Linear Differential Operators

Total Page:16

File Type:pdf, Size:1020Kb

Linear Differential Operators 3 Linear Differential Operators Differential equations seem to be well suited as models for systems. Thus an understanding of differential equations is at least as important as an understanding of matrix equations. In Section 1.5 we inverted matrices and solved matrix equations. In this chapter we explore the analogous inversion and solution process for linear differential equations. Because of the presence of boundary conditions, the process of inverting a differential operator is somewhat more complex than the analogous matrix inversion. The notation ordinarily used for the study of differential equations is designed for easy handling of boundary conditions rather than for understanding of differential operators. As a consequence, the concept of the inverse of a differential operator is not widely understood among engineers. The approach we use in this chapter is one that draws a strong analogy between linear differential equations and matrix equations, thereby placing both these types of models in the same conceptual frame- work. The key concept is the Green’s function. It plays the same role for a linear differential equation as does the inverse matrix for a matrix equa- tion. There are both practical and theoretical reasons for examining the process of inverting differential operators. The inverse (or integral form) of a differential equation displays explicitly the input-output relationship of the system. Furthermore, integral operators are computationally and theoretically less troublesome than differential operators; for example, differentiation emphasizes data errors, whereas integration averages them. Consequently, the theoretical justification for applying many of the com- putational procedures of later chapters to differential systems is based on the inverse (or integral) description of the system. Finally, the application of the optimization techniques of Chapters 6-8 to differential systems often depends upon the prior determination of the integral forms of the systems. One of the reasons that matrix equations are widely used is that we have a practical, automatable scheme, Gaussian elimination, for inverting a matrix or solving a matrix equation. It is also possible to invert certain types of differential equations by computer automation. The greatest progress in understanding and automation has been made for linear, 93 94 Linear Differential Operators constant-coefficient differential equations with initial conditions. These equations are good models for many dynamic systems (systems which evolve with time). In Section 3.4 we examine these linear constant- coefficient models in state-space form and also in the form of nth-order differential equations. The inversion concept can be extended to partial differential equations. 3.1 A Differential Operator and Its Inverse Within the process of inverting a differential operator there is an analogue of the elimination technique for matrix inversion. However, the analogy between the matrix equation and the differential equation is clouded by the presence of the boundary conditions. As an example of a linear differential equation and its associated boundary conditions, we use -f” =u with f(O)= art and f(b) = CQ (3.1) Equation (3.1) can be viewed as a description of the relationship between the steady-state temperature distribution and the sources of heat in an insulated bar of length b. The temperature distribution f varies only as a function of position t along the bar. The temperature distribution is controlled partly by u, the heat generated (say, by induction heating) throughout the bar, and partly by constant temperature baths (of temperatures cyr and [x2, respectively) at the two ends t = 0 and t = b. Thus both the distributed input u and the boundary inputs { ai} have practical significance. The concepts of distributed and boundary inputs extend to other ordinary and partial differential equations. A Discrete Approximation of the Differential System In order to obtain a more transparent analogy to matrix equations and thereby clarify the role of the boundary conditions, we temporarily approximate the differential equation by a set of difference equations.