Trace class operators and Hilbert-Schmidt operators

Jordan Bell
April 18, 2016

1 Introduction

If X,Y are normed spaces, let (X,Y) be the set of all bounded linear maps XY. If T:XY is a linear map, I take it as known that T is bounded if and only if it is continuous if and only if EX being bounded implies that T(E)Y is bounded. I also take it as known that (X,Y) is a normed space with the operator norm, that if Y is a Banach space then (X,Y) is a Banach space, that if X is a Banach space then (X)=(X,X) is a Banach algebra, and that if H is a Hilbert space then (H) is a C*-algebra. An ideal I of a Banach algebra is an ideal of the algebra: to say that I is an ideal does not demand that I is a Banach subalgebra, i.e. does not demand that I is a closed subset of the Banach algebra. I is a *-ideal of a C*-algebra if I is an ideal of the algebra and if AI implies that A*I.

If X and Y are normed spaces, we take as known that the strong operator topology on (X,Y) is coarser than the norm topology on (X,Y), and thus if TnT in the operator norm, then TnT in the strong operator topology.

If X is a normed space, M is a dense subspace of X, Y is a Banach space and T:MY is a bounded linear operator, then there is a unique element of (X,Y) whose restriction to M is equal to T, and we also denote this by T.11 1 This is an instance of a result about topological vector spaces and Fréchet spaces. See Walter Rudin, Functional Analysis, second ed., p. 40, chapter 1, ex. 19.

If X is a normed space, define ,:X×X* by

x,λ=λ(x),xX,λX*.

This is called the dual pairing. If X and Y are normed spaces and T(X,Y), it can be proved that there is a unique T*(Y*,X*) such that

Tx,λ=x,T*λ,xX,λY*,

called the adjoint of T, and that the adjoint satisfies T=T*.22 2 Walter Rudin, Functional Analysis, second ed., p.  98, Theorem 4.10.

I give precise statements of any statement that I want to use without proof. If a fact is not straightforward to prove and is not easy to look up (perhaps because it does not have a standardized name), I give a citation to a statement that is the precise version I use. I am making a point to write out full proofs of some tedious but essential arguments about Hilbert-Schmidt operators and trace class operators.33 3 Many of the proofs that I give are vastly expanded from what is written in the references I used; I simply decided to write down every step that I did, and thus a reader should be able to read this note without needing having to work out calculations on paper or without realizing partway through that I was tacitly identifying things or that what I said is true only under conditions I left unstated because I thought them too obvious. Indeed there is no royal road through mathematics, but we do not need to break up the asphalt and destroy the signage to make travelling what roads there are a trial of one’s skill.

2 Finite rank operators

In this section, X and Y denote Banach spaces. We say that a linear map T:XY has finite rank if T(X) is a finite dimensional subspace of Y. A finite rank operator need not be bounded: If X is infinite dimensional, let be a Hamel basis for X and let {en:n} be a subset of this basis. Defining a linear map on a Hamel basis determines it on the vector space. Define T:X by Ten=nen and Te=0 if e is not in the countable subset. Then T is not bounded; but its range has dimension 1, so T has finite rank. We define 00(X,Y) to be the set of T(X,Y) that have finite rank. It is apparent that 00(X,Y) is a vector space.

If fX* and yY, we define yf:XY by

yf(x)=f(x)y.

yf:XY is linear, and

yf=supx1(yf)(x)=supx1f(x)y=ysupx1|f(x)|ysupx1fx,

so yffy, so yf is bounded. (yf)(H)span{y}, so yf has finite rank. Therefore yf00(X,Y).

The following theorem gives a representation for bounded finite rank operators.44 4 Y. A. Abramovich and C. D. Aliprantis, An Invitation to Operator Theory, p. 124, Lemma 4.2. In this reference, what we write as yf they write as fy.

Theorem 1.

If TB00(X,Y) and w1,,wk is a basis for T(X), then there are unique f1,,fkX* such that

T=j=1kwjfj.

For yY, define Fy:Y* by Fy(λ)=λ(y). One checks that Fy(Y*)*. Often one writes y=Fy, which is fine as long as we keep in mind whether we are using y as an element of Y or as an element of (Y*)*.

Theorem 2.

If there are w1,,wkY and f1,,fkX* such that

T=j=1kwjfj,

then

T*=j=1kfjwj.
Proof.

Let yY and fX*. If xX and gY*, then

(yf)x,g=f(x)y,g=f(x)y,g=f(x)g(y),

where ,:Y×Y* is the dual pairing. But

f(x)g(y)=x,g(y)f=x,Fy(g)f=x,(Fyf)(g).

where ,:X×X* is the dual pairing. We have Fyf(Y*,X*), and hence

(yf)*=Fyf=yf.

This shows that the adjoint of each term wjfj in T is fjwj, and the adjoint of a sum is the sum of the adjoints of the terms, completing the proof. ∎

The above two theorems together show that if T00(X,Y) then T*00(Y*,X*).

Theorem 3.

If X is a Banach space, then B00(X) is a two sided ideal in the Banach algebra B(X).

Proof.

We have stated already that 00(X,Y) is a vector space, and here Y=X. If A00(X) and T(X), then AT(X), and

A(T(X))A(X),

which is finite dimensional, so AT00(X). TA(X), and T(A(X)) is the image of a finite dimensional subspace under T, and so is itself finite dimensional. Hence TA00(X). ∎

3 Compact operators

We say that a metric space M is totally bounded if for every ϵ there are finitely many balls of radius ϵ whose union equals M. The Heine-Borel theorem states that a metric space is compact if and only if it is complete and totally bounded. If S is a subset of a complete metric space M and S¯ is compact, then by the Heine-Borel theorem it is totally bounded, and any subset of a totally bounded metric space is itself a totally bounded metric space, so S is totally bounded. On the other hand, if SM is totally bounded, then one proves that S¯ is also totally bounded. As S¯ is a closed subset of the complete metric space M, S¯ is a complete metric space, and hence by the Heine-Borel theorem it is compact. We say that a subset of a metric space is precompact if its closure is compact, and thus a subset of a complete metric space is precompact if and only if it is totally bounded.

