Skip to main content

Banach Spaces

  • Chapter
  • First Online:
Real Analysis

Part of the book series: Birkhäuser Advanced Texts Basler Lehrbücher ((BAT))

  • 4428 Accesses

Abstract

Let X be a vector space and let \(\varTheta \) be its zero element.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 49.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 49.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Compare with Proposition 10.1 of Chap. 6.

  2. 2.

    Compare with Proposition 16.1 of Chap. 6.

  3. 3.

    \(\mathcal {S}\) need not be countable.

  4. 4.

    These remarks on existence and nonexistence of unbounded linear functionals in metric spaces were suggested by Allen Devinatz\({}^\dag \).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emmanuele DiBenedetto .

Appendices

Problems and Complements

1c Normed Spaces

1.1.:

Let E be a bounded, open subset of \(\mathbb {R}^N\). The space \(C({\bar{E}})\) endowed with the norm of \(L^p(E)\) is not a Banach space.

1.2.:

Every normed space is homeomorphic to its open unit ball.

1.3.:

A normed space \(\{X;\Vert \cdot \Vert \}\) is complete if and only if the intersection of a countable family of nested, closed balls is nonempty.

1.4.:

The \(L^2\)-norm and the sup-norm on C[0, 1] are not equivalent. In particular C[0, 1] is not complete in \(L^p[0,1]\) for all \(1\le p <\infty \). The norms of \(L^p(E)\) and \(L^q(E)\) for \(1\le q<p<\infty \), are not equivalent.

The next proposition provides a criterion for a normed space to be a Banach space.

Proposition 1.1c

A normed space \(\{X;\Vert \cdot \Vert \}\) is complete if and only if every absolutely convergent series converges to an element of X.

Proof

Let \(\{X;\Vert \cdot \Vert \}\) be complete and let \(\mathop {\textstyle {\sum }}\limits x_n\) be absolutely convergent in \(\{X;\Vert \cdot \Vert \}\), so that \(\mathop {\textstyle {\sum }}\limits \Vert x_n\Vert \le M\) for some \(M>0\). Then \(\{\sum _{j=1}^n x_j\}\) is Cauchy and hence convergent to some \(x\in X\).

Conversely, let \(\{x_n\}\) be Cauchy in \(\{X;\Vert \cdot \Vert \}\), so that for each \(j\in \mathbb {N}\) there exists \(n_j\) such that

$$\begin{aligned} \Vert x_n-x_m\Vert \le 2^{-j}\quad \text { for }\> n,m\ge n_j. \end{aligned}$$

Set

$$\begin{aligned} y_j=x_{n_{j+1}}-x_{n_j}. \end{aligned}$$

The series \(\mathop {\textstyle {\sum }}\limits y_j\) is absolutely convergent and we let x denote its limit. Thus the subsequence \(\{x_{n_j}\}\subset \{x_n\}\) converges to x. Since \(\{x_n\}\) is Cauchy, the whole sequence converges to x. \(\blacksquare \)

In the context of \(L^p(E)\) the criterion has been used in the proof of Theorem 5.1 of Chap. 6.

Proposition 1.2c

Every norm p on \(\mathbb {R}^N\) is equivalent to the Euclidean norm \(\Vert \cdot \Vert \).

Remark 1.1c

The statement is a particular case of Proposition 12.1 of Chap. 2. The proof below is more direct using that the topology of \(\mathbb {R}^N\) is generated by a norm \(p(\cdot )\).

Proof

(of Proposition 1.2c) Let \(\{\mathbf {e}_1,\dots ,\mathbf {e}_N\}\) be a basis of \(\mathbb {R}^N\). Then

$$\begin{aligned} x=\mathop {\textstyle {\sum }}\limits _{j=1}^N c_j \mathbf {e}_j\quad \Longrightarrow \quad p(x)\le \mathop {\textstyle {\sum }}\limits _{j=1}^N |c_j| p(\mathbf {e}_j) \le C\Vert x\Vert \end{aligned}$$

for a constant C independent of x. This implies that

$$\begin{aligned} |p(x)-p(y)|\le p(x-y)\le C\Vert x-y\Vert \quad \text { for all }\> x,y\in \mathbb {R}^N. \end{aligned}$$

Thus \(p(\cdot )\) is continuous in the topology of the Euclidean norm. Its restriction to the Euclidean unit sphere \(S_1\) is a continuous function pointwise strictly bounded below in \(S_1\). Since the \(S_1\) is compact in \(\mathbb {R}^N\) with its Euclidean topology, \(p(\cdot )\) attains its minimum \(c>0\) there. From this

$$\begin{aligned} p(x)=p\Big (\Vert x\Vert \frac{x}{\Vert x\Vert }\Big ) =\Vert x\Vert p\Big (\frac{x}{\Vert x\Vert }\Big )\ge c\Vert x\Vert \quad \text { for all }\> x\in \mathbb {R}^N-\{0\}. \end{aligned}$$

\(\blacksquare \)

1.1 1.1c Semi-Norms and Quotients

1.5.:

The quantities \(\mathcal {V}_f[a,b]\) in (1.2), and \([f]_{\alpha }\) in (1.3) are semi-norms in their respective spaces.

1.6.:

Let \(\{p_{\alpha }\}\) be a collection of semi-norms on X such that

$$\begin{aligned} p_\infty (x)=\sup _{\alpha } p_\alpha (x) <\infty \quad \text { for all }\> x\in X. \end{aligned}$$

Then \(p_\infty (\cdot )\) defines a semi-norm on X.

1.7.:

Let \(\{p_j\}\) for \(j=1,\dots ,n\) be a finite collection of semi-norms on X. Then

$$\begin{aligned} q_1(x)=\mathop {\textstyle {\sum }}\limits _{j=1}^n p_j(x);\quad q_2(x)=\sqrt{\mathop {\textstyle {\sum }}\limits _{j=1}^n p_j^2(x)};\quad q_\infty (x)=\max _{1\le j\le n} p_j(x), \end{aligned}$$

are also semi-norms.

1.8.:

Prove that a semi-norm p on X is convex. Conversely a convex function \(f:X\rightarrow \mathbb {R}^+\) which is homogeneous of order 1 in X, defines a semi-norm in X. Hint: Homogeneous of order \(\alpha \in \mathbb {R}\) means that \(f(tx)=|t|^{\alpha } f(x)\) for all \(t\in \mathbb {R}\).

2c Finite and Infinite Dimensional Normed Spaces

2.1.:

An infinite dimensional Banach space \(\{X;\Vert \cdot \Vert \}\) cannot have a countable Hamel basis (16.9 of the Complements of Chap. 2).

2.2.:

Let \(\ell _o\) be the collection of all sequences of real numbers \(\{c_n\}\) with only finitely many nonzero elements. There is no norm on \(\ell _o\) by which \(\ell _o\) would be a Banach space. See also § 9.6c of the Complements of Chap. 2.

2.3.:

\(L^p(E)\) and \(\ell _p\) are of infinite dimension for all \(1\le p\le \infty \) and their dimension is larger than \(\aleph _o\).

2.4.:

Let \(C^1[0,1]\) be the collections of all continuously differentiable functions in [0, 1] with norm

$$\begin{aligned} C^1[0,1]\ni f\rightarrow \Vert f\Vert _{C^1[0,1]}=\sup _{[0,1]}|f| +\sup _{[0,1]} |f^\prime |. \end{aligned}$$

Let \(X_o\) be a closed subspace of \(C^1[0,1]\) which is also closed in \(L^2[0,1]\). Prove that

i.:

The norms \(\Vert \cdot \Vert _{C^1[0,1]}\) and \(\Vert \cdot \Vert _{L^2[0,1]}\) are equivalent on \(X_o\), i.e., there exists two positive constants \(c\le C\) such that

$$\begin{aligned} c\,\Vert f\Vert _{C^1[0,1]}\le \Vert f\Vert _{L^2[0,1]}\le C\,\Vert f\Vert _{C^1[0,1]}. \end{aligned}$$
ii.:

\(X_o\) is a finite dimensional subspace of \(L^2[0,1]\).

2.5.:

The dimension of C[0, 1] and BV[0, 1] is uncountably infinite. The collection \(\{x^n\}\) is an infinite set of linearly independent elements of C[0, 1] and BV[0, 1].

2.6.:

An infinite dimensional Banach space \(\{X;\Vert \cdot \Vert \}\) has an infinite dimensional non-closed subspace. Fix a Hamel basis \(\{x_\alpha \}\), for X, where \(\alpha \) is an uncountable index, select an infinite, countable collection \(\{x_n\}\subset \{x_\alpha \}\) of linearly independent elements and set \(Y=\{x_n\}\). Then Y is non-closed, for otherwise it would be an infinite dimensional Banach space with a countable Hamel basis.

3c Linear Maps and Functionals

3.1.:

A linear map T from \(\{X;\Vert \cdot \Vert _X\}\) into \(\{Y;\Vert \cdot \Vert _Y\}\) is continuous if and only if it maps sequences \(\{x_n\}\) converging to \(\varTheta _X\) into bounded sequences of \(\{Y;\Vert \cdot \Vert _Y\}\).

3.2.:

Two normed spaces \(\{X;\Vert \!\cdot \!\Vert _1\}\) and \(\{X;\Vert \!\cdot \!\Vert _2\}\) are homeomorphic if and only if there exist positive constants \(0<c_o\le 1\le c_1\) such that

$$\begin{aligned} c_o\Vert x\Vert _1\le \Vert x\Vert _2\le c_1\Vert x\Vert _1\quad \text { for all }\> x\in X. \end{aligned}$$
3.3.:

A linear functional on a finite dimensional normed space is continuous.