* Let b = 4, substitute into (3.1) the finite-difference approximation *The approximation of derivatives by finite differences is a practical numerical approach to the solution of ordinary and partial differential equations. The error owing to the finite- difference approximation can be made as small as desired by using a sufficiently fine approximation to the derivatives. (See Forsythe and Wasow [3.3].) Special techniques are usually used to solve the resulting algebraic equations. See P&C 3.3 and Varga [3.12]. Sec. 3.1 A Differential Operator and Its Inverse 95 and evaluate the equation at t = 1, 2, and 3: -f(O)+2f(l) -f(2) =u(l) -f(l)+2f(2) -f(3) = u(2) - f(2) + 2f(3) - f(4) = u(3) (3.4) f(O) =a f(4) =a: It is obvious that this set of algebraic equations would not be invertible without the boundary conditions. We can view the boundary conditions either as an increase in the number of equations or as a decrease in the number of unknowns. The left side of (3.2), including the boundary conditions, is a matrix multiplication of the general vector (f(0) f( 1) * ’ * f(4))= in the space %5x ‘. The corresponding right-hand side of (3.2) is (u(1) u(2) u(3) (xi (~2)~; the boundary values increase the dimension of the range of definition by two. On the other hand, if we use the boundary conditions to eliminate two variables, we reduce the dimension of the domain of the matrix operator by two, f(0) and f(4) become part of the right-hand side, and the reduced matrix operates on the general vector (f(l) f(2) f(3))’ in 9R,3x ‘. By either the “expanded” or the “reduced” view, the transformation with its boundary conditions is invertible. In the next section we explore the differential equation and its boundary conditions along the same lines as we have used for this discrete approximation. The Role of the Boundary Conditions A differential operator without boundary conditions is like a matrix with fewer rows than columns: it leads to an underdetermined differential equation. In the same manner as in the discrete approximation (3.2), appropriate boundary conditions make a linear differential operator invert- ible. In order that we be able to denote the inverse of (3.1) in a simple manner as we do for matrix equations, we must combine the differential operator - D2 and the two boundary conditions into a single operator on a vector space. We can do so using the “increased equations” view of the boundary conditions. Let f be a function in the space e2(0,b) of twice continuously differentiable functions; then -f” will be in C?(O,b), the space of continuous functions. Define the differential system operator T: e2(0, b)-+ e (0, b) x ‘JR2 by Tf 9 (-f”,f(O),f(b)) (3.3) The system equations become Tf = (u, “1,(x2) (3.4) 96 Linear Differential Operators We are seeking an explicit expression of T- ’ such that f = T- ‘(u, ai, (w,). Because of the abstractness of T, an operation which produces a mixture of a distributed quantity u and discrete quantities {(xi}, it is not clear how to proceed to determine T-l. Standard techniques for solution of differential equations are more consistent with the “decreased unknowns” interpretation of the boundary conditions. Ordinarily, we solve the differential equation, - f” = u, ignoring the boundary conditions. Then we apply the boundary conditions to eliminate the arbitrary constants in the solution. If we think of the operator - D2 as being restricted through the whole solution process to act only on functions which satisfy the boundary conditions, then the “arbitrary” constants in the solution to the differential equation are not arbitrary; rather, they are specific (but unknown) functions of the boundary values, { aj}. We develop this interpretation of the inversion process into an explicit expression for the inverse of the operator T of (3.3). How do we express the “restriction” of -D2 in terms of an operator on a vector space? The set of functions which satisfy the boundary conditions is not a subspace of e2(0,b); it does not include the zero function (unless a 1 =a2=O). The analogue of this set of functions in the three-dimensional arrow space is a plane which does not pass through the origin. We frame the problem in terms of vector space concepts by separating the effects of the distributed and boundary inputs. In point of fact, it is the difference in the nature of these two types of inputs that has prevented the differential equation and the boundary conditions from being expressed as a single equation.* Decompose the differential system (3.1) into two parts, one involving only the distributed input, the other only the boundary inputs: -fpu with f,(O) = fd( b) = 0 (3.5) -fp(j with f,(O) = LYE, fb( b) = a2 (3.6) Equations (3.5) and (3.6) possess unique solutions. By superposition, these solutions combine to yield the unique solution f to (3.1); that is, f = fd+ fb. Each of these differential systems can be expressed as a single operator on a vector space. We invert the two systems separately. We work first with (3.5). The operator -D2 is onto C?(O, b); that is, we can obtain any continuous function by twice differentiating some function in e2(0, b). However, -D2 is singular; the general vector in nullspace (- D2) is of the form f(t) = ci + c,t. We modify the definition of the operator - D2 by reducing its domain. Let ?f be the subspace of functions in e2(0,b) which satisfy the homogeneous boundary conditions of (3.5), *Friedman [3.4] does include the boundary conditions in the differential equation by treating the boundary conditions as delta functions superimposed on the distributed input.
