The Hardy-Weyl algebra
Abstract
We study the algebra generated by the Hardy operator and the operator of multiplication by on . We call the Hardy-Weyl algebra. We show that its quotient by the compact operators is isomorphic to the algebra of functions that are continuous on and analytic on the interior of for a planar set = , which we call the lollipop. We find a Toeplitz-like short exact sequence for the -algebra generated by .
We study the operator , show that its point spectrum is , and that the eigenvalues grow in multiplicity as the points move to from the left.
1 Introduction
The classical Weyl algebra is generated by the operators of multiplication by , denoted , and differentiation, denoted by . These operators satisfy the commutator relation
(1.1) |
The algebra generated by these relations has been studied extensively in both algebra and operator theoryβsee e.g. the books [14, 12, 19, 9]. In operator theory, the study is complicated by the fact that no bounded operators satisfy (1.1). In this note, we shall study the associated algebras that arise when one replaces the differentiation operator by a bounded integration operator.
The Hardy operator is defined on by
Let denote the Volterra operator
which is a right inverse of . These operators give rise to a new set of relations instead of (1.1), namely
We shall use to mean throughout.
Definition 1.2.
Let denote the closure of the unital algebra generated by and in the norm topology of . We call the Hardy-Weyl algebra.
When is an operator in ? Can all the elements be described? Since the commutator of and is compact, we first describe the quotient of by the compact operators. Define , a subset of the plane, by
We call the lollipop. The lollipop algebra is the Banach algebra of functions that are continuous on and analytic on , equipped with the maximum modulus norm. Let denote the compact operators on , and let ,
Theorem 1.3.
There is a Banach algebra isomorphism from onto . It is given by
Let be defined by , where is the projection of onto . As a corollary to Theorem 1.3 we obtain that every element of can be written uniquely as
(1.4) |
where , is in , , and .
If we look at , the -algebra generated by , we get something similar to the Toeplitz algebra short exact sequence. See [4, 8, 18] for some recent results on the Toeplitz algebra, and [3, 10, 15, 16] for some applications. The lollipop algebra is replaced by the functions that are continuous on the boundary.
Theorem 1.5.
There is a short exact sequence of -algebras
In addition to studying the algebra generated by and , one may also study the smaller algebra generated by and .
Definition 1.6.
The algebra is the norm-closed unital algebra generated by and .
One can see that is a proper subalgebra of by noting that its quotient by the compact operators is isomorphic to .
The lattice of closed invariant subspaces of the Volterra operator was shown independently by Brodskii and Donoghue [5, 6] to be
We let denote the set of bounded operators on that leave invariant every element of . This is a large algebraβit includes all right translation operators for example. It is described by the following theorem of Radjavi and Rosenthal [17, Example 9.26].
Theorem 1.7.
The algebra is the weak operator topology closure of the algebra generated by and .
It follows from Theorem 1.7 that is in the WOT closure of , and hence and have the same WOT closure. However, no extra compact operators are added in the WOT closure.
Theorem 1.8.
In Section 6 we consider the operator . It follows from Theorem 1.3 that generates . We show has a surprisingly rich collection of eigenvectors.
Theorem 1.9.
Let . Then
The algebraic multiplicity of the eigenvalues of on the stick increases as , and hence operators in the closed algebra generated by have the property that is not just in , but is smoother. We prove:
Theorem 1.10.
Let be in the norm-closed algebra generated by . Then is on .
2 Preliminaries
Definition 2.1.
A monomial operator is a bounded linear operator with the property that there exist constants and so that
(2.2) |
We call it a flat monomial operator if there exists some so that for all .
In [2] we showed that every flat monomial operator is in , and hence by Theorem 1.7 in the weak closure of . The Volterra operator is Hilbert-Schmidt, and hence compact. See e.g. [11] for a proof.
Let denote the Hardy space of the unit disk, and the SzegΕ kernel. We shall let denote the unilateral shift on . Let
and let denote the composition operator of composing with .
There is a unitary that is defined on monomials by
(2.3) |
and extended by linearity and continuity to the whole space. If , we shall let denote . It is easy to see that is unitary, as it preserves inner products. In [1] we prove that is given by the formula
(2.4) |
and show that
If is a compact subset of , we shall let denote the Banach algebra of functions that are continuous on , with the maximum modulus norm. We shall let denote the subalgebra of functions that are continuous on and analytic on the interior of , and denote the closure of the polynomials in . A theorem of Mergelyan [13] says that if the complement of is connected, then .
