The Holomorphic Functional Calculus Approach to Operator Semigroups

Total Page:16

File Type:pdf, Size:1020Kb

The Holomorphic Functional Calculus Approach to Operator Semigroups Chapter 8 The Holomorphic Functional Calculus Approach to Operator Semigroups In Chapter 6 we have associated with a given C0-semigroup an operator, called its generator, and a functional calculus, the Hille{Phillips calculus. Here we address the converse problem: which operators are generators? The classical answer to this question and one of our goals in this chapter is the so- called Hille{Yosida theorem, which gives a characterization of the generator property in terms of resolvent estimates. However, instead of following the standard texts like [1], we shall approach the problem via a suitable functional calculus construction. Our method is generic, in that we learn here how to use Cauchy integrals over infinite con- tours in order to define functional calculi for unbounded operators satisfying resolvent estimates. (The same approach shall be taken later for defining holomorphic functional calculi for sectorial and strip-type operators.) This chapter is heavily based on the first part of [2], which in turn goes back to the unpublished note [3]. 8.1 Operators of Half-Plane Type Recall from Chapter 6 that if B generates a C0-semigroup of type (M; !), then s(B) is contained in the closed half-plane [ Re z ≤ ! ] and M kR(λ, B)k ≤ (Re λ > !): Re λ − ! In particular, if ! < α then R(·;A) is uniformly bounded for Re λ ≥ α. This \spectral data" is our model for what we shall call below an operator of \left half-plane type". Recall, however, that the associated Hille-Phillips calculus for the semi- group generated by B is actually a calculus for the negative generator 135 136 8 The Holomorphic Functional Calculus Approach to Operator Semigroups A := −B. So in order to stay in line with the Hille{Phillips picture we shall construct our new functional calculus for operators of right half-plane type. Let us turn to the precise definitions. For ! 2 [−∞; 1] we let L! := fz 2 C j Re z < !g and R! := fz 2 C j Re z > !g be the open left and right half-plane at the abscissa !. (So R! = C++! in the terminology of Chapter 6.) An operator A on a Banach space X is said to be of right half-plane type ! 2 (−∞; 1] if s(A) ⊆ R! and M(A; α) := supfkR(z; A)k j Re z ≤ αg < 1 for every α < !. We say that A is of right half-plane type if A is of right half-plane type ! for some !. For an operator A of right half-plane type we call !hp(A) := sup α sup kR(z; A)k < 1 Re z≤α the abscissa of uniform boundedness of its resolvent. An operator B is called of left half-plane type ! if A = −B is of right half-plane type −!. The abscissa of uniform boundedness of the resolvent of such an operator is denoted by s0(B) := inf α sup kR(λ, B)k < 1 2 R [ {−∞} Re λ≥α as it is common in semigroup theory. (See [4, p.342] where this number is called the pseudo-spectral bound of B.) Remark 8.1. The definition of a half-plane type operator may appear com- plicated and unintuitive at first glance. Why do we not define A to be of right half-plane type ! simply by s(A) ⊆ R!; sup kR(z; A)k < 1 ? (8:1) Re z<! The answer is that, with our definition, A is of right half-plane type !hp(A), whereas this would be false with the alternative definition. See Exercise 8.1. In the following, when we speak of half-plane type operators, it has to be read as right half-plane type, unless explicitly noted otherwise. 