Recommended publications
  • Linear Differential Equations in N Th-Order ;
    Ordinary Differential Equations Second Course Babylon University 2016-2017 Linear Differential Equations in n th-order ; An nth-order linear differential equation has the form ; …. (1) Where the coefficients ;(j=0,1,2,…,n-1) ,and depends solely on the variable (not on or any derivative of ) If then Eq.(1) is homogenous ,if not then is nonhomogeneous . …… 2 A linear differential equation has constant coefficients if all the coefficients are constants ,if one or more is not constant then has variable coefficients . Examples on linear differential equations ; first order nonhomogeneous second order homogeneous third order nonhomogeneous fifth order homogeneous second order nonhomogeneous Theorem 1: Consider the initial value problem given by the linear differential equation(1)and the n initial Conditions; … …(3) Define the differential operator L(y) by …. (4) Where ;(j=0,1,2,…,n-1) is continuous on some interval of interest then L(y) = and a linear homogeneous differential equation written as; L(y) =0 Definition: Linearly Independent Solution; A set of functions … is linearly dependent on if there exists constants … not all zero ,such that …. (5) on 1 Ordinary Differential Equations Second Course Babylon University 2016-2017 A set of functions is linearly dependent if there exists another set of constants not all zero, that is satisfy eq.(5). If the only solution to eq.(5) is ,then the set of functions is linearly independent on The functions are linearly independent Theorem 2: If the functions y1 ,y2, …, yn are solutions for differential
    [Show full text]
  • The Wronskian
    Jim Lambers MAT 285 Spring Semester 2012-13 Lecture 16 Notes These notes correspond to Section 3.2 in the text. The Wronskian Now that we know how to solve a linear second-order homogeneous ODE y00 + p(t)y0 + q(t)y = 0 in certain cases, we establish some theory about general equations of this form. First, we introduce some notation. We define a second-order linear differential operator L by L[y] = y00 + p(t)y0 + q(t)y: Then, a initial value problem with a second-order homogeneous linear ODE can be stated as 0 L[y] = 0; y(t0) = y0; y (t0) = z0: We state a result concerning existence and uniqueness of solutions to such ODE, analogous to the Existence-Uniqueness Theorem for first-order ODE. Theorem (Existence-Uniqueness) The initial value problem 00 0 0 L[y] = y + p(t)y + q(t)y = g(t); y(t0) = y0; y (t0) = z0 has a unique solution on an open interval I containing the point t0 if p, q and g are continuous on I. The solution is twice differentiable on I. Now, suppose that we have obtained two solutions y1 and y2 of the equation L[y] = 0. Then 00 0 00 0 y1 + p(t)y1 + q(t)y1 = 0; y2 + p(t)y2 + q(t)y2 = 0: Let y = c1y1 + c2y2, where c1 and c2 are constants. Then L[y] = L[c1y1 + c2y2] 00 0 = (c1y1 + c2y2) + p(t)(c1y1 + c2y2) + q(t)(c1y1 + c2y2) 00 00 0 0 = c1y1 + c2y2 + c1p(t)y1 + c2p(t)y2 + c1q(t)y1 + c2q(t)y2 00 0 00 0 = c1(y1 + p(t)y1 + q(t)y1) + c2(y2 + p(t)y2 + q(t)y2) = 0: We have just established the following theorem.