3 The Calkin Hardy-Weyl Algebra
We let denote the ideal of compact operators acting on and set
Evidently, is a 2-sided ideal in . Consequently, we may define an algebra , the Calkin Hardy-Weyl algebra, by
If we let denote the coset of in , i.e.,
Proposition 3.1.
is an abelian Banach algebra.
Proof.
That is a Banach algebra follows from the fact that is closed in . To see that is abelian, observe that as and generate , and generate . Furthermore, as ,
(3.2) |
Likewise, as ,
(3.3) |
so that in particular we have that
As and commute and generate , is abelian. β
3.1 A Uniform Algebra Homeomorphically Isomorphic to
We begin by defining an algebra by gluing together two simpler algebras whose maximal ideal spaces overlap at a single point. Let
where we view as an algebra with the operations
and the norm
We abuse notation by letting
when .
We note that if , then as is self-adjoint and has spectrum equal to , we may form the operator . Likewise, as is cosubnormal and has spectrum equal to , if , then we may form the operator . Concretely,
where , and denotes multiplication by .
Lemma 3.4.
If and , then
Proof.
If we define by the formula
We also define by the formula
Proposition 3.5.
is a continuous unital homomorphism.
Proof.
is linear and . Therefore, is linear and . Also,
so is continuous.
Lemma 3.6.
If is a polynomial in two variables and we define by letting
then
Proof.
If is a polynomial in two variables and we let
then
and
Therefore,
β
Corollary 3.7.
The range of is dense in .
Proof.
Lemma 3.6 suggests that we consider the subset of defined by
We note that it follows from the facts that the polynomials are dense in both and that is dense in .
Lemma 3.8.
If , then
(3.9) |
for all .
Proof.
As is continuous, it suffices to prove the lemma under the assumption that . For satisfying we define a unit vector by the formula
We observe that the mean value theorem for integrals implies that
whenever . Also, as weakly,
whenever is a compact operator acting on . In particular, as is compact and when ,
More generally, if is a polynomial and , write , and we get
Now fix and a compact operator acting on . Using the observations in the previous paragraph we have that
Therefore, as ,
for all and any compact operator acting on . Hence,
for all . As is continuous and is dense in , it follows that (3.9) holds for all . β
Lemma 3.10.
If , then
(3.11) |
for all .
Proof.
We first observe that as , by the Maximum Modulus Theorem it suffices to prove the lemma under the assumption that where . For , let
where is as in (2.3). Clearly, as is a unit vector and is unitary, is a unit vector. Also, as
it follows that , and more generally,
(3.12) |
for all .
Now notice that (2.3) implies that
Claim 3.13.
If and , then
(3.14) |
Proof.
Observe that if is a polynomial and , then Claim 3.13 implies that as . In particular,
(3.15) |
whenever and is a polynomial satisfying .
Lemma 3.16.
is a homeomorphism.
Proof.
Putting together the results of Subsection 3.1 we get the following theorem.
Theorem 3.17.
The map is a homeomorphic unital isomorphism from onto .
3.2 Some Observations on the Gelfand Theory of
If , then there is an isometric isomorphism from onto the lollipop algebra given by
So one could just as well state Theorem 3.17 with replaced by and replaced with the map defined by
Definition 3.18.
Define by
Then Theorem 3.17 says that there is a short exact sequence
Remark 3.19.
By Mergelyanβs theorem, , and since generates , it follows that
generates . We shall examine in Section 6.
4
We shall let denote the -algebra generated by . Since it is irreducible, contains all the compact operators.
The Toeplitz -algebra is the algebra generated by the shift . There is a short exact sequence
(See e.g. [7, 7.23]). A cross-section of is the map that sends a function to the Toeplitz operator on with symbol .
Since , the -algebra generated by is unitarily equivalent to . We wish to think of it as living on , so we must shift things over. Let . For any function defined on some domain in , let be its reflection in the real axis.
Definition 4.1.
Let . Let be defined by
The map is unital and linear, but not multiplicative. One checks that if , then , and if , then .
Theorem 4.2.
There is a short exact sequence
(4.3) |
For every in , its coset in can be written uniquely as
(4.4) |
where , , and . The essential spectrum of as in (4.4) is . If , then the Fredhom index is given by the winding number of about :
Proof: For , we shall let denote its equivalence class in . We have
and
So is abelian. Moreover, for any polynomial in 3 variables, there are polynomials in one variable so that
(4.5) |
Therefore operators of the form (4.5) are dense in .
