This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Inner vectors for Toeplitz operators

Raymond Cheng Department of Mathematics and Statistics, Old Dominion University, Norfolk, VA 23529 rcheng@odu.edu Javad Mashreghi Département de mathématiques et de statistique, Université laval, Québec, QC, Canada, G1V 0A6 javad.mashreghi@mat.ulaval.ca  and  William T. Ross Department of Mathematics and Computer Science, University of Richmond, Richmond, VA 23173, USA wross@richmond.edu Dedicated to Thomas Ransford on the occasion of his sixtieth birthday.
Abstract.

In this paper we survey and bring together several approaches to obtaining inner functions for Toeplitz operators. These approaches include the classical definition, the Wold decomposition, the operator-valued Poisson Integral, and Clark measures. We then extend these notions somewhat to inner functions on model spaces. Along the way we present some novel examples.

This work was supported by NSERC (Canada).

1. Introduction

For φH\varphi\in H^{\infty}, the bounded analytic functions on the open unit disk 𝔻\mathbb{D}, let

(1.1) Tφ:H2H2,Tφf=φf,T_{\varphi}:H^{2}\to H^{2},\quad T_{\varphi}f=\varphi f,

denote the analytic Toeplitz operator on the classical Hardy space H2H^{2}. In this paper we survey, continue, and synthesize some discussions begun in [4, 10, 11] dealing with the notion of an “inner vector” for TφT_{\varphi} along with the general notion of an inner vector for a contraction on a Hilbert space. We connect these results with the operator-valued Poisson kernel and some work from [2, 3] concerning “factoring an L1L^{1} function through a contraction”. Along the way we also produce some interesting examples and reformulations of these connections.

2. Basic definitions and facts

We begin with the definition of an inner vector for a Toeplitz operator from [10]. Recall that the inner product on the Hardy space H2H^{2} is

(2.1) f,g:=𝕋fg¯𝑑m,\langle f,g\rangle:=\int_{\mathbb{T}}f\overline{g}\,dm,

where mm is normalized Lebesgue measure on the unit circle 𝕋\mathbb{T}. As is tradition, we equate an fH2f\in H^{2} with its L2=L2(𝕋,m)L^{2}=L^{2}(\mathbb{T},m) radial boundary function, i.e.,

f(ζ)=limr1f(rζ)f(\zeta)=\lim_{r\to 1^{-}}f(r\zeta)

for almost every ζ𝕋\zeta\in\mathbb{T}. We will also use the term inner function (without any qualifiers like in Definition 2.2 below) to denote an HH^{\infty} function that has unimodular boundary values almost everywhere. Classical theory [6] says that an inner function II can be factored uniquely as I=ξBSμI=\xi BS_{\mu}, where ξ\xi is a unimodular constant, BB is a Blaschke product, and SμS_{\mu} is a singular inner function associated with a positive measure μ\mu on 𝕋\mathbb{T} that is singular with respect to mm. We say the degree of II is equal to dd if II is a finite Blaschke product of order dd, and equal to infinity otherwise. Furthermore, any function fH2f\in H^{2} can be factored, uniquely up to multiplicative constants, as f=IGf=IG, where II is an inner function and GH2G\in H^{2} is an outer function.

For φH\varphi\in H^{\infty} the analytic Toeplitz operator TφT_{\varphi} from (1.1) is a bounded operator on H2H^{2} whose norm Tφ\|T_{\varphi}\| satisfies

Tφ=φ:=esssup{|φ(ξ)|:ξ𝕋}.\|T_{\varphi}\|=\|\varphi\|_{\infty}:=\operatorname{ess-sup\{}|\varphi(\xi)|:\xi\in\mathbb{T}\}.

Also recall that the adjoint TφT_{\varphi}^{*} of TφT_{\varphi} satisfies Tφ=Tφ¯T_{\varphi}^{*}=T_{\overline{\varphi}}, where Tφ¯f=P(φ¯f)T_{\overline{\varphi}}f=P(\overline{\varphi}f) and PP is the Riesz projection of L2L^{2} onto H2H^{2}. When φ\varphi is an inner function, observe from (2.1) that TφT_{\varphi} is an isometry. See [8, Ch. 4] for the details of these basic Toeplitz operator facts and [1] for a definitive treatise.

Definition 2.2.

For φH\varphi\in H^{\infty} we say a unit vector fH2f\in H^{2} is TφT_{\varphi}-inner if Tφnf,f=0\langle T^{n}_{\varphi}f,f\rangle=0 for all n1n\geqslant 1.

When φ(z)=z\varphi(z)=z, one can see from Fourier analysis that the TzT_{z}-inner vectors are precisely the inner functions. Also observe that replacing φ\varphi with cφc\varphi, where c>0c>0, in Definition 2.2 does not change whether or not a function ff is TφT_{\varphi}-inner. Thus we can always assume, by scaling φ\varphi, that

φb(H):={gH:g1},\varphi\in b(H^{\infty}):=\{g\in H^{\infty}:\|g\|_{\infty}\leqslant 1\},

the closed unit ball of HH^{\infty}. This normalization will be important when we need TφT_{\varphi} to be a contraction operator since in this case Tφ=φ1\|T_{\varphi}\|=\|\varphi\|_{\infty}\leqslant 1. Immediate from Definition 2.2 and the inner product formula from (2.1) are the following facts.

Proposition 2.3.

Let φb(H)\varphi\in b(H^{\infty}).

  1. (1)

    If fH2f\in H^{2} is TφT_{\varphi}-inner and II is any inner function, then IfIf is TφT_{\varphi}-inner.

  2. (2)

    If fH2f\in H^{2} is TφT_{\varphi}-inner and Θ\Theta is any inner divisor of ff, i.e., f/ΘH2f/\Theta\in H^{2}, then f/Θf/\Theta is TφT_{\varphi}-inner.

  3. (3)

    Any unit vector belonging to kerTφ¯\ker T_{\overline{\varphi}} is TφT_{\varphi}-inner.

If uu denotes the inner factor of φ\varphi, it is known [8, p. 108] that

kerTφ¯=𝒦u:=(uH2),\ker T_{\overline{\varphi}}=\mathcal{K}_{u}:=(uH^{2})^{\perp},

the model space corresponding to uu. Thus we have the simple corollary.

Corollary 2.4.

If II is any inner function and uu is the inner factor of φb(H)\varphi\in b(H^{\infty}), then any unit vector from I𝒦uI\mathcal{K}_{u} is TφT_{\varphi}-inner.

This corollary gives us many specific examples of TφT_{\varphi}-inner vectors. For example, if λ𝔻\lambda\in\mathbb{D}, the reproducing kernel functions

kλ(z):=1u(λ)¯u(z)1λ¯zk_{\lambda}(z):=\frac{1-\overline{u(\lambda)}u(z)}{1-\overline{\lambda}z}

belong to 𝒦u\mathcal{K}_{u}. In fact, finite linear combinations of these functions are dense in 𝒦u\mathcal{K}_{u} [8, Ch. 5]. Since

kλ=kλ(λ)=1|u(λ)|21|λ|2,\|k_{\lambda}\|=\sqrt{k_{\lambda}(\lambda)}=\sqrt{\frac{1-|u(\lambda)|^{2}}{1-|\lambda|^{2}}},

then

I1|λ|21|u(λ)|21u(λ)¯u(z)1λ¯z,λ𝔻,I inner,I\sqrt{\frac{1-|\lambda|^{2}}{1-|u(\lambda)|^{2}}}\frac{1-\overline{u(\lambda)}u(z)}{1-\overline{\lambda}z},\quad\lambda\in\mathbb{D},\;\;\mbox{$I$ inner},

are TφT_{\varphi}-inner functions.

When φ=u\varphi=u is a finite Blaschke product, then the model space 𝒦u\mathcal{K}_{u} is a certain finite dimensional space of rational functions that are analytic in a neighborhood of 𝔻¯\overline{\mathbb{D}} [8, p. 117]. Furthermore, as we will see in a moment in Theorem 3.12, every TuT_{u}-inner function is bounded. However, when uu is not a finite Blaschke product then 𝒦u\mathcal{K}_{u} is infinite dimensional [8, p. 117] and, since multiplication by an inner function II is an isometry on H2H^{2} (see (2.1)), I𝒦uI\mathcal{K}_{u} is a closed infinite dimensional subspace of L2L^{2}. By a theorem of Grothendieck, it will contain an unbounded function. Putting this all together, we obtain the following.

Corollary 2.5.

If the inner factor of φb(H)\varphi\in b(H^{\infty}) is not a finite Blaschke product, then there are unbounded TφT_{\varphi}-inner functions.

A specific version of this was pointed out in [10, p. 103].

Of course one needs to discuss the case when φ\varphi is an outer function. Since φH2\varphi H^{2} is dense in H2H^{2} [8, p. 86], we see that kerTφ¯={0}\ker T_{\overline{\varphi}}=\{0\}. In this case, it is not clear that there are any TφT_{\varphi}-inner functions. Indeed, we do not see any obvious ones like IkerTφ¯I\ker T_{\overline{\varphi}} since, in this case, kerTφ¯={0}\ker T_{\overline{\varphi}}=\{0\}.