    [Show full text]
  • An Introduction to Pseudo-Differential Operators
    An introduction to pseudo-differential operators Jean-Marc Bouclet1 Universit´ede Toulouse 3 Institut de Math´ematiquesde Toulouse [email protected] 2 Contents 1 Background on analysis on manifolds 7 2 The Weyl law: statement of the problem 13 3 Pseudodifferential calculus 19 3.1 The Fourier transform . 19 3.2 Definition of pseudo-differential operators . 21 3.3 Symbolic calculus . 24 3.4 Proofs . 27 4 Some tools of spectral theory 41 4.1 Hilbert-Schmidt operators . 41 4.2 Trace class operators . 44 4.3 Functional calculus via the Helffer-Sj¨ostrandformula . 50 5 L2 bounds for pseudo-differential operators 55 5.1 L2 estimates . 55 5.2 Hilbert-Schmidt estimates . 60 5.3 Trace class estimates . 61 6 Elliptic parametrix and applications 65 n 6.1 Parametrix on R ................................ 65 6.2 Localization of the parametrix . 71 7 Proof of the Weyl law 75 7.1 The resolvent of the Laplacian on a compact manifold . 75 7.2 Diagonalization of ∆g .............................. 78 7.3 Proof of the Weyl law . 81 A Proof of the Peetre Theorem 85 3 4 CONTENTS Introduction The spirit of these notes is to use the famous Weyl law (on the asymptotic distribution of eigenvalues of the Laplace operator on a compact manifold) as a case study to introduce and illustrate one of the many applications of the pseudo-differential calculus. The material presented here corresponds to a 24 hours course taught in Toulouse in 2012 and 2013. We introduce all tools required to give a complete proof of the Weyl law, mainly the semiclassical pseudo-differential calculus, and then of course prove it! The price to pay is that we avoid presenting many classical concepts or results which are not necessary for our purpose (such as Borel summations, principal symbols, invariance by diffeomorphism or the G˚ardinginequality).
    [Show full text]
  • First Integrals of Differential Operators from SL(2, R) Symmetries
    mathematics Article First Integrals of Differential Operators from SL(2, R) Symmetries Paola Morando 1 , Concepción Muriel 2,* and Adrián Ruiz 3 1 Dipartimento di Scienze Agrarie e Ambientali-Produzione, Territorio, Agroenergia, Università Degli Studi di Milano, 20133 Milano, Italy; [email protected] 2 Departamento de Matemáticas, Facultad de Ciencias, Universidad de Cádiz, 11510 Puerto Real, Spain 3 Departamento de Matemáticas, Escuela de Ingenierías Marina, Náutica y Radioelectrónica, Universidad de Cádiz, 11510 Puerto Real, Spain; [email protected] * Correspondence: [email protected] Received: 30 September 2020; Accepted: 27 November 2020; Published: 4 December 2020 Abstract: The construction of first integrals for SL(2, R)-invariant nth-order ordinary differential equations is a non-trivial problem due to the nonsolvability of the underlying symmetry algebra sl(2, R). Firstly, we provide for n = 2 an explicit expression for two non-constant first integrals through algebraic operations involving the symmetry generators of sl(2, R), and without any kind of integration. Moreover, although there are cases when the two first integrals are functionally independent, it is proved that a second functionally independent first integral arises by a single quadrature. This result is extended for n > 2, provided that a solvable structure for an integrable distribution generated by the differential operator associated to the equation and one of the prolonged symmetry generators of sl(2, R) is known. Several examples illustrate the procedures. Keywords: differential operator; first integral; solvable structure; integrable distribution 1. Introduction The study of nth-order ordinary differential equations admitting the unimodular Lie group SL(2, R) is a non-trivial problem due to the nonsolvability of the underlying symmetry algebra sl(2, R).
    [Show full text]
  • 1 Function Spaces, Linear Operators, and Green's Functions
    1 1 Function Spaces, Linear Operators, and Green’s Functions 1.1 Function Spaces Consider the set of all complex-valued functions of the real variable x, denoted by f (x), g(x), ..., and defined on the interval (a, b). We shall restrict ourselves to those functions which are square-integrable.Definetheinner product of any two of the latter functions by b ∗ (f , g) ≡ f (x) g (x) dx, (1.1.1) a in which f ∗(x) is the complex conjugate of f (x). The following properties of the inner product follow from definition (1.1.1): (f , g)∗ = (g, f ), ( f , g + h) = ( f , g) + ( f , h), (1.1.2) ( f , αg) = α( f , g), (αf , g) = α∗( f , g), with α a complex scalar. While the inner product of any two functions is in general a complex number, the inner product of a function with itself is a real number and is nonnegative. This prompts us to define the norm of a function by 1 b 2 ∗ f ≡ ( f , f ) = f (x) f (x)dx , (1.1.3) a provided that f is square-integrable , i.e., f < ∞. Equation (1.1.3) constitutes a proper definition for a norm since it satisfies the following conditions: scalar = | | · (i) αf α f , for all comlex α, multiplication (ii) positivity f > 0, for all f = 0, = = f 0, if and only if f 0, triangular (iii) f + g ≤ f + g . inequality (1.1.4) Applied Mathematical Methods in Theoretical Physics, Second Edition. Michio Masujima Copyright 2009 WILEY-VCH Verlag GmbH & Co.
