Distinguished Varieties
Jim Agler Partially supported by the National Science Foundation
U.C. San Diego La Jolla, California 92093
John E. Mc Carthy Partially supported by National Science Foundation Grant DMS 0070639
Washington University St. Louis, Missouri 63130
August 20 2004
Abstract
A distinguished variety is a variety that exits the bidisk through the distinguished boundary. We show that Andô's inequality for commuting matrix contractions can be sharpened to looking at the maximum modulus on a distinguished variety, not the whole bidisk. We show that uniqueness sets for extremal Pick problems on the bidisk always contain a distinguished variety.
0 Introduction
In this paper, we shall be looking at a special class of bordered (algebraic) varieties that are contained in the bidisk
in
.
Definition 0.1
A non-empty set
in
is a distinguished variety if there is a polynomial
in
such that
and such that
|
(0.2)
|
Condition ( 0.2 ) means that the variety exits the bidisk through the distinguished boundary of the bidisk, the torus. We shall use
to denote the set given by ( 0.2 ): topologically, it is the boundary of
within
, the zero set of
, rather than in all of
. We shall always assume that
is chosen to be minimal, i.e. so that no irreducible component of
is disjoint from
and so that
has no repeated irreducible factors. Why should one single out distinguished varieties from other bordered varieties?
One of the most important results in operator theory is T. Andô's inequality [7] (see also [12] and [24] ). This says that if
and
are commuting operators, and both of them are of norm
or less, then for any polynomial
in two variables, the inequality
|
(0.3)
|
holds. Andô's inequality is essentially equivalent to the commutant lifting theorem of B. Sz.-Nagy and C. Foiaş [23] — see e.g. [20] for a discussion of this.
Our first main result, Theorem 3.1 , is that if
and
are matrices, then the inequality ( 0.3 ) can be improved to
where
is some distinguished variety depending on
and
. Indeed, in the proof of the theorem, we construct co-isometric extensions of the matrices that naturally live on this distinguished variety. So when studying bivariable matrix theory, rather than operator theory, one is led inexorably to study distinguished varieties.
Conversely, in Theorem 1.12 , we show that all distinguished varieties can be represented as
for some analytic matrix-valued function
on the disk that is unitary on
.
This shows that the study of distinguished varieties leads back to operator theory. Consider the natural notion of isomorphism of two distinguished varieties, namely that there is a biholomorphic bijection between them.
Definition 0.4
A function
is holomorphic on a set
in
if, at every point
in
, there is a non-empty ball
centered at
and an analytic mapping of two variables defined on
that agrees with
on
.
Definition 0.5
Two distinguished varieties
and
are isomorphic if there is a function
that is holomorphic on
and continuous on
such that
is a bijection from
onto
and such that
is holomorphic on
.
(The requirement that
be holomorphic does not follow automatically from the holomorphicity of
— consider e.g.
and
, which are not isomorphic.) By the maximum modulus principle,
must map the boundary of
onto the boundary of
. It follows that
is a pair of inner functions, i.e. a pair of holomorphic scalar-valued functions that each have modulus one on
. So studying isomorphism classes of distinguished varieties is closely connected to the rich structure of inner functions.
W. Rudin has studied when an arbitrary finite Riemann surface
is isomorphic to a distinguished variety, in the sense that there is an unramified pair of separating inner functions on
that are continuous on
[22] . His results show, for example, that a finitely connected planar domain is isomorphic to a distinguished variety if and only if the domain is either a disk or an annulus. He also showed that for every
, there is a finite Riemann surface
that is topologically an
-holed torus minus one disk, and such that
is isomorphic to a distinguished variety.
In Section 2 we show that, under fairly general conditions, a pair of “inner” functions
on a set
must map
into a distinguished variety (i.e. the algebraic relation on the
's comes for free).
A third reason to study distinguished varieties comes from considering the Pick problem on the bidisk. This is the problem of deciding, given points
in
, and values
in
, whether there is a function in
, the bounded analytic functions on
, that interpolates the data and is of norm at most one. The problem is called extremal if there is an interpolating function of norm exactly one, but not less.
If an extremal Pick problem is given, the solution may or may not be unique (see Section 4 for an example). Our second main result is Theorem 4.1 , where we show that there is always a distinguished variety on which the solution is unique.
