Functional Analysis

From testwiki
Revision as of 07:14, 27 January 2008 by imported>TakuyaMurata (2. Normed space)
(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)
Jump to navigation Jump to search

0. Preface

Functional Analysis can mean many different things to different mathematicians. The core of the subject, however, is to study linear spaces with some topology which allows us to do analysis; ones like spaces of functions, spaces of operators acting on the space of functions, etc.

The book consists of two parts. The first part covers the basics of Banach spaces theory with the emphasis on its applications. The second part covers topological vector spaces, especially locally convex ones, generalization of banach spaces. In both parts, we give principal results e.g., the closed graph theorem, resulting in some repetition. One reason for doing this is that one often only needs a Banach-version of such results. Another reason is that the approach seems more pedagogically sound; the statement of the results in their full generality may (at first) look cryptical to students new to them.

Knowledge of measure theory will not be needed except for Chapter 3.4 5 (Lp spaces). On the other hand, solid knowledge in general topology is mandatory.

1. Preliminaries

In this chapter we gather some standard results, some of which are not necessarily in functional analysis proper yet will be used later, primarily to fix language and formulation. We prefer to omit proofs for them since they can be easily found somewhere else.

1. Theorem (Ascoli) Let be a family of continuous real-valued functions on a compact space K. If

(i) For each xK, supf|f(x)|<.
(ii) For each ϵ>0, every point xK has a neighborhood ω such that |f(y)f(x)|<ϵ for every yω and every f (i.e., is equicontinuous),

then

is totally bounded in the metric d(f,g)=supK|fg|.

1. Theorem (Heine-Borel) A totally bounded subset of a complete metric space is relatively compact.

As a corollary, the conclusion of the Ascoli's theorem thus says the closure of is compact since C(K) is complete.

A subset of a topological space is said to be meager if it is a union of countably many subsets of the space that have dense complement.

1. Theorem (Baire) Every complete metric space is non-meager.
Proof: (TODO: we should provide one.)

By modifying the proof so to use compact sets instead of balls we can show that every locally compact (Hausdorff?) space is also non-meager, though we will not be needing this.

1. Exercise Show by contradiction that the set of real numbers is uncountable, using the theorem.

In particular, a dense set may be meager.

1. Exercise Give an example of a dense but meager set. (Hint: you can find such an example in the book.)

1. Theorem (Tychonoff) Every product space of a nonempty collection of compact spaces is compact.

An linear operator from a linear space to a scalar field is called a linear functional.

1. Theorem (Hahn-Banach) Let 𝒳 be a real vector space and p be a function on 𝒳 such that p(x+y)p(x)+p(y) and p(tx)=tp(x) for any x𝒳 and any t>0. If 𝒳 is a closed subspace and f is a linear functional on such that fp, then f admits a linear extension F defined in 𝒳 such that Fp.
Proof: First suppose that 𝒳={x+tz;x,t} for some z∉. By hypothesis we have:

f(x)+f(y)p(xz)+p(y+z) for all x,y,

which is equivalent to:

sup{f(x)p(xz);xM}inf{p(y+z)f(y);yM}.

Let c be some number in between the sup and the inf. Define F(x+tz)=f(x)+tc for x,t>0. It follows that F is an desired extension. Indeed, f=F on being clear, we also have:

F(x+tz)tp(xt+z) if t>0

and

F(x+tz)tp(xtz) if t<0.

Let Ω be the collection of pairs (H,gH) where H is linear space with X and gH is a linear function on that extends f and is dominated by p. It can be shown that Ω is partially ordered and the union of every totally ordered sub-collection of Ω is in Ω (TODO: need more details). Hence, by Zorn's Lemma, we can find the maximal element (L,gL) and by the early part of the proof we can show that =𝒳.

We remark that a different choice of c in the proof results in a different extension. Thus, an extension given by the Hahn-Banach theorem in general is not unique.

1. Exercise State the analog of the theorem for complex vector spaces and prove that this version can be reduced to the real version. (Hint: Ref(ix)=Imf(x))

(TODO: mention moment problem.)

1 Lemma Let g,f1,...,fn be linear functionals on the same linear space. g is a linear combination of f1,...,fn if and only if j=1nkerfjkerg.
Proof: The direct part is clear. We shall show the converse by induction. Suppose n=1. We may suppose that there is y such that f1(y)=1. For any x, since xf1(x)ykerf1kerg,

g(xf1(x)y)=0 or g(x)=g(y)f1(x).

The basic case is thus proven. Now, suppose the lemma holds for some n1. As before, we may suppose that there is some y such that

fn(y)=1 while f1(y)=f2(y)=...=fn1(y)=0.

For any xj=1n1kerfj, since fk(xfn(x)y)=0 for every k=1,2,...,n, g(xfn(x)y)=0. Hence, the application of the inductive hypothesis to the linear functional xg(xfn(x)y) gives:

g(xfn(x)y)=j=1n1ajfj(x)

for some scalars a1,...,an. (TODO: the proof can be clearer and simpler).

Part 1: Banach spaces

2. Normed space

Template:Stage

Let 𝒳 be a linear space. A norm is a real-valued function f on X, with the notation =f(), such that

  • (i) x+yx+y (w:triangular inequality)
  • (ii) λx=|λ|x for any scalar λ
  • (iii) x=0 implies x=0.

(ii) implies that 0=0. This and (ii) then implies 0=xxx+x=2x for all x; that is, norms are always non-negative. With the metric d(x,y)=xy a normed space is a metric space. If only (i) and only (ii) hold, the function is called a seminorm.

A linear space with a norm is called a normed space. We define the operator norm of a continuous linear operator f between normed spaces 𝒳 and 𝒴, denoted by f, by

f=supx𝒳1f(x)𝒴

2 Theorem Let T be a linear operator from a normed space 𝒳 to a normed space 𝒴.

  • (i) T is continuous if and only if there is a constant C>0 such that TxCx for all xX
  • (ii) f=inf{ any C as in (i) }, and f=supx=1f(x) if 𝒳 has nonzero element.

Proof: See w:bounded operator and see w:operator norm.

A complete normed space is called a Banach space. While there is seemingly no prototypical example of a Banach space, we still give one example of a Banach space: 𝒞(K), the space of all continuous functions on a compact space 𝒦, can be identified with a Banach space by introducing the norm:

=supK||

It is a routine exercise to check that this is indeed a norm. The completeness holds since, from real analysis, we know that a uniform limit of a sequence of continuous functions is continuous. In concrete spaces like this one, one can directly show the completeness. More often than that, however, we will see that the completeness is a necessary condition for some results (especially, reflexivity), and thus the space has to be complete. The matter will be picked up in the later chapter.

