element Model Let (w^ Mxx, Myy, Mxy) be interpolated by expressions of the form wo = 52«^,(1\ MA,V = J^Mxi^\ Myy - £A^f \ i=l /=! i~l Mxy = J^Mxyif?\ (10.134) /=! where ifrr1 \ (a = 1, 2, 3, 4) are appropriate interpolation functions. Substituting Eq. (10.134) into Eqs, (10.128)-(10.131) [or substituting Eq. (10.134) into 7i and setting the partial variations of J\ with respect to wj> MXj, Myj> and Mxyj to zero separately], we obtain (10.135) n*11] symm. [Ki2] [K]3] [ATl4]1 [K22] [K23] [K24] [K33] [K34] [Ku]_ f {w] 1 Wv} {My} [{MXy)\ \{F'} {F2} \{F3} [{F4} where K}}=0t ij= 1,2,..., r, „ f dxlr} of2. 7 Jae 3jc dx „ f dxfrj of3 1 Jti, Jy .,s, dy i = l,2,. ,r; j = 1,2, ...,p, i = l,2,...,r; j — \,2,...,q, jx ay oy ax i K22 = j {-Dn)*ftfdxdy, i, j = 1,2,...,s, ^Lm+ff)^ Kj?= [ (-Dl2)^2fjdxdy, i-1,2, ...,s; j = l,2,...,p, Kff = Q, i = l,2,...,s; j = \,2,...,q, KV = / {-DuWftfdxdy, i,j=\,2,...,p, (10.136) Kff = 0t i = l,2. 5 • • • > J-'J y = l,2,...,9, Jcie J
MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 531 F/ = / qyfrj- dxdy + <b Qnf\ dst i = 1, 2,,,., i\ Ff=fb 9xnxjr?ds, i = l,2>...,s, F? = <b Oyfiy \frf ds\ z = 1,2,..., p, F*=j> (0ynx + exnyW? ds, 1 = 1,2,...,$. An examination of the weak forms in Eqs. (10.128)—(10.131) shows that the following minimum continuity conditions are required of the interpolations functions irf (a = 1, 2, 3, 4) used in Eq, (10.134): Vy — linear in x and linear in y, tyf = linear in x and constant in y, 'tf/j — linear in y and constant in a- , tyf = linear in x and linear in y. (10.137) For a rectangular element, the interpolation functions ijrf that meet the minimum requirements are tf?=l a b b and ty} and \/rf are the bilinear functions of a rectangular element [see Eq. (9.119)]. The rectangular element associated with this choice of interpolation functions is shown in Fig. 10.1b. The coefficient matrices in Eq. (10.136) can be easily evaluated for this element. For instance, we have (a and b are the side lengths of the ^^~ 4 ••••"• '.'■■: at/each node.: t."-,:- (a) Mixed model IA Mx{ Ms yz trr T W ' ^—— "-.. w0, MryQt four :: corner, nodes ' ffi —f ^7 T iu Mv: (b) Mixed model IB Figure 10.1 Mixed plate-bending elements based on CPT. (a) Rectangular element I A. (b) Rectangular element IB.
532 MIXED VARIATIONAL FORMULATIONS rectangular element): 2a [K22] = -D22^ r i -l -l i -n i i -ij [K"l a 2b ' 1 1 -1 -1 -1 -1 1 1 2 1 1 2 r b W] = -DU 2 1 1 2 (10.139) etc. The element stiffness matrix is of the order 12 x 12. Another choice of \jrf is provided by (sec Fig. 10.1 a): ^r.1 = \jrf — i//? = -ftf = bilinear functions of a rectangular element. (10.140) Computation of the coefficient matrices is simple and straightforward. The resulting stiffness matrix is of the order 16 x 16. 10.5.3 Mixed Model II Governing Equations A simplified mixed model can be derived by eliminating the twisting moment Mxy from equations (10.123) and (10.126). We have [cf. Eq. (8.132)J 82MX 47)66 34wq B2M- dx2 8x2dy2 d2w0 + yy By2 XvWq •(' + A/3 JV, dx2 + 2N6 32wo dxdy + N2 d2w0 dy2 ) =1, (10.141a) _Ji = _ (Dl2Mxx + DnMyy) d2u>o dy2 = -(Di.2M.xx+DuMyy), (10.141b) where, in the interest of solving the natural vibration frequencies and buckling loads, the contributions of the inertia term (neglecting the rotary inertia) and in- plane compressive and shear forces are added through the parameters kv and Xi,, with do = Ph) A„ = !r,ar X,,= N„ = Nyy = Nx, N\ N2 N6 (10.141c) Here h denotes the plate thickness and (Nxx, Nyy, Nxy) the in-plane compressive and shear forces.
MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 533 Weak Forms The weak forms of Eqs< (10.141a,b) over a typical element Qe are 0: w0 dMxx dSw0 BMyy B2Sw0 B2w0 \ — + — ^- + 4D66 —-r- ^-^~ ~ (?SwQ dxdy dx dy dy dxdy dxdy } wo^wo + \b ' BSwqBwq dSwodwo jVj —-- ■ — \- i\2- dx dx By By + N*\^^ + ^^)\\dxdy *(" dx dy dy dx 2 )])• f Sw0Q„ - 2Df,6-^-(80xny + 80ynx)] ds, (10.142) °-/j£^-«*.<»»«»-wh dy — d> SMyxBxnx ds, 0 = L I dy dy "SMyy (DnMyy " DnM**)\ dxdy. — (I) SMyyOylly ds. The primary and secondary variables of the formulation are WO, MXXi Myyy (10.143) 0xnx = -r— nx, 9yny = ——ny, dx oy (10.144) (10.145a) (10.145b) where V„ is the effective shear force (Kirchhoff free edge condition) [cf. (8.142)]: dMns V„ = Q„ + ds Qn = QxTlX + Qylly, (10.145c) Qx = Qx~*b[Ni~+N6: dw0\ The functional associated with Eqs. (10.142)-(10.144) is J2(wo, Mx, My) = J ^(-D22M^X + 2DnMxxMyy - DnM^, - X„wg) (10.145d) ~2kh
534 MIXED VARIATIONAL FORMULATIONS dWQdMXfX t SwodMyy + Bx Bx By By rd2w0\2 ] , , ■2D66t—-) -qw0\dxdy - f (MXX0XHX + MyyOyfly + V„ Wq) <tS. (10.146) Finite Element Model An examination of the weak forms (10.143) and (10.144) reveals thatthe minimum continuity conditions of the interpolation functions yjrf (a — 1,2,3) are the same as those listed in Eq. (10.137). Therefore, the interpolation functions of Eq. (10.138) or Eq. (10.140) can be used. The corresponding rectangular elements are shown in Fig. 10.2. The finite element model of Eqs. (10.142M10.J44) is obtained by substituting the approximations in Eq. (10.134). We obtain r[Jr' ]] [Ki2\ [KB] [K22] \K2i] [symm. [AT33j_ = • \{Fi)] {F2) [{F3}} ' +K • < I \[S]{w} (0) I «o } {w} 1 (M,) {My}\ > \ t + h < \[G]{w}] [0} } [ {°} 1 where 9V/ d2f} K}j = 4D66 / Kn= f dlt°Ji 42=/(- Bx Bx djr} Bf) By By dxdy, ij= 1,2,..., r, dxdy, i = l,2, ...,r; .7 = 1,2,. dxdy, D22)if?ir]dxdy, i = 1,2 r; j = lt2, i,j = 1,2, ...^j, (10.147a) >s9 4. ■■:"■: ;:.:;: "3 . : W& M^ My.;.■'■■■ ■: at eaich node.. (a) Mixed model I1A K .-■■3 >$ •7'.-. 8 at/each node5T (b) Mixed model IIB Kim. A£ ■J-2 4- .•.:••• " v-:- ;.-...■ 3. $ . w/0at.faur . V ?■.!: comer nodes' MvZ Myl (c) Mixed model IIC Figure 10.2 Mixed rectanglular plate-bending elements based on CPT. (a) Mixed model IIA. (b) Mixed model IIB. (c) Mixed model IIC.
MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 535 Kff = f {~Du)fff) dxdy, 1 = 1,2,..., s; j=\,2,...,p, Kff = f (-Du)fltfdxdy, l,j = 1,2,..., p, i = / Hfi / dxdy + (p Vn \jf/ <:/.?, i = 1,2,..., r, f)2 = A Oxnxtfds, i = 1,2 j, J}3 = A Oynyxlrfds\ / = 1,2 /?, 5'7 - / ^/^/ rf*<ty • z\ j = 1, 2,..., r, a^i a^j av// »*j /a*/ a*j a*/ a^j ox ox oy oy \ ax ay ay ox ij = 1,2,...,}-. This completes the development of the mixed finite element models. A number of plate problems are analyzed using the mixed rectangular elements developed in this section, and the results are discussed in the examples presented below. Example 10.4 The four mixed plate elements discussed above are used to analyze bending of rectangular plates with simply supported and clamped boundary conditions. The bending analysis is carried out using the following types of mixed elements. Mixed Model I Bilinear approximations are used for deflection w and moments Mv, My, and Mxy (mixed model IA). Thus the clement has four nodes and four degrees of freedom at each node (see Fig. 10,2a), resulting in a 16 x 16 element stiffness matrix. Mixed Model II This model is subdivided into three models depending on the type of interpolation used. In mixed model 1IA, bilinear interpolation of wq, MXXy and Myy is used; in mixed model IIB, quadratic (serendipity) interpolation is used (see Fig. 10.2b,c), These elements have stiffness matrices of order 12 x 12 and 24 x 24, respectively. In mixed model TIC, bilinear approximations are used for deflection wq and partial linear approximations for Mxx and Myy. The four corner nodes each have one deflection degree of freedom. Midnodes on the sides perpendicular to the y-axis have one degree of freedom Myy, and midnodes on the sides perpendicular to the x-axis have one degree of freedom Mxx. This element results in a 8 x 8 element stiffness matrix.
536 IWXED VARIATIONAL FORMULATIONS Table 10.2 Maximum deflection and bending moment of simply supported, square, isotropic (v = 0.3) plate under uniformly distributed load . „ , Mixed Models Mesh Size IA IIA IIB IIC Herrmann [22] Central Deflection, (w0D/rq0a4)U)2 (0.4062)a 1 x 1 0.4613 {16)b 0.3906 (12) 0.3867 (24) 0.4943 (8) 0.9018 (6) 2x2 0.4237(36) 0.4082(27) 0.4053(63) 0.4289(21) 0.5127(15) 4x4 0.4106(100) 0.4069(75) 0.4062(195) 0.4117(65) 0.4316(45) 6x6 0.4082 (196) 0.4066 (147) 0.4062 (399) 0,4087 (133) 0.4174 (101) Bending Moment at the Center, (Mxx/qoa2)\0 (0.4789)C 1 xl 2x2 4x4 6x6 0.7196 0.5246 0.4891 0.4834 0.6094 0.5049 0.4849 0.4851 0.3813 0.4818 0.4788 0.4789 0.3482 0.4498 0.4721 0.4579 0.328 0.446 0.471 0.476 aExact solution from Eq. (8.162a). bNumbcr of degrees of freedom in ihc mesh of a quarter plate model. cExact solution from Eq. (8.162b). Table 10.3 Maximum deflection and bending moment of clamped, square, isotropic (v = 0.3) plate under uniformly distributed load Mesh Size 1 x 1 2x2 4x4 6x6 1 x 1 2x2 4x4 6x6 IA 0.1664 (16)b 0.1529(36) 0.1339(100) 0.1299(196) Mixed Models IIA IIB IIC Central Deflection, (w0D/q0a4)H}2 (0.1265)a 0.1563 (12) 0.1466 (24) 0.2278 (8) 0.1480(27) 0.1260(63) 0.1627(21) 0.1325 (75) 0.1264 (195) 0.1359 (65) 0.1292 (147) 0.1265 (399) 0.1307 (133) Bending Moment at the Center, (Mxx/q§a2)\Q (0.231)» 0.5193 0.3166 0.2478 0.2374 0.4875 0.2056 0.2487 0.2899 0.2248 0.2432 0.2443 0.2287 0.2339 0.2358 0.2290 0.2313 Herrmann [22] 0.7440 (6) 0.2854(15) 0.1696(45) 0.1463 (101) 0.208 0.242 0.235 0.232 aFrom Timoshenko and Woinowski-Kriegcr [23]; also see Example 8,12. ^Number of degrees of freedom in the mesh of a quarter plate model. The mixed finite element solutions are presented in Tables 10.2 and 10.3 for isotropic plates and in Table 10.4 for orthotropic plates. The orthotropic properties used are: Glass-epoxy plates: E\ = 7.8 x 106psi, E2 = 2.6 x I06pst> v[2 = 0.25, Gi2 = 13 x 106 psi.
