Abstract
We prove a Julia inequality for bounded non-commutative functions on polynomial polyhedra. We use this to deduce a Julia inequality for holomorphic functions on classical domains in ℂd. We look at differentiability at a boundary point for functions that have a certain regularity there.
1 Introduction
The classical Julia inequality asserts that if a holomorphic function ϕ maps the unit disk 𝔻 to itself, and if at some boundary point τ ∈ ∂𝔻 one has
| (1.1) |
then there exists ω ∈ ∂𝔻 such that
| (1.2) |
The inequality was proved, with an extra regularity hypothesis on φ, by G. Julia in [22], and in the form stated by C. Carathéodory in [16]. D. Sarason found a proof using model theory [33, Chap VI].
Generalizations of Julia’s inequality have been found for functions on the ball by M. Hervé [20], W. Rudin [32, Sec. 8.5] and M. Jury [23], and on the polydisk by K. Wlodarczyk [37], F. Jafari [21] and M. Abate [1]. In the case of the bidisk, a detailed analysis of points for which the analogue of the Julia quotient (1.1) remains bounded (these are called B-points) has been carried out in [7, 8, 10, 15].
It is the purpose of this note to extend Julia’s inequality, and the study of B-points, to non-commutative functions (which we shall define in Subsection 1.1 below) that are bounded on polynomial polyhedra. Our methods rely on the model-theoretic ideas of [8].
Our results are of interest even in the commutative case, because they provide a unified approach to proving boundary versions of the Schwarz-Pick lemma in the Schur-Agler class of various domains, such as the polydisk, or the multipliers of the Drury-Arveson space. The methods also show that at B-points where the function is not analytic, it does have directional derivatives in all directions pointing into the set, and the derivative is a holomorphic (but not necessarily linear) function of the direction. This is explained in Section 6 below.
1.1 Non-commutative Functions
Non-commutative function theory, which originated in the work of J.L. Taylor [34,35], has recently started to flourish [6,11,12,14,17–19,25–28,30,31]. The foundations are developed in the book [24].
The idea is to study functions of non-commuting variables that are generalized non-commuting polynomials in an analogous fashion to thinking of a holomorphic function as a generalization of a polynomial. Our domains are domains of d-tuples of matrices, but they don’t reside in just one dimension. We let 𝕄n denote the n-by-n complex matrices, and let
We shall call a function ϕ defined on a subset of 𝕄[d] graded if, whenever , then ϕ(x) ∈ 𝕄n. If and , we shall let x ⊕ y denote the d-tuple in 𝕄n+m obtained by direct summing each component. If and s is an invertible matrix in 𝕄n, then s−1xs denotes the d-tuple (s−1x1s, …, s−1xds).
Definition 1.3
An nc-function ϕ on a set Ω ⊆ 𝕄[d] is a graded function that respects direct sums and joint similiarities, i.e.
where the equations are only required to hold when the arguments on both sides are in Ω, and in the second one if , then s is invertible in 𝕄n.
(We use superscripts for components, since we shall have many sequences indexed by subscripts.) Notice that every non-commutative polynomial is an nc-function on 𝕄[d]. A particularly nice class of domains on which to study nc-functions are polynomial polyhedra. These are defined in terms of a matrix δ, each of whose entries is a non-commutative polynomial in d variables. Then 𝒢δ is defined as
| (1.4) |
where δ(x), a matrix of matrices, is given the operator norm. A primary example is the d-dimensional noncommutative polydisk, which is the set
| (1.5) |
For this set, we can take δ to be the diagonal d-by-d matrix with the co-ordinate functions on the diagonal. Another well-studied example (see e.g. [29]) is the non-commutative ball, which we shall take to be the column contractions1,
| (1.6) |
which is obtained by letting
1.2 Principal Results
Let 𝒢δ be defined by (1.4). We shall let ℬδ denote the topological boundary of 𝒢δ. If x ∈ ℬδ, then ‖δ(x)‖ = 1, but the converse need not hold — e.g. with d = 1, take . Then 𝒢δ is empty, but ‖δ(0)‖ = 1.
The distinguished boundary of 𝒢δ, which we shall denote ℐδ, is
| (1.7) |
The reader should keep in mind the example of the non-commutative polydisk (1.5), in which case ℐδ is 𝒰[d], the set of d-tuples of unitary matrices in 𝕄[d], and ℬδ is the set of contractive d-tuples such that at least one has norm equal to one. For the column-ball (1.6), the distinguished boundary will agree with ℬδ when n = 1, but will be smaller when n > 1.
