• Keine Ergebnisse gefunden

Invariant Subspace Problem

N/A
N/A
Protected

Academic year: 2022

Aktie "Invariant Subspace Problem"

Copied!
28
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Invariant Subspace Problem

Bachelor Thesis

Johannes Kaiser, Matrikelnummer 1225752

Technische Universit¨ at Wien

February 3, 2016

(2)

Contents

1 Introduction 2

2 Classical Results about the Existence of Invariant Subspaces 2 2.1 Preliminary Notes . . . 2 2.2 The Aronszajn-Smith and Bernstein-Robinson Theorems . . . 3 2.3 Lomonosov’s Theorem . . . 9

3 Similarity and Quasisimilarity 11

3.1 Hyperinvariant Subspaces . . . 14 3.2 Contractions Quasisimilar to Unitary Operators . . . 15

4 Negative Conclusions 19

4.1 Polynomially Boundedness . . . 19 4.2 Other Counterexamples . . . 26

(3)

1 Introduction

In this paper we want to present a few results related to the invariant subspace problem. That is, the question whether an operator on a certain space, usually a Banach or Hilbert space, has a nontrivial invariant subspace. This problem seems to have been stated by Beurling and von Neumann and there are quite a lot of different theorems about it. Some of them give answers under conditions on the Banach or Hilbert space, others give their results under assumptions on the operator.

We divide the paper in two parts. In sections 2 and 3 some of the most popular positive answers to our question are shown and in section 4 we discuss one counterexample in detail, and present a few others in addition.

The chapter about similarity and quasisimilarity closely follows Chapter 4 of [1], while the chapter about polynomial boundedness follows Chapter 10 of [2]. The proof of Lomonosov’s theorem is extracted from Chapter 4.4 of [3]. The Bernstein-Robinson and the Aronszajn-Smith theorem are taken from [5] and [4].

2 Classical Results about the Existence of Invariant Subspaces

In the first section we want to present a few important theorems, which give the existence of an invariant subspace under certain conditions. To start this section we present a few results, which should already be known, or have quite simple proofs. We continue with the classic Aronszajn-Smith theorem given in [4], and the Bernstein-Robinson theorem, extracted from [5], whose proves share some common features. Then we introduce hyperinvariant subspaces and Lomonosov’s theorem, whose proof is different to the proofs of the first two.

2.1 Preliminary Notes

Before we start with the first theorem we recall and clarify the use of a few definitions

Definition 2.1. GenerallyH and K denote Hilbert spaces, while Bdenote a Banach space.

B(H) denote all bounded operators of H into itself. Further:

(i) We call an invariant subspaceSofH under an operatorT, a proper or nontrivial invariant subspace, if {0} 6=S6=H.

(ii) A closed linear subspaceM of H is said to be reducing forT ifM andM are invariant subspaces for T.

(iii) An operator with kTk ≤1 is called a contraction.

Now we want to present a few conclusions already known from linear algebra and funda- mental functional analysis, and one with a quite simple proof.

Theorem 2.2. On finite dimensional vector spaces with dimension greater than one every nonzero operator has an eigenvector and hence a nontrivial invariant subspace.

Theorem 2.3. On a Hilbert spaceH that is not separable every operator has a proper invariant subspace.

Proof. It is easily seen that the sequence {Tn :n∈N}spans a nontrivial subspace of H, and is invariant.

We obtain another conclusion through the spectral theorem for normal operators:

(4)

Theorem 2.4. Due to the spectral theorem every normal operator has an invariant subspace.

Theorem 2.5. Let T and L be nonzero operators on H. If LT = 0, then ker(L) and ran(T) are nontrivial invariant subspaces for T and L.

Proof. If LT = 0, then ran(T) ⊆ ker(L). Hence T(ker(L)) ⊆ T(H) = ran(T) ⊆ ker(L).

Since T 6= 0, we get that ker(L) 6= 0 and on the other hand we have that L 6= 0 so that ker(L)6=H. Therefore ker(L) is a nontrivial invariant subspace forT. Dually sinceTL = 0, L 6= 0 and T6= 0, it follows that ker(T) is a nontrivial invariant subspace for L, and hence ran(T) = ker(T) is a nontrivial invariant subspace for L. Finally recall that ker(L) and ran(T) are trivially invariant subspaces forL and T respectively.

Corollary 2.6. Every nilpotent operator has a nontrivial invariant subspace.

We want to finish this section with another theorem, which gives us an example of a Banach space on which every continuous operator has a proper nontrivial invariant subspace. The proof can be found in [6], but it is quite long so it is omitted.

Theorem 2.7. (Argyros,Haydon) There is an indecomposable Banach space with its dual space being isomorphic to`1. Every bounded linear operator on this space is expressible asλI+K with λa scalar and K compact. In particular every continuous operator has an invariant subspace.

2.2 The Aronszajn-Smith and Bernstein-Robinson Theorems

We now want to present two of the most classical results in invariant subspace theory. We start with the older Aronszajn-Smith Theorem and the proof we present follows the original one, presented in [4].

Theorem 2.8 (Aronszajn, Smith). Let T be a compact operator in a Banach Space B. Then there exist proper invariant subspaces of T.

In the proof of this theorem we use a map P on an arbitrary finite dimensional subspace C⊆B, defined through

kx−P xk=ρ(x, C) = min

y∈Ckx−yk.

We are able to limit ourselves to a separable Banach space, because of Theorem 2.3, therefore we can define an equivalent strictly convex norm on our Banach space, according to [7], Theorem 9. Because of that, we shall suppose that our norm on B is strictly convex. Therefore, and because of the separability, there exists a unique point P x ∈ C which realizes the minimal distance.

