Riesz-Fredhölm Theory

Total Page:16

File Type:pdf, Size:1020Kb

Riesz-Fredhölm Theory Riesz-Fredh¨olmTheory T. Muthukumar [email protected] Contents 1 Introduction1 2 Integral Operators1 3 Compact Operators7 4 Fredh¨olmAlternative 14 Appendices 18 A Ascoli-Arzel´aResult 18 B Normed Spaces and Bounded Operators 20 1 Introduction The aim of this lecture note is to show the existence and uniqueness of Fredh¨olmintegral operators of second kind, i.e., show the existence of a solution x of x − T x = y for any given y, in appropriate function spaces. 2 Integral Operators Let E be a compact subset of Rn and C(E) denote the space of complex valued continuous functions on E endowed with the uniform norm kfk1 = supx2E jf(x)j. Recall that C(E) is a Banach space with the uniform norm. 1 Definition 2.1. Any continuous function K : E × E ! C is called a con- tinuous kernel. Since K is continuous on a compact set, K is both bounded, i.e., there is a κ such that jK(x; y)j ≤ κ 8x; y 2 E and uniformly continuous. In particular, for each " > 0 there is a δ > 0 such that jK(x1; y) − K(x2; y)j ≤ " 8y 2 E whenever jx1 − x2j < δ. Example 2.1 (Fredh¨olmintegral operator). For any f 2 C(E), we define Z T (f)(x) = K(x; y)f(y) dy E where x 2 E and K : E × E ! R is a continuous function. For each " > 0, Z jT f(x1) − T f(x2)j ≤ jK(x1; y) − K(x2; y)jjf(y)j dy ≤ "kfk1jEj E whenever jx1 − x2j < δ. Thus, T f 2 C(E) and T defines a map from C(E) to C(E). Example 2.2 (Volterra integral operator). For any f 2 C[a; b], we define Z x T (f)(x) = K(x; y)f(y) dy a where x 2 [a; b] and K :[a; b] × [a; b] ! R is a continuous function. Note that, for x1; x2 2 [a; b], Z x1 Z x2 T f(x1) − T f(x2) = K(x1; y)f(y) dy − K(x2; y)f(y) dy a a Z x1 = [K(x1; y) − K(x2; y)]f(y) dy a Z x1 + K(x2; y)f(y) dy: x2 2 Therefore, for each " > 0, Z x1 jT f(x1) − T f(x2)j ≤ jK(x1; y) − K(x2; y)jjf(y)j dy a Z x1 + jK(x2; y)jjf(y)j dy x2 ≤ "kfk1(x1 − a) + κ(x1 − x2)kfk1 ≤ "kfk1(b − a) + κδkfk1 ≤ [(b − a) + κ]"kfk1 whenever jx1 − x2j < δ and, without loss of generality, we have assumed δ ≤ ". Thus, T f 2 C[a; b] and T defines a map from C[a; b] to C[a; b]. One can think of Volterra integral operator as a special case of Fredh¨olm integral operators by considering a K :[a; b] × [a; b] such that K(x; y) = 0 for y > x. Geometrically this means, in the square [a; b] × [a; b], K takes the value zero in the region above y = x line. Thus, Volterra integral operator is a Fredh¨olmoperator for a K which may be discontinuous on the line y = x in the square. In fact, this particular K is a special case of weakly singular kernel. Definition 2.2. Let E ⊂ Rn be a compact subset. A function K : E×E ! C is said to be weakly singular kernel if it is continuous for all x; y 2 E such that x 6= y and there exist positive constants M and α 2 (0; n) such that jK(x; y)j ≤ Mjx − yjα−n 8x; y 2 E; x 6= y: Theorem 2.3. Let E ⊂ Rn be a compact subset with non-empty interior. Let K : E ×E ! C be a continuous function. Then the operator T : C(E) ! C(E), defined as, Z (T f)(x) := K(x; y)f(y) dy; for each x 2 E E is a bounded linear operator, i.e., is in B(C(E)) and Z kT k = sup jK(x; y)j dy: x2E E 3 Proof. The fact that T is linear is obvious. For each f 2 C(E) with kfk1 ≤ 1 and x 2 E, we have Z j(T f)xj ≤ jK(x; y)j dy: E Thus, Z kT k = sup kT fk1 ≤ sup jK(x; y)j dy: f2C(E) x2E E kfk1≤1 It now only remains to show the other inequality. Since K