8.2 Functional Calculus on Half-Planes As already announced, we are aiming at defining a functional calculus for operators A of half-plane type. There is a \cheap" way of doing this: pick 8.2 Functional Calculus on Half-Planes 137 some λ 2 r(A) and define Aλ := R(λ, A). Then set up the Dunford-Riesz calculus for Aλ and play it back to A. (See Exercise 8.2 for details.) This approach, elegant as it appears, does not use at all the resolvent estimates which are characteristic for operators of half-plane type. Moreover, it is principally unsuitable to reach our goal, because the functions e−tz have an essential singularity at 1 and hence are not accessible by that calculus nor by its canonical extension. So one has to take a different route. The idea is to mimic the construction of the Dunford{Riesz calculus but with contours leading into the possible singularity, which here is the point at infinity. Since the spectrum of an oper- ator A of half-plane type may be a whole right half-plane, a straightforward choice for such contours are vertical straight lines, oriented from top to down. (Picture this on the Riemann sphere and you will see that this comes as close as possible to \surrounding once the spectrum counterclockwise".) s(A) ! δ !hp(A) R γ Fig. 8.1 The path γ runs from δ + i1 to δ − i1, where ! < δ < !hp(A). In the following, we shall work out this plan. Let A be an operator of half- plane type on a Banach space X and fix ! < δ < !hp(A). We want to define 1 Z Φ(f) := f(z)R(z; A) dz 2pi γ 138 8 The Holomorphic Functional Calculus Approach to Operator Semigroups for suitable functions f 2 Hol(R!), where γ is the top-to-down parameriza- tion of the vertical line [ Re z = δ ], see Figure 8.2. (We use the alternative expressions 1 Z −1 Z δ+i1 f(z)R(z; A) dz and f(z)R(z; A) dz 2pi Re z=δ 2pi δ−i1 to denote this integral. Observe that the orientation of the contour is implicit in the first of these.) Of course, as the contour is infinite, not every holomorphic function f on R! will be suitable. Rather, the properties of f and R(z;A) should match in order to guarantee that the integral is convergent. As the resolvent is assumed to be bounded, it is reasonable to require that f is integrable on vertical lines. This is the reason for the following definition. For ! 2 R we let1 1 1 E(R!) := f 2 H (R!) j 8 δ > ! : f 2 L (δ + iR) and call it the algebra of elementary functions on the half-plane R!. By writing f 2 L1(δ + iR) we mean that the function f(δ + ix) is in L1(R), i.e., that Z jf(δ + ix)j dx < 1: R 1 It is easily checked that E(R!) is a (non-unital) subalgebra of H (R!). For elementary functions we have the following version of Cauchy's integral the- orem. Lemma 8.2. Let f 2 E(R!), δ > !, and a 2 C with Re a 6= δ. Then ( 1 Z f(z) f(a) if δ < Re a; dz = (8:2) 2pi Re z=δ z − a 0 if δ > Re a: Moreover, f 2 C0(Rδ). Proof. Consider first the case Re a < δ. Define Z f(z) g(a) := dz (Re a < δ): Re z=δ z − a Then, by some standard arguments, g is holomorphic and Z 0 f(z) g (a) = 2 dz: Re z=δ (z − a) We claim that for fixed a this integral does not depend on δ > Re a. Indeed, 1 For an open set O ⊆ C we denote by H1(O) the space of bounded and holomorphic functions. It is a unital Banach algebra with respect to the sup-norm. 8.2 Functional Calculus on Half-Planes 139 1 Z f(z) 0 = 2 dz; 2pi γR (z − a) by Cauchy's theorem, where γR is the positively oriented boundary of the rectangle [δ; δ0]×[−R; R] with δ0 > δ and R > 0. Since f is uniformly bounded, the integrals over the horizontal line segments tend to zero as R ! 1, resulting in 1 Z f(z) 1 Z f(z) 0 = 2 dz − 2 dz: 2pi Re z=δ (z − a) 2pi Re z=δ0 (z − a) But now we can let δ0 ! 1 and, again by the boundedness of f, we conclude that g0(a) = 0. Hence, g is constant. On the other hand, g(a) ! 0 as Re a ! −∞, which yields g = 0. Consider now the case Re a > δ. Again we employ Cauchy's theorem with 0 the contour γR as above, but now with δ < Re a < δ and R > jIm aj. This yields 1 Z f(z) f(a) = dz: 2pi γR z − a Again, the integrals over the horizontal line segments tend to zero as R ! 1. It follows that 1 Z f(z) 1 Z f(z) f(a) = dz − dz: 2pi Re z=δ z − a 2pi Re z=δ0 z − a By what we have proved first, the second integral is equal to zero, and we are done. For the last assertion, fix δ0 > δ. Then it follows from (8.2) and the dominated 0 convergence theorem that f(a) ! 0 as a ! 1 within Rδ0 . As δ > δ > ! were arbitrary, the claim follows. For f 2 E(R!) we now let (as announced above) 1 Z ΦA(f) := f(z)R(z; A) dz; (8:3) 2pi Re z=δ where A is the given half-plane type operator and ! < δ < !hp(A), see also Figure 8.2.
