II[IIIgmnrgnflnmnlqmnII new“ A}. Univtfli‘)’ g :- This is to certify that the thesis entitled DESIGN SENSITIVITY IN DYNAMICAL SYSTEMS presented by JOSEPH EUGENE WHITESELL, JR. has been accepted towards fulfillment of the requirements for PH.D. degreein MECHANICAL ENGINEERING I W f Major professor Date 8-8-80 0-7639 ovanoug FINES: 25¢ per day per in: , hng'mue Liamv mains: Place in book return to rénové charge from circulation records .I I ' cc- DESIGN SENSITIVITY IN DYNAMICAL SYSTEMS By Joseph Eugene Nhitesell, Jr. A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 1980 ABSTRACT DESIGN SENSITIVITY IN DYNAMICAL SYSTEMS By Joseph Eugene Whitesell, Jr. Modal analysis is a standard tool used in the design of dynamical systems. By numerically determining certain eigen- values and eigenvectors associated with a linear structural model, its vibratory behavior can be investigated. The difficulty with this technique is that it may be costly in computer time if several redesigns (and re-evaluations of the eigenvalues and eigenvectors) are needed to find a suitable design. The thesis presents a technique to represent selected eigen- values and eigenvectors by a Taylor series in a design variable. Since, with the technique, the Taylor coefficients can be computed efficiently and to arbitrary order, the Taylor series can be used to accurately estimate the consequences of a design change. The technique, therefore, has the potential to substantially reduce the number of calculations necessary to re-evaluate the eigensystem after a design change. Each Taylor coefficient is determined by evaluating a recursive formula which is derived through an application of generalized in- verse theory. If the eigenvector is found through an inverse itera- tion, the technique is especially efficient. Most of computational effort spent to find the eigenvector can then be re-applied to find the Taylor coefficients. Two representative examples illustrate the power of the method for practical design problems. DEDICATION To my father and mother. ii ACKNOWLEDGEMENTS There are several people who deserve acknowledgement for helping me during my studies and through the preparation of this dissertation. The advice and friendship of Dr. James Bernard, my major professor, has been invaluable. His quick mathematical insight and his engineering intuition as well as his high personal values and his good humor have made working with him stimulating and enjoyable. It was Dr. Albert Andry who first stimulated my interest in vibration problems and in the topic of this thesis. His ability as a lecturer and his good advice and friendship were major influences during the past three years. Both Drs. Andry and Bernard were cited last year by Michigan State University for excellence-in-teaching. Dr. Robert Schlueter, who was my advisor during my Master's degree studies, and a member of my thesis committee, has helped me numerous times during my studies here and I am greatly indebted to him. The West Coast Branch of my thesis committee consists of Herb, Margie, Marta, Erik, Evan and Lorin Rauch who have been good friends and they have enriched my life in many ways. I want to thank especially Dr. J.S. Frame for introducing me to linear algebra through his extraordinary teaching ability. Attending his course was a rare privilege. For the difficult job of typing this thesis I thank Jan Swift of the Department of Mechanical Engineering. I appreciate the programming help of Mr. Mark Zykin of the Case Center for Computer-Aided Design. I have much enjoyed working with Mark. Finally, I would like to thank Dr. John Brighton, Chairman of the Department of Mechanical Engineering, for his consistant support and leadership and for his advice. iv Chapter I - l.l l.2 Chapter 2 - 2.l 2.2 Chapter 3 - 3.1 3.2 3.3 3.4 3.5 Chapter 4 - 4.l 4.2 4.3 Chapter 5 - 5.1 5.2 .3 Chapter 6 - 6.1 6.2 6.3 Bibliography TABLE OF CONTENTS Introduction Some Historical Notes Thesis Overview Mathematical Preliminaries Concepts from Linear Algebra Concepts from Functional Analysis Eigensystem Differentiability Preliminary Differentiability Results The Reduction Process Eigenvector Differentiability Differentiation with Respect to Several Parameters Conclusions Differentiation of Eigensystems Differentiation of Fully Analytic Matrices Eigensystem Derivatives with Generalized Inverse Numerical Approach Design Sensitivity in Finite Element Models of Dynamical Structures The Finite Element Method Dynamical Problems Examples Discussion and Further Study Discussion Methods Suggestions for Further Study (JON l9 25 25 28 32 37 42 43 43 49 59 64 64 7o 78 94 94 95 96 97 LIST OF TABLES Table Page 5.3-6 The Effect of a Design Change on an Eigenvalue 85 vi Figure 5.l-ll 5.3-7 5.3—8 5.3-9 5.3-10 5.3-ll 5.3-12 LIST OF FIGURES The Triangulation of a Domain Triangular Element Eigenvalue Dependence on Design Baseline Design First Design Change Second Design Change Third Design Change Fourth Design Change vii Page 68 80 86 89 90 91 92 93 Chapter 1 Introduction The numerical determination of static and dynamic responses for linear structural models has become a routine procedure in engineering design [l,2]. Typically a sequence of design and redesign is followed until certain features such as weight, natural frequency, mode shape, and deflection and stress magnitudes attain suitable values. At each stage the designer changes the design and performs a numerical analysis to evaluate the consequences of the changes. This approach relies heavily on human judgement to modify the design at each stage and since the appropriate design modi- fications may not be obvious, several redesigns and reanalyses are often necessary. Alternative methods to improve this pro- cedure have been proposed. In one method, the full structural model is replaced by a Taylor series which approximates its behavior in a neighborhood of the current design [3]. Combined with interactive computer graphics hardware, this method allows a designer to preview the consequences of a proposed design while avoiding a costly reanalysis. Another approach is the optimal design method [3], which combines nonlinear programming methods with structural design techniques. The procedure identifies a cost functional with certain important design features and attempts to minimize it by changing the design. When the method is successful, a sequence of numerically determined designs con- verge to a feasible design for which the selected cost functional attains a local minimum. A major difficulty arises when these methods are applied to large problems since the determination of the system's state variables and especially their sensitivity values may require an excessive amount of computation. However, the potential benefits of these methods provide a strong incentive to develop improved numerical methods. l.l Some Historical Notes When the design involves dynamical measures such as natural frequencies or mode shapes, efficient methods for determining the sensitivities of eigenvalues and eigenvectors to design changes become important. Although methods for computing the first derivative of an eigenvalue with respect to some system parameter have been known since the nineteenth century work of Jacobi [4], eigenvector derivatives have a much shorter history. The first work in this area was done by Fox and Kapoor [5], who presented two methods for determining an eigenvector derivative for the symmetric eigenvalue problem (K - AM)u = 0. Unfortunately both methods are computationally inefficient. The first method, although it requires only knowledge of the specified eigenvalue and eigenvector, involves the multiplication of an (n + l)xn matrix by its transpose and the subsequent solution of a fully p0pulated symmetric system. The matrix multiplication is a lengthy computation and usually leads to a loss of any sparseness possessed by the factor matrices. The second method avoids these problems, but it expresses the derivative of an eigenvector in terms of a complete set of eigenvalues and eigenvectors. This formulation has only theoretical value when the eigensystem is large since it is difficult to extract a full set of eigenvec- tors. In refs. [6-8] the work of Fox and Kapoor was extended to non-symmetric systems but no improvement in computational ef- ficiency was made. These computational difficulties were first discussed by Nelson [9]. As a remedy, he proposed a technique by which the rank (n - l) matrix L — AI is modified by zeroing certain of its entries. The modified matrix, which describes a system of equations which must be solved to determine the eigenvector deri- vative, is well conditioned and retains the sparseness of the original system. This is a significant improvement since the problems of mathematical physics and engineering often involve sparse matrices. In this thesis further improvement in computing eigensystem sensitivities are presented. Two representative examples illustrate the power of the method for practical design problems. l.2 Thesis Overview Notation and preliminary mathematical concepts are presented in Chapter II. In Chapter III a survey of eigensystem differen- tiability results are presented. The presentation is based on the literature of the perturbation theory for linear operators. After two example problems the chapter closes with a discussion of eigensystem dependence on several parameters. Chapter IV is concerned with methods for differentiating eigenvalues and eigenvectors. The chapter states some basic results, followed by an application of generalized inverse theory which leads to further improvements in computing eigensystem derivatives. A recursive numerical method is then described for efficiently computing eigensystems derivatives of arbitrary order. Each eigenvalue and eigenvector derivative is determined by solving a sparse triangular system and involves no more than 0(n2) multiplications. Chapter