MIXED FINITE ELEMENT MODELS OF THE CLASSICAL PLATE THEORY 537 Table 10,4 Maximum deflection and bending moments of simply supported, square, orthotopic plates under uniformly distributed transverse load [w= {wQH/q0a4^03, H= £?12 + 2D66; Mxx = (My^/q^a2)^ Myy « {Myyfq0a*)M)2] Mesh 1 X 1 2x2 4x4 6x6 8x8 Exacta i x 1 2x2 4x4 6x6 8x8 Exacta w 2.9737 3.1041 3.0930 3.0901 3.0890 3.0876 0.9348 0.9346 0.9220 0.9208 0.9204 0.9200 aFromReddy[24]. IIA Mxx Myy w Gkiss-Epoxy Plates 0.9435 0.8078 0.7737 0.7676 0.7654 0.7628 3.6290 2.8937 2.7851 2.7685 2.7628 2.7556 2.9412 3.3081 3.0872 3.0876 3.0876 3.0876 Graphite-Epoxy Plates 1.6152 1.3570 1.2845 1.2740 1.2704 1.2659 0.6720 0.2860 0.3143 0.3252 0.3270 0.3271 Table 10.5 Natural frequencies {Xv = co2ph a*/D) of \ clamped square isotropic (v = 0.3) plates Mesh 1 x 1 2x2 3x3 4x4 Hinged (Exact: 389.636) IIA 576.000 431.529 407.809 399.766 Graphite-epoxy plates: im 408.439 390.461 389.792 389.685 £l = 31,8x 106psi, vi2 = 0.3J, 0.9207 0.9224 0.9198 0.9199 0.9200 0.9200 IIB Mxx 0.57 J 8 0.7687 0.7627 0.7628 0.7628 0.7628 0.9491 1.3094 1.2661 1.2658 1.2658 1.2659 Myy 2.2297 2.7849 2.7556 2.7556 2.7556 2.7556 0.3800 0.3370 0.3274 0.3272 0.3272 0.3271 simply supported (or hinged) and Clamped (Exact: 1295.28) HA 1440.000 1361.283 1325.837 1312.403 E2=L02 Gn = 0.96 x 106psi, x 106psi. IIB 1355.077 1303.380 1298,051 1296.114 The present finite element solutions for center deflection and bending moments are compared with various finite element solutions available in the literature. From (he numerical results it is clem' that mixed models IIA and IIB are the best among the mixed models discussed here. Example 10*5 Fundamental frequencies and buckling loads of plates are presented for (he following cases: (1) vibration of simply supported and clamped square plates (Table 10.5);
538 MIXED VARIATIONAL FORMULATIONS Mode 1 2 3 4 1 2 3 4 Table 10.6 Mesh 2x 1 4x2 Frequency parameters (k]/2 = coePy/ph/D) of cantilever plates Present Analysis HA 3.27 14.46 22.89 51.28 3.41 14.55 22.64 49.93 IIB 3.46 14,60 22.49 49.07 3.44 14.76 21.51 48.13 Anderson [25] 3.99 1.5.30 21.16 49.47 3.44 14.76 21.60 48.28 Barton [26f Ritz 3.47 14.93 21.26 48.71 3.47 14.93 21.26 48.71 Test 3.42 14.52 20.86 46.90 3.42 14.52 20.86 46.90 P1unketta L27] 3.50 14.50 21.70 48.10 3.50 14.50 21.70 48.10 aResults are independent of mesh. Table 10.7 Buckling coefficients (lb) for simply supported and clamped, square, isotropic (v = 0.3) plates under uniform edge loads Load Case (Exact)a Uniaxial (S: 4.00) (C: 10.07) Biaxial (S: 2.00) (C: 5.30) Shearb (S: 9.34) (C: 14.71) Mesh 2x2 3x3 4x4 8x8 2x2 3x3 4x4 8x8 2x2 3x3 4x4 8x8 Simply Supported HA 4.2095 4.0922 4.0517 4.0129 2.1048 2.0461 2.0258 2.0064 — 14.3220 11.1499 10.2898 IIB 4.0063 4.0012 4.0004 4.0000 2.0031 2.0006 2.0002 2.0000 14.4124 9.6758 9.3816 9.3403 Carson and Newton [28] 4.0158 4.0032 4.0010 4.0007 — — — — 10.016 9.418 — — I1A 11.530 10.688 10.412 10.156 6.0592 5.6339 5.4868 5.3487 — 25.7773 18.5902 16.7104 Clampec IIB 10.2009 10.1107 10.0866 10.0748 5.3746 5.3233 5.3102 5.3040 29.3967 15.4640 14.8361 14.7123 1 Carson and Newton |28] — __ — — 5.3622 5.3660 5.3271 5.3054 23.264 15.043 — — "FromTimoshenko and Gere [29] and Rcddy [30]; S : bObtained using full plate model. - simply supported; C = clamped. (2) vibration of rectangular cantilever plate (Table 10.6); and (3) budding of simply supported and clamped square plates under various edge loads (Table 10.7). In summary, the mixed finite element models of plates give accurate results for deflections, stress, buckling loads, and natural frequencies. The mixed models give more accurate results when compared to certain conforming and nonconforming elements. Further, the mixed finite elements are algebraically less complex, and involve less "element forming" efforts.
EXERCISES 539 10.6 CLOSURE In this chapter, mixed variational formulations of problems of solid mechanics are presented. The construction of functionals associated with the Hellinger-Reissner and Reissner variational principles was developed for simple problems of beams. The use of the variational methods to determine approximate solutions as well as finite element models based on the mixed variational statements is illustrated via beam theories. Alternative mixed variational formulations of the classical plate theory were also developed and their finite element models were derived. Mixed formulations and associated finite element models of the first-order shear deformation theory are not included here, but they can be found in Refs. 31 and 32. There exists a vast literature on mixed and so-called hybrid formulations (see, for example, [33-35]). Finite element models based on these elements have received mixed success. Finite element models based on mixed formulations in which the strains are treated as independent variables (e.g., see Section 10.4.2) have enjoyed better success. EXERCISES Derive the Euler equations of the stationary functionals in Exercises 10.1-10.4. 10,1 w,v) = / (-— v.v + v-gradu-/uj — / quds — f Js2 Jsi dV n • v(w — u)ds. 10.2 103 10.4 /(",*,*) dx — Fqw(L). =f[¥(£)2+/-eH /(«!, "2. P) = / y7yiuU + UJJ)UU - PuU I dx\dX2 - I tjlli ds. I(w, 4>u4>2> *1.*2) 2 7a [V3*1/ ■ \3*2/ "a*i 3*2 * \9x2 dxj f r /3w \ fdw t \ xdxidx2+ Ma 1-01 1+^2(7—H-02)-.; dx\dx2-
540 MIXED VARIATIONAL FORMULATIONS 10>5 Derive the Hellinger-Reissner functional for the bar problem of Section 10.1.2 [i.e., Eqs. (10.2)-(i0,4) and (10.13)]. 10.6 Derive the stationary functional corresponding to the problem of minimizing the functional /(«,■) - / [fajj +Ujti)Uij - fiUj^dV - / fa; ds subject to the constraint w,-if- = 0. 10*7 (Washizu's Principle [4,7-10]). Derive the following generalized Washizu functional: Tlw(uiteij1(Xij)= I {[j(«i,j + ujj +umsjiimj) - et-j]oij j v + pty - fiiti) dV - I (itf — ii})ti ds - I till; ds, where £//> <r\j, «/» and /? are subject to independent variations, and ^ = ^(e/y, T) is the Gibb's free-energy function. 10.8 Derive the Euler equations of the Washizu functional by requiring that the actual state of equilibrium of a body subjected to mechanical and thermal loads is associated with a stationary value of the functional Tl\y in Exercise 10.7. 10»9 Modify the Hellinger-Reissner variational principle for elastic bodies acted upon by mechanical and thermal loads. 10.10 The mixed variational principle for axisymmetric bending of circular plates requires <5TT,„ — 0, where n,n(w0. Mr, M9) = n f'\ KsGi3h U + ^ J + 2Mr^ + *MB<f> + D22M} - 2D{2MrMe + DnM$ - Iqw^rdi% where a is the radius of the plate, Ks the shear correction factor, and Obtain the Euler equations of the functional. 10.11 The principle of minimum total potential energy for axisymmetric bending of polar orthotopic plates according to the first-order shear deformation theory
EXERCISES 541 requires 8T\(wo, 0) = 0, where Sn(Wo,0)=f[(D)1^+/>12J)^+i(z),2^+z)2^)^ + Ass {»%)(»+<7*)-<-. rdr (a) where b is the inner radius and a the outer radius. Derive the displacement finite element model of the equations. In particular, show that the finite clement model is of the form (i.e., define the matrix coefficients of the following equation): [ IK11] [K*2]] Uw}\ _ \{F}\ 12 Derive the mixed finite element model for the annular plates using the following variational statement: ISM, + ^ + (DuMe-D]2Mr) SM$ — cj8wq\ rdr, whereQ = AssW+dwo/dr)andDtJ = Du/D0> D0 = DnD\2-DuD22> Show that the linitc element model is of the form " [*"] [Kn] [03 |01 " \K22] [K2*] [K24] symmetric IA"33] \.K34} [Ku]_ [ lw) • J {</>} j [Mr] [{Mo}, '{F} {0} {0} .(0} 13 Derive the functional in Eq. (10.133a) by identifying 5(A, SA) and 1(8A), where A = (u>o, Mxx, Myy) Mxy), and then using Eq. (7.53): J(A) = ±fl(A,<SA)-/(5A). 14 Verify the element matrices in Eq. (10.) 39). 15 Evaluate the element matrices in Eq. (10.147a,b) for mixed mode) HA.
542 MIXED VARIATIONAL FORMULATIONS REFERENCES 1. Oden, J, T., and Reddy, J. N., Variational Methods in Theoretical Mechanics, 2nd edition, Springer-Verlag, Berlin (1982). 2. Reddy, J. N., Energy and Variational Methods in Applied Mechanics* John Wiley, New York (1984). 3. Hellinger, E., "Die allegerneinen Ansatze dcr Mechanik der Kontinua," Art. 30 in Encyclopddie der Mathematischen Wissenschaften, 4, 654-655, F. Klein and C. Muller (eds.), Leipzig, Teubner (1914). 4. Hu, Hai-Chang, "On some variational principles in the theory of elasticity and the theory of plasticity," Set. Sinica> 4, 33-54 (1955). 5. Reissner, R, "On a variational theorem in elasticity," /. Math. Phy.y 29, 90-95 (1950). 6. Reissner. E., "On variational principles of elasticity," Proc. Symp. Appl. Math., 8, t-6 (1958). 7. Washizu, IC, "The relations between the energy theorems applicable in structural theory/' Phil Mag. J. Set. 7th Ser., 26, 617-635 (1938). 8. Washizu, K., "On the variational principles of elasticity and plasticity," Aeroelastic Research Laboratory, MIT Tech. Rep. 25-18, MIT, Cambridge, MA (1955). 9. Washizu, K., "Variational principles in continuum mechanics," Department of Aeronautical Engineering Report, 62-2, University of Washington, Seattle, WA (1962). 10. Washizu, K., Variational Methods in Elasticity and Plasticity, 3rd edition, Pergamon Press, New York (1982). 11. Gurtin, M. D., "A note on the principle of minimum potential energy for linear anisotropic elastic solids," Q. I Appl. Math, 20, 379-382 (1963). 12. Tonti, E., "On the mathematical structure of a large class of physical theories," Acad. Naz. dei Lincei, Serie III 52, 48-56 (1972). 13. Oden, J. T., and Reddy, J. N., "On dual complementary variational principles in mathematical physics," Int. J. Engng. Sci.y 12,1-29 (1974). 14. Reddy, J. N., "A note on mixed variational principles for initial-value problems," Q. J. Mech. Appl Mech., 28(1), 123-132 (1975). 15. Reddy, J. N., "Modified Gurtin's variational principles in the linear dynamic theory of viscoelasticity," Int. J. Solids Struct., 12, 227-235 (1976). 16. Reddy, J. N., "On complementary variational principles for the linear theory of plates," ./. Struct Mech., 4(4), 417-436 (1976). 17. Reddy, J. NM "On locking-free shear deformablc beam finite elements," Comput, Methods Appl Meek Engng., 149, 113-132 (1997). 18. Reddy, J. N., "On the dynamic behavior of the Timoshenko beam finite elements," Sadhana {Journal of the Indian Academy of Sciences), 24, 175-198 (1999). 19. Reddy, J. N., "On the derivation of the superconvergent Timoshenko beam fi nite element," Int. J. Comput Civ. Struct. Engng., 1(2), 71-84 (2000). 20. Reddy, J. N., and Tsay, C. S., "Mixed rectangular finite elements for plate bending," Proc. Oklahoma Acad. ScLt 47, 144-148 (1977). 21. Reddy, J. R, and Tsay, C. SM "Stability and vibration of thin rectangular plates by simplified mixed finite elements," J. Sound Vibration, 55(2), 289-302 (1977). 22. Herrmann, L, R„ "Finite element bending analysis for plates," Proc. ASCE, J. Engng. Mech. Div., 93(8), 13-26 (1967).
REFERENCES 543 23. Ti moshcnko, S. R , and Woinowsky- Krieger, S., Theory of Plates and Shells, McGraw-Hill, Singapore (1970). 24. Reddy, J. N., Mechanics of Laminated Composite Plates, Theory and Analysis, CRC Press, Boca Raton, FL (1997). 25. Anderson, B. W., "Vibration of triangular cantilever plates," J. Appl. Mech., 18(4), 129- 134(1954). 26. Barton, M. V,, "Vibration of rectangular and skew cantilever plates/' J. Appl. Mech., 18, 129-134(1951). 27. Plunkett, R-, "Natural frequencies of uniform and non-uniform rectangular cantilever plates/* 7. Mech. Engng. Set, 5,146-156 (1963). 28. Carson, W. G, and Newton, R. E., "Plate buckling analysis using a fully compatible finite element/MMA I, 7(3), 527-529 (1969). 29. Timoshenko, S. P., and Gere, J. MM Theory of Elastic Stability, 2nd edition, McGraw-Hill, New York (1961). 30. Reddy, J. N., Theory and Analysis of Elastic Plates, Taylor & Francis, Philadelphia (1999). 31. Putcha, N. S., and Reddy, J. N., "A mixed shear flexible finite clement for the analysis of laminated plates," Comput. Methods Appl. Mech. Engng., 44, 213-227 (1984). 32. Reddy, J. N., and Sandidge, D., "Mixed finite element models for laminated composite plates," J, Engng. Indust, 109, 39^15 (1987). 33. Pian, T. H. H., and Tong, P., "Basis of finite element methods for solid continua,'* Int. J. Numerical Methods Engng., 1, 3-28 (1969). 34. Tong, P., "New displacement hybrid finite element methods for solid continua," Int. J. Numerical Methods Engng., 2, 73-85 (1970). 35. Neai, B. K., Henshell, R, D., and Edwards, G, "Hybrid plate bending elements," J. Sound Vibration, 23(1), 101-112(1972).