Example 2.6.

Suppose that φ\varphi is the outer function φ(z)=1+z\varphi(z)=1+z and that fH2f\in H^{2} is TφT_{\varphi}-inner, i.e.,

Tφnf,f=0,n1.\langle T_{\varphi}^{n}f,f\rangle=0,\ \,\forall n\geqslant 1.

In other words,

(2.7) 𝕋(1+ξ)n|f(ξ)|2𝑑m(ξ)=0,n1.\int_{\mathbb{T}}(1+\xi)^{n}|f(\xi)|^{2}\,dm(\xi)=0,\ \,\forall n\geqslant 1.

Then the L1L^{1} function |f|2|f|^{2} annihilates (1+z)n(1+z)^{n} for all n1n\geqslant 1, along with all their linear combinations. In particular, |f|2|f|^{2} annihilates

(1+z)2(1+z)=1+2z+z21z=z(1+z).(1+z)^{2}-(1+z)=1+2z+z^{2}-1-z=z(1+z).

The above observation will be the first step in a proof by induction. Next, suppose that |f|2|f|^{2} annihilates zk(1+z)z^{k}(1+z) for all 1kn1\leqslant k\leqslant n. Then

zn+1(1+z)\displaystyle z^{n+1}(1+z) =(1+z)n+2[(1+z)n+1zn+1](1+z).\displaystyle=(1+z)^{n+2}-\big{[}(1+z)^{n+1}-z^{n+1}\big{]}(1+z).

By the TφT_{\varphi}-inner property of ff notice that |f|2|f|^{2} annihilates the first term on the right. It also annihilates the subtracted expression, by the induction hypothesis (the expression in square brackets is a polynomial of degree nn). Thus we have shown by induction that |f|2|f|^{2} annihilates {zn(1+z)}n0\{z^{n}(1+z)\}_{n\geqslant 0} (the n=0n=0 case follows from (2.7)). This means that

(2.8) 𝕋ξn(1+ξ)|f(ξ)|2𝑑m(ξ)=0,n0,\int_{\mathbb{T}}\xi^{n}(1+\xi)|f(\xi)|^{2}dm(\xi)=0,\quad n\geqslant 0,

and by complex conjugation,

𝕋ξ¯n(1+ξ¯)|f(ξ)|2𝑑m(ξ)=0,n0.\int_{\mathbb{T}}\overline{\xi}^{n}(1+\overline{\xi})|f(\xi)|^{2}dm(\xi)=0,\quad n\geqslant 0.

A little algebra yields

(2.9) 𝕋ξ¯n+1(1+ξ)|f(ξ)|2𝑑m(ξ),n0.\int_{\mathbb{T}}\overline{\xi}^{n+1}(1+\xi)|f(\xi)|^{2}dm(\xi),\quad n\geqslant 0.

Equations (2.8) and (2.9) say that all of the Fourier coefficients of (1+ξ)|f(ξ)|2(1+\xi)|f(\xi)|^{2} vanish and so (1+ξ)|f(ξ)|2(1+\xi)|f(\xi)|^{2} is zero. Conclusion: there are no TφT_{\varphi}-inner functions when φ(z)=1+z\varphi(z)=1+z.

3. Inner vectors via the Wold decomposition

Using some ideas from [10], we can use the Wold decomposition [9] to explore the inner vectors for certain Toeplitz operators. Observe that when uu is an inner function the Toeplitz operator TuT_{u} is an isometry on H2H^{2}. Thus the Wold decomposition of H2H^{2} with respect to TuT_{u} becomes

H2=X0X1TuX1Tu2X1,H^{2}=X_{0}\oplus X_{1}\oplus T_{u}X_{1}\oplus T_{u}^{2}X_{1}\oplus\cdots,

where

X0:=n=1TunH2={0},X1:=H2TuH2=𝒦u.X_{0}:=\bigcap_{n=1}^{\infty}T_{u}^{n}H^{2}=\{0\},\quad X_{1}:=H^{2}\ominus T_{u}H^{2}=\mathcal{K}_{u}.

Thus

H2=𝒦uu𝒦uu2𝒦u.H^{2}=\mathcal{K}_{u}\oplus u\mathcal{K}_{u}\oplus u^{2}\mathcal{K}_{u}\oplus\cdots.

The above decomposition says that every fH2f\in H^{2} has a unique expansion as

(3.1) f=F0+uF1+u2F2+,Fj𝒦u.f=F_{0}+uF_{1}+u^{2}F_{2}+\cdots,\quad F_{j}\in\mathcal{K}_{u}.

Furthermore, for each integer N1N\geqslant 1,

uNf,f\displaystyle\langle u^{N}f,f\rangle =uNk0ukFk,l0ulFl\displaystyle=\Big{\langle}u^{N}\sum_{k\geqslant 0}u^{k}F_{k},\sum_{l\geqslant 0}u^{l}F_{l}\Big{\rangle}
=k,l0uN+klFk,Fl\displaystyle=\sum_{k,l\geqslant 0}\langle u^{N+k-l}F_{k},F_{l}\rangle
=lk=NFk,Fl.\displaystyle=\sum_{l-k=N}\langle F_{k},F_{l}\rangle.

This leads us to the following.

Proposition 3.2.

A unit vector fH2f\in H^{2} with expansion

f=F0+uF1+u2F2+,Fj𝒦u,f=F_{0}+uF_{1}+u^{2}F_{2}+\cdots,\quad F_{j}\in\mathcal{K}_{u},

as in (3.1) is TuT_{u}-inner if and only if

(3.3) k=0Fk,FN+k=0,N1.\sum_{k=0}^{\infty}\langle F_{k},F_{N+k}\rangle=0,\quad N\geqslant 1.

Though this is just a restatement of the condition for ff to be TuT_{u}-inner, it is useful for producing more tangible examples of TuT_{u}-inner functions.

Example 3.4.

Choose orthogonal vectors Fj,j0F_{j},j\geqslant 0 from 𝒦u\mathcal{K}_{u} so that j0Fj2=1\sum_{j\geqslant 0}\|F_{j}\|^{2}=1. Then the condition (3.3) is easily satisfied and thus the unit vector f=j0ujFjf=\sum_{j\geqslant 0}u^{j}F_{j} is a TuT_{u}-inner function (as is any inner function times this vector).

Example 3.5.

If u(z)=znu(z)=z^{n}, then 𝒦u=span{1,z,z2,zn1}\mathcal{K}_{u}=\operatorname{span}\{1,z,z^{2},\ldots z^{n-1}\} and the vectors

Fj=zjn,0jn1,F_{j}=\frac{z^{j}}{\sqrt{n}},\quad 0\leqslant j\leqslant n-1,

satisfy the conditions of the previous example. Thus

f=j=0n1ujFj=1n+zn+1n+z2n+2n+z3n+3n++z(n1)(n+1)nf=\sum_{j=0}^{n-1}u^{j}F_{j}=\frac{1}{\sqrt{n}}+\frac{z^{n+1}}{\sqrt{n}}+\frac{z^{2n+2}}{\sqrt{n}}+\frac{z^{3n+3}}{\sqrt{n}}+\cdots+\frac{z^{(n-1)(n+1)}}{\sqrt{n}}

is a TznT_{z^{n}}-inner vector.

Example 3.6.

The previous example can be generalized to a finite Blaschke product

u(z)=j=1nzaj1aj¯z,aj𝔻.u(z)=\prod_{j=1}^{n}\frac{z-a_{j}}{1-\overline{a_{j}}z},\quad a_{j}\in\mathbb{D}.

If we define

F0(z)=1|a1|21a1¯z,F_{0}(z)=\frac{\sqrt{1-|a_{1}|^{2}}}{1-\overline{a_{1}}z},
F1(z)=1|a2|21a2¯zza11a1¯z,F_{1}(z)=\frac{\sqrt{1-|a_{2}|^{2}}}{1-\overline{a_{2}}z}\frac{z-a_{1}}{1-\overline{a_{1}}z},
F2(z)=1|a3|21a3¯zza11a1¯zza21a2¯z,F_{2}(z)=\frac{\sqrt{1-|a_{3}|^{2}}}{1-\overline{a_{3}}z}\frac{z-a_{1}}{1-\overline{a_{1}}z}\frac{z-a_{2}}{1-\overline{a_{2}}z},
\vdots
Fn1(z)=1|an|21an¯zj=1n1zaj1aj¯z,F_{n-1}(z)=\frac{\sqrt{1-|a_{n}|^{2}}}{1-\overline{a_{n}}z}\prod_{j=1}^{n-1}\frac{z-a_{j}}{1-\overline{a_{j}}z},

one can show that {F0,,Fn1}\{F_{0},\ldots,F_{n-1}\} is an orthonormal basis for 𝒦u\mathcal{K}_{u}. Now choose α0,,αn1\alpha_{0},\ldots,\alpha_{n-1}\in\mathbb{C} such that j=0n=1|αj|2=1\sum_{j=0}^{n=1}|\alpha_{j}|^{2}=1. Then

f=j=0n1αjujFjf=\sum_{j=0}^{n-1}\alpha_{j}u^{j}F_{j}

is TuT_{u}-inner.