V begins with a statement of the design sensitivity problem in finite element models. The methods of Chapter IV are then extended to the generalized eigenvalue problem (K - AM)u = 0. The chapter ends with two examples which involve finite element formulations. In the first Example (5.3-l), an eigenvalue problem associated with a single element plane elastic vibration problem is considered. After introducing a design variable to the problem, a Taylor series representing an eigenvalue as a function of the design variable is determined with the technique presented in this thesis. The results are compared with direct evaluations of the eigenvalue at various values of the design variable. In the second example (5.3-7) an eigenvalue which depends on the boundary shape of a fixed-fixed vibrating plate is estimated using a first order approximation. The estimated eigenvalue is compared with a direct evaluation of the eigenvalue for several design modifications. Chapter VI presents concluding remarks and suggestions for future work. Chapter 2 Mathematical Preliminaries This chapter contains a summary of definitions and theorems drawn from linear algebra and functional analysis which are used in the following chapters. The reader may find this material in many sources such as refs. [10-14]. In this thesis an n dimensional column vector is denoted by x. Its transpose to a row vector is denoted by xT. The con- jugation of xT is denoted by x*. To avoid confusion between the ith vector in an indexed sequence of vectors and the ith component in a particular vector, 5i denotes an indexed vector and xi denotes the ith component of the vector x. The jth com- ponent of the vector 5i is denoted by xij' An nxm matrix is denoted by A, B, C etc. The entry in the ith row and jth column of A is denoted by aij' The ith matrix in an indexed sequence of matrices is denoted by A1. The entry in the jth row and kth column of the matrix A1 is denoted by aijk' The real (complex) field is denoted by 32(3). The real (complex) vector space of dimension n is denoted by 3?"(@n). The space of real (complex) nxm matrices is denoted by ‘gynxm (gnxm) . 2.l Concepts from Linear Algebra Definition (2.l-l): A linear space X is a set of elements called vectors which is closed under addition (if XeX and yeX then x+y = 2eX) and under multiplication by a real or complex scalar a (if XeX then ax = yeX). Definition (2.l-2): Let X be a linear space over a field F and let V be a set of k vectors {54’52""’5k} of X. Then if for some set of k scalars {c],c2,...,ck} from F, not all zero, the linear combination C154 + c252 + ... + Ckék = 0, the vectors {§I’§Z""’§k} are said to be linearly dependent over F. If instead c154 + c252 + ... + ckxk = 0 only if each ci=0 then the vectors {54’52"°”5k} are linearly independent. Definition (2.1-3): A set of vectors {54’52’°"’5k’°"} is said to span X if every vector XeX can be expressed as a linear combination of the set {54’52’°"’5k""}' Definition (2.l-4): A set of vectors {54’52""’§k} of a linear space V is said to be a Hamel basis if and only if i) the set spans V and ii) the vectors {54’52""’5k} are linearly independent. Definitions (2.l-l) through (2.l-4) are concerned with only the algebraic properties of a linear space (algebraic linear space). If these notions are combined with the topo- logical notions of length, distance and convergence, a space with both algebraic and topological properties results (topo- logical linear space). Definition (2.l-5): A norm on a linear space X is a positive real valued function with the following properties 1) IIXII f 0 if x f O, lIx|| = 0 otherwise 1'1')||9><|L 96F iii) ||§1+§2||1|l51||+|l52l| Definition (2.l-6): An inner product is a scalar function, (-,-), of two elements x,yeX such that X142 + x3) = (51.52) + (51.253) 54: 052) = aIéqgéz), aeF x,x) > 0 if x f 0, (x,x) = 0 otherwise. Definition (2.l-7): Two vectors x and y are orthogonal with respect to an inner product (~,-) if (x,y) = 0. Definition (2.l-8): A set of vectors 54 forms an orthonormal §§t_with respect to an inner product (.,.) if Definition (2.l-9): Two sets of vectors 5i and yfi form a bi; orthonormal set with respect to an inner product (-,-) if (‘é-l :Xj) : 51i- Definition (2.l-lO): The transpose of a matrix A is the matrix T- = A - B such that aij bji' matrix A is the matrix A* = B such that aij = bji' The Hermitian transpose of a Definition (2.l-ll): A matrix A is called symmetric (Hermitian) if A = AT (A = A*) or skew-symmetric (skew-Hermitian) if A = -AT (A = -A*). Definition (2.l-l2): A matrix A is upper triangular if aij = O for i > j, lower triangular if aij = D for i < j and diagonal . = 0 for i f j. if aiJ Definition (2.l-l3): The matrix A whose elements are all zero with the exception that a1. = l is given the special symbol, A = e.. J 1J and is called the ij_matrix unit. Definition (2.l-l4): The product of an mxk matrix A and an kxn matrix B is the matrix C = AB such that k Cij = Z aipbpr p=l Definition (2.l-l5): For positive integers, the powers of an nxn matrix A=A1 are defined by Definition (2.l-l6): A matrix A is idempotent if A2 = A. Definition (2.l-l7): If the negative integral powers of the nxn matrix A exist, they are defined by A.“ = (A-I)n IO where A'1 is called the inverse of A and I If A" exists the matrix A is called non-singular otherwise it is called singular. Definition (2.l-l8): The rank r of an mxn matrix A is the maximum number of linearly independent columns of A. Definition (2.l-l9): A generalized inverse of an mxn matrix A of rank r is any nxm matrix AI of rank r such that AAIA = A. I The mxm matrix AAI and the nxn matrix A A are each idem- potent matrices [10]. Theorem (2.l-20): Any solution X of the linear system AX H .< where A is an mxn matrix of rank r X is an nxk matrix Y is an mxk matrix ll may be expressed as X = A Y + A Z where A satisfies AA = 0 and Z is arbitrary [10]. O 0 Definition (2.1-21): The scalars A and corresponding vectors u which satisfy the equation Lu = Au are called the eigenvalues and corresponding right eigenvectors of the matrix L. The eigenvectors of LT are called the left eigenvectors of L. The subspace consisting of the origin and all the right (left) eigenvectors corresponding to A is called the right (left) eigenspace of L corresponding to A. Theorem (2.1-22): The eigenvalues of a real symmetric matrix are real. Theorem (2.1-23): Corresponding to any real symmetric matrix is a set of orthonormal eigenvectors. Definition (2.1-24): A matrix A is similar to a matrix B if there exists a matrix S such that 12 B = 5' AS. Theorem (2.1-25): If A and B are similar then they have the same eigenvalues. Definition (2.1-26): A matrix A is called diagonable if it is l similar to a diagonal matrix such that 3‘ AS = A, where the diagonal matrix A has the eigenvalues A1 of A as its entries and where the ith column of S and the ith row of S'1 are respective right and left eigenvectors corresponding to A1. Definition (2.1-27): A matrix A such that AA* = A*A is called a normal matrix. Definition (2.1-28): A matrix S is called orthogonal (unitary) if sTs = I (s*s = I). Theorem (2.1-29): Every real symmetric matrix A may be diagonalized by an orthogonal matrix S S AS = A where the diagonal matrix A has the eigenvalues, A1 of A, as its entries. 13 Definition (2.1-30): If f(A) is a polynomial such that m m-l f(A) fox + f1A + ... + f A + fm m-1 then f(A) Am'I + ... + f m fOA + f1 m-lA + fmI is the corresponding matrix polynomial. Theorem (2.1-31): If f(A) is an analytic function defined on a simply connected domain Dg; i? and if A is diagonable matrix such that -1 l S AS = A and A = SAS- then f(A) = Sf(A)S where f(A) = :E:f(i.)e . Theorem (2.1-32): If f(A) and A are as defined in Theorem (2.1-31) then 14 where A, = SeiiS']. The matrix EH is called the ith rank one constituent idempotent matrix and AiAj = 5iin and n 2 A]. = I. Furthermore A]. = g]. y_1.' where u]. is the ith i=1 column of S (right eigenvector) and v: is the ith row of S"1 (left eigenvector). Alternatively, if A has 5 distinct eigenvalues A, with multiplicity mi then S f(A) = 2 mpg, i=1 where mi mi 51: Z Sens-1:2 Aj’ jg{j|ii = A3,}. 3 J The matrix 5i is called the ith rank mi constituent idempotent 5 matrix and AiAj = 61.in and Z A. = I. i=1 Theorem (2.1-33): (Interpolation formula) The ith rank mi con- situent idempotent matrix Ai corresponding to a diagonable matrix A with 5 distinct eigenvalues is given by (A-A.I) S #‘I LAC—J. 15 Definition (2.1-34): The matrix RZ = R(z) = (A-zI)'] is called the resolvent matrix of A. Theorem (2.1-35): (Residue Theorem) Let F be a closed curve in the complex plane enclosing eigenvalues A1 of a matrix A, i = 1,k. Then k _ 1 Z A,- - - 21”.] R(z)dz. i=1 r Theorem (2.1-36): (Dunford-Taylor Integral) Let f(z) be an analytic function defined on a simply connected domain D g g and let P be a closed curve in ‘6’ enclosing all of the eigen- values Ai of a matrix A. Then if R is defined for 21 and 22. Definition (2.1-38): The matrix AA = A - AI is called the charac- teristic matrix of A. The equation det(AA) = 0 is called the characteristic eqpation of A. The notation det(-) denotes the determinant of a matrix [10]. 16 Theorem (2.1-39): (Cayley-Hamilton) Every matrix satisfies its characteristic equation. Definition (2.1-40): A function u(x): €+€n is called a vector-valued function. Definition (2.1-41): A function A(x): 952+ gm" is called a matrix-valued function. Definition (2.1-42): Suppose D is a simply connected domain in if , i‘I‘ L(x) is an nxn matrix-valued function of x e D and {Ai(x) 1 is a set of eigenvalues of L(x). Then if Ai(x) + A(x]) as x + x1 where A(x]) is an eigenvalue of L(x1) with multi- plicity m and x1 2 D, the set {Ai(x)}? is called the A-gY‘OUE. Now let P be a closed curve in i? enclosing a A-group of eigenvalues. Then TIT P(x) = --217 J{.R(z]x)dz where R(z]x) = (L(x) - 21)"1 P is called the total projector1 for the A-group [14]. 