Recommended publications
  • 2 Hilbert Spaces You Should Have Seen Some Examples Last Semester
    2 Hilbert spaces You should have seen some examples last semester. The simplest (finite-dimensional) ex- C • A complex Hilbert space H is a complete normed space over whose norm is derived from an ample is Cn with its standard inner product. It’s worth recalling from linear algebra that if V is inner product. That is, we assume that there is a sesquilinear form ( , ): H H C, linear in · · × → an n-dimensional (complex) vector space, then from any set of n linearly independent vectors we the first variable and conjugate linear in the second, such that can manufacture an orthonormal basis e1, e2,..., en using the Gram-Schmidt process. In terms of this basis we can write any v V in the form (f ,д) = (д, f ), ∈ v = a e , a = (v, e ) (f , f ) 0 f H, and (f , f ) = 0 = f = 0. i i i i ≥ ∀ ∈ ⇒ ∑ The norm and inner product are related by which can be derived by taking the inner product of the equation v = aiei with ei. We also have n ∑ (f , f ) = f 2. ∥ ∥ v 2 = a 2. ∥ ∥ | i | i=1 We will always assume that H is separable (has a countable dense subset). ∑ Standard infinite-dimensional examples are l2(N) or l2(Z), the space of square-summable As usual for a normed space, the distance on H is given byd(f ,д) = f д = (f д, f д). • • ∥ − ∥ − − sequences, and L2(Ω) where Ω is a measurable subset of Rn. The Cauchy-Schwarz and triangle inequalities, • √ (f ,д) f д , f + д f + д , | | ≤ ∥ ∥∥ ∥ ∥ ∥ ≤ ∥ ∥ ∥ ∥ 2.1 Orthogonality can be derived fairly easily from the inner product.
    [Show full text]
  • 18.102 Introduction to Functional Analysis Spring 2009
    MIT OpenCourseWare http://ocw.mit.edu 18.102 Introduction to Functional Analysis Spring 2009 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. 108 LECTURE NOTES FOR 18.102, SPRING 2009 Lecture 19. Thursday, April 16 I am heading towards the spectral theory of self-adjoint compact operators. This is rather similar to the spectral theory of self-adjoint matrices and has many useful applications. There is a very effective spectral theory of general bounded but self- adjoint operators but I do not expect to have time to do this. There is also a pretty satisfactory spectral theory of non-selfadjoint compact operators, which it is more likely I will get to. There is no satisfactory spectral theory for general non-compact and non-self-adjoint operators as you can easily see from examples (such as the shift operator). In some sense compact operators are ‘small’ and rather like finite rank operators. If you accept this, then you will want to say that an operator such as (19.1) Id −K; K 2 K(H) is ‘big’. We are quite interested in this operator because of spectral theory. To say that λ 2 C is an eigenvalue of K is to say that there is a non-trivial solution of (19.2) Ku − λu = 0 where non-trivial means other than than the solution u = 0 which always exists. If λ =6 0 we can divide by λ and we are looking for solutions of −1 (19.3) (Id −λ K)u = 0 −1 which is just (19.1) for another compact operator, namely λ K: What are properties of Id −K which migh show it to be ‘big? Here are three: Proposition 26.