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS Chapter 2 2.1 The required equation is given by C • [A - (A • e^e^] = 0 (or any multiple of it). 2.2 A necessary and sufficient condition for the three vectors to be coplanar is that (A - B) x (B - C) ■ (A - C) = 0, which provides an equation for the required plane. 23 We begin with the left side of the equality and arrive at the right side: A x (B x C) = A/e/ x (ejSjkfBjCfc) = AiBjCk£jMepuhp = (Sjphi - 8jiSkP)A;BjCkep = AiBjQej - A,-B/Qejt = (A • C)B - (A ■ B)C. 2.4 This follows from Exercise 2.3. 2.5 (a) 8u = flu + <522 + 533 = 1 + 1 + 1 = 3. (c) FjjSjk — J^i<5i/(+^2^2it + ^3%'Cleariy,thefirsttermiszerounlessA: — 1. If Jc = 1, the remaining two terms are zero, giving Fj\, Similarly, the second term is zero unless k = 2, which makes the first and third terms zero so that the result is Fa. Finally, the third term is zero unless k = 3, which makes the first two terms zero, giving F/3. Thus, FijSjk is equal to F^. (e) Follows from part (d). 2.6 Note that B x C ■ A = A x B < C. 544
ANSWERSySOLUTIONS TO SELECTED PROBLEMS 545 2.8 Let A = (1,1,0,0), B = (0, 1,0,1), C = (0,0, 1, 1) and D - (1,1, 1, 1). Then the relation aA + pB + yC + fjiD = 0 implies a + //, = 0, a + jSH-At = 0, K + M = 0, p + y + n = Q whose solution is a — y = —/u,. Hence, the vectors arc linearly dependent. 2.10 (a) The set is linearly dependent and does not span ))i3. (b) The set is linearly independent. In particular, the vectors &i, £2, and $3 are represented by the vectors A, B, and C as e! = -iA + ±B + C, e2 = ±A + |B - C, e3 = -£A + ^B, Hence the set (A, B, C) spans 9?3. 2.11 Follows as outlined in the problem statement. 2.12 (a) We have &[ = -^(^i **** £2 + %); ^ = n, the unit normal to the plane; finally, the third basis vector in an orthonormal system is related to the other two vectors by §3 = e'j x ef2. Thus, the two coordinate systems are related by (note the matrix of direction cosines) 1 V3 2 >/l4 -4 V42 -1 V3 3 vT4 1 V42 I -1 V3 I VIS 5 V42-I (b) Transformation matrix relating (ey,, &2> 63) to (ei, £2, £3) is given by [A]: V3 j_ 1 V6 i 1 V6 1 -, V3 V3 7! 0 2 V6- (°y = «/•«/)• (c) [A] = /3 2 1 "2 f 2 V3 2 0 0
ANSWERS/SOLUTIONS TO SELECTED PR08LEMS 2.14 Follows from the definition \A] = 0 4. "I V2 I 2 V2 2 _I _L I L 2 ^ 2 2.15 The direction cosines are (2/3, -2/3, -1/3). 2.16 (a) A • B = 6 - 4 = 2. (b) The angle between A and B is 0 = 82.34°. 2.18 (a) We have (b) We have a - " grad(r") = ei—-(xjxj)"'2 = ^-(xyxy)^2^1^, (n-2)/2 _ wrfl-2r> = nejXi(xjXj) (c) We have, in view of the result in (b), VV) = T^-tr") = n{n - 2)r"-3|^r/ + nrn-% dXjdxj dXj = n(n- 2)r"-3—jcf + 3«r""2 r [n(« - 2) + 3n]r"^ = n(n + l)r »«-2 (e) div(r x A) = e, • —(SjkiXjAkei) = £;*/S„ f TT^fc + xj y^ ) = (0 + 0) = 0. (g) div(,-A) = e,- • ^-(rAjej) = e, • ey (^Ay) - 2A/ = Ir • A. 2.19 (a) Since 'd2FjdXidxj is symmetric in i and 7, we obtain V x (VF) V dxij \J'dxjJ Jk kdxidxj
(d) ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 547 grad(FG) = e,£- (FG) = * [ ^G + F^\ — (FG)-*-(—r ■ -°G> = VFG + FVG. (g) First, show that V(A-B) = VA-B+VB-A (1) and A x curl B = VB . A - A • VB. (2) From Eqs. (1) and (2) the required vector identity follows. (h) . 3 . (3Ai 3Bk div (A x B) = ei—(8jkiAjBkei) - £,•/* ^r-Bk + A;—- 0Xj ' ' \axi dxj — curl A • B — curl B • A, a) 3 A ■ (V x A) x A = e/,-*—^-e* x (Apep) oxi _ dAj . _ dAj A OXj OXj = ^lAiej - **lA& = A . VA - VA . A. 9*,- j dxj J 2.20 Follows from the gradient theorem, Eq. (2.54a). 2.21 Note that dr/dxj = xjr and grad(r2) = 2r. Then use the divergence theorem to obtain the result. 2.23 The integral relations are obvious. For example, the identity in (a) is obtained by substituting A = <f> Vij/ for A into Eq. (2.54b). The identity in (b) follows directly from (a) by interchanging <p and \fs and subtracting the resulting identity from the one in (a). Finally, identity (c) follows from (b) by replacing \fr with V V. 2.24 See Exercise 2.12(a) for the base vectors of the barred coordinate system in terms of the unbarred system; the matrix of direction cosines [A] is given there. Then the components of the dyad in the barred coordinate system are [f] = IA][T][A]T.
548 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 2.25 First establish the identity that the determinant of 3 x 3 matrix [A] is given by \A\ = £ijk<*ua2jazk = £eijke™tai>ajsakt. (1) Then write the characteristic polynomial associated with the three-dimensional stress matrix [a] as \\a — Xl\\ = 0, and expand the determinant as in (1). 2.26 (a) Clearly, k\ = 3 is an eigenvalue of the matrix. The remaining two eigenvalues are A2 — 2(1 + V5), A3 = 2(1 - \/5). The eigenvector associated with X{ = 3 is A<'> = ±(0, 0, 1). Similarly, A(2) = ±(-0.851, 0.526, 0) andA(3> = ±(0.526,0.851,0). (c) The eigenvalues are Ai — 4, X2 = 2, A3 = I. The eigenvectors are A(1) = ±-Uo, 1t -1), A(2) = ±4=(0> U 1), V2 V2 A(3) = ±4=<1,0,0). (e) The eigenvalues are Ai = 11.824, A2 = 1.285, A3 = -7.109. The eigenvector associated with X\ is A(1) = ±(0.5239,0.2462, 0.4396). (f) The eigenvalues arc A| = 3.247, X2= 1.555, A3 = 0.198. 2.27 The invarianLs of a matrix [A] in terms of its coefficients #// as well as its eigenvalues are I\ = an = an -ha22 + *33> 1% - -fauajj ~ atjaji), h = \A\, IX = Xi + X2 + A3, h = -± [(Xi + X2 + A3)2 - A2 - A2 -A2] , /3 = X1X2A3. (a) Using the actual matrix coefficients, we obtain /1 =4 + 0 + 3 = 7, h = ~\ [(7)(7) - 2«22 _ 2a23 - 2af3 - a2, - a\2 - a2,] '= -\ [49 - 2 x 16 - (4)2 - (3)2] = 4, /3 =-(-4) [(-4)3-0] = -48.
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 549 (c) Using the actual matrix coefficients, we obtain /, = 1 + 3 + 3 = 7, h = _ i [49 _ 2(-l)2 - (I)2 - (3)2 - <3)2] = -14, /3 = (1)K3)(3)»(-1)(-1)]=8. (e) Using the matrix coefficients, we obtain h = 3 + 1 + 2 = 6, h = ~k [(6)2 - 2(5)2 - 2(8)2 - (3)2 - (I)2 - (2)2] - -78, /3 = 3[(l)(2)-01-5f(5)(2)-0] + 8[0-(8)(l)]=:-108. Using the eigenvalues, we obtain h = 11.824+1.285-7.109 = 6, I2 = -i [(6)2 - (11.824)2 - (1.285)2 - (-7.109)2] = -78, h = (11.824)(1.285)(-7.109) = -108. 2,28 (i) We have tfi = 2(ei + 7e2 + e3). un = 7.33 ksi and us = 12.26 k&i. (iii) The stress vector is given by tA = (7+\/2)ei -(8-hv/2)e2+4(l+v/2)e3. 2.30 (c) We have (0 x A)T = ^Aks/«(eje/)T = <fr7A/c£^/e/ef- = -0Ue/c) x ((pijejei) = A x (<J>)X. Chapter 3 Exercises 3.1-3.14 in this chapter are designed to review general principles of continuum mechanics. Many of them have specific directions to cany out the proofs. Therefore, the answers are already given in the problem statements, The proofs are easier to establish if one uses the index notation in rectangular Cartesian coordinates. 3.1 The result follows from subtracting Eq. (b) from Eq. (a) of the problem statement. 3.5 Let Q — 1 in Eq. (b) of Exercise 3.1 and obtain — [ dV = l v . xi dS = [ div v dV. Ot JR fs Jr By shrinking the volume to an infinitesimal volume rfV,we have — (dV)^div\dV or diVv= lim ~L — (dV). Dt cfv^o dV Dt
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 3.6 Show that grad(u2/2) — v x curl v — v • grad v. 3.7 Follows from the product rule of differentiation. 3.8 This is an interesting (perhaps difficult) exercise. It requires the establishment of the following identities [see Exercises 2.19(b), 2.19(i), and 2.19(j)l: V-(VxA)=0, (1) (V x A) x A = A . VA - VA • A, (2) Vx (AxB) = B- VA-A- VB +AdivB - BdivA. (3) Next, begin with the definition of the material derivative of the velocity vector to establish the required identity, 3*9 Follows directly from Eq. (c) of Exercise 3 J by replacing Q with p\. 3.10 Follows dii'ectly from Eq, (a) of Exercise 3.4 and the divergence theorem. 3.12 Follows directly from Exercises 3.7 and 3,10. 3.14 We have div(a . v) - e,-— * ((JjkZjh ' vmem) d dajk dvk = — fakVk) = Vk-— + ova — ox; oxi ax,- = div a • v-j- a : (Vv) . 3.16 Check the equilibrium equations for ft = ft = ft = 0. (a) Not possible, (b) Not possible, (c) Possible. 3.17 The body force components are ft = 0, ft — 0, and ft — —4. 3.18 The stresses are ^—£(#-4), 2/3 CT22 3.19 The stresses are #oJCl tut 2\ "° («•-—)■ 22 6/3 2/3 2b 3.20 1 (1,1.3) 1 -3 -14 7' 10 . tn = - y psi, ts = 8.576 psi.
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 551 3.21 The eigenvalues are Ai = 6,856, A2 = -10.533, A3 = -3.323. The eigenvectors are A(1) = ±(0,42, 0.0498, -0.905), A{2) = ±(0.257, -0.964,0.066), A(3)= ±(0.870, 0.261, 0,418). 3.22 The eigenvalues (or principal stresses) are Ai = 11.824 (psi), A2 = -7.109 (psi), A3 - 1.285 (psi), The maximum stress is a\ — 11.824 psi and the plane of maximum stress is given by n(1) = ±(0.730,0.337,0.594). 3.23 The linear strains are £ii =2(\-X])x2\ 2^12 = (3c2+C3)^| "2^1; e22 = -2c3(l-X[)x2. 3.25 The displacements are given by u\ — x\ - Xi = — X2, «2 - *2 - %2 = 0, b from which the strains can be computed. 3.26 The displacements are given by it[ =zx[-X\= \j^J X2% 112 = 0, 3.28 Check for compatibility of strains: (a) yes. (b) no. 3.29 Use the strain transformation equations to obtain efn{—efn) = eo/2b, and e'n(=e's) = 0. 3.30 The principal strains computed using the eigenvalue approach arc A] =0 and A2 = 10 x 10~4 (inVin.). The principal directions are A^l) = -^(2, 1) and 0 = 26.57° (clockwise from the x-axis), 3.31 See answer to Exercise 3.29. 3.32 To express the vector form of the equilibrium equations in cylindrical coordinates (r, 0, z), express a in nonion form, and the del operator V, the body force vector f, and the displacement vector u in the
552 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS cylindrical coordinate system. Next, carry out the operations indicated in the equilibrium equation (i.e., take the divergence of the stress tensor, noting the dot product between the basis vectors) and collect the coefficients of various base vectors to obtain the required result. 3.33 The linear strain tensor in vector form is defined by Ks = \ [Vu + (Vu)7] (1) where _ .9L 3±.3 9c, A dee . dr r 30 dz 30 BO Using Eq. (2) and the displacement vector u — urer + wzez, (3) evaluate the expressions needed in the vector definition of the strain tensor, 3.34 A special case of Exercise 3.33. 3*35 Begin with the requirement that B2Ui _ d2ut dxj'dxk dXkdxj Thus, wjk is symmetric in ./ and k> Hence, dhu 3 (3ui\ 3 0 = a^'w = s^aTk [to]) = ejkl^k(eiJ + ^ Differentiate with respect to xm and multiply with £,-„,„ to obtain V = GjkfGimn\€ij,kni i ^ij,ktn) = &jkt€imn€ij,km* which is the component form the required identity. By multiplying with e„ and e/, one can obtain the vector identity- Chapter 4 4.1 V = (mi + nt2)gLi(i - cos#i) + m2gL2(l - cos02)- 4.2 V = (WL/4)Q02 + 56% + 02) + ifAtff + k2(02 - 0{)2 + /c3(03 - 02fl 4.3 The increment of complementary work done is given by A VF* = (« + vO) x AFV + [v - (L + u)0]AFy (assuming small 0).