1Note that P(xo) is identical to the rank m constituent idempotent matrix corresponding to A(x0). Note also the connection to Theorem (2.1-35) which shows that the total projector is the sum of the in- dividual consituent idempotent matrices of the A-group. That P(x) is also idempotent follows from Theorems (2.1-32) and (2.1-35). 17 Theorem (2.1-43): (Product Rule) i) If A(x) and B(x) are differentiable matrix-valued functions such that C(x) = A(x)B(x) then 0.0. xlci ll 3: 0.0.. XW + 9—0. ><3> CD ii) If u(x) is a differentiable vector-valued function such that v(x) = A(x)u(x) then where gig-ord a—-indicates the differentiation of the individual entries Cij or Vi with respect to x. Theorem (2.1-44):'If A(x) is differentiable and n is a positive integer n WEE: -1 dA An- i i=1 and if A(x) is also non-singular dA'" _ -n 6A" 3;- - - A ai‘ A Definition (2.1-45): An eigenvalue A of a matrix A is called simple if it is not repeated. An eigenvalue A of multiplicity m is called semi-simple if its eigenspace has dimension m. All the 18 -T(I~ ' TF'I. .. eigenvalues of a diagonable matrix are semi-simple. 2.2 Concepts from Functional Analysis Definition (2.2-l): Let {xn} be an infinite sequence of elements in a normed space X such that lim||xn - x|| = 0. Then {xn} n-mo is said to converge (strongly) to x. Definition (2.2-2): A sequence {xn} in a normed space is called a Cauchy sequence if for any 6 > 0 there is an N(e) such that ||xm - xnll < e for every m, n > N. Definition (2.2-3): A normed space X is said to be complete if it contains the limit point of every Cauchy sequence in X. A complete normed linear space is called a Banach space. Definition (2.2-4): An inner product space is a linear space X on which (x,y) is defined for each pair of elements x,y in X. A complete inner product space is called a Hilbert space. (Since every inner product generates a norm, any Hilbert space is also a Banach Space.) 19 Definition (2.2-5): A linear operator L is a mapping between linear spaces such that i) The domain 99(L) of L is a linear space and the range 159(L) lies in a linear space over the same field F ii) For all x,y c @(L) and aeF L(x + y) = Lx + Ly L(aX) = aLx. Definition (2.2-6): The Null Space of L denoted./V(L) is the set of all x c @(L) such that Lx = O. The following standard mapping notation will be used. A function f which maps from a set A into a set B, is denoted by f:A + B l. A function f:A + B is surjective (onto) if and only if each b e B is an image of some element of A. 2. A function f:A + B is injective (one to one) if for each b e 39(f) there is exactly one a e A such that b = f(a). 3. A function f:A + B is bijective (one to one and onto) if and only if it is subjective ggg_injective (that is if and only if every b e B is the unique image of some a e A.) Definition (2.2-7): A function f:A + B is called a homeOmorphism if it is bijective and if both f and f-]:B + A are continuous. 20 I‘OI'J l i‘ |-: : '..‘ Definition (2.2-8): Let X and Y be normed spaces and L:£3(L) + Y a linear operator, where @(L)§ X. The operator is said to be bounded if there is a positive real number c such that for all Xe@(L), IILXII :C ||X||o Theorem (2.2-9): If a normed space X is finite dimensional, then every linear operator on X is bounded. Theorem (2.2-10): If L is a linear operator such that the mapping L: £Z(L) + Y is injective then there exists the mapping -1 L :9I’(L) + @(L) which maps each y e 92(L) onto x 59(L) for which Lx = y. Theorem (2.2-ll): Let X and Y be linear spaces and L:QD(L) + Y. Then a) L']:.§?(L) —> @(L) exists if and only .A/(L) is empty. b) L"1 is a linear operator. Definition (2.2-12): Let X be a linear space, a function f defined for each u e X is a linear functional on X if f[au + 8v] = af[u] + Bf[v] for all u,v of X and a,8 e F. 21 Definition (2.2-13): The linear space consisting of all the bounded linear functionals defined on X is called the dual space of X and is denoted by X'. Theorem (2.2-l4): Let X and Y be Hilbert spaces with inner products (-,-)x and (-,-)y respectively and L:X + Y a linear operator. Then corresponding to L there exists a unique operator L*:Y + X called the adjoint of L which satisfies * = (L v,U)X (v,LU)y for every u e X and v e Y. If L* = L then L is self adjoint. Definition (2.2-15): A linear self-adjoint operator L is called positive definite (u,Lu) > 0 unless u 0. Theorem (2.2-l6): Let A be a non—zero scalar. Let X be a Hilbert space with inner product (-,-)x and L:X + X a bounded linear operator. Then if Lx - Ax = y (I) Lx - Ax = 0 (2) L*f - Af = g (3) L*f - Af = 0 (4) 22 i) (1) has a solution x for every y e X if and only if (2) has only the trivial solution x = 0. ii) (3) has a solution f for every 9 e X if and only if (4) has only the trivial solution f = 0. iii) (1) has a solution x (is normally solvable) if and only if (f,y)x = 0 for all solutions of (4). iv) (3) has a solution f (is normally solvable) if and only if (g,x)X = 0 for all solutions x of (2)- Definition (2.2-l7): Let X and Y be Banach spaces. Then a function f:X + Y is called Frechet differentiable at x e X if there exists a unique linear operator f'(x):X + Y such that Hm llflx + h) - f(x) - f'(x)hll . 0. llh||+O Ilhll Definition (2.2-18): Let X be a Hilbert space with inner product (-,-)X and then if f:X-+3? is a Frechet differentiable real valued function there exists a unique vector in X denoted by vf(x) called the gradient of f at x such that f'(x)h = (h,vf(x))x. Definition (2.2-l9): Let X be a Hilbert space with inner product (-,-)X and f:X-+3? , then the directional derivative of f at x is defined by 23 provided that this limit exists. If this limit exists for any direction h then f is called Gateaux differentiable at XeX. Remark: If f is Frechet differentiable then Df(x,h) exists for any direction h and Df(x,h) = '(x)h = (h,vf(x))x. Theorem (2.2-20): (Projection Theorem [15]) Let X be a Hilbert space with inner product (-,-)X, let ||-||X be the norm generated by (-,-)X and let M be a closed subspace of X. Then corresponding to any x e X there is a unique vector m e M such that ||v - m*||X §_||v - m||X for all m e M. (2.2-21) Furthermore (x - m*,m) = 0 is a necessary and sufficient condition that m* e M is the unique vector which satisfies (2.2-21). 24 Chapter 3 Eigensystem Differentiability In this chapter various results related to the differentia- bility of eigenvalues and eigenvectors are presented. 3.1 Preliminary Differentiability Results Suppose L(x) is an nxn matrix-valued function (See 2.1-4) whose elements are analytic functions of x e D _C_ 93’. It is natural to ask if and when L(x) has eigenvalues and eigenvectors which are also analytic functions of x. This question, which has been studied extensively by Rellich, Kato and others [15-17], un- fortunately does not have a succinct answer, since the problem inherits all of the complications associated with repeated eigen- values and defective eigenspaces from the underlying eigenproblem. Nevertheless it is vital to recognize whether the eigenvalue problem behaves smoothly if eigenvalue and eigenvector derivatives are to be used in computations. The following theorem gives an important sufficient con- dition for eigenvalue differentiability. Its proof may be found in cited literature. Theorem (3.1-l): Let L(x) be an nxn matrix-valued function of x e 3? , which is analytic in some neighborhood of x = 0. Then 25 a) If A(O) is a non-repeated eigenvalue of L(O), A(x) is an analytic function in some neighborhood of x = O and there exist left and right eigenvectors v(x) and u(x) which are analytic vector-valued func- tions of x in some neighborhood of x = 0. b) If A(O) has multiplicity m and {Ai(x)}m 1 is the A-group, Ai(x) may be expanded in a Puiseaux series [18] as 1 2 Ai(x) = A(O) + ulxp + nzxp + ... , p is an integer :_m Proof: [14, 18 and 19]. It is easy to show that the conditions of Theorem (3.1-l) (a) are not necessary by constructing a matrix which has repeated analytic eigenvalues and analytic eigenvectors. Example (3.1-2): Let x) (3.I-3) where the Ai(x) are analytic scalar functions not necessarily distinct for x e D g; Q and the _u_1.(x) are analytic vector- valued functions of x e D such that the matrix S(x), whose ith column is g4(x), is invertible. Then S'I(x) is also 26 analytic [l4], and we may take its ith row as v4(x). Theorem (2.1-32) shows that the Ai(x) are analytic eigen- values of L(x) and the vi(x) and ui(x) are analytic left and right eigenvectors of L(x), even if Ai(x1) = Aj(x]), i f j for some isolated value x1 (exceptional point). So L(x) has the desired properties. The matrix L(x) in example (3.1-2), although it has repeated eigenvalues, is diagonable due to the manner in which it was constructed [10]. That an analytic matrix is diagonable however, may not be taken as a necessary or sufficient condition for eigenvalue analyticity as the next example shows. Example (3.1-4): The matrix [0 x O L(x) = 0 0 x x 0 l .. J is diagonable for x3 f 4/27 yet its eigenvalues cannot be written as a Taylor series in x at x = D [14] whereas the matrix 27 L(x) = is not diagonable for any x 5'8’ but its eigenvalue x is clearly analytic. 3.2 The Reduction Process To gain further insight into the behavior of eigenvalue differentiability the reduction process of Kato [14] is useful. Suppose L(x) is an analytic matrix-valued function with the ex- pansion L(x) = j[: xl L(l)(0). (3.2-l) If A(x) is an analytic eigenvalue of L(x) with multiplicity m then a particular branch of A(x) may be expanded as A(x) = A(O) + xA(])(O) + x2A(2)(O) + ... + ... . (3.2-2) Kato has demonstrated that under some circumstances the sequence of coefficients A(1)(O) may be identified with certain eigen— values of a sequence of matrices, L(i)(x), which are described below. The following theorem is central to the process. 