We refer to the upper map P as ”metric projection”, because it is quite similar to a pro- jection, but not necessarily linear. Before we begin with the proof, we want to sum up a few general properties, which apply to the metric projection:

Lemma 2.9. IfP is a metric projection on a finite dimensional subspaceC, following properties are fulfilled:

1. P is idempotent: P2 =P

2. P is homogeneous: P(αx) =αP x

3. P is quasiadditve: P(y+x) =y+P x for everyy∈C 4. kP x−xk ≤ kxk, kP xk ≤2kxk

(5)

5. |kx−P xk − ky−P yk| ≤ kx−yk

6. If C0 ⊆C andP0 is the metric projection on C0, then kx−P xk ≤ kx−P0xk.

Proof. The properties 1., 4. and 6. are obvious from the definition and 5. is the general property of the shortest distance from x to a fixed set C. To prove 2. consider P(x) = y and observe that kαx−αP(x)k= miny∈Ckαx−αyk=kαx−P(αx)k.

Property3. follows withk(x+y)−P(x+y)k= minz∈Ckx+y−zk=k(x+y)−(y+P(x))k, because y∈C.

Now we can begin with a few constructions for the proof of the Aronszajn-Smith-Theorem:

In finite dimensional spaces our theorem holds true, because of Theorem 2.2. Because of Theorem 2.3 we can restrict ourselves to infinite dimensional separable spaces and we have

(A) B is separable.

Now we are only interested in the case that

span{Tnf :n∈N}=B, f ∈B. (1)

because otherwise we would have already found our invariant subspace in the span of{Tn:n∈ N}. This formula implies the following property:

(B) Tnf 6= 0 and all elements {Tnf :n∈N} are linearly independent.

To prove (B)suppose thatα1Tn1f+α2Tn2f+...+αkTnkf = 0, with 0≤n1 < n2 < ... < nk and αi 6= 0,i∈ {1, ..., k}. We obtain that

Tnkf =− 1

αk

1Tn1f+...+αk−1Tnk−1f+),

and hence that all Tnf would lie in a subspace generated by Tnf with n < nk which is a contradiction to (1) and the infinite dimension ofB.

Now we consider a sequence of closed subspaces Ck∈B. The limes inferior of the sequence is defined as

lim inf

k Ck :={x∈B:∃xk ∈Ck, xk→x}.

The following two properties are easily verified:

(C) lim infkCk is a closed subspace.

(D) If everyCkis finite dimensional, then x∈lim infkCk if and only ifPkx→x, wherePk

denotes the projection on Ck.

With f satisfying (1) we construct the k-dimensional subspace C(k)= [Tnf]k−10 .

We denote by Pk the metric projection onto C(k). By (1) we obtain that lim infCk = B and hence that

Pkx→x, x∈B. (2)

(6)

Now we consider the operatorTk on C(k) defined by Tkx=PkT x, x∈C(k).

Now we prove thatTk is linear. Through homogenity and quasiadditivity of the metric projec- tion and by setting x:=Pk−1

i=0 ξiTif, we get that

Tkx=PkT x=Pk k−1

X

i=0

ξiTi+1f =

k−2

X

i=0

ξiTi+1f +ξk−1PkTkf, and hence that Tk is linear.

Because Tk is a linear operator on ak-dimensional space we can find a triangular matrix to represent Tk. Therefore there exists an increasing sequence of subspaces

0 =Ck0⊂Ck1 ⊂...⊂Ckk=Ck, (3) whereCki is ani-dimensional invariant subspace ofTk.

Lemma 2.10. Let {km}and{im} be sequences, such thatkm→ ∞ and0≤im ≤km,∀m∈N.

Further let xm∈Ckim

m. If T xm →y theny∈lim infCkim

m. Proof. In fact we have PkmT xm = Tkmxm ∈ Ckim

m. On the other hand, by Theorem 2.9,5, we have

kT xm−P(km)T xmk − ky−P(km)T xmk

≤ kT xm−yk. With this and (2) we obtain kT xm−P(km)T xmk ≤ ky−P(km)yk+kT xm−yk →0

ky−P(km)T xmk ≤ ky−T xmk+kT xm−P(km)T xmk →0, which proves the lemma, becauseTkmxm →y and thereforey∈lim infkmCkim

m.

We continue the preparation for the proof of the Aronszajn-Smith Theorem with the follow- ing corollaries:

Corollary 2.11. For any sequences {km} and {im} satisfying the conditions of Lemma 2.10, lim infkCkim

m is an invariant subspace of T.

Proof. If we have x ∈lim infkCkim

m we have by the definition of the lim inf the existence of xm such that xm ∈ Ckim

m, xm → x. By the continuity of T we obtain T xm → T x and by Lemma 2.10 T x∈lim infkCkim

m.

Corollary 2.12. If thelim inf of every subsequence ofCkim

m ={0}then for any bounded sequence xm ∈Ckimm we have T xm→ {0}.

Proof. By compactness of T, the sequencexm is transformed into a relatively compact subse- quenceT xm. Therefore it is enough to prove that if any subsequenceT xmj converges to somey, theny= 0. But this follows from our hypothesis, since by Lemma 2.10 y∈lim infkmCkim

m. Proof. (Aronszajn-Smith) Now we choose an arbitrary α >0 with

0< α <1,kT fk> αkTkkfk. (4) Since f ∈Ck we have by Theorem 2.9-3 and 2.9-6

kfk=kf −Pk,0fk ≥ kf −Pk,1fk ≥...≥ kf −Pk,kfk={0}.

(7)

Therefore there exists a unique indexi(k),0≤i(k)< k for eachk∈N such that

kf−Pki(k)fk ≥αkfk>kf −Pki(k)+1fk. (5) Letuk,k= 1,2, ...be an element ofCki(k)+1 such that

kukk= 1, Pki(k)u(k) = 0. (6) An element with the given property can be obtained from an arbitrary v ∈ Cki(k)+1\Cki(k) by setting uk := kv−Pki(k)vk−1(v−Pki(k)v). The property (6) is now proved by homogenity and quasiadditivity of P.