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 553 4.5 The strain energy and work done by external applied loads are 4.6 The total complementary strain energy is (Ra — $ kN) W* = 74,9 + 0.244/* (N-m), Clearly, the strain energy due to transverse shear is very small compared to that due to bending. 4.8 The complementary strain energy is given by 4.9 The total complementary strain energy of the structure is the sum of the complementary strain energies of the members: «"-£['- (£)'««'« $)'+<$]■ 4.10 The strain energy and complementary strain energy are given by 19 . _ r _. /50\a P3 U = —AEv+fv, -GO 25 \57/ ?>E1A1' 4*11 The admissible virtual displacements are dSwn Swq = 0, =0 at* = 0. dx 4.13 There are several possibilities. For example, if 8P is applied at jc = L (much as P in the figure), then it should be in equilibrium such that 8P - 8RA - 8RC = 0, where 8Ra is the virtual tensile force at the left end and 8RC is the virtual compressive force at the right end. 4.15 One possibility is to apply a vertical force (downward) of 8 P at point D, and 0.55 P (upward) each at points A and B. 4.16 The total virtual work expression for the problem is &W=[ EAC-^^dx+ku0(L)8u0(L)--\ [ f(x)8uodx + PSu0(L) Jo dx dx Uo
554 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 4.18 4.19 The total virtual work is 8 W* = (v - OL cos 0)SFy + (u + 0L sin 6)8 Fx. For small 0, we have v ~0L and u — 0. The total complementary virtual work expression for the problem is 8W" = / —-8N dx - SPbuo(L) + 78FBF,, Jo FA k 4.21 4.23 4.24 where it is assumed that a virtual force 8Pb is applied to the right at x = L. For this problem, we have 8W* = SWf + &W%> where 8W* = 8U* and U* is given in Exercise 4.8* The virtual work done by a virtual force 8P applied at point A is SWg — —wfiSP, where wfi is the vertical deflection at point A in the direction of the virtual load 8P. The total virtual work done is 8W* = L\(Fycos9[ - Fxsm0\)80] + L2{Fy cos 02 ~ F* sin $2)862. Denote the forces and extensions in the three springs by (F\, F2, F3) and (vi, V2, V3), respectively, with similar notation for the virtual forces. Then the total virtual work done is SW*-8F3 (vi - v2)~ + v$-v2\ + 8F(v2 -v) + 8M a L a ~B 4.25 The virtual strain energy is Jo 1 \ dx dx clx J d2S U>0 " dx* dx. 4.27 The linear strain-displacement relations in cylindrical coordinates, for the axisymmetric case, are (see Section 8.2.1 for complete developments): 1 fdur Br?. = 3 r £77 — duz dz * dr )r ) eFe = 0, Gtf = 0. (c) 4.29 4.31 5/ Ja Vi + ("')2 8u' + Jl + («02 to dx. 81= f (u"Su" + Su"u' + u"Su' + u'Su' - Su)dx. Ja
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 555 4.33 81 — I (UX8UX 4" SUjcVx + UX$VX -h dllyVy -h UySVy + VySVy + f8u + g8v)dxdy. 4.35 SI — f [/jl (SiijjUij + UjjSujj) — PSuij — 5P«r\/]jA*irf^2 ./ft — / fiSu{dx]dx2— I U&Uids. 4.37 The total time taken by a particle in going from point A to point B is given by T = -^= f J1 ' *"'—' dx ~= /(«). Jo V w J_ /"* ll + (du/dx)2 4,38 The problem consists of finding the curve of minimum length joining the two given points A: (a, ya) and B: (6, y^). Mathematically, this amounts to minimizing the integral '-jCMs)*"'00- 4.40 4,41 The Euler equations are w"" - u" - 1 = 0, ina<j<i, w'-w"' = 0, w" + w'=:0, atx = a,fc. The Euler equations are lx' = 0, 0<x<L, d I dw dx I dx 2 V dx / dw 1 /dw\2~\] d2 ( d2w\ dx 2\dxJ J J rfjc2 V dx2/ Ev4 !w 1 /du^ dw d ,v=L du dx + 1 /d»;\2] d (EId2u)\\ 2\dx) J dx V dx2)\ -P=0, ,v=Z. -Mo=0.
556 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 4.42 The Euler equations are -UXX ~ 1>JC.V - Vyy + / := 0, -UXX - Uyy ~ Vyy + g = 0 1R ("A" + Vjf)^ + Vytly = 0, «.v«,v + (tiy + Vy)tly — 0 On 4*43 The Euler equations are V- (-liUJ " ";,0") ~ PJ ~ fi = ° in n» ja («/,j' -f My\/) ny - Pti/ — tj — 0 on ,$2. 4,44 The Euler equation and the natural boundary conditions are given by dsw n rs2w d2w\ n , —-: ^l-r-y + ^TTT ^ ° at* =0, a for any y, ox \ dxA oyA J 3Sw „ / 32v) \ —: (1 ~v)D[-jf^-7 ) =° at * =0, a for any y, a2i aSw _ /a2u; aiy — Q,b for any x, 95 w 17"' (1 - v)£> ( ^H = 0 at y = 0, ft for any x. 4.45 The Euler equation and the natural boundary conditions are d2 ( d2w0\ d ( ldw0\ . . / d2WQ ^ dwo\ I D)|r—y + Dyi-— 1=0 at ;• = r,- and r = ro, 4.46 The Euler equations are at r = fj. d,v a? a2£v a2<^, a /a0A- a0y\ a2^, a2&. a /a^ 80v\
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 557 (4) ay and the natural boundary conditions are a wo , n „, &WQ M*: —-+£r=0, Sky: —ii + 0 =0t Ox 9y 84>t 50_ V 9* ty ) ■v: D(\ - v) -r1 + "X2- 1= 0 at x - 0, a tor any y, ' \ 9)' a* / <5d>v: d[^- + v— )=0 at j> = 0, Mor any x, \ d)> 9x ) rv: D(l -v)( -~ + -P- ) =0 at y = 0, Mor any *. fife: 4.47 The Euler equations are (5) (6) (7) (8) ** ^2fe ^ 92fe 'd {3<t>xB4>y\ fdw0^.\ n ^ a>>2 a*ay dx V a^ a* / V 3)' / 4.48 (a) The Euler equation is 1 + Y dx ):fM— dit dx /duY \dx~) = 0. (1) (b) The Euler equations are Su: '4 du dx 'i+ m\ = 0, 8k: cFg 1 afjc - L = 0. (2) (3)
558 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS Clearly, a comparison of Eq, (2) with (1) shows that X = y 4.50 The Euler equations are lh(f, J dx — L 8exx: EIkxx -h A = 0; 5X: kxx d2WQ ~dx^ — 0; Swq: (4) d2X Chapter 5 5.1 The Euler equations are -i(EAj?)-m=0> 0<X<L' duo' EA- dx + ku0(L) - P = 0. 5 A The Euler equations are dN , dQ n dM „ „ dx dx dx The boundary expressions indicate that «o» wo> and <p are the primary variables and jV, M, and Q are the secondary} variables of the problem. Thus the natural boundary conditions involve specifying JV, M, and Q. 5.5 v = (50P/51AE)2. 5,6 The (tensile) forces in the wires are 3M0 Pb = ML' Pc = 4Mq Til' 5.7 The slope atx — L and reactions are 1 /^2 \ <y0L2 _ 3 (qoL2 \ 2 " 5.8 v = (2595/32)(P2/A2E2), where L, = 60 in., L3 = 48 in., and L4 = 36 in. (with the understanding that A and K = £ are in sq. in. and psi, respectively). 5.9 wc = (5«>i*/384E/).
5J0 5.11 ANSWERS/SOLUTIONS TQ SELECTED PROBLEMS 559 00 = -{MoL/AEI) + {qQL3/4ZEI). w< _ u^ ^ _ /25FqL3+^0L4\ 2 ~ \ 768£7 )' 5.12 5.13 Note that this is one of many relations between wc and wc that can be obtained depending on the choice of the virtual forces. It can be verified that the above relationship is exact. wu = (5F0a3/6EI). The force Fc in the cable is -i FC = ErA c , Lb 2 , Lfr . 2 -h ^ cos a + -„ . sin a £/>/U 3EbIh Ml sin a F0. 5.14 5.15 5.16 The slope atB is 0B = (170.67/EI) clockwise. The structure has bending and torsional deformations (see Example 5.14): 1 WA = Qa3 (P + Q)by + ±iQa*b). EI The displacements are 6EI 1 Phb 5.17 5.18 5J9 u^6EI^Pva2b + 2Ph^ + ^i' We have u = 0 and v = (2595/32)(P2/A2E2), where L* = 60 in., L3 = 48 in., and L4 = 36 in. (with the understanding that A and £ are in sq. in. and psi, respectively). (Pc=4Mo/ni). w(L) = \\ + (kL3/3EI)) 30£7 \\ + (kL3/3EI) As a special case, the deflection at the free end of the beam when not supported by a spring is obtained by setting k = 0. 5.20 The displacements are w = PR3K 4EI ' u — PR3 2EI
560 ANSWERS/SOLUTJONS TO SELECTED PROBLEMS 5.21 The displacements are FR^tt 2FR3 s 2EI EI 5.22 The displacements are 2pR* 3jtpR4 u = —eT> v = —2FT' 5.23 The displacement is FRn FR3n fsFRn tin = - H H ■ -. 2EA 2EI 2GA 5.24 The deflection (up) and the slope in clockwise direction at point A are quL4 { ktfy1 _ qpL3 [3 - (kL*/SEI)^ Wa-~JeT\] + JeI) ' °A- 6£7L3 + (^3/^oJ 5.25 The deflection is 1 /F0L3 q0L4\ t . u L , WA ~ ~Fi \ ~Y~+ ~1T~ I (wlthout shcar)- 1 /F0l3 qoL*\ fs ( qQL2\ 5.26 v>c = w0(Q = (dU*/dF0) = (F0b2L/3El). 5.27 (BA = F0ab/6EI). 5.28 The vertical and horizontal displacements are given by 5.29 The reactions are Dy - QA4P and Cy = 0.56?. 5.31 The work done by Fq in moving through the displacement wba is FqMqL2 Wba=F0wba = -££;-. The work done by Mq in moving through the rotation Gab is WAb = Mo6ab = MqFqL2 16E1 Thus, Wba = Wab-
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 561 5.33 The deflection at the midspan is °\2J \48EI ^ 3Z4EI ) 5.34 From Maxwell's theorem, we have Fb • wba + Fc • wCA = FA • wAb + Fa • ™ac = FA • wAy wA = wAB + wAc< Fa = 4,0001b, FB = 4,500 lb, Fc - 2,000lb. Chapter 6 61 The Lagrangian L of the system is L = -U = --[ktxf + (*2 + k3)(x2 - xi)2 + h(xi - xi)2\ 63 The rate of the complementary Lagrangian function is ( F F \ 8L* = -i{8F, -x2SFd-±x8F = (-- +x ]8F. \ *i n ) 6.4 The Euler-Lagrange equations arc given by 501: —W2//1^20102sin(0i - 02) — (mi + m2)gL\ sin0i - j[(mi +m2)/?0i +m2LiL2cos(0i - 02)02] = 0, S02: m2Z,]Z,20i02 sin(0i - 02) - migL% sin 02 - m2^\\\B2 + LiL2cos(0[ - 02)0i] = 0. 6.5 The total kinetic energy of the rigid bar assemblage is K = ^^(20? + 402 + 02 + 90!02 + 30203 + 30i03). 6.6 The Lagrangian function is L =^mi [/202 -hi2 - 2x19 sin0] + \m2x2 H- mig(x - / cos0) + m2gx + ^/c(a' + A)2, where ft is the elongation in the spring due to the masses: h = -{w\ +m2). k
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 6.7 The Lagrangian is given by L = \m\it\ + )pm(u2 + ii\- V2u\U2) + 4jW2£"2* 6.9 The equation of motion is /m[-m2\ 6.10 The equation of motion is (m\ - mz — m-A Xl = —;—~r~^ )8' \m\ +1112 + "13/ 6J1 i'i=*/23. 6.12 The equations of motion arc m(x +10 cos0 - W2 sin0) + Lv = F, m[lx cos 0 + (I2 + £22)£] + m#/ sin 0 = 2aFcos9. 6.13 The natural boundary conditions become p/ „ 0 + m—« —- EI—-?- = 0, at* — L, OA-ar2 a/2 ax \ a*2 / a2^0 , a3w0 n . r EI—-T- + J „ 0 =Q> at„r —L. ax2 a* a/2 6.14 The Euler-Lagrange equations are Suq: f + dx 3 / 3w0\ n The boundary expressions indicate that uq , w;o, and <£ are the primary variables and yv^-, Afvjr, and gA- are the secondary variables of the problem. Thus the natural boundary conditions involve specifying Mxx and Qx. 6.15 The stress resultants are Mxx = EI^-, ax Qx = KsGA(j£ + <ty
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 563 6.17 The equation of motion is a2 / _a2t 3x St ( rdw\ 9 ( JW\ d2Mx< d2 / d3w \ dw r; ICS ^ ,, — C-~ h q = 0. 6.18 The Euler-Lagrange equations are given by Suq: - pAiio + —^ = 0, (a) ox d2w0 B ( T dw0\ d2Mxx SW0: -PAu« + pI-^ + -(N„—) + -^ + q = 0. (b) 6.19 The Euler-Lagrange equations are B2u 3NX , n ax S^: ~C2pI{Ci3X^+C2w)+C2^r-C2Qx=0- 6.20 The Euler-Lagrange equations are x aNXi a ( a«0\ _ ^: - -~ + Or + ^7 6.21 The equation of motion is a / 3«\ a / du\ a ( du\ 6.22 a,! = (9.8766/Z,2)V£///oA. 6.23 wi = (15.45 /L2)jET/pA. 6.24 o>, = (22A5/L2)*S(EI/pA).