3.4.:

Let E be a bounded, open subset of \(\mathbb {R}^N\). A linear map \(T:C({\bar{E}})\rightarrow \mathbb {R}\) is a positive functional, if \(T(f)\ge 0\) whenever \(f\ge 0\). A positive linear functional on \(C({\bar{E}})\) is bounded. Thus positivity implies continuity. Moreover any two of the conditions

$$\begin{aligned} \text {(i)}~~\ \Vert T\Vert =1\qquad \text {(ii)}~~\ T(1)=1\qquad \text {(iii)}~~\ T\ge 0 \end{aligned}$$

implies the remaining one.

3.5.:

Let \(\{X;\Vert \cdot \Vert \}\) be an infinite dimensional Banach space. There exists a discontinuous, linear map \(T:X\rightarrow X\).

Having fixed a Hamel basis \(\{x_\alpha \}\) for X, after a possible renormalization, we may assume that \(\Vert x_\alpha \Vert =1\) for all \(\alpha \). Every element \(x\in X\) can be represented in a unique way, as the finite linear combination of elements of \(\{x_\alpha \}\), i.e., for every \(x\in X\) there exists a unique m-tuple of real numbers \(\{c_1,\dots ,c_m\}\) such that

$$\begin{aligned} x=\mathop {\textstyle {\sum }}\limits _{j=1}^m\,c_jx_{\alpha _j}. \end{aligned}$$
(3.1c)

Since X is of infinite dimension, the index \(\alpha \) ranges over some set A such that \((A)\ge (\mathbb {N})\). Out of \(\{x_\alpha \}\) select a countable collection \(\{x_n\}\subset \{x_\alpha \}\). Then set

$$\begin{aligned} T(x_n)=n x_n \quad \text { and }\quad T(x_\alpha )=\varTheta \quad \text { if }\>\alpha \notin \mathbb {N}. \end{aligned}$$

For \(x\in X\), having determined its representation (3.1c), set also

$$\begin{aligned} T(x)=\mathop {\textstyle {\sum }}\limits _{j=1}^m\,c_j T(x_{\alpha _j}). \end{aligned}$$

In view of the uniqueness of the representation (3.1c), this defines a linear map from X into X. Such a map, however, is discontinuous since \(\Vert T(x_n)\Vert =n\Vert x_n\Vert \rightarrow \infty \) as \(n\rightarrow \infty \).

3.6.:

Let \(\{X;\Vert \cdot \Vert \}\) be an infinite dimensional Banach space. There exists a discontinuous, linear functional \(T:X\rightarrow \mathbb {R}\).

3.7.:

Are the constructions of 3.5 and 3.6 possible if \(\{X;\Vert \cdot \Vert \}\) is an infinite dimensional normed space?

Remark 3.1c

The conclusion of 3.6 is in general false for metric spaces. For example if X is a set, the discrete metric, generates the discrete topology on X. With respect to such a topology there exists no discontinuous maps \(T:X\rightarrow \mathbb {R}\).

However there exist metric, not normed, spaces that admit discontinuous linear functionals. As an example consider \(L^p(E)\) for \(0<p<1\) endowed with the metric (3.1c) of § 3.4c of the Complements of Chap. 6. If E is a Lebesgue measurable subset of \(\mathbb {R}^N\) and \(\mu \) is the Lebesgue measure, then every nontrivial, linear functional on \(L^p(E)\) is discontinuous.Footnote 4

3.8.:

Let the unit ball in \(\mathbb {R}^N\) in some norm \(\Vert \cdot \Vert \), be \(\prod _{j=1}^N[-1,1]\). Compute the unit ball of the dual space.

3.9.:

Let \(X=\bigoplus _{j=1}^k X_j\) where \(X_j\) for \(j=1,\dots ,k\) are Banach spaces. Then \(X^*=\bigoplus _{j=1}^k X_j^*\). In particular \((L^p(E)^k)^*=L^q(E)^k\) where \(1\le p<\infty \) and p and q are Hölder conjugate.

6c Equibounded Families of Linear Maps

Let \(\{X;\Vert \cdot \Vert _X\}\) and \(\{Y;\Vert \cdot \Vert _Y\}\) be Banach spaces.

6.1.:

Let \(T(x,y):X\times Y \rightarrow \mathbb {R}\) be a functional linear and continuous in each of the two variables. Then T is linear and bounded with respect to both variables.

6.2.:

Let \(\{T_n\}\) be a sequence in \(\mathcal {B}(X;Y)\), such that the limit of \(\{T_n(x)\}\) exists for all \(x\in X\). Then \(T(x)=\lim T_n(x)\) defines a bounded, linear map from \(\{X;\Vert \cdot \Vert _X\}\) into \(\{Y;\Vert \cdot \Vert _Y\}\).

6.3.:

Let \(\{T_n\}\) be a sequence in \(\mathcal {B}(X;Y)\), such that \(\Vert T_n\Vert \le C\) for some positive constant C and all \(n\in \mathbb {N}\). Let \(X_o\) be the set of \(x\in X\) for which \(\{T_n(x)\}\) converges. Then \(X_o\) is a closed subspace of \(\{X;\Vert \cdot \Vert _X\}\).

6.4.:

Let \(T_1\) and \(T_2\) be elements of \(\mathcal {B}(X;X)\) and let \(T_1T_2(\cdot )=T_1(T_2(\cdot ))\) be the composition map. Then \(T_1T_2\in \mathcal {B}(X;X)\) and \(\Vert T_1T_2\Vert \le \Vert T_1\Vert \,\Vert T_2\Vert \).

Let \(T\in \mathcal {B}(X;X)\) satisfy \(\Vert T\Vert <1\). Then \((I+T)^{-1}\) exists as an element of \(\mathcal {B}(X;X)\) and

$$\begin{aligned} (I+T)^{-1}=I+\mathop {\textstyle {\sum }}\limits \,(-1)^n T^n. \end{aligned}$$

Hint: Since \(\Vert T\Vert <1\), the series \(\mathop {\textstyle {\sum }}\limits \Vert T^n\Vert \le \mathop {\textstyle {\sum }}\limits \Vert T\Vert ^n\) converges. Since \(\mathcal {B}(X;X)\) is a Banach space, \(\mathop {\textstyle {\sum }}\limits (-T)^n\) converges to an element \(\mathcal {B}(X;X)\).

6.5.:

Let E be a bounded open set in \(\mathbb {R}^N\). Fix \(h\in L^\infty (E)\) and find \(f\in L^\infty (E)\) such that

$$\begin{aligned} h=(I+T)f\qquad \text { i.e., formally }\qquad f=(I+T)^{-1}h \end{aligned}$$
(6.1c)

where T(f) is the Riesz potential introduced in (4.2). Verify that T is a bounded linear map from \(L^\infty (E)\) into itself. Give conditions on E so that such a formal solution formula is actually justified and exhibit explicitly the solution f.

6.6.:

Let E be a Lebesgue measurable subset of \(\mathbb {R}^N\) of finite measure and let \(1\le p,q\le \infty \) be conjugate. Then if \(q>p\) the space \(L^q(E)\) is of first category in \(L^p(E)\). Hint: \(L^q(E)\) is the union of \(\big [\Vert g\Vert _q\le n\big ]\).

8c The Open Mapping Theorem

\(\{X;\Vert \cdot \Vert _X\}\) and \(\{Y;\Vert \cdot \Vert _Y\}\) are Banach spaces:

8.1.:

\(T\in \mathcal {B}(X;Y)\) is a homeomorphism if and only if it is onto, and there exist positive constants \(c_1\le c_2\) such that

$$\begin{aligned} c_1\Vert x\Vert _X\le \Vert T(x)\Vert _Y\le c_2\Vert x\Vert _X\qquad \text { for all }\> x\in X. \end{aligned}$$
8.2.:

A bounded linear map T from a subspace \(X_o\subset X\) into Y, has closed graph if and only if its domain is closed.

8.3.:

The sup-norm on C[0, 1] generates a strictly stronger topology than the \(L^2\)-norm.

8.4.:

Let \(T\in \mathcal {B}(X;Y)\). If T(X) is of second category in Y, then T is onto.

9c The Hahn–Banach Theorem

Let X be a vector space over the complex field \(\mathbb {C}\). The norm \(\Vert \!\cdot \!\Vert \) is defined as in (i)–(iii) of § 1, except that \(\lambda \in \mathbb {C}\). In such a case \(|\lambda |\) is the modulus of \(\lambda \) as element of \(\mathbb {C}\).

Denote by \(X_{\mathbb {R}}\) the vector space X when multiplication is restricted to scalars in \(\mathbb {R}\). A linear functional \(T:X\rightarrow \mathbb {C}\) is separated into its real and imaginary part by

$$\begin{aligned} T(x)=T_{\mathbb {R}}(x)+iT_i(x) \end{aligned}$$
(9.1c)

where the maps \(T_{\mathbb {R}}\) and \(T_i\) are functionals from \(X_{\mathbb {R}}\) into \(\mathbb {R}\). Since \(T:X\rightarrow \mathbb {C}\) is linear, \(T(i x)=i T(x)\) for all \(x\in X\). From this compute

$$\begin{aligned} T(ix)=T_{\mathbb {R}}(ix)+iT_i(ix)\qquad iT(x)=iT_{\mathbb {R}}(x)-T_i(x). \end{aligned}$$

This implies

$$\begin{aligned} T_i(x)=-T_{\mathbb {R}}(ix)\qquad \text { for all }\> x\in X. \end{aligned}$$
(9.2c)

Thus \(T:X\rightarrow \mathbb {C}\) is identified by its real part \(T_{\mathbb {R}}\) regarded as a linear functional from \(X_{\mathbb {R}}\) into \(\mathbb {R}\). Vice versa any such real-valued functional \(T_{\mathbb {R}}:X_{\mathbb {R}}\rightarrow \mathbb {R}\) identifies a linear functional \(T:X\rightarrow \mathbb {C}\) by the formulae (9.1c)–(9.2c).