Since the dimensions of Cki(k)+1 and Cki(k) differ by one, every element y∈Cki(k)+1 is repre- sentable in a unique way in the form y =x+βuk with x =Pki(k)y. Correspondingly we shall put

Pki(k)+1f =xkkuk, Pki(k)+1T f =x0kk0uk, xk, x0k∈Cki(k). (7) We have by Theorem 2.9-4,

kxkk=kPki(k)Pki(k)+1fk ≤4kfk, .kx0kk ≤4kT fk. (8) Now we prove the following statements:

(E) For every sequencekm→ ∞,lim infkmCki(km)

m 6=B. (F) For some sequence km0 → ∞,lim infkmCki(k0 0m)+1

m 6= 0.

(G) If for every sequence km → ∞,lim infkmCki(km)

m = {0} then for every sequence km0

∞,lim infkmCki(k0 m0 )+1 m 6=B. If lim infkmCki(km)

m = B, then by (B) Pki(km)

m f → f which contradicts (5), hence (E) holds true.

If(F)were not true we would have by Corollary 2.12 that the bounded subsequencePki(k)+1f (see Theorem 2.9-4) is transformed into a sequence T Pki(k)+1f converging to 0. Since T f = T(f−Pki(k)+1f) +T Pki(k)+1f we getkT fk= limkT(f−Pki(k)+1f)k ≤lim infkkTkkf−Pki(k)+1fk which by (5) giveskT fk ≤4kTkkfk in contradiction to (4).

Suppose that for some km0 → ∞, lim infk0

mCki(k0 0m)+1

m =B. By (B) we havePki(k0 0m)+1 m f → f and Pki(k0 0m)+1

m T f → T f. By (7) we have f = limkm(xk0mk0mukm0 ) and T f = limk0m(x0k0 m + βk00

muk0m). Further we obtainT f = limk0m(T xkm0k0mT uk0m) andT2f = limk0m(T x0k0 mk00

mT uk0m).

By (8) and Corollary 2.12 it follows T f = limk0mβk0mT uk0m and T2f = limkm0 βk00

mT ukm0 . There- foreβk00

mk0

m converges to someγ andT2f =γT f in contradiction to(B), hence(G)is proved.

Now we obtain the proof of our theorem as follows. If there is any sequence km → ∞ such that Λ = lim infkmCki(kmm)6={0} then in view of(E)and Corollary 2.11, Λ is a proper invariant subspace. If there is no such sequence, then by (F) we choose a sequencek0m → ∞ such that Λ0 = lim infkmCki(k0 0m)+1

m 6={0}. By(G) and Corollary 2.11 Λ0 is a proper invariant subspace.

The Bernstein-Robinson theorem is an extension to the Aronszajn-Smith theorem, originally proved by using nonstandard analysis, in [5]. However we present the proof given by Halmos in [8], which still has a lot of features in common with Aronszajn-Smith’s proof, but does not use nonstandard analysis.

(8)

Theorem 2.13 (Bernstein-Robinson). If A is an operator on a Hilbert space H of dimension greater than one and if p is a nonzero polynomial such that p(A) is compact, there exists a nontrivial subspace of H underA.

Before we can begin with the proof we need a short definition:

Definition 2.14. Iffn and gn are sequences onH we shall write fn∼gn forkfn−gnk →0.

Proof. According to Theorem 2.2, we can assume the existence of a nonzero vectoresuch that e,Ae,A2e,... are linearly independent and have H as their closed linear span. Otherwise the closed linear span would already be an invariant subspace.

Through Gram-Schmidt orthogonalization we can obtain an orthonormal basis {e1, e2, ...}

with the property, that{e1, e2, ..., em}has the same linear span as{e, Ae, ..., Am−1e}, form∈N.

Ifam,n := (Aen, em), it follows that am,n = 0, ifm > n+ 1. The matrix entries of thekth power ofAare given bya(k)m,n = (Aken, em) Through induction one can see thata(k)m,n = 0, ifm > n+k and

a(k)n+k,n = Y

1≤j≤k

an+j,n+j−1.

Let k ≥ 1 be the degree of our given polynomial p. If the matrix entries of p are given by a(p)m,n = (p(A)en, em), then a(p)n+k,n is a constant multiple ofa(k)n+k,n. This is verified because the coefficient a(l)n+k,n = 0 if l < k. Since kp(A)enk → 0, for n → ∞, which we have because of the compactness ofp(A), there exists an increasing sequence{k(n)}n∈Nof postive integers such that the corresponding subdiagonal termsak(n)+1,k(n) converge to 0 forn→ ∞.

IfHnis the span of{e1, ..., ek(n)}, then{Hn}n∈Nis an increasing sequence of finite-dimensional subspaces of H withH as their span. IfPn is the projection with range Hn, thenPns I (I being the identity operator). Since, for each operator, the identityPnAPn leaves Hn invariant, it follows that for eachn there exists a chain of subspaces invariant underPnAPn

{0}=Hn(0)⊂Hn(1)⊂...⊂Hn(k(n))=Hn.

with dimHn(i)=i,i= 0,1, ..., k(n), a construction similar to the proof of the Aronszajn-Smith Theorem.

Iffn∈H is a bounded sequence of vectors, we want to prove that

APnfn∼PnAPnfn. (9)

For the proof of (9), we have that Pnf =Pk(n)

j=1(f, ej)ej, iff ∈H and that APnfn−PnAPnfn=

k(n)

X

j=1

(fn, ej)

X

i=k(n)+1

aijei.