From Corollary 2.4 we know, for an inner function II, that any unit vector from the set {IkerTu¯:I is inner}\{I\ker T_{\overline{u}}:\mbox{$I$ is inner}\} is a Tu¯T_{\overline{u}}-inner vector. Perhaps one might think we have equality here. Indeed, sometimes we do. For example, if u(z)=zu(z)=z, then kerTz¯=\ker T_{\overline{z}}=\mathbb{C} and, as discussed earlier, the TzT_{z}-inner vectors are precisely the inner functions. Here is another positive example of when the unit vectors from {IkerTu¯:I is inner}\{I\ker T_{\overline{u}}:\mbox{$I$ is inner}\} constitute the complete set of TuT_{u}-inner vectors.

Example 3.7.

If the inner function uu is the single Blaschke factor

u(z)=za1a¯z,a𝔻,u(z)=\frac{z-a}{1-\overline{a}z},\quad a\in\mathbb{D},

one can show [8, Ch. 5] that

kerTu¯=𝒦u=11a¯z.\ker T_{\overline{u}}=\mathcal{K}_{u}=\mathbb{C}\frac{1}{1-\overline{a}z}.

As shown in [4], the TuT_{u}-inner vectors are

I1|a|21a¯z,I inner.I\frac{\sqrt{1-|a|^{2}}}{1-\overline{a}z},\quad\mbox{$I$ inner}.

However, in general, the unit vectors from {IkerTu¯:I is inner}\{I\ker T_{\overline{u}}:\mbox{$I$ is inner}\} form a proper subset of the TuT_{u}-inner vectors. One can see this with the following example.

Example 3.8.

Using the technique from Example 3.5, we see that when u(z)=znu(z)=z^{n} the vector

f=12+zn+12f=\frac{1}{\sqrt{2}}+\frac{z^{n+1}}{\sqrt{2}}

is TuT_{u}-inner. However, ff is not of the form IgIg, where II is inner and g𝒦ug\in\mathcal{K}_{u}. This follows from the fact that ff is outer and does not belong to 𝒦u=span{1,z,z2,,zn1}\mathcal{K}_{u}=\operatorname{span}\{1,z,z^{2},\ldots,z^{n-1}\}.

The papers [10, 11] yield a description of the TuT_{u}-inner vectors. From the Wold decomposition (3.1) we see that any fH2f\in H^{2} can be written as

f=k=0Fkuk.f=\sum_{k=0}^{\infty}F_{k}u^{k}.

If {vj}j1\{v_{j}\}_{j\geqslant 1} is an orthonormal basis for 𝒦u\mathcal{K}_{u}, then we can expand things a bit further and write

f\displaystyle f =k=0Fkuk\displaystyle=\sum_{k=0}^{\infty}F_{k}u^{k}
=k=0uk(j1cj,kvj)\displaystyle=\sum_{k=0}^{\infty}u^{k}\Big{(}\sum_{j\geqslant 1}c_{j,k}v_{j}\Big{)}
=j1vj(k=0cj,kuk).\displaystyle=\sum_{j\geqslant 1}v_{j}\Big{(}\sum_{k=0}^{\infty}c_{j,k}u^{k}\Big{)}.

Observe that

j1|cj,k|2=Fk2\sum_{j\geqslant 1}|c_{j,k}|^{2}=\|F_{k}\|^{2}

and that

f2\displaystyle\|f\|^{2} =k=0Fk2\displaystyle=\sum_{k=0}^{\infty}\|F_{k}\|^{2}
=k=0j1|cj,k|2\displaystyle=\sum_{k=0}^{\infty}\sum_{j\geqslant 1}|c_{j,k}|^{2}
=j1k=0|cj,k|2.\displaystyle=\sum_{j\geqslant 1}\sum_{k=0}^{\infty}|c_{j,k}|^{2}.

Thus for each jj, k0|cj,k|2<\sum_{k\geqslant 0}|c_{j,k}|^{2}<\infty and so

fj(z)=k=0cj,kzkf_{j}(z)=\sum_{k=0}^{\infty}c_{j,k}z^{k}

defines a function in H2H^{2} (square summable power series). By the Littlewood subordination principle [8, p. 126], fjuf_{j}\circ u also belongs to H2H^{2}.

Thus every unit vector fH2f\in H^{2} has the unique representation

(3.9) f(z)=j1vj(z)fj(u(z)),f(z)=\sum_{j\geqslant 1}v_{j}(z)f_{j}(u(z)),

where fjH2f_{j}\in H^{2} with j1fj2<\sum_{j\geqslant 1}\|f_{j}\|^{2}<\infty, and {vj}j1\{v_{j}\}_{j\geqslant 1} is an orthonormal basis for 𝒦u\mathcal{K}_{u}. Furthermore, as observed in [10, Prop. 1] (and can be proved using the above calculation), if

(3.10) f=j1vjfj(u),g=j1vjgj(u),f=\sum_{j\geqslant 1}v_{j}f_{j}(u),\quad g=\sum_{j\geqslant 1}v_{j}g_{j}(u),

as in (3.9), then

(3.11) f,g=j1fj,gj.\langle f,g\rangle=\sum_{j\geqslant 1}\langle f_{j},g_{j}\rangle.
Theorem 3.12.

A unit vector ff written as in (3.9) is TuT_{u}-inner if and only if

j1|fj(ξ)|2=1\sum_{j\geqslant 1}^{\infty}|f_{j}(\xi)|^{2}=1

for almost every ξ𝕋\xi\in\mathbb{T}.

Proof.

Here is the original proof from [10]. With

f=j1vjfj(u),f=\sum_{j\geqslant 1}v_{j}f_{j}(u),

and n1n\geqslant 1, (3.11) yields

Tunf,f\displaystyle\langle T_{u}^{n}f,f\rangle =fun,f\displaystyle=\langle fu^{n},f\rangle
=jvjunfj(u),kvkfk(u)\displaystyle=\big{\langle}\sum_{j}v_{j}u^{n}f_{j}(u),\sum_{k}v_{k}f_{k}(u)\big{\rangle}
=j1znfj,fj\displaystyle=\sum_{j\geqslant 1}\langle z^{n}f_{j},f_{j}\rangle
=j1𝕋ξn|fj(ξ)|2𝑑m(ξ)\displaystyle=\sum_{j\geqslant 1}\int_{\mathbb{T}}\xi^{n}|f_{j}(\xi)|^{2}dm(\xi)
(3.13) =𝕋ξn(j1|fj(ξ)|2)𝑑m(ξ).\displaystyle=\int_{\mathbb{T}}\xi^{n}\Big{(}\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}\Big{)}dm(\xi).

Then Tunf,f=0\langle T_{u}^{n}f,f\rangle=0 for all n=1,2,n=1,2,\ldots if and only if, by Fourier analysis, j1|fj|2\sum_{j\geqslant 1}|f_{j}|^{2} is constant almost everywhere. But since we assuming that ff is a unit vector, we see, by putting n=0n=0 in (3.13), that j1|fj|2=1\sum_{j\geqslant 1}|f_{j}|^{2}=1 almost everywhere. ∎

When uu is a finite Blaschke product, then 𝒦u\mathcal{K}_{u} is finite dimensional. In this case (3.9) is finite and each basis vector vjv_{j} is a rational function that is analytic in a neighborhood of 𝔻¯\overline{\mathbb{D}} [8, Ch. 5]. From here it follows that every TuT_{u}-inner vector is a bounded function. Contrast this with Corollary 2.5 which says that when uu is not a finite Blaschke product there are always TuT_{u}-inner vectors that are unbounded functions.

The two papers [10, 11] go further and discuss an “inner-outer” factorization of any fH2f\in H^{2} in terms of TuT_{u}-inner and TuT_{u}-outer vectors. They also discuss the concept of TuT_{u}-inner in HpH^{p}, for p>1p>1, along with some properties of the norms of TuT_{u}-inner vectors as well as their growth near 𝕋\mathbb{T}.