1.1 9.1c The Complex Hahn–Banach Theorem

Theorem 9.1c

([17, 151]) Let X be a complex vector space and let \(p:X\rightarrow \mathbb {R}\) be a semi-norm on X. Then, every linear functional \(T_o:X_o\rightarrow \mathbb {C}\) defined on a subspace \(X_o\) of X and satisfying \(|T_o(x)|\le p(x)\) for all \(x\in X_o\), admits an extension \(T:X\rightarrow \mathbf {C}\) such that

$$\begin{aligned} |T(x)|\le p(x)\quad \text { for all }\>x\in X\quad \text { and }\quad T(x)=T_o(x)\quad \text { for }\> x\in X_o. \end{aligned}$$

Proof

Denote by \(X_{o,\mathbb {R}}\) the real subspace of \(X_o\) and by \(T_{o,\mathbb {R}}:X_{o,\mathbb {R}}\rightarrow \mathbb {R}\) the real part of T. By the representation (9.1c)–(9.2c) it suffices to extend \(T_{o,\mathbb {R}}\) into a linear map \(T_{\mathbb {R}}:X_{\mathbb {R}}\rightarrow \mathbb {R}\). This follows from the Hahn–Banach theorem, since

$$\begin{aligned} T_{o,\mathbb {R}}(x)\le |T(x)|\le p(x)\qquad \text { for all } \> x\in X. \end{aligned}$$

\(\blacksquare \)

1.2 9.2c Linear Functionals in \(L^\infty (E)\)

The Riesz representation theorem for the bounded linear functionals in \(L^p(E)\), fails for \(p=\infty \). A counterexample can be constructed as follows.

On \(C[-1,1]\) let \(T_o(f)=f(0)\). This is bounded, linear functional on \(C[-1,1]\). The boundedness of \(T_o\) is meant in the sense of \(L^\infty [-1,1]\), i.e.,

$$\begin{aligned} \Vert T_o\Vert =\sup _{\genfrac{}{}{0.0pt}{}{\varphi \in C[-1,1]}{\Vert \varphi \Vert _\infty =1}}\, |T_o(\varphi )|. \end{aligned}$$

Then, by the Hahn–Banach theorem, \(T_o\) can be extended to a bounded linear functional T in \(L^\infty [-1,1]\) coinciding with \(T_o\) on \(C[-1,1]\) and such that \(\Vert T\Vert =\Vert T_o\Vert \). For such an extension there exists no function \(g\in L^1[-1,1]\) such that

$$\begin{aligned} T(f)=\int _{-1}^1 fgdx\qquad \text { for all }\quad f\in L^\infty [-1,1]. \end{aligned}$$

11c Separating Convex Subsets of X

1.1 11.1c A Counterexample of Tukey [164]

The nonempty interior assumption is essential. We produce two disjoint, closed, convex sets \(C_1\) and \(C_2\) in \(\ell _2\), such that \(C_1-C_2\) is dense in \(\ell _2\). This implies that \(C_1-C_2\) cannot be separated from \(\varTheta \) and hence \(C_1\) and \(C_2\) cannot be separated. Identify \(x\in \ell _2\) with its corresponding sequence \(\{x_n\}\) and define

$$\begin{aligned} \begin{aligned} C_1&=\Big \{x\in \ell _2\,\bigm |\, x_1>\big |n^2\big (x_n-{\textstyle \frac{1}{n}}\big )\big |, \text { for all }\> n>1\Big \}\\ C_2&=\big \{x\in \ell _2\,\bigm |\, x_n=0 \text { for all }\> n>1\Big \}. \end{aligned} \end{aligned}$$

One verifies that \(C_1\) and \(C_2\) are convex and closed. They are also disjoint. Indeed if \(y\in C_1\cap C_2\)

$$\begin{aligned} y_1\ge \big |n^2\big (0-{\textstyle \frac{1}{n}}\big )\big |=n \quad \text { for all }\> n\ge 2. \end{aligned}$$

Thus \(y_1=\infty \) and hence \(y\notin \ell _2\). To show that \(C_1-C_2\) is dense in \(\ell _2\), fix \(z\in \ell _2\) and \(\varepsilon >0\) and pick \(n_\varepsilon \) so large that

$$\begin{aligned} \mathop {\textstyle {\sum }}\limits _{n\ge n_\varepsilon } n^{-2}+ \mathop {\textstyle {\sum }}\limits _{n\ge n_\varepsilon } z_n^2<\frac{\varepsilon ^2}{4}. \end{aligned}$$

Choose a number

$$\begin{aligned} x_1>\big |n^2\big (z_n-{\textstyle \frac{1}{n}}\big )\big | \qquad \text { for }\> 2\le n_\varepsilon <n \end{aligned}$$

and let \(x=\{x_n\}\in C_1\) and \(y=\{y_n\}\in C_2\) be defined by

$$\begin{aligned} x_n=\left\{ \begin{array}{ll} x_1\quad &{}\text { for }\> n=1;\\ z_n\quad &{}\text { for }\> 2\le n <n_\varepsilon ;\\ \frac{1}{n}\quad &{}\text { for }\> n \ge n_\varepsilon ; \end{array}\right. \quad \qquad y_n=\left\{ \begin{array}{ll} z_1+x_1\quad &{}\text { for }\> n=1;\\ 0&{}\text { for }\> n>1. \end{array}\right. \end{aligned}$$

By the triangle inequality \(\Vert z-(x-y)\Vert <\varepsilon \).

1.1.1 11.1.1c A Variant of Tukey’s Counterexample

Corollary 11.1c

Every infinite dimensional Banach space X contains a 1-dimensional subspace \(C_1\) and a closed, convex set \(C_2\) that cannot be separated by a bounded linear functional.

Proof

Let \(\{\mathbf {e}_n\}\) be a sequence of linearly independent elements on the unit sphere of X and set \(C_2=\{\mathbf {e}_1\}\) and

$$\begin{aligned} C_1=\left\{ \genfrac{}{}{0.0pt}{}{x\in X\text { of the form }\> x=\mathop {\textstyle {\sum }}\limits x_n\mathbf {e}_n\> \text { with}}{x_1\ge n^3\big |x_n-n^{-2}\big |\>\text { for all }\> n\ge 2} \right\} \end{aligned}$$

\(\blacksquare \)

1.2 11.2c A Counterexample of Goffman and Pedrick [56]

Let X be a linear space with a countably infinite Hamel basis \(\{x_n\}\). Let \(C_1\) be the collections of all elements \(x\in X\) whose last nonzero coefficient of its representations in terms of \(\{x_n\}\) is positive. The set \(C_1\) is convex and \(\varTheta \notin C_1\). Any functional T that separates \(\{\varTheta \}\) from \(C_1\) must be either non-positive or nonnegative on \(C_1\). Let then T be a linear functional on X, which is nonnegative on \(C_1\). For every \(x_j\in \{x_n\}\) and all \(\alpha \in \mathbb {R}\) the element \(\alpha x_j+x_{j+1}\in C_1\). Therefore

$$\begin{aligned} T(\alpha x_j+x_{j+1})=\alpha T(x_j)+T(x_{j+1}) \ge 0 \quad \text { for all }\>\alpha \in \mathbb {R}. \end{aligned}$$

Thus \(T(x_j)=0\). Since \(x_j\in \{x_n\}\) is arbitrary, T vanishes on all basis elements of X and therefore is the zero functional.

1.3 11.3c Extreme Points of a Convex Set

Let C be a convex set in a linear, normed space. A point \(x\in C\) is an extreme point of C if there do not exist distinct points \(u,v\in C\) and \(t\in (0,1)\), such that \(x=t u+(1-t)v\), that is, if no line segment in C has x as an interior point.

11.1.:

The extreme points of a convex, closed polyhedron in \(\mathbb {R}^N\) are its vertices.

11.2.:

An open convex polyhedron in \(\mathbb {R}^N\) has no extreme points.

11.3.:

A closed \(\frac{1}{2}\)-space in \(\mathbb {R}^N\) has no extreme points.

11.4.:

The set of extreme points of the closed unit ball of a uniformly convex linear normed space \(\{X;\Vert \cdot \Vert \}\) is the unit sphere. In particular the extreme points of \(L^p(E)\) for \(1<p<\infty \) are all those functions in \(L^p(E)\) such that \(\Vert f\Vert _p=1\).

11.5.:

The set of the extreme points of the closed unit ball in \(L^\infty (E)\) are those \(f\in L^\infty (E)\) such that \(|f|=1\) a.e. in E.

11.6.:

The set of the extreme points of the closed unit ball in \(L^1(E)\) is empty.

11.7.:

Let E be a bounded, connected, open set in \(\mathbb {R}^N\). The closed unit ball of \(C({\bar{E}})\) has no, non constant, extreme points. Examine the case when E has several connected components.

A set \(E\subset C\) is an extremal subset of C if it satisfies the property:

$$\begin{aligned} \left\{ \genfrac{}{}{0.0pt}{}{\text {if there exist}\ u,v\in C\ \text {and}\ t\in (0,1)\ \text {such that}}{tu+(1-t)v\in E,\ \text {then}\ u,v\in E.} \right\} \end{aligned}$$

Extreme points are extremal sets. Convex portions of a face of a closed, convex polyhedron are extremal sets of that polyhedron. Prove the following:

Lemma 11.1c

If \(E_1\) is an extremal subset of C and \(E_2\) is an extremal subset of \(E_1\), then \(E_2\) is an extremal subset of C.

Lemma 11.2c

A nonempty, convex, compact subset C of a linear, normed space X, has extreme points.