    [Show full text]
  • Curl, Divergence and Laplacian
    Curl, Divergence and Laplacian What to know: 1. The definition of curl and it two properties, that is, theorem 1, and be able to predict qualitatively how the curl of a vector field behaves from a picture. 2. The definition of divergence and it two properties, that is, if div F~ 6= 0 then F~ can't be written as the curl of another field, and be able to tell a vector field of clearly nonzero,positive or negative divergence from the picture. 3. Know the definition of the Laplace operator 4. Know what kind of objects those operator take as input and what they give as output. The curl operator Let's look at two plots of vector fields: Figure 1: The vector field Figure 2: The vector field h−y; x; 0i: h1; 1; 0i We can observe that the second one looks like it is rotating around the z axis. We'd like to be able to predict this kind of behavior without having to look at a picture. We also promised to find a criterion that checks whether a vector field is conservative in R3. Both of those goals are accomplished using a tool called the curl operator, even though neither of those two properties is exactly obvious from the definition we'll give. Definition 1. Let F~ = hP; Q; Ri be a vector field in R3, where P , Q and R are continuously differentiable. We define the curl operator: @R @Q @P @R @Q @P curl F~ = − ~i + − ~j + − ~k: (1) @y @z @z @x @x @y Remarks: 1.
    [Show full text]
  • Numerical Operator Calculus in Higher Dimensions
    Numerical operator calculus in higher dimensions Gregory Beylkin* and Martin J. Mohlenkamp Applied Mathematics, University of Colorado, Boulder, CO 80309 Communicated by Ronald R. Coifman, Yale University, New Haven, CT, May 31, 2002 (received for review July 31, 2001) When an algorithm in dimension one is extended to dimension d, equation using only one-dimensional operations and thus avoid- in nearly every case its computational cost is taken to the power d. ing the exponential dependence on d. However, if the best This fundamental difficulty is the single greatest impediment to approximate solution of the form (Eq. 1) is not good enough, solving many important problems and has been dubbed the curse there is no way to improve the accuracy. of dimensionality. For numerical analysis in dimension d,we The natural extension of Eq. 1 is the form propose to use a representation for vectors and matrices that generalizes separation of variables while allowing controlled ac- r ͑ ͒ ϭ ͸ ␾l ͑ ͒ ␾l ͑ ͒ curacy. Basic linear algebra operations can be performed in this f x1,...,xd sl 1 x1 ··· d xd . [2] representation using one-dimensional operations, thus bypassing lϭ1 the exponential scaling with respect to the dimension. Although not all operators and algorithms may be compatible with this The key quantity in Eq. 2 is r, which we call the separation rank. representation, we believe that many of the most important ones By increasing r, the approximate solution can be made as are. We prove that the multiparticle Schro¨dinger operator, as well accurate as desired.
    [Show full text]
  • Maxwell's Equations in Differential Form
    M03_RAO3333_1_SE_CHO3.QXD 4/9/08 1:17 PM Page 71 CHAPTER Maxwell’s Equations in Differential Form 3 In Chapter 2 we introduced Maxwell’s equations in integral form. We learned that the quantities involved in the formulation of these equations are the scalar quantities, elec- tromotive force, magnetomotive force, magnetic flux, displacement flux, charge, and current, which are related to the field vectors and source densities through line, surface, and volume integrals. Thus, the integral forms of Maxwell’s equations,while containing all the information pertinent to the interdependence of the field and source quantities over a given region in space, do not permit us to study directly the interaction between the field vectors and their relationships with the source densities at individual points. It is our goal in this chapter to derive the differential forms of Maxwell’s equations that apply directly to the field vectors and source densities at a given point. We shall derive Maxwell’s equations in differential form by applying Maxwell’s equations in integral form to infinitesimal closed paths, surfaces, and volumes, in the limit that they shrink to points. We will find that the differential equations relate the spatial variations of the field vectors at a given point to their temporal variations and to the charge and current densities at that point. In this process we shall also learn two important operations in vector calculus, known as curl and divergence, and two related theorems, known as Stokes’ and divergence theorems. 3.1 FARADAY’S LAW We recall from the previous chapter that Faraday’s law is given in integral form by d E # dl =- B # dS (3.1) CC dt LS where S is any surface bounded by the closed path C.In the most general case,the elec- tric and magnetic fields have all three components (x, y,and z) and are dependent on all three coordinates (x, y,and z) in addition to time (t).