Definition 1.8
The Schur class of 𝒢δ, denoted 𝒮(𝒢δ), is the set of nc functions ϕ on 𝒢δ such that ‖ϕ(x)‖ < 1 ∀ x ∈ 𝒢δ.
A B-point for ϕ is a point in the boundary where a certain regularity occurs.
Definition 1.9
Let ϕ ∈ 𝒮(𝒢δ), and let T ∈ ℬδ. Then T is a B-point for ϕ if
| (1.10) |
for some sequence Zj converging to T.
Here is our non-commutative Julia inequality, which says that at a B-point, one has a boundary version of the Schwarz-Pick inequality, akin to (1.2).
Theorem 3.6
Suppose ϕ ∈ 𝒮(𝒢δ) and T ∈ ℬδ. If
then there exists W ∈ 𝒰n such that for all Z in 𝒢δ
We note that the noncommutative structure then gives that if T is a B-point, then so is U*TU for any unitary matrix U. Furthermore, given two B-points S and T, then S ⊕ T must also be a B-point.
Our second main result, Theorem 4.17, gives a characterization of when a point in ℐδ is a B-point, in terms of a realization of a δ nc-model for ϕ. We shall defer an exact statement until Section 4.
Our third result, Theorem 5.3, holds under the assumption that there are a lot of inward directions at T. For now, we shall just give a special case.
Theorem 1.11
Suppose 𝒢δ is either (1.5) or (1.6). Suppose T ∈ ℐδ is a B-point of ϕ. Then
exists for all H satisfying T + tH ∈ 𝒢δ for t small and positive. Moreover η is a holomorphic function of H, which is homogeneous of degree 1.
There is compatibility between the directional derivative at a B-point T, given by the function η in Theorem 1.11, and the directional derivative at a direct sum of m copies of T, which we will denote by T(m). In particular, if a scalar point is a B-point, then one obtains B-points at all levels.
Corollary 1.12
Suppose 𝒢δ is either (1.5) or (1.6). Suppose is a B-point of ϕ. Then
exists for all m, and for all H satisfying T(m) + tH ∈ 𝒢δ for t small and positive. Moreover η is a free function of H, which is homogeneous of degree 1.
Corollary 1.12 follows immediately from the definition of a free function, and the fact that T is a d-tuple of scalars. We leave the details to the interested reader, but emphasize that the noncommutative Julia-Carathéodory really captures a regularity that holds not only within each level , but also between levels.
In Example 7.1 we consider the function
which we show is in the Schur class of (1.4).
2 Background material
2.1 The one variable Julia-Carathéodory theorem
The Julia-Carathéodory Theorem, due to G. Julia [22] in 1920 and C. Carathéodory [16] in 1929, is the following.
Theorem 2.1
Let φ : 𝔻 → 𝔻 be a holomorphic function, and τ ∈ 𝕋. The following are equivalent:
.
The quotient has a non-tangential limit as z tends to τ.
- The function φ has both a non-tangential limit ω ∈ 𝕋 at τ and also an angular derivative η ∈ ℂ, that is the difference quotient
has a non-tangential limit η at τ. There exist ω in 𝕋 and η in ℂ so that at τ, φ(z) tends to ω non-tangentially and ϕ′(ζ) tends to η non-tangentially.
Furthermore, if (1.1) holds, then (1.2) does.
On the bidisk, the analogue of (B) does not imply (C); but it is proved in [8] that (C) implies (D). Moreover, it is shown that even when ϕ does not have a holomorphic differential pointing into the bidisk, the one sided derivative exists and is holomorphic in the direction.
2.2 Background on free holomorphic functions
A free holomorphic function is an nc function that is locally bounded with respect to the topology generated by all the sets 𝒢δ, as δ ranges over all matrices with entries that are free polynomials. These functions are studied in [3], and two principal results are obtained. One is that a bounded function on 𝒢δ is nc if and only if it is the pointwise limit of a sequence of non-commutative polynomials. The other is that every bounded nc-function on 𝒢δ has an nc δ-model. Alternative proofs of both these results have been found in [13] and [2]. Before explaining what this is, we need to slightly expand definition 1.3. Let ℰ1 and ℰ2 be Hilbert spaces, and let ℒ(ℰ1, ℰ2) denote the bounded linear operators from ℰ1 to ℰ2. Following [28], we shall write tensor products vertically to enhance readability and condense realization formulas, so represents the same object as A⊗B. We shall assume that the domain Ω of any nc function is closed w.r.t. direct sums. Also, we shall for notational convenience assume that δ is a square J-by-J matrix — we can always add rows or columns of zeroes to ensure this. We shall let Ωn denote , and ℐn denote the invertible matrices in 𝕄n.