28 Theorem (3.2-3): Let L(x) be an analytic matrix with expansion L(x) = Z x‘LI‘)(0) (3.2-4) i=0 and let A(O) = A be a semi-simple eigenvalue of L(D) = L of multiplicity m with total projector P(D) = P. If Agl) is an eigenvalue of L(]) = PL(I)P in the subspace 99(P) with corresponding constituent idempotent matrix Pg1) then L(x) has exactly m§1) = dim Pgl) repeated eigenvalues of the form A + XA§]) + 0(x). Proof: From the first resolvent equation Theorem (2.1-37) we find that (L(x) - AI) R(z,x) = I + (z - A) R(z,x) . (3.2-5) Suppose that r is a closed curve in ‘8’ which encloses the A-group eigenvalues of L(x) then (L(x) - AI) P(x) = - I “/7 (z - A)R(z,x)dz (3.2-6) P an follows from applying theorem (2.1-35) to (3.2-5) and re- calling Definition (2.l-42). Since R(z,x) is analytic in x [14] (3.2-6) can be written as 29 (L(x) - AI) P(x) = jg: xl£(l)(0) (3.2—7) where L(1)(0) = — 1"}f (z - A) R(i)(z,0) dz . (3.2-8) Since A is a semi-simple eigenvalue both sides of (3.2-7) must vanish at x = D so L(O)(O) = 0 and “,8 (L(x) )2 AUPIX) = L(1)(O) X—> Now if uj(x) is an eigenvector of Aj(x) where Aj(x) is a member of the A-group then lim (L(x) - All P(x)u.(x) = lim j . X+0 X J X+0 X J which shows that A§I)(O) is an eigenvalue of L(])(O). To complete the proof we must evaluate (3.2-8) for i = 1. From Theorem (2.1-44) R(I)(z,0) = - R(z,0) L(])(O) R(z,0) we may expand R(z,0) as 30 R(z,0) = (z — A)-]P(O) + 5(2) where S(z) has no singularities so —1 Amen) = [(2 - A>“P<0) + S(zn Wm) [(2 - A) 9(0) + 5(2)]. Integrating (z - A)R(I)(z,0) along r shows that This theorem can be re-applied to L(])(x) if its eigenvalues are also semi-simple to derive a higher order expansion for A(x). In this case we have A(x) = A+ Xigll + x2 g2) + 0(x2), k = i,m§‘) (3.2-5) where the Aji are the repeated eigenvalues of ~(2) _ (l) (2) (l) (1)(1)+(l) (l) L Pj L Pj - Pj L LAL Pj (3.2-6) in the subspace 3?(P§])). The matrix L: is defined by n -l . . . L: = - ii: (A - A11) Li if L 15 diagonable. (3.2-7) i=1 The reduction process then consists of repeated applications of Theorem (3.1-8) to the matrices L(i). 31 The reduction process may also be used to generate expan- sions for the constituent idempotent matrices, so if it can be shown that each L(i) has semi-simple eigenvalues then the process may be continued indefinitely to show that the eigenvalues of L(x) are analytic and to establish the existance of analytic eigenvectors (See Theorem (4.1-4)). In the next section, the question of eigenvector differentia- bility is discussed through examples involving the interpolation formula Theorem (2.1-33) for constituent idempotent matrices. 3.3 Eigenvector Differentiability Eigenvector differentiability is more complicated than eigen- value differentiability. This is due, in part, to the possibility that the dimension of the eigenspace associated with a particular eigenvalue may change abruptly or disappear entirely. In this section we shall use the interpolation formula of Theorem (2.1-33) to study the behavior of eigenvectors corresponding to analytic semi-simple eigenvalues in two examples. In the first example the matrix L(x) has no exceptional point (See Example (3.2-2)) in its domain D. Example (3.3-l): Suppose the eigenvalues of L(x) can be represented by the s analytic functions A](x), A2(x),...,AS(x) where Ai(x) is assumed to have multiplicity mi and where L(x) is analytic and diagonable for x e D. 32 Since L(x) is diagonable, we may express the rank mi con- stituent idempotent matrices of L by the interpolation formula of Theorem 2.1-33. For the ith constituent idempotent matrix, Li’ we have L.(x) = j , x e 9 (3.3-2) “ LI (L(x)-i (x)) If there is no exceptional point in D, the Lj(x) are analytic on D. Any column of L4(x) which is non-zero on D represents an analytic eigenvector gg(x) corresponding to Ai(x) (See Theorem (2.132)). We now assume, without loss of generality, that x = 0 is in D. Then if g4(0) is an eigen- vector of L(D) corresponding to Ai(0), Li(x)gg(0) is an analytic eigenvector corresponding to Ai(x) in some neigh- borhood of x = 0. This is true since (1) S (A.(0)-A L.(D) .(0) = ‘ ‘1 3‘ U (A1.(O)-A .fl. .(0 J u.(O) =u.(0)7‘0 3‘0 LIL]. and the continuity of L4(x) imply that g4(x) is non-zero in a neighborhood of x = 0 and since (2) by the Cayley-Hamilton theorem, Li(x)g4(0) is in the null space of L(x) - Ai(x)I. 33 Now suppose that x = 0 is an exceptional point in D such that Ak(0) = Ai(0). Then L4(x) is not continuous at x = 0. It may have a removable discontinuity at x = 0 however.1 In the fol- lowing example we study the behavior of L4(x) in a deleted neighborhood of an exceptional point. Example (3.3-3): Suppose L(x) is an nxn matrix which is dia- gonable for x e D with n eigenvalues Ai(x) which are distinct analytic functions on D. If A 0) = Ai(0) for x = 0 e D k( then x = 0 is an exceptional point in D since by hypothesis Ak(x) and Ai(x) are distinct analytic functions. We will now investigate the behavior of S (L )-A.( I Li(x)y_(0) = n (X 3 X) ) 2(0) (3-3-4) i=1 (A1(X)-Aj(X)) Jfl in a deleted neighborhood of x = 0, where g(0) is any eigenvector taken from the two dimensional eigenspace cor- responding to Ai(0) = Ak(0). Both the numerator and denominator of L(x)u(0) become zero at x = D so lim L1.(x)y_(0) = lim d" dx 3(0) (3.3—5) X+O X+O 9&3IX) dA (x) de - dxk 1For example if L(x) is a normal matrix on x e D then L§(x) = Li(x) [10] with ||Li(x)|| = l implying that Li(x) has no singularities on D [14]. 34 after applying L'Hopital's rule. If by hypothesis, Ai(x), Ak(x) have different slopes at x = O, the denominator of (3.3-4) is non-zero at O and the limit (3.3-5) exists.1 However, (3.3-5) is an eigenvector only if its numerator is also non- zero. To study the behavior of the numerator of (3.3—5) let P = P(O) be the total projector corresponding to Ai(0) = Ak(0). Then (3.3-15) may be written lim Li(x)g(0) = lim d ( )- ( ) 3(0) (3.3-6) +0 +0 . d x x 3%] x _ Hgk x since g(0) = Pg(0). Now suppose that g(0) is an eigenvector of ” _ dL(O) . . . dA.(0) 2 corresponding to its eigenvalue ail (See Theorem (2.l-21)). Then d d I d o P 3%”) Pu(0) = aé-l(0)u(0) = 3%”) Pu(0) (3.3-7) So the numerator of (3.3-6) is non-zero since otherwise 1Note also that the matrix L. (x ) has only rank one in the limit as x+D since A. (x ) is non- repeated in the deleted neighborhood of x = 0. An eigenvector of L may be selected from the eigenspace of A. (0) since the range of P is precisely that subspace. 35 ppm) 139(0) = %%(0) P_u_(0) (3.3-8) and P gh—(O) Pg(0) = 3353(0) Pgm) (3.3-9) would contradict (3.3-7). If g(0) is an eigenvector of L corresponding to (0) a similar analysis shows that the numerator of AQQ boxy .1-16) does vanish. So if g(0) contains a component which is an eigenvector corresponding to %§i(0), (3.3-5) is an analytic eigenvector passing through the exceptional point x = 0 which corresponds to Ai(x). An analytic eigenvector cor- responding to Ak(x) can be constructed similarly. As the final t0pic of this chapter, we consider the question of differentiability of eigensystems which are functions of several parameters. As we shall see, the Frechet differentia- bility of such eigensystems may not be assumed for repeated eigenvalues even if they are analytic in a single variable. 36 3.4 Differentiation with Respect to Several Parameters The preceding sections concerned differentiability of eigen- systems with respect to a single variable. The situation is more complicated when several variables are involved. Consider the following example [14]: Let L(x],x2) = , (x],x2) e 3.2. (3.4-l) Even though L(x],x2) is symmetric and Frechet differentiable, 2 2)l/2 its eigenvalues A1 2 = :(x1 + x2 are only Gateaux differentiable when x1 = x2 = 0 [20]. Using Definition (2.2-l9) the directional derivative of A1 when x1 = x2 = O is determined as ((X 'I' th )2 + (X + th )2)I/2 I (X2 + X2)]/2 I I 2 2 l 2 D(A1 2.11) - Ilm t ’ t+0 - :_(h( + h§)I/2. where n = (h,,n2). (3.4-2) Since D(A1 2,h) exists for all h, A1 2 are Gateaux differentiable, but they are not Frechet differentiable since D(A h) is not 1,2’ linear in h. 37 This complication has serious consequences if these deri- vatives are to be used in non-linear programming routines which use gradients to determine a descent direction. Since the defini- tion of the gradient of a function depends crucially on the linearity of the Frechet differential. Recalling Definition (2.2-20), the Frechet differential of f may be represented in terms of the gradient of f as D(f,h) = (Vf,h). (3.4-3) where (-,-) is the inner product. It is this equivalence which provides the justification for using the negative gradient as the direction of steepest descent since if |Ivf|| = l and ||h|| = l D(f,h) = (vf,h) attains its minimum value for h* = - vf. When the Gateaux derivative is non-linear in the direction h, the descent direction is not so easily determined [21, 27]. We may also use Theorem (3.2-3) to compute the directional derivative for A1 or A2. Since L(x],x2) is Frechet differentiable we may represent D(L,h) in terms of its gradient VL as 38 Then D(L,h) = to which Theorem (3.2-3) may be applied since L(t) is an ana- lytic symmetric function of t. The directional derivatives of the eigenvalues A1 2 at x1 = x2 = O are the eigenvalues of L = E_D (L,h) P = P %%-P = D(L,h) SO _ 2 2 1/2 D(A12,h) - :(h1 + h2) as before. Remark (3.4-4): If D(L,h) is a symmetric matrix and if B is any matrix representing an orthonormal basis for 99(P) then the directional derivatives are also the eigenvalues of T“ T 8 L8 = B PD(L,h)PB = BT D(L,h)B (3 4-5) which is a formulation given by Haug and Rousselet [20] which may be easier to apply than Theorem (3.2-3). The non-linear dependance of the directional derivatives of eigenvalues can perhaps best understood by writing L as 39 where although the matrix L is linear in hi its eigenvalues cannot expected to be, unless .9P(P) has dimension one (if the eigen- value is simple). The rank one constituent idempotent matrices and eigenvectors or L(x1,x2) are even 'less' regular than the eigenvalues. To show this we construct an expansion of L(x],x2) in terms of its con- stituent idempotent matrices Li(x],x2) (See Theorem (2.l-33).) L(x],x2) = A](x],x2) L](x],x2) + A2(x],x2) L2(x],x2) (3.4-6) where L - A2 L - A] L1 = [ifjriigl , L2 = A2 _ A] . (3.4—7) or L (x x ) 1 X1 + (XI+X2)1/2 X2 1 1’ 2 g(x$ + x§)I/2 x2 -X] + (x$+x§)]/2 (3.4—8) 1 Xi ' (XI+X§)]/2 X2 L2(x1.X2) 2(X§ + x91/2 x2 _x] _ (x§+x§)i/2 (3.4-9) 4O However, the idempotents Li(x],x2) possess no limit as (x],x2)+(0,0) since “3...: limo L](O,x2) 2 1 1 whereas lim L1(x1,0) = (3.4-ll) x +0 1 O 0 and similarly N|-—* limo L2(0,x2) 2 -l 1 whereas lim L,(x ,0) X1+O 2 1 This illustrates that although in this example the rank gpg_ constituent idempotent matrices and therefore the eigenvectors can be continued smoothly through the origin along the x] or x2 axis, their total limits do not exist at the origin. Con— sequently, the rank one constituent idempotent matrices and eigen- vectors at not even continuous at x1 = x2 = 0. They do have directional derivatives since L(t) is symmetric and analytic in t (see Theorem (4.1-4)). 