One can think then of the Pick problem as having two parts:
(a) Solve the problem on the distinguished variety where the solution is unique.
(b) Parametrize all the extensions of the solution to the whole bidisk.
We give a formula ( 4.10 ) for Problem (a). The extension problem (b) is non-trivial: unless the distinguished variety is isomorphic to a disk, there will always be some functions that cannot be extended to the whole bidisk without increasing the norm [5] . Obviously a function arising from a Pick problem will be extendable, but what distinguishes such functions remains mysterious.
If one starts with an inner function on
and wants to extend this to a rational inner function on
, there can be more than one extension.
However, there is a restriction on the degree, given by Theorem 2.8 . If the variety is of rank
, i.e. there are generically
sheets above every first coordinate and
above every second coordinate, then any regular rational inner extension of degree
must have
equal to the number of zeroes the original function had on
.
An admissible kernel
on a set
in
is an
-by-
positive definite matrix such that
It is known [1, 4] that studying all the admissible kernels on a set is essential to understanding the Pick problem. A key idea in the proof of Theorem 4.1 is that every admissible kernel automatically extends to a distinguished variety.
Distinguished varieties have been studied in a somewhat more abstract and general setting by J. Ball and V. Vinnikov [9] . They have a determinental representation that is analogous to Theorem 1.12 .
We would like to thank the referees for many valuable remarks.
1 Representing Distinguished Varieties
Let
be a distinguished variety. We say a function
is holomorphic on
if, for every point of
, there is an open ball
in
containing the point, and a holomorphic function
of two variables on
, such that
. We shall use
to denote the Banach algebra of functions that are holomorphic on
and continuous on
. This is a uniform algebra on
, i.e. a closed unital subalgebra of
that separates points. The maximal ideal space of
is
.
If
is a finite measure on a distinguished variety
, let
denote the closure in
of the polynomials. If
is an open subset of a Riemann surface
, and
is a finite measure on
, let
denote the closure in
of
, the functions that are holomorphic on
and continuous on
. We say a point
is a bounded point evaluation for
(or
) if evaluation at
, a priori defined only for a dense set of analytic functions, extends continuously to the whole Hilbert space. If
is a bounded point evaluation, we call the function
that has the property that
the evaluation functional at
.
The following lemma is well-known. It is valid in much greater generality, but this will suffice for our purposes. If the boundary of
consists of closed analytic curves, the lemma follows from J. Wermer's proof [25] that
is hypo-Dirichlet, and the description of representing measures for hypo-Dirichlet algebras given by P. Ahern and D. Sarason in [6] . (Actually Wermer's proof extends without difficulty to the case where the boundary is just piecewise
, but we shall not need this fact). For a detailed description of the measures in this case, see K. Clancey's paper [10] .
Lemma 1.1
Let
be a compact Riemann surface. Let
be a domain whose boundary is a finite union of piecewise smooth Jordan curves. Then there exists a measure
on
such that every
in
is a bounded point evaluation for
, and such that the linear span of the evaluation functionals is dense in
.
Proof: Because its boundary is nice,
is regular for the Dirichlet problem (see e.g. [15,SectionIV.2] ). Let
be harmonic measure for
with respect to some fixed base-point. Then by Harnack's inequality, harmonic measure for any other point in the domain is boundedly absolutely continuous with respect to
. As harmonic evaluation functionals are a fortiori analytic evaluation functionals, we get that every point of
is a bounded point evaluation (with an
evaluation functional) for
.
Ahern and Sarason [6,p.159] proved that the span of the evaluation functionals is dense. Their argument, in brief, was to find an exhaustion
of
, i.e. an increasing family of open sets, each contained compactly in the next, whose union was
. Let
be harmonic measure for each
, with respect to the same fixed base-point. Then they showed that for every
in
, its norm was equal to
where
is the harmonic extension of
to
. In particular, any function in
that vanishes identically on
must be the zero function.
Lemma 1.2
Let
be a distinguished variety. There is a measure
on
such that every point in
is a bounded point evaluation for
, and such that the span of the evaluation functionals is dense in
.
Proof: Let
be the minimal polynomial such that
is the intersection of
with
.