is continuous, there is a x0 2 E such that Z Z jK(x0; y)j dy = max jK(x; y)j dy: E x2E E For each " > 0, choose g" 2 C(E) as K(x0; y) g"(y) := ; for y 2 E: jK(x0; y)j + " Then kg"k1 ≤ 1 and Z 2 jK(x0; y)j kT g"k1 ≥ j(T g")(x0)j = dy E jK(x0; y)j + " Z 2 2 Z jK(x0; y)j − " ≥ dy = jK(x0; y)j dy − "jEj: E jK(x0; y)j + " E Hence Z kT k = sup kT fk1 ≥ kT g"k1 ≥ jK(x0; y)j dy − "jEj; f2C(E) E kfk1≤1 and, since " is arbitrary, we have Z Z kT k ≥ jK(x0; y)j dy = max jK(x; y)j dy: E x2E E For any operator T : X ! X, one can define the iterated operators T k : X ! X as the k-times composition of T with itself. 4 Theorem 2.4 (Neumann Series). Let X be a Banach space, T 2 B(X) with kT k < 1 (contraction) and I : X ! X be the identity operator. Then the map I − T : X ! X is invertible given by 1 X (I − T )−1 = T k k=0 and the inverse is bounded 1 k(I − T )−1k ≤ : 1 − kT k Proof. Since X is a Banach space, by Theorem B.12, B(X) is a Banach space. P1 k Thus, owing to Theorem B.4, to show the convergence of the series k=0 T it is enough to show that it is absolutely convergent. Note that 1 m m 1 X X X X kT kk = lim kT kk ≤ lim kT kk = kT kk: m!1 m!1 k=0 k=0 k=0 k=0 Since kT k < 1, the geometric series converges to (1 − kT k)−1 and hence the P1 k series k=0 T is absolutely convergent and, thus, convergence in B(X). Let S : X ! X be the limit of the series, i.e., 1 X S := T k: k=0 Note that kSk ≤ (1 − kT k)−1, hence, is a bounded linear operator on X, i.e., S 2 B(X). It only remains to show that S is the inverse of I − T . Note that m X (I − T )S = (I − T ) lim T k = lim (I − T )(I + T + ::: + T m) m!1 m!1 k=0 = lim (I − T m+1) = I: m!1 The last equality is due to the fact that the sequence T m+1 converges to 0 in m+1 m+1 m+1 B(X) because kT k ≤ kT k and, since kT k < 1, limm!1 kT k ! 0. Similarly, m X S(I − T ) = lim T k(I − T ) = lim (I + T + ::: + T m)(I − T ) m!1 m!1 k=0 = lim (I − T m+1) = I: m!1 5 If x 2 X is the solution of (I − T )x = y, for any given y 2 X, then x can be computed as 1 m −1 X k X k x = (I − T ) y = T y = lim T y = lim xm: m!1 m!1 k=0 k=0 Note that xm+1 = T xm + y, for m ≥ 0. Theorem 2.5. Let X be a Banach space, T 2 B(X) with kT k < 1 (con- traction) and I : X ! X be the identity operator. For any given y 2 X and arbitrary x0 2 X the sequence xm+1 := T xm + y for m = 0; 1; 2;::: converges to a unique solution x 2 X of (I − T )x = y. 2 Proof. Note that x1 = T x0 + y, x2 = T x1 + y = T x0 + (I + T )y. Thus, by induction, for m = 1; 2;:::, m−1 m X k xm = T x0 + T y: k=0 Hence, 1 X k −1 lim xm = T y = (I − T ) y: m!1 k=0 Corollary 2.6. Let K be a continuous kernel satisfying Z max jK(x; y)j dy < 1: x2E E Then, for each g 2 C(E), the integral equation of the second kind Z f(x) − K(x; y)f(y) dy = g(x) x 2 E E has a unique solution f 2 C(E). Further, for any f0 2 C(E), the sequence Z fm+1(x) := K(x; y)fm(y) dy + g(x); m = 0; 1; 2;::: E converges uniformly to f 2 C(E). 6 3 Compact Operators Theorem 3.1. A subset of a normed space is compact iff it is sequentially compact. Definition 3.2. A subset E of a normed space X is said to be relatively compact if E, the closure of E, is compact in X. Theorem 3.3. Any bounded and finite dimensional subset of a normed space is relatively compact. Definition 3.4. Let X and Y be normed spaces. A linear operator T : X ! Y is said to be compact if T (E) is relatively compact in Y , for every bounded subset E ⊂ X.