Definition 2.2
An ℒ(ℰ1, ℰ2)-valued nc function F on a set Ω ⊆ 𝕄[d] is a function satisfying
Definition 2.3
Let δ be a J-by-J matrix of free polynomials. Let ϕ be an nc-function on 𝒢δ. A δ nc-model for ϕ is an -valued nc function u on that Ω satisfies
| (2.4) |
Theorem 2.5
[3] An nc function ϕ defined on 𝒢δ is bounded by 1 in norm if and only if it has a δ nc-model.
To help the reader, let us rewrite Theorem 2.5 in the case that 𝒢δ is the non-commutative d-polydisk, given by 1.5, and both ℰ1 and ℰ2 are one dimensional. Then J = d, and Theorem 2.5 becomes:
Theorem 2.6
Let Ω be the non-commutative d-polydisk, (1.5). The graded function ϕ is in the Schur class of Ω if and only if there are d ℒ(ℂ, ℰ)-valued nc functions u1, … ud so that, for all n, for all , we have
3 Julia’s Inequality and Consequences
We shall assume for the remainder of the paper that δ is a J-by-J matrix of non-commutative polynomials, that ϕ ∈ 𝒮(𝒢δ), and u is a δ nc-model for ϕ, with values in . We shall further assume that T ∈ ℬδ is in . We shall let 𝒰n denote the n-by-n unitaries.
If ϕ is in 𝒮(𝒢δ), it follows from Theorem 2.5 that
So if (1.10) holds, then {‖u(Zj)‖} is bounded, and the following definition is non-vacuous.
Definition 3.1
Let T be a B-point for ϕ. Let YT = YT (u) denote the set of all weak-limits of u(Zj), where Zj is a sequence in 𝒢δ that converges to T and has
| (3.2) |
bounded. We shall call YT the cluster set of the model u.
Proposition 3.3
Let T be a B-point for ϕ. Then there exists W ∈ 𝒰n such that for all υ ∈ YT, for all Z in 𝒢δ, we have
| (3.4) |
Moreover, if
| (3.5) |
holds for some sequence, then there exists υ ∈ YT with ‖υ‖2 ≤ α.
Proof
Suppose (1.10) holds and u(Zj) tends weakly to υ. Then ‖I − ϕ(Zj)*ϕ(Zj)‖ → 0, so by passing to a subsequence we can assume that ϕ(Zj) tends to some unitary W. Taking the limit in
we get (3.4). To see that W is unique, if another sequence tending to T with weakly convergent had , then letting in (3.4) we get
Since ϕ(Zj)′ and act on finite dimensional spaces, we can pass to a subsequence so that they converge in norm, and converges weakly. In the limit, we get
so W = W′.
For the latter part, note
so
Taking υ to be any weak cluster point of u(Zj), we get ‖υ‖2 ≤ α.
If T is a B-point for ϕ, we shall let ϕ(T) denote the matrix W that satisifies (3.4).
Here is the nc Julia inequality.
Theorem 3.6
Suppose T ∈ ℬδ, and
Then there exists W ∈ 𝒰n such that for all Z in 𝒢δ
| (3.7) |
Proof
By Proposition 3.3, we can choose υ in YT with ‖υ‖2 ≤ α. From (3.4) we have
so
| (3.8) |
Now
| (3.9) |
Combining (3.8) and (3.9), we get
which yields (3.7).
If T ∈ ℐδ, a non-tangential approach region is a region of the form
| (3.10) |
A corollary of Julia’s lemma is that a function’s behavior is controlled non-tangentially at a B-point on the distinguished boundary.
Proposition 3.11
If T ∈ ℐδ is a B-point for ϕ, then
is bounded on all sets that approach T non-tangentially.