41 (3.4-10) (3.4-12) (3.4-13) 3.5 Conclusions The theorems of this chapter provide sufficient conditions needed to justify the differentiation of eigenvalues and eigen- vectors in a number of situations important to applications. In the next chapter formulations for the actual eigensystem deri- vatives are developed. 42 Chapter 4 Differentiation of Eigensystems As the remarks of the previous chapter illustrate, the question of eigenvalue and eigenvector differentiability in its full generality is difficult. Some matrices however, have properties which lead to relatively simple eigensystem deriva- tives. 4.1 Differentiation of Fully Analytic Matrices Definition (4.1-1): Let the diagonable matrix-valued function L(x) have the decomposition ' n L(x) = ZAi(x)ui(x)vT(x) = z Ai(x)L1.(x) x c 0 gig (4.1-2) -— -n i=1 i=1 where the g4(x) and 34(x) belong to an analytic biorthogonal set of eigenvectors such that 14(x)uj(x)=6ij’ Ai(x) are corresponding analytic eigenvalues and the Li(x) are rank one constituent idempotent matrices. Then we say, in this thesis, that L(x) is fully analytic on D. Remark (4.1-3): Theorem (3.1-l) gives sufficient conditions for full analyticity of L(x) on D if L(x) has no repeated eigen- values for any x e D. 43 The following theorems form a basis for a study of eigensystem derivatives of fully analytic matrices and methods to compute them. The first of these (Theorem (4.1-4)) states a sufficient condition for full analyticity which is important in applications. The theorem was originally stated by Rellich [16] but the proof here follows Kato [14]. (See also [14] for extensions to normal matrices.) Theorem (4.1-4) [14, 16]: If L(x) is a Hermitian analytic matrix- valued function for x e D C_i .9? then L(x) is fully analytic on D. Proof: If L(x) is Hermitian on D then the L(i) in Theorem (3.2-2) are Hermitian and therefore diagonable. The reduction process can then be continued indefinitely.D The next two theorems (Theorems (4.1-5 and 6)) present various algebraic properties of eigensystems derivatives. Part (e) is essentially the result given by Jacobi [4] although the derivation is different. Theorem (4.1-5): Let L(x) be an nxn fully analytic matrix-valued function for x e D. Then ll Tl dL _ dL. dA. a> o‘z'ziiaiz”z;fi”i i=1 44 C) ggflj: 48:5 d) I %%-gj = (AJ Ai)!I ggj = (A1 - Aj) %%4 95’ i f j 9’ VI%ui=%i where dependence on x is understood. Proof: To prove part (a) apply the product rule of differentiation to equation (4.1-2). Part (b) follows after differentiating L? = Li' For part (c), differentiate v§(x)uj(x) = 6ij' Part (d) and (e) are proved by forming the product T dL _ dA T T dL -"-i 632% ‘ 2314‘ liLkij + 24k ii as?" 13' and applying (b) and (C).D Theorem (4.1-6): Let L(x) be an nxn analytic real matrix-valued function for x e D g_ 8? such that LT(x) = L(x). Then l’l ll dL _ dL. dA. 9) 37-2 A1371+Za§iLi i=1 i=1 45 a?” 93 7 24 dx d) TdL =(A-A)uT99—=(A A)-d—u-Iu —i 6% J i—i dx 1 j dx —j e) 1i 2179—1 =33? Proof: Similar to (4.1-5) after noting that Theorem (4.1-4) guarantees that L(x) is fully analytic.o Remark (4.1-7): In part (c) if i = j we have g; 3&4 = 0, showing that 3i is orthogonal to ggi. This results from the constant norm condition imposed by gg(x)gj(x) = 6ij' Note also that L(x):9P:>3ann. Further insight into the behavior of eigensystem derivatives may be obtained by studying the effect of changing an individual element of a matrix. In the following theorem we will assume that L(O) is a constant matrix such that L(x) = L(O) + eij x is fully analytic for x e D. Theorem (4.1-8): Let L(x) = L(O) + eijx be a fully analytic matrix-valued function for x e D. Let Ai(x) be an analytic eigenvalue of L(x), let 34(x) and g (x) belong to an analytic l 46 set of left and right eigenvectors such that g}(x)pj(x) = 6ij and let Li(x) = gi(x) 31(x) be the corresponding consti- tuent idempotent matrix. Then T dL = dA ___ = a) it 3734. a‘xi‘ lki Bio Lka‘i T dL _ T du = b) is $541; ' (At ' As) is aszt lei y—tj Proof: To prove part a) apply Theorem (3.2-4) e) with g%(x) = eij and note that xki Ekj is the jith entry in Lk' For part b) apply Theorem (3.2-4) d). Part c) follows from part a) and Theorem (2.1-32). Since the sum of eigenvalue deriva- tives is a constant function of x, the sum of second deri- vatives with respect to x is zero proving part d).a Remark (4.1-9): Since the non-zero columns of the idempotent matrices Li are eigenvectors corresponding to A1, part (a) above es- tablishes a strong relationship between an eigenvalue's deri- vatives and its eigenvectors. 47 The following Theorem (4.1-10) states a formulation, given by Fox and Kapoor [5] (See also [12]), which expresses the deri- vative of an eigenvector of a symmetric matrix as a linear com- bination of eigenvectors. Theorem (4.1-l0) [5]: Let L(x) be an analytic real matrix-valued function for x e 05; 3? such that LT(x) = L(x) Then n Elfin) = Z] CjIijIX) (4.1-ll) J with < ) 0 1: J C. X = J (xxx) - A360)“ ujm $359460 = gJT-(X) 334m 1 7‘J . d . . . . . Proof: Since agfl is a vector in an n dimenSional space, we can express it as a linear combination of n orthonormal eigen- vectors of L(x) Tl %%i(x) = 2 ck(x)9k(x) k=1 —j ij j a;— = 0 (See Remark (4.1-6), we have 48 TdU-_ T : Tg_U_'= : ° Ej ail — cj Ed Ed cj and g4 dxi Ci 0, applying Theorem (4.1-5) (d) completes the proof.D Although this formulation has theoretical value it is not useful when a complete set of eigenvectors is not available [9]. However, we may rewrite (4.1-ll) as n du. _ -l T dL 6)? ‘ '2. (ii ‘ ii) -”-i 1.13741 i=1 n = ' IE: [Iii ‘ *i)-]Lj] gF'Ei = ‘ L: 8%”!i (4'1'12) i=1 The matrix T] L: = - jg; [(Aj - Ai)-]Lj] j=l as we shall see, in Section 4.3, can be easily computed and will lead to an efficient means to compute ggi. 4.2 Eigensystem Derivatives with Generalized Inverse If L(x) is a diagonable fully analytic nxn matrix-valued function in some neighborhood of x = 0, then we may write 49 [L(x) — A(x)I]u(x) = (2 x1(L(1) - x(1)1))(2 x1u(1)) = o by expanding L(x), A(x) and u(x) in a Taylor series about x = 0. Collecting terms in x1 m i Z1 x Z (L (111 - A(1'11I)u(~11= o. (4.2—2) i=0 j=0 Each coefficient of x1 must vanish separately so i c o 1 o c c 2(LI1'31- A(1'311)u(11= o i = O,l,2,... (4.2-3) i=0 Ol‘ 2LI111uI =2A11‘1)u( i O,l,2,... (4.2-4) After subtracting the i = j term from (4.2-3) the eigensystem derivatives are given recursively by (L - AI)u = o, i = o (4.2—5) . 1-] . o o o c (L - AI)u(1) = - (L(1'11- A(1’J)I)u(31, i = l,2,3,... i=0 (4.2-6) 50 Since we have assumed that L(x) is fully analytic, these equations must be normally solvable (See Theorem (2.2-16)) for each i. Hence if v is any vector such that v c.AV(LT - AI), then 1:] c c o o o vT :E: (L(l-J) _ A(i-J)I)U(J) z 0 (4.2-7) J=0 follows from Theorem (2.2-l6). Solving for A111 results in A(1) _ vT L(1) u _ __________ 4.2-8) VTU ( and i-l 1_] VT<2 L(i-j)u(.i) _ A(i-i>u(i)) A(i) = J=0 T 3‘1 , i = 2,3,4,... V U where v is any vector such that vTu f 0 and v euV(LT - AI) and u = u(O) lies on u(x), an analytic eigenvector. We now turn our attention to solving equations (4.2-6) for u(1). Since the matrix LA = L - AI has rank n - m we use a generalized inverse L; and apply Theorem (2.1-20) to find that i-l ”(1): _ Li Z (L(i-j) _ A(1-j)I)U(j) + Z, (4.2_‘|0) j=0 51 where z 530(LA) and Li is any rank n - m matrix satisfying I _ LALALA - LA' We shall have immediate use for the following: Lemma (4.2-ll): Let A, AR and AL be nxn matrices satisfying AA = 0 and A A = 0. Then if AI R L of A, is any generalized inverse (4.2-12) is also a generalized inverse of A. Proof. Multiplying (4.2-12) on both sides by A results in # AA A = A(I - A 1 )A1(I - A )A = AA A = A.D R L Now suppose P(O) = P is the total projector associated with A then since L is diagonable. From Lemma (4.2-ll) if L: is any generalized inverse of LA then >2 A H l 'U v A 43 P0 . -l3) 52 is also a generalized inverse of Lx. Furthermore, since Pu = u T T and v P = v , L: satisfies Liu = 0 (4.2-14) and VTL: = o . (4.2-15) If 1-] o u u o 0 5(1) z _ L: :2: (L(l-J) A(l-J)I)UIJ), 6(0) _ u j=0 and u(1) = 9(1) + c.u, i = l,2,3,... (4.2-16) then Theorem (2.1-20) guarantees that u(11 is a solution of (4.2-6) for any constants c1 since ciu euV(LA). We may use the properties (4.2-14 and 15) to simplify equations (4.2-8 and 9). If c1 = 0 for all i, then equations (4.2-8 and 9) become - T (4.2-17) 53 where u is the value of an analytic eigenvector at x = 0 and v is any vector satisfying v euW(LT - AI) and vTu # 0. If the eigenvector derivatives are to be used explicitly, then the constants Ci should be selected to satisfy some normal- ization conditions such as x) u(x) = 1 (4.2-18) and (u(x))* u(x) = 1. (4.2-19) If we assume (4.2-l8), we obtain v u = v u + civ u (4.2-20) also T follows from (4.2.16) since v L: = 0. Now suppose u is also normalized so that (4.2-l9) holds. Then (u + xu(1) + x2u(2) + ...)