Let
be the projective closure of
in
. Let
be the desingularization of
. This means
is a compact Riemann surface (not connected if
is not irreducible) and there is a holomorphic function
that is biholomorphic from
onto
and finite-to-one from
onto
.
Here
is the set of non-singular points in
, and
is the preimage of
.
See e.g. [14] or [17] for details of the desingularization.
Let
. Then
is a finite union of disjoint curves, each of which is analytic except possibly at a finite number of cusps. Let
be the measure from Lemma 1.1 (or the sum of these if
is not connected).
The desired measure
is the push-forward of
by
, i.e. it is defined by
. Indeed, if
is in
and
, let
be a representing measure for
in
. Then the function
is defined
-a.e., and satisfies
Note that
is a finite codimensional subalgebra of
. For a description of what finite codimensional subalgebras look like, see Gamelin's paper [16] .
For positive integers
and
, let
|
(1.3)
|
be an
-by
unitary matrix. Let
|
(1.4)
|
be the
-by-
matrix valued function defined on the unit disk
by the entries of
. This is called the transfer function of
. Because
, a calculation yields
|
(1.5)
|
so
is a rational matrix-valued function that is unitary on the unit circle and contractive on the unit disk. Such functions are called rational matrix inner functions, and it is well-known that all rational matrix inner functions have the form ( 1.4 ) for some unitary matrix decomposed as in ( 1.3 ) — see e.g. [4] for a proof.
Let
be the set
|
(1.6)
|
We shall show that
is a distinguished variety, and that every distinguished variety arises this way.
Lemma 1.7
Let
let
and let
Then
.
Proof: The point
is in
iff there is a non-zero vector
in
such that
|
(1.8)
|
Claim: ( 1.8 ) holds if and only if there is a non-zero vector
in
such that
|
(1.9)
|
Proof of Claim: If ( 1.9 ) holds, then solving gives ( 1.8 ). Conversely, if ( 1.8 ) holds, define
Then ( 1.9 ) holds. Moreover, if
were
, then
would be in the kernel of
and be a
-eigenvector of
. As
, this would force
, contradicting the fact that
.
Given the claim, the point
is in
iff there are non-zero vectors
and
such that
|
(1.10)
|
Interchanging coordinates, ( 1.10 ) becomes
|
(1.11)
|
Clearly, ( 1.9 ) and ( 1.11 ) are equivalent.
Note that if
has a non-trivial kernel
, then ( 1.5 ) shows that
is isometric on
for all
, so by the maximum principle is equal to a constant isometry with initial space
. If
has a trivial kernel, we say
is pure. Every rational inner function decomposes into the direct sum of a pure rational inner function and a unitary matrix — see e.g. [24] .
Since
, we see that
has no kernel iff
. Since
, this in turn is equivalent to
having no kernel. Therefore
is pure iff
is.
Theorem 1.12
The set
, defined by ( 1.6 ) for some rational matrix inner function
, is a distinguished variety. Moreover, every distinguished variety can be represented in this form.
Proof: Suppose
is given by ( 1.6 ), and that
is in
. Without loss of generality, we can assume that
is pure. Indeed, any unitary summand of
would add sheets to the variety
of the type
, for some unimodular
. These sheets are all disjoint from the open bidisk
.
If
, equation ( 1.5 ) then shows that
is a strict contraction, so all its eigenvalues must have modulus less than
, and so
also.
To prove that
implies
, just apply the same argument to
.
Therefore ( 0.2 ) holds, and
is a distinguished variety.
To prove that all distinguished varieties arise in this way, let
be a distinguished variety. Let
be the measure from Lemma 1.2 , and let
be the closure of the polynomials in
. The set of bounded point evaluations for
is precisely
. (It cannot be larger, because
is polynomially convex, and Lemma 1.2 ensures that it is not smaller). Let
be the pair of operators on
given by multiplication by the coordinate functions. They are pure commuting isometries
because the span of the evaluation functionals is dense. The joint eigenfunctions of their adjoints are the evaluation functionals.
By the Sz.-Nagy-Foiaş model theory [24] ,
can be modelled as
, multiplication by the independent variable
on
, a vector-valued Hardy space on the unit circle. In this model,
can be modelled as
, multiplication by
for some pure rational matrix inner function
. A point
in
is a bounded point evaluation for
iff
is a joint eigenvalue for
. In terms of the unitarily equivalent Sz.-Nagy-Foiaş model, this is equivalent to
being an eigenvalue of
.