564 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS Chapter 7 7.1 (a) A subspace. (b) Not a subspace. (c) A subspace. (d) A subspace. 7.2 (a) Linearly dependent, (b) Linearly independent, (c) Linearly independent 7.3 (a) Linearly independent, (b) Linearly independent, (c) Linearly dependent; same as Exercise 7.2(c). 7.4 (a) ||u||o = y l - £, sup-norm = 1. (c) II"Ho = 7i-£, sup-norm = 0.207. (e) ||u||o = \/50, sup-norm = 100. 7.6 (a) Qualifies as an inner product, (b) Qualifies as an inner product (note that it is symmetric and positive-definite). 7.7 (a)(«,i;)o = 47r31 (utv){ = £(i + ^). (c) (m, v)o = 0, (k, u)i=0. (e) («, v)0 - £( - l + £ + £ + £), (w, vh - (ii, u)0 + $. 7.8 ||«-i;||§ = (33/35). 7»9 The functions are not orthogonal 7,10 The functions are not orthogonal. Zll a = 1037/270 and b = 362/135. 7.12 (a) C = 1/3. (b) C = (-7 ± V50)/2. 7.14 (a) LineaT. (b) Nonlinear, (c) Linear. 7.15 (a) (7i + T2)(x) = U2 - *3»3*i - x2 - *3, *i + x2 + *3>. (c) (727,)(x) = (*, -x2 -2x3,0:3 -xu0). 7.16 The transformation matrix is 12 0 0 0 2 10 10 10 7J7 (a) Bilinear; it is a linear functional in v for fixed it. (b) Linear functional in v for fixed u. 7.18 (a) A bilinear form, (b) A bilinear form, 7,20 (a) 81 is linear in it because / (u) is quadratic in w. The first variation of any functional is always linear in Sir, thus, 81 is linear in both u and 8u. (b) 81 is always linear in Su. Since /(«) is a functional, its first variation is also functional, linear in Su.
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 565 7.21 0o — (h/L)x, 0i = x(L - x), and 02 = x2(L — x). The trigonometric functions are , . nx , Jinx 0o = h stn —, 0„ = sin ——. 7,23 0i = x(L - x) and 02 00 = x\\L - x) or x{\l? - x2). 7,25 See Exercise 7.23. 7*27 0i = (x —a)(y — fo), which violates the biaxial symmetry, 02 = (x2 — a2)x (y2-b2i 7.28 The solution becomes 7.29 The solution is W2(x) = (q0L4/24EI)(5(x2/L2) - 2(jc3/L3)). 7.30 The Ritz solution is r4 / v -2 „3' W/ ^°L (AX X AX \ 7.31 The Ritz solution is ™ FoL3 f.x2 x3 x4\ W2 = 64£7(7F-127J + 5FJ' 7.32 The Ritz solution is gpL4/, , l2SEI\fx x2\ q0L4 / 30EA/3* x3\ 7.33 See Example 7.16. 7.34 The one-parameter solution is given by The two-parameter solution is given by The one-parameter solution with a trigonometric function is 32/0 7i x ity U] = -r- COS —- COS . 7T4 2 2
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 7.36 (a) Let 4>\ = x(L — x), 02 ~ *2(L — x). The solution becomes 12A V 420/ /,2LI2 u A = 12(EI)2L4 + — -f —ElkL*. * 25200 105 7.38 The critical buckling load is Ncr = 9.87(E //L2). 7.39 The critical buckling load is ,2 Ncr~ L2 V (n^D/L^ + s)1 7.40 For a beam simply supported at both ends and subjected to a suddenly applied unifomily distributed load q = qQH(t), the solution is wo(x, 0 = -^ T — (1 - cos V) sin —- (7) n ~ nKn L n—\ for n odd. 7.41 The two-parameter Ritz solution is ui{x, t) — C[(t)x +C2(t)x2t where cx(t) = 4 C2(t) = — ^ (1 - cos 00 + % (1 - cos y0 J , ]• A2 „ _ , B2 3 IP2 yl where t*2 = 30, ^ = 24, £2 = 4.155, y2 = 38.512, y2 - £2 y2 - £2 yl - fiz Yl - ft* 7A3 The one-parameter solution is ™-354£W£)2-
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 567 7*44 (a) For one-parameter approximation w(r) «* U\(r) — c\cos{7xr/2a)y one has X = 5.832/a 2. (b) For two-parameter approximation U2O4) = c\ cos(7rr/2a) + C2Cos(37rr/ 2a), the smaller root is X - 5192/a2. IAS The solution is /71 7 \ 7.46 The solution is W\(x) = 0.8111(1 -x)(2-x) -> 1/i(jc)== 0.8111(1-*)(2-*)+*- 7.47 This is a special case of Exercise 7.34. 7.49 For the choice U2(x) = ci(2x — x2) + C2@x — x3), one obtains 516-126* 120-1307T C\ = , CO — 433* 433tt 7.50 (b) For two-parameter Galerldn approximation with 01 = (x - L)2{x2 + 2Lx + 3L2), <f>2(x) = (x - L)3(3x2 + 4Lx + 3L2), one obtains c\ = 0.01374 and c2 = 0.00228. 7.51 Approximation of the type U2 — ci0i + C202 + <£o with 00 = (1 — )') sin 71x, 0i = sin nx sin 7T)\ 02 = sin 2tt.v sin 2jtj. yields c.\ — —jt/8 and C2 = tt/24. 7.52 (a) The solution is 42 0 4 (b) For^i — jc, we obtain two values of c\. The choice c\ = V2 —V3 yields the smaller value of £ = J0 7£dx, 7.53 The eigenvalues are X \ = 4.212, X2 = 34.188. 7.57 The solution is given by U2(x) = 0.1763(2* - x2) - 0.15(3* - *3). 7.58 Let u = a/2 + (1 — x2)c\, and obtain £/i = \/2 - 0.326 + 0.326jc2.
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 7.59 The eigenvalues are X\ = 4.212 and X2 - 34.188. 7*60 See Exercise 7.54. 7.61 Let W^(x) = c\x2 -J- c2x3 4- C3X4, and obtain C3 — 0 and foL2 f0L c\ - —— and c2 = -T7—* 4an 36c/o 7.62 The solution is U\ = (-2V2 + V6)(l - x2) + V2 =: 1.0355 + 0.3789a:2. 7.63 See Exercises 7.53 and 7.59. The eigenvalues are Ai — 3.785andX2 — 19,753. 7.64 The one-parameter approximation t/i — C[ \ sin nx sin ny yields c\ \ = (fo/2n2), which is about half the value of the first term in Exercise 7.54 (c„ =8/b/*4). 7.65 The solution is U\ (x> y) = e~^0y (x - r2). 7.67 The solution is given by W\ (jc, t) — ci(Q) cos at (1 - cos2jtjc) . 7*68 For solution of the form U\ = c\ (x)(y2 — b2), one obtains 7T2 10 *l =7^? + (2a)2 (2b)2' whereas the exact value is X = 2 2 n 7r (2a)2 (2b)2' 7.69 For the two-parameter approximation of the form U2 = [ci(x) + c2(x)y2](y2-b\ one obtains n2 9.871 Al ~ m2 + (2/7)2 • 7.70 A two-parameter approximation of the form (which is more complete than the one-parameter approximation, although it has no effect on the derivative of (he solution) *2 = ci + c2(x*-6x2y2 + y4)
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 569 yields 53 /• 2 Cl=180/oa' --J-k C2~ 144 a2' Chapter 8 8.1 The Euler equation and the natural boundary conditions at the inner edge are given by Id2/ d2w0\ i d {Dndwo\ 7-d^{D^^-)---Tr\—^)-q=Q d ( r, d2w0\ , DndwQ rr{rD"-d?r) + d "dr „ d2wo ^ = 0> = 0, b < r < a, at r = by at r - b. r dr dw0 dr2 + °22^ Note that the terms involving D\2 cancel out. 8.3 The Euler equation is n a / a\ l a2 in a / 3w0\ i a2u;o"| , , 8.4 The deflection due to the applied edge moment Ma at r = a is W0(r) = V,Vi f sr> ( ' 2 ) • 2(1 + v)D \ a1} 8.5 The solution is given by wo M,, 64D [(l + v)D + p a l(l + v)D + Pa\\a) W J" 16 ' , (3 + v)D + pa Mfl6 = 16 d + f) (l+p)D + /3a (1+3 O2] 8.6 This is a special case of the loading shown in Fig. 830 (set qi — 0). For a lineaily distributed load of the type *<D = *>(!-£)
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS one obtains d - ---?- (F'\a) + V-F'(a)) = ---- /ZI±___\ , 2 (1 + v) V w ^ a K }) (1 + v) V 360 ; ' c4 - - [Fia) + -c2j - -^ (^^~j • 8.7 This is a special case of the loading shown in Fig. 8.30 (set qo — 0). One obtains _ 2q\a2 /4+iA _ q\aA (& + v\ ;2" "(i + v) \~45~)' C4" o + v) V^so"/' <?2 8.8 c2 - ~(\2q0a2/516) and c4 = {2q0a4/576). S3 c2 = -(29c/0fl2/360) and c4 = (I29qaa4/14400). 8,10 C2 = -(2gi«2/45) and c4 = (3q{a4/450). 8*11 The solution is .4 / a < 8,12 The solution is ™~„0-S)- Clearly, the solution is no? acceptable. 8>13 Assume solution of the form (for n = 0) r2 Wi(r) = c^ifr, 0) = ci/Kr), /Kr) - 1 - --, 8.14 and obtain See Example 8,7, CO = 4.899 : a2 ]j ph , The fundamental frequency is M-- The eigenvector is ph<o2 = D 10439 ftW = (,- £)' + 0.325 (l-£)\
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 571 8.15 The center deflection of a simply supported plate under asymmetric load is .2 wc = %a SD 8.16 The deflection is wc 8.17 The deflection is /3j-_u\ a2 _ a2! _ qoa*_ /5 + iA \l + v/ 4 8 J "" (AD \l+v)' _ g0a4 /5 + v 6 + v\ = (l + v)D V"~64 150") ' 43 ?0a4 ' 4800 D The solution coincides with the solution at r = 0 given in Exercise 8.9. 8.20 The solution is given by c\ = C2 = 7ta2k Go „ , l/32Z>(l + iQ , 1^ -1 7Ttf4/< 8>22 The boundary conditions are at j? = 0: u;o ~ 0, Swq — 0; at y — b: wq = 0, B2w0 By ~' ""•/ "'*" ~w "' By2 = 0. In view of the boundary conditions, the constants A,n, Bmi Cm, and Dm are found to be An7 — 4ff0a4 /727T **w — ^//ii (--m — An = 2<7o« 2 cosh /?„, — 2 cosh fim — fim sinh #„ m5n5D cosh /}„, sinh /8,„ — fim Am 2(^) sinh fl„ cosh fl„ - (^)sinh fim - (^)2frcoshfl,: 2 cosher sinh pm -/3m where /?in — (mitb/a). In the-case of a square plate the maximum deflection and maximum bending moment are found to be wo(a/2t b/2) « 0,0028 MVA-(a/2,0) ^0.08(?0«2-
572 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 8*24 This is a CCCS plate. Hence we choose 8.26 This is a CCCF plate. Hence, we choose 8.28 The choice Wi(x, y) = en 1 — cos sin coO =cu f 1 -cos j b gives 8?0fl4 en = [I6D11 + 8 (Dl2 + 2Z)66) j2/*2 + 3D22*4rt4] nx5' where.? = a/fc. 8.30 We obtain _ c/pa4 __ a C" " 4tt4[3Du +2(Z)12 + 2D66).92 + 3D2254] ' * ~ 6" For isotropic case, it reduces to qoa4 a Cil = 4tt4D(3 + 2s2 + 3*4) ' S==b' 8.31 The only difference belwecn this exercise and Exercise 8.30 is in the load. One obtains = Qoa4 , _ £ CU ^7r4[3Du+2(Di2 + 2D66)^ + 3D2254] ' ' " b' 8.32 One obtains en = 0?oa4)/(27.984jr4Z>o)anduW' = (qoa4)/(6.9967i4D0) = 0.00l41qoa4/D0. 8.33 One has (A0 = a2/*/3) c\ = Qoa2/(36<>/3D). 8.34 Evaluate F| = <f> Mo—- ds, — = n* — + nyr-i
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 573 and the direction cosines nx and ny on BC, CA, and AB are Icnown from n = -eA- 2ev i- 2 e3, 2eA- 2 e3, on BC on CA on AB. Thus 373 on AB, BC, and CA. Hence, d\ = M0a2/4D (A0 = a2/^D). 8.35 Then Galerkin's method yields = 00 24[5(Di{/a4) + 2{Dl2/a2b2) + (D22/b4)]' For the isotropic case, the solution is (s = a/b) , , qoa4(x/a) ( x2 y2\2 m(X>y) = 24D(s* + 2s2 + 5){l~^-l?) • 8*36 For the isotropic case, the one-parameter Ritz solution is given by (s = a/b) »o(xj) = V a2 b2) 8D(l + 2vs2 + s4) 8.40 The resulting equation is d2w<) 9f2 2 Chapter 9 9.1 The FE solution is t/2 = i, t/3 = i, P,(1) = 2.5t/,-2.5£/2 = §. The exact solution is Ho(x) = J - y1,"*"?. no(0.5) = 0.7605, ho(I.O) = 0.5906. 1 + log 2
574 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 9.3 The solution is t/2=0.2xl0-3in., f/3 = -0.3333 xl0"3in., U4 = -0.8667 x 10-3 in. The forces in each member can be computed from the element equations: p([) r= 3,0001b, P2(2) = -2,0001b, />2(2) = -2,0001b. 9.4 The compression is U2 = -0.002996 a* 0.003 in. The clement forces PJb) and pP are />f} = -551.59 lb and P2('° = -778.41 lb. The stresses in steel and brass are ax = 22.47 ksi (compressive), Ob = 11.24ksi (compressive). 9.5 The displacements are U2 = 0.4444 in. and U3 — 1.9444 in. 9.6 The aluminum member is elongated and the steel member is compressed by the amount U2 — 0.0134 in. The reaction forces and stresses are />2(1) = -21,052.6 lb, a(1) = 6701.25 psi, P2(2) = -78,947 lb, cr(2) = -44,675.1 psi. 9.7 The displacement is U% = 0.3 mm. The forces and stresses in steel and aluminum pipes are P2(1) = 36.364 kN, cs = 606.06 MPa; p™ = -63.636 kN, aa = -106.06 MPa. 9.8 First we must find the force acting at point B to be P = 50001b downward. The solution is U2 = uB = 0.01167 in., U3 = «c = 0.01967 in. 9.9 The solution is U2 = uB= 0.01163 in., I/3 = uc = 0.01956 in. Note that if we had taken k — 105 lb/in., we would have obtained U2 = ub = 0.00957 in., f/3 = uc = 0.01255 in. Thus one may use the value of k to simulate elastic restraint at point C 9.10 Exploit the symmetry about the middle of the beam and use two beam elements to analyze the problem. The generalized displacements are £/2 = -0.00312, <73 = 0.01077in., tf4 = -0.001833, U5 =0.01673 in.