    [Show full text]
  • Sturm Liouville Theory
    1 Sturm-Liouville System 1.1 Sturm-Liouville Dierential Equation Denition 1. A second ordered dierential equation of the form d d − p(x) y + q(x)y = λω(x)y; x 2 [a; b] (1) dx dx with p; q and ! specied such that p(x) > 0 and !(x) > 0 for x 2 (a; b), is called a Sturm- Liouville(SL) dierential equation. Note that SL dierential equation is essentially an eigenvalue problem since λ is not specied. While solving SL equation, both λ and y must be determined. Example 2. The Schrodinger equation 2 − ~ + V (x) = E 2m on an interval [a; b] is a SL dierential equation. 1.2 Boundary Conditions Boundary conditions for a solution y of a dierential equation on interval [a; b] are classied as follows: Mixed Boundary Conditions Boundary conditions of the form 0 cay(a) + day (a) = α 0 cby(b) + dby (b) = β (2) where, ca; da; cb; db; α and β are constants, are called mixed Dirichlet-Neumann boundary conditions. When both α = β = 0 the boundary conditions are said to be homogeneous. Special cases are Dirichlet BC (da = db = 0) and Neumann BC (ca = cb = 0) Periodic Boundary Conditions Boundary conditions of the form y(a) = y(b) y0(a) = y0(b) (3) are called periodic boundary conditions. Example 3. Some examples are 1. Stretched vibrating string clamped at two ends: Dirichlet BC 2. Electrostatic potential on the surface of a volume: Dirichlet BC 3. Electrostatic eld on the surface of a volume: Neumann BC 4. A heat-conducting rod with two ends in heat baths: Dirichlet BC 1 1.3 Sturm-Liouville System (Problem) Denition 4.
    [Show full text]
  • Step Functions, Delta Functions, and the Variation of Parameters Formula
    STEP FUNCTIONS, DELTA FUNCTIONS, AND THE VARIATION OF PARAMETERS FORMULA STEPHEN SCHECTER 1. The unit step function and piecewise continuous functions The Heaviside unit step function u(t) is given by 0 if t< 0, u(t)= (1 if t> 0. The function u(t) is not defined at t = 0. Often we will not worry about the value of a function at a point where it is discontinuous, since often it doesn’t matter. u(t) u(t−a) 1−u(t−b) u(t−a)−u(t−b) 1 1 1 1 a b a b Figure 1.1. Heaviside unit step function. The function u(t) turns on at t = 0. The function u(t − a) is just u(t) shifted, or dragged, so that it turns on at t = a. The function 1 − u(t − b) turns off at t = b. The function u(t − a) − u(t − b), with a < b, turns on a t = a and turns off at t = b. We can use the unit step function to crop and shift functions. Multiplying f(t) by u(t−a) crops f(t) so that it turns on at t = a: 0 if t < a, u(t − a)f(t)= (f(t) if t > a. Multiplying f(t) by u(t − a) − u(t − b), with a < b, crops f(t) so that it turns on at t = a and turns off at t = b: 0 if t < a, (u(t − a) − u(t − b))f(t)= f(t) if a<t<b, 0 if t > b.