2 Theorem (completion) Every normed space can be identified with a dense subspace of some Banach space.

The completeness is essential, for example, in the next result.

2 Theorem Let 𝒳 be a normed linear space, and 𝒴 be a Banach space. Let 𝒳 be a subspace. If is dense, then a continuous linear operator T:𝒴 admits a unique extension to 𝒳.
Proof: For each x𝒳, let xn be Cauchy such that xnx. By continuity TxnTxmTxnxm0 as n,m. Since 𝒴 is complete, limnTxn exists in 𝒴, and we denote it by T~x. From this definition It is clear that the extension T~ is linear. Moreover, if there is another extension S, then Sx is the limit of Txn; whence, the extension is unique. Finally, T~ is continuous since

T~x=T~limnxnlimnTxn=Tx.

2 Theorem (open mapping theorem) Let 𝒳,𝒴 be Banach spaces. If T:𝒳𝒴 is a continuous linear surjection, then it is an open mapping; i.e., it maps open sets to open sets.
Proof: Let B(r)={x𝒳;x<r}. Since T is surjective, n=1T(B(n))=T(n=1B(n))=T(X)=Y. Then by Baire's Theorem

A linear operator from a normed space 𝒳 to 𝒴 is said to be closed if its graph, that is the set {(x,Tx);xdom(𝒳)}, is closed in 𝒳×𝒴.

2 Corollary If T is a continuous linear operator between Banach spaces with closed range, then there exists a K>0 such that if yim(T) then xKy for some x with Tx=y.
Proof: Since T is an open mapping, there exists a δ>0 such that

Bδ(T(0))T(B1(0))

where by Br(x) we denote an open ball of radius r about x. By replacing δ with δ/2 we can also archive that

Bδ(0)T(B1(0))

Let yim(T)0 be given. Then there is z with Tz=y. Since T(δyz)=δ, there exists z0B1(0) such that T(z0)=T(δyz). Let x=yδz0. Then

x=yδz01δy and Tx=y

Finally, if y=0, just take x=0.

2 Exercise Use a quotient map to simplify the proof. (Hint: the answer will be given in the chapter 4)

2 Corollary If (𝒳,1) and (𝒳,2) are Banach spaces, then the norms 1 and 2 are equivalent; i.e., each norm is dominated by the other.
Proof: Let I:(𝒳,1+2)(𝒳,1) be the identity map. Then we have:

I1=1(1+2).

This is to say, I is continuous. Since Cauchy sequences apparently converge in the norm 1+2, the open mapping theorem says that the inverse of I is also continuous, which means explicitly:

1+2=I11+I12I11.

By the same argument we can show that 1+2 is dominated by 2

The technique using the corollary will be exploited in the proof of the next result.

2 Theorem (closed graph theorem) Let 𝒳,𝒴 be Banach spaces, and T:𝒳𝒴 a linear operator. The following are equivalent.

  • (i) T is continuos.
  • (ii) If xj0 and Txj is convergent, then Txj0.
  • (iii) The graph of T is closed.

Proof: That (i) implies (ii) is clear. To show (iii), suppose (xj,Txj) is convergent in X. Then xj converges to some x0 or xjx00, and TxjTx is convergent. Thus, if (ii) holds, T(xjx)0. Finally, to prove (iii) (i), we note that Corollary 2.something gives the inequality:

+TK

since by hypothesis the norm in the left-hand side is complete. Hence, if xjx, then TxjTx.

Note that when the domain of a linear operator is not a Banach space (e.g., just dense in a Banach space), the condition (ii) is not sufficient for the graph of the operator to be closed. A function is said to be densely defined if its domain is dense.

2 Exercise Let T be a closed densely defined operator and S be a linear operator with dom(T)dom(S). If there are constants a,b such that (i) 0a<1 and b>0 and (ii) S(u)aT(u)+bu for every udom(T), then T+S is a closed densely defined operator.

2 Corollary A linear functional u is continuous if and only if it has closed kernel.
Proof: ker(u)=u1({0}) is closed if u is continuous. To show the converse, suppose xj0 and u(xj) is convergent. Since the corollary is obvious when u is identically zero, we may suppose that there is z such that u(z)=1. Then the sequence xju(xj)z has a limit in the kernel of u, since the kernel is closed. It follows:

0=u(limjxju(xj)z)=u(limjxj)+u(limju(xj)z)=limju(xj).

When 𝒳,𝒴 are normed spaces, by L(𝒳,𝒴) we denote the space of all continuous linear operators from 𝒳 to 𝒴.

2 Theorem If 𝒴 is complete, then every Cauchy sequence Tn in L(𝒳,𝒴) converges to a limit T and T=limnTn.
Proof: Let Tn be a Cauchy sequence in operator norm. For each x𝒳, since

Tn(x)Tm(x)TnTmx

and 𝒴 is complete, there is a limit y to which Tn(x) converges. Define T(x)=y. T is linear since the limit operations are linear. It is also continuous since T(x)=limTn(x)limTnx. Finally, limTnT=supx1limTn(x)T(x) and |TnT|TnT0 as n.

2 Theorem (uniform boundedness principle) Let be a family of continuos functions f:XY where Y is a normed linear space. Suppose that MX is non-meager and that:

sup{f(x):f}< for each xM

It then follows: there is some GX open and such that

(a) sup{f(x):f,xG}<

If we assume in addition that each member of is a linear operator and X is a normed linear space, then

(b) sup{f:f}<

Proof: Let Ej=f{xX;f(x)j} be a sequence. By hypothesis, Mj=1Ej and each Ej is closed since {xX;f(x)>j} is open by continuity. It then follows that some EN has an interior point y; otherwise, M fails to be non-meager. Hence, (a) holds. To show (b), making additional assumptions, we can find an open ball B=B(2r,y)EN. It then follows: for any f and any xX with x=1,

f(x)=r1f(rx+y)f(y)2r1N.

A family Γ of linear operators is said to be equicontinuous if given any neighborhood W of 0 we can find a neighborhood V of 0 such that:

f(V)W for every fΓ

The conclusion of the theorem, therefore, means that the family satisfying the hypothesis of the theorem is equicontinuous.