Proof
Suppose υ and W are such that (3.4) holds:
Fix c, and let S be the non-tangential approach region
By (3.7), we have for Z in S that
| (3.12) |
Now
| (3.13) |
Squaring and using (3.12), we get
| (3.14) |
Since T ∈ ℐδ, we have δ(T) is an isometry, so
Therefore if Z ∈ S, the expression in parantheses on the right-hand side of (3.14) is bounded by c2(1 − ‖δ(Z)‖2), so we get
as required.
4 Models and B-points
In this section we shall study how being a B-point is related to properties of the δ nc model. In the case of the bidisk, these results are in [8].
Proposition 4.1
Let T ∈ ℐδ, and suppose Zj in 𝒢δ converges to T non-tangentially in the region (3.10). The following are equivalent:
‖I − ϕ(Zj)*ϕ(Zj)‖ ≤ M ‖I − δ(Zj)*δ(Zj)‖.
‖I − ϕ(Zj)*ϕ(Zj)‖ ≤ M′ (1 − ‖δ(Zj)‖2).
‖u(Zj)‖2 ≤ M″ for some u satisfying (2.4).
‖u(Zj)‖2 ≤ M″ for every u satisfying (2.4).
Proof
(i) ⇒ (iv)
Therefore
But
Therefore we get
(iii) ⇒ (i) Since
taking norms we get
(i) ⇔ (ii) In the region (3.10), we have
We can now give a different characterization of B-points that are on the distinguished boundary.
Corollary 4.2
Let T ∈ ℐδ, and let u be a δ nc-model for ϕ. Suppose: (NT) There exists some sequence in 𝒢δ that approaches T non-tangentially.
Then The following are equivalent:
T is a B-point of ϕ.
u(Zj) is bounded on some sequence Zj that approaches T non-tangentially.
u(Z) is bounded on every set that approaches T non-tangentially.
is bounded on every set that approaches T non-tangentially.
Proof
(i) ⇒ (iv) by Proposition 3.11, and (iii) ⇒ (ii) is trivial.
(iv) ⇒ (iii) and (ii) ⇒ (i) both follow from Proposition 4.1, and the observation that the proof shows that all the constants M, M′, M″ are comparable once the aperture of the non-tangential approach region is fixed.
Remark
Condition (NT) is very mild. It will hold if Γ(T) (see Def. 4.6 below) is non-empty.
If u is a δ nc model for ϕ, then by [3, Cor. 8.2], there is an isometry (which is called a realization of the model)
| (4.3) |
so that for ,
| (4.4) |
and
| (4.5) |
For T ∈ ℐδ, the inward directions for T are those H such that T + tH is inside 𝒢δ for t small and positive. Formally, if , and , let
denote the derivative of δ at T in the direction H. If A is a self-adjoint matrix, we write A < 0 to mean A is negative definite.
Definition 4.6
Let be in ℐδ. The inward set of T is the set
The transverse inward set of T is the subset of Γ(T) defined by
We have the following elementary result.
Lemma 4.7
Let H ∈ Γ(T). Then there exists ε > 0 such that
Moreover, T + tH approaches T non-tangentially as t ↓ 0.
Proof
Let V = δ(T), an isometry since T ∈ ℐδ. Then
so
| (4.8) |
This yields the first assertion, and the second follows from this and the mean value theorem, which implies that
For the rest of this section, we shall make the following assumption:
-
(A1)The set Δ(T) is non-empty, so there exists and β > 0 so that
(4.9)
Lemma 4.10
Let T ∈ ℐδ, and u be a δ nc model for ϕ. Suppose that T is a B-point for ϕ, and that K ∈ Δ(T) satisfies (4.9). Let Zj = T + tjK, where 0 < tj < 1 and tj → 0. If u(Zj) converges weakly to υ, then u(Zj) converges in norm to υ.
Proof
We have
| (4.11) |
| (4.12) |
Let V = δ(T) and X = ∇δ(T)[K]. Let Z = T + tK, so δ(Z) = V + tX + O(t2), and recall that V*V = I since T ∈ ℐδ. So the lower parts of the terms in parentheses on the right-hand sides of (4.11) and (4.12) are, respectively, −tV*X + O(t2) and −2tV*X + O(t2). From (4.12) we get
Take the real part, and subtract and add twice the real part of (4.11) to get
Therefore
| (4.13) |
As Zj → T within a non-tangential approach region, by Theorem 3.6 and Proposition 4.1 (iii) ⇒ (i), we get some constant M so that
Therefore the right-hand side of (4.13) is O(t), and we conclude that, since V*X ≤ − βI,
so
as desired.