Since the largestj isk(n) and the smallestiisk(n) + 1 and since aij = 0 ifi > j+ 1, it follows that kAPnfn−PnAPnfnk ≤ kfnkak(n)+1,k(n) → 0, and therefore we have proved (9). (9) can be generalized to higher exponents:

AkPnfn∼(PnAPn)kfn k= 1,2, ... (10) which can again be proved by induction. For k= 0, (10) says thatkPnfn−fnk →0, which is a stringent condition on the bounded sequence fn. If that is satisfied, then (10) implies that

p(A)Pnfn∼p(PnAPn)fn. (11)

(9)

Now we return to our vector e. Since Pne = e for every n, it follows that p(A)Pne ∼ p(PnAPn)e. Sincep(A)e6= 0, which follows because the vectors (e, Ae, ...) are linearly indepen- dent, we have

=kp(PnAPn)ek=kp(A)ek>0.

Consider for eachnthe numbers

kp(PnAPn)e−p(PnAPn)Pn(0)ek, kp(PnAPn)e−p(PnAPn)Pn(1)ek,

...

kp(PnAPn)e−p(PnAPn)Pn(k(n))ek.

wherePn(i) is the projection with range Hn(i). SincePn(0) is the zero projection the first of these numbers tends to . Since, on the other hand Pn(k(n))=Pn, the last of these numbers is always 0. In view of these facts it is possible to choose for eachnwith a finite number of exceptions a positive integer i(n), 1≤i(n)≤k(n), such that

kp(PnAPn)e−p(PnAPn)Pn(i(n)−1)ek ≥

2 (12)

and

kp(PnAPn)e−p(PnAPn)Pn(i(n))ek<

2. (13)

Further leti(n) be the smallest integer for which these inequalities hold true.

Since bothPni(n)−1 andPni(n) are bounded sequences of operators, there exists an increasing sequence nj of integers such that both Pni(nj)−1 and Pni(nj) are weakly convergent. To simplify notation we set Qj := Pni(nj)−1 and Q+j := Pni(nj). Let M := {f ∈ H | Qj f →s f}, and M+:={f ∈H|Q+jf →s f}.

Now we are going to prove that M and M+ are subspaces of H that are both invariant underA, and that at least one of them is nontrivial. To prove that M is closed, suppose that g is in the closure ofM. We have to show thatg∈M and therefore that Qjg →g. Given a positive number δ, one has to find f ∈ M so that kf −gk < δ3 and then find j0 so that kQjf−fk< δ3 forj≥j0. It follows that, if j≥j0, then kQj g−gk ≤ kQjg−Qj fk+kQj f− fk+kf −gk< δ. This proves thatM is closed, the proof forM+ is the same.

To prove that M is invariant under A, we suppose that f ∈ M, so that Qj f → f and infer, first, that AQj f → Af, because A is bounded and second, that Qj AQj f ∼ QjAf, because Qj is uniformly bounded. Then we reason as follows:

Qj Af ∼QjAQj f =(a)QjPnjAPnjQj f =(b) PnjAPnjQj f ∼(c)APnjQjf =AQj f →Af.

(a) is valid because Qj ≤ Pnj, (b) because the range of Qj is invariant under PnjAPnj and (c) because of (10). This proves that M is invariant under A, the prove that M+ is invariant follows the same scheme.

The next step is to prove that M 6=H. This is done by showing thate /∈M. For this purpose observe first that the operators p(PnAPn) are uniformly bounded as one can see, by observing that

k(PnAPn)kk ≤ kPnAPnkk≤ kAkk.

(10)

and by using the polynomial whose coefficients are the absolut values of the coefficients of p.

Now, because of (12) we have

2 ≤ kPnjAPnjkke−Qj ek.

SincekPnjAPnjk is bounded from above, its reciprocal is bounded from zero, and consequently ke−Qjek is bounded away from zero, which makes the convergenceQj e→eimpossible.

The corresponding step for M+ says that M+ 6= {0}, but with a quite different proof.

The choice of the sequence {nj} implies that the sequence {Q+je} is weakly convergent. The compactness of p(A) implies therefore that the sequence{p(A)Q+je} is strongly convergent to, say,f. The proof that follows, we have to show two parts:

1. f 6= 0 2. f ∈M+

To prove 1. we have p(A)Q+j e ∼p(PnjAPnj)Q+j e by (11), which is within 2 of p(PnjAPnj)e, by (13), whose norm tends to . It follows thatk{p(A)Q+je}k can not tend to zero, and hence thatf 6= 0.

To prove 2. we have thatQ+j f ∼Q+jp(A)Q+j e, sinceQ+j is uniformly bounded. Then we have Q+j p(A)Q+je∼Q+jp(PnjAPnj)Q+j eby (11) and uniform boundedness. Because the range ofQ+j is invariant underp(PnjAPnj), we haveQ+jp(PnjAPnj)Q+j e=p(PnjAPnj)Q+j e∼p(A)Q+j ewith the last relation holding true because of (11). At last p(A)Q+j e→f by definition.

If M+ 6= H all is well. It remains to be proved that if M+ = H then M 6= {0}. If M+ = H then Q+j f → f for all f, at least weakly. At the same time the sequence {Qj} is known to be weakly convergent to, say, Q. The operators Qj and Q+j are projections such thatQj ≤Q+j and such thatQ+j −Qj has rank 1. It follows that for eachjthere exists a unit vectorfj, such that (Q+j −Qj )f = (f, fj)fj for all f.

Observe now that Qj e cannot weakly tend to e, for if it did, it would tend strongly to e, which is a property of projections, but was proved to not be the case. This implies that Qe6=e, or, equivalently that (1−Q)e6= 0. Now the numbers |(e, fj)|can not be arbitrarily small. This follows because, since|((Q+j −Qj)e, g)| ≤ |(e, fj)|kgkfor allg, an affirmative answer would imply that ((1−Q)e, g) = 0 for all g, so that (1−Q)e= 0 - a contradiction. So we have obtained that the numbers |(e, fj)|are bounded away from zero, which makes it possible to prove that M6={0}.