Proof

Consider the collection \(\mathcal {E}\) of all closed, extremal subsets of C, partially ordered by inclusion. Such a collection is not empty since \(C\in \mathcal {E}\). Any linearly ordered subcollection \(\mathcal {E}_1\subset \mathcal {E}\) has the finite intersection property (§ 5 of Chap. 2) and therefore

$$\begin{aligned} E_o=\mathop {\textstyle {\bigcap }}\limits \{E\,|\,E\in \mathcal {E}_1\}\quad \text { is not empty, closed and compact}. \end{aligned}$$

We claim that \(E_o\) is a singleton which is extreme of C. If \(E_o\) contains more than one point, pick \(p,q\in E_o\) with \(p\ne q\), let \(T\in X^*\) be such that \(T(p)>T(q)\) (Corollary 10.1), and set

$$\begin{aligned} E_o^\prime = \Big \{x\in E_o\,\big |\, T(x)=\inf _{y\in E_o}T(y)\Big \}. \end{aligned}$$

The set \(E_o^\prime \) is well defined, since T is continuous and \(E_o\) is compact, and is a proper subset of \(E_o\), since \(T(p)>T(q)\). The proof consists in verifying that \(E_o^\prime \) is a convex, closed, extremal subset of \(E_o\). Thus by Lemma 11.1c it would be a closed, extremal subset of C, properly contained in \(E_o\), thereby contradicting the definition of \(E_o\).

If there exists \(u,v\in C\) and \(t\in (0,1)\) such that \(tu+(1-t)v\in E_o^\prime \subset E_o\), then \(u,v\in E_o\) and

$$\begin{aligned} tT(u)+(1-t)T(v)=T(tu+(1-t)v)=\inf _{y\in E_o}T(y). \end{aligned}$$

Now

$$\begin{aligned} T(u)>\inf _{y\in E_o}T(y)\quad \text { implies }\quad T(v)<\inf _{y\in E_o}T(y) \end{aligned}$$

contradicting that \(v\in E_o\). Thus

$$\begin{aligned} T(u)=T(v)=\inf _{y\in E_o}T(y) \quad \text { which implies }\> u,v\in E_o^\prime . \end{aligned}$$

\(\blacksquare \)

Theorem 11.1c

(Krein–Milman [86]) A compact, convex set C in a normed linear space is the closed convex hull of its extreme points.

Proof

Let E be the set of the extreme points of C. By the definition of convex hull, \(\overline{c(E)}\subset C\). To prove the converse inclusion, proceed by contradiction, assuming that there exists \(x_o\in C\) such that \(x_o\notin \overline{c(E)}\). By Proposition 11.3 there exists \(T\in X^*\) such that \(T(x_o)<T(\overline{c(E)})\). Set

$$\begin{aligned} C_1=\Big \{x\in C\,\bigm |\, T(x)=\inf _{y\in C} T(y)\Big \}. \end{aligned}$$

Proceeding as in the proof of Lemma 11.2c one verifies that \(C_1\) is a nonempty, convex, closed, extremal subset of C. Moreover \(C_1\cap E=\emptyset \). Since \(C_1\) is convex and compact it has at least one extreme point \(y_o\), which by Lemma 11.1c is also an extreme point of C. \(\blacksquare \)

11.8.:

Give an example to show that the compactness assumption on C is essential.

1.4 11.4c A General Version of the Krein–Milman Theorem

Prove that Theorem 11.1c continues to hold, with essentially the same proof if X is a Hausdorff topological vector space on which \(X^*\) separates points. Compactness of C is referred to the topology of X.

12c Weak Topologies

Proposition 12.1c

Let \(\{X;\Vert \cdot \Vert \}\) be a normed space and let \(\mathcal {W}\) denote its weak topology. Denote also by \(\mathcal {O}\) a weak neighborhood of the origin of X. Then:

(i):

For every \(V\in \mathcal {W}\) and \(x_o\in V\), there exists \(\mathcal {O}\) such that \(x_o+\mathcal {O}\subset V\).

(ii):

For every \(V\in \mathcal {W}\) and \(x_o\in V\), there exists \(\mathcal {O}\) such that \(\mathcal {O}+\mathcal {O}+x_o\subset V\).

(iii) :

For every \(\mathcal {O}\) there exists \(\mathcal {O}^\prime \) such that \(\lambda \mathcal {O}^\prime \subset \mathcal {O}\), for all \(|\lambda |\le 1\).

Proof

(of (i)) An open weak neighborhood V of \(x_o\in X\) contains an open set of the form

$$\begin{aligned} V_o=\mathop {\textstyle {\bigcap }}\limits _{j=1}^m T^{-1}_j (T_j(x_o)-\alpha _j,T_j(x_o)+\alpha _j) \end{aligned}$$

for some finite m and \(\alpha _j>0\). Equivalently

$$\begin{aligned} V_o=\{x\in X\bigm | |T_j(x-x_o)|<\alpha _j \quad \text { for }\> j=1,\dots ,n\}. \end{aligned}$$

The neighborhood of the origin

$$\begin{aligned} \mathcal {O}=\{x\in X\bigm | |T_j(x)|<\alpha _j \quad \text { for }\> j=1,\dots ,n\big \} \end{aligned}$$

is such that \(x_o+\mathcal {O}\subset V_o\). \(\blacksquare \)

The remaining statements are proved similarly.

12.1.:

In a finite dimensional normed linear space, the notions of weak and strong convergence coincide.

12.2.:

Construct examples and counterexamples for the following statements:

i.:

A weakly closed subset of a normed linear space is also strongly closed. The converse is false.

ii.:

A strongly sequentially compact subset of a normed linear space is also weakly sequentially compact. The converse is false.

12.3.:

A normed linear space is weakly complete if and only if it is complete in its strong topology. A weakly dense set in \(\{X;\Vert \cdot \Vert \}\) is also strongly dense.

12.4.:

The weak topology on a normed linear space is Hausdorff (Proposition 11.3).

1.1 12.1c Infinite Dimensional Normed Spaces

Let \(\{X;\Vert \cdot \Vert \}\) be an infinite dimensional Banach space. There exists a countable collection \(\{X_n\}\) of infinite dimensional, closed subspaces of X such that \(X_{n+1}\subset X_n\) with strict inclusion. For example one might take a nonzero functional \(T_1\in X^*\) and set \(X_1=\ker \{T_1\}\). Such a subspace is infinite dimensional (Hint: Proposition 5.1). Then select a nonzero functional \(T_2\in X_1^*\), set \(X_2=\ker \{T_2\}\), and proceed by induction.

For each n, select an element \(x_n\in X_{n+1}-X_n\) so that \(\Vert x_n\Vert =2^{-n}\). This generates a sequence \(\{x_n\}\) of linearly independent elements of X whose span \(X_o\) is isomorphic to \(\ell _\infty \) by the representation

$$\begin{aligned} \ell _\infty \ni \{c_n\}\>\longrightarrow \>\mathop {\textstyle {\sum }}\limits \, c_n x_n\in X \end{aligned}$$

Since the dimension of \(\ell _\infty \) is not less than the cardinality of \(\mathbb {R}\) the dimension of an infinite dimensional Banach space is at least the cardinality of \(\mathbb {R}\). Compare with 2.1.

12.5.:

A Banach space \(\{X;\Vert \cdot \Vert \}\) is finite dimensional if and only if every linear subspace is closed (2.6). However an infinite dimensional Banach space X contains infinitely many, infinite dimensional, closed subspaces.

12.6.:

The weak topology of an infinite dimensional normed space X is not normable, i.e., there exists no norm \(\Vert \cdot \Vert _w\) on X that generates the weak topology. This follows from Corollary 12.3. Give an alternate proof. Hint: If such \(\Vert \cdot \Vert _w\) exists, the open unit ball, with respect to such a norm, would be an open neighborhood of the origin.

1.2 12.2c About Corollary 12.5

In the context of \(L^p(E)\) spaces the corollary had been established in [14]. When \(p=2\) the coefficients \(\alpha _j\) can be given an elegant form as established by the following proposition.

Proposition 12.2c

Let \(\{f_n\}\) be a sequence of functions in \(L^2(E)\) weakly convergent to some \(f\in L^2(E)\). There exists a subsequence \(\{f_{n_j}\}\) such that setting

$$\begin{aligned} \varphi _m=\frac{f_{n_1}+f_{n_2}+\cdots +f_{n_m}}{m} \end{aligned}$$

the sequence \(\{\varphi _m\}\) converges to f strongly in \(L^2(E)\).

Proof

By possibly replacing \(f_n\) with \(f_n-f\), we may assume that \(f=0\). Fix \(n_1=1\). Since \(\{f_n\}\rightarrow 0\) weakly in \(L^2(E)\), there exists an index \(n_2\) such that

$$\begin{aligned} \Big |\int _Ef_{n_1}f_{n_2}d\mu \Big |\le \frac{1}{2}. \end{aligned}$$

Then there is an index \(n_3\) such that

$$\begin{aligned} \Big |\int _Ef_{n_1}f_{n_3}d\mu \Big | \le \frac{1}{3}\qquad \text { and }\qquad \Big |\int _Ef_{n_2}f_{n_3}d\mu \Big |\le \frac{1}{3}. \end{aligned}$$

Proceeding in this fashion we extract, out of \(\{f_n\}\), a subsequence \(\{f_{n_j}\}\), such that, for all \(k\ge 2\)

$$\begin{aligned} \Big |\int _Ef_{n_\ell }f_{n_k}d\mu \Big | \le \frac{1}{k}\quad \text { for all }\>\ell =1,\dots ,k-1. \end{aligned}$$

Denoting by M the upper bound of \(\Vert f_n\Vert _2\), compute

$$\begin{aligned} \int _E\varphi _{m}^2d\mu&=\frac{1}{m^2}\int _E(f_{n_1}+f_{n_2}+\cdots f_{n_m})^2d\mu \\&\le \frac{1}{m^2}\Big (m M^2+2+4\frac{1}{2}+\cdots +2m\frac{1}{m}\Big )\\&\le \frac{M^2+2}{m}\>\longrightarrow \>0\quad \text { as }\> m\rightarrow \infty . \end{aligned}$$

\(\blacksquare \)

1.3 12.3c Weak Closure and Weak Sequential Closure

For a subset E of a normed space \(\{X;\Vert \cdot \Vert \}\) denote by \(\bar{E}_{w-\sigma }\) its is weak sequential closure, that is the set of all limit points of weakly convergent sequences in E. Equivalently, by \(\bar{E}_{w-\sigma }\) is the set of all points \(x\in X\), for which there exists a sequence \(\{x_n\}\subset E\) weakly convergent to x. By the definition \(\bar{E}_{w-\sigma }\subset \bar{E}_w\). The inclusion is in general strict, as shown by the following examples.