Lemma 4.14
Under the assumptions of Lemma 4.10, there exists a unique uT such that
| (4.15) |
and
| (4.16) |
Proof
By Corollary 4.2, as t decreases to 0, the vectors u(T + tK) stay bounded; so there is some sequence tj so that u(T + tjK) converges weakly, Choose a sequence rj increasing to 1 so that u(rjT) converges weakly, and hence, by Lemma 4.10, also in norm, to a vector υ. Writing Zj = T + tjK,
so taking the limit we get
Since is in the range of , there exists a a unique vector uT satisfying (4.16) and (4.15).
We can now give a characterization of B-points, for homogeneous δ’s, in terms of realizations.
Theorem 4.17
Let ϕ ∈ 𝒮(𝒢δ), let u be a δ nc model for ϕ, and let be a realization as in (4.3 – 4.5). Let T ∈ ℑδ, and assume that (A1) holds. Then T is a B-point for ϕ if and only if
Proof
If T is a B-point, then the inclusion follows from Lemma 4.14. Conversely, suppose
for some vector υ. By (4.4), for any Z in 𝒢δ we have
Then
| (4.18) |
Now let Z approach T non-tangentially (such as along T + tK), and the right-hand side of (4.18) stays bounded; therefore T is a B-point by Corollary 4.2.
If δ is homogeneous, then the norm of the vector uT introduced in Lemma 4.14 is limr↑1 ‖u(rT)‖. We shall only prove it when δ is homogeneous of order 1, though the argument can be modified for any positive homogeneity.
Proposition 4.19
Assume that δ(rZ) = rδ(Z). Let T ∈ ℑδ be a B-point for ϕ. Then uT satisfies
| (4.20) |
| (4.21) |
Proof
Let rj be a sequence increasing to 1 so that u(rjT) converge weakly, and hence by Lemma 4.10 in norm, to some vector υ. By continuity,
so
As is a contraction, we have that
So each u(rjT) is perpendicular to ker and hence υ is also. Therefore υ is the vector uT from Lemma 4.14, and (4.20) holds.
To prove (4.21), we need to show that
| (4.22) |
Let α denote the left-hand side of (4.22). By Theorem 3.6, we have
As in (3.13), we have ‖I − ϕ(rT)*ϕ(rT)‖ ≤ 2‖W − ϕ(rT)‖. So we have
Dividing by
we get
So we have proved that (4.22) holds.
5 Derivatives at B-points
Let T be a B-point of ϕ in ℐδ, and W = ϕ(T). We will keep these fixed for the remainder of the section. Let us make the following assumption:
(A) The complex span of Δ(T) is all of .
This condition ensures that we have a full set of transverse directions pointing into Gδ.. Assumption (A) is equivalent to the following two conditions holding:
(A1) The set Δ(T) is non-empty.
- (A2) The complex span of
is all of .
Let H ∈ Γ(T); we want to show that
exists and is holomorphic in H.
First, let us sharpen Lemma 4.7.
Lemma 5.1
Let β > 0. Then there exists ε > 0 such that, if
| (5.2) |
then
Proof
This follows from (4.8), and the observation that the error term can be bounded by some absolute constant (which depends on δ and its derivatives in a neighborhood of T) times t2.
Let
Consider the set of functions
These functions are all defined on U, and are locally bounded by Corollary 4.2, and are holomorphic in both z and H. So they form a normal family by Montel’s theorem. Let S = (tn) be a sequence decreasing to 0 such that
exists; call this limit ηS(z, H). We wish to show that ηS does not, in fact, depend on the choice of sequence S.
Let K be in Δ(T). Multiplying K by a small positive number if necessary, we can assume that
Let υ be a unit vector in ℂn, and define the function
Then f : 𝔻(1, 1) → 𝔻, where 𝔻(1, 1) is the disk centered at 1 of radius 1. Moreover, 0 is a B-point for f, because, for t ∈ (0, 1),
and, letting M denote ‖∇δ(T)‖,
so
and the right-hand side is bounded since T is a B-point of ϕ.