    [Show full text]
  • Inverse and Implicit Function Theorems for Noncommutative
    Inverse and Implicit Function Theorems for Noncommutative Functions on Operator Domains Mark E. Mancuso Abstract Classically, a noncommutative function is defined on a graded domain of tuples of square matrices. In this note, we introduce a notion of a noncommutative function defined on a domain Ω ⊂ B(H)d, where H is an infinite dimensional Hilbert space. Inverse and implicit function theorems in this setting are established. When these operatorial noncommutative functions are suitably continuous in the strong operator topology, a noncommutative dilation-theoretic construction is used to show that the assumptions on their derivatives may be relaxed from boundedness below to injectivity. Keywords: Noncommutive functions, operator noncommutative functions, free anal- ysis, inverse and implicit function theorems, strong operator topology, dilation theory. MSC (2010): Primary 46L52; Secondary 47A56, 47J07. INTRODUCTION Polynomials in d noncommuting indeterminates can naturally be evaluated on d-tuples of square matrices of any size. The resulting function is graded (tuples of n × n ma- trices are mapped to n × n matrices) and preserves direct sums and similarities. Along with polynomials, noncommutative rational functions and power series, the convergence of which has been studied for example in [9], [14], [15], serve as prototypical examples of a more general class of functions called noncommutative functions. The theory of non- commutative functions finds its origin in the 1973 work of J. L. Taylor [17], who studied arXiv:1804.01040v2 [math.FA] 7 Aug 2019 the functional calculus of noncommuting operators. Roughly speaking, noncommutative functions are to polynomials in noncommuting variables as holomorphic functions from complex analysis are to polynomials in commuting variables.
    [Show full text]
  • 23. Kernel, Rank, Range
    23. Kernel, Rank, Range We now study linear transformations in more detail. First, we establish some important vocabulary. The range of a linear transformation f : V ! W is the set of vectors the linear transformation maps to. This set is also often called the image of f, written ran(f) = Im(f) = L(V ) = fL(v)jv 2 V g ⊂ W: The domain of a linear transformation is often called the pre-image of f. We can also talk about the pre-image of any subset of vectors U 2 W : L−1(U) = fv 2 V jL(v) 2 Ug ⊂ V: A linear transformation f is one-to-one if for any x 6= y 2 V , f(x) 6= f(y). In other words, different vector in V always map to different vectors in W . One-to-one transformations are also known as injective transformations. Notice that injectivity is a condition on the pre-image of f. A linear transformation f is onto if for every w 2 W , there exists an x 2 V such that f(x) = w. In other words, every vector in W is the image of some vector in V . An onto transformation is also known as an surjective transformation. Notice that surjectivity is a condition on the image of f. 1 Suppose L : V ! W is not injective. Then we can find v1 6= v2 such that Lv1 = Lv2. Then v1 − v2 6= 0, but L(v1 − v2) = 0: Definition Let L : V ! W be a linear transformation. The set of all vectors v such that Lv = 0W is called the kernel of L: ker L = fv 2 V jLv = 0g: 1 The notions of one-to-one and onto can be generalized to arbitrary functions on sets.