12.7.:

Let \(E=\{\sqrt{n}\mathbf {e}_n\}\subset \ell _2\), where \(\mathbf {e}_n\) is the infinite sequence consisting of zeroes except the nth entry which is one. Denoting by \(\mathbf {0}\) the zero element of \(\ell _2\)

$$\begin{aligned} \mathbf {0}\in \overline{\{\sqrt{n}\mathbf {e}_n\}}_{w} \qquad \text { but }\qquad \mathbf {0}\notin \overline{\{\sqrt{n}\mathbf {e}_n\}}_{w-\sigma }. \end{aligned}$$
(12.1c)

To prove the first of these, pick \(T\in \ell _2^*\) and for a fixed \(\alpha >0\), consider the weak, open neighborhood of the origin

$$\begin{aligned} \mathcal {O}_{T,\alpha }=\{x\in \ell _2\,\bigm |\, |T(x)|<\alpha \}. \end{aligned}$$

By the Riesz representation theorem any such T is identified by some \(\mathbf {t}\in \ell _2\) acting on elements \(\mathbf {a}\in \ell _2\) by the formula

$$\begin{aligned} T(\mathbf {a})=\mathbf {t}\cdot \mathbf {a}=\mathop {\textstyle {\sum }}\limits t_na_n. \end{aligned}$$

Since \(\mathbf {t}\in \ell _2\), for all \(M>0\), there exists \(n>M\) such that \(|t_n|\le \alpha /\sqrt{n}\). Indeed otherwise

$$\begin{aligned} \Vert \mathbf {t}\Vert ^2\ge \mathop {\textstyle {\sum }}\limits _{n>M} t_n^2>\alpha \mathop {\textstyle {\sum }}\limits \frac{1}{n}. \end{aligned}$$

For such an index, \(|T(\sqrt{n}\mathbf {e}_n)|<\alpha \), and hence \(\sqrt{n}\mathbf {e}_n\in \mathcal {O}_{T,\alpha }\). Let now

$$\begin{aligned} \mathcal {O}=\mathop {\textstyle {\bigcap }}\limits _{j=1}^kT_j^{-1}(-\alpha _j,\alpha _j), \quad \text { for }\>\alpha _j>0\quad \text { for some }\>k\in \mathbb {N}. \end{aligned}$$
(12.2c)

Prove that n can be chosen so large that \(\sqrt{n}\mathbf {e}_n\in \mathcal {O}\). Therefore any weak, open neighborhood of the origin intersects \(\{\sqrt{n}\mathbf {e}_n\}\).

To prove the second of (12.1c) we establish that there exists no subsequence \(\{\sqrt{m}\mathbf {e}_m\}\subset \{\sqrt{n}\mathbf {e}_n\}\), weakly convergent to \(\mathbf {0}\). Having picked such a subsequence assume first that some fixed index \(j\in \mathbb {N}\) occurs infinitely many times in \(\{\sqrt{m}\mathbf {e}_m\}\). Pick \(T_j\in \ell _2^*\) corresponding to \(\mathbf {e}_j\in \ell _2\). For such a choice the sequence \(\{T_j(\sqrt{m}\mathbf {e}_m)\}\) contains the value \(\sqrt{j}\) infinitely many times, and thus it cannot converge to zero. If no index \(j\in \mathbb {N}\) occurs infinitely many times in \(\{\sqrt{m}\mathbf {e}_m\}\) pick a subsequence

$$\begin{aligned} \{\sqrt{m_j}\mathbf {e}_{m_j}\}\subset \{\sqrt{m}\mathbf {e}_m\}\quad \text { so that }\quad m_j\ge 2^j \end{aligned}$$

and set

$$\begin{aligned} \mathbf {t}=\mathop {\textstyle {\sum }}\limits _{j\in \mathbb {N}}\frac{1}{\sqrt{m_j}}\mathbf {e}_{m_j}\quad \text { satisfying }\quad \Vert \mathbf {t}\Vert ^2=\mathop {\textstyle {\sum }}\limits _{j\in \mathbb {N}}\frac{1}{m_j}<\infty . \end{aligned}$$

Therefore \(\mathbf {t}\in \ell _2\) and we let \(T\in \ell _2^*\) be its corresponding functional. For such a functional \(T(\sqrt{m_j}\mathbf {e}_{m_j})=1\). Therefore the sequence \(\{T(\sqrt{m}\mathbf {e}_m)\}\) contains 1 infinitely many times and thus it cannot converge to zero.

Remark 12.1c

The set E is countable, so that countability alone is not sufficient to identifty weak closure with weak sequential closure.

Remark 12.2c

The set \(\{\sqrt{n}\mathbf {e}_n\}\) is unbounded in \(\ell _2\). However the strict inclusion \(\bar{E}_{w-\sigma }\subset \bar{E}_w\) continues to hold even for bounded sets, as shown by the next example.

12.8.:

Consider the set \(E\subset \ell _1\)

$$\begin{aligned} E=\mathop {\textstyle {\bigcup }}\limits _{n,m\in \mathbb {N}; m\ne n} \{\mathbf {e}_m-\mathbf {e}_n\}. \end{aligned}$$

We claim that E is bounded in \(\ell _1\), its weak closure contains \(\mathbf {0}\) but its weak sequential closure does not contain \(\mathbf {0}\).

To establish the first claim let \(\mathcal {O}\) be a weakly open set of the form (12.2c). Every \(T_j\in \ell _1^*\) is identified with an element \(\mathbf {t}_j\in \ell _\infty \) acting on elements \(\mathbf {a}\in \ell _1\) by the formula

$$\begin{aligned} T_j(\mathbf {a})=\mathbf {t}_j\cdot \mathbf {a}=\mathop {\textstyle {\sum }}\limits t_{j,n}a_n. \end{aligned}$$

Since \(\mathbf {t}_j\in \ell _\infty \) for \(j=1,\dots ,k\), as n ranges over \(\mathbb {N}\), the k-tuple \((t_{1,n},\dots , t_{k,n})\) ranges over a bounded set \(K\subset \mathbb {R}^k\), which, without loss of generality we may assume to be a closed cube with faces parallel to the coordinate planes. Pick \(\varepsilon =\min \{\alpha _1,\dots ,\alpha _k\}\) and subdivide K into no less than \(\varepsilon ^{-k}\) homotetic closed sub-cubes \(K_\varepsilon \) of edge not exceeding \(\varepsilon \). At least one of these sub-cubes must contain infinitely many k-tuples \((t_{1,n},\dots , t_{k,n})\). Select one such cube and relabel the corresponding k-tuples as \(\{(t_{1,n_i},\dots , t_{k,n_i})\}_{i\in \mathbb {N}}\). For the element \(\mathbf {e}_{m_i}-\mathbf {e}_{n_i}\in \{\mathbf {e}_m-\mathbf {e}_n\}\) compute

$$\begin{aligned} |T_j(\mathbf {e}_{m_i}-\mathbf {e}_{n_i})|=|t_{j,m_i}-t_{j,n_i}|\le \varepsilon \le \alpha _j \quad \text { for }\> j=1,\dots k. \end{aligned}$$

Therefore \(\mathbf {e}_{m_i}-\mathbf {e}_{n_i}\in \mathcal {O}\). Thus every weak open neighborhood of \(\mathbf {0}\) intersects \(\{\mathbf {e}_m-\mathbf {e}_n\}\) and hence \(\mathbf {0}\in \overline{\{\mathbf {e}_m-\mathbf {e}_n\}}_w\).

To show that no subsequence \(\{\mathbf {e}_{m_j}-\mathbf {e}_{n_j}\}\subset \{\mathbf {e}_m-\mathbf {e}_n\}\) converges weakly to \(\mathbf {0}\) it suffices for any such subsequence, to construct a functional \(T\in \ell _1^*\) such that \(\{T(\mathbf {e}_{m_j}-\mathbf {e}_{n_j})\}\) does not converge to zero. Construct one such T by examining separately the following cases

i.:

Some pair \((m_j,n_j)\) occurs infinitely many times;

ii.:

Some index \(m_j\) occurs infinitely many times and \(n_j\rightarrow \infty \), or vice versa.

iii.:

Both \(m_j, n_j\rightarrow \infty \).

12.9.:

Let \(E=\ell _2\) be defined by (von Neuman [113])

$$\begin{aligned} E=\{\mathbf {e}_n+n\mathbf {e}_m\}\quad \text { for }\>0\le n<m. \end{aligned}$$

Prove that E is strongly closed (Hint: all the sequences in E, Cauchy in \(\ell _2\) are constant).

Prove that the origin of \(\ell _2\) is in the weak closure but not in the weak sequential closure of E.

12.10.:

Prove that \(\ell _2\) and \(\ell _1\) equipped with their weak topology do not satisfy the first axiom of countability.

14c Weak Compactness

14.1.:

A weakly compact subset of a normed linear space is weakly closed. See also Proposition 5.1 of Chap. 2.

14.2. :

A weakly compact, convex subset of a normed linear space is strongly closed.

1.1 14.1c Linear Functionals on Subspaces of \(C({\bar{E}})\)

Let E be a bounded, open set in \(\mathbb {R}^N\) and let \(C({\bar{E}})\) denote the space of the continuous functions in \({\bar{E}}\) equipped with the sup-norm.

