The singular value decomposition of compact operators on Hilbert spaces
1 Preliminaries
The purpose of these notes is to present material about compact operators on Hilbert spaces that is special to Hilbert spaces, rather than what applies to all Banach spaces. We use statements about compact operators on Banach spaces without proof. For instance, any compact operator from one Banach space to another has separable image; the set of compact operators from one Banach space to another is a closed subspace of the set of all bounded linear operators; a Banach space is reflexive if and only if the closed unit ball is weakly compact (Kakutani’s theorem); etc. We do however state precisely each result that we are using for Banach spaces and show that its hypotheses are satisfied.
Let be the set of positive integers. We say that a set is countable if it is bijective with a subset of . In this note I do not presume unless I say so that any set is countable or that any Hilbert space is separable. A neighborhood of a point in a topological space is a set that contains an open set that contains the point; one reason why it can be handy to speak about neighborhoods of a point rather than just open sets that contain the point is that the set of all neighborhoods of a point is a filter, whereas it is unlikely that the set of all open sets that contain a point is a filter. If , we denote .
2 Bounded linear operators
An advantage of working with normed spaces rather than merely topological vector spaces is that continuous linear maps between normed spaces have a simple characterization. If and are normed spaces and is linear, the operator norm of is
If , then we say that is bounded.
Theorem 1.
If and are normed spaces, a linear map is continuous if and only if it is bounded.
Proof.
Suppose that is continuous. In particular is continuous at , so there is some such that if then . If then, as is linear,
Thus , so is bounded.
Suppose that is bounded. Let , and let . If , then
Hence is continuous at , and so is continuous. ∎
If and are normed spaces, we denote by the set of bounded linear maps . It is straightforward to check that is a normed space with the operator norm. One proves that if is a Banach space, then is a Banach space,11 1 Walter Rudin, Functional Analysis, second ed., p. 92, Theorem 4.1. and if is a Banach space one then checks that is a Banach algebra. If is a normed space, we define , which is a Banach space, called the dual space of .
If and are normed spaces and is linear, we say that has finite rank if
is finite. If is infinite dimensional and , let be a Hamel basis for , let be a countable subset of , and let be nonzero. If we define by and if , then is a linear map with finite rank yet is unbounded. Thus a finite rank linear map is not necessarily bounded. We denote by the set of bounded finite rank linear maps , and check that is a vector space. If is a Banach space, one checks that is an ideal of the algebra (if we either pre- or postcompose a linear map with a finite rank linear map, the image will be finite dimensional).
If and are Banach spaces, we say that is compact if the image of any bounded set under is precompact (has compact closure). One checks that if a linear map is compact then it is bounded (unlike a finite rank linear map, which is not necessarily bounded). There are several ways to state that a linear map is compact that one proves are equivalent: is compact if and only if the image of the closed unit ball is precompact; is compact if and only if the image of the open unit ball is precompact; is compact if and only if the image under it of any bounded sequence has a convergent subsequence. In a complete metric space, the Heine-Borel theorem asserts that a set is precompact if and only if it is totally bounded (for any , the set can be covered by a finite number of balls of radius ). We denote by the set of compact linear maps , and it is straightforward to check that this is a vector space. One proves that if an operator is in the closure of the compact operators then the image of the closed unit ball under it is totally bounded, and from this it follows that is a closed subspace of . is an ideal of the algebra : if and , one checks that and .
Let and be Banach spaces. Using the fact that a bounded set in a finite dimensional normed vector space is precompact, we can prove that a bounded finite rank operator is compact: . Also, it doesn’t take long to prove that the image of a compact operator is separable: if then has a countable dense subset. (We can prove this using the fact that a compact metric space is separable.)
If is a Hilbert space and are subsets of , we define to be the closure of the span of . We say that is an orthonormal basis for if and .
A sesquilinear form on is a function that is linear in its first argument and that satisfies . If is a sesquilinear form on , we say that is bounded if
The Riesz representation theorem22 2 Walter Rudin, Functional Analysis, second ed., p. 310, Theorem 12.8. states that if is a bounded sesquilinear form on , then there is a unique such that
and . It follows from the Riesz representation that if , then there is a unique such that
and . is a -algebra: if and then , , , , and .
We say that is normal if , and self-adjoint if . One proves using the parallelogram law that is self-adjoint if and only if for all . If is self-adjoint, we say that is positive if for all .