So we can apply the one variable Julia-Carathéodory Theorem 2.1 to conclude that
exists — indeed the limit exists as 0 is approached non-tangentially from within 𝔻(1, 1). Since this holds for every unit vector υ, by polarization we can conclude that
exists, so every function ηS agrees on points of the form (t, K), and, by holomorphicity, on points in U of the form (z, K), whenever K ∈ Δ(T).
Now, fix some element K ∈ Δ(T) such that (2, K) ∈ U. Then, for some ε > 0, if H is in Σ(T) and ‖H‖ < ε, then K + H is in Δ(T). So all the ηS agree on points in U of the form (t, K + H), with t > 0 and H in Σ(T). By assumption (A2), since ηS is holomorphic in H, we get that in fact ηS is independent of the choice of S.
Thus we have proved:
Theorem 5.3
Suppose T ∈ ℐδ is a B-point of ϕ, and assumption (A) holds. Then
| (5.4) |
exists for all H ∈ Γ(T). Moreover η is a holomorphic function of H, homogeneous of degree 1.
6 Deducing the scalar case from the nc theorem
Knowing Theorem 3.6, how could we deduce the classical Julia inequality? We would need to know that any holomorphic function ψ : 𝔻 → 𝔻 could be extended to a function in the Schur class of {x ∈ 𝕄[1] : ‖x‖ < 1}. This indeed holds, by von Neumann’s inequality [36].
More generally, suppose Ω is a domain in ℂd, and ψ : Ω → 𝔻 is holomorphic. If we wish to deduce a Julia inequality using the results of the previous section, first we need to find a matrix γ of polynomials in d commuting variables so that Ω = {z ∈ ℂd : ‖γ(z)‖ < 1}. This of course may not be possible, though it is for the polydisk, or any polynomial polyhedron, and for the ball.
We can define Gγ to be the subset of 𝒢γ consisting of commuting d-tuples of matrices x = (x1, …, xd) for which ‖γ(x)‖ < 1. The original function ψ, which is holomorphic and bounded on , can be extended to all of Gγ, either by approximating ψ by polynomials, or using the Taylor functional calculus (see [4, 9] for a discussion). Let us define H∞(Gγ) to be those holomorphic functions ψ so that
is finite. (If Ω were the ball, we would get the multiplier algebra of the Drury-Arveson space). By [5], any function in H∞(Gγ) can be extended to a bounded function on 𝒢γ of the same norm. So we can deduce the following corollary of Theorem 3.6.
Corollary 6.1
Let be a domain in ℂd, and let ψ be a holomorphic function on Ω. Assume that ‖ψ‖H∞(Gγ) = 1. Suppose τ ∈ ∂Ω satisfies
Then there exists a complex number ω of modulus 1 so that
In the case of the ball, Corollary 6.1 is proved as Theorem 8.5.3 in [32], though with the weaker assumption that ψ is bounded by 1 in the sup-norm, not in the multiplier norm of the Drury-Arveson space. In [37], K. Wlodarczyk obtained a version for the unit ball of any J*-algebra, which includes the polydisk.
Theorem 5.3 also can be applied to the scalar case. Assumptions (A1) and (A2) can be checked in many concrete cases, such as the ball or the polydisk.
Corollary 6.2
Assume Ω, ψ and τ are as in Corollary 6.1, and that (A) holds at τ. Then ψ has a directional derivative in all inward directions at τ, and moreover this directional derivative is a holomorphic function of the direction.
Of course, if ψ were regular at τ, the directional derivative would be a linear function of the direction.
7 Examples
Example 7.1
Suppose
Then T ∈ ℐδ if and only if each Tr is an isometry. We have
and this is non-empty (e.g. take K = −T). The set Σ(T) is the set of H such that each Hr is Tr times a self-adjoint, so (A2) is also satisfied.
Let d = 2, and consider the scalar rational inner function
| (7.2) |
This has a B-point at (1, 1). By Andô’s theorem, f is of norm one on Gδ, so by [5], we can extend f to a function of norm one on 𝒢δ. It is not immediately obvious how to do so.
Claim
The function
| (7.3) |
is in 𝒮(𝒢δ) and agrees with f on commuting variables.
Let us temporarily accept the claim. For each n, the point (In, In) is a B-point, because
Theorem 3.6 then says that for ϕ as in (7.3), and Z a pair of contractions,
| (7.4) |
If Z1 = Z2, we get equality in (7.4).
If we calculate η(H) as in (5.4), we get that for all H with Re (H1) and Re (H2) negative definite, the derivative of ϕ at (In, In) in the direction H is
which is clearly holomorphic and homogeneous of degree 1.