Recommended publications
  • Chapter 4 the Lebesgue Spaces
    Chapter 4 The Lebesgue Spaces In this chapter we study Lp-integrable functions as a function space. Knowledge on functional analysis required for our study is briefly reviewed in the first two sections. In Section 1 the notions of normed and inner product spaces and their properties such as completeness, separability, the Heine-Borel property and espe- cially the so-called projection property are discussed. Section 2 is concerned with bounded linear functionals and the dual space of a normed space. The Lp-space is introduced in Section 3, where its completeness and various density assertions by simple or continuous functions are covered. The dual space of the Lp-space is determined in Section 4 where the key notion of uniform convexity is introduced and established for the Lp-spaces. Finally, we study strong and weak conver- gence of Lp-sequences respectively in Sections 5 and 6. Both are important for applications. 4.1 Normed Spaces In this and the next section we review essential elements of functional analysis that are relevant to our study of the Lp-spaces. You may look up any book on functional analysis or my notes on this subject attached in this webpage. Let X be a vector space over R. A norm on X is a map from X ! [0; 1) satisfying the following three \axioms": For 8x; y; z 2 X, (i) kxk ≥ 0 and is equal to 0 if and only if x = 0; (ii) kαxk = jαj kxk, 8α 2 R; and (iii) kx + yk ≤ kxk + kyk. The pair (X; k·k) is called a normed vector space or normed space for short.
    [Show full text]
  • 01B. Norms and Metrics on Vectorspaces 1. Normed Vector Spaces
    (October 2, 2018) 01b. Norms and metrics on vectorspaces Paul Garrett [email protected] http:=/www.math.umn.edu/egarrett/ [This document is http:=/www.math.umn.edu/egarrett/m/real/notes 2018-19/01b Banach.pdf] 1. Normed vector spaces 2. Inner-product spaces and Cauchy-Schwarz-Bunyakowsky 3. Normed spaces of continuous linear maps 4. Dual spaces of normed spaces 5. Extension by continuity Many natural real or complex vector spaces of functions, such as Co[a; b] and Ck[a; b], have (several!) natural metrics d(; ) ncoming from norms j · j by the recipe d(v; w) = jv − wj A real or complex vector space with an inner product h; i always has an associated norm 1 jvj = hv; vi 2 If so, the vector space has much additional geometric structure, including notions of orthogonality and projections. However, very often the natural norm on a vector space of functions does not come from an inner product, which creates complications. When the vector space is complete with respect to the associated metric, the vector space is a Banach space. Abstractly, Banach spaces are less convenient than Hilbert spaces (complete inner-product spaces), and normed spaces are less convenient than inner-product spaces. 1. Normed vector spaces A real or complex [1] vectorspace V with a non-negative, real-valued function, the norm, j · j : V −! R with properties jx + yj ≤ jxj + jyj (triangle inequality) jαxj = jαj · jxj (α real/complex, x 2 V ) jxj = 0 ) x = 0 (positivity) is a normed (real or complex) vectorspace, or simply normed space.