2 Corollary Let 𝒳,𝒴,𝒵 be normed spaces. Let T:𝒳×𝒴 be a bilinear or sesquilinear operator. If T is separately continuous (i.e., the function is continuous when all but one variables are fixed) and 𝒴 is complete, then T is continuous.
Proof: For each y𝒴,

sup{T(x,y)𝒵;x𝒳1}=T(,y)

where the right-hand side is finite by continuity. Hence, the application of the principle of uniform boundedness to the family {T(x,);x𝒳1} shows the family is equicontinuous. That is, there is K>0 such that:

T(x,y)𝒵Ky𝒴 for every x𝒳le1 and every y𝒴.

The theorem now follows since 𝒳×𝒴 is a metric space.

Since scalar multiplication is a continuous operation in normed spaces, the corollary says, in particular, that every linear operator on finite dimensional normed spaces is continuous. The next is one more example of the techniques discussed so far.

2. Theorem (Hahn-Banach) Let (𝒳,) be normed space and 𝒳 be a linear subspace. If z is a linear functional continuous on , then there exists a continuos linear functional w on 𝒳 such that z=w on and z=w.
Proof: Apply the Hahn-Banach stated in Chapter 1 with z as a sublinear functional dominating z. Then:

z=sup{w(x);x,x1}sup{w(x);x𝒳,x1}=wz;

that is, z=w.

2. Corollary Let be a subspace of a normed linear space 𝒳. Then x is in the closure of if and only if z(x) = 0 for any z𝒳* that vanishes on .
Proof: By continuity z()z(). Thus, if x, then z(x)z()={0}. Conversely, suppose x∉. Then there is a δ>0 such that yxδ for every y. Define a linear functional z(y+λx)=λ for y and scalars λ. For any λ0, since λ1y,

|z(y+λx)|=|λ|δ1δδ1|λ||λ1y+x=y+λx.

Since the inequality holds for λ=0 as well, z is continuous. Hence, in view of the Hahn-Banach theorem, z𝒳 while we still have z=0 on and z(x)0.

2 Corollary If (𝒳,) is a normed space and z(x)=z(y) for all z𝒳*, then x=y
Proof: Suppose xy, and define z(λ(xy))=λxy for scalars λ. Now, z is continuous since its domain is finite-dimensional, and so by the Hahn-Banach theorem z𝒳*. Thus, |z(x)z(y)|=xy0, or z(x)z(y).

Here is a classic application.

2 Theorem Let 𝒳,𝒴 be Banach spaces, T:𝒳𝒴 be a linear operator. If xn0 implies that (zT)xn0 for every z𝒳*, then T is continuous.
Proof: Suppose xn0 and Txny. For every z𝒳*, by hypothesis and the continuity of z,

0=limnz(Txn)=z(y).

Now, by the preceding corollary y=0 and the continuity follows from the closed graph theorem.

2 Theorem Let 𝒳 be a Banach space and E𝒳. Then E is bounded if and only if supE|f|< for every f𝒳*
Proof: By continuity,

sup{|f(x)|;xE}fsupE.

This proves the direct part. For the converse, define Txf=f(x) for xE,f𝒳*. By hypothesis

|Txf|supE|f| for every xE.

Thus, by the principle of uniform boundedness, there is K>0 such that:

|Txf|Kf for every xE,f𝒳*

Hence, in view of Theorem 2.something, for xE,

x=sup{|f(x)|;f𝒳*,f1}K.

2. Corollary Let (𝒳,) be Banach, fj𝒳* and 𝒳 dense and linear. Then fj(x)0 for every if and only if (1) supjfj< and fj(y)0 for every y.
Proof: Since fj is Cauchy, it is bounded. This shows the direct part. To show the converse, let x𝒳. If yj, then

|fj(x)||fj(xyj)|+|fj(yj)|=supjfjxyj+|fj(yj)|

By denseness, we can take yj so that yjx0.

2 Exercise Show the corollary may fail if fj is not bounded.

2 Theorem Let T be a continuous linear operator into a Banach space. If IT<1 where I is the identity operator, then the inverse T1 exists, is continuous and can be written by:

T1(x)=k=0(IT)k(x) for each x in the range of T.

Proof: For nm, we have:

k=mn(IT)k(x)xk=mnITk.

Since the series is geometric by hypothesis, the right-hand side is finite. Let Sn=k=0n(IT)k. By the above, each time x is fixed, Sn(x) is a Cauchy sequence and the assumed completeness implies that the sequence converges to the limit, which we denote by S(x). Since for each x supn1Sn(x)<, it follows from the principle of uniform boundedness that:

supn1Sn.

Thus, by the continuity of norms,

S(x)=limnSn(x)(supn1Sn)x.

This shows that S is a continuous linear operator since the linearity is easily checked. Finally,

TS(x)x=limn(IT)n+1(x)xlimnITn+1=0.

Hence, S is the inverse to T.

2 Corollary The space of invertible continuous linear operators 𝒳 is an open subspace of L(𝒳,𝒳).
Proof: If TL(𝒳,𝒳) and ST<1T1, then S is invertible.

If 𝐅 is a scalar field and 𝒳 is a normed space, then L(𝒳,𝐅) is called a dual of 𝒳 and is denoted by 𝒳*. In view of Theorem 2.something, it is a Banach space.

3. Hilbert space

Template:Stage

A Banach space is called a Hilbert space if for each ordered pair (x,y) of elements in the space there is a unique complex (or real) number called an inner product of x and y, denoted by x,y, subject to the following conditions:

  • (i) The functional f(x)=x,y is linear.
  • (ii) x,y=y,x
  • (iii) x,x=0 implies that x=0

We define x=x,x1/2 and this can be shown to be a norm. Indeed, it is clear that αx=|α|x and (iii) is the reason that x=0 implies that x=0. Finally, the triangular inequality follows from the next lemma.

3 Lemma (Schwarz's inequality) |x,y|xy where the equality holds if and only if we can write x=λy for some scalar λ.

If we assume the lemma for a moment, since Re(α)|α| for any complex number α it then follows:

x+y2 =x2+2Rex,y+y2
x2+2|x,y|+y2
(x+y)2

Proof of Lemma: First suppose x=1. If α=x,y, it then follows:

0αxy2=|α|22Re(αx,y)+y2=|α|2+y2

where the equation becomes 0 if and only if x=λy. Since we may suppose that x0, the general case follows easily.