Proof of claim
We shall write down a realization for f, and extend it to non-commutative variables.
Let
![]() |
(7.5) |
Let
and let
Since (7.5) is a unitary matrix, the function A + Bδ(z, w)u(z, w) is a rational inner function on 𝔻2, which by inspection agrees with f in (7.2). Now, we keep the same unitary, and using (4.3) – (4.5) (where ℰ is just ℂ, and J = 2), we get a formula for ϕ, which, after some algebra, becomes
| (7.6) |
Inverting the matrix on the right-hand side of (7.6) using the Boltz-Banachiewicz formula does not lead to a nice formula; but if one expands the inverse in a Neumann series, and observes that
then for k ≥ 1, we get
Then (7.6) becomes
and summing the Neumann series we get (7.3), as claimed.
Remark
The function
is another nc extension of f, but it is not in 𝒮(𝒢δ). Indeed, evaluating it on the pair of unitaries
we get
which has norm . So by continuity
Example 7.7
Let Ω be the classical Cartan domain of symmetric J-by-J contractive matrices in dimensions. There is an obvious embedding δ that takes d numbers and writes them as a J-by-J symmetric matrix, and we can extend this map to matrices, giving 𝒢δ and the commutative version Gδ. A point is in the distinguished boundary of 𝒢δ when δ(T) is a symmetric isometry. (A1) holds at every distinguished boundary point, and so does (A2), since Σ(T) is the set of H such that δ(H) can be written as the sum of d(T) times a self-adjoint matrix and (I − δ(T)*δ(T)) times anything. So if ϕ is in the Schur class of 𝒢δ or Gδ, we can apply both Theorem 3.6 and Theorem 5.3.
Footnotes
Partially supported by National Science Foundation Grant DMS 1565243
Partially supported by National Science Foundation Fellowship DMS 1606260
It is more common to consider the row contractions, but we choose column contractions so that what we call the distinguished boundary will be non-empty. It is easy to pass between these two sets, since the column contractions are just the adjoints of the row contractions.
References
- 1.Abate M. The Julia-Wolff-Carathéodory theorem in polydisks. J. Anal. Math. 1998;74:275–306. [Google Scholar]
- 2.Agler J, McCarthy JE. Global holomorphic functions in several non-commuting variables II. Canadian Math. Bull., To appear [Google Scholar]
- 3.Agler J, McCarthy JE. Global holomorphic functions in several non-commuting variables. Canad. J. Math. 2015;67(2):241–285. [Google Scholar]
- 4.Agler J, McCarthy JE. Operator theory and the Oka extension theorem. Hiroshima Math. J. 2015;45(1):9–34. [Google Scholar]
- 5.Agler J, McCarthy JE. Pick interpolation for free holomorphic functions. Amer. J. Math. 2015;137(6):1685–1701. [Google Scholar]
- 6.Agler J, McCarthy JE. Aspects of non-commutative function theory. Concrete Operators. 2016;3:15–24. [Google Scholar]
- 7.Agler J, McCarthy JE, Young NJ. Facial behavior of analytic functions on the bidisk. 2011;43:478–494. [Google Scholar]
- 8.Agler J, McCarthy JE, Young NJ. A Carathéodory theorem for the bidisk via Hilbert space methods. Math. Ann. 2012;352(3):581–624. [Google Scholar]
- 9.Agler J, McCarthy JE, Young NJ. On the representation of holomorphic functions on polyhedra. Michigan Math. J. 2013;62(4):675–689. [Google Scholar]
- 10.Agler J, Tully-Doyle R, Young NJ. Boundary behavior of analytic functions of two variables via generalized models. Indag. Math. (N.S.) 2012;23(4):995–1027. [Google Scholar]
- 11.Alpay D, Kalyuzhnyi-Verbovetzkii DS. Matrix-J-unitary non-commutative rational formal power series. The state space method generalizations and applications. 2006:49–113. [Google Scholar]
- 12.Balasubramanian Sriram. Toeplitz corona and the Douglas property for free functions. J. Math. Anal. Appl. 2015;428(1):1–11. [Google Scholar]