Let X and Y be Banach spaces. If T:XY is a linear map and U is the open unit ball in X, we say that T is compact if the closure of T(U) in Y is compact. Therefore, to say that a linear map T:XY is compact is to say that T(U) is totally bounded.

It doesn’t take long to show that if T:XY is linear and compact then it is bounded, so there is no difference between stating that something is a bounded compact operator and stating that it is a compact operator. The following is often a convenient characterization of a compact operator.

Theorem 4.

A linear map T:XY is compact if and only if for every bounded sequence xnX there is a subsequence xa(n) such that Txa(n) converges in Y.

We denote the set of compact operators XY by 0(X,Y). It is apparent that 0(X,Y) is a vector space. For T(X,Y), it is a fact that T0(X,Y) if and only if T*0(Y*,X*).55 5 Walter Rudin, Functional Analysis, second ed., p. 105, Theorem 4.19.

The following theorem states if a sequence of compact operators converges to a bounded operator, then that operator is compact.66 6 Walter Rudin, Functional Analysis, second ed., p. 104, Theorem 4.18 (c). Since (X,Y) is a Banach space, this implies that 0(X,Y) is a Banach space with the operator norm.

Theorem 5.

If X and Y are Banach spaces, then B0(X,Y) is a closed subspace of the Banach space B(X,Y).

The following theorem shows that a bounded finite rank operator is a compact operator. Since a limit of compact operators is a compact operator, it follows from this that a limit of bounded finite rank operators is a compact operator.

Theorem 6.

If TB00(X,Y) then TB0(X,Y).

Proof.

Let U be the open unit ball in X. Since T is bounded and U is a bounded set in X, T(U) is a bounded set in Y. But T(X) is finite dimensional, so T(U) is a bounded set in a finite dimensional vector space and hence by the Heine-Borel theorem, the closure of T(U) in T(X) is a compact subset of T(X). T(X) is finite dimensional so is a closed subset of Y, and hence the closure of T(U) in T(X) is equal to the closure of T(U) in Y. Hence the closure of T(U) in Y is a compact subset of Y. (If E is a topological space and CDE, then the subspace topology C inherits from E is the same as the subspace topology it inherits from D, so if KT(X)Y then to say that K is compact in T(X) is equivalent to saying that K is compact in Y.) ∎

Theorem 7.

If X is a Banach space, then B0(X) is a two sided ideal in the Banach algebra B(X).

Proof.

We have stated already that 0(X,Y) is a closed subspace of (X,Y), and here Y=X. Let K0(X) and T(X). On the one hand, if xn is a bounded sequence in X, then TxnTxn, so Txn is a bounded sequence in X. Hence there is a subsequence Txa(n) such that K(Txa(n))=(KT)xa(n) converges, showing that KT is compact.

On the other hand, if xn is a bounded sequence in X, then there is a subsequence xa(n) such that Kxa(n) converges to some x. T is continuous, so T(Kxa(n)) converges to Tx, showing that TK is compact. ∎

If X is not separable, then the image of the identity map idX is not separable, so the image of a bounded linear operator need not be separable. However, the following theorem shows that the image of a compact operator is separable. Check that if a subset of a metric space is separable then its closure is separable. From this and Theorem 8 we get that the closure of the image of a compact operator is separable.

Theorem 8.

If KB0(X,Y), then K(X) with the subspace topology from Y is separable.

Proof.

Let Un={xX:x<n}. Then K(Un)¯ is compact. It is a fact that a compact metric space is separable, hence K(Un)¯ is separable. A subset of a separable metric space is itself separable with the subspace topology, so let Ln be a countable dense subset of K(Un). Let L=n=1Ln, which is countable. It is straightforward to verify that L is a dense subset of

K(X)=n=1K(Un).

Therefore, K(X) is separable. ∎

4 Hilbert spaces

We showed that if X is a Banach space then both 00(X) and 0(X) are two sided ideals in the Banach algebra (X). We also showed that if A00(X) then A*00(X*), and that if A0(X) then A*0(X*). If H is a Hilbert space then (H) is a C*-algebra (as the adjoint of A(H) is not just an element of (H*), but can be identified with an element of (H)) and what we have shown implies that 00(H) and 0(H) are two sided *-ideals in the C*-algebra (H).

If Sα,αI are subsets of a Hilbert space H, we denote by αISα the closure of the span of αISα. If S1,S2 are subsets of H, we write S1S2 if for every s1S1 and s2S2 we have s1,s2=0. If V is a closed subspace of H, then H=VV, and the orthogonal projection onto V is the map P:HH defined by P(v+w)=v for vV,wV. P(H)=V, so rather than first fixing a closed subspace and then talking about the orthogonal projection onto that subspace, one often speaks about an orthogonal projection, which is the orthogonal projection onto its image. It is straightforward to check that if an orthogonal projection is nonzero then it has norm 1, and that an orthogonal projection is a positive operator.

The following theorem is an explicit version of Theorem 1 for orthogonal projections.

Theorem 9.

Let {e1,,en} be orthonormal in H and let M=k=1n{ej}. If P is the orthogonal projection onto M, then

P=k=1nekek,Ph=k=1nh,ekek,hH.
Proof.