3 Spectrum in Banach spaces
If and are Banach spaces and is a bijection, then its inverse function is linear, since the inverse of a linear bijection is itself linear. Because is a surjective bounded linear map, by the open mapping theorem it is an open map: if is an open subset of then is an open subset of , and it follows that . That is, if a bounded linear operator from one Banach space to another is bijective then its inverse function is also a bounded linear operator.
If is a Banach space and , the spectrum of is the set of those such that the map is not a bijection. One proves that is nonempty (the proof uses Liouville’s theorem, which states that a bounded entire function is constant). One also proves that if then . We define the spectral radius of to be
and so . Because is an ideal in the algebra , if is invertible then is compact. One checks that if is compact then is finite dimensional (a locally compact topological vector space is finite dimensional), and therefore, if is an infinite dimensional Banach space and , then .
The resolvent set of is . One proves that is open,33 3 Gert K. Pedersen, Analysis Now, revised printing, p. 131, Theorem 4.1.13. from which it then follows that is a compact set. For , we define
called the resolvent of .
If is a Banach space and , then the point spectrum of is the set of those such that is not injective. In other words, to say that is to say that
If , we say that is an eigenvalue of , and call its geometric multiplicity.44 4 There is also a notion of algebraic multiplicity of an eigenvalue: the algebraic multiplicity of is defined to be For self-adjoint operators this is equal to the geometric multiplicity of , while for a operator that is not self-adjoint the algebraic multiplicity of an eingevalue may be greater than its geometric multiplicity. Nonzero elements of are called generalized eigenvectors or root vectors. See I. C. Gohberg and M. G. Krein, Introduction to the Theory of Linear Nonselfadjoint Operators in Hilbert Space. It is a fact that each nonzero eigenvalue of has finite geometric multiplicity, and it is also a fact that if , then is a bounded countable set and that if has a limit point that limit point is . The Fredholm alternative tells us that
If is infinite dimensional then , and might or might not include . If , check that is a finite set.
If is a Hilbert space, using the fact that a bounded linear operator is invertible if and only if both and are bounded below ( is bounded below if there is some such that for all ), one can prove that the spectrum of a bounded self-adjoint operator is a set of real numbers, and the spectrum of a bounded positive operator is a set of nonnegative real numbers.
4 Numerical radius
If is a Hilbert space and , the numerical range of is the set .55 5 The Toeplitz-Hausdorff theorem states that the numerical range of any bounded linear operator is a convex set. See Paul R. Halmos, A Hilbert Space Problem Book, Problem 166. The closure of the numerical range contains the spectrum of .66 6 Paul R. Halmos, A Hilbert Space Problem Book, Problem 169. If is normal, then the closure of its numerical range is the convex hull of the spectrum: Problem 171. The numerical radius of is the supremum of the numerical range of : . If is self-adjoint, one can prove that77 7 John B. Conway, A Course in Functional Analysis, second ed., p. 34, Proposition 2.13.
The following theorem asserts that a compact self-adjoint operator has an eigenvalue whose absolute value is equal to the norm of the operator. Thus in particular, the spectral radius of a compact self-adjoint operator is equal to its numerical radius. Since a self-adjoint operator has real spectrum, to say that is to say that either or . A compact operator on a Hilbert space can have empty point spectrum (e.g. the Volterra operator on ) and a bounded self-adjoint operator can have empty point spectrum (e.g. the multiplication operator on ), but this theorem shows that if an operator is compact and self-adjoint then its point spectrum is nonempty.
Theorem 2.
If is compact and self-adjoint then at least one of is an eigenvalue of .
Proof.
Because is self-adjoint, . Also, as is self-adjoint, is a real number, and thus either or . In the first case, due to being a supremum there is a sequence , all with norm , such that . Using that is compact, there is a subsequence such that converges to some , and because each has norm . Using ,
Therefore, as , the sequence tends to , so , i.e. is an eigenvector for the eigenvalue . If the argument goes the same. ∎
5 Polar decomposition
If is a Hilbert space and is positive, there is a unique positive element of , denoted , satisfying , which we call the positive square root of .88 8 Gert K. Pedersen, Analysis Now, revised printing, p. 92, Proposition 3.2.11. If one checks that is positive, and hence has a positive square root, which we denote by and call the absolute value of . One proves that is the unique positive operator in satisfying
An element of is said to be a partial isometry if there is a closed subspace of such that the restriction of to is an isometry and . One proves that is the orthogonal projection of onto . It can be proved that if then there is a unique partial isometry satisfying both and .99 9 Gert K. Pedersen, Analysis Now, revised printing, p. 96, Theorem 3.2.17. This is called the polar decomposition of . The polar decomposition satisfies
6 Spectral theorem
If , we define by . is linear, and
so . Depending on whether the image of is or the span of , and in either case . If either of or is then has rank , and otherwise has rank , and it is an orthogonal projection precisely when is a multiple of .