We wish to prove that is isomorphic to the abelian -algebra . We will use a similar strategy to the proof of Theorem 3.17. Let
The algebra is just , but it is easier to define the functional calculus on it. Define
Let . The following lemma is straightforward to prove.
Lemma 4.6.
(i) Let . Then .
(ii) Let and . Then
Using Lemma 4.6, one can check that is a unital *-homomorphism from into . Its range is dense, so if we can show it has no kernel, then it is a C*-isomorphism.
Lemma 4.7.
If , then
(4.8) |
for all .
Proof.
As is continuous, it suffices to prove the lemma under the assumption that , and as is continuous, we can assume that is a polynomial, and that where and are polynomials.
Therefore,
β
If , by Lemma 4.7 we must have . So must be compact. But is unitarily equivalent to a Toeplitz operator, and there are no non-zero compact Toeplitz operators. Therefore has a trivial kernel, and hence is a -isomorphism.
The claim about the spectrum of now follows from the fact that the spectrum of a function in equals its range. Finally, the claim about the Fredholm index follows from the fact that the Fredholm index at will be unchanged under any homotopy of that keeps outside its range. Then can be homotoped to where and for some integer , and the Fredholm index of is .
5 Compact operators in the little algebra
Recall from Definition 1.6 that is the norm-closed algebra generated by and . We shall prove that every compact operator in lies not just in but in .
For an interval in , let us write for the subspace of that vanishes a.e. off , and let denote projection onto . For we write to denote the rank one operator
The key observation is the following:
Lemma 5.1.
Suppose and . Then .
Proof: We have
The right-hand side is if , and if .
Lemma 5.2.
Every finite rank operator on can be written as an integral operator whose kernel is in .
Proof: Let . Define
Then , and is in .
Lemma 5.3.
Let be in , and let . Then is in if and only if for .
Proof: Sufficiency is clear. To prove necessity, assume that for some , the kernel
is not a.e. As an integral operator is zero if and only if the kernel is a.e., this means that the corresponding integral operator is non-zero, and hence maps a function in to a function that is not a.e. on .
Lemma 5.4.
Let be a rank-one operator. Then is in if and only if for some , the support of is in (i.e. a.e. on ) and the support of is in . In this case, .
Proof: The first part follows from Lemma 5.3. For the second part, observe that if the supports of are in and respectively, then . If and are both in , this proves that .
For the general case, choose continuous functions and that converge to and respectively in . It follows from Lemma 5.1 that converges to in norm as , and that converges to . Therefore .
Theorem 5.5.
Let be a compact operator in . Then , and can be approximated in norm by finite rank operators in .
Proof: Note that means that for all , we have .
Let . First, consider . This can be approximated within by a finite rank operator that is a sum of rank one operators that map to . By Lemma 5.4, this means that this finite rank operator is in .
A similar argument shows that and can both be approximated by finite rank operators in within . Iterating, we get that if is a power of , we can approximate
within by a finite rank operator in .
6 The operator
Let us write for the operator . We know that generates the Calkin Hardy-Weyl algebra . By Theorem 3.17 we know that the spectrum of in is . It is not surprising that are eigenvalues of , since they are eigenvalues of . It is perhaps surprising that every point in the stick, except , is also an eigenvalue. Moreover as we move up the stick to the bulb of the lollipop, the eigenvalues increase in multiplicity.
Theorem 6.1.
(i) .
(ii) The point spectrum of is empty.
(iii) The spectrum of is .
Proof: (i) Suppose . Let . Then we have
As , we get the equation
(6.2) |
with the boundary condition
(6.3) |
The function is continuous. Let denote the relatively open subset of on which it is non-zero.
We get that the solution of (6.2), with , is
(6.4) |
where the constant can a priori be different on different components of (though we show below that is actually connected). Hence
(6.5) |
Case: .
For to be in with , we need
which is the same as . For , we get the eigenfunctions
(6.6) |
When , we get
and
(6.7) |
Now suppose that is not all of . Decompose as a union of disjoint non-empty intervals. On each interval, we have that is given by (6.5), with some constant that can depend on the interval. Choose such an interval, . Then cannot be an end-point of , or we would have that is given by (6.4), and this is discontinuous at the right-hand end-point of .
So assume that the left-hand end-point of is . Then the boundary condition (6.3) is replaced by . On , we have
so to have we need . By continuity, cannot vanish again, so we conclude that and that for the function
(6.8) |
is an eigenvector of with eigenvalue .