As M is finite dimensional it is closed, and hence H=MM. Let h=h1+h2, h1M,h2M; as P is the orthogonal projection onto M, we have Ph=h1.

Let Qh=k=1nh,ekek. We have to show that Qh=Ph. For each 1jn, using that ek,ej=δk,j we get

Qh,ej=k=1nh,ekek,ej=h,ej.

Hence, if 1jn then h-Qh,ej=0. As {ej} are an orthonormal basis for M, this implies that h-QhM, and so h-QhM. That is,

h1+h2-QhM,

and as h2M it follows that h1-QhM. But h1-QhM (h1M by definition, and Qh is a sum of elements in M), so h1-Qh=0. As Ph=h1, this means that Ph=Qh, completing the proof. ∎

We say that a Banach space X has the approximation property if for each A0(X) there is a sequence An00(X) such that AnA. A result of Per Enflo shows that there are Banach spaces that do not have the approximation property. However, in the following theorem we show that every Hilbert space has the approximation property.

Theorem 10.

If H is a Hilbert space and AB0(H), then there is a sequence AnB00(H) such that AnA.

Proof.

By Theorem 8, V=A(H)¯ is separable. V is a closed subspace of the Hilbert space H, so is itself a Hilbert space. If V has finite dimension then A is itself finite rank. Otherwise, let {en:n1} be an orthonormal basis for V, and let Pn be the orthogonal projection onto j=1n{ej}. Pn00(H), and define An=PnA00(H).

In any Hilbert space, if vα is an orthonormal basis then idH=αvαvα, where the series converges in the strong operator topology (see §5). Thus, for any hV we have

Ah=k=1Ah,ekek,

where the series converges in H. By Theorem 9,

Anh=PnAh=k=1nAh,ekek,

and therefore Anh-Ah0 as n. What we have shown is that AnA in the strong operator topology.

Let B be the closed unit ball in H. Because A is a compact operator, by the Heine-Borel theorem A(B) is totally bounded: If ϵ>0, then there is some m and h1,,hmH such that

A(B)j=1mBϵ(Ahj).

If hB, there is some 1jm with AhBϵ(Ahj), i.e. Ah-Ahj<ϵ. If n1, then, as Pn1,

Ah-Anh Ah-Ahj+Ahj-Anhj+Anhj-Anh
= Ah-Ahj+Ahj-Anhj+Pn(Ahj-Ah)
2Ah-Ahj+Ahj-Anhj
2ϵ+Ahj-Anhj.

As AnA in the strong operator topology, for each 1jm there is some N(j) such that if nN(j) then Ahj-Anhj<ϵ and so, if nN(j),

Ah-Anh<3ϵ.

Let N=max1jmN(j), whence for all hB, if nN then

Ah-Anh<3ϵ.

But

A-An=suph1(A-An)h,

so if nN then

A-An<3ϵ,

showing that AnA. ∎

5 Diagonalizable operators

If is an orthonormal set in a Hilbert space H, which we do not demand be separable, then is an orthonormal basis for H if and only if for every hH we have h=eh,ee.77 7 John B. Conway, A Course in Functional Analysis, second ed., p. 16, Theorem 4.13. In other words, if is an orthonormal set in H, then is an orthonormal basis for H if and only if

idH=eee,

where the series converges in the strong operator topology.

We say that a linear map A:HH is diagonalizable if there is an orthonormal basis of H each element of which is an eigenvector of A. If A is a bounded linear operator on H, A is diagonalizable if and only if there is an orthonormal basis ei,iI, of H and λi such that the series

iIλieiei

converges to A in the strong operator topology. One checks that the series iIλi¯eiei converges to A* in the strong operator topology.

If A(H) is diagonalizable with eigenvalues {λi:iI}, it is a fact that

A=supiI|λi|. (1)
Theorem 11.

Let H be a Hilbert space with orthonormal basis {ei:iI}, let λiC, and define a linear map A:span{ei:iI}span{ei:iI} by Aei=λiei. If supiI|λi|<, then A extends to a unique element of B(H).

Proof.

Let M=supiI|λi|<. If J is a finite subset of I and x=iJαiei, then, as the ei are orthonormal,

Ax2=iJαiAei2=iJαiλiei2=iJαiλiei2M2iJ|αi|2.

But

x2=iJαiei2=iJαiei2=iJ|αi|2.

So

AxMx.

It follows that A is a bounded operator on span{ei:iI}, which is dense in H. Then there is a unique element of (H) whose restriction to span{ei:iI} is equal to A, and we denote this element of (H) by A. ∎

In Theorem 8 we proved that the image of a compact operator is separable. Hence if a compact operator is diagonalizable then it has only countably many nonzero eigenvalues.

Theorem 12.

If H is a separable Hilbert space with orthonormal basis en,n1, and if A:HH is linear and Aen=λnen for all n, then AB0(H) if and only if λn0 as n.

Proof.

Suppose that A is compact. Then A* is compact, and as A is diagonalizable so is A*. The proof of Theorem 10 wasn’t generally useful enough to be worth putting into a lemma, but to understand the following it will be necessary to read that proof. Let λa(n) be the nonzero eigenvalues of A, and let μn=λa(n), fn=ea(n). Check that fn,n1 is an orthonormal basis for A*(H)¯, Pn be the projection onto j=1n{fj}. Then using the argument in the proof of Theorem 10 we get PnA*A*; this completes our trip back to that proof. Let An=A-APn. As PnA*-A*0 as n,

An = (A-APn)*
= A*-Pn*A*
= A*-PnA*
0.

