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

Shift operators, Cauchy integrals and approximations

Bartosz Malman Division of Mathematics and Physics, Mälardalen University, Västerås, Sweden [email protected]
Abstract.

This article consists of two connected parts. In the first part, we study the shift invariant subspaces in certain 𝒫2(μ)\mathcal{P}^{2}(\mu)-spaces, which are the closures of analytic polynomials in the Lebesgue spaces 2(μ)\mathcal{L}^{2}(\mu) defined by a class of measures μ\mu living on the closed unit disk 𝔻¯\overline{\mathbb{D}}. The measures μ\mu which occur in our study have a part on the open disk 𝔻\mathbb{D} which is radial and decreases at least exponentially fast near the boundary. Our focus is on those shift invariant subspaces which are generated by a bounded function in HH^{\infty}. In this context, our results are definitive. We give a characterization of the cyclic singular inner functions by an explicit and readily verifiable condition, and we establish certain permanence properties of non-cyclic ones which are important in the applications. The applications take up the second part of the article. We prove that if a function g1(𝕋)g\in\mathcal{L}^{1}(\mathbb{T}) on the unit circle 𝕋\mathbb{T} has a Cauchy transform with Taylor coefficients of order 𝒪(exp(cn))\mathcal{O}\big{(}\exp(-c\sqrt{n})\big{)} for some c>0c>0, then the set U={x𝕋:|g(x)|>0}U=\{x\in\mathbb{T}:|g(x)|>0\} is essentially open and log|g|\log|g| is locally integrable on UU. We establish also a simple characterization of analytic functions b:𝔻𝔻b:\mathbb{D}\to\mathbb{D} with the property that the de Branges-Rovnyak space (b)\mathcal{H}(b) contains a dense subset of functions which, in a sense, just barely fail to have an analytic continuation to a disk of radius larger than 1. We indicate how close our results are to being optimal and pose a few questions.

1. Introduction and main results

1.1. Some background

We will study spaces of analytic functions corresponding to Borel measures of the form

(1.1) dμ(z)=G(1|z|)dA(z)+w(z)dm(z),d\mu(z)=G(1-|z|)\,dA(z)+w(z)d\textit{m}(z),

where dAdA and dmdm are the area and arc-length measures on, respectively, the unit disk 𝔻:={z:|z|<1}\mathbb{D}:=\{z\in\mathbb{C}:|z|<1\} and its boundary circle 𝕋:={z:|z|=1}\mathbb{T}:=\{z\in\mathbb{C}:|z|=1\}. The radial weight G(1|z|)G(1-|z|) living on 𝔻\mathbb{D} is defined in terms of a continuous, increasing and positive function GG, and the weight ww living on 𝕋\mathbb{T} is a general Borel measurable non-negative integrable function. Given such a measure, we may construct first the Lebesgue space 2(μ)\mathcal{L}^{2}(\mu) of (equivalence classes of) Borel measurable functions living on the carrier of μ\mu, and next consider its subspace 𝒫2(μ)\mathcal{P}^{2}(\mu), by which we denote the smallest closed subspace of 2(μ)\mathcal{L}^{2}(\mu) which contains the set 𝒫\mathcal{P} of analytic polynomials. The space 𝒫2(μ)\mathcal{P}^{2}(\mu) will be the setting for the first part of our study.

The shift operator Mz:𝒫2(μ)𝒫2(μ)M_{z}:\mathcal{P}^{2}(\mu)\to\mathcal{P}^{2}(\mu), which takes a function f(z)f(z) to zf(z)zf(z), is a subnormal operator, in the sense that it is the restriction of a normal operator, namely Mz:2(μ)2(μ)M_{z}:\mathcal{L}^{2}(\mu)\to\mathcal{L}^{2}(\mu), to an invariant subspace. From the point of view of an operator theorist, the significance of the pair (𝒫2(μ),Mz)(\mathcal{P}^{2}(\mu),M_{z}) lies in the fact the study of subnormal operators can essentially be reduced to the study of the operator Mz:𝒫2(μ)𝒫2(μ)M_{z}:\mathcal{P}^{2}(\mu)\to\mathcal{P}^{2}(\mu) for some measure μ\mu which is compactly supported in the plane. The monograph [6] by Conway is an excellent source of information on this topic.