3 Theorem A normed linear space is a pre-Hilbert space if and only if xy2=2x2+2y2x+y2 (polarization identity)
Proof: The direct part is clear. To show the converse, we define

x,y=41(x+y2xy2+ix+iy2ixiy2.

It is then immediate that x,y=y,x, x,y=x,y and ix,y=ix,y. Moreover, since the calculation:

x1+x2+y2x1+x2y2 =2x1+y22x1y2x1x2+y2x1x2y2
=j=12xj+y2xjy2,

we have: x1+x2,y=x1,y+x2,y. If α is a real scalar and αj is a sequence of rational numbers converging to α, then by continuity and the above, we get: αx,y=limjαjx,y=αx,y.

3 Lemma Let be a pre-Hilbert. Then xjx in norm if and only if for any y xjx and xjx,y0 as j.
Proof: The direct part holds since:

|xjx|+|xjx,y|xjx(1+y)0 as j.

Conversely, we have:

xjx2=xj22Rexj,x+x20 as j

3 Lemma Let D be a convex closed subset of a Hilbert space. Then D admits a unique element z such that

z=inf{x;xD}.

Proof: For each n=1,2,..., there is some xnD such that 0xnδn1. That is, δ=limnxn. Since D is convex,

xn+xm2D and so δ12xn+xm.

It follows:

xnxm2 =2xn2+2xm2xn+xm2
2xn2+2xm24δ2
2δ2+2δ24δ2=0 as n,m

This is to say, xn is Cauchy. Since D is a closed subset of a complete metric space, whence it is complete, there is a limit zD with z=δ. The uniqueness follows since if w=δ we have

zw2=2z2+2w2z+w2

where the right side is 0 for the same reason as before.

3 Theorem (orthogonal decomposition) Let be a Hilbert space and be a subspace. For every xH we can write

x=y+z

where y and z, and y and z are uniquely determined by x
Proof: Clearly x is convex, and it is also closed since the image of a closed set under translation is again closed. Now, the application of the above lemma gives an element y such that xy=inf{xw;w}. Let z=xy. This z turns out to be orthogonal to . Indeed, for any w with w=1, if λ=w,z, we have

z2x(y+λw)2=zλw2=z22Rez,λw+|λ|2=z2|λ|2

since y+λw. That is, 0|λ|2 or λ=0. The uniqueness holds since if x=y1+y2=y2+z2 and y1,y2 and z1,z2, then

y1y22=y1y2,y1y2=y1y2,z1z2 (since y1y2=z1z2).

The right-hand side is equal to 0 since z1z2. Thus, y1=y2.

3 Corollary Let be a subspace of a Hilbert space . Then

  • (i) ={0} if and only if is dense in .
  • (ii) =.

Proof: Left as an exercise.

3 Theorem (representation theorem) Every continuous linear functional f on a Hilbert space has the form:

f(x)=x,y with a unique y and f=y


Proof: Let =f1({0}). Since f is continuous and any finite set is closed in or , is closed. If =, then take y=0. If not, by the above corollary, there is a nonzero z orthogonal to . By replacing z with zz1 we may suppose that z=1. For any x, since zf(x)f(z)x is in the kernel of f and thus is orthogonal to z, we have:

0=zf(x)f(z)x,z=z,zf(x)f(z)x,z

and so:

f(x)=x,f(z)z

The uniqueness follows since x,y1=x,y2 for all x means that y1y2={0}. Finally, we have the identity:

y=|yy,y|fy

where the last inequality is Schwarz's inequality.

3 Exercise Using Lemma 1.6 give an alternative proof of the preceding theorem.

In view of Theorem 3.something, for each x, we can write: x=y+z where y, a closed subspace of , and z. Denote each y, which is uniquely determined by x, by π(x). The function π then turns out to be a linear operator. Indeed, for given x1,x2, we write:

x1=y1+z1,x2=y2+z2 and x1+x2=y3+z3

where yj and zj for j=1,2,3. By the uniqueness of decomposition

π(x1)+π(x2)=y1+y2=y3=π(x1+x2).

The similar reasoning shows that π commutes with scalars. Now, for x=y+z (where y and z), we have:

x2=π(x)2+z2π(x)2

That is, π is continuous with π1. In particular, when is a nonzero space, there is x0 with π(x0)=x0 and x0=1 and consequently pi=1. Such π is called an orthogonal projection.

The next theorem gives an alternative to the Hahn-Banach theorem.

3 Theorem Let be a linear (not necessarily closed) subspace of a Hilbert space. Every continuous linear functional on can be extended to a unique continuous linear functional on that has the same norm and vanishes on
Proof: Define g=fπ. By the same argument used in the proof of Theorem 2.something (Hahn-Banach) and the fact that π=1, we obtain f=g. Since g=0 on , it remains to show the uniqueness. For this, let h be another extension with the desired properties. Since the kernel of fh is closed and thus contain , f=h on . Hence, for any x,

h(x)=(hπ)x=(fπ)x=g(x).

The extension g is thus unique.

3 Theorem Let n be an increasing sequence of closed subspaces, and be the closure of 12.... If π is an orthogonal projection onto , then for every x πn(x)x. Proof: Let 𝒩={x;πn(x)x(n)}. Then 𝒩 is closed. Indeed, if xj𝒩 and xjx, then

πn(x)x2xxj+πn(xj)xj

and so x𝒩. Since 𝒩, the proof is complete.

3 Lemma (Bessel's inequality) If uk is a sequence orthonormal in a Hilbert space , then

k=1|x,uk|2x2 for any x.

Proof: If x,y=0, then x+y2=x2+y2. Thus,

xk=1nx,uk2=x22Rek=1n|x,uk|2+k=1n|x,uk|2=x2k=1n|x,uk|2.

Letting n completes the proof. .

2 Lemma A normed space (𝒳,) is complete (in the sense of metric spaces) if and only if j=1xj< implies j=1xj exists.
Proof: If j=1xj<, then since

j=1nxjj=1mxj=j=m+1nxjj=m+1nxj0 as n>m and n,m,

j=1xj exists by completeness. Conversely, suppose xj is a Cauchy sequence. Thus, for each j=1,2,..., there exists an index kj such that xnxm<2j for any n,mkj. Let xk0=0. Then j=0xkj+1xkj<2. Hence, by assumption we can get the limit x=j=0xkj+1xkj, and since

xnkx=j=1nxkj+1xkjx0 as n,

we conclude that xj has a subsequence converging to x; thus, it converges to x.