    [Show full text]
  • Gradients and Directional Derivatives R Horan & M Lavelle
    Intermediate Mathematics Gradients and Directional Derivatives R Horan & M Lavelle The aim of this package is to provide a short self assessment programme for students who want to obtain an ability in vector calculus to calculate gradients and directional derivatives. Copyright c 2004 [email protected] , [email protected] Last Revision Date: August 24, 2004 Version 1.0 Table of Contents 1. Introduction (Vectors) 2. Gradient (Grad) 3. Directional Derivatives 4. Final Quiz Solutions to Exercises Solutions to Quizzes The full range of these packages and some instructions, should they be required, can be obtained from our web page Mathematics Support Materials. Section 1: Introduction (Vectors) 3 1. Introduction (Vectors) The base vectors in two dimensional Cartesian coordinates are the unit vector i in the positive direction of the x axis and the unit vector j in the y direction. See Diagram 1. (In three dimensions we also require k, the unit vector in the z direction.) The position vector of a point P (x, y) in two dimensions is xi + yj . We will often denote this important vector by r . See Diagram 2. (In three dimensions the position vector is r = xi + yj + zk .) y 6 Diagram 1 y 6 Diagram 2 P (x, y) 3 r 6 j yj 6- - - - 0 i x 0 xi x Section 1: Introduction (Vectors) 4 The vector differential operator ∇ , called “del” or “nabla”, is defined in three dimensions to be: ∂ ∂ ∂ ∇ = i + j + k . ∂x ∂y ∂z Note that these are partial derivatives! This vector operator may be applied to (differentiable) scalar func- tions (scalar fields) and the result is a special case of a vector field, called a gradient vector field.
    [Show full text]
  • Differential Operators Commuting with Convolution Integral Operators
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOURNAL OF MATHEMATICAL ANALYSIS AND APPLICATIONS 91, 80-93 (1983) Differential Operators Commuting with Convolution Integral Operators F. ALBERTOGRUNBAUM * Department of Mathematics and Lawrence Berkeley Laboratory, University of California, Berkeley, California 94720 Submitted by P. D. Lax Received March 13, 1981 1. INTRODUCTION The need to make a detailed study of the spectral structure of the convolution type integral operator acting on L*(B) arises in a number of applications, including optics [ 11, radio astronomy [2,3], electron microscopy [4], and x-ray tomography [5,6]. For instance, consider the problem of determining the values of a function g(x), x E R, which is known to be supported in the set B, from the knowledge of its Fourier transform g(L) for A E A (both A and B are compact and have nonempty interiors). While in principle g(x) is completely determined by analytic continuation, the practical solution of this problem is a different matter. One is led to the consideration of the operator (1), where k(r) stands for the Fourier transform of the characteristic function of the set A. K turns out to be a compact operator with spectrum in (0, 1), and a detailed spectral decom- position of K exhibits the directions along which the problem is ill posed [6, 71. In the case when A and B are intervals centered at the origin the detailed structure of the eigenfunctions and eigenvalues of (1) has been displayed in * Partly supported by NSF Grant MCS 78-06718 and by the Director, Oflice of Energy Research, Office of Basic Energy Sciences, Engineering, Mathematical, and Geosciences Division of the U.S.
    [Show full text]
  • Differential Operator Method of Finding a Particular Solution to an Ordinary Nonhomogeneous Linear Differential Equation with Constant Coefficients
    Differential Operator Method of Finding A Particular Solution to An Ordinary Nonhomogeneous Linear Differential Equation with Constant Coefficients Wenfeng Chen Department of Mathematics and Physics, College of Arts and Sciences, SUNY Polytechnic Institute, Utica, NY 13502, USA We systematically introduce the idea of applying differential operator method to find a par- ticular solution of an ordinary nonhomogeneous linear differential equation with constant coefficients when the nonhomogeneous term is a polynomial function, exponential function, sine function, cosine function or any possible product of these functions. In particular, differ- ent from the differential operator method introduced in literature, we propose and highlight utilizing the definition of the inverse of differential operator to determine a particular solu- tion. We suggest that this method should be introduced in textbooks and widely used for determining a particular solution of an ordinary nonhomogeneous linear differential equation with constant coefficients in parallel to the method of undetermined coefficients. arXiv:1802.09343v1 [math.GM] 16 Feb 2018 2 I. INTRODUCTION When we determine a particular solution of an ordinary nonhomogeneous linear differential equa- tion of constant coefficients with nonhomogeneous terms being a polynomial function, an exponen- tial function, a sine function, a cosine function or any possible products of these functions, usually only the method of undetermined coefficients is introduced in undergraduate textbooks [1–3]. The basic idea of this method is first assuming the general form of a particular solution with the co- efficients undetermined, based on the nonhomogeneous term, and then substituting the assumed solution to determine the coefficients. The calculation process of this method is very lengthy and time-consuming.
