• 沒有找到結果。

For f, g ∈ F(S, R) and a ∈ R, we define the sum f + g : S → R of f and g and the scalar multiplication (af

N/A
N/A
Protected

Academic year: 2022

Share "For f, g ∈ F(S, R) and a ∈ R, we define the sum f + g : S → R of f and g and the scalar multiplication (af"

Copied!
5
0
0

加載中.... (立即查看全文)

全文

(1)

AN INTRODUCTION TO BANACH AND HILBERT SPACES

FRANK LIOU

1. Some Important Function Spaces:

Let S be a nonempty set. The set of all real valued functions from S to R is denoted by F (S, R). For convenience, F(S, R) is denoted by F(S). For f, g ∈ F(S, R) and a ∈ R, we define the sum f + g : S → R of f and g and the scalar multiplication (af ) : S → R of f by a as follows. For s ∈ S,

(f + g)(s) = f (s) + g(s), (af )(s) = af (s).

The zero function 0 : S → R is defined to be 0(s) = 0 for all s ∈ S.

Proposition 1.1. The set F (S, R) forms a real vector space.

Proof. The proof is left to the reader. 

Proposition 1.2. Let S be a finite set and N be the number of elements of S. Then F (S) is linear isomorphic to RN :

F (S) ∼= RN.

Proof. Assume that S contains N distinct element and denote S by {s1, · · · , sN}. For each f ∈ F (S), we define

T (f ) = (f (s1), · · · , f (sN)).

Let us check that T : F (S) → RN is a linear isomorphism. Checking the linearity of T is routine. Let us show that the map is bijective. For a = (a1, · · · , aN) ∈ RN, we define a function f : S → R by f (si) = ai. Then T (f ) = a. For f ∈ ker T, then f (s1) = · · · = f (sN) = 0. Thus f is the zero function. We find ker T = {0}. 

When S = {1, · · · , N }, there is a canonical choice of linear isomorphism between F ({1, · · · , N }) and RN. For f ∈ F ({1, · · · , N }), we denote f (i) by fi. The canonical choice of linear iso- morphism T : F ({1, · · · , N }) → RN is defined by T (f ) = (f1, · · · , fN). Thus RN can be defined as the space of functions a : {1, · · · , N } → R.

Definition 1.1. A function f ∈ F (S) is said to be bounded if there exists M > 0 such that

|f (s)| ≤ M for all s ∈ S.

Let B(S) be the subset of F (S) consisting of bounded functions.

Proposition 1.3. The subset B(S) forms a real vector subspace of F (S).

Proof. Let f, g ∈ B(S). Choose two positive real numbers M1 and M2 such that

|f (s)| ≤ M1, |g(s)| ≤ M2 for all s ∈ S. By triangle inequality,

|f (s) + g(s)| ≤ |f (s)| + |g(s)| ≤ M1+ M2 for any s ∈ S.

1

(2)

Hence M1+ M2 is an upper bound for {|f (s) + g(s)| : s ∈ S}. Thus f + g : S → R is a bounded function, i.e. f + g ∈ B(S). Since |0(s)| = |0| ≤ 1 for any s ∈ S, 0 ∈ B(S). For any a ∈ R, |af (s)| ≤ |a|M1 for any s ∈ S, af ∈ B(S).

 Definition 1.2. A norm on a real vector space V is a function

k · k : V → R such that

(1) kavk = |a|kvk for any a ∈ R and for any v ∈ V ; (2) kvk ≥ 0; kvk = 0 if and only if v = 0;

(3) kv + wk ≤ kvk + kwk for any v, w.

A real vector space together with a norm (V, k · k) is called a real normed (vector) space.

Lemma 1.1. For each f ∈ B(S), we set

kf k= sup

s∈S

|f (s)|.

The function k · k: B(S) → R defines a norm on B(S).

Proof. For a ∈ R,

kaf k= sup

s∈S

|af (s)| = sup

s∈S

|a||f (s)| = |a| sup

s∈S

|f (s)| = |a|kf k.