Therefore
as desired.
If
is the transfer function of a unitary
as in ( 1.3 ), and
is pure, we shall say that
is of rank
. This means that generically there are
sheets above each
, and
sheets above each
.
2 Inner Functions
Rudin's results [22] show that planar annuli can be mapped isomorphically into distinguished varieties by a pair of inner functions. The advantage of doing this is that the coordinate functions are then easier to deal with than the original inner functions. Inner functions on a finite bordered Riemann surface can be shown to satisfy an algebraic equation. In this section, we show that even without the Riemann surface structure, inner functions must satisfy an algebraic equation. The result is reminiscent of Livsic's Cayley-Hamilton theorem for a pair of commuting operators with finite rank imaginary parts — see e.g. the book [19] .
Let
be a set. By a kernel on
we mean a self-adjoint map
that is positive definite, in the sense that for any finite set
of distinct points in
, the self-adjoint matrix
is positive definite.
Given any kernel
, there is a Hilbert space
of functions on
for which
is the reproducing kernel, i.e.
(For details of the passage between a kernel and a Hilbert function space, see e.g. [4] ).
Let
and
be functions on
with modulus less than one at every point. Assume that we can find some kernel
on
so that multiplication by each
is a pure isometry on
with finite dimensional cokernel. For example,
could be a distinguished variety, the
's could be the coordinate functions, and
could be the closure of the polynomials in
. Or,
could be a smoothly bounded planar domain, the
's could be inner functions that are continuous on
and have finitely many zeroes, and
could be the closure in
of the rational functions with poles off
.
Let
be an orthonormal basis for
. Then
is an orthonormal basis for
. So by Bergman's formula [4,Prop2.18] ,
| |
|
(2.1)
|
Similarly, if
is an orthonormal basis for
, we get
|
(2.2)
|
Equating the right-hand sides of ( 2.1 ) and ( 2.2 ) and cross-multiplying, we get
| |
|
(2.3)
|
Let
be the vector in
with components
, and let
. Then ( 2.3 ) can be rewritten as saying that the map
| |
| |
is an isometry on the linear span of the vectors
Even if these vectors do not span all of
, we can always extend
to be a unitary from
onto
, and we shall assume that we have done this.
Let
|
(2.4)
|
be the
-by-
matrix valued function defined on the unit disk
that is the transfer function of
. Moreover, we have
Therefore the points
all lie in the set
|
(2.5)
|
which we know from Theorem 1.12 is a distinguished variety. Thus we have proved:
Theorem 2.6
Let
be a reproducing kernel Hilbert space on a set
. Let
and
be multipliers of
such that multiplication by each
is a pure isometry with finite dimensional cokernel, and such that
. With notation as above, the function
maps
into the distinguished variety
given by ( 2.5 ).
Note that applying Theorem 2.6 to
, the space in Lemma 1.2 , we get the second part of Theorem 1.12 .
If
is a distinguished variety, an inner function on
may or may not extend to an inner function on
. If it does extend, the extension may not be unique. It is curious, however, that there is a rigidity in the degree of this extension. Let
be a rational inner function on
. Then it can be represented as
|
(2.7)
|
for some polynomial
that does not vanish on
[21] , where
and
is a multi-index. The representaion is not unique — e.g. taking
and
, one gets the constant function
. The representation will be unique if
is restricted so that
is finite. In this event, we shall call
the degree of
.
If
is an inner function in
, then it is rational and moreover the function
will not vanish on
[21,Thm.5.2.5] ; we shall call such a function regular.
Theorem 2.8
Let
be a variety of rank
, and let
be a regular rational inner function on
of degree
. Then
restricted to
has exactly
zeroes, counting multiplicities.
Proof: By applying an automorphism of
, we can assume that
is not in
and that all points with first or second coordinate
are regular.
Consider first the case
, i.e.
in ( 2.7 ). Then at each of the
points in
with second coordinate
has a zero of multiplicity
, and each of the
points in
with first coordinate
has a zero of multiplicity
.
Now let
be an arbitrary polynomial that does not vanish on
, normalized so that
. Let
, and
As
increases from
to
, the function
changes continuously from
to
. As each
is in
and is inner, the number of zeroes must remain constant.