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 575 The reaction force at the left support and the internal bending moment at the center are q[{) = -1200 lb (up), Qf] = 8400 lb-in. (counterclockwise). 9.11 The solution is 03 = -0.03529 cm, 04 = 0.01438, 06 = 0.06376. 9.12 The solution is 02 = -0.001792, C/4 = 0,000512, 05 = 0.03686 in., 06 = -0.001408. 9.13 One-element mesh can be used. The solution is given by [)L4 „ qoL3 Ui = qui u4. 6EI 1 + ('♦£) 9,14 This problem can be modeled with four elements with h\ — A2 = A3 = A4 = 5'. The main objective here is to represent the applied loads appropriately. The global node 2 will have a downward load of 1,000 lb and bending moment of 1,000 ft-lb (CCW). The total size of the assembled global stiffness matrix is 10 x 10. The solution is given by 02 = 0.12187, 03 = -0.45660ft, 04 = 0.03021, 05 =-0,32986 ft, 06 = -0,06979, 08=0.001042, 09 = -0.39583 ft, 0IO =0.10521, where 0/ = Uj(EI x 10~5). The bending moment at a: — 2.5 ft, for example, is given by A-=2.5 d2<fif ;=i + U3 dx2 d24>± Tidl4>\ x=2.5h 2^T dx2 + 04 dx2 = 1833.33 Ib-ft. *=2.5 The shear force at x = 2.5 ft is given by Vc = 733.33 lb. 9.15 The primary objccLive of this problem is to compute the force vector for element 1. One obtains (qo = 500 and h — 5) {q ' 60 ( 9 1 \-2h 1 21 1 [ 3/i ' 375.00] -416.67 | 875.00 1 625.00 J
576 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS The solution is given by U2 = -0.24826, t/3 = 0.99537, U4 = -0.1111 i, U5 = 0.98380, t/6 = 0.11806, t/8 = 0.23611, where {/,• = £/;(£/ x 10"5). 9.16 The components of the force vector due to the distributed load are <*».- Q. The solution is 26 1-4*1 14 [ 3fti t/3 = —0.94151, £4 = 0.19413, . = ■ r 3,466.7 -4,266.7 1,866.7 3,200.0 U5 = = -2.2503 where £/,- ~ U; (EI x 10 6). The bending moment and shear force at x = 3 ft, for example, are Mc = -28,133ft-lb and Vc = 3,866.71b. The values at x = 0 are M(0) = -39,733 ft-lb and V(0) = 3,866.71b. 9.17 The global stiffness matrix is 6 x 6, which can be expressed as EI u 24 + EI 0 -12 -6L k\I? ' EI 0 0 8L2 6L 2L2 0 0 12 + -12 6L /c2L3 6L 0 0 EI -6L 2L2 6L 4L2 0 0 hi? EI 0 12 + hL3 6L EI 0 0 0 0 6L 42 9.18 The finite element model is [Ke]{Ae}-X[Ge]{AG} = {Qe}, where {A6"} and {Qe} are the columns of nodal generalized displacement and force degrees of freedom of the Euler-Bemoulli beam element. The matrices \Ke] and [Ge], known as the stiffness and stability matrices, are defined by pA—^-^dx, ax ax (5) and <j>? are the Hermite cubic interpolation functions.
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 577 9.19 The numerical form of the stiffness matrices [Ke] is given in Eq. (9.61). The expression for the stability matrix [Ge] for the Hermite cubic interpolation is given by [Ge] = 30/2 36 -3k -36 -3h -3/i Ah2 3h -h2 -36 3h 36 3h -3A -h2 3h Ah2 9.20 The finite element model of the equation is of the form [Ke]{Ae)~Ncr[GettAe) = {Qe}. . Using one element in half beam, the condensed form of the above equation for the case at hand is /2£7j~2ft2 3/i V h3 L3/* 6J . 1 ~Ncr3Qh 'Ah2 3ft' 3/? 36 DISH!} Then, Ncr is given by the smallest eigenvalue JVcr = 9.9439(£7/L2). 9.21 The finite element model is given by where rn7 [Ke][ue] = {fe} + {Qe} " d2<p,- d24>j I (dfr d2<j>j d2<pi d<t>j /■''* qf = / rq<t>i Jr„ +D22 dr. 1 d<pi d<f>j~\ T2 dr dr J r dr, Set D\\ = D21 = D and Dn — vD for the isotropic case. 9.22 The finite element model is given by substituting the approximation m n and V{ = yjfj and i>2 = <f>i into the weak forms. We obtain \[K"] [^12]1|W]_|{F1}}
578 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS where <8'£'**%%*• *4'-jf>£**-'* + (I~+rA55\fii,j P1' Ffmf rffi dr + in{ra)Qx + ^(r6)j23, Fil = 4>i(ra)Q2 + 4>i(rh)Q4. Note that both yf/} and fa are the Lagrange interpolation functions. 9.23 Let wq(x) ^ c\ + c^x + c^x2 + C4X3 -J- C5JC4 + ^6-t5, where if is the local coordinate with the origin at node 1. Evaluating wq, 0 == dwo/dx, and k =s d2Wo/dx2 at nodes 1 and 2 (i.e., at a; = 0 and x — /?), we obtain wi 01 K\ W2 <h\ «2 \ ' "10 0 0 1 0 0 0 2 1 ft ft2 0 1 2ft 0 0 2 0 0 0 ft3 3ft2 6ft 0 0 0 ft4 4ft3 12ft2 0 0 0 ft5 5ft4 20ft3 [ci c2 C3 CA C5 C6 Invert the equations to solve for c\ and substitute into wo(x) ^ C[ + <?2* + C3X2 + C4X3 4- C5X4 + C6^5 to obtain the result. 9.24 The stiffness matrix is [JH = £7 70ft3 " 1200 600ft 30ft2 -1200 600ft -30ft2 600ft 384ft2 22ft3 -600ft 216ft2 -8ft3 30ft2 22ft3 6ft4 -30ft2 8ft3 ft4 -1200 -600ft -30ft2 1200 -600ft 30ft2 600ft 216ft2 8ft3 -600ft 384ft2 -22ft3 -30ft2" -8ft3 ft4 30ft2 -22ft3 6ft4 9.25 The finite element model is £^>5 = /,' + Gf, j=i
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 579 where *H,t dx dx ox dy dy dx dfi dirt \ dy dy ) ft= [ ftf dxdy, ef = <£ Vi^ds. Chapter 10 10.1 The Euler equations are Sy: v + grad u = 0, k Su: ~ (V • v + f) = 0, in V, 8u: n • v — q = 0 onS2, 8\: u — h = 0, on Si. 10.2 The Euler equations are 50: — (e!^-\ + A. = 0, 0 < * < L, Sw: - — + / = 0, 0<;t<L, SA.: 0 +—=0, 0<.x<L, J* 50: £7 — =0 atx = 0,L, d* * Su;: X(0) = 0 and k{L) - Vb = 0. 10.3 The natural boundary conditions are (since Suj is arbitrary on S = Si + S2) f^(uij + u-jyOnj ~ Pni — 0 on Si, /x(w,,j 4- ttjj)nj — Pnt — tj = 0 on S2. 10.4 See answer to Exercise 4.46.
580 ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 10.5 The Euier equations of this functional arc 8e\\: EAbxx + X = 0, in 0 < x < L> du SX: sxx =0, in 0 < x < L, ax dk Su: — = /, in 0 < x < L, dx Su: - X(L) - P = 0, Su: «(0) — «o — 0. 10.7 Follows from the discussion presented in Section 102.2, where Il(}ii9Sij,T)= J \pV(eu,T)-fiiii)dV- f fads. Jv Js2 Note that U0 in Eqs. (10.28) is replaced by p*. 10.9 ThefunctionalinEq. (10«28)shouldbemodifiedtoincludetheheatconduction equations and Uo should be replaced by p^. The functional is given by f fl ST BT I n,iW(uitEfj,<Tij, T) = Tlw(iti>Sij,0ij, T)+ I -kij — — -QT UIV, where dissipation terms are neglected. 10.10 The Eulcr equations are d Swq 84> KsGnh dr K*+S)] qr = 0, : KsGi3hU + ^y-£{rMr) + M$=0, d<p SMr: -I- + D21Mr - Di2Me = 0, dr SMB: 4> + r(DnM9 - D12Mr) = 0. 10.11 The matrix coeffici ents are
ANSWERS/SOLUTIONS TO SELECTED PROBLEMS 581 fi = / q^i rdr. 10.12 The coefficients K?f(a, 0 = 1,2) are defined in Exercise 10,11; additional coefficients are defined as follows: K}j = f Diirinti dr9 K% = - P Di21n1rjr<tr, K# = fb D,i^; r dr} Kff = f * As^fj r dr,
Addition: parallelogram law of, 10 of vectors, 207 Adjoint, 260 Alternating (permutation) symbol, 17 Angular momentum, conservation of, 53-54 Anisotropic bodies, 62 Approximation, 4 Approximation functions, 225-226 Archimedes, 2 Aristotle, 2 Banach spaces, 212 Bars, finite element analysis of, 434-454 approximation over an clement, 436-^441 assembly (connectivity) of elements, 444-447 boundary conditions, imposition of, 448 domain representation by finite elements, 435-436 finite element equations, 442-444 governing equation, 434-435 reactions/derivatives of solution, calculation of, 449-454 weak form of governing equation, 441-442 Basis: dual, 21 orlhonormal, 16 reciprocal, 21 Bending; circular plates: analytical solutions, 312-317 variational solutions, 320-329,421^24 in classical plate theory: circular plates, 312-317,320-329 fully discretized finite element models, 475 Levy solutions of rectangular plates, 359-362 Navier solutions of rectangular plates, 344-350 and shear deformation theory, 412-421 variational solutions. 362-379 rectangular plates: Levy solulions, 359-362 Navier solutions, 344-350 Index shear deformation plate theory, 407-410 variational solutions, 421^-24 Bernoulli, Jacques, 3 Bernoulli, Jean, 2 Bcsscl differential equation of order n, 318 Betti, Enrico, 3 Betti's reciprocity theorem, 166-167 Bilinear forms, 216 Body (volume) forces, 51 Boundary conditions: bars, finite element analysis of, 448 bendiug of plates, 362-368 Dirichlet, 113,250 essential (geometric), 96, U 2-117,249 Enler-Bernoulli beam theory, 461-463 natural (force), 96,112-117.250 Boundary-value problems, 245-259 Ritz approximations, 254-259 variational formulations, 245-254 Bounded from above/below, 206 Brachistochrone, 2-3 Buckling: circular plates: analytical solutions, 317-320 variational solutions, 333-334 in classical plate theory: circular plates, 317-320, 333-334 fully discretized finite element models, 475 Navier solutions of rectangular plates, 352-354 variational solutions, 384-396 rectangular plates: Navier solutions, 352-354 shear deformation plate theory, 411-412 variational solutions, 384-396 Buckling load, criticul, 264, 352 Calculus of variations, see Variational calculus Cartesian, 19 Cartesian spaces, 207 Castigliano, 3 Castigliano's Theorem I, 146-151 583
584 INDEX Castigliano's Theorem II, 156-164 Cauchy stress formula, 31 Cauchy stress tensor, 3 i CCCC plates, 367- 368 CCSS plates, 365 CFFF plates, 367 CFSF plates, 367 CFSS plates, 365-366 Change, rate of, 22 Characteristic equation, 37 Characteristic vectors, see Eigenvectors C1E, see Consistent interpolation element Circular plates: bending in: analytical solutions, 312-317 variational solutions, 320-329,421-424 buckling in: analytical solutions, 317-320 variational solutions, 333-334 shear deformation plate theoiy: exact solutions of axisymmetric circular plates, 401-405 governing equations, 396-398 vibration in, 329-333 Ciapeyron, 3 Classical plate theory (CPT), 128, 301-302, 305-396 bending: circular plates, 312-317, 320-329 fully discretized finite element models, 475 Le"vy solutions of rectangular plates, 359-362 Navier solutions of rectangular plates, 344-350 and shear deformation theory, 412^121 variational solutions, 362-379 buckling: circular plates, 317-320, 333 334 fully discretized finite element models, 475 Navier solutions of rectangular plates, 352-354 variational solutions, 384-396 circular plates, 305-334 bending, 312-317, 320-329 buckling, 317-320, 333-334 governing equations, 305-312 vibration, 329-333 finite element models of, 471-479 conforming/nonconformmg plate elements, 474-475 fully discretized finite element models. 475-479 general formulation, 472^474 governing equations: of circular plates, 305-312 in rectangular coordinates, 334-342 Le'vy solutions of rectangular plates, 359-362 mixed finite clement models of, 527-538 mixed model 1,527-532 mixed model IT, 532-535 Navier solutions of rectangular plates, 342-356 bending, 344-350 buckling analysis, 352-354 natural vibration, 350-352 transient analysis, 354-356 variational solutions: bending, 362-379 buckling, 384-396 vibration, 379-384 vibration; circular plates, 317-320,329-333 fully discretized finite element models, 476 Navier solutions of rectangular plates, 350-352 variational solutions, 379-384 Collocation method, 263-264 Complementary potential energy, principle of maximum, 156 Complementary strain energy, 85-95 of an elastic body, 88 definition of. 86 Complementary strain energy density, 85 Complementary virtual work, 97, 151-155 Completeness: and convergence, 225 of normed vector spaces, 212 of set of functions, 215 Components, 225 contravariant, 20 covaiiant.21,22 physical, 16 scalar, 16 of u vector, 15 Configurations. 95 Conforming elements, 474-475 Conjugate (trans(x>se) of dyadic, 32 Conservation of angular momentum, 53-54 Conservation of linear momentum, 50-53 Conservative stresses. 84 Consistent inleipolatioii clement (CIE), 467^68 Constitutive equations, 49,62-69 generalized Hooke's law, 62-65 plane stress, 65 thermoelustic, 65-69 Constrained systems, Hamilton's principle for, 190-196 Constraints, 190 Continuous systems, 177 Continuum, 48, 183-190 Contracted notation, 63 Contravariant components, 20 Convergence, 225-226 Coordinate functions, 225 Coordinate systems, 8 Comer behavior, 3 Cosines, direction, 18 Covariant components, 21,22 CPT, xee Classical plate theory Critical buckling load, 264, 362 Cross product, 13 Curl of a vector, 24 Curl theorem, 27 Damping coefficient, 182 Deformable bodies, 3 Deformation, kinematics of, 54-62 strain compatibility equations, 58-62 strain tensor, 54-58 Hegrocs of freedom, 459 Del operator, 23,26 Descriptions of motion, 49-50 Deviatoric tensor, 39 Differential operators, 23 Directional derivative, 22 Direction cosines. 18 Direct variational methods (direct methods), 204-289 functional analysis, 205-221 linear vector spaces, 206-211