    [Show full text]
  • Sturm–Liouville Theory
    2 Sturm–Liouville Theory So far, we’ve examined the Fourier decomposition of functions defined on some interval (often scaled to be from π to π). We viewed this expansion as an infinite dimensional − analogue of expanding a finite dimensional vector into its components in an orthonormal basis. But this is just the tip of the iceberg. Recalling other games we play in linear algebra, you might well be wondering whether we couldn’t have found some other basis in which to expand our functions. You might also wonder whether there shouldn’t be some role for matrices in this story. If so, read on! 2.1 Self-adjoint matrices We’ll begin by reviewing some facts about matrices. Let V and W be finite dimensional vector spaces (defined, say, over the complex numbers) with dim V = n and dim W = m. Suppose we have a linear map M : V W . By linearity, we know what M does to any → vector v V if we know what it does to a complete set v , v ,..., v of basis vectors ∈ { 1 2 n} in V . Furthermore, given a basis w , w ,..., w of W we can represent the map M in { 1 2 m} terms of an m n matrix M whose components are × Mai =(wa,Mvi) for a =1, . , m and i =1, . , n , (2.1) where ( , ) is the inner product in W . We’ll be particularly interested in the case m = n, when the matrix M is square and the map M takes M : V W = V is an isomorphism of vector spaces.
    [Show full text]
  • Functional Analysis
    Principles of Functional Analysis SECOND EDITION Martin Schechter Graduate Studies in Mathematics Volume 36 American Mathematical Society http://dx.doi.org/10.1090/gsm/036 Selected Titles in This Series 36 Martin Schechter, Principles of functional analysis, second edition, 2002 35 JamesF.DavisandPaulKirk, Lecture notes in algebraic topology, 2001 34 Sigurdur Helgason, Differential geometry, Lie groups, and symmetric spaces, 2001 33 Dmitri Burago, Yuri Burago, and Sergei Ivanov, A course in metric geometry, 2001 32 Robert G. Bartle, A modern theory of integration, 2001 31 Ralf Korn and Elke Korn, Option pricing and portfolio optimization: Modern methods of financial mathematics, 2001 30 J. C. McConnell and J. C. Robson, Noncommutative Noetherian rings, 2001 29 Javier Duoandikoetxea, Fourier analysis, 2001 28 Liviu I. Nicolaescu, Notes on Seiberg-Witten theory, 2000 27 Thierry Aubin, A course in differential geometry, 2001 26 Rolf Berndt, An introduction to symplectic geometry, 2001 25 Thomas Friedrich, Dirac operators in Riemannian geometry, 2000 24 Helmut Koch, Number theory: Algebraic numbers and functions, 2000 23 Alberto Candel and Lawrence Conlon, Foliations I, 2000 22 G¨unter R. Krause and Thomas H. Lenagan, Growth of algebras and Gelfand-Kirillov dimension, 2000 21 John B. Conway, A course in operator theory, 2000 20 Robert E. Gompf and Andr´as I. Stipsicz, 4-manifolds and Kirby calculus, 1999 19 Lawrence C. Evans, Partial differential equations, 1998 18 Winfried Just and Martin Weese, Discovering modern set theory. II: Set-theoretic tools for every mathematician, 1997 17 Henryk Iwaniec, Topics in classical automorphic forms, 1997 16 Richard V. Kadison and John R. Ringrose, Fundamentals of the theory of operator algebras.
    [Show full text]