*(u + xu(1) + x2u12) + ... ) - l 54 must be true and the coefficients of x1, i :_1, must be zero so 2 (u(1'J))*u(J) _ O, 1:] j=0 or 1-1 (u(0))*u(1) + (u(1))*U(O) = _ (”(1-3))*UIJ) j=1 Equating real parts of both sides Re[u*u(1)] = - %-Re[:E: (u(1-1))*u(j)] u*u - u*fl + ciu*u and Ci = u*u(1) _ u“1(1). Then, Re[c]] = - Re[u*fi(1)] and 55 (4.2-21) i-1 Re[ci] = - %—Re[ :2: (u(1'j))*u(j)] - Re[u*d(1)], i 3_2 (4.2-22) j=1 must be satisfied for (4.2-18 and 19) to hold. Remark (4.2-23): If L(x) is a normal matrix then u*u(1) = 0 since u* = vT. If L(x) is a real symmetric matrix then uT = VT, c1 = 0 and 111 > ( > _ l (i-J T J C, - - 2—[ (u ) u 1. i 3_2 To determine u in the event that A = A(D) has multiplicity m in (4.2-l) we note that since v in (4.2-13) may be selected as any vector in the m-dimensional null space of LT(0) - A(0)I and since 1( L(x) is fully analytic, JV(L 0) - A(0)I) is spanned by the values at x = O of m independent analytic vectors, which we may take as the rows of the m x n matrix VT(0) = VT. Then from (4.2-6) 1.-.! c o 0 o o vT (L(1'11 - A(1'111)u(11 = 0 (4.2-23) i=0 0?“ i-l . . . i-l v1 IE: LII-J)u(J) z VT :5: A(i-J)U(J) (4.2_24) i=0 i=0 56 Similarly we form an n x m matrix U(0) = U where the columns of U are derived from the analytic right eigenvectors of A and i-l i-l v1 2 L(i-J’)U(.i) = VT 2 Um A(i-i) z A(i) (4.2-25) 3:0 1:0 In particular, if i = 1, VTL(1)U = VTPL(1)PU = v1u A(11 = A(11 (4.2-26) where A(11 is an m x m diagonal matrix consisting of the m first derivatives of A and the columns of V and U are normalized so that VTU = I. Equation (4.2-26) shows that V and U must diagonalize the matrices L(11 and L111 = PL(1)P. Furthermore, if A(l) has distinct values on its diagonal then V and U are unique. We now turn to an example. Example (4.2-27): Let The eigenvalues of L(x) are A12 = l :_x and 112) = :_l for all XlEig 1 Analytic eigenvectors of L(x) may be found from the non-zero columns of the rank one constituent idempotent matrices L1 and L2. Using Theorem (2.1-33) we arrive at 57 J. I.- l -11 12 2 () 2 2 L (x) = L x = 1 41 1_ 2 _i 1 L2" 2 L 2' '2 Then [./2 ./2 () 7 (1 7 U x = u x = ' 52 b7... are the analytic unit eigenvectors of L(x) and they are the unique unit vectors that form the columns in the matrix S which diagonalizes ./2 /2 - i ./2 ./2 i /2 /2 ./2 ./2 L2 ”'21 .1 0. f2- '7] -0 4. Furthermore if v is any vector such that vTu1 # 0 or vTuZ # 0 and v e .4/(L1 - AI) then 0 l o l vT u1 = l or vT u2 = -1 l 0 l i T T 58 Since if vT = (x,y) then (x,y) 0 l 13-1 ./2 ./2 /2 x ——-+ y —- l 0 7 2 2 d: :1 /2 ./2 ./2 ./2 7+y7 X7+Y7 OY‘ (x,y) 0 l %- / /" _/2 -x —%—+ y —%— l 0 ‘2' = =.." /2 1% £_ :2 X z'y 2 X 2 Y 2 and we have agreement with (4.2-l7) for i = 1. Note that it is essential that u (or v) lie on the trajectory of an analytic eigen- vector for this formulation to work. 4.3 Numerical Approach In numerical work the algorithm (inverse iteration) [22], (L - AI)ui+] = V, (4.3-1) Vi = ui/lluilloo (4-3-2) 59 is often used to determine an eigenvector u corresponding to the eigenvalue A where A = A + 8. Solving (4.3-1) we have = (L - AI)'1v. 1 (4.3-3) ui+i In practice, the matrix (L - AI) is given an LU factorization and the iteration (4.3-l) is carried out as a series of back sub- stitution stages. Since each back substitution stage only in- volves solving a triangular linear system the LU decomposition stage is the major effort in the computation [23]. These practical considerations provide a strong incentive to study the structure of (L - AI)"1 with the hope that it can be utilized in constructing LI. If L = L(0) is a diagonable matrix with 5 distinct eigen- values then 5 ~ -1 _ ~ -1 J'=1 s ” -1 _ -1 ” -l (L - Ail) - - 6 L1 i- )E: (AJ - A1) Lj (4.3-4) J=1 J'i‘i follows from Theorem (2.1-32). Now consider 6O . -1 . -1 113(1‘ Ll)“- - 411) (1' L1) = 2:31:62 (1' £14.,“ " Li) S ~ -1 + (Aj — x,) [j]. (4.3-5) 1 LIL]. ';i The first term of the right side of (4.3-5) vanishes since (I - L4) Li (I — Li) = 0 is a constant and we have L1 = (1 — L ) (L — A 1)‘1 (1 — L ) Ai -i i -4 s -l = - . - . . 4.3-6 IE: (AJ A,) La ( ) J=1 if] Furthermore, Li is a generalized inverse of LA = L - AiI since i i (L - AiI) (1 - Li)(L — xi(1)'1(1 - Lfi)(L - x11) = L - AiI (4.3-7) Finally since PiIO)’ the total projector of the Ai - group, is equal to Li we have (dropping the subscript i), L: = (1 - P)(L - AI)'1(I - P) (4.3-8) From Lemma (4.2-11) however, (I - P)L (I - P) = L: (4.3-9) I A 61 is also a generalized inverse of LA but since I - P is idempotent L: = Li. Therefore we write i _ -1 LA - (I - P)(L - AI) (I - P) (4.3-10) and note that L: satisfies conditions (4.2-14) and (4.2-15), namely This form of L: is especially convenient since (L - AI)’1 may be computed previously to determine the eigenvector u and since the idempotent matrix I - P is easily multiplied by a vector particularly if P has only rank one, i.e., (I - uvT)b = b - (va)u (4.3-ll) OY‘ 61(1 - uvT) = 61 - (bTu)vT. (4.3-12) In practice, instead of directly inverting (L - AI)'1, an LU decomposition of L - AI is formed. The following has proved to be an efficient procedure if A is non-repeated. 62 Form an LU decomposition of L - AI. Use the LU decomposition in an inverse iteration scheme to find an eigenvector u. T) dL (I - uv ax'“ . Solve for z in LLz Solve for y in Luy 1) Z . y = y - (va)u + c1u iii-(I -UV where Lu and LL are upper and lower triangular matrices computed in the LU factorization of L - AI. If A is a repeated eigenvalue with multiplicity m then it may be necessary to resort to a technique such as the method of diagonalizing PL(1)P given in Section 4.2 to find an analytic eigenvector u, before proceeding with, 3. . dL Solve for z in LLz (I - P) 32‘” Solve for y in Luy II N 0.9.. X: = (I - P)y + clu. 63 Chapter 5 Design Sensitivity in Finite Element Models of Dynamical Structures The finite element method is used extensively for solving boundary value problems involving partial differential operators. Its success depends largely on its ability to deal effectively with complicated domains and boundary conditions. Its major disadvantage is that it often requires large amounts of computer time to find solutions. It makes sense then to develop efficient methods to update existing solutions when small design changes are made in order to avoid a complete re-analysis. In this chapter, methods for computing a linear dynamical system's sensitivity to design changes are discussed and two illustrative examples are presented. The methods described are directly applicable to solutions generated by the finite element method [1, 24-26] or similar methods. To begin we shall outline how the finite element method can be derived for a static homo- geneous boundary value problem in the plane. 5.1 The Finite Element Method Consider the boundary value problem Lu = f in 0, (5.1-l) u = 0 on an 64 where L: H1(Q) + L2(Q) is the 2nd order partial differential operator: = _ 3 9E. - 2.. 9!. Lu Sin ax] 3y [q 8y] + cu, 1 : 'flz 9—1.1: H (0) {u c L2(Q). ax uX and By uy c L2(Q)}. p9Q:C and f e L2(Q), and 9 is a connected, closed, and bounded domain in 922 with boundary 39. Now suppose that we define the following inner product on H1(n) to be (u,v)Hl =j;(uxvx + uyvy + uv) (5.1-2) for any u, v e H1(n), and consider the quantity A (LU.v) = (u,v)L with v = O on 39. If (u,v)L is integrated by parts over the domain 0, an application of Green's identities yields (u,v)L .1; (puxvx + quyvy + cuv) + J£;(-puxv + quyv). .1; (puxvx + quVy + cuv). (5.1-4) 65 It can be shown that [25] if p,q and c are strictly positive and continuous then positive constants a and B exist such that OLLIUNHUZ 1 [(u.u)L]1/2 : 8[(U.U)]1/2 (5.1-s) and so (u,u)L = 0 if and only if u = 0. Since (u,v)L = (v,u)L (symmetric), and (oLu1 + BU1,V)L = a(u],v)L + B(u2,v) (linear) we have established that (u,v)L is an inner product on H;(n) = {u e H1(n): u = 0 on an}. Since (u,v)L is an inner product on H;(Q) we may now apply the Projection Theorem (2.2-20) to find a numerical solution for u. To find an approximate solution to 5.1.1 let M C H; be spanned by the Hamel basis 8 = 1¢l’¢2""’¢m1’ then m u = :5: 21¢, (5.1-8) is a vector in M C Hg. We seek G such that ||LG - f|| 5_ ||Lfi - f|| , for all 8 e M By Theorem (2.2-20), 6 must satisfy 66 (u. 4.), = (i.ej) J Tl (12::1 2,4,. oJlL = (1243-) ll 12;] (4,. 63.)sz = (f,¢j) for all oj e B (5.1-9) which we write in matrix form as Kz II H‘ (5.1-10) where kij = (¢i’¢j)L' The functions ¢i e H$(Q) may be constructed by triangulating the domain a [25] as shown in Figure (5.1-ll). We then require ¢i(x,y) to satisfy where "j = (xj’yj) is a vertex in the triangulation of n. This choice of the 1i will tend to cause K to have a banded structure 67 since K is populated by inner products (¢i’ ¢j1L which are non- zero only if "i and nj are sufficiently close in the triangulation. Since the matrix K was generated by using the Projection Theorem (2.2-20) with the inner product (-,-)L on the space Hg(9), we would, in general expect K to change if the domain 9 were to change. Thus, it is natural to consider the variable finite ele- ment problem K(9)z(9) = f(9) (5.1-12) where 9 is now considered to be a variable domain. Suppose T(9) = 9', T: 99:2 +592 (5.1-13) is a mapping operating on the domain 9 such that 9' = 9(6) = 9(0) + on, o e D (5.1-l4) where A<:992 is a connected, closed and bounded set whose points are such that every point in A corresponds to some point in 9(0) (2 392 where the addition operation in (5.1-l3) is vector addition of the points in 9(0) to the corresponding points in A and where T is a topological homeomorphism for a e 0. Then if 9(0) is given a triangulation, we may construct a matrix K(a) such that 69 and the variable problem K(a)x(a) = f(a) results. We now define the directional derivative of K(a) as DK(9(0),A) = lim “9(0) + 0‘4) ' W401) (5.1-15) a 9+0 (If this limit exists for any 9 as defined above then K(a) is Gateaux differentiable and DK(9(O),A) is its Gateaux derivative.) 