    [Show full text]
  • Algebra Homomorphisms and the Functional Calculus
    Pacific Journal of Mathematics ALGEBRA HOMOMORPHISMS AND THE FUNCTIONAL CALCULUS MARK PHILLIP THOMAS Vol. 79, No. 1 May 1978 PACIFIC JOURNAL OF MATHEMATICS Vol. 79, No. 1, 1978 ALGEBRA HOMOMORPHISMS AND THE FUNCTIONAL CALCULUS MARC THOMAS Let b be a fixed element of a commutative Banach algebra with unit. Suppose σ(b) has at most countably many connected components. We give necessary and sufficient conditions for b to possess a discontinuous functional calculus. Throughout, let B be a commutative Banach algebra with unit 1 and let rad (B) denote the radical of B. Let b be a fixed element of B. Let έ? denote the LF space of germs of functions analytic in a neighborhood of σ(6). By a functional calculus for b we mean an algebra homomorphism θr from έ? to B such that θ\z) = b and θ\l) = 1. We do not require θr to be continuous. It is well-known that if θ' is continuous, then it is equal to θ, the usual functional calculus obtained by integration around contours i.e., θ{f) = -±τ \ f(t)(f - ]dt for f eέ?, Γ a contour about σ(b) [1, 1.4.8, Theorem 3]. In this paper we investigate the conditions under which a functional calculus & is necessarily continuous, i.e., when θ is the unique functional calculus. In the first section we work with sufficient conditions. If S is any closed subspace of B such that bS Q S, we let D(b, S) denote the largest algebraic subspace of S satisfying (6 — X)D(b, S) = D(b, S)f all λeC.
    [Show full text]
  • The Weyl Functional Calculus F(4)
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOURNAL OF FUNCTIONAL ANALYSIS 4, 240-267 (1969) The Weyl Functional Calculus ROBERT F. V. ANDERSON Department of Mathematics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139 Communicated by Edward Nelson Received July, 1968 I. INTRODUCTION In this paper a functional calculus for an n-tuple of noncommuting self-adjoint operators on a Banach space will be proposed and examined. The von Neumann spectral theorem for self-adjoint operators on a Hilbert space leads to a functional calculus which assigns to a self- adjoint operator A and a real Borel-measurable function f of a real variable, another self-adjoint operatorf(A). This calculus generalizes in a natural way to an n-tuple of com- muting self-adjoint operators A = (A, ,..., A,), because their spectral families commute. In particular, if v is an eigenvector of A, ,..., A, with eigenvalues A, ,..., A, , respectively, and f a continuous function of n real variables, f(A 1 ,..., A,) u = f(b , . U 0. This principle fails when the operators don’t commute; instead we have the Uncertainty Principle, and so on. However, the Fourier inversion formula can also be used to define the von Neumann functional cakuius for a single operator: f(4) = (37F2 1 (Sf)(&) exp( -2 f A,) 4 E’ where This definition generalizes naturally to an n-tuple of operators, commutative or not, as follows: in the n-dimensional Fourier inversion 240 THE WEYL FUNCTIONAL CALCULUS 241 formula, the free variables x 1 ,..,, x, are replaced by self-adjoint operators A, ,..., A, .
    [Show full text]
  • Some C*-Algebras with a Single Generator*1 )
    transactions of the american mathematical society Volume 215, 1976 SOMEC*-ALGEBRAS WITH A SINGLEGENERATOR*1 ) BY CATHERINE L. OLSEN AND WILLIAM R. ZAME ABSTRACT. This paper grew out of the following question: If X is a com- pact subset of Cn, is C(X) ® M„ (the C*-algebra of n x n matrices with entries from C(X)) singly generated? It is shown that the answer is affirmative; in fact, A ® M„ is singly generated whenever A is a C*-algebra with identity, generated by a set of n(n + l)/2 elements of which n(n - l)/2 are selfadjoint. If A is a separable C*-algebra with identity, then A ® K and A ® U are shown to be sing- ly generated, where K is the algebra of compact operators in a separable, infinite- dimensional Hubert space, and U is any UHF algebra. In all these cases, the gen- erator is explicitly constructed. 1. Introduction. This paper grew out of a question raised by Claude Scho- chet and communicated to us by J. A. Deddens: If X is a compact subset of C, is C(X) ® M„ (the C*-algebra ofnxn matrices with entries from C(X)) singly generated? We show that the answer is affirmative; in fact, A ® Mn is singly gen- erated whenever A is a C*-algebra with identity, generated by a set of n(n + l)/2 elements of which n(n - l)/2 are selfadjoint. Working towards a converse, we show that A ® M2 need not be singly generated if A is generated by a set con- sisting of four elements.