4. Inner vectors via the operator-valued Poisson kernel

We can rephrase the language of inner vectors for Toeplitz operators in terms of operator-valued Poisson kernels [2]. Moreover, using this new language, we can extend our discussion to inner vectors for contractions on Hilbert spaces. For λ𝔻\lambda\in\mathbb{D} and ξ𝕋\xi\in\mathbb{T}, define

(4.1) Pλ(ξ):=11λ¯ξ+11λξ¯1P_{\lambda}(\xi):=\frac{1}{1-\overline{\lambda}\xi}+\frac{1}{1-\lambda\overline{\xi}}-1

and observe that this can be written as

Pλ(ξ)=1|λ|2|ξλ|2,P_{\lambda}(\xi)=\frac{1-|\lambda|^{2}}{|\xi-\lambda|^{2}},

which is the standard Poisson kernel. Classical theory says that for any gL1=L1(𝕋,m)g\in L^{1}=L^{1}(\mathbb{T},m) the function

𝕋Pλ(ξ)f(ξ)𝑑m(ξ)\int_{\mathbb{T}}P_{\lambda}(\xi)f(\xi)dm(\xi)

is harmonic on 𝔻\mathbb{D} with

(4.2) limr1𝕋Prζ(ξ)f(ξ)𝑑m(ξ)=f(ζ)\lim_{r\to 1^{-}}\int_{\mathbb{T}}P_{r\zeta}(\xi)f(\xi)dm(\xi)=f(\zeta)

for almost every ζ𝕋\zeta\in\mathbb{T}. Furthermore, if μ\mu is a finite complex measure on 𝕋\mathbb{T}, we have

(4.3) 𝕋Pλ(ξ)𝑑μ(ξ)=μ^(0)+n1μ^(n)λn+n1μ^(n)λ¯n,\int_{\mathbb{T}}P_{\lambda}(\xi)d\mu(\xi)=\widehat{\mu}(0)+\sum_{n\geqslant 1}\widehat{\mu}(n)\lambda^{n}+\sum_{n\geqslant 1}\widehat{\mu}(-n)\overline{\lambda}^{n},

where

μ^(n):=𝕋ξ¯n𝑑μ(ξ),n,\widehat{\mu}(n):=\int_{\mathbb{T}}\overline{\xi}^{n}d\mu(\xi),\quad n\in\mathbb{Z},

are the Fourier coefficients of μ\mu. We will now discuss an operator version of the Poisson kernel.

For a contraction TT on a Hilbert space \mathcal{H}, we imitate the formula in (4.1) and define, for λ𝔻\lambda\in\mathbb{D}, the operator-valued Poisson kernel Kλ(T)K_{\lambda}(T) as

Kλ(T):=(IλT)1+(Iλ¯T)1I.K_{\lambda}(T):=(I-\lambda T^{*})^{-1}+(I-\overline{\lambda}T)^{-1}-I.

By the spectral radius formula, notice how σ(T)𝔻¯\sigma(T)\subseteq\overline{\mathbb{D}} and thus the formula for Kλ(T)K_{\lambda}(T) above makes sense. A computation with Neumann series will show that for r[0,1)r\in[0,1) and θ[0,2π)\theta\in[0,2\pi)

(4.4) Kreiθ(T)=n=0rneinθTn+n=0rneinθTnI.K_{re^{i\theta}}(T)=\sum_{n=0}^{\infty}r^{n}e^{in\theta}T^{*n}+\sum_{n=0}^{\infty}r^{n}e^{-in\theta}T^{n}-I.

The operator identity

Kλ(T)=(Iλ¯T)1(I|λ|2TT)(IλT)1K_{\lambda}(T)=(I-\overline{\lambda}T)^{-1}(I-|\lambda|^{2}TT^{*})(I-\lambda T^{*})^{-1}

from [2, Lemma 2.4] shows that for each 𝐱\mathbf{x}\in\mathcal{H}

Kλ(T)𝐱,𝐱0,λ𝔻.\langle K_{\lambda}(T)\mathbf{x},\mathbf{x}\rangle\geqslant 0,\quad\lambda\in\mathbb{D}.

Moreover, the function

λKλ(T)𝐱,𝐱\lambda\mapsto\langle K_{\lambda}(T)\mathbf{x},\mathbf{x}\rangle

is harmonic on 𝔻\mathbb{D}. Hence, a classical harmonic analysis result of Herglotz ([6, p. 10] or [8, p. 17]) produces a unique positive finite Borel measure μT,𝐱\mu_{T,\mathbf{x}} on 𝕋\mathbb{T} such that

(4.5) Kλ(T)𝐱,𝐱=𝕋Pλ(ζ)𝑑μT,𝐱(ζ).\langle K_{\lambda}(T)\mathbf{x},\mathbf{x}\rangle=\int_{\mathbb{T}}P_{\lambda}(\zeta)d\mu_{T,\mathbf{x}}(\zeta).

Since K0(T)=IK_{0}(T)=I we have

1=𝐱,𝐱=K0(T)𝐱,𝐱=𝕋𝑑μT,𝐱1=\langle\mathbf{x},\mathbf{x}\rangle=\langle K_{0}(T)\mathbf{x},\mathbf{x}\rangle=\int_{\mathbb{T}}d\mu_{T,\mathbf{x}}

and so μT,𝐱\mu_{T,\mathbf{x}} is a probability measure.

As we defined for Toeplitz operators earlier in Definition 2.2, we say that a unit vector 𝐱\mathbf{x} is TT-inner if

Tn𝐱,𝐱=0,n1.\langle T^{n}\mathbf{x},\mathbf{x}\rangle=0,\quad n\geqslant 1.

Note that 𝐱\mathbf{x} is TT-inner if and only if 𝐱\mathbf{x} is TT^{*}-inner. From (4.4) we see that 𝐱\mathbf{x} is TT-inner if and only if Kλ(T)𝐱,𝐱=1\langle K_{\lambda}(T)\mathbf{x},\mathbf{x}\rangle=1 for all λ𝔻\lambda\in\mathbb{D}, or equivalently,

1=𝕋Pλ(ζ)𝑑μT,𝐱(ζ),λ𝔻.1=\int_{\mathbb{T}}P_{\lambda}(\zeta)d\mu_{T,\mathbf{x}}(\zeta),\quad\lambda\in\mathbb{D}.

By (4.3) this is equivalent to the condition μT,𝐱=m\mu_{T,\mathbf{x}}=m. This gives us the following.

Proposition 4.6.

Suppose that TT is a contraction on a Hilbert space \mathcal{H} and 𝐱\mathbf{x} is unit vector in \mathcal{H}. Then 𝐱\mathbf{x} is TT-inner if and only if μT,𝐱=m\mu_{T,\mathbf{x}}=m, where μT,𝐱\mu_{T,\mathbf{x}} is defined as in (4.5).

For an inner function uu, note that TuT_{u} is an isometry, hence a contraction. Thus we can apply the above analysis to μTu,f\mu_{T_{u},f}.

Proposition 4.7.

If

f=j1vjfj(u)f=\sum_{j\geqslant 1}v_{j}f_{j}(u)

is a vector from H2H^{2} as in (3.9), then

(4.8) dμTu,f=j1|fj|2dm.d\mu_{T_{u},f}=\sum_{j\geqslant 1}|f_{j}|^{2}\;dm.
Proof.

If

f=j1vjfj(u),f=\sum_{j\geqslant 1}v_{j}f_{j}(u),

then

f2=j1fj2=j1𝕋|fj|2𝑑m=𝕋j1|fj|2dm\|f\|^{2}=\sum_{j\geqslant 1}\|f_{j}\|^{2}=\sum_{j\geqslant 1}\int_{\mathbb{T}}|f_{j}|^{2}dm=\int_{\mathbb{T}}\sum_{j\geqslant 1}|f_{j}|^{2}dm

and the calculation used to prove Theorem 3.12 yields

Tunf,f=𝕋ξn(j1|fj(ξ)|2)𝑑m(ξ),\langle T_{u}^{n}f,f\rangle=\int_{\mathbb{T}}\xi^{n}\Big{(}\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}\Big{)}dm(\xi),
Tunf,f=𝕋ξ¯n(j1|fj(ξ)|2)𝑑m(ξ).\langle T_{u}^{*n}f,f\rangle=\int_{\mathbb{T}}\overline{\xi}^{n}\Big{(}\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}\Big{)}dm(\xi).

From here we observe

𝕋Pλ(ξ)𝑑μTu,f(ξ)\displaystyle\int_{\mathbb{T}}P_{\lambda}(\xi)d\mu_{T_{u},f}(\xi) =Kλ(Tu)f,f\displaystyle=\langle K_{\lambda}(T_{u})f,f\rangle
=n0λnTunf,f+n0λ¯nTunf,ff,f.\displaystyle=\sum_{n\geqslant 0}\lambda^{n}\langle T_{u}^{*n}f,f\rangle+\sum_{n\geqslant 0}\overline{\lambda}^{n}\langle T_{u}^{n}f,f\rangle-\langle f,f\rangle.
=n0λn𝕋ξ¯n(j1|fj(ξ)|2)𝑑m(ξ)\displaystyle=\sum_{n\geqslant 0}\lambda^{n}\int_{\mathbb{T}}\overline{\xi}^{n}\Big{(}\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}\Big{)}dm(\xi)
+n0λ¯n𝕋ξn(j1|fj(ξ)|2)𝑑m(ξ)\displaystyle\quad\quad+\sum_{n\geqslant 0}\overline{\lambda}^{n}\int_{\mathbb{T}}\xi^{n}\Big{(}\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}\Big{)}dm(\xi)
j1𝕋|fj(ξ)|2𝑑m\displaystyle\quad\quad\quad-\sum_{j\geqslant 1}\int_{\mathbb{T}}|f_{j}(\xi)|^{2}dm
=𝕋(11λξ¯+11λ¯ξ1)j1|fj(ξ)|2dm(ξ)\displaystyle=\int_{\mathbb{T}}(\frac{1}{1-\lambda\overline{\xi}}+\frac{1}{1-\overline{\lambda}\xi}-1)\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}dm(\xi)
=𝕋Pλ(ξ)j1|fj(ξ)|2dm(ξ)\displaystyle=\int_{\mathbb{T}}P_{\lambda}(\xi)\sum_{j\geqslant 1}|f_{j}(\xi)|^{2}dm(\xi)

Now use the uniqueness of the Fourier coefficients of a measure along with (4.3) to obtain (4.8). ∎

Notice how this gives us another way of thinking about Theorem 3.12: a unit vector fH2f\in H^{2} is TuT_{u}-inner if and only if μTu,f=m\mu_{T_{u},f}=m.