If 1jn then Anfj=Afj-Afj=0, and if j>n then Anfj=Afj=μjfj. Hence An is a diagonalizable operator, and by (1) we get An=supj>n|μj|. Together with An0 as n, this means that lim supn|μn|=0. As μn are precisely the nonzero λn, we obtain from this that limn|λn|=0.

On the other hand, suppose that λn0 as n. Because λn0, the absolute values of the eigenvalues of A are bounded and hence A(H). Let Pn be the projection onto j=1n{ej}. If 1jn then APnej=Aej=λjej, and if j>n then APnej=0. Therefore APn00(H). For An=A-APn, we have An=supj>n|λj|, from which it follows that

limnAn=0.

Hence APnA, and as APn are bounded finite rank operators, A is a compact operator. ∎

If A(H) is diagonalizable, say Aei=λiei, iI, then A*ei=λi¯ei, iI, and for jI,

AA*ej=AiIλi¯ej,eiei=A(λj¯ej)=λj¯λjej=|λj|2ej.

Likewise we get A*Aej=|λj|2ej, so AA*=A*A, that is, a bounded diagonalizable operator is normal. The following theorem states an implication in the other direction.88 8 Gert K. Pedersen, Analysis Now, revised printing, p. 108, Theorem 3.3.8. This is an instance of the spectral theorem.

Theorem 13.

If H is a Hilbert space over C and T is a normal compact operator on H, then T is diagonalizable.

6 Polar decomposition

It is a fact that A(H) is self-adjoint if and only if Ax,x for all xH. A(H) is said to be positive if A is self-adjoint and Ax,x0 for all xH. It is a fact that if A(H) is positive then there is a unique positive element of (H), call it A1/2, such that (A1/2)2=A; namely, a bounded positive operator has a unique positive square root in (H).99 9 Gert K. Pedersen, Analysis Now, revised printing, p. 92, Proposition 3.2.11. It is straightforward to check that if A(H) then A*A is positive, and hence has a square root (A*A)1/2(H), which we denote by |A|. |A| satisfies Ax=|A|x for all xH.

An isometry from one Hilbert space to another is a linear map A:H1H2 such that if xH1 then Ax=x. If A:HH is linear and the restriction of A to (kerA) is an isometry, then we say that A is a partial isometry. One checks that a partial isometry is an element of (H), if it is not the zero map then it has norm 1, and that its image is closed. We call (kerA) the initial space of A, and A(H) the final space of A. We can prove that if A is a partial isometry then A* is a partial isometry whose initial space is the final space of A and whose final space is the initial space of A. For example, an orthogonal projection is a partial isometry whose initial space is the image of the orthogonal projection and whose final space is equal to its initial space.

It is a fact that if A(H) then there is a partial isometry U with kerU=kerA satisfying A=U|A|, and that if V is a partial isometry with kerV=kerA that satisfies A=V|A|, then V=U.1010 10 Gert K. Pedersen, Analysis Now, revised printing, p. 96, Theorem 3.2.17. A=U|A| is called the polar decomposition of A. The polar decomposition satisfies

U*U|A|=|A|,U*A=|A|,UU*A=A,A*=|A|U*,|A*|=U|A|U*. (2)

We will use these formulas repeatedly when we are working with trace class operators, and we have numbered the above equation to refer to it and also to draw the eye to it.

If I is an ideal of (H) and AI then |A|=U*AI. In particular, if A00(H) then |A|00(H) and if A0(H) then |A|0(H).

7 Hilbert-Schmidt operators

Let H be a Hilbert space, and {ei:iI} an orthonormal basis of H. We say that A(H) is a Hilbert-Schmidt operator if

iIAei2<.

The following theorem shows that if A(H) is a Hilbert-Schmidt operator using one basis it will also be one using any other basis, that if A is not a Hilbert-Schmidt operator using one basis it will not be one using any other basis, and that A is Hilbert-Schmidt if and only if its adjoint A* is Hilbert-Schmidt.

Theorem 14.

Let H be a Hilbert space. If {eα:αI} and {fα:αJ} are orthonormal bases for H and AB(H), then

αIAeα2=αJA*fα2=αJAfα2.
Proof.

For each βJ, using Parseval’s identity we have

Afβ=αJAfβ,fαfα,Afβ2=αJ|Afβ,fα|2.

Using this we get

βJAfβ2 = βJαJ|Afβ,fα|2
= αJβJ|Afβ,fα|2
= αJβJ|fβ,A*fα|2
= αJβJ|A*fα,fβ|2
= αJA*fα2.

For each βI, Parseval’s identity gives us

Aeβ2=αJ|Aeβ,fα|2,

and using this we obtain

βIAeβ2 = βIαJ|Aeβ,fα|2
= αJβI|Aeβ,fα|2
= αJβI|eβ,A*fα|2
= αJβI|A*fα,eβ|2
= αJA*fα2.

Let 2(H) be the set of Hilbert-Schmidt operators in (H). If {ei:iI} is an orthonormal basis for H and A(H), we define

A2=(iIAei2)1/2.

To say that A2(H) is to say that A(H) and A2<. Using the triangle inequality in 2(I), one checks that 2(H) is a vector space and that 2 is a norm on 2(H), which we call the Hilbert-Schmidt norm.

Because |A|x=Ax for all xH, A is Hilbert-Schmidt if and only if |A| is Hilbert-Schmidt, and A2=|A|2. From Theorem 14 we obtain that if A2(H) then A*2(H), and A2=A*2.

Theorem 15.

B2(H) is a two sided *-ideal in the C*-algebra B(H), and if AB2(H) and TB(H) then

AT2A2T,TA2TA2.
Proof.