- 13.Ball JA, Marx G, Vinnikov V. Interpolation and transfer function realization for the non-commutative Schur-Agler class. To appear. arXiv:1602.00762 [Google Scholar]
- 14.Ball Joseph A, Groenewald Gilbert, Malakorn Tanit. Conservative structured noncommutative multidimensional linear systems. The state space method generalizations and applications. 2006:179–223. [Google Scholar]
- 15.Bickel Kelly, Knese Greg. Inner functions on the bidisk and associated Hilbert spaces. J. Funct. Anal. 2013;265(11):2753–2790. [Google Scholar]
- 16.Carathéodory C. Über die Winkelderivierten von beschränkten analytischen Funktionen. Sitzunber. Preuss. Akad. Wiss. 1929:39–52. [Google Scholar]
- 17.Cimpric Jakob, Helton J William, McCullough Scott, Nelson Christopher. A noncommutative real nullstellensatz corresponds to a noncommutative real ideal: algorithms. Proc. Lond. Math. Soc. (3) 2013;106(5):1060–1086. [Google Scholar]
- 18.Helton J William, Klep Igor, McCullough Scott. Proper analytic free maps. J. Funct. Anal. 2011;260(5):1476–1490. [Google Scholar]
- 19.Helton J William, McCullough Scott. Every convex free basic semi-algebraic set has an LMI representation. Ann. of Math. (2) 2012;176(2):979–1013. [Google Scholar]
- 20.Hervé Michel. Quelques propriétés des applications analytiques d’une boule à m dimensions dan elle-même. J. Math. Pures Appl. (9) 1963;42:117–147. [Google Scholar]
- 21.Jafari F. Angular derivatives in polydisks. Indian J. Math. 1993;35:197–212. [Google Scholar]
- 22.Julia G. Extension nouvelle d’un lemme de Schwarz. Acta Math. 1920;42:349–355. [Google Scholar]
- 23.Jury MT. An improved Julia-Caratheodory theorem for Schur-Agler mappings of the unit ball. arXiv:0707.3423 [Google Scholar]
- 24.Kaliuzhnyi-Verbovetskyi Dmitry S, Vinnikov Victor. Foundations of free noncommutative function theory. AMS, Providence. 2014 [Google Scholar]
- 25.McCarthy John E, Timoney R. Non-commutative automorphisms of bounded non-commutative domains. Proc. Roy. Soc. Edinburgh Sect. A. 2016;146(5):1037–1045. [Google Scholar]
- 26.Muhly Paul S, Solel Baruch. Tensorial function theory: from Berezin transforms to Taylor’s Taylor series and back. Integral Equations Operator Theory. 2013;76(4):463–508. [Google Scholar]
- 27.Pascoe JE. The inverse function theorem and the Jacobian conjecture for free analysis. Math. Z. 2014;278(3–4):987–994. [Google Scholar]
- 28.Pascoe JE, Tully-Doyle R. Free Pick functions: representations, asymptotic behavior and matrix monotonicity in several noncommuting variables. J. Funct. Anal. 2017;273(1):283–328. [Google Scholar]
- 29.Popescu G. Von Neumann inequality for (B(ℋ)n)1. Math. Scand. 1991;68:292– 304. [Google Scholar]
- 30.Popescu Gelu. Free holomorphic functions on the unit ball of B(ℋ)n. J. Funct. Anal. 2006;241(1):268–333. [Google Scholar]
- 31.Popescu Gelu. Free biholomorphic classification of noncommutative domains. Int. Math. Res. Not. IMRN. 2011;4:784–850. [Google Scholar]
- 32.Rudin W. Zeros of holomorphic functions in balls. Indag. Math. 1976;38:57–65. [Google Scholar]
- 33.Sarason D. University of Arkansas Lecture Notes. Wiley; New York: 1994. Sub-Hardy Hilbert spaces in the unit disk. [Google Scholar]
- 34.Taylor JL. A general framework for a multi-operator functional calculus. Advances in Math. 1972;9:183–252. [Google Scholar]
- 35.Taylor JL. Functions of several non-commuting variables. Bull. Amer. Math. Soc. 1973;79:1–34. [Google Scholar]
- 36.von Neumann J. Eine Spektraltheorie für allgemeine Operatoren eines unitären Raumes. Math. Nachr. 1951;4:258–281. [Google Scholar]
- 37.Wlodarczyk K. Julia’s lemma and Wolff’s theorem for J*-algebras. Proc. Amer. Math. Soc. 1987;99:472–476. [Google Scholar]