For any s ∈ S, |f (s)| ≥ 0 and hence kf k ≥ 0. If kf k = 0, then 0 ≤ |f (s)| ≤ sups∈S|f (s)| = 0 for any s ∈ S. Therefore |f (s)| = 0 for all s ∈ S. This implies that f is the zero function, i.e. f (s) = 0 for all s ∈ S. For any f, g ∈ B(S) and any s ∈ S,

|f (s) + g(s)| ≤ |f (s)| + |g(s)| ≤ kf k+ kgk.

We find that kf k+ kgk is an upper bound of {|f (s) + g(s)| : s ∈ S}. By definition of

sup, kf + gk≤ kf k+ kgk. 

Any norm k · k on a vector space V defines a metric d on V by d(v, w) = kv − wk.

This metric is called the metric induced from k · k.

Definition 1.3. A real normed space V is called a real Banach space if V together with the metric induced from the norm is a complete metric space. A real Banach space is a complete real normed space.

Lemma 1.2. Let (an) be a sequence of real numbers convergent to a ∈ R. Suppose that M > 0 is a real number such that |an| ≤ M for all n ≥ 1. Then |a| ≤ M.

Proof. Suppose not. Assume that |a| > M. Take  = (|a| − M )/2. Since (an) is convergent to a in R, we can choose N ∈ N so that |an− a| <  = |a|−M2 < |a| − M when n ≥ N. This implies that |an| ≥ |a| − |an− a| > M whenever n ≥ N. This leads to the contradiction to

our assumption. 

Theorem 1.1. The normed space (B(S), k · k) is a real Banach space.

(3)

Proof. Let (fn) be a Cauchy sequence in B(S). For any  > 0, there exists N ∈ N such that kfn− fmk< /2. whenever n, m ≥ N. For s ∈ S,

|fn(s) − fm(s)| ≤ kfn− fmk<  2

whenever n, m ≥ N. This shows that The sequence (fn(s)) of real numbers is a Cauchy sequence in R. By completeness of R, (fn(s)) is convergent in R. We denote the limit of (fn(s)) by f (s), i.e. f (s) = limn→∞fn(s). We obtain a function f : S → R. Let us prove that f ∈ B(S) and (fn) is convergent to f in B(S).

Since (fn) is a Cauchy sequence in B(S), it is bounded in B(S). There exists M > 0 so that kfnk≤ M for all n ≥ 1. For s ∈ S,

|fn(s)| ≤ kfnk≤ M.

We see that the sequence (fn(s)) of real numbers is bounded. By Lemma 1.2, we find

|f (s)| ≤ M for all s ∈ S. Thus f : S → R is a bounded function. We see that f ∈ B(S). By taking m → ∞ of |fn(s) − fm(s)| < /2, we find that for s ∈ S

|fn(s) − f (s)| ≤  2

whenever n ≥ N. This shows that kfn− f k ≤ /2 <  when n ≥ N. This proves that

(fn) is convergent to f in B(S). 

When S = {s1, · · · , sN} is a finite set, we can see that B(S) = F (S). In fact, for each f ∈ F (S), Im f = {f (s1), · · · , f (sN)} is a finite subset of R and hence Im f is bounded.

Moreover,

kf k= max{|f (s1)|, · · · , |f (sN)|}.

Using the isomorphism F (S) ∼= RN in Proposition 1.2, we can define the infinity norm on RN as follows. For each a = (a1, · · · , aN) in RN, we set

kak= max{|a1|, · · · , |aN|}.

We leave it to the reader to verify that (RN, k · k) is a real normed space. On RN, there are also other different norms. Let us recall that the Euclidean norm of a is given by kak = Pn

i=1|ai|21/2

. For each 1 ≤ p < ∞, define

kakp=

n

X

i=1

|ai|p

!1/p

.

Proposition 1.4. For 1 ≤ p ≤ ∞, (RN, k · kp) is a real Banach space. Moreover, kak= limp→∞kakp.

Proof. We leave it to the reader to verify that k · kp defines a norm on RN. See hw 7.