Example. Let
be the distinguished variety
, of rank
.
The inner function
can be extended to either the function
of degree
or
of degree
. In either event,
.
3 A sharpening of Andô's inequality
Theorem 3.1
Let
and
be commuting contractive matrices, neither of which has eigenvalues of modulus
. Then there is a distinguished variety
such that, for any polynomial
in two variables, the inequality
|
(3.2)
|
holds.
Proof: Let the dimension of the space on which the matrices act be
.
(i) First, let us assume that each
has
linearly independent unit eigenvectors,
. So we have
for some set of scalars
. As each
is a contraction, we have
is positive semidefinite, so
|
(3.3)
|
As the matrix in ( 3.3 ) is positive semidefinite, it can be represented as the Grammian of vectors
, which can be chosen to lie in a Hilbert space of dimension
equal to the defect of
(the defect of
is the rank of
). So we have
|
(3.4)
|
|
(3.5)
|
Multiplying the first equation by
and the second equation by
, we see that they are equal. Therefore
|
(3.6)
|
Reordering equation ( 3.6 ), we get
|
(3.7)
|
Equation 3.7 says that there is some unitary matrix
|
(3.8)
|
such that
|
(3.9)
|
If the linear span of the vectors
is not all of
, then
will not be unique. In this event, we just choose one such
. Define the
matrix-valued analytic function
by
|
(3.10)
|
For any function
of two variables, scalar or matrix-valued, define
Let
, so
Equation 3.9 implies that
|
(3.11)
|
Let
be the Szegő kernel in the Hardy space
of the unit disk, so
|
(3.12)
|
Let
be the vector in
given by
Consider the pair of isometries
on
, where
is multiplication by the coordinate function (times the identity matrix on
) and
is multiplication by the matrix function
. Then
| |
Therefore the map that sends each
to
gives a unitary equivalence between
and the pair
restricted to the span of the vectors
. Therefore the pair
, acting on the full space
, is a co-isometric extension of
.
Let
be any polynomial (scalar or matrix valued) in two variables. We have
| |
| |
| |
| |
where
and
are the sets
| |
|
(3.14)
|
Equality ( 3.13 ) follows from the observation that
|
(3.15)
|
where the norm on the right is the operator norm on the
matrices.
Equation ( 1.5 ) shows that, except possibly for the finite set
, the matrix
is unitary, and so the norm of any polynomial applied to
is just the maximum value of the norm of the polynomial on the spectrum of
. By continuity, we obtain ( 3.13 ). Taking complex conjugates, ( 3.13 ) gives
the desired inequality.
By Theorem 1.12 , we see that
and
are distinguished varieties, and by construction,
contains the points
.
(ii) Now, we drop the assumption that
be diagonizable. J. Holbrook proved that the set of diagonizable commuting matrices is dense in the set of all commuting matrices [18] . So we can assume that there is a sequence
of commuting matrices that converges to
in norm and such that each pair satisfies the hypotheses of (i), i.e. each
is a pair of commuting contractions that have
linearly independent eigenvectors and no unimodular eigenvalues. Each
has a unitary
associated to it as in ( 3.8 ). By passing to a subsequence if necessary, we can assume that the defects
and
are constant, and that the matrices
converge to a unitary
. The corresponding functions
from ( 3.10 ) will converge to some function
. Let
, and
. Let
be defined by ( 3.14 ) for this
, and
be the variety corresponding to
. Notice that the degrees of
are uniformly bounded.
Claim:
is non-empty.
Indeed, otherwise it would contain no points of the form
for
.
That would mean that
, and so
and
would be zero. That in turn would mean that the submatrices
in
would have all their eigenvalues tending to
, and hence by ( 3.9 ), the eigenvalues of
would all tend to
. Therefore
would have a unimodular eigenvalue, contradicting the hypotheses.
Claim:
is a distinguished variety.
This follows from Theorem 1.12 .
Claim: Inequality ( 3.2 ) holds.
This follows from continuity. Indeed, fix some polynomial
. For every
, for every
, we have
| |
We wish to show that
Suppose not. Then there is some sequence
in
such that
|
(3.16)
|
for some
. Moreover, we can assume that
converges to some point
in
. The point
is in the zero set of
, so if it were in
, then it would be in
. Otherwise,
must be in
. To ensure that
is in
, we must rule out the possibility that some sheet of the zero set of
just grazes the boundary of
without ever coming inside.