    [Show full text]
  • Arxiv:Math/0606537V2 [Math.CA] 7 Dec 2007 Rpitdcme ,20.T Perin Appear to 2007
    Preprint December 7, 2007. To appear in Real Analysis Exchange. The distributional Denjoy integral Erik Talvila1 Department of Mathematics and Statistics University College of the Fraser Valley Abbotsford, BC Canada V2S 7M8 [email protected] Abstract. Let f be a distribution (generalised function) on the real line. If there is a con- tinuous function F with real limits at infinity such that F ′ = f (distributional derivative) ∞ then the distributional integral of f is defined as f = F ( ) F ( ). It is shown that −∞ ∞ − −∞ this simple definition gives an integral that includesR the Lebesgue and Henstock–Kurzweil integrals. The Alexiewicz norm leads to a Banach space of integrable distributions that is isometrically isomorphic to the space of continuous functions on the extended real line with uniform norm. The dual space is identified with the functions of bounded variation. Basic properties of integrals are established using elementary properties of distributions: integration by parts, H¨older inequality, change of variables, convergence theorems, Banach lattice structure, Hake theorem, Taylor theorem, second mean value theorem. Applications are made to the half plane Poisson integral and Laplace transform. The paper includes a short history of Denjoy’s descriptive integral definitions. Distributional integrals in Eu- clidean spaces are discussed and a more general distributional integral that also integrates Radon measures is proposed. 2000 subject classification: 26A39, 46B42, 46E15, 46F05, 46G12 Key words: distributional Denjoy integral; continuous primitive integral; Henstock– Kurzweil integral; Schwartz distributions; Alexiewicz norm; Banach lattice. arXiv:math/0606537v2 [math.CA] 7 Dec 2007 1 Introduction We are fortunate to live in a richly diverse universe in which there are many integrals and many interesting ways of defining these integrals.
    [Show full text]
  • Topological Vector Spaces and Continuous Linear Functionals
    CHAPTER III TOPOLOGICAL VECTOR SPACES AND CONTINUOUS LINEAR FUNCTIONALS The marvelous interaction between linearity and topology is introduced in this chapter. Although the most familiar examples of this interaction may be normed linear spaces, we have in mind here the more subtle, and perhaps more important, topological vector spaces whose topologies are defined as the weakest topologies making certain collections of functions continuous. See the examples in Exercises 3.8 and 3.9, and particularly the Schwartz space S discussed in Exercise 3.10. DEFINITION. A topological vector space is a real (or complex) vector space X on which there is a Hausdorff topology such that: (1) The map (x, y) → x+y is continuous from X×X into X. (Addition is continuous.) and (2) The map (t, x) → tx is continuous from R × X into X (or C × X into X). (Scalar multiplication is continuous.) We say that a topological vector space X is a real or complex topolog- ical vector space according to which field of scalars we are considering. A complex topological vector space is obviously also a real topological vector space. A metric d on a vector space X is called translation-invariant if d(x+ z, y + z) = d(x, y) for all x, y, z ∈ X. If the topology on a topological vector space X is determined by a translation-invariant metric d, we call X (or (X, d)) a metrizable vector space. If x is an element of a metrizable vector space (X, d), we denote by B(x) the ball of radius 43 44 CHAPTER III around x; i.e., B(x) = {y : d(x, y) < }.
    [Show full text]
  • Functional Analysis
    Functional Analysis Lyudmila Turowska Department of Mathematical Sciences Chalmers University of Technology and the University of Gothenburg These lectures concern the theory of normed spaces and bounded linear oper- ators between normed spaces. We shall define these terms and study their prop- erties. You can consult the following books. 1. G. Folland: Real Analysis. Modern Techniques and their Applications, John Wiley & Sons, 1999, Chapters 5-7 and parts of Chapter 4. 2 1 Vector Spaces Definition 1.1. A complex vector space (or a vector space over the field C) (or a complex linear space) is a set V with addition V × V ! V; (x; y) 7! x + y the sum of x and y; and scalar multiplication C × V ! V; (λ, x) 7! λx the product of λ and x; satisfying the following rules: x + (y + z) = (x + y) + z 8x; y; z 2 V; x + y = y + x 8x; y 2 V; 90 2 V : 0 + x = x 8x 2 V; 8x 2 V 9 − x 2 V : x + (−x) = 0; λ(x + y) = λx + λy 8x; y 2 V; 8λ 2 C; (λ + µ)x = λx + µx 8x 2 V; 8λ, µ 2 C; λ(µx) = (λµ)x 8x 2 V; 8λ, µ 2 C; 1x = x 8x 2 V: It is easy to see that the usual arithmetic holds in V . Remark 1.2. Similarly, we can define a real vector space by replacing C with R in the definition. n Examples 1.3. 1. C = f(x1; : : : ; xn): xi 2 C 8ig, the complex vector space of n-tuples of complex numbers with coordinatewise addition n (x1; : : : ; xn) + (y1; : : : ; yn) = (x1 + y1; : : : ; xn + yn); (x1; : : : ; xn); (y1; : : : ; yn) 2 C and coordinatewise scalar multiplication n λ(x1; : : : ; xn) = (λx1; : : : ; λxn); λ 2 C; (x1; : : : ; xn) 2 C : 2.