(ii) As , the eigenvalue equation becomes
where
As , we get the differential equation
(6.9) |
Solving for , we get
The only solution on an interval that vanishes on the right end-point is the zero solution.
(iii) We know
If is a point in , it must be a Fredholm point. By (i), it is not an eigenvalue of , and by (ii) it is not an eigenvalue of . Therefore has trivial kernel and cokernel, and closed range. Therefore it is invertible, and is in the resolvent of .
Not only are points on the stick of the lollipop eigenvalues, there is some additional smoothness. By a generalized eigenvector of order we mean a vector that satisfies but .
We shall prove the case first.
Lemma 6.10.
At , the operator has generalized eigenvectors of all orders.
Proof: We want to show that if we let from (6.7), then for every there exists so that
(6.11) |
Claim 6.12.
For every there exists a polynomial of degree , with lowest order term of degree , so that the functions
satisfy (6.11).
We prove this by induction on . It is true when . Assume we have proved it up to level , and we want to prove it for . So we wish to solve the equation
(6.13) |
and show that the solution is of the form
(6.14) |
Writing for , equation (6.13) is
This gives us the linear differential equation
Multiply by the integrating factor to get
Therefore
where is a polynomial of degree that may have a constant term, and whose next lowest order term is of degree . This gives
Let
The degree of is two higher than , so it is . There may be a term of order ; the next lowest order term is . But as , one can subtract a multiple of from without changing (6.13), so we can assume that has no term of order .
For points in the stick, a similar method works, but there are restrictions when requiring the generalized eigenvectors to be in . Here is one result.
Lemma 6.15.
Let . Then has a generalized eigenvalue of order at if and only if .
Proof: Let .
(6.16) |
All the functions below are supported on . We wish to find a function that satisfies
(6.17) |
Writing for , this becomes
or
(6.18) |
An integrating factor for (6.18) is . This yields
Integrating, we get
Dividing through by the integrating factor and differentiating, we get
We can choose , since it is the coefficient of . This gives
(6.19) |
Examining this expression, we see that is smooth on , and the first term
vanishes at for every . However the second term
has a singularity that grows like . This is integrable for every , but it is only in for . So we have shown that (6.17) has a solution in if and only if .
One can repeat the argument of Lemma 6.15 to get higher order generalized eigenvectors, as gets closer to .
Lemma 6.20.
Let . Let lie in the interval . Then has generalized eigenvectors up to order at .
Proof: We shall inductively find functions satisfying , with as in (6.16). Let , and write
Then we have
Writing for , we want to solve
(6.21) |
After multiplying by the integrating factor , we have
(6.22) |
Claim 6.23.
There are constants such that the functions satisfy
(6.24) |
Proof of Claim 6.23: By induction on . It is true when . Assume it is true up to . From (6.22) we get for :
(6.25) |
By the inductive hypothesis, the integrand in (6.25) is , so the integral is . So satisfies
(6.26) |
From (6.21) we have
(6.27) |
When we use (6.26) for , we get
Now the claim follows from the inductive hypothesis on .
It follows from (6.27) that that is continuous on , and Claim 6.23 shows that its singularity at is of order . This means is in provided which is the same as .
For later use, let us note that if you track the constants in Claim 6.23, you can show:
Lemma 6.28.
In Claim 6.23 one can take and the consants satisfy
Lemma 6.29.
Let . On the interval one can choose the generalized eigenvectors of order of continuously in , for every , and satisfying , for every .
Proof: Let us write for the choice of generalized eigenvector of order at . Write
We have
On every compact subset , of the functions are uniformly bounded, and a.e., so the map is continuous as a map from into .
For higher , we find as in Lemma 6.20. At each stage, we take the constant in (6.26) to be . (We can do this because will be in the kernel of .) This gives us
Moreover we have if and
By Lemma 6.28, we have that each can be chosen uniformly in for in . For any interval of length , we get
Therefore, as ranges over any compact subset of , the functions are uniformly integrable in . So, by the Vitali convergence theorem, the map
is continuous.
These lemmas say that operators in the closed algebra generated by have certain smoothness properties when mapped by into . The functions get smoother as we get closer to .
Theorem 6.30.
Let be in the norm-closed algebra generated by . Then is on .
Proof: Let . Let be a sequence of polynomials so that . It follows from Theorem 3.17 that converges to uniformly on .