But this cannot happen. For every
in
, there are
roots of
, and all of these occur in
. So as
tends to
from inside
, one of the
branches of
must tend to
from inside the disk too. Therefore
is in the closure of
, and ( 3.16 ) cannot happen.
Remark 1. If
has a unimodular eigenvalue
, then the corresponding eigenspace
will be reducing for
. Indeed, writing
the commutativity of
and
means
. As
is not in the spectrum of
, it follows that
.
Therefore for any polynomial
, we have
|
(3.17)
|
By von Neumann's inequality for one matrix, the first entry on the right-hand side of ( 3.17 ) is majorized by
So if we allow the matrices to have unimodular eigenvalues, we can still obtain ( 3.2 ) by adding to
a finite number of disks in the boundary of
.
The new
, however, will not be a distinguished variety.
Remark 2. Once one knows Andô's inequality for matrices, then it follows for all commuting contractions by approximating them by matrices — see [13] for an explicit construction. Of course, the set
must be replaced by the limit points of the sets that occur at each stage of the approximation, and in general this may be the whole bidisk.
Remark 3. We have actually constructed a co-isometric extension of
that is localized to
, and a unitary dilation of
with spectrum contained in
.
4 The uniqueness variety
A solvable Pick problem on
is a set
of points in
and a set
of complex numbers such that there is some function
of norm less than or equal to one in
that interpolates (satisfies
). An extremal Pick problem is a solvable Pick problem for which no function of norm less than one interpolates. The points
are called the nodes, and
are called the values. By interpolating function we mean any function in the closed unit ball of
that interpolates.
Consider the two following examples, in the case
.
Example 1. Let
. Then a moment's thought reveals that the interpolating function is unique, and is given by
.
Example 2. Let
. Then the interpolating function is far from unique — either coordinate function will do, as will any convex combination of them. (A complete description of all solutions is given by J. Ball and T. Trent in [8] ). But on the distinguished variety
, all solutions coincide by Schwarz's lemma. For an arbitrary solvable Pick problem, let
be the set of points in
on which all the interpolating functions in the closed unit ball of
have the same value. The preceding examples show that
may be either the whole bidisk or a proper subset. In the event that
is not the whole bidisk, it is a variety. Indeed, for any
not in
, there are two distinct values
and
so that the corresponding
point Pick problem has a solution.
By [8, 2] these problems have interpolating functions that are rational, of degree bounded by
. The set
must lie in the zero set of the difference of these rational functions. Taking the intersection over all
not in
, one gets that
is the intersection of the zero sets of polynomials. Therefore
is a variety, and indeed, by factoring these polynomials into their irreducible factors, we see that
is the intersection with the bidisk of the zero set of one polynomial, together with possibly a finite number of isolated points. We shall call
the uniqueness variety. (If the problem is not extremal,
is just the original set of nodes).
We shall say that an
-point extremal Pick problem is minimal if none of the
point subproblems is extremal. The main result of this section is that if the uniqueness variety is not the whole bidisk, then it at least contains a distinguished variety running through the nodes. If
, it is shown in [3] that either
or the minimal extremal problem has a solution that is a function of one coordinate function only.
Theorem 4.1
Let
, and let
and
be the data for a minimal extremal Pick problem on the bidisk. The uniqueness variety
contains a distinguished variety
that contains each of the nodes.
For a point
in
, we shall write
and
for the first and second coordinates, respectively. Given a set of points
in
, an admissible kernel
is an
-by-
positive definite matrix, with all the diagonal entries
, such that
|
(4.2)
|
A theorem of the first author [1] asserts that a Pick problem on
is solvable if and only if, for every admissible kernel
, the matrix
|
(4.3)
|
is positive semi-definite (see [11, 8, 2] for alternative proofs). We shall say that an admissible kernel is active if the matrix ( 4.3 ) has a non-trivial null-space, i.e. if it is positive semi-definite but not positive definite.
Lemma 4.4
A solvable Pick problem has an active kernel if and only if it is extremal.