    [Show full text]
  • On Local Finite Dimensional Approximation of C^*-Algebras
    pacific journal of mathematics Vol. 181, No. 1, 1997 ON LOCAL FINITE DIMENSIONAL APPROXIMATION OF C∗-ALGEBRAS Sorin Popa To Ed Effros, on the occasion of his 60th birthday. We prove a local finite dimensional approximation result for unital C*-algebras that are simple, have a trace, contain sufficiently many projections and are quasidiagonal. 1. Introduction. Recall that a C∗-algebra A is called quasidiagonal if it can be represented faithfully on a Hilbert space H such that there exist finite dimensional vector subspaces Hi ⊂H, with Hi %Hand lim k[projH ,x]k =0,∀x∈A(see i i e.g., [V2]). By Voiculescu’s theorem ([V1]), if A is simple then this property doesn’t in fact depend on the Hilbert space on which A is represented, so that it can be reformulated in terms of the following local finite dimensional approximation property: A simple C∗-algebra is quasidiagonal if for any representation of A on a Hilbert space H, any finite set F ⊂ A and any ε>0, there exists a finite dimensional vector subspace 0 6= H0 ⊂Hsuch that k[proj ,x]k<ε,∀x∈F. H0 The result that we obtain in this paper gives an alternative, more intrinsic characterization of quasidiagonality, for unital simple C∗-algebras A that have sufficiently many projections (abreviated s.m.p.), i.e., satisfy (1.1) “There exists an abelian subalgebra A0 ⊂ A such that for any x = ∗ x ∈ A and any ε>0, there exist a partition of 1A with projections in A0, {pj}1≤j≤n ⊂P(A0), and scalars {αj}1≤j≤n ⊂ C, such that kpjxpj − αjpjk <ε,1≤j≤n.” Note that in case A is separable this condition is easily seen to be implied by: (1.10)“Foranyx=x∗ ∈A,anyp∈P(A) and any ε>0, there exist a partition of p with projections in A, {qi}1≤i≤k ⊂P(pAp), and scalars {αi}1≤i≤k ⊂ C such that kqixqi − αiqik <ε,1≤i≤k.” 141 142 SORIN POPA and this last property is clearly satisfied by C∗-algebras with real rank zero (see 2.2).
    [Show full text]
  • Banach Spaces
    90 Chapter 5 Banach Spaces Many linear equations may be formulated in terms of a suitable linear operator acting on a Banach space. In this chapter, we study Banach spaces and linear oper- ators acting on Banach spaces in greater detail. We give the definition of a Banach space and illustrate it with a number of examples. We show that a linear operator is continuous if and only if it is bounded, define the norm of a bounded linear op- erator, and study some properties of bounded linear operators. Unbounded linear operators are also important in applications: for example, differential operators are typically unbounded. We will study them in later chapters, in the simpler context of Hilbert spaces. 5.1 Banach spaces A normed linear space is a metric space with respect to the metric d derived from its norm, where d(x; y) = kx − yk. Definition 5.1 A Banach space is a normed linear space that is a complete metric space with respect to the metric derived from its norm. The following examples illustrate the definition. We will study many of these examples in greater detail later on, so we do not present proofs here. Example 5.2 For 1 ≤ p < 1, we define the p-norm on Rn (or Cn) by p p p 1=p k(x1; x2; : : : ; xn)kp = (jx1j + jx2j + : : : + jxnj ) : For p = 1, we define the 1, or maximum, norm by k(x1; x2; : : : ; xn)k1 = max fjx1j; jx2j; : : : ; jxnjg : Then Rn equipped with the p-norm is a finite-dimensional Banach space for 1 ≤ p ≤ 1.