Let be an orthonormal basis for H. If A2(H) and T(H), then

TA22=eTAe2e(TAe)2=T2A22<.

On the other hand, using the above and the fact that if A is Hilbert-Schmidt then A* is Hilbert-Schmidt,

AT22=(AT)*22=T*A*22=T*2A*22<.

The following theorem shows that the operator norm is dominated by the Hilbert-Schmidt norm, and therefore that the topology on the normed space 2(H) with the Hilbert-Schmidt norm is finer than the subspace topology it inherits from (H) (i.e. its topology as a normed space with the operator norm).

Theorem 16.

If AB2(H) then AA2.

Proof.

Let ϵ>0. We have

A=supv=1Av.

Take vH, v=1, with Av2+ϵ>A2. There is an orthonormal basis {eα:αI} for H that includes v; one proves this using Zorn’s lemma. Then

A2<ϵ+Av2ϵ+Av2+eαvAeα2=ϵ+αIAeα2=ϵ+A22.

As this is true for all ϵ>0, it follows that AA2. ∎

The following theorem states that every bounded finite rank operator is a bounded Hilbert-Schmidt operator, and that every bounded Hilbert-Schmidt operator is the limit in the Hilbert-Schmidt norm of a sequence of bounded finite rank operators.

Theorem 17.

00(H) is a dense subset of the normed space B2(H) with the Hilbert-Schmidt norm.

Proof.

If A00(H), then there is an orthonormal basis {ei:iI} for H and a finite subset J of I such that if iIJ then Aei=0. From this it follows that A2<, and thus A2(H), so 00(H) is a subset of 2(H).

Let {ei:iI} be an orthonormal basis for H, let A2(H), and let ϵ>0. As iIAei2<, there is some finite subset J of I such that

iIJAei2<ϵ.

Let P be the orthogonal projection onto span{ei:iJ} (which is finite dimensional and hence closed), and define B00(H) by B=AP. We have

A-B22=iIJAei2<ϵ,

showing that 00(H) is dense in 2(H). ∎

If A2(H), then by the above theorem there is a sequence of bounded finite rank operators An such that An-A20 as n. But by Theorem 16, An-AAn-A2, and by Theorem 6, a limit of bounded finite rank operators is a compact operator, so A is compact. Thus, a bounded Hilbert-Schmidt operator is a compact operator.

We are going to define an inner product on 2(H) and we will show that with this inner product 2(H) is a Hilbert space. However the cleanest way I know to do this is by defining the trace of an operator. Moreover, we care just as much about trace class operators as we do Hilbert-Schmidt operators.

8 Trace class operators

If {eα:αI} and {fα:αJ} are orthonormal bases for H and A(H), then using Theorem 14 we have

αI|A|eα,eα = αI|A|1/2eα,|A|1/2eα
= αI|A|1/2eα2
= αJ|A|1/2fα2
= αJ|A|1/2fα,|A|1/2fα
= αJ|A|fα,fα.

If {ei:iI} is an orthonormal basis for H, we say that A(H) is trace class if

iI|A|ei,ei<.

We denote the set of trace class operators in (H) by 1(H). For A(H), define

A1=iI|A|ei,ei.

To say that A1(H) is to say that A(H) and that A1<. As |A|1/2 is self-adjoint, it is apparent that A1=|A|1/222. We will prove that 1(H) is a vector space and that 1 is a norm on this vector space, but this takes a surprising amount of work and we will not do this yet.

The following theorem gives different characterizations of bounded trace class operators.1111 11 John B. Conway, A Course in Operator Theory, p. 88, Proposition 18.8. This theorem shows in particular that if A1(H) then A is the product of two bounded Hilbert-Schmidt operators, and thus, as 2(H) is an ideal in (H), that A2(H). In particular, as a consequence Theorem 16, every bounded Hilbert-Schmidt operator is compact, so every bounded trace class operator is compact.

Theorem 18.

If AB(H), then the following are equivalent.

  • A1(H).

  • |A|1/22(H).

  • A is the product of two elements of 2(H).

  • |A| is the product of two elements of 2(H).

Proof.

Let ei,iI be an orthonormal basis for H. Suppose that A1(H). We have

|A|1/222=iI|A|1/2ei2=iI|A|1/2ei,|A|1/2ei=iI|A|ei,ei=A1,

so |A|1/22(H).

Suppose that |A|1/22(H). A=U|A|=(U|A|1/2)|A|1/2. As 2(H) is an ideal, we get U|A|1/22(H), hence A is the product of two elements of 2(H).

Suppose that A=BC, with B,C2(H). Let A=U|A| be the polar decomposition of A. By (2), the polar decomposition satisfies U*U|A|=|A|. But U|A|=BC implies that U*U|A|=U*BC, hence |A|=U*BC=(U*B)C. As 2(H) is an ideal, we have U*B2(H), so we have written |A| as a product of two elements of 2(H).

Suppose that |A|=BC, with B,C2(H); so B*2(H) too. We have, using the Cauchy-Schwarz inequality first in H and next in 2(I),

A1 = iI|A|ei,ei
= iIBCei,ei
= iICei,B*ei
iICeiB*ei
(iICei2)1/2(iIB*ei2)1/2
= C2B*2
< .

Hence A1(H), completing the proof. ∎

The following theorem shows that if A1(H) then sums similar to A1 are also finite, and that the series eAe,e does not depend on the orthonormal basis .1212 12 John B. Conway, A Course in Operator Theory, p. 88, Proposition 18.9.

Theorem 19.

If AB1(H) and E is an orthonormal basis for H, then

e|Ae,e|<,

and if F is an orthonormal basis for H then

eAe,e=fAf,f.
Proof.