INDEX 585 minimum of a quadratic functional, 216-221 normcd and inner product spaces, 211-215 transformations, lincnr and bilinear. 215-216 general boundary-value problems, 245-259 Ritz approximations, 254-259 variational formulations, 245-254 Ritz method, 221-245 approximation functions, properties of, 225-226 description of, 222-225 examples illustrating, 232-245 with general boundary-value problems, 254-259 general features of, 230 -232 parameters, equations for, 226-230 weighted-residual methods, 259-288 collocation method, 263-264 eigenvalue and time-dependent problems, 264-266 examples illustrating, 267-288 Gherkin's method, 262-263 least-squares method. 263 for undetermined parameters, 266-267 Dirichlet boundary conditions, 113, 250 Dirichlcl boundary-value problems, 113 Discrete structural systems, 3 Displacements: generalized, 146,396 virtual, 3. 83, 95, 133-136 Displacement-based shear tie formation theories, 302-304 Displacement finite clement models. 474 Divergence: of second-oider tensor, 36 of a vector, 23 Divergence tlworcm, 27 Dot product: double-, 35 of dyadic and vector, 32 Double-dot product, 35 Dual basis, 21 Dummy index, 16 Dyads, 32 Dyadic(s), 29-31 conjugate (transpose) of, 32 divergence of, 36 dot product of, with vector, 32 double-dot product between two, 35 eigenvectors associated with, 36-41 nonion form of, 32-36 stress, 31 symmetiic, 37 -38 Dynamic (Neumann) boundary conditions, 113 Dynamics, laws of, 3 £-5 identity, 17-18 Eigenvalues, 36-40,264-267 Eigenvectors, 36-41 Elastica (Euler), 3 Elaslicily, 3 hypcrelasticity, 63 ideal, 62 Elements (of a stf), 205 Element nodes, 436 Energy: complementary strain energy, 85-95 definition of, 83 kinetic, 177 potential, 81, 177. principle of minimum total, 142-146 total, 143 strain energy, 84-95 Energy methods, 1 Engesser, 3, 146 Entropy, specific* 85 Equations for the Ritz parameters, 227 Equilibrium problems, 266 Erdmann, 3 Essential (geometric) boundary conditions, 96, 112-117,249 Euclidean norm, 212 Euler* Leonhurd, 3 Euler-Bemoulli beam theory, 67, 152-153,164, 170, 194,201 examples illustrating, 90-94, 135-136,187 finite element analysis of, 454-463 approximation functions, derivation of, 456- 458 assembly of element equations, 459-460 boundary conditions, imposition of. 461-463 finite element model, 458-459 governing equation, 454^455 weak form over an element, 455-456 mixed finite element models of. 517-522 governing equatious and weak forms, 517-519 weak form finite element model, 519-521 wcighied-residual finite element models, 521-522 Eider equations, 110-132 Eulerian (spatial) description, 49 Kulerian (substantial) derivative, 71 Euler-Lagrange equations, 188 -190 External forces, 51 Extrema, 3 Extremum of a functional, 108 ! 10, 505-506 Pennat, Pierre de, 2 EFSS plates, 366 Finite element mesh, 436 Finite element method, 433-493 bars, analysis of, 434-454 approximation over an element, 436-441 assembly (connectivity) of elements, 444-447 calculation of reactions and derivatives of solution: postprocessing, 449-454 domain representation by finite elements, 435^36 Hnile element equations* 442-444 governing equation, 434-435 imposition of boundary conditions, 448 weak form, 44M42 classical plate theory models, 471^179 conforming and nonconforming plate elements, 474-^175 fully discretized linile element models, 475-479 general formulation, 472-474 Euler-Bernoulli beam theoiy, 454-463 assembly of clcmcni equations, 459-460 derivation of approximation functions, 456-458 finite clement model, 458^59 governing equation, 454-455 imposition of boundary conditions, 461^463 weak form over an element, 455-456
586 INDEX Finite element method (continued) first-order shear deformation plate theory models, 479-493 finite element model, 480-483 governing equations and weak forms, 479-480 numerical integration, 483-493 mixed variational formulations: beam*, 517-527 classical plate tlieory, 527-538 Timoshenko beam theory models, 463-471 consistent interpolation element, 467-468 displacement fmiie clement models, 464—466 governing equations, 463-464 reduced integration element, 466-467 sirpcrconvergait clement, 468-471 Finite elements, 433 First-order shear deformation plate theory (FSDT), 303-304 boundary conditions for, 405 and classical plate theory, 412 finite element models of, 479-493 governing equations and weak forms, 479-480 numerical integration, 483-493 potential energy of, 131 First-order tensors, 9, 33. See also Vector(s) First variation, 103,106-107 Fixed region, 70 Fixed spatial location, 49 Flexibility matrix, 164 Forces: body vs. surface, 51 friction, 180-181 internal vs. external, 51 principle of complementary virtual, 152-155 Force (natural) boundaiy conditions, 96, 112-117, 249, 250 Force systems, self-equilibrating, 97 Formulation, variational, 2 Fourier coefficients, 225 Free energy function, 85 Frequency, fundamental, 351 Friction forces, 180-181 FSDT, see First-order shear deformation plate theory Functional(s), L 106-107, 2(5 Euler equations with, 110-112 externum of a, 108-110,505-506 first variation of a, 106-107 linear. 106,215 minimization of, with equality constraints, 117-123 Lagrange multiplier method, 118-119 penalty function method, 120-123 minimizing curves of, 3 and natural/essential boundary conditions, 112-117 quadratic, 106 Funclionn? analysis, 205-221 inner product, 213-214 linear vector spaces, 206-211 and linear independence, 209-210 scalar multiplication in, 207 vector addition in, 207 minimum of a quadratic functional, 216-221 norm, 211-213 orthogonality, 214-215 transformations, linear and bilinear, 215-216 Fundamental frequency, 351 Fundamental lemma of variational calculus, 107-108, 215 Galerkin,B.G.,4 Galerkin method, 262-263 Galileo, 2 Generalized displacements, 146,396 Generalized Hookc's law, 62-65 Geometric (Diricblct) boundary conditions, 113 Geometric (essential) boundary conditions, 96, 112-117, 249 Gibb's free energy, 85 Global nodes, 436 Gradient vector, 22 Greatest lower bound, 206 Green, George, 3 Green's first and second theorems, 45 Hamilton, William, 3 Hamilton's principle, 177-198 for constrained systems, 190-196 for a continuum, 183-190 general foim of, 178 for particles, 177-183 Rayleigh's method, 196-198 for rigid bodies, 183 Heaviside step function, 356 Hel linger, li.,4 Hellinger-Rcissner variation principle, 508-513 Hermitc cubic interpolation functions. 457-458 Heterogeneous bodies, 62 Hilbcrt space, 214,215 Holonomic constraints, 190 Homogenous bodies, 62 Homogenous form, 225 Hooke's law, generated, 62-65 Hybrid variational formulations, see Mixed variational formulations Hyperelastic materials, 63 Ideally elastic bodies, 62 Ideal systems, 97 TIF (interdependent interpolation element), 469 Infinitesimal distance, 79 infinitesimal strain, 56 inner product, 213-214 Inner product spaces, 214 Integral relations, 24-28 Interdependent inteipolalion clement (UK), 469 Internal forces, 51 Interpolation functions, 146 Tsopcrimetric problem, 2 Isotropic bodies, 62 Isotropic materials, 6*1 Jacobi, Carl Gustav, 3 Kantorovich method, 278 Kelvin, Baron, 3 Kinematics, 49 of defonnalion, 54-62 strain compatibility equations, 58-62 strain tensor, 54-58 Kinetic energy, 177 Kirchhoff, Gustav Robert, 3 Kirchhoff hypotheses, 301-302 Kirchhoff plate theory, 301-302. See also Classical plate theory Kronecker delta, 17
INDEX 587 Lagrange, Joseph-Louis, 3 Lagrange equations of motion, 179-180 Lagrange interpolation functions, 438-439 Lagrange multiplier, IJ 8-119 Lagrangian function, 178-179 Lagrangian (material) description, 49-50 Lame*, Gabriel , 3 Laminated plates, 477 Laplacian operators, 23,26 Law of minimal properly of potential energy for stable equilibrium, 3 Least action, principle of, 3 Least-squares method, 263 Least Time, Principle of, 2 Least upper bound, 206 Lebcsgue spaces, 211 Legendrc, Adrien-Marie, 3 Leibniz, Gottfried, 2,3 Leibniz rule, 69 l^vel surfaces, 22 Levy method, 278, 359-362 L'H6pital,3 Limit (of a sequence), 225 Linear combinations, 219 Linear dependence, 219 Linear form, 250 Linear functionals, 106,215 Linear independence, 15,219-220 Linear momentum, conservation of, 50-53 Linear relations, 219 Linear vector spaces, 206-211 addition in, 207 and linear independence, 209-210 scalar multiplication in, 207 Lower bound, 206 Mapping, 215 Master element, 484 Material derivative, 71 Material (Lagrangian) description, 49-50 Maupertuis, Picrre-J^ouis dc, 3 Maxima. 3 Maximum complementary potential energy, principle of, 156 Maxwell, James Clerk, 3 Maxwell's reciprocity theorem, 167-169 Me~chamque analyilque (J ^grange), 3 Members (of a set), 205 Memoires de VAcademic des Sciences (Maupertuis), 3 Menabrea, 3 Miudlin plate theory, 303 Minima, 3 Minimizing curves, 3 Minimum of a quadratic functional, 216-221 Minimum potential energyx 3 Minimum total potential energy, 2 Minimum total potential energy, principle of, 142-146 Mixed boundary-value problems, 113 Mixed variational formulations, 502-539 extrcmum and stationary behavior of functionais, 505-506 mixed finite element models; of beams, 517-527 of classical plate theory. 527-538 principles of, 502-505 stationary variational principles, 506-513 Heliinger-Reissner variation principle, 508-513 minimum of total potential energy, 506-508 Rcissner variation principle, 513 variational solutions based on, 514-517 Moment, II, 12 Momentum; conservation of angular, 53-54 conservation of linear, 50-53 Momentum-flux tensor, 73 Motion: descriptions of, 49-50 equations of, 50-53 Lagrange equations of, 179-180 rigid-body, 54 Multiplication, vector dot product, 32, 35 inner product, 213-214 scalar multiplication, 207 Natural (force) boundary conditions, 96, 112-117, 249, 250 Natuiallawsv8 Natural metric, 211 Natural (noi-malivied) coordinate, 439,484 Natural norms, 213 Natural vibration, 350-352,410-411,476 Navier solutions of rectangular plates, 342-356 bending, 344-350 buckling analysis, 352-354 biaxial compression, 353 uniaxial compression, 353,354 natural vibration, 350-352 transient analysis, 354-356 Neumann boundary-value problems, 113 Neumann (dynamic) boundary conditions, 113 Ncwmark family, 476 Newton, Tsaac, 2.3 Newton-Raphson method, 231 Newton's second law of motion, 30,50-53,72, 177, 183 Newton's third law of motion, 29 Nodes: clement, 436 global, 436 Nonconforming elements, 474 Nonholonomic constraints, 190 Nonion form (dyadics), 32-36 Nonscalars, 9 Normal derivative* 27 Normalized (natural) coordinate, 439,484 Normal stress, 29 Nomiccl vector spaces, 211, 212 Norm (of a vector), 206, 211-213 Null (zero) vector, 11 Operators, 215 strictly positive, 216 symmetric, 216 Ordered pairs, 209 Orthogonality, 214-215 Ormonormal basis, 16 Ortlionormal vectors, 214 Orthotropic materials, 63 Outer product, 13 Parallelepiped, 13 Parallelogram law of addition, 10