For measures such as (1.1), the space 𝒫2(μ)\mathcal{P}^{2}(\mu) is, like 2(μ)\mathcal{L}^{2}(\mu), a space of Borel measurable functions on the closed disk 𝔻¯=𝔻𝕋\overline{\mathbb{D}}=\mathbb{D}\cup\mathbb{T}. In certain cases it is even a space of analytic functions on 𝔻\mathbb{D}. In such a case, each element f𝒫2(μ)f\in\mathcal{P}^{2}(\mu), a priori interpreted as a function on 𝔻¯\overline{\mathbb{D}}, has a unique restriction f𝔻f_{\mathbb{D}} to the disk 𝔻\mathbb{D}. The restriction f𝔻f_{\mathbb{D}} must be an analytic function by the virtue of it being a locally uniform limit of analytic polynomials. We will below use the term irreducible for such a space which is in this sense ”analytic”. It is a difficult problem (and in general open) to determine which weight pairs (G,w)(G,w) as in (1.1) produce an irreducible space. Khrushchev in the article [16] solved certain special cases of the problem. For instance, his results apply to G(t)=tnG(t)=t^{n} for some n>0n>0, and w=1Ew=1_{E} being a characteristic function of a set E𝕋E\subset\mathbb{T} in a certain class (defined in terms of Beurling-Carleson conditions). Already these results have fascinating applications to function theory, of which there are plenty in [16]. The article [25] builds on Khrushchev’s work, explains the structure of 𝒫2(μ)\mathcal{P}^{2}(\mu) when w=1Ew=1_{E} and EE is a general subset of 𝕋\mathbb{T}, and showcases further applications to the theory of the Cauchy integral operator and de Branges-Rovnyak spaces.

1.2. Irreducible 𝒫2(μ)\mathcal{P}^{2}(\mu)-spaces

Recently, the author found in [24] an exact condition for irreducibility of 𝒫2(μ)\mathcal{P}^{2}(\mu) in the case when G(t)G(t) decays at least exponentially as t0+t\to 0^{+}, thus confirming a conjecture by Kriete and MacCluer from [20]. Roughly speaking, if G(t)G(t) is smaller than the weight exp(ct1)\exp(-ct^{-1}) for some c>0c>0, or more precisely if

(ExpDec) lim infx0+xlog1/G(x)>0,\liminf_{x\to 0^{+}}\,x\log 1/G(x)>0,

but large enough to satisfy

(LogLogInt) 0dloglog(1/G(x))dx<,\int_{0}^{d}\log\log(1/G(x))\,dx<\infty,

for some d>0d>0, then the space 𝒫2(μ)\mathcal{P}^{2}(\mu) is irreducible if and only if the carrier set of the measure dμ𝕋=wdmd\mu_{\mathbb{T}}=w\,d\textit{m} on 𝕋\mathbb{T} can be covered by intervals II satisfying the condition

(1.2) Ilogwdm>.\int_{I}\log w\,d\textit{m}>-\infty.

In order to properly state the result we will need to define the following concept of core sets. For our purposes this concept is critical, and it will appear frequently throughout the article.

Definition 1.1.

(Core sets of weights) \thlabelCoreDef Let ww be a non-negative integrable function on 𝕋\mathbb{T}. We define core(w)\text{core}(w) to be the union of all open intervals II for which (1.2) holds. In other words,

(1.3) core(w)={x𝕋: there exists open I containing x for which (1.2) holds }\text{core}(w)=\{x\in\mathbb{T}:\text{ there exists open }I\text{ containing }x\text{ for which }\eqref{logwIntIntegr}\text{ holds }\}

The set core(w)\text{core}(w) is open, and it does not depend on the particular representative of ww in the space 1(𝕋)\mathcal{L}^{1}(\mathbb{T}) of equivalence classes of functions which are Lebesgue integrable on 𝕋\mathbb{T} with respect to dmd\textit{m}.

Definition 1.2.

(Carrier sets) Let η\eta be a non-negative Borel measure on 𝕋\mathbb{T}. A Borel subset EE of 𝕋\mathbb{T} is a carrier for η\eta if

η(𝕋E)=0.\eta(\mathbb{T}\setminus E)=0.

If ww is a Borel measurable function on 𝕋\mathbb{T}, then we say that a set EE is a carrier for ww if it is a carrier for the Borel measure wdmw\,d\textit{m}.

Carriers are obviously not unique. The set

(1.4) {x𝕋:w(x)>0}\{x\in\mathbb{T}:w(x)>0\}

is a carrier for ww. If ww is only defined up to a set of mm-measure zero, then we may take as a carrier for ww any set differing from (1.4) by a set of mm-measure zero. Since log0=\log 0=-\infty, it is obvious from (1.2) that core(w)\text{core}(w) is essentially contained in any carrier of ww.

Irreducibility of 𝒫2(μ)\mathcal{P}^{2}(\mu)-spaces of the form (1.1) with GG satisfying (ExpDec) and (LogLogInt) can be characterized in terms of core sets. The next theorem, fundamental to our study, follows from [24, Theorem A], with the non-trivial part being the equivalence of the third condition and the other two.

Theorem 1.3.
\thlabel

IrrDef For a space 𝒫2(μ)\mathcal{P}^{2}(\mu) defined by a measure μ\mu of the form (1.1), with GG satisfying (ExpDec) and (LogLogInt), the following three conditions are equivalent:

  1. (i)

    the space 𝒫2(μ)\mathcal{P}^{2}(\mu) contains no non-trivial characteristic function of a measurable subset of 𝔻¯\overline{\mathbb{D}}: if AA is a Borel subset of 𝔻¯\overline{\mathbb{D}} and 1A𝒫2(μ)1_{A}\in\mathcal{P}^{2}(\mu) is not the zero element, then 1A=1𝔻¯1_{A}=1_{\overline{\mathbb{D}}}.