By Theorem 18, there are B,C2(H) such that A=C*B. If λ and e then

(B-λC)e2 = (B-λC)e,(B-λC)e
= Be,Be-Be,λCe-λCe,Be+λCe,λCe
= Be2-Be,λCe-Be,λCe¯+|λ|2Ce2
= Be2-2ReBe,λCe+|λ|2Ce2.

As (B-λC)e20,

2ReBe,λCeBe2+|λ|2Ce2,

so

2Re(λ¯Be,Ce)Be2+|λ|2Ce2.

This is true for any λ and e. Take |λ|=1, depending on e, such that

λ¯Be,Ce=|Be,Ce|,

which gives

2Re|Be,Ce|Be2+Ce2,

i.e.,

|Be,Ce|12(Be2+Ce2).

Since this inequality doesn’t involve λ, the fact that we chose λ depending on e doesn’t matter, and the above inequality holds for any e. Therefore

e|Ae,e|=e|C*Be,e|=e|Be,Ce|12B22+12C22<,

which is the first statement we wanted to prove and which was necessary to prove even to make sense of the second statement.

Because e|Ae,e|<, the series eAe,e converges. We have to show that its value does not depend on the orthonormal basis . If e, then

(B+C)e2-(B-C)e2 = Be2+Be,Ce+Ce,Be+Ce2
-(Be2-Be,Ce-Ce,Be+Ce2)
= 2Be,Ce+2Ce,Be
= 4ReBe,Ce,

which gives us

eReAe,e = eReBe,Ce
= 14e(B+C)e2-(B-C)e2
= 14B+C22-14B-C22.

Applying this to iA=iC*B gives us, as (iC*)*=-iC,

eReiAe,e=14B-iC22-14B+iC22.

But ReiAe,e=Re(iAe,e)=-ImAe,e, as Re(i(a+ib))=Re(ia-b)=-b. Therefore

e-ImAe,e=14B-iC22-14B+iC22,

i.e.

eImAe,e=14B+iC22-14B-iC22.

Thus

eAe,e=14B+C22-14B-C22+i4B+iC22-i4B-iC22,

and the right-hand side does not depend on the orthonormal basis , completing the proof. ∎

If A1(H) and is an orthonormal basis for H, we define the trace of A, written trA, to be

trA=eAe,e.

It is apparent that tr:1(H) is a positive linear functional: tr is a linear functional 1(H), and if A1(H) is a positive operator, then trA is real and 0. If A1(H) is a positive operator then it is diagonalizable (being a bounded trace class operator implies that it is compact): there is an orthonormal basis {ei:iI} for H such that

A=iIAei,eieiei,

where the series converges in the strong operator topology. Since A is positive, Aei,ei is a real nonnegative number for each iI. trA=0 means that

iIAei,ei=0,

and as this is a series of nonnegative terms they must all be 0. Putting these into the expression for A gives us A=0, showing that tr:1(H) is a positive definite linear functional. We haven’t yet proved that tr is a bounded linear functional. This follows from Theorem 23, which we prove later in this section.

In the following we show that the set of bounded trace class operators is a normed vector space with norm 1, which we call the trace norm.1313 13 John B. Conway, A Course in Operator Theory, p. 89, Theorem 18.11 (a).

Theorem 20.

1(H) is a normed vector space with the norm 1.

Proof.

Let A,B1(H), and let their polar decompositions be A=U|A|,B=V|B|, and A+B=W|A+B|. As bounded trace class operators are compact and as the compact operators are a vector space, A+B is a compact operator. We have already stated that if T is compact then |T| is compact, which follows from the polar decomposition of T and the fact that the compact operators are an ideal. Thus |A+B| is a compact operator. |A+B| is a positive operator, so by the spectral theorem for normal compact operators that we stated as Theorem 13, |A+B| is diagonalizable: there is a countable subset {en:n1} of an orthonormal basis for H, and λn, such that |A+B|en=λnen and |A+B|e=0 if e is not a member of this countable subset. As |A+B| is a positive operator, the eigenvalues λn are real and nonnegative. We have, in the strong operator topology,

|A+B|=n=1λnenen.

First, as |A+B|=W*(A+B) and using the Cauchy-Schwarz inequality in H,

A+B1 = e|A+B|e,e
= n1|A+B|en,en
= n1(A+B)en,Wen
= n1U|A|en,Wen+V|B|en,Wen
= n1|A|en,U*Wen+|B|en,V*Wen
= n1|A|1/2en,|A|1/2U*Wen+|B|1/2en,|B|1/2V*Wen
n1|A|1/2en|A|1/2U*Wen+|B|1/2en|B|1/2V*Wen.

Applying the Cauchy-Schwarz inequality in 2 to this gives

(n1|A|1/2en2)1/2(n1|A|1/2U*Wen2)1/2
+(n1|B|1/2en2)1/2(n1|B|1/2V*Wen2)1/2.
(e|A|1/2e2)1/2(e|A|1/2U*We2)1/2
+(e|B|1/2e2)1/2(e|B|1/2V*We2)1/2.

So we have

A+B1|A|1/22|A|1/2U*W2+|B|1/22|B|1/2V*W2.

By Theorem 18, |A|1/2,|B|1/22(H), and then using the inequality in Theorem 15 gives us

|A|1/2U*W2|A|1/22U*W|A|1/22,

where we used the fact that U* and W are partial isometries and hence either have norm 1 or 0, depending on whether they are the zero map. Likewise,

|B|1/2V*W2|B|1/22.

Therefore we have obtained

A+B1|A|1/222+|B|1/222=A1+B1.

so A+B1(H), and 1 satisfies the triangle inequality.