Proof: (
) If the problem were not extremal, then for some
one would have
|
(4.5)
|
for all admissible kernels. Take
to be an active kernel, with
a non-zero vector in the null-space of
. Then taking the inner product of the left-hand side of ( 4.5 ) applied to
with
gives
, which is negative.
(
) As the problem is extremal, for each
there is some admissible kernel
such that
is not positive semi-definite. By compactness of the set of
-by-
positive semi-definite matrices with
's down the diagonal, there therefore exists some positive semi-definite
, satisfying ( 4.2 ), and such that ( 4.3 ) is not positive definite. It just remains to show that this
is actually positive definite, and therefore a kernel.
Suppose it were not, so for some non-zero vector
, we have
. By ( 4.2 ), for each
, the vector
(i.e. the vector whose
component is
) is also in the null-space of
. Iterating this observation, one gets that for any polynomial
, the vector
is in the null-space of
. Taking
to be a polynomial that is
at
and zero on the other nodes, we get
, a contradiction.
Lemma 4.6
Every admissible kernel on a set
can be extended to a continuous admissible kernel on a distinguished variety that contains the points
.
Proof: Let
be an admissible kernel on the set
. As it is positive definite, there are vectors
in
such that
. Because
is admissible, Equations ( 3.4 ) and ( 3.5 ) hold.
Following the proof of Theorem 3.1 , one gets that for every point
in the variety
given by ( 3.14 ), one has non-zero vectors
and
such that
Moreover, as the vector
must just be chosen in the null-space of
it can be chosen continuously. When
is one of the nodes
, we choose
| |
| |
Normalize the vectors so that
Now let
where
is the Szegő kernel on the disk as in ( 3.12 ).
The desired extension of
to
is given by
This is obviously a kernel that extends
, it is continuous on
by construction, and the fact that it is admissible follows, in the language of Theorem 3.1 , from the fact that
and
are contractions.
Proof of Theorem 4.1 :
(Step 1.) By Lemma 4.4 , the problem has an extremal kernel, and by Lemma 4.6 , this kernel can be extended to a distinguished variety
that contains all the nodes. Let us call the extended kernel
. Let
be a non-zero vector in the null-space of
. Let
be any point in
that is not one of the original nodes. Let
be some possible value that an interpolating function can take at
. As the
point Pick problem with nodes
and values
is solvable, and as
is admissible, we must have that
Therefore, for every
, we have
|
(4.7)
|
As
is in the null-space of
, Inequality ( 4.7 ) reduces to
|
(4.8)
|
As this holds for all
, we must have that the linear term vanishes, and so we can solve for
and get
|
(4.9)
|
|
(4.10)
|
As long as both sides of ( 4.9 ) do not reduce to zero, this gives a formula for
, which must therefore be unique. (Step 2.) So far, we have not used the minimality of the problem. Minimality ensures that no component of
can be zero, for otherwise an
point subproblem would have an active kernel.
Fix one of the nodes,
say, and consider what happens when
tends to
along some sheet of
. By continuity,
tends to
for each
.
If
then by continuity
for
in
and close to
, and so formula ( 4.10 ) gives the unique value that the interpolating function must take at
.
Assume instead that
|
(4.11)
|
Consider the
point Pick problem with nodes
, and values
for some
in
. If this problem were solvable, then, since
is an admissible kernel, one would have
|
(4.12)
|
where
Take the inner product of the left-hand side of ( 4.12 ) applied to
with
.
We get
| |
|
(4.13)
|
The first sum in ( 4.13 ) vanishes because
is in the null space of
. The second sum vanishes by hypothesis ( 4.11 ). Therefore for any
, ( 4.13 ) is negative. This means that the value
at
is uniquely determined by the choice of the other
values at
. Therefore this
point subproblem must be extremal, contradicting the minimality hypothesis.
We conclude therefore that ( 4.10 ) gives a well-defined formula for the unique value of
at points
in
near the nodes. As we know that some solution to the problem is given by a rational function, we therefore know that this rational function gives the unique solution near the nodes.
Therefore the union of the irreducible components of
that contain the nodes is a distinguished variety contained in
.