588 INDEX Panicles, Hamilton's principle for, 177-183 Penalty function method, 120-123 Penalty parameter, 120 Pendulum, 181-182 Permutation (alternating) symbol, 17 Perpendicular to a level surface, 22 Physical components, 16 Physical vectors, 8,9 Plane stress constitutive relutions, 65 Plane stress-reduced stiffnesses, 65 Plane stress stales, 65 Plates, 299-300,477 Plate theories, 3, 301^24 bending solutions in, 412-421 classical, sec Classical plate theory displacement-based shear deformation theories, 302-304 high-order, 303-304 Kirchhoff plate theory, 301-302 Mindlin plate theory, 303 Reissner plate theory, 304 shear deformation, 396-424 exact solutions of rectangular plates, 405-412 governing equations in rectangular coordinates, 398-401 governing equations of circular plates, 396-398 variations solutions of circular and rectangular plates, 421^24 stress-based theories, 304-305 Poisson's ratio, 64 Polyadics, 33 Postfactors, 32 Potential energy, 81, 177 minimum of total. 506-508 principle of maximum complementary, 156 principle of minimum total, 142-146 total, 143 Prefactors, 32 Primary variables, 114 Product space, 209 Pythagorean theorem, 34 Quadratic functional (s). 106,216-221 Rate of change* 22 Rayleigh, Lord, 4 Rayleigh quotient, 197 Rayleigh's method, 196-198 Real linear vector spaces, 207 Reciprocal basis, 21 Reciprocity theorem(s): Betti's, 166-167 Maxwell's, 167-169 Rectangular Cartesian, 19 Rectangular plates: bending in: Levy solutions, 359-362 Navier solutions, 344-350 variational solutions, 421-424 buckling in; Navier solutions, 352- -354 variational solutions, 384-396 vibration in, 350-352 Reduced integration, 467 Reduced integration element (RIE), 466-467 Reissner. II., 4 Reissner plate theory, 304 Reissner variation principle, 513 Residuals, 248 RIE, see Reduced integration element Rigid bodies. Hamilton's principle for, 183 Rigid-body motion, 54 Ritz, W., 4,222 Ritz method, 197,221-245 approximation functions, properties of, 225-226 description or, 222-225 examples illustrating, 232-245 with general boundary-value problems, 254-259 general features of, 230-232 parameters, equations for the, 226-230 weighted-residual methods vs., 262 Scalar components, 16 Scalar differential operators. 23 Scalar products, 11-15 SCSS plates, 366 Secondary variables, 114,250 Second-order tensors, 29. See also Dyadics Self-adjoint operators, 216 Self-equilibrating force systems, 97 Self-equilibrating stresses/forces, 151 Scmidiscreti/alion method, 278 Seminorm, 211 Sense of travel, 29 Sequence, limit of a, 225 Sets, 205-206 Set of admissible configurations, 95 Set of competing (admissible) functions, 109 SFSS plates, 366-367 Shear deformation plate Ihcory, 396-424 bending solution of classical plate Ihcory vs., 412-421 exact solutions of axisymmelric circular plates, 401-405 exact solutions of rectangular plates, 405-412 bending analysis, 407^110 buckling analysis, 411-412 natural vibration, 410-411 first-order, 131 governing equations in rectangular coordinates, 398-401 governing equations of circular plates, 396-398 variations solutions of circular and rectangular plates, 421^24 Shear stress, 29 Shells, 299-300 Shell theories, 301-305 displacement-based shear deformation theories, 302-304 and Kirchhoff plate theory, 301-302 •stress-based theories, 304-305 Skew product, 13 Sobolev norm, 212-213 Sobolev spaces, 212-213 Solid continuum, 48 Solid mechanics, equations of, 48-69 angular momentum, conservation of, 53-54 classification of, 48-49 constitutive equations, 62-69 generalized Hooke's law, 62-65 plane stress, 65 thermoelastic, 65-69 deformation, kinematics of, 54-62
INDEX 589 strain compatibility equations, 58-62 strain tensor, 54-58 Hooke's law, generalized, 62-65 linear momentum, conservation of, 50-53 motion: descriptions of, 49-50 equations of, 50-53 plane stress constitutive relations, 65 strain compatibility equations, 58-62 strain tensor, 54-58 stress tensor, symmetry of, 53-54 thcrmoelaslic constitutive relations, 66-69 Space-time approximations, 266 Spatial (Eulerian) description, 49 Specific entropy, 85 Spring-mass systems, 81-82 Square-integrable functions, 209 Statically admissible field of variation, 151 Stationary variational principles, 506-513 Hellinger-Reissncr variation principle, 508-513 minimum of total potential energy, 506-508 Reissner variation principle, 513 Stake's law, 182 Strain, infinitesimal, 56 Strain compatibility equations. 58-62 Strain energy, 84-95 of an elastic body, 88 definition of, 86 Strain energy density, 85, 86 Strain energy density function, 63 Strain tensor, 54-58 Stress(es): conservative, 84 normal, 29 shear, 29 Stress-based plate theories, 304-305 Stress tensor, 31 Cauchy,31 definition of, 31 symmetry of, 53-54 Stress vector, 29.33-34 Strictly positive operators, 216 Structural mechanics, 133-169 potential energy: complementary. 156-164 total, 142-151 reciprocity theorems, 164-169 virtual forces, 151-155 virtual work, 133-142 principle of virtual displacements, 133-142 unit-dummy-dHplacement method, 137—142 Subdomain method, 273 Subspaces, 207 Substantial (Eulerian) derivative, 71 Summation of vectors, 16-19 Supeiconvcrgent elements, 459.469-471 Sup-norm, 211 Surface forces, 51 Surfaces* level, 22 Symmetric bilinear forms, 216 Symmetric dyadic, 37-38 Symmetric uperatoi-s, 216 Symmetry (of stress tensor), 53-54 Tensor(s), 9 deviatoric, 39 first-order, 9,33. See afro Vector(s) momentum-Rux, 73 second-order (dyadics), 29-41 eigenvectors associated with, 36-41 genera! properties of, 32 nonion form of, 32-36 strain, 54-58 stress, 31,53-54 symmetry of stress, 53-54 third-order, 33 trace of a, 35 unit, 34-35 zcroth-order, 33 Thermodynamic form (of energy equation for a continuum), 73 Thermoclaslic constitutive relations, 66-69 Thin shells and plates, 299 Third-order beam theory, 187-190 Third-oitler tensors. 33 Time-dependent problems, weighted-residual methods with, 264-267 Timoshenkobeam theory, 67, 127, 170,190, 256 displacement finite clement models of, 463-471 consistent intej-polation clement, 467-468 displacement finite element models, 464-466 governing equations, 463-464 reduced integration element, 466-467 supcrconvergent element, 468-471 mixed finite clement models of, 523-527 ASD-HQLC element, 526-527 ASD-LLCC clement, 524-525 ASD-QLCC element, 525-526 general finite clement model, 523-524 governing equations. 523-527 Total potential energy, 1, 2,143,506-508 Trace of a tensor, 35 Transformations, 215-216 Transient analysis, 354-356 Transient response, 354 Transpose (conjugate) of dyadic, 32 Trefftz method, 285 Triple product, vector, 13-14 Undetennined parameters, 204, 266-267 Unitary system, 19 Unit-dummy-displuccment method, 137-142 Unit-dummy-load techniques, 3 Unit tensor, 34-35 Unit vector, 10-11 Upper bound, 206 Variables: primary, 114 secondai"y, 114 Vaiiution(s): definition of, 103 first, 103 Variational calculus, 2, 102-123 boundary conditions, natural and essential, 112417 Euler equations, 110 412 functional, 106-107 extremum of, 108-110 first variation of, 106-107 fundamental lemma of, 107-108,215
590 INDEX Variational calculus (combined) minimization uf functional with equality constraints, 117-123 Lagrange multiplier method, 118-119 penally function method, 118-119 mixed variational formulations, 514-517 operator, variational, 102-106 Variational formulation, 2 Variational methods, 1 direct, see Direct variational methods history of, 2-4 Variational operator, 102-106 Variational principles, 1,2 Variational problem, 223 Veclor(s), 8-28 abstract, 8-9 addition of, 207 characteristic, see Eigenvectors components of, 15 curl of a, 24 defined, 9-11 divergence of a. 23 as firsi-ordcr tensors, 9 gradient, 22 inner product of, 213-214 integral relations, 24-28 multiplication of; dot product, 32, 35 inner product, 213-214 scalar multiplication, 207 as nonscHlars. 9 norm of, 206. 21 i-213 notation for, 9-10 null, 11 orthogonal, 214-215 orthonormal, 214 physical, R, 9 products of, 11-15 scalar multiplication of, 207 stress, 29, 33-34 summation convention, 16-19 unit, 10-11 vorticity, 71 zero, 11 Vector calculus, 19-26 Vector components, 16 Vector differential operators, 23 Vector from h linear vector space, 8 Vector products, 32-15 Vector spaces: linear, 206-211 linear independence, 209-210 scular multiplication, 207 vector addition, 207 normed,206,211 subspaces of, 207 Vibration: circular plates, 317-320, 329-333 natural, 350-352.410-411,476 Navier solutions of rectangular plates, 350-352 rectangular plates: Navier solutions, 350-352 shear deformation plate theory* 410-411 shear deformation plate theory, 4-10-411 variational solutions, 379-384 Virtual displacements, 3> 83,95. 133-136 Virtual forces, principle of, 151-155 Vutual velocities, 2 Virtual work, 95-102 complementary, 97,151-155 definition of, 96 and principle of minimum total potential energy, 142-146 in unit-dummy-displacement method, 137-142 and virtual displacements, 133-136 Voigt-Kelvin notation, 63 Volume (body) forces, 51 Vorticity vector, 71 Weak forms, 245 Weierstrass, Karl Theodor, 3 Weighted-residual methods, 259-288 collocation method, 263-264 eigenvalue and time-dependent problemSj 264—266 examples illustrating, 267-288 Galcrkin's method, 262-263 least-squares method, 263 Ritz method vs., 262 for undetermined parameters, 266-267 Weight function(s), 248v 250, 261 Work, 11,79-83 definition of, 79-81 by external vs. internal forces, 81-83 virtual, 95-102. See also Virtual work Young's modulus, 64 Zero (null) vector, 11 Zeroth-order tensors, 33
ABOUT THE AUTHOR J. N. Reddy is a D istinguished Professor and the holder of the Oscar S. Wyatt Endowed Chair in the Department of Mechanical Engineering at Texas A&M University, College Station, Texas. Prior to this current position, he was the Clifton C. Garvin Professor in the Department of Engineering Science and Mechanics at Virginia Polytechnic Institute and State University (VPI&SU), Blacksburg, Virginia, Professor Reddy is internationally known for his contributions to variational principles and methods, refined theories and finite element models of laminated composite plates and shells, and finite element analysis of problems in fluid mechanics and heat transfer, as well as in solid mechanics. He is the author of over 260 journal papers and 12 other books, including An Inti'oduction to the Finite Element Method, McGraw-Hill, 1984 (second edition, 1993); Energy and Variational Methods in Applied Mechanics, John Wiley & Sons, 1984; Applied Functional Analysis and Variational Methods in Engineering, McGraw-Hill, 1986; Mechanics of Laminated Composite Plates: Theory and Analysis, CRC Press, 1997; and Theory and Analysis of Elastic Plates, Taylor & Francis, 1999....... Professor Reddy has received both teaching and research awards from the institutions where he taught. Hehas received, numerous national awards including the Walter L. Huber Civil Engineering;: Research Prize of the American Society of Civil Engineers (ASCE), the Worcester Reed ...Warner Medal and the Charles Russ Richards Memorial Award of.the! American: Society, of Mechanical Engineers (ASME), the 1997 Archie Higdori Distinguished Educator Award from the American Society of Engineering Education (ASEE), the 1998 Nathan M. Newmark Medal from ASCE, arid the 2000 Excellence in the Field of Composites Award from the American Society of Composites. He is a Fellow of ASCE, ASME, the American Academy of Mechanics (AAM), the U.S. Association of Computational Mechanics (USACM), the International Association of Computational Mechanics (IACM), and the Aeronautical Society of India (ASI). He is- the main editor of the journal Mechanics oj Composite Materials and Structures, editor of International Journal 591
592 ABOUT THE AUTHOR of Computational Engineering Science (World Scientific Publishers, Singapore) and International Journal of Structural Stability and Dynamics (World Scienliiic Publishers, Singapore), and serves on the editorial boards of over two dozen journals, including International Journal for Numerical Methods in Engineering, International Journal for Numerical Methods in Fluids, and Engineering Structures.