5.2 Dynamical Problems The distributed parameter forced vibration problem (without damping) can be written as 2 L1 Lg-(xgc) + L2u(x,t) = f(x,t) (5.2-1) 3t B(u) = g(x,t) on 39x1 (5.2-2) C(u) = r(x,0) in 9 (5.2-3) where L1 and L2 are self adjoint positive definite operators on 9, u is taken from a function space V(9xT) defined on the product space 9xT derived from the spacial domain 9 (k = 2,3) and T is the temporal domain [O,t]. In addition, boundary conditions 70 (5.2-2) defined on a9xT, where 39 is the spatial boundary of 9, may be given along with initial conditions (5.2-3). When the finite element method is used to obtain an approxi- mate solution to (5.2-1, 2 and 3), it is usual to arrive at a lumped parameter model of the form M'z'(t) + Kz(t) = f(t) (5.2-4) where M is an n x n matrix K is an n x n matrix z(t) is an n-vector consisting of n = km time varying coefficients of the spline basis functions 11 e B related to the triangulation of 9 and f(t) is an n vector resulting from projecting f(x,t) onto the finite subspace spanned by B. We are mainly concerned with the variable problem M(e)'z'(a,t) + K(o()z(a,t) = f(a,t) (5.2-5) and the related generalized eigenvalue problem [28] [K(a) — A(a) M(a)]u(a) = 0 (5.2-6) where 6 accounts for a changing spacial domain 0(a) = 9(0) + 9A 71 This problem can be related to the symmetric eigenvalue problem in standard form through the use of the following transformation [8], u = M‘1/2v, (5.2-7) 1/2 where M' is the positive definite square root of M'1 and dependence on a is understood. Then M'I/ZKM_1/2v = Av (5-2-8) Since W”2 is positive definite and since K and M are analytic functions of a e D, W”2 and therefore M'1/2KM'1/2 are analytic on D. Thus (5.2-6) is equivalent to the variable eigenvalue problem in standard form of the previous section and the eigen- values Ai(a) and particular corresponding eigenvectors vi(a) i = l, n, are analytic for a e D. It is not convenient to determine the derivatives of Ai(a) and Vi(a) from (5.2-8) since the transformation M'1/2(a) may be difficult to compute and the transformed matrix in (5.2-8) is usually difficult to differentiate. Instead (5.2-6) is used. After differentiating and rearranging (5.2-6) we arrive at due- 91:. 9’1-“ - [K-AMIE; [dd Au a;M]u (5.2 9) 72 Solving for gE-with a generalized inverse Di, we obtain du _ I dK dM dA d'oT"DA[doT' *T‘aa‘mu” where DA=(K-AM) and DI-(K MI . A - - A ) satisfy 0 DID =9 and z e uV(K - AM). We may assume that analytic eigen- vectors exist such that 1 .i' 1.1 we have 'T 1/2 1/2 = T _ ui M M ”j ”i M ”j - aij‘ Differentiating (5.2-12) gives 73 (5.2-10) (5.2-11) (5.2-12) (5.2-13) (5.2-14) Now suppose u satisfies (5.2-12). Then from Lemma (4.2-ll) o: = (1 - uuTM)D:(I - MuuT) (5.2-15) is also a generalized inverse of DA since DA uuTM = MuuT DA = 0. (5.2-16) Substituting 0: in (5.2-10) with z = cu gives = - D:[%§-- Agg]u + cu (5.2-l7) Multiplying (5.2-l7) by uT M and applying (5.2-12) shows that _ l d C - - §1 U a“; U (5.2-18) Again we must determine 0: to complete the formulation. The approach is similar to the one used in the standard eigenvalue problem. Suppose S is a matrix such that ST[K - AM]S = A - AI (5.2-l9) 74 Then we may expand K - AM as [29] n K - AM = S(A - AI)ST = jg: (A1 - A)Gi (5.2—20) i=1 where U.U.T _ i 1 G1 - T u.Mu l and GiMGj = GijGi‘ (5.2-22) T1 (K - AM)'1 = IE: (A, - i)'1e. (5.2-23) If PG is the sum of the m matrices Gk associated with eigenvalue A of multiplicity m then D: = (I - PGM)(K - AM)‘1(I - MPG) (5.2-24) is a generalized inverse of K - AM such that DIMu = 0 (5.2-25) uTMD: = 0. (5.2-26) 75 If u is determined through the inverse iteration algorithm [26] (K - AM)u = Mv i+1 i (5.2-27) < ll uj/IIU-iII0° then the algorithm given in (4.3) is essentially unchanged. The eigenvalue derivative may be found by multiplying (5.2-9) by vT, v e 34(K - AM) and solving for dA/da dx v1 (§§-- x§§)u T a...: T , where v Mu f 0. (5-2-28) “ v Mu The system (5.1-4) can be modified to include dissipative terms by adding the term C2 to the left side [28]. We then have the damped system n(a)2(u,t) + Ci(a,t) + Kz(o,t) = f(a,t) (5.2-29) and the related generalized eigenvalue problem a)M(a) + A(a)C(a) + K(a))u(a) = 0 (5.2-30) where the positive semi-definite matrix C accounts for viscous damping. 76 The 'damped' eigensystem of equation (5.2-30) is usually more difficult to solve than the corresponding system (5.2-6) without damping. However, under special circumstances the eigen- vectors of the damped system are also eigenvectors of the cor- responding undamped system. This is true, for example, if C is some linear combination of K and M, i.e., C = 8K + yM, 8,y e32 (5.2-31) It has been shown that [30] KM'1c = CM'1K (5.2-32) is a necessary and sufficient condition for the eigensystem of (5.2-30) to be uncoupled when transformed to the modal coordinates of the corresponding eigensystem of (5.2-6). If (5.2-32) holds for a e D 999 then the eigenvectors of the damped system do not depend on C(u) and their derivatives may be computed directly from (5.2-l7). The eigenvalue derivatives may be determined by differentiating (5.2-30) and solving for dA/do to obtain T 2 dM dC dK (A ETC-1.1+ 1138-1113;)” vTCu + 2AvTMu V (5.2-33) 0.10. 9 >2 ll 77 Remark: Some caution must be taken when using this formula since the eigenvalues are not differentiable for particular values of the matrix C. For example, suppose u is an eigenvector of (5.2-30) then AZUTMU + AuTCu + uTKu = 0 (5.2-34) is satisfied by -uTCu :_[(uTCu)2 — 4uTMu uTKu]1/2 A = T ’ (5.2-35) 2u Mu If (uTCu)2 = 4uTMu uTKu (5.2-36) then the mode corresponding to u is said to be critically damped [31] and the eigenvalue A is not analytic. This can be verified by attempting to differentiate (5.2-35). 5.3 Examples This section is devoted to two design sensitivity examples involving finite element vibration models. In the first example we will use the methods presented in Chapters 4 and 5 to compute a Taylor series for the lowest natural frequency of a triangular plane elastic element. The design variation consists of a changing node (vertex) position. The 78 second example involves a fixed-fixed plate assembled from several plane elastic triangular elements which is undergoing boundary shape variations. Example (5.3-l): tqrf Figure (5.3-2). the stiffness matrix is given as a function of x by K(x) = 4 0 2(x-l) 4 0 x2-2x+2 symmetric -2 —2(x+1) 2 2(x+l) 0 -2(x+1) -(x-1) -x2 x-l x -2x+2 x+l -x x2+2x+2 -(x+l) x2+2x+2 79 _1 Consider the plane elastic element shown in Following standard procedures and notation [1,2], (5.3—3) (XZ’yZ) (x3’y3) ("190) (190) Figure 5.3-2 Triangular Element (shown in three design positions) 80 where hld t is the plate thickness, b1 0 _ l ‘11 b1 a] = x3 - x1 2 a2 = x] - x3 x1 - 1 a3 = x2 - x1 -x1 - 1 2A = b1x1 + b2x2 + b3x3 = 2 V = At. (1) z d_K _ K dx - 0 0 2 0 0 0 2 2x-l -1 2x+2 symmetric 81 and y2'5’3 y3'yi =3’1‘3’2 -2x 2x+2 (5.3-4) I0 0 0 0 o 0 o o o o o (2) _ 1 de _ t 2 o -2 0 K 779 dx 2 0 -2 2 o symmetric ' 2 K(11=o i>3 For the mass matrix we use which is a lumped mass formulation [2]. Since the area A is a constant If the plate thickness t is set equal to 3, M = I and D (x) = K(x) - A(x)I . A The smallest non-zero eigenvalue of K(O) is determined as 82 (5.3-5) and its corresponding unit eigenvector is We may now use (4.2-16 and 17) to determine a Taylor series for A(x). For L: = D: we use (4.2-13), i.e. o: = ([1 - uu1]) (K - i1)'1 (1 - uu1) to find that 83 The first 10 Taylor series coefficients to single precision accuracy are A(11 = 0.0 A(Z) = -.500000 A(31 = 0.0 A(4) = .138890 A15) = 0.0 A(6) = .001543 A(7) = 0.0 A18) = -.052641 A(91 = 0.0 AI10): .061548 With these coefficients, A(x) may be approximated as n A(x) = 2 x1AI1). i=0 In Table (5.3e6a) a Taylor series representing A(x) is evaluated for x = 0 to x = 1.5 and for n = 2, 4, 6, 8, 10 and 20. In Table (5.3-6b) a direct numerical solution for A(x) is presented for the same values of x. Figure (5.3-7) presents these data graphically. A comparison between the values of A(x) given by the Taylor series and the values of A(x) obtained by direct evaluation show excellent agreement even when n = 2. 84 .o=_m>cmmwm mo cowpmzpm>m pomcwo mpm—mw.o mkppom.o mmmmmm.o mvmmoo.F Kmmmoo.F mmwmfip.— mmoomF.F cmommm._ __mmmm.P Nm¢mmm.P wvmmwm._ mmmmm¢.~ mmpom¢.F mmmom¢.P mpomm¢.~ ooooom._ Ax? 3V om.P o¢.~ om.F om.— op.P 00.? om.o ow.o o~.o om. om.o o¢.o om.o om.o o—.o 00.0 x mvmmmp.o¢ momom¢.o— mvamo.m omomo¢._ okump.p mwowmp.~ omomnp.— mmxmmm.~ mrmowm._ Fmvmmm._ mqmmmm._ Nmmmm¢.P mmpmm¢._ mmmow¢._ mpommv.F OOOOOm.— 8h: monmmm.m commmo.m mommmv.— m¢mm-.— mkmuvy._ mmmm¢—.F mexmm—.F cmommm.— NmNNmN.F mmmmmm.~ mmmmwm.P emmmm¢.F mmpom¢.— mmmom¢.F mpomm¢._ ooooom._ OP": n_¢mmN.Ou NmemN.o «mmmmm.o memowm.o vammm.o pmmnmo.p vwmeo_.P poemNN.F «memmm.p mm—nmm.— mmvmmm._ mmmmm¢.— mmpom¢._ mmmow¢.F mfiommv.y ooooom.— m n : mim.m mpnmp Fonmmo.F ompmoo._ mmpmmo.P mommno.F owopop._ Nm¢o¢P.— ewmomF.F mommmm.~ wmmmmm._ mmowmm.~ ¢ommmm._ mommmv.F mNFom¢.P mmmow¢.P vpomm¢.— ooooom.P m h c mmpwno.F mmmmmo.~ ommpmo.~ ooowoo.P mvmwmo.F mwwmm~.F mNme~.P mmmmmm.~ memwwm.— ooowmm.— omommm._ mmmmmv.F mmpmmv.F mmmow¢._ e—omme.p ooooom.P a n : .mscmp c mo mmwcmm Lopxmh cpwz mmpms_pmm mzfim>cmm_m momemm.o ammo—m.o mmmvmo.o oooowm.o ooommw.o oooooo.~ ooommo._ oooom—.F ooommm.~ oooomm.~ ooommm.— oooo~v.~ QQOmm¢._ oooom¢._ ooomm¢._ ooooom.~ N n c 3 om.— o¢.F om om F — op._ oo.P om.o om.o om.o on o om o o¢.o om.o om.o o_.o oo.o .m 85 :mwmwo co mocmvcmamo m:_m>cmmwm mim.m mczmvu m.F o.— m. m.i O.Fi m.