It turns out that ifg⊥(1−Q)e, theng∈M. Indeed since (e, fj)(fj, g)→((1−Q)e, g) = 0, it follows that (fj, g) → 0 and hence that (f, fj)(fj, g) → 0 for all f. This implies that ((1−Q)f, g) = 0 for all f, and hence that (1−Qg) = 0. In other words Qj g→ g weakly, and therefore strongly. From this it follows thatg∈M.

2.3 Lomonosov’s Theorem

Lomonosov’s Theorem is another classical result in invariant subspace theory, but it has a com- pletely different proof to the Aronszajn-Smith Theorem and the Bernstein-Robinson Theorem.

The proof we present follows the one in [3], Chapter 6.

Actually Lomonosov’s Theorem is split in two parts, a lemma and the actual theorem.

Before we get to these two we need to prove another theorem and provide a few definitions:

(11)

Definition 2.15. Let B(H,K) be the space of all bounded operators from H to K, with H and K being Hilbert spaces.

(i) With K(H,K) we denote all compact operators from H to K, and with K(H) all compact operators fromH into itself.

(ii) Let Lat(T) denote the set of all invariant subspaces forT. (iii) If A ⊆ B(H) then let Lat(A) :=T

{Lat(T) :T ∈ A}.

Definition 2.16. If T is a linear operator, we shall call a subspace C ⊆H hyperinvariant if S(C)⊆C for allS commuting withT.

Definition 2.17. We denote the convex hull of a set S with co(S) and have co(S) := {xt+ (1−t)y:x, y∈S, t∈[0,1]}.

Theorem 2.18. (Mazur’s Theorem) IfH is a Banach space andK is a compact subset ofH, thenco(K) is compact.

Proof. Obviously, it suffices to show that co(K) is totally bounded. Let > 0 and choose x1, ..., xn in K such that K ⊆ Sn

j=1B(xj, /3). Put C = co{x1, ..., xn}, then C is obviously compact. Hence there are vectorsy1, ..., yn inC such thatC ⊆Sn

j=1B(yj, /3). Ifw∈co(K), there is a z in co(K) with kw−zk < /3. Thus z =Pl

p=1αpkp, where kp ∈ K, αk ≥ 0 and Pαk = 1. Now for eachkp there is anxj(k) with kkp−xj(p)k< /3. Therefore

z−

l

X

p=1

αpxj(p)

=

l

X

p=1

αk(kp−xj(p))

l

X

p=1

αkkkp−xj(p)k< /3.

But Pl

p=1αpxj(p) ∈ C, so there is an yi with kPl

p=1αpxj(p)−yik < /3. With the triangle inequality we get kw−yjk ≤ kw−zk+kz−Pl

p=1αpxj(p)k+kPl

p=1αpxj(p)−yik< , which shows us that co(K)⊆Sn

j=1B(yj, ) and so co(K) is totally bounded.

Lemma 2.19. If A is a subalgebra toB(H), such that I ∈ Aand LatA={∅,H} and ifK is a nonzero compact operator onH, then there is an A∈ Asuch that ker(AK−I)6= 0.

Proof. It may be assumed that kKk= 1. Fix x0 ∈H such that kKx0k>1 and putS ={x∈ H :kx−x0k ≤1}. If we have 0∈ S, we get 1< kK(x0)k ≤ kKkkx0−0k ≤ 1 and hence a contradiction.

On the other hand, if 0∈(K(S)) we have that there exists a sequence yn∈S,n∈Nsuch that limnK(yn) = 0 and hence there exists an N ∈N such that 1<kKx0k − kK(yN)k. Now we have 1<kKx0k − kK(yN)k ≤ kKkkx0−yNk ≤1 and therefore another contradiction.

In conclusion we have

0∈/ S and 0∈/K(S). (14)

Now ifx∈H andx6= 0,{T x:T ∈ A}is an invariant subspace forA, becauseAis an algebra, and it contains the nonzero vectorx, becauseI ∈ A. By hypothesis we get{T x:T ∈ A}=H. By (14) this says that for every y∈K(S), there is aT inA withkT y−x0k<1. Equivalently

K(S)⊆ [

T∈A

{y:kT y−x0k<1}.

(12)

Because K(S) is compact, there areT1, ..., Tn inAsuch that K(S)⊆

n

[

j=1

{y:kTjy−x0k<1}. (15)

Fory∈K(S) and 1≤j≤n, letaj(y) = max{0,1− kTjy−x0k}. By (15),Pn

j=1aj(y)>0, for all y∈K(S). Define bj :K(S)→R, by

bj(y) = aj(y) Pn

i=1ai(y), and define Ψ :S →H by

Ψ(x) =

n

X

j=1

bj(Kx)TjKx.

It is easy to see that aj : K(S) → [0,1] is a continuous function and hence bj and Ψ are continuous too.

If x∈S, thenKx∈K(S). Ifbj(Kx)>0, thenaj(Kx)>0 and sokTjKx−x0k<1. That isTjKx∈S, whenever bj(Kx)>0. Since S is a convex set andPn

j=1bj(Kx) = 1 for x∈S, Ψ(S)⊆S.

Note that TjK ∈ K(H) for eachj so that S

j≤nTjK(S) has compact closure. By Mazur’s Theorem 2.18 co(S

TjK(S)), is compact. But this convex set contains Ψ(S) so that Ψ(S) is compact. This is, Ψ is a linear map. By the Schauder Fixed-Point Theorem, there is a vector x1 ∈ S such that Ψ(x1) = x1. Let βj = bj(Kx1) and put A = Pn

j=1βiTj. So A ∈ A and AKx1= Ψ(x1) =x1. Sincex16= 0, becausex1 is in S, we obtain ker(AK−I)6= 0.