This brings us to an interesting related question. One can also show that for any f,gH2f,g\in H^{2}, we can define the harmonic function Kλ(Tu)f,g\langle K_{\lambda}(T_{u})f,g\rangle on 𝔻\mathbb{D} and prove this function also has bounded integral means. This yields, via Herglotz’s theorem, a complex valued measure μTu,f,g\mu_{T_{u},f,g} on 𝕋\mathbb{T} for which

(4.9) Kλ(Tu)f,g=𝕋Pλ(ξ)𝑑μT,f,g(ξ),λ𝔻.\langle K_{\lambda}(T_{u})f,g\rangle=\int_{\mathbb{T}}P_{\lambda}(\xi)d\mu_{T,f,g}(\xi),\quad\lambda\in\mathbb{D}.

See [2, Prop. 2.6] for details. A similar calculation used to prove Proposition 4.6 shows that

(4.10) dμTu,f,g=j1fjgj¯dm.d\mu_{T_{u},f,g}=\sum_{j\geqslant 1}f_{j}\overline{g_{j}}\;dm.

In the above formula, fjf_{j} and gjg_{j} come from the representations of ff and gg from (3.10). A general result from [3] says that given any FL1F\in L^{1} and a non-constant inner function uu that is not an automorphism, there are f,gH2f,g\in H^{2} for which

(4.11) F(ζ)=dμTu,f,gdm(ζ)F(\zeta)=\frac{d\mu_{T_{u},f,g}}{dm}(\zeta)

mm-almost everywhere. In the language of [3] this says that any FL1F\in L^{1} can be “factored through TuT_{u}”. Equivalently stated, using (4.10) and (4.11), we have

F(ζ)=j1fj(ζ)gj(ζ)¯.F(\zeta)=\sum_{j\geqslant 1}f_{j}(\zeta)\overline{g_{j}(\zeta)}.

This is an interesting representation for L1L^{1} functions and a refinement of the one from [3].

Question 4.12.

Proposition 4.7 shows that when φ\varphi is an inner function and f,gH2f,g\in H^{2}, then dμTφ,f,gd\mu_{T_{\varphi},f,g} is absolutely continuous with respect to mm. When φb(H)\varphi\in b(H^{\infty}) is this still the case? For this to be true we would need to know that φnf,g,n1\langle\varphi^{n}f,g\rangle,n\geqslant 1, are the Fourier coefficients of an L1L^{1} function.

5. Inner vectors via Clark measures

For any fixed α𝕋\alpha\in\mathbb{T} and inner function uu, the function

z1|u(z)|2|αu(z)|2=(α+u(z)αu(z))z\mapsto\frac{1-|u(z)|^{2}}{|\alpha-u(z)|^{2}}=\Re\Big{(}\frac{\alpha+u(z)}{\alpha-u(z)}\big{)}

is a positive harmonic function on 𝔻\mathbb{D}. Thus by Herglotz’s theorem, there is a unique positive measure σα\sigma_{\alpha} on 𝕋\mathbb{T} for which

1|u(z)|2|αu(z)|2=𝕋Pλ(ξ)𝑑σα(ξ).\frac{1-|u(z)|^{2}}{|\alpha-u(z)|^{2}}=\int_{\mathbb{T}}P_{\lambda}(\xi)d\sigma_{\alpha}(\xi).

The family of measures {σα:α𝕋}\{\sigma_{\alpha}:\alpha\in\mathbb{T}\} is called the family of Clark measures corresponding to uu. Let us record some important facts about this family of measures. Proofs can be found in [5].

First, one can use the fact that uu is an inner function, along with standard harmonic analysis, to prove that each σα\sigma_{\alpha} is singular with respect to mm. Second, if EαE_{\alpha} is defined to be the set of ξ𝕋\xi\in\mathbb{T} for which

limr1u(rξ)=α,\lim_{r\to 1^{-}}u(r\xi)=\alpha,

then EαE_{\alpha} is a Borel subset of 𝕋\mathbb{T} with

(5.1) σα(𝕋Eα)=0.\sigma_{\alpha}(\mathbb{T}\setminus E_{\alpha})=0.

In other words, σα\sigma_{\alpha} is “carried” by EαE_{\alpha}. From this we also see that the measures {σα:α𝕋}\{\sigma_{\alpha}:\alpha\in\mathbb{T}\} are singular with respect to each other. Third, a beautiful disintegration theorem of Aleksandrov says that if gL1g\in L^{1} then for mm-almost every α𝕋\alpha\in\mathbb{T}, integral

𝕋g(ξ)𝑑σα(ξ)\int_{\mathbb{T}}g(\xi)d\sigma_{\alpha}(\xi)

is well defined. Moreover this almost everywhere defined function

α𝕋g(ξ)𝑑σα(ξ)\alpha\mapsto\int_{\mathbb{T}}g(\xi)d\sigma_{\alpha}(\xi)

is integrable with respect to mm and

(5.2) 𝕋(𝕋g(ξ)𝑑σα(ξ))𝑑m(α)=𝕋g(ζ)𝑑m(ζ).\int_{\mathbb{T}}\Big{(}\int_{\mathbb{T}}g(\xi)d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha)=\int_{\mathbb{T}}g(\zeta)dm(\zeta).

Using Clark measures, we can use a technique from [11] to compute a formula for Kλ(Tu)f,f\langle K_{\lambda}(T_{u})f,f\rangle along with the measure dμTu,f/dmd\mu_{T_{u},f}/dm. This gives us another way to think about the formula (4.11). The result here is the following.

Theorem 5.3.

For an inner function uu and fH2f\in H^{2} we have

dμTu,f(α)=(𝕋|f(ξ)|2𝑑σα(ξ))dm(α).d\mu_{T_{u},f}(\alpha)=\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)\Big{)}\;dm(\alpha).
Proof.

For any fH2f\in H^{2} use the formulas from (5.1) and (5.2) to obtain

Tunf,f\displaystyle\langle T_{u}^{n}f,f\rangle =𝕋|f(ξ)|2u(ξ)n𝑑m(ξ)\displaystyle=\int_{\mathbb{T}}|f(\xi)|^{2}u(\xi)^{n}dm(\xi)
=𝕋(𝕋|f(ξ)|2u(ξ)n𝑑σα(ξ))𝑑m(α)\displaystyle=\int_{\mathbb{T}}\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}u(\xi)^{n}d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha)
=𝕋(𝕋|f(ξ)|2αn𝑑σα(ξ))𝑑m(α)\displaystyle=\int_{\mathbb{T}}\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}\alpha^{n}d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha)
=𝕋αn(𝕋|f(ξ)|2𝑑σα(ξ))𝑑m(α).\displaystyle=\int_{\mathbb{T}}\alpha^{n}\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha).

In a similar way

Tunf,f=𝕋α¯n(𝕋|f(ξ)|2𝑑σα(ξ))𝑑m(α).\langle T_{u}^{*n}f,f\rangle=\int_{\mathbb{T}}\overline{\alpha}^{n}\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha).

Now follow the proof of Proposition 4.7 to get

𝕋Pλ(ξ)𝑑μTu,f(ξ)\displaystyle\int_{\mathbb{T}}P_{\lambda}(\xi)d\mu_{T_{u},f}(\xi)
=Kλ(Tu)f,f\displaystyle=\langle K_{\lambda}(T_{u})f,f\rangle
=n0λnTunf,f+n0λ¯nTunf,ff,f\displaystyle=\sum_{n\geqslant 0}\lambda^{n}\langle T_{u}^{*n}f,f\rangle+\sum_{n\geqslant 0}\overline{\lambda}^{n}\langle T_{u}^{n}f,f\rangle-\langle f,f\rangle
=𝕋((11λα¯+11λ¯α1)(𝕋|f(ξ)|2𝑑σα(ξ)))𝑑m(α)\displaystyle=\int_{\mathbb{T}}\left(\Big{(}\frac{1}{1-\lambda\overline{\alpha}}+\frac{1}{1-\overline{\lambda}\alpha}-1\Big{)}\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)\Big{)}\right)dm(\alpha)
=𝕋Pλ(α)(𝕋|f(ξ)|2𝑑σα(ξ))𝑑m(α).\displaystyle=\int_{\mathbb{T}}P_{\lambda}(\alpha)\Big{(}\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha).