Let \(X_o\) be a subspace of \(C({\bar{E}})\) closed in the topology of \(L^2(E)\). Prove that there exist positive constants \(C_o\le C_1\), such that

$$\begin{aligned} C_o\Vert f\Vert _\infty \le \Vert f\Vert _2\le C_1\Vert f\Vert _\infty . \end{aligned}$$

Let \(T_{o,y}\in X_o^*\) be the evaluation map at y

$$\begin{aligned} T_{o,y}(f)=f(y)\qquad \text { for all }\> f\in X_o. \end{aligned}$$

By the Hahn–Banach theorem there exists a functional \(T_y\in L^2(E)^*\) such that \(T_y=T_{o,y}\) on \(X_o\). By the Riesz representation of the bounded linear functionals in \(L^2(E)\), there exists a function \(K(\cdot ,y)\in L^2(E)\), such that

$$\begin{aligned} f(y)=\int _EK(x,y)f(x)d\mu \qquad \text { for all }\> f\in X_o. \end{aligned}$$
(14.1c)

Proposition 14.1c

The unit ball of \(X_o\) is compact.

Proof

The closed unit ball \(\bar{B}_{o,1}\) of \(X_o\) is also weakly closed. Since \(L^2(E)\) is reflexive, also \(X_o\) is reflexive. Therefore \(\bar{B}_{o,1}\) is bounded and weakly closed and hence sequentially compact. In particular every sequence \(\{f_n\}\subset \bar{B}_{o,1}\) contains in turn a subsequence \(\{f_{n^\prime }\}\) weakly convergent to some \(f\in \bar{B}_{o,1}\). By (14.1c) such a sequence converges pointwise to f and

$$\begin{aligned} \Vert f_n\Vert _\infty \le (\text {const})\Vert f_n\Vert _2\le (\text {const})^\prime \end{aligned}$$

for a constant independent of n. Therefore, by the Lebesgue dominated convergence theorem, \(\{f_{n^\prime }\}\rightarrow f\) strongly in \(L^2(E)\).

Thus every sequence \(\{f_n\}\subset \bar{B}_{o,1}\) contains in turn a strongly convergent subsequence. This implies that \(\bar{B}_{o,1}\) is compact in the strong topology inherited form \(L^2(E)\). \(\blacksquare \)

Corollary 14.1c

Every subspace \(X_o\subset C({\bar{E}})\), closed in \(L^2(E)\) is finite dimensional.

1.2 14.2c Weak Compactness and Boundedness

14.3.:

Give a different proof of Corollary 14.1 by means of the uniform boundedness principle. Hint: For all \(T\in X^*\), the image T(E) is a compact and hence bounded subset of \(\mathbb {R}\).

15c The Weak\({}^*\) Topology of \(X^*\)

1.1 15.1c Total Sets of X

The linear span \(X_*\) of a collection of linear functionals on a linear vector space X is a total set of X, if it separates its points, that is, if for any two distinct elements \(x,y\in X\), there exists \(T\in X_*\) such that \(T(x)\ne T(y)\). The dual \(X^*\) of a normed space \(\{X;\Vert \cdot \Vert \}\) separates the points of X and therefore is a total set of X. However the notion of total set does not require a topology being placed on X nor the continuity of the linear maps in \(X_*\).

If \(X_*\) is total for X, one might endow X with the weakest topology \(\mathcal {W}_*\), for which all the elements of \(X_*\) are continuous. A \(\mathcal {W}_*\)-open base neighborhood of the origin is of the form

$$\begin{aligned} \mathcal {O}_* =\left\{ \genfrac{}{}{0.0pt}{}{x\in X\,\bigm |\,|T_j(x)|<\alpha _j\> \text { for }\> T_j\in X_*\>\text { and }\>\alpha _j\in \mathbb {R}^+}{\text {for }\> j=1,\dots ,n\> \text { for some finite }\> n.}\right\} \end{aligned}$$
(15.1c)

The construction is in all similar to the construction of the weak topology of a normed space \(\{X;\Vert \cdot \Vert \}\). Since \(X_*\) is total, the \(\mathcal {W}_*\) topology on X is Hausdorff and one verifies that the operations of sum and product by scalars are continuous in the indicated topology. Thus \(\{X;\mathcal {W}_*\}\) is a Hausdorff, topological vector space. By construction the elements of \(X_*\) are continuous from \(\{X;\mathcal {W}_*\}\) into \(\mathbb {R}\). The next proposition asserts that these are the only bounded linear functional on \(\{X;\mathcal {W}_*\}\).

Proposition 15.1c

Let \(T:\{X;\mathcal {W}_*\}\rightarrow \mathbb {R}\) be nonidentically zero, linear and continuous. Then \(T\in X_*\).

Proof

By the continuity of T, there exists a \(\mathcal {W}_*\)-open set of the form (15.1c) such that \(\mathcal {O}_*\subset T^{-1}(-1,1)\). Let \(T_j\in X_*\) for \(j=1,\dots ,n\) be the functionals in \(X_*\) that identify \(\mathcal {O}_*\). If \(x\in {\mathop {\textstyle {\bigcap }}\limits }_{j=1}^n\ker \{T_j\}\) then \(x\in T^{-1}(-\varepsilon ,\varepsilon )\) for all \(\varepsilon >0\). Therefore T vanishes on \({\mathop {\textstyle {\bigcap }}\limits }_{j=1}^n\ker \{T_j\}\), and by Proposition 11.2 of Chap. 2, T is a linear combination of the \(T_j\). \(\blacksquare \)

15.1.:

Let \(\{X;\Vert \cdot \Vert \}\) be a normed space. Then \(X_{**}\) as defined in (13.3) separates the points in \(X^*\) and is total for \(X^*\). The elements of \(X_{**}\) are the only bounded linear functionals on \(X^*\) equipped with its \(\text {weak}^*\) topology. In particular every \(T_*\in X^{**}-X_{**}\) is discontinuous with respect to the \(\text {weak}^*\) topology of \(X^*\),

15.2.:

Let \(\{X;\Vert \cdot \Vert \}\) be a normed space. Then \(X^*\) equipped with its \(\text {weak}^*\) topology is a topological vector space whose dual is isometrically isomorphic to X.

15.3.:

Let \(\{X;\Vert \cdot \Vert \}\) be a normed space. Then every convex, \(\text {weak}^*\) compact subset of \(X^*\) is the \(\text {weak}^*\) closed convex hull of its extreme points (§ 11.4c).

1.2 15.2c Metrization Properties of Weak\({}^*\) Compact Subsets of \(X^*\)

Proposition 15.2c

Let \(\{X;\Vert \cdot \Vert \}\) be a Banach space. Then the weak\(^{*}\) topology of a \(\text {weak}^*\) compact set \(K\subset X^*\) is metric if and only if \(\{X;\Vert \cdot \Vert \}\) is separable.

Proof

(sufficient condition) Assume first that X is separable and let \(\{x_n\}\subset X\) be a sequence dense in X. For \(T_1,T_2\in X^*\) set

$$\begin{aligned} d(T_1,T_2)=\mathop {\textstyle {\sum }}\limits \frac{1}{2^n} \frac{|(T_1-T_2)(x_n)|}{1+|(T_1-T_2)(x_n)|}. \end{aligned}$$

Keeping in mind that X as a subset of \(X^{**}\) separates the points of \(X^*\), one verifies that

$$\begin{aligned} d:X^*\times X^*\rightarrow \mathbb {R}^+ \end{aligned}$$

is a metric which generates a metric topology on \(X^*\) (§ 13.11c of the Complements of Chap. 2). One also verifies that every ball \(B_\varepsilon (T)\subset X^*\) of center \(T\in X^*\) and radius \(\varepsilon >0\) in the metric \(d(\cdot ,\cdot )\) contains a weak\({}^*\) open neighborhood of T. Therefore the identity map from \(X^*\) equipped with its weak\({}^*\) topology onto \(X^*\) equipped with the metric topology of \(d(\cdot ,\cdot )\), is continuous. Let \(\{K;\text {weak}^*\}\) and \(\{K;d\}\) be the topological spaces formed by K equipped with the topologies inherited respectively from the weak\({}^*\) and metric topologies of \(X^*\). The identity map from the compact space \(\{K;\text {weak}^*\}\) onto \(\{K;d\}\) is continuous and one-to-one. To show that it is a homeomorphism we appeal to Proposition 5.1 of Chap. 2. A \(\text {weak}^*\)-closed set \(E\subset \{K;\text {weak}^*\}\) is \(\text {weak}^*\) compact. Since the identity map from \(\{K;\text {weak}^*\}\) onto \(\{K;d\}\) is continuous, the image of E in \(\{K;d\}\) is compact and hence closed in the metric topology of \(\{K;d\}\), since the latter is Hausdorff. \(\blacksquare \)

Remark 15.1c

The identity map from \(\{K;\text {weak}^*\}\) onto \(\{K;d\}\), being a homeomorphism, identifies the two topologies. For this is essential that \(\{K;\text {weak}^*\}\) be \(\text {weak}^*\) compact. In particular, the proposition does not imply that the \(\text {weak}^*\) topology of the whole \(X^*\) is metrizable.

Proof

(necessary condition) Assume conversely that the weak\(^{*}\) topology of \(K\subset X^*\) is metric. Up to a translation, may assume that \(\varTheta \in K\). There exists a countable collection \(\{\mathcal {O}^*_n\}\) of weak\({}^*\) open neighborhoods of \(\varTheta ^*\) such that \(\mathop {\textstyle {\bigcap }}\limits \mathcal {O}^*_n=\varTheta ^*\). Without loss of generality the \(\mathcal {O}^*_n\) are of the form

$$\begin{aligned} \mathcal {O}^*_n=\left\{ \genfrac{}{}{0.0pt}{}{T\in X^*\bigm | |x_{j_n}(T)|<\alpha _{j_n}\>\text { for }\> \alpha _{j_n}>0\> \text { and }\> x_{j_n}\in X}{\text {for }\> j_n\in \{j_{n,1},\dots ,j_{n,k}\}\subset \mathbb {N}}\right\} . \end{aligned}$$

The collection of \(\{x_{j_n}\}\) is countable and its closed linear span is separable. We claim that \(\overline{\{x_{j_n}\}}=X\). If not, there exists a nontrivial \(T\in X^*\) vanishing on \(\overline{\{x_{j_n}\}}\) (Corollary 10.3). However if \(T(x_{j_n})=0\) for all \(x_{j_n}\) then \(T\in \mathcal {O}_n^*\) for all n and thus \(T=\varTheta ^*\).\(\blacksquare \)

16c The Alaoglu Theorem

16.1.:

Let E be a bounded open set in \(\mathbb {R}^N\). Prove that there is no linear normed space \(\{X;\Vert \cdot \Vert \}\) whose dual is \(L^1(E)\) with respect to the Lebesgue measure. Hint: Problems 11.6, and 15.3.