Theorem 2.20. (Lomonosov’s Theorem) IfH is a Banach space over C,T ∈ B(H),T is not a multiple of the identity and T K =KT, for some nonzero compact operator K, then T has a nontrivial hyperinvariant subspace.

Proof. Let A = {T}0. We want to show that LatA 6= {0,H}. If this is not the case then Lomonosov’s Lemma implies that there is an operatorAinAsuch thatN = ker(AK−I)6= 0.

ButN ∈Lat(AK) andAK|N is the identity operator. SinceAK ∈ K(H), we get dimN <∞.

Since AK ∈ A = {T}0 for any x ∈ N. AK(T x) = T A(Kx) = T x, hence TN ⊆ N. But dimN <∞, so thatT|N must have an eigenvalueλ. Thus ker(T−λ) =M 6= 0. ButM 6=H, sinceT is not a multiple of the identity. It is easy to check that M is hyperinvariant forT:

x∈ker(T−λ)⇒0 = (T −λ)x=K(T−λ)x= (T−λ)Kx⇒Kx∈ker(T−λ)

3 Similarity and Quasisimilarity

In this section we want to take a look from a different angle at the invariant subspace problem. So far we have proved theorems which give us the existence of an invariant subspace under certain conditions. Now we want to take a look at how far the existence of an invariant subspace for one operator, carries over to another operator with a certain relation to the first one. These relations are going to be similarity and quasisimilarity. We are going to take a closer look at hyperinvariant subspaces, which we have already introduced at Lomonosov’s theorem. This section closely follows Chapter 4 of [1]. We can now start with a few definitions:

(13)

Definition 3.1. (i) T ∈ B[H,K] is called quasiinvertible, if it is an injective operator with dense range.

(ii) An operator T ∈ B[H,K] is called quasiaffine transform of L∈ B[H,K] if there exists a quasiinvertible X∈ B[K,H] such that XT =LX.

(iii) Two operatorsT ∈ B[H] andL∈ B[K] are called quasisimilar, if there exist two operators X∈ B[H,K] and Y ∈ B[K,H] such thatXT =LX and Y L=T Y.

(iv) An operatorX ∈ B[H,K] intertwines T ∈ B[H] toL∈ B[K] ifXT =LX.

Lemma 3.2. Let T ∈ B[H], L ∈ B[K] and X ∈ B[H,K] such that XT = LX. Suppose C(K is a nontrivial invariant subspace forL. If

ran(X) =K and ran(X)∩C6={0}, thenX−1(C) is a nontrivial invariant subspace forT.

Proof. Let C ( K be an nontrivial invariant subspace for L. Since X : H → K, is linear and continuous, X−1(C) is a subspace of H. Moreover, sinceX(X−1(C))⊆C it follows that LXX−1(C) ⊆L(C). Hence, since L(C) ⊆ C, we have LXX−1(C) ⊆C. Because LX = XT, we get XT X−1(C) ⊆ C and so X−1XT X−1(C) ⊆ X−1(C). On the other hand, we get T X−1(C)⊆X−1XT X−1(C), becauseA⊆X−1(X(A)) for all setsA⊆H. Therefore we have

T X−1(C)⊆X−1(C),

in other wordsX−1(C) is an invariant subspace forT. Now we shall verify that the assumptions on ran(X) are enough to ensure that the invariant subspace X−1(C) is nontrivial. Take an arbitrary y ∈ ran(X)∩C so that y = Xu ∈ C, with u ∈ H. If X−1(C) = {0}, then u = 0 because

X−1(C) ={0} ⇔ {x∈H :Xx∈C}={0}.

Hence we obtain thaty= 0, becauseXis linear. We conclude that, ifX−1(C) ={0}, it follows that ran(X)∩C={0}. Equivalently, we have

ran(X)∩C6={0} ⇒X−1(C)6={0}.

If X−1(C) = H, then ran(X) = X(H) = XX−1(C). Thus since X(X−1(C)) ⊆ C we get ran(X) ⊆ C = C 6= K. We conclude that X−1(C) = H ⇒ ran(X) 6= K, or equivalently ran(X) =K ⇒X−1(C)6=H

If the intertwining operator X is surjective, then X−1(C) is a nontrivial invariant subspace forT, wheneverC is a nontrivial invariant subspace for L. An even more particular case reads as follows:

Corollary 3.3. If two operators are similar and one of them has an invariant subspace then so has the other.

Proposition 3.4. Let K be a Hilbert space, let M be a finite-dimensional subspace of K and let R be a linear manifold on K. If R=K, then

(R∩M) =M.

(14)

Proof. The result holds trivially if K is a finite-dimensional Hilbert space, for in such a case R =K and thereforeR∩M =M. It also holds trivially if dim(M) = 0, because then we haveM ={0}and thereforeM=K, which gives usR∩M =R=K =M. So from now on we can assume that K is infinite-dimensional and that m := dim(M) ≥1. First we shall verify that the above result holds true form= 1. Ifm= 1 thenM = span(e) for{0} 6=e∈K. Since R is dense inK there existsx∈K, such that

(e, x)6= 0,

because otherwise, if (e, x) = 0 for all x∈R, we would have that e∈R, or e={0}, because R is dense in K. Now take an arbitrary z ∈M. Since R = K and z ∈ K, there exists a sequence {zj ∈R:j ≥1} such thatzj →zas j→ ∞.

For each j≥1 set

yj =zj−(e, zj) (e, x)x,

and note thatyj ∈R for everyj≥1, because zj, x∈R and R is a linear manifold. Further we have yj ∈ M for every j ≥1, because (e, yj) = (e, zj)− (e,z(e,x)j)(e, x) = 0, so that yj ⊥ e and hence yj ∈M and yj → z, j → ∞, since (e, z) = 0, because z ∈ M ={e}. Therefore for everyz∈M, there exists an (R∩M)-valued sequence converging to z. Hence (R∩M) is dense in M and we conclude that the result holds for m= 1.