If A(H) and α then |αA|1/2=|α|1/2|A|1/2, and from this it follows that if A1(H) and α then αA1=|α|A1 and so αA1(H). Therefore 1(H) is a vector space.

If A1=0, then tr|A|=0. We have shown that tr is a positive definite linear functional: hence |A|=0, and so A=0. This completes the proof that 1 is a norm on 1(H). ∎

We have shown that 1(H) with the trace norm 1 is a normed space. In the following theorem we show that a bounded finite rank operator is a bounded trace class operator, and that 00(H) is a dense subset of 1(H).

Theorem 21.

00(H) is a dense subset of the normed space B1(H) with the trace norm.

Proof.

If A00(H), then there is an orthonormal basis {ei:iI} for H and a finite subset J of I such that such that if iIJ then Aei=0, and so |A|ei=0. This gives A1<, so A1(H).

Let {ei:iI} be an orthonormal basis for H. We’ve shown that the bounded finite rank operators are contained in the bounded trace class operators, and now we have to show that if A1(H) and ϵ>0 then there is some B00(H) satisfying A-B1<ϵ. As iI|A|ei,ei<, there is a finite subset J of I such that

iIJ|A|ei,ei<ϵ.

Let P be the orthogonal projection onto span{ei:iJ}, and define B00(H) by B=AP, which gives us

A-B1=iI|A-B|ei,ei=iIJ|A|ei,ei<ϵ.

Taking an adjoint of an operator on a Hilbert space and taking the complex conjugate of a complex number ought to be interchangeable where it makes sense. We show in the following theorem that the adjoint of an element of 1(H) is also an element of 1(H) and that the trace of the adjoint is the complex conjugate of the trace.

Theorem 22.

If AB1(H) then A*B1(H) and

trA*=trA¯.
Proof.

As A1(H), by Theorem 18 there are B,C2(H) such that A=C*B. Then A*=B*C is a product of two bounded Hilbert-Schmidt operators, and so by the same theorem is itself an element of 1(H).

We’re going to extract something from our proof of Theorem 19, which showed that the trace of an operator does not depend on the orthonormal basis that we use: We proved that, with A=C*B,

tr(A)=14B+C22-14B-C22+i4B+iC22-i4B-iC22,

Applying this to the adjoint A*=B*C, and as i(C+iB)=iC-B,

trA* = 14C+B22-14C-B22+i4C+iB22-i4B-iC22
= 14B+C22-14B-C22+i4iC-B22-i4iB+C22
= trA¯.

It is familiar to us that if A and B are matrices then tr(AB)=tr(BA). In the next theorem we show that this is true for bounded linear operators providing one of the two is a bounded trace class operator.1414 14 John B. Conway, A Course in Operator Theory, p. 89, Theorem 18.11 (e). The first thing we prove in this theorem is that 1(H) is an ideal of the algebra (H), so that it makes sense to talk about the trace of a product of two bounded linear operators only one of which is a bounded trace class operator. The theorem also shows that tr:1(H), which we have already shown is a positive definite linear functional, is bounded.

Theorem 23.

If AB1(H) and TB(H), then AT,TAB1(H), and tr(AT)=tr(TA), and |tr(TA)|TA1.

Proof.

A is the product of two bounded Hilbert-Schmidt operators, say A=C*B (we write it this way because this will be handy later in the proof). Hence AT=C*(BT). As 2(H) is an ideal of (H), we have BT2(H), showing that AT is a product of two bounded Hilbert-Schmidt operators, which implies that AT1(H). Similarly, TA1(H). We use from the proof of Theorem 19 the following: (we wrote A=C*B to match the way we wrote A in that theorem)

tr(C*B)=14B+C22-14B-C22+i4B+iC22-i4B-iC22.

Applying this to CB*, and using that the norm of the adjoint of an operator is equal to the norm of the operator itself and that (iC*)*=-iC,

tr(CB*) = 14B*+C*22-14B*-C*22+i4B*+iC*22-i4B*-iC*22
= 14B+C22-14B-C22+i4B-iC22-i4B+iC22
= tr(C*B)¯.

Using this we obtain

tr(TA) = tr((TC*)B)
= tr((TC*)*B*)¯
= tr(CT*B*)¯
= tr(C(BT)*)¯
= tr(C*(BT))
= tr(AT),

which we wanted to show.

We still have to prove that |tr(TA)|TA1. Let A=U|A| be the polar decomposition of A, and let be an orthonormal basis for H. By Theorem 18, |A|1/22(H). (We mention this to justify talking about the Hilbert-Schmidt norm of |A|1/2.) We have, using the Cauchy-Schwarz inequality in H and in 2,

|tr(TA)| = |eTU|A|e,e|
= |e|A|1/2e,|A|1/2U*T*e|
e||A|1/2e,|A|1/2U*T*e|
e|A|1/2e|A|1/2U*T*e
(e|A|1/2e2)1/2(e|A|1/2U*T*e2)1/2
= |A|1/22|A|1/2U*T*2.

By Theorem 15, and as U* is a partial isometry,

|A|1/2U*T*2|A|1/22U*T*|A|1/22T*.

Therefore

|tr(TA)||A|1/222T*=|A|1/222T=A1T,

completing the proof. ∎

In Theorem 22 we proved that the adjoint of a bounded trace class operator is itself trace class, and we now prove that they have the same trace norm.

Theorem 24.

If AB1(H) then

A*1=A1.
Proof.

The polar decomposition A=U|A| satisfies |A*|=U|A|U* and U*U|A|=|A|, as we stated in (2). We have, using this and Theorem 23,

A*1=tr|A*|=tr((U|A|)U*)=tr(U*(U|A|))=tr|A|=A1.