3 Theorem (Parseval) Let uk be a orthonormal sequence in a Hilbert space . Then the following are equivalent:

  • (i) span{u1,u2,...} is dense in .
  • (ii) For each x, x=k=1x,ukuk.
  • (iii) For each x,y, x,y=k=1x,uky,uk.
  • (iv) x2=k=1|x,uk|2 (the Parseval equality).

Proof: Let =span{u1,u2,...}. If v, then it has the form: v=k=1αkuk for some scalars αk. Since v,uj=k=1ajuk,uj=aj we can also write: v=k=1v,ukuk. Let y=k=1x,ukuk. Bessel's inequality and that is complete ensure that y exists. Since

y,v=k=1x,ukuk,v=k=1x,v,ukuk=x,k=1v,ukuk=x,v

for all v, we have xy={0}, proving (i) (ii). Now (ii) (iii) follows since

|x,yk=1nx,uky,uk|=|x,yk=1ny,ukuk|0 as n

To get (iii) (iv), take x=y. To show (iv) (i), suppose that (i) is false. Then there exists a z(span{u1,u2,...}) with z0. Then

k=1|z,uk|2=0<z2.

Thus, (iv) is false.

3 Theorem (adjoint) Let 𝒳,𝒴 be Hilbert spaces. Let T be a linear operator from a subspace of 𝒳 to 𝒴. For each y𝒴, Tx,y𝒴 is continuous for all x in the domain of T if and only if there exists some z𝒳 such that:

Tx,y𝒴=x,z𝒳 for all x in the domain of T.

Here z is uniquely determined by y if and only if the domain of T is dense.
Proof: The existence follows from the Hahn-Banach theorem (or the projection argument for Hilbert spaces) and the Riesz representation theorem. (FIXME: need more detail). Suppose the uniqueness of z. It then follows: if xdomT and w(domT), then:

Tx,y=x,z=x,z+w

and the uniqueness shows that z=z+w and w=0. Conversely, if domT is dense, then

Tx,y=x,z=x,w for xdomT implies that zw(domT)=0.

When the domain of T is dense, for each y satisfying the continuity condition, we denote z in the theorem by T*y. T* can be shown to be linear (since T is densely defined but not because of linearlity), and we call it the adjoint of T.

3 Theorem Let 𝒳,𝒴 be Hilbert spaces and T:𝒳𝒴 be a closed densely defined (i.e., having dense domain and closed graph) operator. Then:

  • (i) T* is a closed densely defined.
  • (ii) T**=T.

Proof: Let z(domT*). If uy is orthogonal to the graph of T, then

0=uy,xTx=u,x+y,Tx for all x in the domain of T;

that is, y(domT*). Thus, :

0=0,u+z,y for uy(graT).

This is to say: 0z(graT)=graT and z=T(0)=0. (ii) (i) implies the existence of T** and the equality holds since the following are equivalent:

  • (a) wzgraT**
  • (b) 0=T*y,w+y,z=(T*y)y,wz for uygra(T).
  • (c) 0=uy,wz for uygra(T).
  • (d) wzgraT.

Note that the equivalence of (b) and (c) follows, as in the proof of (i), from that y is in the domain of T* and so T*y=u. We conclude that T=T**, and (by applying (ii)) that T* has closed graph.

3 Corollary Let 1,2 be Hilbert spaces, and T:H1H2/math>acloseddenselydefinedlinearoperator.<math>udom if and only if there is some K>0 such that:

T*f,uKf for every fdom*

Proof: If Tu is defined, then take K=Tu2. As for the converse, the estimate means that the map fT*f,u is continuous; hence, udom**=dom.

3 Theorem Let 1,2,3 be Hilbert spaces, and Tj:HjHj+1(j=1,2) densely defined linear operators. Suppose ranT1domT2. Then graT1*T2*gra(T2T1)* where the equality holds if Tj**=Tj(j=1,2)
Proof: Let zdomT1*T2* be given. Then

(T2T1)u,z=T1u,T2*z==u,T1*T2*z for every udomT1. But, by definition, (T1T2)* denotes an element such as T1*T2*z.

Hence, (T1T2)*=T1*T2* in domT1*T2*. Finally, if Tj**=Tj(j=1,2), then, by the fact we have just proved

graT1*T2*gra(T2T1)*=gra(T2**T1**)*graT1*T2*.

3 Theorem Let be a closed linear subspace of a Hilbert space . P is an orthogonal projection onto if and only if P=P*=P2 and the range of P is .
Proof: The direct part is clear except for P=P*. To see this, note that for any x Px,x is a real number since Px,x=Px,xPx+Px,Px. Hence,

Px,x=x,Px=P*x,x for every x.

By Lemma 3.something P=P*. For the converse we only have to verify xPx. But it holds since kerP=kerP*=(ranP) and for any x P(xPx)=PxP2x=0.

As an application we shall prove:

3 Theorem Let 1,2 be Hilbert spaces, T:12 be a closed densely defined linear operator. Then T is surjective if and only if there is a K>0 such that

f2KT*f1 for every fdomT*.

Proof: Suppose T is surjective. It suffices to show that

E={f;fdomT*,T*f11}

is bounded. If g2, then we can write g=Tu and so:

|f,g|2=|T*f,u|1u for every fE.

That E is bounded now follows from Theorem 2.something. To show the converse, let g2 be given. Since T* is injective, we can define a linear functional L by L(T*f)=f,g2 for f2. For every fdomT*,

|L(T*f)|=|f,g2|KT*f.

Thus, L is continuous on the range of T*. It follows from the Hahn-Banach theorem that we may assume that L is defined and continuous on 1. Thus, by Theorem 3.something, we can write L()=,u1 in 1. Then:

L(T*f)=f,g2=T*f,u1=f,Tu2 for every f2,

and so Tu=g.

3 Corollary Let T,1,2 be as given in the preceding theorem. Then ranT is closed if and only if ranT* is closed.
Proof: Suppose ranT is closed. Since T* restricted to (kerT*) has the same range as T*, we only have to show this operator has the closed range. Thus, suppose T*fj is convergent and fj(kerT*)=ranT. Applying the preceding theorem with =ranT, we get:

fjfk2KT*(fjfk)10 as j,k.