Now suppose it holds for somem≥1, that is, suppose R∩M=M,

for any m-dimensional subspace M of K. Take an arbitrary (m+ 1)-dimensional subspace of K, sayN. Let{el: 0≤l≤m}, be an orthonormal basis for N, so that

N =

m

M

l=0

[el] Take an arbitrary integerk∈ {0, ..., m}. Set

Mk =

m

M

l=0,l6=k

[el],

so that R∩Mk=Mk, once dimMk=m. Note that there exists xk∈R∩Mk such that (ek, xk)6= 0,

because, if (ek, xk) = 0 for allxk ∈ R∩Mk, then ek ∈ (R∩Mk) = (R∩Mk) =Mk⊥⊥ = Mk = Mk, which contradicts the fact that 0 6= ek ⊥ Mk. Take an arbitrary z ∈ N = (Lm

l=0[ek]). Since R = K and z ∈ K, there exists a sequence {zj ∈ R :j ≥ 1} such that zj →zas j→ ∞.

For each j≥1 set

yj =zj

m

X

k=0

(ek, zj) (ek, xk)xk

and note that yj ∈ R for every j ≥ 1, because zj, xk ∈ R and R is a linear manifold in K. Further yj ∈ N for every j ≥ 1, since xk ∈Mk = (Lm

l=0,l6=k[el]) ⊆ [en] for every n 6=k, n∈ {0, ..., m}, it follows that (el, xk) = 0, for every l6=k,l∈ {0, ..., m}. Hence

(el, yj) = el, zj

m

X

k=0

(ek, zj) (ek, xk)xk

!

=−

m

X

k=0,k6=l

(ek, zj)

(ek, xk)(el, xk) = 0

(15)

for everyj≥1 and everyl∈ {, ..., m}. Therefore yj ∈Tm

l=0[el]= (Lm

l=0[ek])=N for every j ≥ 1. Moreover yj → z as j → ∞, since (ek, z) = 0, because z ∈N = Tm

l=0[el]. Thus for every z ∈ N there exists an (R∩N)-valued sequence converging to z. Hence (R∩N) is dense in N and we conclude that the result holds form+ 1, whenever it holds for m, which ends the proof by induction.

This proposition gives us the following two corollaries:

Corollary 3.5. TakeT ∈ B[H], L∈ B[K] andX ∈ B[H,K]such that XT =LX.

Let M ⊂K be a nontrivial finite-dimensional reducing subspace for L. If ran(X) = K, then X−1(M) is a nontrivial invariant subspace forT.

Proof. This is an immediate conclusion from Proposition 3.4 and Lemma 3.2.

Corollary 3.6. If an operator T is a quasiaffine transform of another operator L, that has a nontrivial finite-dimensional reducing subspace, then T has a nontrivial invariant subspace.

3.1 Hyperinvariant Subspaces

Definition 3.7. The commutant {T}0 of T ∈ B[H] is the set of all operators in B[H] that commute with T, or equally

{T}0 :={U ∈ B[H] :U T =T U}.

Further letTx:={y∈H :y=U x for someU ∈ {T}0}.

Proposition 3.8. For each x∈H, Tx is a subspace of H which is hyperinvariant for T. Proof. Take any x ∈ H and consider the set Tx ⊆ H. If y1, y2 ∈ Tx, then there exist U1, U2∈ {T}0, such thaty1 =U1x andy2 =U2x. Therefore we havey1+y2 = (U1+U2)x∈Tx, because obviously U1, U2 ∈ {T}0 ⇒ U1+U2 ∈ {T}0. Moreover αy ∈ Tx for every α ∈ C and everyy ∈Tx, trivially. Therefore Tx is a linear manifold on H. Now takeU ∈ {T}0 arbitrary.

If y∈Tx, then y=U0x for some U0 ∈ {T}0, so thatU y =U U0x ∈Tx, forU U0 is obviously in {T}0. Thus U(Tx)⊆Tx, and henceU(Tx)⊆Tx, becauseU is continuous.

Because the closure of a manifold is a linear subspace, we conclude that Tx is an invariant subspace for every U ∈ {T}0 and equivalently Tx is an hyperinvariant subspace for T.

Lemma 3.9. Let T ∈ B[H],L∈ B[K], X∈ B[H,K]and Y ∈ B[K,H]be such that XT =LX and Y L=T Y.

Suppose C is a nontrivial hyperinvariant subspace of L. If

ran(X) =K and ker(Y)∩C ={0},

then Y(C) 6= {0} and for each nonzero x ∈ Y(C), Tx is a nontrivial hyperinvariant subspace for T.

(16)

Proof. According to Proposition 3.8 it is enough to verify, that under the above hypothesis, {0} 6= Tx 6= H, for every 0 6= x ∈ Y(C) 6= {0}. First note that XU Y ∈ {L}0 for every U ∈ {T}0. Indeed, if Y L=T Y,T U =U T and XT =LX, then

(XU Y)L=XU T Y =XT U Y =L(XU Y).

SinceCis hyperinvariant forL, it follows thatCis invariant forXU Y, wheneverU ∈ {T}0. Now take x∈Y(C) arbitrary so that x=Y ufor someu∈C⊂K. Ify∈Tx, then y=U x=U Y u for some U ∈ {T}0 and hence Xy = XU Y u. But u ∈ C and C is invariant for XU Y. Thus Xy∈C. ThereforeX(Tx)⊆C so that, since X is continuous, we have

X(Tx)⊆C =C.

IfTx=H, then ran(X) =X(H) =X(Tx)⊆C6=K. We conclude that ran(X) =K ⇒Tx 6=H ∀x∈Y(C)

Finally, if Y(C) = {0}, then obviously C ⊆ ker(Y) and hence C ∩ker(Y) = C 6= {0}.

Therefore

ker(y)∩C={0} ⇒Y(C)6={0}.