—i — LI (1+ ” _ a w H: N N: COWPGJFQ>U pooewo 0.7... ‘ OUCQVHC . OoNIlI OF H : ON n : o.mI:i 86 The efficiency of the technique may make design sensitivity calculations of eigenvalues and eigenvectors feasible in many practical situations. For example, if an inverse iteration method is used to determine the eigenvector at x = 0 then subsequent determinations of each of the Taylor coefficients A(11 and u(11 involve little extra effort since the generalized inverse D: is easily determined from 4.2-13 (see Section 4.3). (The method is especially efficient if the matrix K(x) is only linear or quad~ ratic in x since in these cases most of the matrices K(11 = L111 in Equations (4.2-16 and 17) are null.) Since the Taylor coef- ficients can be obtained so economically, there is a strong in- centive to represent the design with a Taylor series. Then subsequent evaluations of the eigenvalue and eigenvector as the design is modified can be carried out with substantially less computational effort. The next example involves a finite element model assembled from thirty two plane elastic elements similar to the element of example (5.3-l). The example illustrates the results of a design sensitivity calculation for an eigenvalue depending on a vari- able boundary shape. Example (5.3-7): Figures (5.3-8) through (5.3-12) illustrate the sensitivity on the lowest eigenvalue of a dynamic finite model to changes of the boundary shape. Figure (5.3-8a) shows the baseline design. The lowest eigenvalue for the baseline design is calculated as 14.44. (Figure (5.3-8b) illustrates 87 the mode shape corresponding to this eigenvalue.) In Figures (5.3-9) through (5.3-12) the shape of the lower edge of the model is changed. In each case an estimate of the new eigenvalue is compared with the result of a finite element run for the new configuration. The estimates are based on a single first order directional derivative of the eigenvalue with respect to some direction h (see Definition (2.2-10)) which is related to the variable boundary. The direction h is determined by a gradient projection method as a direction (in the design space) which causes the eigen- value to increase without increasing the volume (mass) of the plate. (For a comprehensive discussion of this and similar techniques see [3].) Each subsequent variation in the design is determined by moving along the direction h an increased distance. A comparison between a direct evaluation of A and the first order estimate of A shows good agreement for moderate changes in the design but for extreme design changes the estimate is substantially higher than the direct evaluation. 88 cmwmmo wcwpwmmm mum.m wczmvm omqq.¢_ n msfim>zmmwm op mcwucoammccou mamzm one: ADV 4223:8491 ’> moo_a coxwa-ooxea to eo_pe_=memeee Aev 89 mmcmsu cmwmmo pmcwd m-m.m mczmwm mnmm.o~ o=_e>comem oopsanu oe¢o.o~ o=_e>eomem eoeeeeemm 9O mmcmno cmwmmo ucoomm opim.m mesmwd emem.e~ o=Pe>=om2m eooaaeoo eeee.m~ o=_e>eom_m eooee_omu 91 omemeo ememoo eeeee __-m.m seamen vomm.vm m¢¢m.~m mspm>cwmwm umpaqsou mzpm>cmmwm uwumawumm 92 mucosa :mwmwo sensed Npum.m mczmwm umpp.ee ospe>eomem 66336268 _mem.em o=_n>comem emceewomm 93 Chapter 6 Discussion and Further Study 6.1 Discussion Modal analysis has become a standard tool used in the design of dynamical systems. By numerically determining certain eigenvalues and eigenvectors associated with a linear structural model a designer can investigate how a particular structural design will perform dynamically. The eigenvalues are related to the natural frequencies of the structure and the eigenvectors indicate the associated vibratory modes. If the designer is concerned with dynamical performance, several redesigns and re-evaluations of the eigenvalues and eigenvectors may be necessary before a suitable design is found. Since this procedure is costly in computer time, there has been an incentive to find improved methods. This thesis presented a technique to represent an eigenvalue and eigenvector by a Taylor series in a design variable. Since the Taylor coefficients can be determined to arbitrary order, excellent estimates may be easily calculated for design changes which do not exceed the radius of convergence of the series. Since the coefficients of the Taylor series can be computed with substantially less effort than recomputing the eigenvalue or eigenvector, the procedure has the potential to reduce the number 94 of calculations necessary to carry out the design of dynamical structures. There is also the potential to use the method in the implementation of optimal design algorithms. Two representative example problems were presented. In one of them, twenty Taylor coefficients of a specified eigenvalue were determined after a single finite element run. These coef- ficients were used to represent the eigenvalue as a power series in the design variable. Comparison between separate finite element runs and the Taylor series showed excellent agreement. 6.2 Methods Each Taylor coefficient is determined by repeatedly evaluating a simple recursive formula which is derived through an application of generalized inverse theory. The procedure is simplified by using a generalized inverse matrix specially selected to annihilate certain terms in the formulation. The generalized inverse itself need not be formed explicitly if the eigenvector is determined through the inverse iteration method, with the result that each Taylor series coefficient is computed with no more than 0(n2) multiplications. The technique was extended to compute derivatives for the eigenvalues and eigenvectors of the generalized eigenvalue problem (K a AM)u = 0 and there was some discussion of the damped 2 eigenvalue problem (MA! + AC + K)u = 0. 95 6.3 Suggestions for Further Study There are several questions involving eigensystem derivatives which deserve further study. Necessary conditions for eigenvalue differentiability are in short supply. It may be possible to use Equation (4.2-25) to construct a sequence of necessary conditions for the full analy- ticity of diagonable matrices similar to the sufficient condition given by the reduction process. (Equation (4.2-26) amounts to such a necessary condition.) A convenient computational method for determining the radius of convergence for Taylor series representing eigenvalues and eigenvectors would be useful in applications. Lower bounds have been given [14] which unfortunately often severely underestimate the actual convergence radius. Most of the results of this thesis can be extended to eigen- value problems involving linear operators on Hilbert spaces. Ex- cellent references for continued study are Refs [10-14 and 19]. 96 BIBLIOGRAPHY [l] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] BIBLIOGRAPHY Zienkiewicz, O.C., The Finite Element Method, McGraw-Hill, New York, 1978. Desai, C.S. and Abel, J.F., Introduction to the Finite Element Method, Van Nostrand Reinhold Co., New York, 1972. Haug, E.J., and Arora, J.S., Applied Optimal Design, John Wiley & Sons, New York, 1976. Jacobi, C.G.J., Crelle's Journal Far Die Reine Und Angewand Mathematik, Vol. 30, Berlin: De Gruyter, 1846, pp. 51-95. Fox, R.L. and Kapoor, M.P., ”Rates of Change of Eigenvalues and Eigenvectors", AIAA Journal, Vol. 6, Dec. 1968, pp. 2426-2429. Plaut, R.H. and Huseyin, K., "Derivatives of Eigenvalues and Eigenvectors in Non-Self Adjoint Systems", AIAA Journal, Vol. II, Feb. 1973, pp. 250-251. Rogers, L.C., "Derivatives of Eigenvalues and Eigenvectors", AIAA Journal, Vol. 8, May 1970, pp. 943-944. Rudisill, C.S., "Derivatives of Eigenvalues and Eigenvectors for a General Matrix", AIAA Journal, Vol. 12, May 1974, pp. 721-722. Nelson, R.B., "Simplified Calculation of Eigenvector Derivatives", AIAA Journal, Vol. 14, Sept. 1976, pp. 1201-1205. Frame, J.S., "Matrix Functions and Applications, Part I", IEEE Spectrum, March, 1964. Frame, J.S., "Matrix Functions and Applications, Part II", IEEE Spectrum, April, 1964. Wilkinson, J.H., The Algebraic Eigenvalue Problem, Clarendon Press, London, 1965. Dunford, N. and Schwartz, Linear Operators, Part I, II and III, John Wiley & Sons, New York, 1971. Kato, T., Perturbation Theory for Linear Operators, Springer- Verlag, New York, 1976, 2nd Ed. 97 EEK—...— L___‘S:T" Kl [15] Luenberger, D.G., Optimization by Vector Space Methods, John Wiley & Sons, Inc., New York, 1969. [16] Rellich, F., Perturbation Theory of Eigenvalue Problems, Lecture Notes, New York University, 1953. [17] Baumgarter, H., Englichdimensionale analytische Storungtheorie, Academie-Verlag, 1973. [18] Knopp, K., Theory of Functions, Part II, Dover, New York, 1947 and 1975. [19] Ben-Israel, A. and Greville, T.N.E., Generalized Inverses: Theory and Applications, John Wiley & Sons, New York, 1974. [20] Haug, E.J. and Rousselet, B., "Design Sensitivity Analysis in Structural Mechanics: Eigenvalue Variations", Technical Report No. 52, NSF Project No. ENG 77-19967, Division of Material Engineering, College of Engineering, The University of Iowa, Iowa 52242. [21] Choi, Kyung, Personal Conversation, May, 1980. [22] Wilkinson, J.H. and Reinsch, C., Handbook for Automatic Compu- tation, Vol. II, Linear Algebra, Springer-Verlag, New York, 1971. [23] Forsythe, G.F. and Moler, C.B., Computer Solutions of Linear Algebraic Systems, Prentice-Hall, Inc., [24] Strang, G. and Fix, G.J., An Analysis of the Finite Element Method, Prentice Hall, Inc., Englewood Cliffs, N.J., 1973. [25] Prenter, P.M., Splines and Variational Methods, John Wiley & Sons, New York, 1975. [26] Bathe, K.J., Wilson, E.L., Numerical Methods in Finite Element Analysis, Prentice-Hall, Englewood Cliffs, N.J., 1976. [27] Haug, E.J. and Rousselet, B., Personal Conversation, May, 1980. [28] Meirovitch, L., Analytic Methods in Vibrations, MacMillan, New York, 1967. [29] Lancaster, P. Lambda-Matrices and Vibrating Systems, Pergamon Press, New York, 1965. [30] Caughey, T.K. and O'Kelley, M.E.S., "Classical Normal Modes in Damped Linear Dynamic Systems", Journal of Applied Mechanics, Vol. 32, 1965, pp. 583-588. [31] Inman, D.J. and Andry, A.N., "Some Results on the Nature of Eigenvalues of Discrete Damped Linear Systems", Journal of Applied Mechanics, ASME, to appear. 98