Thus, there is a limit f of fj and since T* is a closed operator, T*fjT*f. Hence, ranT* is closed. The converse holds since T**=T.

3 Lemma Let S,T be linear operators from a Hilbert space over to itself. If Tx,x=Sx,x for x, then S=T.
Proof: Let R=TS. We have 0=R(x+y),x+y=Rx,y+Ry,x and 0=iR(x+iy),x+iy=Rx,y+i2Ry,x. Summing the two we get: 0=2Rx,y for x,y. Taking y=Rx gives 0=Rx|2 for all x or R=0.

3 Lemma Let T be a continuous linear operator on a Hilbert space . Then TT*=T*T if and only if Tx=T*x for all x.

Continuous linear operators with the above equivalent conditions are said to be normal. For example, an orthogonal projection is normal. See w:normal operator for additional examples and the proof of the above lemma.

3 Theorem Let N be a normal operator. If α and β are distinct eigenvalues of N, then the respective eigenspaces of α and β are orthogonal to each other.
Proof: Let I be the identity operator, and x,y be arbitrary eigenvectors for α,β, respectively. Since the adjoint of αI is α¯I, we have:

0=(NαI)x=(NαI)*x=N*xα¯x.

That is, N*x=α¯x, and we thus have:

α¯x,y=N*x,y=x,Ny=β¯x,y

If x,y is nonzero, we must have α=β.

A sequence xn in a Hilbert space is said to converge weakly to x if xnx,y0

3 Theorem Let xk be an orthogonal sequence in a Hilbert space . The series k=1xk converges if and only if the series k=1xk converges weakly. Proof: (TODO: write one using the uniform boundedness principle and completeness.)

3 Exercise Let be a Hilbert space with orthogonal basis e1,e2,..., and xn be a sequence with xnK. Prove that there is a subsequence of xn that converges weakly to some x and that xK. (Hint: Since xn,ek is bounded, by Cantor's diagonal argument, we can find a sequence xnk such that xnk,ek is convergent for every k.)

We will obtain a more general version of this in the next chapter.

4 Geometry of Banach spaces

The major topic of the chapter is the notion of reflexibility of Banach spaces, which, as it turns out, is a geometric property.

If 𝒳 is a Banach space, we can introduce a topology called the "weak topology" in addition to the original norm one, as follows.

From the definitions follows immediately:

4. Corollary

  • (i) xn converges weakly (i.e., in the terms of a weak topology) to x in 𝒳 if and only if f(xn) converges to f(x) for every f𝒳*.
  • (ii) fn converges weak-* to f in 𝒳* if and only if fn(x) converges to f(x) for every x𝒳.

In particular, when 𝒳 is a Hilbert space, in view of Theorem 3.something the above discussion remains valid if we replace each occurrence of f(x) by x,y for some y.

We remark this easy-to-remember characterization does not sufficiently tell the whole story.

4. Corollary The continuous dual in terms of a weak topology is the same as the continuous dual in terms of a norm topology.
Proof: Let τ1 be the weak topology and tau2 be the norm topology. By definition τ1τ2 and so: (E,τ2)*(E,τ1)*. Conversely, if f(E,τ1)*, then f is norm-continuous; thus, f(E,τ2)*.

4. Theorem Let 𝒳 be a Banach space. A linear functional z is continuous on 𝒳* if and only if it is continuous on the closed unit ball of 𝒳*

4 Theorem Every finite dimensional Banach space is reflexive (thus, complete).
Proof: (TODO)

A linear operator T is said to be a compact operator if the image of the open unit ball under T is relatively compact. We recall that if a linear operator between normed spaces maps bounded sets to bounded sets, then it is continuous. Thus, every compact operator is continuous.

4. Theorem Let 𝒳 be a reflexive Banach space and 𝒴 be a Banach space. Then a linear operator T:𝒳𝒴 is a compact operator if and only if T sends weakly convergent sequence to norm convergent ones.
Proof: [1] Let xn converges weakly to 0, and suppose Txn is not convergent. That is, there is an ϵ>0 such that Txnϵ for infinitely many n. Denote this subsequence by yn. By hypothesis we can then show (TODO: do this indeed) that it contains a subsequence ynk such that Tynk converges in norm, which is a contradiction. To show the converse, let E be a bounded set. Then since 𝒳 is reflexive every countable subset of E contains a sequence xn that is Cauchy in the weak topology and so by the hypothesis Txn is a Cauchy sequence in norm. Thus, T(E) is contained in a compact subset of 𝒴.

4 Lemma Let r>0. A normed space 𝒳 is finite-dimensional if and only if its closed ball of radius r is compact.
Proof: If 𝒳 is not finite dimensional, using w:Riesz's_lemma, we can construct a sequence xj such that:

1=xjxjk=1j1akxk for any sequence of scalars ak.

Thus, in particular, xjxk1 for all j,k. (TODO: fill gaps)

4 Corollary

  • (i) Every finite-rank linear operator T (i.e., a linear operator with finite-dimensional range) is a compact operator.
  • (ii) Every linear operator T with the finite-dimensional domain is continuous.

Proof: (i) is clear, and (ii) follows from (i) since the range of a linear operator has dimension less than that of the domain.

4 Theorem The set of all compact operators into a Banach space forms a closed subspace of the set of all continuous linear operators in operator norm.
Proof: Let T be a linear operator and ω be the open unit ball in the domain of T. If T is compact, then T(ω) is bounded (try scalar multiplication); thus, T is continuous. Since the sum of two compacts sets is again compact, the sum of two compact operators is again compact. For the similar reason, αT is compact for any scalar α. We conclude that the set of all compact operators, which we denote by E, forms a subspace of continuous linear operators. To show the closedness, suppose S is in the closure of E. Let ϵ>0 be given. Then there is some compact operator T such that ST<ϵ/2. Also, since T is a compact operator, we can cover T(ω) by a finite number of open balls of radius ϵ/2 centered at z1,z2,...zn, respectively. It then follows: for xω, we can find some j so that Txzj<ϵ/2 and so SxTxSxzj+zjTx<ϵ. This is to say, S(ω) is totally bounded and since the completeness its closure is compact.

4. Corollary If Tn is a sequence of compact operators which converges in operator norm, then its limit is a compact operator.