    [Show full text]
  • Stat 8112 Lecture Notes Weak Convergence in Metric Spaces Charles J
    Stat 8112 Lecture Notes Weak Convergence in Metric Spaces Charles J. Geyer January 23, 2013 1 Metric Spaces Let X be an arbitrary set. A function d : X × X ! R is called a metric if it satisfies the following properties (i) d(x; y) = d(y; x), (ii) d(x; y) ≥ 0, (iii) d(x; y) = 0 if and only if x = y, (iv) d(x; z) ≤ d(x; y) + d(y; z), it being understood that these are identities holding for all values of the free variables. The identity (iv) is called the triangle inequality. A set equipped with a metric is called a metric space. The most familiar metric spaces are the real numbers R with the metric d(x; y) = jx − yj d and finite-dimensional vector spaces R with the metric d(x; y) = kx − yk (1) where the norm k · k can be defined in a number of different ways d X kxk = jxij (2a) i=1 d !1=2 X 2 kxk = jxij (2b) i=1 kxk = max jxij (2c) 1≤i≤d which are called the L1 norm, the L2 norm (or the Euclidean norm), and the L1 norm (or the supremum norm), respectively. 1 2 Topology of Metric Spaces A sequence xn in a metric space X converges to x if d(xn; x) ! 0; as n ! 1: For any x and " > 0, the set B"(x) = f y 2 X : d(x; y) < " g (3) is called the open ball centered at x having radius ". A subset of a metric space is called open if it is a union of open balls or if it is empty.
    [Show full text]
  • Norm (Mathematics) - Wikipedia, the Free Encyclopedia
    Norm (mathematics) - Wikipedia, the free encyclopedia http://en.wikipedia.org/wiki/Norm_(mathematics) Norm (mathematics) From Wikipedia, the free encyclopedia In linear algebra, functional analysis and related areas of mathematics, a norm is a function that assigns a strictly positive length or size to all vectors in a vector space, other than the zero vector. A seminorm , on the other hand, is allowed to assign zero length to some non-zero vectors. A simple example is the 2-dimensional Euclidean space R2 equipped with the Euclidean norm. Elements in this vector space (e.g., (3, 7)) are usually drawn as arrows in a 2-dimensional cartesian coordinate system starting at the origin (0, 0). The Euclidean norm assigns to each vector the length of its arrow. Because of this, the Euclidean norm is often known as the magnitude. A vector space with a norm is called a normed vector space. Similarly, a vector space with a seminorm is called a seminormed vector space. Contents 1 Definition 2 Examples 2.1 Euclidean norm 2.1.1 Euclidean norm of a complex number 2.2 Taxicab norm or Manhattan norm 2.3 p-norm 2.4 Maximum norm (special case of: infinity norm, uniform norm, or supremum norm) 2.5 Zero norm 2.5.1 Hamming distance of a vector from zero 2.6 Other norms 2.7 Infinite-dimensional case 3 Properties 4 Classification of seminorms: Absolutely convex absorbing sets 5 See also 6 Notes 7 References Definition Given a vector space V over a subfield F of the complex numbers a norm on V is a function p: V → F with the following properties: [1] For all a ∈ F and all u, v ∈ V, 1.
    [Show full text]
  • MTH 303: Real Analysis
    MTH 303: Real Analysis Semester 1, 2013-14 1. Real numbers 1.1. Decimal and ternary expansion of real numbers. (i) [0; 1] and R are uncountable. (ii) Cantor set is an uncounbtable set of measure zero. 1.2. Order properties of real numbers. (i) Least upper bound axiom. (ii) Every subset of R that has a lower bound also has a greatest lower bound. (iii) Archimedean property. (iv) Density of Q. 2. Sequences of real numbers 2.1. Limits of sequences. (i) Sequences and their limits with examples. (ii) A sequence in R has at most one limit. (iii) A sequence converges if and only if its tail converges. (iv) Let (xn) be a sequence of real numbers and let x 2 R. If (an) is a sequence of positive reals with lim(an) = 0 and if we have a constant C > 0 and some m 2 N such that jxn − xj ≤ Can, for n ≥ m, then lim(xn) = x. 2.2. Limit Therorems. (i) Bounded sequences with examples. (ii) A convergent sequence of real numbers is bounded. (iii) Convergence of the sum, difference, product, and quotient of sequences. (iv) The limit of a non-negative sequence is non-negative. (v) Squeeze Theorem: Suppose that X = (xn), Y = (yn), and Z = (zn) are sequences such that xn ≤ yn ≤ zn, 8 n, and lim(xn) = lim(zn). Then Y converges and lim(xn) = lim(yn) = lim(zn). (vi) Let X = (x ) be a sequence of positive reals such that L = lim xn+1 . n xn If L < 1, then X converges and lim(xn) = 0.