Theorem 25.

If AB1(H) and TB(H), then

AT1A1T,TA1TA1.
Proof.

Let A have the polar decomposition A=U|A| and let TA have the polar decomposition TA=W|TA|. We have, from (2),

|TA|=W*(TA)=W*TU|A|.

It follows from Theorem 23 that, putting S=W*TU,

TA1=tr(|TA|)=tr(S|A|)SA1.

As W* and U are partial isometries, ST, thus

TA1TA1. (3)

On the other hand, Theorem 24 tells us

AT1=T*A*1.

As A*1(H), by (3) we get

T*A*1T*A*1=TA1,

using Theorem 24 again. ∎

In Theorem 16 we proved that if A2(H) then AA2. Now we prove that the trace norm also dominates the operator norm, so that the topology on the normed space 1(H) with the trace norm is finer than its topology as a subspace of (H).

Theorem 26.

If AB1(H), then

AA1.
Proof.

Let A have polar decomposition A=U|A|. As |A| is a compact positive, it is diagonalizable: there is an orthonormal basis {ei:iI} for H and λi such that |A|ei=λiei. On the one hand, the operator norm of a diagonalizable operator is the supremum of the absolute values of its eigenvalues, as we stated in (1). On the other hand,

|A|1=iI|λi|.

Certainly then |A||A|1. But

AU|A||A||A|1=A1,

which is what we wanted to prove. ∎

We have already shown that the trace class operators with the trace norm are a normed space. We now prove that they are a Banach space. We do this by showing that there is an isometric isomorphism ρ:1(H)0(H)*.1515 15 John B. Conway, A Course in Operator Theory, p. 93, Theorem 19.1. The latter space is a Banach space, so if An1(H) is a Cauchy sequence, its image ρ(An) is a Cauchy sequence in 0(H)* and hence has a limit, call it B. Since ρ is surjective, there is some A1(H) such that ρ(A)=B, and one checks that AnA.

For A1(H) and C0(H), define

ΦA(C)=tr(CA).
Theorem 27.

The map ρ:B1(H)B0(H)* defined by

ρ(A)=ΦA,A1(H),

is an isometric isomorphism.

Proof.

Let A1(H). It is apparent that ΦA:0(H) is a linear map. Using Theorem 23, its operator norm is

ΦA=supC1|ΦA(C)|=supC1|tr(CA)|supC1CA1=A1,

where the supremum is taken over compact operators. Hence ΦA0(H)* (if H={0} then the final equality is ). We have ΦAA1, so to show that ρ is an isometric isomorphism, we have to show that if A1(H) then ΦAA1, and that ρ is surjective (as being injective is implied by being an isometry).

Let Φ0(H)*. For g,hH, define

B(g,h)=Φ(gh),B(g,h)v=gh(v)=v,hg.

It is apparent from this that B is a sesquilinear form on H. A sesquilinear form B is said to be bounded if M=supg,h=1|B(g,h)|<. For g=h=1,

|B(g,h)|=|Φ(gh)|ΦghΦgh=Φ.

Thus B is a bounded sesquilinear form on H, and we can therefore apply the Riesz representation theorem,1616 16 Walter Rudin, Functional Analysis, second ed., p. 310, Theorem 12.8. which states that there is a unique T(H) such that

B(g,h)=g,Th,g,hH,

and that this T satisfies T=M.

Let A=T*, let A=U|A| be the polar decomposition of A, and let be an orthonormal basis for H. If 0 is a finite subset of , define

C0=(e0ee)U*=e0eUe.

It is apparent that C000(H), and one checks that C01. We have

e0|A|e,e = |e0|A|e,e|
= |e0U*Ae,e|
= |e0e,TUe|
= |e0B(e,Ue)|
= |e0Φ(eUe)|
= |Φ(C0)|.

Then

e0|A|e,eΦC0Φ.

This is true for any finite subset 0 of , and it follows that

A1Φ, (4)

and thus A1(H).

Let C00(H). Then there are some g1,,gn, h1,,hnH such that (cf. Theorem 1)

C=k=1ngkhk.

We have

Φ(C) = k=1nΦ(gkhk)
= k=1nB(gk,hk)
= k=1nAgk,hk
= k=1ntr(A(gkhk))
= tr(AC)
= ΦA(C).

Since the bounded linear functionals Φ and ΦA agree on 00(H), a dense subset of 1(H) with the trace norm, they are equal. Therefore, ρ is surjective. Moreover, we showed in (4) that A1Φ, so ΦAA1, so ρ is an isometry, which completes the proof. ∎

We have shown that the 1(H) is the dual of 0(H). It can further be shown that (H) is the dual of 1(H). For T(H) we define ΨT:1(H) by ΨT(A)=tr(TA), A1(H). Then the map TΨT is an isometric isomorphism (H)1(H)*.1717 17 John B. Conway, A Course in Operator Theory, p. 94, Theorem 19.2.

9 The Hilbert-Schmidt inner product

If A,B2(H), we define

A,B=tr(B*A),

which makes sense because, by Theorem 18, AB1(H). As tr:1(H) is a positive definite linear functional, we get that ,:2(H)×2(H) is an inner product. We call this the Hilbert-Schmidt inner product. Check that A,A=A22. It is a fact that 2(H) is a complete metric space with the Hilbert-Schmidt norm,1818 18 Gert K. Pedersen, Analysis Now, revised printing, p. 119, Theorem 3.4.9. and hence 2(H) with the Hilbert-Schmidt inner product is a Hilbert space.

Theorem 28.

2(H) with the Hilbert-Schmidt inner product is a Hilbert space.