4 Theorem (transpose) Let 𝒳,𝒴 be Banach spaces, and u:𝒳𝒴 be a continuous linear operator. Define tu:𝒴*𝒳* by the identity tu(z)(x)=u(z(x)). Then tu is continuous both in operator norm and the weak-* topology, and tu=u.
Proof: For any z𝒴*

tu(z)=supx1|(uz)(x)|uz

Thus, tuu and tu is continuous in operator norm. To show the opposite inequality, let ϵ>0 be given. Then there is x0𝒳 with (1ϵ)u|u(x0)|. Using the Hahn-Banach theorem we can also find z0=1 and z0(u(x0))=|u(x0)|. Hence,

tu=supz1tu(z)tu(z0)=|z0(u(x))|=|u(x0)|(1ϵ)u.

We conclude tu=u. To show weak-* continuity let V be a neighborhood of 0 in 𝒳*; that is, V={z;z𝒳*,|z(x1)|<ϵ,...,|z(xn)|<ϵ} for some ϵ>0,x1,...,xn𝒳. If we let yj=u(xj), then

tu({z;z𝒴*,|z(y1)|<ϵ,...,|z(yn)|<ϵ})V

since z(yj)=tu(z)(xj). This is to say, tu is weak-* continuous.

4 Exercise State and prove the analog of Theorem 3.something using the transpose instead of a Hilbert adjoint operator.

5 Lp spaces

Template:Stage

5 Theorem (Minkowski's inequality) If f,gLp(dμ) and p>1, then:

f+gpfp+gp.

Proof (See Chapter 2 of An Introduction to Analysis for more details): The theorem is a consequence of Hölder's inequality:

|fgdμ|(|f|pdμ)1/p(|f|qdμ)1/q (where we let q by 1/p+1/q=1)

which is then a simple consequence of the following inequality:

a1pb1q=inft>0(1pt1qa+1qt1pb) (where a, b > 0)

To simplify notation we denote |f|,|g| by f and g, respectively. First we have:

(f+g)pdμ=f(f+g)p1+g(f+g)p1dμ

Then Hölder's inequality followed by division gives:

(f+g)pdμ((f+g)(p1)qdμ)1/q(fpdμ)1/p+(gpdμ)1/p

where 1p+1q=1. The desired inequality follows after noting (p1)q=p and 11q=1p.

By letting μ be a counting measure we also obtain the analog for the series:

5 Corollary (k=1|fk+gk|p)1/p(k=1|fk|p)1/p+(k=1|gk|p)1/p

5 Theorem Let 1p, 1q, and LpLq a linear subspace. If is closed in both Lp and Lq, then the topology for inherited from Lp and Lq coincide.
Proof: The identity map I:(E,Lq)(E,LpLq) is continuous. Since I is a bijection between Banach spaces, the open mapping theorem says I is a homeomorphism.

(Remark: the theorem is just an exercise 2.something.)

5 Theorem Lp spaces are uniformly convex (thus, complete and reflexive).
Proof: (TODO)

5 Theorem (lifting property of l1) Let 1,2 be Banach spaces, and Q:12 be a linear surjection. If T:l12 is a a continuous linear operator, there exists a continuous linear operator T~:l11 such that QT~=T.
Proof: Let en be a canonical basis for l1. Using Corollary 2.something to the open mapping theorem we obtain a sequence xn and constant K so that:

Qxn=Ten, and xnKQxn.

Then supn1xnKsupn1QxnKTsupn1en<. Define T~ by T~en=xn. Then QT~=T, and T~ is continuous since supn1T~en<.

6 Banach algebras

Part 2: Topological vector spaces

Topological vector space

A vector space endowed by a topology that makes translations (i.e., x+y) and dilations (i.e., αx) continuous is called a topological vector space or TVS for short.

A subset E of a TVS is said to be:

  • bounded if for every neighborhood V of 0 there exist s>0 such that EtV for every t>s
  • balanced if λEE for every scalar λ with |λ|1
  • convex if λ1x+λ2yE for any x,yE and any λ1,λ20 with λ1x+λ2y=1.

1 Corollary (s+t)E=sE+tE for any s,t>0 if and only if E is convex.
Proof: Supposing s+t=1 we obtain sx+tyE for all x,yE. Conversely, if E is convex,

ss+tx+ts+tyE, or sx+sy(s+t)E for any x,yE.

Since (s+t)EsE+tE holds in general, the proof is complete.

Define f(λ,x)=λx for scalars λ, vectors x. If E is a balanced set, for any |λ|1, by continuity,

f(λ,E)f(λ,E)E.

Hence, the closure of a balanced set is again balanced. In the similar manner, if E is convex, for s,t>0

f((s+t),E)=f((s+t),E)=sE+tE,

meaning the closure of a convex set is again convex. Here the first equality holds since f(λ,) is injective if λ0. Moreover, the interior of E, denoted by E, is also convex. Indeed, for λ1,λ20 with λ1+λ2=1

λ1E+λ2EE,

and since the left-hand side is open it is contained in E. Finally, a subspace of a TVS is a subset that is simultaneously a linear subspace and a topological subspace. Let be a subspace of a TVS. Then is a topological subspace, and it is stable under scalar multiplication, as shown by the argument similar to the above. Let g(x,y)=x+y. If is a subspace of a TVS, by continuity and linearity,

g(,)g(,)=.

Hence, is a linear subspace. We conclude that the closure of a subspace is a subspace.

Let V be a neighborhood of 0. By continuity there exists a δ>0 and a neighborhood W of 0 such that:

f({λ;|λ|<δ},W)V

It follows that the set {λ;|λ|<δ}W is a union of open sets, contained in V and is balanced. In other words, every TVS admits a local base consisting of balanced sets.

1 Theorem Let 𝒳 be a TVS, and E𝒳. The following are equivalent.

  • (i) E is bounded.
  • (ii) Every countable subset of E is bounded.
  • (iii) for every balanced neighborhood V of 0 there exists a t>0 such that EtV.

Proof: That (i) implies (ii) is clear. If (iii) is false, there exists a balanced neighborhood V such that E⊄nV for every n=1,2,.... That is, there is a unbounded sequence x1,x2,... in E. Finally, to show that (iii) implies (i), let U be a neighborhood of 0, and V be a balanced open set with 0VU. Choose t so that EtV, using the hypothesis. Then for any s>t, we have:

EtV=stsVsVsU

1 Corollary Every Cauchy sequence and every compact set in a TVS are bounded.
Proof: If the set is not bounded, it contains a sequence that is not Cauchy and does not have a convergent subsequence.