Use (4.3) along with the uniqueness of Fourier coefficients of a measure to compute the proof. ∎

Combing Theorem 5.3 and Proposition 4.6 yields the following result from [11].

Corollary 5.4.

A unit vector fH2f\in H^{2} is TuT_{u}-inner if and only if

𝕋|f(ξ)|2𝑑σα(ξ)=1\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)=1

for mm-almost every α𝕋\alpha\in\mathbb{T}.

Recall the notation from (4.9) that for a given inner function uu and f,gH2f,g\in H^{2}

Kλ(Tu)f,g=𝕋Pλ(ξ)𝑑μTu,f,g(ξ).\langle K_{\lambda}(T_{u})f,g\rangle=\int_{\mathbb{T}}P_{\lambda}(\xi)d\mu_{T_{u},f,g}(\xi).

Moreover, if deg(u)2\operatorname{deg}(u)\geqslant 2, any FL1F\in L^{1} can be written as dμTu,f,g(ξ)/dmd\mu_{T_{u},f,g}(\xi)/dm for some f,gH2f,g\in H^{2}. Here is another way of thinking about this via Clark measures. The same argument used to prove Theorem 5.3 shows that

(5.5) dμTu,f,g=𝕋f(ξ)g(ξ)¯𝑑σα(ξ)𝑑md\mu_{T_{u},f,g}=\int_{\mathbb{T}}f(\xi)\overline{g(\xi)}d\sigma_{\alpha}(\xi)\;dm

Since any FL1F\in L^{1} is equal to dμTu,f,g/dmd\mu_{T_{u},f,g}/dm for some f,gH2f,g\in H^{2} [3], we see that any FL1F\in L^{1} can be written as

F(α)=𝕋f(ξ)g(ξ)¯𝑑σα(ξ).F(\alpha)=\int_{\mathbb{T}}f(\xi)\overline{g(\xi)}d\sigma_{\alpha}(\xi).

This Clark measure viewpoint has the additional feature, via Aleksandrov’s theorem, that

𝕋F(α)𝑑m(α)\displaystyle\int_{\mathbb{T}}F(\alpha)dm(\alpha) =𝕋(𝕋f(ξ)g(ξ)¯𝑑σα(ξ))𝑑m(α)\displaystyle=\int_{\mathbb{T}}\Big{(}\int_{\mathbb{T}}f(\xi)\overline{g(\xi)}d\sigma_{\alpha}(\xi)\Big{)}dm(\alpha)
=𝕋f(ζ)g(ζ)𝑑m(ζ).\displaystyle=\int_{\mathbb{T}}f(\zeta)g(\zeta)dm(\zeta).
Example 5.6.

If uu is a finite Blaschke product of degree dd and α𝕋\alpha\in\mathbb{T}, then one can compute (see [5, p. 209] for the details) the Clark measure to be

dσα=j=1d1|u(ζj)|δζj,d\sigma_{\alpha}=\sum_{j=1}^{d}\frac{1}{|u^{\prime}(\zeta_{j})|}\delta_{\zeta_{j}},

where ζ1,,ζd\zeta_{1},\ldots,\zeta_{d} are the dd distinct solutions to the equation u(z)=αu(z)=\alpha and δζj\delta_{\zeta_{j}} is the unit point pass as ζj\zeta_{j}. The denominators in the above expression may look troublesome but at the end of the day we have u0u^{\prime}\not=0 on 𝕋\mathbb{T}. By Theorem 5.3 we see that

dμTu,fdm(α)=𝕋|f(ξ)|2𝑑σα(ξ)=j=1d|f(ζj)|2|u(ζj)|.\frac{d\mu_{T_{u},f}}{dm}(\alpha)=\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{\alpha}(\xi)=\sum_{j=1}^{d}\frac{|f(\zeta_{j})|^{2}}{|u^{\prime}(\zeta_{j})|}.

Thus the criterion for a unit vector fH2f\in H^{2} to be a TuT_{u}-inner vector is that the above sum is equal to 11 for mm-almost every α𝕋\alpha\in\mathbb{T}.

Furthermore, by (5.5), given FL1F\in L^{1}, there are f,gH2f,g\in H^{2} so that

F(α)=j=1df(ζj)g(ζj)¯|u(ζj)|F(\alpha)=\sum_{j=1}^{d}\frac{f(\zeta_{j})\overline{g(\zeta_{j})}}{|u^{\prime}(\zeta_{j})|}

for mm-almost every α𝕋\alpha\in\mathbb{T}. This formula appears in [3].

Example 5.7.

Let us apply this to the simple case where u(z)=z2u(z)=z^{2}. Given any α𝕋\alpha\in\mathbb{T}, the two solutions ζ1,ζ2\zeta_{1},\zeta_{2} to the equation z2=αz^{2}=\alpha are

ζ1=eiargα/2,ζ2=eiargα/2.\zeta_{1}=e^{i\arg\alpha/2},\quad\zeta_{2}=-e^{i\arg\alpha/2}.

Thus the condition that a unit ff is a Tz2T_{z^{2}}-inner vector becomes

|f(eiargα/2)|2+|f(eiargα/2)|2=2,m-a.e. α𝕋.|f(e^{i\arg\alpha/2})|^{2}+|f(-e^{i\arg\alpha/2})|^{2}=2,\quad\mbox{$m$-a.e. $\alpha\in\mathbb{T}$}.

Furthermore, given any FL1F\in L^{1}, there are f,gH2f,g\in H^{2} for which

F(α)=12f(eiargα/2)g(eiargα/2)¯+12f(eiargα/2)g(eiargα/2)¯.F(\alpha)=\tfrac{1}{2}f(e^{i\arg\alpha/2})\overline{g(e^{i\arg\alpha/2})}+\tfrac{1}{2}f(-e^{i\arg\alpha/2})\overline{g(-e^{i\arg\alpha/2})}.

This second fact was first observed in [3].

Example 5.8.

Consider the atomic inner function

u(z)=exp(z+1z1).u(z)=\exp\Big{(}\frac{z+1}{z-1}\Big{)}.

For a fixed t[0,2π)t\in[0,2\pi), the solutions to u(z)=eitu(z)=e^{it} are

ζk=i(t+2πk)+1i(t+2πk)1,k.\zeta_{k}=\frac{i(t+2\pi k)+1}{i(t+2\pi k)-1},\quad k\in\mathbb{Z}.

Noting that

|u(ζk)|=2|ζk1|2,|u^{\prime}(\zeta_{k})|=\frac{2}{|\zeta_{k}-1|^{2}},

a similar computation as in Example 5.6 shows that

dσeit=12kδζk|ζk1|2.d\sigma_{e^{it}}=\frac{1}{2}\sum_{k\in\mathbb{Z}}\delta_{\zeta_{k}}|\zeta_{k}-1|^{2}.

Thus

dμTu,fdm(eit)\displaystyle\frac{d\mu_{T_{u},f}}{dm}(e^{it}) =𝕋|f(ξ)|2𝑑σeit(ξ)\displaystyle=\int_{\mathbb{T}}|f(\xi)|^{2}d\sigma_{e^{it}}(\xi)
=12k|f(ζk)|2|ζk1|2\displaystyle=\frac{1}{2}\sum_{k\in\mathbb{Z}}|f(\zeta_{k})|^{2}|\zeta_{k}-1|^{2}
=k|f(i[t+2πk]+1i[t+2πk]1)|22|i(t+2πk)1|2.\displaystyle=\sum_{k\in\mathbb{Z}}\Big{|}f\Big{(}\frac{i[t+2\pi k]+1}{i[t+2\pi k]-1}\Big{)}\Big{|}^{2}\,\frac{2}{|i(t+2\pi k)-1|^{2}}.

To create a TuT_{u}-inner function, we need to find a unit vector fH2f\in H^{2} so that the above expression is equal to one for almost every tt. Let us create a specific example of when this happens. In fact we can even make ff unbounded. We already knew we could do this from Corollary 2.5 but our example below will be explicit, while the proof of Corollary 2.5 needed Grothendieck’s theorem and is not an explicit construction.

To see how to do this, fix β(12,1)\beta\in(\tfrac{1}{2},1), and let aka_{k}, kk\in\mathbb{Z}, be the collection of coefficients

(5.9) ak=11+|k|β.a_{k}=\frac{1}{1+|k|^{\beta}}.

Note that k|ak|2<\sum_{k\in\mathbb{Z}}|a_{k}|^{2}<\infty.

Let IkI_{k} be the indicator function of the interval [π+2πk,π+2πk)[-\pi+2\pi k,\pi+2\pi k), kk\in\mathbb{Z}. Now define FF on 𝕋\mathbb{T} by

F(eiθ):=2kakeiθ1Ik(i1+eiθ1eiθ).F(e^{i\theta}):=\sqrt{2}\sum_{k\in\mathbb{Z}}\frac{a_{k}}{e^{i\theta}-1}\,I_{k}\Big{(}i\frac{1+e^{i\theta}}{1-e^{i\theta}}\Big{)}.