If is an orthonormal set in a Hilbert space , then is an orthonormal basis for if and only if the unordered sum
converges strongly to .1010 10 John B. Conway, A Course in Functional Analysis, second ed., p. 16, Theorem 4.13.
Let’s summarize what we have stated so far about the spectrum and point spectrum of a compact self-adjoint operator on a Hilbert space.
Theorem 3 (Spectrum of compact self-adjoint operators).
If is a Hilbert space and is self-adjoint, then:
-
•
is a nonempty compact subset of .
-
•
If is infinite dimensional, then .
-
•
.
-
•
is countable.
-
•
If is a limit point of , then .
-
•
At least one of is an element of .
-
•
Each nonzero eigenvalue of has finite geometric multiplicity: If and , then .
-
•
If , then is a finite set.
We say that is diagonalizable if there is an orthonormal basis for and a bounded set such that the unordered sum
converges strongly to .
The following is the spectral theorem for normal compact operators.1111 11 Gert K. Pedersen, Analysis Now, revised printing, p. 108, Theorem 3.3.8.
Theorem 4 (Spectral theorem).
If is normal, then is diagonalizable.
The last assertion of Theorem 3 is that a bounded self-adjoint finite rank operator on a Hilbert space has finitely many elements in its point spectrum. Using the spectral theorem, we get that if is self-adjoint and is finite, then is finite rank. In the notation we introduce in the following definition, precisely when has finite rank.
Definition 5.
If is self-adjoint, define to be the sum of the geometric multiplicities of the nonzero eigenvalues of :
Define
to be the sequence whose first term is the element of with largest absolute value repeated as many times as its geometric multiplicity. If are both nonzero elements of , we put the positive one first. We repeat this for the remaining elements of . If , we define for .
Using the spectral theorem and the notation in the above definition we get the following.
Theorem 6.
If is self-adjoint, then there is an orthonormal set in such that
converges strongly to .
If , then its absolute value is a positive compact operator, and for all .
Definition 7.
If and is an eigenvalue of , we call a singular value of , and we define
Because the absolute value of the absolute value of an operator is the absolute value of the operator, if and then . If and , one proves that is an eigenvalue of if and only if is an eigenvalue of and that they have the same geometric multiplicity. From this we get that for all .
7 Finite rank operators
Theorem 8 (Singular value decomposition).
If is a Hilbert space and has , then there is an orthonormal set and an orthonormal set such that
Proof.
is a positive operator with , and according to Theorem 6, there is an orthonormal set in such that
Using the polar decomposition ,
Define . As and as ,
showing that is an orthonormal set. ∎
If , then
so .
Theorem 9.
If then .
Proof.
Let have the singular value decomposition
Taking the adjoint, and because ,
is a sum of finite rank operators and is therefore itself a finite rank operator. ∎
8 Compact operators
If and are Banach spaces, . But if is a Hilbert space we can say much more: is a dense subset of . In other words, any compact operator on a Hilbert space can be approximated by a sequence of bounded finite rank operators.1212 12 John B. Conway, A Course in Functional Analysis, second ed., p. 41, Theorem 4.4. As the adjoint of each of these finite rank operators is itself a bounded finite rank operator,
so . Because each bounded finite rank operator is compact and is closed, this establishes that . (In fact, it is true that the adjoint of a compact linear operator between Banach spaces is itself compact, but there we don’t have the tool of showing that the adjoint is the limit of the adjoints of finite rank operators.)
If is a Hilbert space, the weak topology is the topology on such that a net converges to weakly if for all the net converges to in . Let be the closed unit ball in , and let it be a topological space with the subspace topology inherited from with the weak topology. Thus, a net converges to if and only if for all the net converges to .
Theorem 10.
If is a Hilbert space, , and is the closed unit ball in with the subspace topology inherited from with the weak topology, then is compact if and only if is continuous.
Proof.
Suppose that is compact and let be a net in that converges weakly to some . If , then there is some with . Let have the singular value decomposition
We have, using that the are orthonormal,
Eventually this is , and for such ,
We have shown that in the no rm of , and this shows that is continuous.