1 Lemma Let f be a linear operator between TVSs. If f(V) is bounded for some neighborhood V of 0, then f is continuous.

1 Theorem Let f be a linear functional on a TVS 𝒳.

  • (i) f has either closed or dense kernel.
  • (ii) f is continuous if and only if kerf is closed.

Proof: To show (i), suppose the kernel of f is not closed. That means: there is a y which is in the closure of kerf but f(y)0. For any x𝒳, xf(x)f(y)y is in the kernel of f. This is to say, every element of 𝒳 is a linear combination of y and some other element in ker. Thus, kerf is dense. (ii) If f is continuous, kerf=f1({0}) is closed. Conversely, suppose kerf is closed. Since f is continuous when f is identically zero, suppose there is a point y with f(y)=1. Then there is a balanced neighborhood V of 0 such that y+V(kerf)c. It then follows that supV|f|<1. Indeed, suppose |f(x)|1. Then

yxf(x)ker(f)(y+V) if xV, which is a contradiction.

The continuity of f now follows from the lemma.

1 Lemma Let V0,V1,... be a sequence of subsets of a a linear space containing 0 such that Vn+1+Vn+1Vn for every n0. If xVn1+...+Vnk and 2n1+...+2nk2m, then xVm.
Proof: We shall prove the lemma by induction over k. The basic case k=1 holds since VnVn+Vn for every n. Thus, assume that the lemma has been proven until k1. First, suppose n1,...,nk are not all distinct. By permutation, we may then assume that n1=n2. It then follows:

xVn1+Vn2+...+VnkVn21+...+Vnk and 2n1+...2nk=2(n21)+...+2nk2m.

The inductive hypothesis now gives: xVm. Next, suppose n1,...,nk are all distinct. Again by permutation, we may assume that n1<n2<...nk. Since no carry-over occurs then and m<n1, m+1<n2 and so:

2n2+...+2nk2(m+1).

Hence, by inductive hypothesis, xVn1+Vm+1Vm.

1 Theorem Let 𝒳 be a TVS.

  • (i) If 𝒳 is Hausdorff and has a countable local base, 𝒳 is metrizable with the metric d such that
d(x,y)=d(x+z,y+z) and d(λx,0)d(x,0) for every |λ|1
  • (ii) For every neighborhood V𝒳 of 0, there is a continuous function g such that
g(0)=0, g=1 on Vc and g(x+y)g(x)+g(y) for any x,y.

Proof: To show (ii), let V0,V1,... be a sequence of neighborhoods of 0 satisfying the condition in the lemma and V=V0. Define g=1 on Vc and g(x)=inf{2n1+...+2nk;xVn1+...+Vnk} for every xV. To show the triangular inequality, we may assume that g(x) and g(y) are both <1, and thus suppose xVn1+...+Vnk and yVm1+...+Vmj. Then

x+yVn1+...+Vnk+Vm1+...+Vmj

Thus, g(x+y)2n1+...+2nk+2m1+...+2mj. Taking inf over all such n1,...,nk we obtain:

g(x+y)g(x)+2m1+...+2mj

and do the same for the rest we conclude g(x+y)g(x)+g(y). This proves (ii) since g is continuous at 0 and it is then continuous everywhere by the triangular inequality. Now, to show (i), choose a sequence of balanced sets V0,V1,... that is a local base, satisfies the condition in the lemma and is such that V0=𝒳. As above, define f(x)=inf{2n1+...+2nk;xVn1+...+Vnk} for each x𝒳. For the same reason as before, the triangular inequality holds. Clearly, f(0)=0. If f(x)2m, then there are n1,...,nk such that 2n1+...+2nk2m and xVn1+...+Vnk. Thus, xVm by the lemma. In particular, if f(x)2m for "every" m, then x=0 since 𝒳 is Hausdorff. Since Vn are balanced, if |λ|1,

λxVn1+...+Vnk for every n1,...,nk with xVn1+...+Vnk.

That means f(λx)f(x), and in particular f(x)=f((x))f(x)f(x). Defining d(x,y)=f(xy) will complete the proof of (i). In fact, the properties of f we have collected shows the function d is a metric with the desired properties. The lemma then shows that given any m, {x;f(x)<δ}Vm for some δ2m. That is, the sets {x;f(x)<δ} over δ>0 forms a local base for the original topology.

The second property of d in (i) implies that open ball about the origin in terms of this d is balanced, and when 𝒳 has a countable local base consisting of convex sets it can be strengthened to:d(λx,y)λd(x,y), which implies open balls about the origin are convex. Indeed, if x,yVn1+...+Vnk, and if λ10 and λ20 with λ1+λ2, then

λ1x+λ2yVn1+...+Vnk

since the sum of convex sets is again convex. This is to say,

f(λ1x+λ2y)min{f(x),f(y)}f(x)+f(y)2

and by iteration and continuity it can be shown that f(λx)λf(x) for every |λ|1.

Corollary For every neighborhood V of some point x, there is a neighborhood of x with WV
Proof: Since we may assume that x=0, take W={x;g(x)<21}.

Corollary If every finite set of a TVS 𝒳 is closed, 𝒳 is Hausdorff.
Proof: Let x,y be given. By the preceding corollary we find an open set VV{y}c containing x.

Locally convex spaces

A TVS with a local base consisting of convex sets is said to be locally convex. Since in this book we will never study non-Hausdorff locally convex spaces, we shall assume tacitly that every finite subset of every locally convex is closed, hence Hausdorff in view of Theorem something.

Lemma Let 𝒳 be locally convex. The convex hull of a bounded set is bounded.

Distributions

The support of u, denoted by supp(u), is the complement of the union of all the open sets ω such that u(w)=0 for every w𝒟(Ω).

Theorem Let u𝒟'(Ω). Then u(w)=0 for every w𝒟((supp(u))c)

Theorem A distribution u𝒟'(Ω) has compact support if and only if it can be extended to an continuous linear functional on 𝒞(Ω).

Theorem If uL2(n), then

u^L2=uL2

Proof: When u𝒟(n), the theorem is a special case of Theorem something. Using Theorem something we can find a sequence uj𝒟(n) such that uju in L2-norm. It follows:

uj^uk^L2=ujukL20.

Hence, there is a limit vL2(n) such that uj^v in L2-norm and v=u^.

Template:Subject Template:Alphabetical Template:Shelf Template:DDC Template:LOC

  1. This proof and a few more related results appear in [1]