Question 4.14 Is the distinguished variety constructed in the proof equal to all of
? Given any function on any subset of the bidisk, the result in [1] tells whether it can be extended to a function in the closed unit ball of
. If the set is a distinguished variety, and the function is analytic on it, is there a better criterion, which one might think of as solving Problem (b) in the Introduction? Question 4.15 How can one tell whether a function on a distinguished variety extends to all of
without increasing its norm?
References
-
J. Agler. Some interpolation theorems of Nevanlinna-Pick type. Preprint, 1988.
-
J. Agler and J.E. Mc Carthy. Nevanlinna-Pick interpolation on the bidisk. J. Reine Angew. Math., 506:191–204, 1999.
-
J. Agler and J.E. Mc Carthy. The three point Pick problem on the bidisk. New York Journal of Mathematics, 6:227–236, 2000.
-
J. Agler and J.E. Mc Carthy. Pick Interpolation and Hilbert Function Spaces. American Mathematical Society, Providence, 2002.
-
J. Agler and J.E. Mc Carthy. Norm preserving extensions of holomorphic functions from subvarieties of the bidisk. Ann. of Math., 157(1):289–312, 2003.
-
P. Ahern and D. Sarason. The
spaces of a class of function algebras. Acta Math., 117:123–163, 1967.
-
T. Andô. On a pair of commutative contractions. Acta Sci. Math. (Szeged), 24:88–90, 1963.
-
J.A. Ball and T.T. Trent. Unitary colligations, reproducing kernel Hilbert spaces, and Nevanlinna-Pick interpolation in several variables. J. Funct. Anal., 197:1–61, 1998.
-
J.A. Ball and V. Vinnikov. Hardy spaces on a finite bordered Riemann surface, multivariable operator theory and Fourier analysis along a unimodular curve. In Operator Theory Advances and Applications, volume 129, pages 37–56. Birkhäuser, Basel, 2000.
-
K. Clancey. Representing measures on multiply connected planar domains. Illinois Math. J., 35:286–311, 1991.
-
B.J. Cole and J. Wermer. Pick interpolation, von Neumann inequalities, and hyperconvex sets. In Complex Potential Theory, pages 89–129. Kluwer Acad. Publ., Dordrecht, 1994.
-
B.J. Cole and J. Wermer. Andô's theorem and sums of squares. Indiana Math. J., 48:767–791, 1999.
-
S.W. Drury. Remarks on von Neumann's inequality. In R.C. Blei and S.J. Sidney, editors, Banach spaces, Harmonic analysis, and Probability theory, volume 995 of Lecture notes in Math., pages 14–32. Springer Verlag, Berlin, 1983.
-
G. Fischer. Plane algebraic curves. American Mathematical Society, Providence, 2001.
-
H.M. Frakas and I. Kra. Riemann Surfaces. Springer, New York, 1991.
-
T.W. Gamelin. Embedding Riemann surfaces in maximal ideal spaces. J. Funct. Anal., 2:123–146, 1968.
-
P. Griffiths and J. Harris. Principles of algebraic geometry. Wiley, New York, 1978.
-
J.A.R. Holbrook. Polynomials in a matrix and its commutant. Linear Algebra Appl., 48:293–301, 1982.
-
M.S. Livsic, N. Kravitsky, A.S. Markus, and V. Vinnikov. Theorem of commuting nonselfadjoint operators. Kluwer, Dordrecht, 1995.
-
N. K. Nikol'skiĭ and V. Vasyunin. Elements of spectral theory in terms of the free function model Part I: Basic constructions. In S. Axler, J.E. Mc Carthy, and D. Sarason, editors, Holomorphic Spaces, volume 33, pages 211–302. Mathematical Sciences Research Institute Publications, 1998.
-
W. Rudin. Function Theory in Polydiscs. Benjamin, New York, 1969.
-
W. Rudin. Pairs of inner functions on finite Riemann surfaces. Trans. Amer. Math. Soc., 140:423–434, 1969.
-
B. Szokefalvi-Nagy and C. Foiaş. Commutants de certains opérateurs. Acta Sci. Math. (Szeged), 29:1–17, 1968.
-
B. Szokefalvi-Nagy and C. Foiaş. Harmonic Analysis of Operators on Hilbert Space. North Holland, Amsterdam, 1970.
-
J. Wermer. Analytic disks in maximal ideal spaces. Amer. J. Math., 86:161–170, 1964.