16.2.:

Let E be a bounded open set in \(\mathbb {R}^N\). Prove that \(C({\bar{E}})\) is not the dual of any linear normed space. Hint: Problems 11.7, and 15.3.

16.3.:

The weak\({}^*\) topology of the closed unit ball of \(L^\infty (E)\) and \(\ell _\infty \) are metrizable.

16.4.:

A bounded and weakly closed subset E of a Banach space X, need not be weakly compact. In particular the closed unit balls of \(L^1(E)\), \(L^\infty (E)\), \(\ell _1\) and \(\ell _\infty \) are weakly closed but not weakly compact.

16.5.:

Let \(\{X;\Vert \cdot \Vert \}\) be a normed space and let

$$\begin{aligned} B_1=\{x\in X\bigm | \Vert x\Vert \le 1\},\qquad S_1=\{x\in X\bigm | \Vert x\Vert = 1\} \end{aligned}$$

be respectively the unit ball and the unit sphere in X. Prove or disprove by a counterexample the following statements:

i.:

\(B_1\) is strongly and weakly bounded.

ii.:

\(B_1\) is weakly sequentially closed.

iii.:

\(S_1\) is weakly sequentially closed.

iv.:

If \(\{X;\Vert \cdot \Vert \}\) is reflexive then \(S_1\) is weakly sequentially closed.

v.:

If \(\{X;\Vert \cdot \Vert \}\) is reflexive then \(B_1\) is weakly sequentially compact.

vi.:

If \(\{X;\Vert \cdot \Vert \}\) is reflexive then it is Banach.

vii.:

If \(\{X;\Vert \cdot \Vert \}\) is Banach then it is reflexive.

viii.:

The weak topology of \(\{X;\Vert \cdot \Vert \}\) is Hausdorff.

1.1 16.1c The Weak\({}^*\) Topology of \(X^{**}\)

Denote by \(X^{***}\) the dual of \(X^{**}\), that is the collection of all bounded linear functionals \(f:X^{**}\rightarrow \mathbb {R}\). It is a Banach space by the norm

$$\begin{aligned} \Vert f\Vert =\sup _{x^{**}\in X^{**}; x^{**}\ne 0} \frac{f(x^{**})}{\Vert x^{**}\Vert }= \sup _{x^{**}\in X^{**}; \Vert x^{**}\Vert =1}f(x^{**}). \end{aligned}$$
(16.1c)

Every element \(x^*\in X^*\) identifies an element \(f_{x^*}\in X^{***}\) by the formula

$$\begin{aligned} X^{**}\ni x^{**}\longrightarrow f_{x^*}(x^{**})= x^{**}(x^*). \end{aligned}$$
(16.2c)

Let \(X_{***}\) denote the collection of all such functionals, that is

$$\begin{aligned} X_{***}=\left\{ \genfrac{}{}{0.0pt}{}{\text {the collection of all functionals}\ f_{x^*}\in X^{***}}{\text {of the form (13.2) as}\ x^*\ \text {ranges over}\ X^*}\right\} . \end{aligned}$$

From Corollary 10.2 and (16.1c) it follows that \(\Vert f_{x^*}\Vert =\Vert x^*\Vert \). Therefore the injection map

$$\begin{aligned} X^*\ni x^*\longrightarrow f_{x^*}\in X_{***}\subset X^{***} \end{aligned}$$
(16.3c)

is an isometric isomorphism between \(X^*\) and \(X_{***}\). In general, not all the bounded linear functionals \(f:X^{**}\rightarrow \mathbb {R}\) are derived from the injection map (16.3c); otherwise said, the inclusion \(X_{***}\subset X^{***}\) is in general strict. The set \(X_{***}\) is total for \(X^{**}\) and it generates a topology \(\mathcal {W}_{**}\) on \(X^{**}\) which is the \(\text {weak}^*\) topology generated by \(X^*\) on \(X^{**}\) (§ 15.1c). By this topology, \(\{X^{**};\mathcal {W}_{**}\}\) turns into a Hausdorff, linear, topological vector space (§ 15.1c), and for such a space the separation Proposition 11.3 holds. In particular \(\mathcal {W}_{**}\)-closed sets are separated from points by a bounded linear functional on \(\{X^{**};\mathcal {W}_{**}\}\). By Proposition 15.1c, any such a functional is of the form (16.2c). Let

$$\begin{aligned} \begin{array}{ll} B\>&{}\text {the unit ball of}\ X\ \text {closed in the norm of}\ \{X;\Vert \cdot \Vert _X\}\\ B^{**}\>&{}\text {the unit ball of}\ X^{**}\ \text {closed in the norm of}\ \{X^{**};\Vert \cdot \Vert _{**}\}. \end{array} \end{aligned}$$

Since the natural injection \(X\ni x\rightarrow f_x\in X^{**}\) defined by (14.2) is an isometric isomorphism, the ball B when regarded as a subset of \(B^{**}\) is a closed subset of \(B^{**}\) and the inclusion is proper unless X is reflexive. If \(B^{**}\) is given the \(\mathcal {W}_{**}\) topology, by the Alaoglu Theorem is weak\({}^*\) compact, and since \(\mathcal {W}_{**}\) is Hausdorff, \(B^{**}\) is both \(\Vert \cdot \Vert _{**}\)-closed and \(\mathcal {W}_{**}\)-closed (Proposition 5.1-(iii) of Chap. 2). Set

$$\begin{aligned} B_{**}=\left\{ \genfrac{}{}{0.0pt}{}{\text {the}\ \mathcal {W}_{**}\text {-closure of norm-closed unit ball of}\ X}{\text {regarded as a subset of}\ \{X^{**};\mathcal {W}_{**}\}} \right\} \end{aligned}$$

The next proposition asserts that while B with its norm topology is a subset \(B^{**}\), with in general strict inclusion, when it is given the \(\mathcal {W}_{**}\) topology, it is actually dense in \(B^{**}\).

Proposition 16.1c

([57]) \(B_{**}=B^{**}\).

Proof

\(B_{**}\) is a closed and convex subset of \(B^{**}\) (10.1-(iii) of the Complements of Chap. 2). If \(B_{**}\subset B^{**}\) with proper inclusion, select and fix \(x^{**}\in B^{**}-B_{**}\). By Proposition 11.3 there exists a nontrivial, bounded, linear functional f on \(\{X^{**};\mathcal {W}_{**}\}\) and real numbers \(\alpha <\beta \) such that

$$\begin{aligned} f(y)\le \alpha <\beta \le f(x^{**})\quad \text { for all }\>y\in B_{**}. \end{aligned}$$

By Proposition 15.1c, such a functional f is of the form (16.2c). Therefore, there exists \(x^*\in X^*\) such that

$$\begin{aligned} x^*(y)\le \alpha <\beta \le x^{**}(x^*)\quad \text { for all }\>y\in B. \end{aligned}$$
(16.4c)

Now \(y\in B\) implies \(-y\in B\). Therefore

$$\begin{aligned} \Vert x^*\Vert =\sup _{\Vert y\Vert =1}x^*(y)\le \alpha . \end{aligned}$$

On the other hand, since \(x^{**}\in B^{**}\)

$$\begin{aligned} \alpha <\beta \le x^{**}(x^*)\le \Vert x^{**}\Vert \,\Vert x^*\Vert \le \alpha . \end{aligned}$$

This contradicts (16.4c) and proves the proposition. \(\blacksquare \)

Corollary 16.1c

The Banach space \(\{X;\Vert \cdot \Vert \}\) when regarded as a subspace of \(X^{**}\) and equipped with the weak\({}^*\) topology of \(X^{**}\), is dense in \(X^{**}\).

1.2 16.2c Characterizing Reflexive Banach Spaces

The next statements follow from the previous proposition, upon observing that the \(\mathcal {W}_{**}\) topology of X, as a subset of \(X^{**}\), is precisely the weak topology generated on X by \(X^*\).

Corollary 16.2c

A Banach space \(\{X;\Vert \cdot \Vert \}\) is reflexive if and only if its closed unit ball is weakly compact.

Corollary 16.3c

A Banach space \(\{X;\Vert \cdot \Vert \}\) is reflexive if and only if a bounded and weakly closed set is weakly compact.

Thus in some sense, reflexive Banach spaces are those for which a version of the Heine–Borel Theorem holds (Proposition 6.4 of Chap. 2).

16.6.:

Some of the following statements contain fallacies. Identify and disprove them, by an argument or a counterexample.

i.:

The closed unit ball B of a normed space \(\{X;\Vert \cdot \Vert \}\), is convex and hence weakly closed (Proposition 12.2).

ii.:

Regarding B as a subset of \(B^{**}\) its \(\mathcal {W}_{**}\) topology coincides precisely with its weak topology.

iii.:

Therefore B as a subset of \(B^{**}\) is \(\mathcal {W}_{**}\) closed and hence \(\mathcal {W}_{**}\)-compact, since it is a closed subset of a compact set.

iv.:

Since the \(\mathcal {W}_{**}\) topology on \(B\subset B^{**}\) coincides with its weak topology, B is weakly compact.