    [Show full text]
  • Functional Analysis
    Functional Analysis Irena Swanson, Reed College, Spring 2011 This is work in progress. I am still adding, subtracting, modifying. Any comments, solutions, corrections are welcome. Table of contents: Section 1: Overview, background 1 Section 2: Definition of Banach spaces 2 Section 3: Examples of Banach spaces 2 Section 4: Lp spaces 6 Section 5: L∞ spaces 9 Section 6: Some spaces that are almost Banach, but aren’t 10 Section 7: Normed vector spaces are metric spaces 11 Section 8: How to make new spaces out of existent ones (or not) 13 Section 9: Finite-dimensional vector spaces are special 18 Section 10: Examples of (infinite dimensional) subspaces 20 Section 11: Sublinear functional and the Hahn–Banach Theorem 22 Section 12: A section of big theorems 27 Section 13: Hilbert spaces 35 Section 14: Perpendicularity 37 Section 15: The Riesz Representation Theorem 41 Section 16: Orthonormal bases 41 Section 17: Section I.5 in the book 46 Section 18: Adjoints of linear transformations 47 Section 19: Self-adjoint operators 49 Section 20: Spectrum 51 Section 21: Holomorphic Banach space-valued functions 55 Appendix A: Some topology facts 59 Appendix B: Homework and exam sets 60 1 1 Overview, background Throughout, F will stand for either R or C. Thus F is a complete (normed) field. [Conway’s book assumes that all topological spaces are Hausdorff. Let’s try to not impose that unless needed.] Much of the material and inspiration came from Larry Brown’s lectures on functional analysis at Purdue University in the 1990s, and some came from my Reed thesis 1987.
    [Show full text]
  • G14FUN Functional Analysis
    G14FUN Functional Analysis Dr J. F. Feinstein March 13, 2007 Contents 1 Complete metric spaces 2 1.1 Completeness and complete metrizability . 2 1.2 The Baire Category Theorem . 2 2 Infinite products and Tychonoff’s Theorem 5 2.1 Infinite products . 5 2.2 Tychonoff’s Theorem . 7 3 Basic Functional Analysis 9 3.1 Normed spaces and Banach spaces . 9 3.2 Equivalence of norms . 11 3.3 Linear maps . 12 3.4 Sequence spaces . 13 3.5 Isomorphisms . 14 3.6 Sums and quotients of vector spaces . 15 3.7 Dual spaces . 17 3.8 Extensions of linear maps . 19 3.9 Completions, quotients and Riesz’s lemma . 21 3.9.1 Completions . 21 3.9.2 Quotient spaces and Riesz’s lemma . 21 4 The weak-* topology and the Banach-Alaoglu Theorem 23 4.1 The weak-* topology . 23 4.2 The Banach-Alaoglu Theorem . 24 5 Open mappings and their applications 25 5.1 Open mappings and the Open Mapping Lemma . 25 5.2 Urysohn’s Lemma and the Tietze Extension Theorem . 26 5.3 The Open Mapping Theorem and the Banach Isomorphism Theorem . 27 6 The Banach-Steinhaus Theorem (Uniform Boundedness) 29 1 1 Complete metric spaces 1.1 Completeness and complete metrizability We begin by recalling the definition of complete. Definition 1.1 A metric space (X, d) is complete if every Cauchy sequence in X converges in X. Otherwise X is incomplete: this means that there is at least one Cauchy sequence in X which does not converge in X. We also describe the metric d as either a complete metric or an incomplete metric, accordingly.
    [Show full text]