Then

𝕋|F|2𝑑m\displaystyle\int_{\mathbb{T}}|F|^{2}\,dm =2ππ|F(eiθ)|2dθ2π\displaystyle=2\int_{-\pi}^{\pi}|F(e^{i\theta})|^{2}\,\frac{d\theta}{2\pi}
=2kππ|ak|2|eiθ1|2Ik(i1+eiθ1eiθ)dθ2π\displaystyle=2\sum_{k\in\mathbb{Z}}\int_{-\pi}^{\pi}\frac{|a_{k}|^{2}}{|e^{i\theta}-1|^{2}}\,I_{k}\Big{(}i\frac{1+e^{i\theta}}{1-e^{i\theta}}\Big{)}\,\frac{d\theta}{2\pi}
=2k|ak|2|it1|222Ik(t)2dt2π|it1|2\displaystyle=2\sum_{k\in\mathbb{Z}}\int_{-\infty}^{\infty}|a_{k}|^{2}\frac{|it-1|^{2}}{2^{2}}\,I_{k}(t)\,\frac{2\,dt}{2\pi|it-1|^{2}}
=kππ|ak|2|i[t+2πk]1|2dt2π|i[t+2πk]1|2\displaystyle=\sum_{k\in\mathbb{Z}}\int_{-\pi}^{\pi}|a_{k}|^{2}\big{|}i[t+2\pi k]-1\big{|}^{2}\,\frac{dt}{2\pi|i[t+2\pi k]-1|^{2}}
=k|ak|2<,\displaystyle=\sum_{k\in\mathbb{Z}}|a_{k}|^{2}<\infty,

i.e., FF is square integrable on 𝕋\mathbb{T} with

(5.10) F2=k|ak|2.\|F\|^{2}=\sum_{k\in\mathbb{Z}}|a_{k}|^{2}.

Next we establish that log|F|\log|F| is integrable. We’ll need the following estimates, which hold for all k0k\neq 0. First note that for k0k\neq 0,

|ak||i(t+2πk)1|\displaystyle|a_{k}|\big{|}i(t+2\pi k)-1\big{|} =|ak|([π+2π|k|]2+1)1/2\displaystyle=|a_{k}|\big{(}[\pi+2\pi|k|]^{2}+1\big{)}^{1/2}
|ak|2π|k|\displaystyle\geqslant|a_{k}|\cdot 2\pi|k|
2π|k|1+|k|β\displaystyle\geqslant\frac{2\pi|k|}{1+|k|^{\beta}}
1.\displaystyle\geqslant 1.

Consequently, for k0k\neq 0 and t[π,π)t\in[-\pi,\pi),

|log(|ak|i(t+2πk)1)|\displaystyle\Big{|}\log\Big{(}|a_{k}|\big{|}i(t+2\pi k)-1\Big{)}\Big{|} =log|ak||i(t+2πk)1|\displaystyle=\log{|a_{k}||i(t+2\pi k)-1|}
log|i(π+2π|k|)1|1+|k|β\displaystyle\leqslant\log\frac{|i(\pi+2\pi|k|)-1|}{1+|k|^{\beta}}
log([2π(|k|+1/2)]2+1)1/21+|k|β\displaystyle\leqslant\log\frac{([2\pi(|k|+1/2)]^{2}+1)^{1/2}}{1+|k|^{\beta}}
log([2π(|k|+|k|/2)]2+|k|2)1/2|k|β\displaystyle\leqslant\log\frac{([2\pi(|k|+|k|/2)]^{2}+|k|^{2})^{1/2}}{|k|^{\beta}}
log(|k|1β9π2+1).\displaystyle\leqslant\log(|k|^{1-\beta}\sqrt{9\pi^{2}+1}).

We now have

𝕋|log|F||𝑑m\displaystyle\int_{\mathbb{T}}\big{|}\log|F|\big{|}\,dm
=ππ|log|F(eiθ)||dθ2π\displaystyle=\int_{-\pi}^{\pi}\Big{|}\log|F(e^{i\theta})|\Big{|}\,\frac{d\theta}{2\pi}
=kππ|log|ak|2|eiθ1||Ik(i1+eiθ1eiθ)dθ2π\displaystyle=\sum_{k\in\mathbb{Z}}\int_{-\pi}^{\pi}\Big{|}\log\frac{|a_{k}|\sqrt{2}}{|e^{i\theta}-1|}\Big{|}\,I_{k}\Big{(}i\frac{1+e^{i\theta}}{1-e^{i\theta}}\Big{)}\,\frac{d\theta}{2\pi}
=k|log(|ak||it1|2/2)|Ik(t)dt2π|it1|2\displaystyle=\sum_{k\in\mathbb{Z}}\int_{-\infty}^{\infty}\Big{|}\log\big{(}|a_{k}|\big{|}it-1\big{|}\sqrt{2}/2\big{)}\Big{|}\,I_{k}(t)\,\frac{dt}{2\pi|it-1|^{2}}
=kππ|log(6π|k|1β|i[t+2πk]1|/2)|dt2π|i[t+2πk]1|2.\displaystyle=\sum_{k\in\mathbb{Z}}\int_{-\pi}^{\pi}\Big{|}\log\big{(}6\pi|k|^{1-\beta}\big{|}i[t+2\pi k]-1\big{|}/\sqrt{2}\big{)}\Big{|}\,\frac{dt}{2\pi|i[t+2\pi k]-1|^{2}}.

The series is summable, because the terms behave like (log|k|)/|k|2(\log|k|)/|k|^{2}.

It follows that there exists an outer function gH2g\in H^{2} with radial limit function satisfying |g|=|F||g|=|F| almost everywhere on 𝕋\mathbb{T}, namely

g(z):=exp(𝕋eiθ+zeiθzlog|F(eiθ)|dm(eiθ)).g(z):=\exp\Big{(}\int_{\mathbb{T}}\frac{e^{i\theta}+z}{e^{i\theta}-z}\log|F(e^{i\theta})|\,dm(e^{i\theta})\Big{)}.

Finally, let JJ be any classical inner function, and define f=gJf=gJ. Then

dμTu,fdm(eit)\displaystyle\frac{d\mu_{T_{u},f}}{dm}(e^{it}) =k|f(i[t+2πk]+1i[t+2πk]1)|22|i(t+2πk)1|2\displaystyle=\sum_{k\in\mathbb{Z}}\Big{|}f\Big{(}\frac{i[t+2\pi k]+1}{i[t+2\pi k]-1}\Big{)}\Big{|}^{2}\,\frac{2}{|i(t+2\pi k)-1|^{2}}
=k|F(i[t+2πk]+1i[t+2πk]1)|22|i(t+2πk)1|2\displaystyle=\sum_{k\in\mathbb{Z}}\Big{|}F\Big{(}\frac{i[t+2\pi k]+1}{i[t+2\pi k]-1}\Big{)}\Big{|}^{2}\,\frac{2}{|i(t+2\pi k)-1|^{2}}
=k|ak(i[t+2πk]1)2|2222|i(t+2πk)1|2\displaystyle=\sum_{k\in\mathbb{Z}}\frac{|a_{k}(i[t+2\pi k]-1)\sqrt{2}|^{2}}{2^{2}}\,\frac{2}{|i(t+2\pi k)-1|^{2}}
=k|ak|2.\displaystyle=\sum_{k\in\mathbb{Z}}|a_{k}|^{2}.

Notice from (5.10) that

dμTu,fdm(eit)=F2\frac{d\mu_{T_{u},f}}{dm}(e^{it})=\|F\|^{2}

and so one can scale FF so that it (and hence ff) is a unit vector This also gives us dμTu,f/dm(eit)=1d\mu_{T_{u},f}/dm(e^{it})=1 for almost every tt. Any such ff will be a TuT_{u}-inner function.

As a bonus, we get that the ff we just constructed is unbounded. To see this, note that FF is unbounded, since for θ\theta approaching zero, F(eiθ)F(e^{i\theta}) takes values

F(i[t+2πk]+1i[t+2πk]1)=ak1i[t+2πk]+1i[t+2πk]1=i[t+2πk]+12+2|k|βF\Big{(}\frac{i[t+2\pi k]+1}{i[t+2\pi k]-1}\Big{)}=\frac{a_{k}}{1-\frac{i[t+2\pi k]+1}{i[t+2\pi k]-1}}=\frac{-i[t+2\pi k]+1}{2+2|k|^{\beta}}

where t[π,π)t\in[-\pi,\pi). Since β<1\beta<1, this expression is unbounded as |k||k|\to\infty.