    [Show full text]
  • And Varifold-Based Registration of Lung Vessel and Airway Trees
    Current- and Varifold-Based Registration of Lung Vessel and Airway Trees Yue Pan, Gary E. Christensen, Oguz C. Durumeric, Sarah E. Gerard, Joseph M. Reinhardt The University of Iowa, IA 52242 {yue-pan-1, gary-christensen, oguz-durumeric, sarah-gerard, joe-reinhardt}@uiowa.edu Geoffrey D. Hugo Virginia Commonwealth University, Richmond, VA 23298 [email protected] Abstract 1. Introduction Registration of lung CT images is important for many radiation oncology applications including assessing and Registering lung CT images is an important problem adapting to anatomical changes, accumulating radiation for many applications including tracking lung motion over dose for planning or assessment, and managing respiratory the breathing cycle, tracking anatomical and function motion. For example, variation in the anatomy during radio- changes over time, and detecting abnormal mechanical therapy introduces uncertainty between the planned and de- properties of the lung. This paper compares and con- livered radiation dose and may impact the appropriateness trasts current- and varifold-based diffeomorphic image of the originally-designed treatment plan. Frequent imaging registration approaches for registering tree-like structures during radiotherapy accompanied by accurate longitudinal of the lung. In these approaches, curve-like structures image registration facilitates measurement of such variation in the lung—for example, the skeletons of vessels and and its effect on the treatment plan. The cumulative dose airways segmentation—are represented by currents or to the target and normal tissue can be assessed by mapping varifolds in the dual space of a Reproducing Kernel Hilbert delivered dose to a common reference anatomy and com- Space (RKHS). Current and varifold representations are paring to the prescribed dose.
    [Show full text]
  • October 7, 2015 1 Adjoint of a Linear Transformation
    Mathematical Toolkit Autumn 2015 Lecture 4: October 7, 2015 Lecturer: Madhur Tulsiani The following is an extremely useful inequality. Prove it for a general inner product space (not neccessarily finite dimensional). Exercise 0.1 (Cauchy-Schwarz-Bunyakowsky inequality) Let u; v be any two vectors in an inner product space V . Then jhu; vij2 ≤ hu; ui · hv; vi Let V be a finite dimensional inner product space and let fw1; : : : ; wng be an orthonormal basis for P V . Then for any v 2 V , there exist c1; : : : ; cn 2 F such that v = i ci · wi. The coefficients ci are often called Fourier coefficients. Using the orthonormality and the properties of the inner product, we get ci = hv; wii. This can be used to prove the following Proposition 0.2 (Parseval's identity) Let V be a finite dimensional inner product space and let fw1; : : : ; wng be an orthonormal basis for V . Then, for any u; v 2 V n X hu; vi = hu; wii · hv; wii : i=1 1 Adjoint of a linear transformation Definition 1.1 Let V; W be inner product spaces over the same field F and let ' : V ! W be a linear transformation. A transformation '∗ : W ! V is called an adjoint of ' if h'(v); wi = hv; '∗(w)i 8v 2 V; w 2 W: n Pn Example 1.2 Let V = W = C with the inner product hu; vi = i=1 ui · vi. Let ' : V ! V be represented by the matrix A. Then '∗ is represented by the matrix AT . Exercise 1.3 Let 'left : Fib ! Fib be the left shift operator as before, and let hf; gi for f; g 2 Fib P1 f(n)g(n) ∗ be defined as hf; gi = n=0 Cn for C > 4.
    [Show full text]