Let (an) be a Cauchy sequence in (RN, k · kp). For any  > 0, there exists N ∈ N such that kan− amkp < /2√p

N for any n, m ≥ N. Denote an by (a1n, · · · , aNn) for n ≥ 1. For 1 ≤ i ≤ N,

(1.1) |ain− aim| ≤ kan− amkp<  2√p

N

whenever n, m ≥ N. This shows that (ain) is a Cauchy sequence in R for 1 ≤ i ≤ N. By completeness of R, (ain) are all convergent. Let ai be the limit of (ain) for all i. By taking the limit of (1.1) as m → ∞, we find

|ain− ai| ≤  2√p

N < 

p

N

(4)

whenever n ≥ N. Let a = (a1, · · · , an) ∈ RN. Then

kan− akpp =

N

X

i=1

|ain− ai|p < p

N · N = p

whenever n ≥ N. Thus kan− akp <  whenever n ≥ N. We prove that (an) is convergent

to a in (RN, k · kp). 

When S = N, we call F(N) the space of sequences of real numbers. In this case, we denote the Banach space (B(N), k · k) by l(N). For 1 ≤ p < ∞. let lp(N) be the subset of F (N) consisting of sequences a : N → R such thatP

i=1|a(i)|p < ∞. For each a ∈ lp(N ), we define

kakp =

X

i=1

|a(i)|p

!1/p

Lemma 1.3. For each 1 ≤ p < ∞, we obtain a normed space (lp(N), k · kp).

Proof. The proof basically follows from Holder inequality and the Minkowski inequality for infinite series and is almost the same as that of (RN, k · kp).  We simply denote the norm space (lp(N), k · kp) by lp(N). In fact, we can prove further that:

Theorem 1.2. For 1 ≤ p ≤ ∞, the normed space lp(N) is a real Banach space.

Proof. The proof is very similar to that of the completeness of (RN, k · kp). The details are

left to the reader as an exercise. See hw 7. 

Lemma 1.4. For a, b ∈ l2(N), the infinite seriesP

i=1a(i)b(i) is absolutely convergent.

For a, b ∈ l2(N), we define

ha, bi =

X

i=1

a(i)b(i).

Then h·, ·i defines an inner product on l2(N).

Definition 1.4. Let V be a real vector space. An inner product on V is a function h·, ·i : V × V → R such that for any u, v, w ∈ V and a, b ∈ R,

(1) hau + bv, w·i = ahu, wi + bhv, wi, (2) hu, vi = hv, ui,

(3) hu, ui ≥ 0 and = 0 if and only if u = 0.

A real vector space V together with an inner product h·, ·i is called a real inner product space.

An inner product h·, ·i on V induces a norm on V defined by kvk =phv, vi for v ∈ V . We call k · k the norm induced from h·, ·i.

Definition 1.5. An inner product space (V, h·, ·i) is called a real Hilbert space if V together with the normed induced from h·, ·i is a real Banach space. A real Hilbert space is a complete inner product space.

(5)

For a ∈ l2(N), we see that

ha, ai = kak2. This implies that

Theorem 1.3. The inner product space (l2(N), h·, ·i) is a real Hilbert space.

We can also consider complex Banach spaces and complex Hilbert spaces. We leave the definitions to the readers.

參考文獻

相關文件

The question of whether every Cauchy sequence in a given inner product space must converge is very important, just as it is in a metric or normed space.. We give the following name

We complete the proof of

Proof of Inverse Function Theorem..

The ordered basis β ∗ is called the dual basis to β..

Theorem (Comparison Theorem For Functions) Suppose that a ∈ R, that I is an open interval that contains a, and that f,g are real functions defined everywhere on I except possibly at

 This theorem implies that the completion of a metric space is unique up to isomorphisms..

Let {f n } be a sequence of real-valued continuous functions on [a, b] which converges uniformly to a continuous function f on

A subset N of a metric space (M, d) is called sequentially compact if the metric subspace (N, d N ) is sequentially compact.. We obtain another equivalence of