1.3 16.3c Metrization Properties of the Weak Topology of the Closed Unit Ball of a Banach Space

Proposition 16.2c

Let \(\{X;\Vert \cdot \Vert \}\) be a Banach space. Then the weak topology of the closed unit ball \(B\subset X\) is metric, if and only if the dual \(X^*\) is separable.

Proof

If \(X^*\) is separable, the weak\({}^*\) topology of \(B^{**}\subset X^{**}\) is metric. Identify X as a subset of \(X^{**}\) by the natural injection map \(X\ni x\rightarrow f_x\in X^{**}\) and observe that the weak\({}^*\) topology inherited by X as a subset of \(X^{**}\) is precisely the weak topology of X.

Assume next that the weak topology of the closed unit ball \(B\subset X\) is metric. There exists a countable collection \(\{\mathcal {O}_n\}\) of weakly open neighborhoods of the origin such that every weak neighborhood of the origin contains some \(\mathcal {O}_n\). Without loss of generality the \(\mathcal {O}_n\) can be taken of the form

$$\begin{aligned} \mathcal {O}_n=\left\{ \genfrac{}{}{0.0pt}{}{x\in X\bigm | |T_{j_n}(x)|<\varepsilon _n\>\text { for } \> \varepsilon _n>0\> \text { and }\> T_{j_n}\in X^*}{\text {for }\> j_n\in \{j_{n,1},\dots ,j_{n,k}\}\subset \mathbb {N}}\right\} \end{aligned}$$

The collection of \(\{T_{j_n}\}\) is countable and its closed linear span is separable. We claim that \(\overline{\{T_{j_n}\}}=X^*\). If not pick \(\eta ^*\in X^*-\overline{\{T_{j_n}\}}\) at distance \(\delta \in (0,1)\) from \(\overline{\{T_{j_n}\}}\) and determine a nontrivial, bounded linear functional \(x^{**}\in X^{**}\), of norm \(\Vert x^{**}\Vert \le 1\), vanishing on \(\overline{\{T_{j_n}\}}\) and such that \(x^{**}(\eta ^*)=\delta \) (Proposition 10.3). The set

$$\begin{aligned} \mathcal {O}_\delta =\big \{x\in X\bigm | |\eta ^*(x)|<{\textstyle \frac{1}{2}}\delta \big \} \end{aligned}$$

is a weak neighborhood of the origin of X and hence \(\mathcal {O}_n\subset \mathcal {O}_\delta \) for some n. Since \(x^{**}\in B^{**}\), by the density Proposition 16.1c, there exists \(x\in B\) such that

$$\begin{aligned} |x^{**}(T_{j_n})-T_{j_n}(x)|<\varepsilon _n\>\text { for all }\> j_n\in \{j_{n,1},\dots ,j_{n,k}\} \end{aligned}$$

and simultaneously

$$\begin{aligned} |\delta -\eta ^*(x)|=|x^{**}(\eta ^*)-\eta ^*(x)|<\varepsilon _n. \end{aligned}$$

By picking \(\varepsilon _n\) sufficiently small we may insure that \(\varepsilon _n<\frac{1}{2}\delta \). Then, since \(x^{**}\) vanishes on \(\overline{\{T_{j_n}\}}\), the first of these gives,

$$\begin{aligned} |T_{j_n}(x)|<\varepsilon _n\quad \text { for all }\> j_n\in \{j_{n,1},\dots ,j_{n,k}\} \end{aligned}$$

which implies that \(x\in \mathcal {O}_n\). However the second of these implies \(|\eta ^*(x)|>\frac{1}{2}\delta \). That is \(x\notin \mathcal {O}_\delta \subset \mathcal {O}_n\). \(\blacksquare \)

16.7.:

The weak topology of the closed unit ball of \(L^1(E)\) and \(\ell _1\) are not metrizable.

1.4 16.4c Separating Closed Sets in a Reflexive Banach Space

Proposition 16.3c

Let \(C_1\) and \(C_2\) be two nonempty, disjoint, closed subsets of a reflexive Banach space \(\{X;\Vert \cdot \Vert \}\). Assume that at least one of them is bounded. There exists a bounded linear functional T on X and real numbers \(0<\alpha <\beta \) such that (11.2) holds. Equivalently, any two closed subsets of a reflexive Banach space, can be strictly separated, provided at least one of them is bounded.

Proof

If \(C_1\) is closed and bounded it is \(\text {weak}^*\) compact. The statement then follows from Proposition 11.3. \(\blacksquare \)

Remark 16.1c

The requirement that at least one of the two closed sets \(C_1\) and \(C_2\) be bounded is essential, in view of the counterexample in § 11.1c

17c Hilbert Spaces

1.1 17.1c On the Parallelogram identity

The parallelogram identity is equivalent to the existence of an inner product on a vector space X in the following sense.

A scalar product \(\langle \cdot ,\cdot \rangle \) on a vector space X generates a norm \(\Vert \!\cdot \!\Vert \) on X that satisfies the parallelogram identity. Vice versa, let \(\{X;\Vert \cdot \Vert \}\) be a normed space whose norm \(\Vert \cdot \Vert \) satisfies the parallelogram identity. Then setting

$$\begin{aligned} 4\langle x,y\rangle =\Vert x+y\Vert ^2-\Vert x-y\Vert ^2 \end{aligned}$$

defines a scalar product in X.

17.2.:

If \(p\ne 2\) then \(L^p(E)\) is not a Hilbert space.

17.3.:

Let \(\{x_n\}\) and \(\{y_n\}\) be Cauchy sequences in a Hilbert space H. Then \(\{\langle x_n,y_n\rangle \}\) is a Cauchy sequence in \(\mathbb {R}\).

18c Orthogonal Sets, Representations and Functionals

18.1.:

For an n-tuple \(\{x_1,x_2,\dots ,x_n\}\) of orthogonal elements in a Hilbert space H

$$\begin{aligned} \Big \Vert \mathop {\textstyle {\sum }}\limits _{i=1}^n x_i\Big \Vert ^2= \mathop {\textstyle {\sum }}\limits _{i=1}^n\Vert x_i\Vert ^2\qquad \text {(Pythagora's theorem)} \end{aligned}$$
18.2.:

Let E be a subset of H. Then \(E^\perp \) is a linear subspace of H and \(\left( E^\perp \right) ^\perp \) is the smallest, closed, linear subspace of H containing E.

18.3.:

Every closed convex set of H has a unique element of least norm. More generally, let C be a weakly closed subset of a reflexive Banach space. Then the functional \(h(x)=\Vert x\Vert \) takes its minimum on C, i.e., there exists \(x_o\in C\) such that \(\inf _{x\in C}\Vert x\Vert =\Vert x_o\Vert \).

18.4.:

Let E be a bounded open set in \(\mathbb {R}^N\) and let \(f\in C({\bar{E}})\). Denote by \(\mathcal {P}_n\) the collections of polynomials of degree at most n in the coordinate variables. There exists a unique \(P_o\in \mathcal {P}_n\) such that

$$\begin{aligned} \int _E|f-P|^2dx\ge \int _E|f-P_o|^2dx \qquad \text { for all }\> P\in \mathcal {P}_n. \end{aligned}$$

19c Orthonormal Systems

19.1.:

Let H be a Hilbert space and let \(\mathcal {S}\) be an orthonormal system in H. Then for any pair xy of elements in H

$$\begin{aligned} \mathop {\textstyle {\sum }}\limits _{\mathbf {u}\in \mathcal {S}}|\langle \mathbf {u},x\rangle | |\langle \mathbf {u},y\rangle |\le \Vert x\Vert \Vert y\Vert . \end{aligned}$$
19.2.:

Let \(\mathcal {S}\) be an orthonormal system in H and denote by \(H_o\) the closure of the linear span of \(\mathcal {S}\). The projection of an element \(x\in H\) into \(H_o\) is defined by

$$\begin{aligned} x_o=\mathop {\textstyle {\sum }}\limits _{\mathbf {u}\in \mathcal {S}}\,\langle x,\mathbf {u}\rangle \mathbf {u}. \end{aligned}$$

Such a formula defines \(x_o\) uniquely. Moreover \(x_o\in H_o\) and \((x-x_o)\perp H_o\).

19.3.:

A Non Separable Hilbert Space: In \(L^2_{{\text {loc}}}(\mathbb {R})\) with respect to the Lebesgue measure, define

$$\begin{aligned} L^2_{{\text {loc}}}(\mathbb {R}) \ni f,g \rightarrow \langle f,g \rangle \buildrel {\text {def}}\over {=} \lim _{\rho \rightarrow \infty } \frac{1}{\rho } \int _{-\rho }^{\rho } fgdx. \end{aligned}$$

Set

$$\begin{aligned} \begin{aligned} H_o&=\{f\in L^2_{{\text {loc}}}(\mathbb {R})\bigm | \Vert f\Vert =0\}\\ H_1&=\{ f\in L^2_{{\text {loc}}}(\mathbb {R})\bigm |\Vert f\Vert <\infty \}. \end{aligned} \end{aligned}$$

Since \(H_o\) contains nonzero elements, \(\Vert \cdot \Vert \) is not a norm. Introduce the quotient space

$$\begin{aligned} H=\frac{H_1}{H_o}\quad \text { of equivalence classes } \> f+H_o\quad \text { for }\> f\in H_1. \end{aligned}$$

Verify that \(\langle \cdot ,\cdot \rangle \) is an inner product and \(\Vert \cdot \Vert \) is a norm on H. The system

$$\begin{aligned} \big \{\sin \alpha x+H_o\big \}_{\alpha \in \mathbb {R}} \end{aligned}$$

is orthonormal in H, and uncountable. Therefore H is non separable, by Proposition 19.2.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer Science+Business Media New York

About this chapter

Cite this chapter

DiBenedetto, E. (2016). Banach Spaces. In: Real Analysis. Birkhäuser Advanced Texts Basler Lehrbücher. Birkhäuser, New York, NY. https://doi.org/10.1007/978-1-4939-4005-9_7

Download citation

Publish with us

Policies and ethics