Suppose that is continuous. Kakutani’s theorem states that a Banach space is reflexive if and only if the closed unit ball is weakly compact. A Hilbert space is reflexive, hence , the closed unit ball with the weak topology, is a compact topological space.1313 13 cf. Paul R. Halmos, A Hilbert Space Problem Book, Problem 17. Since is continuous and is compact, the image is compact (the image of a compact set under a continuous map is a compact set). We have shown that the image of the closed unit ball is a compact subset of , and this shows that is compact; in fact, to have shown that is compact we merely needed to show that the image of the closed unit ball is precompact, and is a Hausdorff space so a compact set is precompact. ∎
A compact linear operator on an infinite dimensional Hilbert space is not invertible, lest be compact. However, operators of the form may indeed be invertible.1414 14 Ward Cheney, Analysis for Applied Mathematics, p. 94, Theorem 2.
Theorem 11.
If is a normal operator with diagonalization
and , then is invertible and
where the series converges in the strong operator topology.
Proof.
As we have , and as we have . Define
and if , then, for any ,
By Bessel’s inequality, , hence as ; this depends on , and this is why the claim is stated merely for the strong operator topology and not the norm topology. We have shown that is a Cauchy sequence in and hence converges. We define to be this limit. For ,
whence
showing that . It is straightforward to check that is linear, thus . (Thus is a strong limit of finite rank operators. But if is infinite dimensional then is in fact not the norm limit of the sequence: for if it were it would be compact, and we will show that is invertible, which would tell us that is compact, contradicting being infinite dimensional.)
For ,
where the final equality is because the series is the diagonalization of . On the other hand,
showing that . ∎
We can start with a function and ask what kind of series it can be expanded into, or we can start with a series and ask what kind of function it defines. The following theorem does the latter. It shows that if and are each orthonormal sequences and is a sequence of complex numbers whose limit of , then the series
converges and is an element of .
Theorem 12.
If is a Hilbert space, is an orthonormal set, is an orthonormal set, and is a sequence tending to , then the sequence
converges to an element of .
Proof.
Let and let be such that if then . If and , then, as the are orthonormal,
By Bessel’s inequality, , and hence
As this holds for all ,
showing that is a Cauchy sequence, which therefore converges in . As each term in the sequence is finite rank and so compact, the limit is a compact operator. ∎
Continuing the analogy we used with the above theorem, now we start with a function and ask what kind of series it can be expanded into. This is called the singular value decomposition of a compact operator. Helemskii calls the series in the following theorem the Schmidt series of the operator.1515 15 A. Ya. Helemskii, Lectures and Exercises on Functional Analysis, p. 215, Theorem 1. We have already presented the singular value decomposition for finite rank operators in Theorem 8.
Theorem 13 (Singular value decomposition).
If is a Hilbert space and
then there is an orthonormal set and an orthonormal set such that , where
Proof.
As is self-adjoint and compact, by Theorem 6 there is an orthonormal set such that
That is, with defined by
we have .
Let be the polar decomposition of , and define . As and as (because is not finite rank), we have , and hence
showing that is an orthonormal set. Define
and we have . As and , we get
showing that . ∎
9 Courant min-max theorem
Theorem 14 (Courant min-max theorem).
Let be an infinite dimensional Hilbert space and let be a positive operator. If then
and
Proof.
is compact and positive, so according to Theorem 6 there is an orthonormal set such that
For , let . , so has codimension . (The codimension of a closed subspace of a Hilbert space is the dimension of its orthogonal complement.) If is a dimensional subspace of , then there is some with . This is because if is a closed subspace with codimension of a Hilbert space and is a dimensional subspace of the Hilbert space, then their intersection is a subspace of nonzero dimension. As , there are , , with
As the sequence is nonincreasing,
where we write
This shows that if then
Let , and let , , with . As is a finite dimensional Hilbert space, the unit sphere in it is compact, so there a a subsequence that converges to some , . We have
As , we get . As , we get
As and , we have in fact
This is true for any dimensional subspace of , so
If then , , and
so in fact
which is the first of the two formulas that we want to prove.
For , let . If is a dimensional subspace of , then is a closed subspace with codimension , so the intersection of and has nonzero dimension, and so there is some with . As there are with , giving
This shows that
Define . Because is a supremum, there is a sequence on the unit sphere in such that . The unit sphere in is compact because is finite dimensional, so this sequence has a convergent subsequence . As
and , we get
whence
As this is true for any dimensional subspace ,
But for we have , , and
which implies that
∎
If is compact, then the eigenvalues of are equal to the singular values of . Therefore the Courant min-max theorem gives expressions for the singular values of a compact linear operator on a Hilbert space, whether or not the operator is itself self-adjoint.