6. Inner vectors in model spaces

In this section we depart slightly from Toeplitz operators on H2H^{2} to the related topic of compressions of Toeplitz operators on model spaces. For an inner function Θ\Theta, recall the model space 𝒦Θ=(ΘH2)\mathcal{K}_{\Theta}=(\Theta H^{2})^{\perp}. An important operator to study here is the compressed shift operator

SΘ:𝒦u𝒦u,SΘf=PΘ(zf),S_{\Theta}:\mathcal{K}_{u}\to\mathcal{K}_{u},\quad S_{\Theta}f=P_{\Theta}(zf),

where PΘP_{\Theta} is the orthogonal projection of L2L^{2} onto 𝒦u\mathcal{K}_{u}. This operator is used to model a certain class of contraction operators on Hilbert space [8, Ch. 9] – hence the use of the phrase “model space.”

As a generalization of our discussion of classifying the TzT_{z}-inner vectors in H2H^{2}, one can ask for a description of the SΘS_{\Theta}-inner vectors in 𝒦Θ\mathcal{K}_{\Theta}, i.e., those unit vectors f𝒦Θf\in\mathcal{K}_{\Theta} for which

SΘnf,f=0,n1.\langle S_{\Theta}^{n}f,f\rangle=0,\quad n\geqslant 1.

Before continuing, let us make a few comments about SΘS_{\Theta}. For the proofs, see [8, Ch. 9]. First note that since SΘS_{\Theta} is a compression of TzT_{z} to 𝒦Θ\mathcal{K}_{\Theta} we have the identity

SΘn=PΘTzn|𝒦Θ.S_{\Theta}^{n}=P_{\Theta}T_{z^{n}}|_{\mathcal{K}_{\Theta}}.

Furthermore, we have the adjoint formula

SΘ=Tz¯|𝒦Θ.S_{\Theta}^{*}=T_{\overline{z}}|_{\mathcal{K}_{\Theta}}.

For any φH\varphi\in H^{\infty} there is the functional calculus for SΘS_{\Theta} which allows us to define

φ(SΘ)=PΘTφ|𝒦Θ\varphi(S_{\Theta})=P_{\Theta}T_{\varphi}|_{\mathcal{K}_{\Theta}}

along with the adjoint formula

φ(SΘ)=PΘTφ¯|𝒦Θ.\varphi(S_{\Theta})^{*}=P_{\Theta}T_{\overline{\varphi}}|_{\mathcal{K}_{\Theta}}.

One can actually compute the SΘS_{\Theta}-inner vectors with the following result from [8, p. 177].

Theorem 6.1.

Any SΘS_{\Theta}-inner function is an inner function. Moreover, 𝒦Θ\mathcal{K}_{\Theta} contains an inner function if and only if u(0)=0u(0)=0 and the inner functions belonging to 𝒦Θ\mathcal{K}_{\Theta} are precisely the inner divisors of Θ(z)/z\Theta(z)/z.

So now the question becomes the following.

Question 6.2.

What are the φ(SΘ)\varphi(S_{\Theta})-inner functions?

As we did before with Toeplitz operators, we focus our attention on the case where φ\varphi is inner. It is clear that the inner vectors for φ(SΘ)\varphi(S_{\Theta}) are the same as those for φ(SΘ)\varphi(S_{\Theta})^{*}. As observed with an analogous result in Proposition 2.3, we see that any (unit) vector in kerφ(SΘ)\ker\varphi(S_{\Theta})^{*} is a φ(SΘ)\varphi(S_{\Theta})^{*}-inner vector. It is well-known [8] that (assuming φ\varphi is an inner function)

kerφ(SΘ)=𝒦Θ𝒦φ=𝒦gcd(Θ,φ),\ker\varphi(S_{\Theta})^{*}=\mathcal{K}_{\Theta}\cap\mathcal{K}_{\varphi}=\mathcal{K}_{\operatorname{gcd}(\Theta,\varphi)},

where gcd(Θ,φ)\operatorname{gcd}(\Theta,\varphi) is the greatest common inner divisor of the inner functions Θ\Theta and φ\varphi.

At this point, it might the case that gcd(Θ,φ)\operatorname{gcd}(\Theta,\varphi) is a unimodular constant function whence 𝒦gcd(Θ,φ)={0}\mathcal{K}_{\operatorname{gcd}(\Theta,\varphi)}=\{0\} and it is not clear as to whether or not there are any φ(SΘ)\varphi(S_{\Theta})-inner vectors.

Question 6.3.

We know that if gcd(Θ,φ)\operatorname{gcd}(\Theta,\varphi) is non-constant, then there are φ(SΘ)\varphi(S_{\Theta})-inner vectors. Is the converse true?

For the special case where φ|Θ\varphi|\Theta, let us find a class of φ(SΘ)\varphi(S_{\Theta})-inner vectors. Define

I:=ΘφI:=\frac{\Theta}{\varphi}

and observe from a result in [7] that an analytic function gg on 𝔻\mathbb{D} multiplies 𝒦φ\mathcal{K}_{\varphi} to 𝒦Θ\mathcal{K}_{\Theta} if and only g𝒦zIg\in\mathcal{K}_{zI}. Recall from Theorem 6.1 that the inner functions in 𝒦zI\mathcal{K}_{zI} are precisely the inner divisors of II. Here is our result about some of the φ(SΘ)\varphi(S_{\Theta})-inner vectors.

Theorem 6.4.

With the notation above, any unit vector from

{v𝒦φ:v|I}\{v\mathcal{K}_{\varphi}:v|I\}

is a φ(SΘ)\varphi(S_{\Theta})-inner vector.

Proof.

Let ff be a unit vector from 𝒦φ\mathcal{K}_{\varphi} and note that vf𝒦Θvf\in\mathcal{K}_{\Theta} and hence PΘ(vf)=vfP_{\Theta}(vf)=vf. Thus for all n1n\geqslant 1 we have

(φ(SΘ))n(vf),vf\displaystyle\langle(\varphi(S_{\Theta}))^{n}(vf),vf\rangle =PΘ(φnfv),vf\displaystyle=\langle P_{\Theta}(\varphi^{n}fv),vf\rangle
=φnvf,PΘ(vf)\displaystyle=\langle\varphi^{n}vf,P_{\Theta}(vf)\rangle
=φnvf,vf\displaystyle=\langle\varphi^{n}vf,vf\rangle
=φnf,f\displaystyle=\langle\varphi^{n}f,f\rangle
=f,Tφ¯nf.\displaystyle=\langle f,T_{\overline{\varphi}}^{n}f\rangle.

But since f𝒦φ=kerTφ¯f\in\mathcal{K}_{\varphi}=\ker T_{\overline{\varphi}}, this last quantity is equal to zero. This shows that vfvf is a φ(SΘ)\varphi(S_{\Theta})-inner vector. ∎

When Θ(0)=0\Theta(0)=0 and φ(z)=z\varphi(z)=z, notice how this recovers Theorem 6.1. At the other extreme, notice that when φ=Θ\varphi=\Theta then II is a unimodular constant inner function and the theorem above yields 𝒦Θ\mathcal{K}_{\Theta} as the complete set of TΘT_{\Theta}-inner functions. Of course this result is obvious once one realizes that TΘf,f=0\langle T_{\Theta}f,f\rangle=0 for any f𝒦Θf\in\mathcal{K}_{\Theta} by the definition of the model space 𝒦Θ=(ΘH2)\mathcal{K}_{\Theta}=(\Theta H^{2})^{\perp}.

Also observe that one can relax the assumption that φ|Θ\varphi|\Theta and set I=u/gcd(Θ,φ)I=u/\operatorname{gcd}(\Theta,\varphi) and give a more general version of the theorem above.

References

  • [1] A. Böttcher and Silbermann B. Analysis of Toeplitz operators. Springer Monographs in Mathematics. Springer-Verlag, Berlin, second edition, 2006.
  • [2] Isabelle Chalendar. The operator-valued Poisson kernel and its applications. Irish Math. Soc. Bull., (51):21–44, 2003.
  • [3] Isabelle Chalendar and Jonathan R. Partington. Spectral density for multiplication operators with applications to factorization of L1L^{1} functions. J. Operator Theory, 50(2):411–422, 2003.
  • [4] R. Cheng, J. Mashreghi, and William T. Ross. Inner functions in reproducing kernel spaces. preprint.
  • [5] Joseph A. Cima, Alec L. Matheson, and William T. Ross. The Cauchy transform, volume 125 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2006.
  • [6] P. L. Duren. Theory of Hp{H}^{p} spaces. Academic Press, New York, 1970.
  • [7] Emmanuel Fricain, Andreas Hartmann, and William T. Ross. Multipliers between model spaces. Studia Math., 240(2):177–191, 2018.
  • [8] Stephan Ramon Garcia, Javad Mashreghi, and William T. Ross. Introduction to model spaces and their operators, volume 148 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2016.
  • [9] Paul R. Halmos. Shifts on Hilbert spaces. J. Reine Angew. Math., 208:102–112, 1961.
  • [10] T. L. Lance and M. I. Stessin. Multiplication invariant subspaces of Hardy spaces. Canad. J. Math., 49(1):100–118, 1997.
  • [11] Michael Stessin and Kehe Zhu. Generalized factorization in Hardy spaces and the commutant of Toeplitz operators. Canad. J. Math., 55(2):379–400, 2003.