Case: . Let be as in Lemma 6.29. For any polynomial , we have
Moreover,
Let be the linear combination of and that satisfies and . Then
So as functions on , we get that converges to some function .
Claim 6.31.
For all in , we have , and is continuous.
We have , where the coefficients and solve the linear system
By Lemma 6.29, since and are continuous in , so as they are linearly independent, we have that is also continuous in . Therefore is continuous, and converges to locally uniformly on . As converges to , we get that .
We have shown that is in . A similar argument with higher derivatives proves that is in .
Of course one can also find generalized eigenvectors for of all orders at points in , but we already know that is analytic on for every .
7 Open Questions
Question 7.1.
Is there a good description of , the compact operators in ?
Question 7.2.
Is ? To prove this, it is sufficient to show that is irreducible, since then it would contain all the compacts, and its quotient by the compacts would be all of .
Question 7.3.
Do the eigenvectors of span ?
References
- [1] Jim Agler and JohnΒ E. McCarthy. The asymptotic MΓΌntz-SzΓ‘sz theorem. Preprint, 2022.
- [2] Jim Agler and JohnΒ E. McCarthy. A generalization of Hardyβs operator and an asymptotic MΓΌntz-SzΓ‘sz theorem. Preprint, 2022.
- [3] Francesca Arici and Bram Mesland. Toeplitz extensions in noncommutative topology and mathematical physics. In Geometric methods in physics XXXVIII, Trends Math., pages 3β29. BirkhΓ€user/Springer, Cham, [2020] Β©2020.
- [4] Laura Brake and Wilhelm Winter. The Toeplitz algebra has nuclear dimension one. Bull. Lond. Math. Soc., 51(3):554β562, 2019.
- [5] M.Β S. Brodskii. On a problem of I. M. Gelfand. Uspehi Mat. Nauk (N.S.), 12(2(74)):129β132, 1957.
- [6] W.Β F. Donoghue, Jr. The lattice of invariant subspaces of a completely continuous quasi-nilpotent transformation. Pacific J. Math., 7:1031β1035, 1957.
- [7] R.Β G. Douglas. Banach Algebra Techniques in Operator Theory. Academic Press, New York, 1972.
- [8] Philip Easo, Esperanza Garijo, Sarunas Kaubrys, David Nkansah, Martin Vrabec, David Watt, Cameron Wilson, Christian BΓΆnicke, Samuel Evington, Marzieh Forough, Sergio GirΓ³nΒ Pacheco, Nicholas Seaton, Stuart White, MichaelΒ F. Whittaker, and Joachim Zacharias. The Cuntz-Toeplitz algebras have nuclear dimension one. J. Funct. Anal., 279(7):108690, 14, 2020.
- [9] Franco Gallone. Hilbert space and quantum mechanics. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2015.
- [10] Shin Hayashi. Topological invariants and corner states for Hamiltonians on a three-dimensional lattice. Comm. Math. Phys., 364(1):343β356, 2018.
- [11] Peter Lax. Functional Analysis. Wiley, 2002.
- [12] Toufik Mansour and Matthias Schork. Commutation relations, normal ordering, and Stirling numbers. Discrete Mathematics and its Applications (Boca Raton). CRC Press, Boca Raton, FL, 2016.
- [13] S.Β N. Mergelyan. Uniform approximations of functions of a complex variable. Uspehi Matem. Nauk (N.S.), 7(2(48)):31β122, 1952.
- [14] Toshio Oshima. Fractional calculus of Weyl algebra and Fuchsian differential equations, volumeΒ 28 of MSJ Memoirs. Mathematical Society of Japan, Tokyo, 2012.
- [15] Efton Park. Index theory and Toeplitz algebras on certain cones in . J. Operator Theory, 23(1):125β146, 1990.
- [16] Efton Park and Claude Schochet. On the -theory of quarter-plane Toeplitz algebras. Internat. J. Math., 2(2):195β204, 1991.
- [17] Heydar Radjavi and Peter Rosenthal. Invariant subspaces. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 77. Springer-Verlag, New York-Heidelberg, 1973.
- [18] Adam Rennie, David Robertson, and Aidan Sims. PoincarΓ© duality for Cuntz-Pimsner algebras. Adv. Math., 347:1112β1172, 2019.
- [19] Konrad SchmΓΌdgen. An invitation to unbounded representations of -algebras on Hilbert space, volume 285 of Graduate Texts in Mathematics. Springer, Cham, [2020].