Hence Tx 6={0}, for every nonzero x∈Y(C), because Tx={0} if, and only if,x= 0.

In particular, if XT =LX andY L=T Y, with ran(X) =K andker(Y) ={0}, then there exists x ∈ Y(C), such that Tx is a nontrivial hyperinvariant subspace for T, whenever C is a nontrivial hyperinvariant subspace forL. An even more particular case reads:

Corollary 3.10. If two operators are quasisimilar, and one has a nontrivial hyperinvariant subspace then so has the other.

3.2 Contractions Quasisimilar to Unitary Operators

Throgh this section T shall be a contraction on a Hilbert space H. Now we classify these contractions, but first we need a short definition:

Definition 3.11. We call a contraction strongly stable if {Tn :n∈ N} converges strongly to the zero operator or in signs{Tn:n∈N}→s 0.

LetC0.be the class of all strongly stable contractions and let C.0 be the class of all contrac- tions, whose adjoint is strongly stable. LetC1. and C.1 be the classes of contractions such that Tnx 90 and T∗nx90, respectively for every nonzero x∈H. All combinations are possible and lead to classes C00, C10, C01 and C11. If T is a contraction, we have that the sequence {T∗nTn:n∈N}is a bounded monotone sequence of self-adjoint operators, so that it converges strongly. Therefore it has a strong limit A, while {TnT∗n:n∈N} has a strong limitA. Proposition 3.12. Let T∗nTn s→A. Then A has the following properties:

(i) A is nonnegative and kAk ≤1.

(ii) kTnk → kA12k asn→ ∞ for allx∈H. (iii) ker(A) ={x∈H :Tnx→0}.

(iv) T∗nATn=A for every n≥1.

(17)

Proof. Ais nonnegative because it is the strong limit of a nonnegative sequence. Ais a contrac- tion, because it is the strong limit of a sequence of contractions, indeedkAxk= limnkT∗nTnxk ≤ kxkfor all x∈H, and we have proven (i).

We obtain(ii) by observing

kTnxk2 = (T∗nTnx, x)→(Ax, x) =kA12xk2.

Further we have that (ii) implies (iii), because ker(A) = ker(A12), for every nonnegative operatorA.

Note that T∗k+nTk+n = T∗nT∗kTkTn, for every k, n ≥1, and also that T∗k+nTk+ns A and T∗nT∗kTkTn →T∗nATn as k→ ∞, for everyn≥1. Thus the identity in (iv) follows by uniqueness of the strong limit.

With Proposition 3.12, it follows that T ∈C0. if and only ifA= 0 andT ∈C1.if and only if ker(A) ={0}. Therefore

T ∈C00⇔A=A = 0,

T ∈C01⇔A= 0 and ker(A) ={0}, T ∈C10⇔A = 0 and ker(A) ={0},

T ∈C11⇔ker(A) = ker(A) ={0}.

These properties enable us to show a few theorems, about which contractions possess non- trivial invariant subspaces. To accomplish this we start with the following proposition:

Proposition 3.13. If a contraction is quasisimilar to a unitary operator, then it is of class C11.

Proof. If T ∈ B[H] is a contraction and U ∈ B[K] is a unitary operator such thatXT =U X and Y U =T Y, for a pair of quasiinvertible operatorsX ∈ B[H,K] andY ∈ B[K,H], then XTn=UnX andYT∗n=U∗nYfor everyn≥1. Therefore ifx∈H such that limnTnx= 0, then limnUnXx= 0. HenceXx= 0, so thatx= 0. That is, by Proposition 3.12, ker(A) ={0}.

If one repeats this argument by putting U in the position of U and Y in the position of X, one obtains, that ker(A) ={0}. Thus if a contraction T is quasismilar to a unitary operator, then ker(A) = ker(A) ={0}, or equivalently T ∈C11.

In the following theorem we want to show that also the conversion to that proposition holds.

Therefore we need to construct an isometryV associated to T such that the equality V A12 =A12T

holds. Recall that ker(A) = ker(A), ran(A) ⊆ran(A12) and ran(A) = ran(A12) = ker(A12) = ker(A).

Now we consider the decomposition H = ker(A)⊕ker(A) = ker(A)⊕ran(A) and define a map

V : ran(A)→ran(A),

as follows. Take an arbitraryy∈ran(A) so that there exists anxy ∈ran(A12) withy=A12xy. Note that xy is unique, for A is injective when acting on ran(A) and so is A12. Set V0y = A12T xy, which defines a transformationV0 : ran(A)→ ran(A12) = ran(A) that is clearly linear, for A12 and T are linear. Extend it to ran(A) to get the linear transformation V : ran(A) → ran(A).

Referenzen

ÄHNLICHE DOKUMENTE

In particular, by using rank metric codes in different variants, we are able to obtain CDCs that either give improved lower bounds for many parameters, including A 2 (12, 4; 4), A

Constant dimension codes (CDC) are treated in Section 3, where the currently implemented lower bounds, constructions, and upper bounds are described in Subsection 3.1 and

This family of designs was considered recently for network coding solution for a family of multicast networks called the gen- eralized combination networks.. The goal of this work is

• After T 7 , all original assignments y = e; with y 6∈ Vars e are assignments to dead variables and thus can always be eliminated :-). • By this, it can be proven that

• After T 7 , all original assignments y = e; with y 6∈ Vars e are assignments to dead variables and thus can always be eliminated :-). • By this, it can be proven that

More precisely, we consider an operator family (A(ρ)) ρ∈X of closed densely defined operators on a Banach space E, where X is a locally compact

the field of formal power series in two variables R ((X, Y )); second, function fields of curves over the power series field R ((t)) or, more generally, over a field with a

Abstract We obtain a new upper bound on the dimensions of anisotropic quadratic torsion forms over a field that is an extension of finite transcendence degree of a real