Allahverdiev’s theorem1616 16 I. C. Gohberg and M. G. Krein, Introduction to the Theory of Linear Nonselfadjoint Operators in Hilbert Space, p. 28, Theorem 2.1; cf. J. R. Retherford, Hilbert Space: Compact Operators and the Trace Theorem, p. 75 and p. 106. gives an expression for the singular values of a compact operator that does not involve orthonormal sets, unlike Courant’s min-max theorem. Thus this formula makes sense for a compact operator from one Banach space to another.
Theorem 15 (Allahverdiev’s theorem).
Let be a Hilbert space and let be the set of bounded finite rank operators of rank . If and , then
10 Schatten class operators
If and , we define
and define to be those with . In other words, an element of is a compact operator whose sequence of singular values is an element of . We call an element of a Schatten class operator. We call elements of trace class operators and elements of Hilbert-Schmidt operators.
If is positive, then, according to Theorem 6, there is an orthonormal set such that
where the series converges in the strong operator topology. As the are orthonormal, we have
which is itself a positive compact operator, and thus for . Therefore, if is a positive compact operator, then .
If and , then and . Hence, if then
As is compact and self-adjoint, it has an eigenvalue with absolute value , from which it follows that if then .
Theorem 16.
If , , and , then
Proof.
For all ,
Applying the Courant min-max theorem to the positive operators and , if then
But
and
so taking the square root,
∎
Using Theorem 16, if then
The following theorem states that the Schatten class operators are Banach spaces.1717 17 Gert K. Pedersen, Analysis Now, revised printing, p. 124, E 3.4.4
Theorem 17.
If , then is a Banach space with the norm .
11 Weyl’s inequality
Weyl’s inequality relates the eigenvalues of a self-adjoint compact operator with its singular values.1818 18 Peter D. Lax, Functional Analysis, p. 336, chapter 30, Lemma 7. We use the notation from Definition 5. For the left hand side is equal to so the inequality is certainly true then.
Theorem 18 (Weyl’s inequality).
If is self-adjoint and , then
Proof.
Let
which is finite dimensional. Check that is an invariant subspace of , and let be the restriction of to . is a positive operator. As is spanned by eigenvectors for nonzero eigenvalues of it follows that , and as is finite dimensional, we get that is invertible. If has polar decomposition , then is invertible; if a partial isometry is invertible then it is unitary, so is unitary, and therefore the eigenvalues of all have absolute value . As the determinant of a linear operator on a finite dimensional vector space is the product of its eigenvalues counting algebraic multiplicity,
(1) |
Theorem 19.
If , is self-adjoint, and , then
Proof.
Schur’s majorization inequality1919 19 Peter D. Lax, Functional Analysis, p. 337, chapter 30, Lemma 8; cf. J. Michael Steele, The Cauchy-Schwarz Master Class, p. 201, Problem 13.4. states that if and are nonincreasing sequences of real numbers satisfying, for each ,
and is a convex function with , then for every ,
With the hypotheses of Theorem 18, for , define and and let . By Theorem 18 these satisfy the conditions of Schur’s majorization inequality, which then gives us for that
If then . ∎
12 Rayleigh quotients for self-adjoint operators
If is self-adjoint, we define the Rayleigh quotient of by
Let and be normed spaces, an open subset of , and a function. If and there is some such that
(2) |
then is said to be Fréchet differentiable at , and is called the Fréchet derivative of at ;2020 20 Ward Cheney, Analysis for Applied Mathematics, p. 149. it does not take long to prove that if both satisfy (2) then . We denote the Fréchet derivative of at by . is a map from the set of all points at which is Fréchet differentiable to .
To say that is a stationary point of is to say that is Fréchet differentiable at and that the Fréchet derivative of at is the zero map. One proves that if are Fréchet derivatives of at then , and thus speak about the Fréchet derivative of at
Theorem 20.
If is self-adjoint, then each eigenvector of is a stationary point of the Rayleigh quotient of .
Proof.
If is an eigenvalue of then, as is self-adjoint, . Let satisfy . We have
For , using that is self-adjoint and that ,
Therefore
As the right-hand side tends to (one of the terms tends to , one doesn’t depend on , and the denominator is bounded below in terms just of for sufficiently small ), showing that is the Fréchet derivative of at . ∎