  2. (ii)

    the space 𝒫2(μ)\mathcal{P}^{2}(\mu) is a space of analytic functions on 𝔻\mathbb{D} in which the analytic polynomials are dense,

  3. (iii)

    the set core(w)\text{core}(w) is a carrier for ww, or in other words it coincides with (1.4), up to a set of mm-measure zero.

Definition 1.4.
\thlabel

IrrDef2(Irreducible spaces) A space 𝒫2(μ)\mathcal{P}^{2}(\mu) is irreducible if it satisfies the three equivalent conditions stated in \threfIrrDef.

In particular, the following measures μ\mu correspond to irreducible 𝒫2(μ)\mathcal{P}^{2}(\mu):

(T1T1) dμ(z)=exp(c(1|z|)β)dA(z)+w(z)dm(z),c>0,β1d\mu(z)=\exp\Big{(}-\frac{c}{(1-|z|)^{\beta}}\Big{)}dA(z)+w(z)d\textit{m}(z),\quad c>0,\,\beta\geq 1

and

(T2T2) dμ(z)=exp(cexp(1(1|z|)α))dA(z)+w(z)dm(z),c>0,α(0,1).d\mu(z)=\exp\Bigg{(}-c\exp\Bigg{(}\frac{1}{(1-|z|)^{\alpha}}\Bigg{)}\Bigg{)}dA(z)+w(z)d\textit{m}(z),\quad c>0,\,\alpha\in(0,1).

If the core(w)\text{core}(w) is not a carrier of ww, then the space 𝒫2(μ)\mathcal{P}^{2}(\mu) will contain a full Lebesgue space 2(wrdm)\mathcal{L}^{2}(w_{r}d\textit{m}), members of which live only on 𝕋\mathbb{T}. Here wrw_{r} denotes a certain residual weight. The residuals play no role in the statements of our main results, but will be important in the proofs. Their definition is postponed to coming sections.

The reader might wonder what happens in the case β<1\beta<1 in (T1T1). Then (ExpDec) is violated, and condition (iii)(iii) in \threfIrrDef implies (ii)(ii), but the converse is false. This can be inferred from work of Khruschev in [16], and this idea is further elaborated on in [25]. Also one might ask what happens if α1\alpha\geq 1 in (T2T2), which means that (LogInt) is violated. This case is less interesting: Volberg’s theorem in [30] implies that 𝒫2(μ)\mathcal{P}^{2}(\mu) is then either a close cousin of the Hardy space H2H^{2} (this happens when 𝕋logwdm>\int_{\mathbb{T}}\log w\,d\textit{m}>-\infty) or it is not a space of analytic functions at all (if 𝕋logwdm=\int_{\mathbb{T}}\log w\,d\textit{m}=-\infty). See also the introductory section to [24] for a more detailed account.

1.3. Invariant subspaces generated by singular inner functions

Having established fairly sharp conditions for irreducibility, a way opens to an operator and function theoretic study of this class of spaces. Motivated by certain applications which will soon be detailed, in the first part of the article we study the structure of MzM_{z}-invariant subspaces of 𝒫2(μ)\mathcal{P}^{2}(\mu) generated by functions in HH^{\infty}, the algebra of bounded analytic functions in 𝔻\mathbb{D}. This question readily reduces to the study of invariant subspaces generated by singular inner functions

(1.5) Sν(z)=exp(𝕋x+zxz𝑑ν(x)),z𝔻,S_{\nu}(z)=\exp\Big{(}-\int_{\mathbb{T}}\frac{x+z}{x-z}d\nu(x)\Big{)},\quad z\in\mathbb{D},

where ν\nu is a finite positive singular Borel measure on 𝕋\mathbb{T}. For hHh\in H^{\infty}, we will denote by [h][h] the smallest MzM_{z}-invariant subspace containing hh. It is well-known that any singular inner function generates a non-trivial invariant subspace in the classical Hardy space H2H^{2} of square-summable Taylor series, and it is almost as well-known that in order for SνS_{\nu} to generate a non-trivial invariant subspace in the standard weighted Bergman spaces (which are 𝒫2(μ)\mathcal{P}^{2}(\mu)-spaces of the kind (1.1) themselves, with G(t)=tnG(t)=t^{n} for some n>1n>-1, and w0w\equiv 0) we must have ν(A)>0\nu(A)>0 for some Beurling-Carleson set AA (see [18], [19], [27]).

Our first main result characterizes the cyclic singular inner functions in the considered class of 𝒫2(μ)\mathcal{P}^{2}(\mu)-spaces. By cyclicity we mean that [Sν]=𝒫2(μ)[S_{\nu}]=\mathcal{P}^{2}(\mu). It is not hard to see that the minimal considered rate of decay (ExpDec) of the part of μ\mu living on 𝔻\mathbb{D} makes every non-vanishing bounded function be cyclic in 𝒫2(μ)\mathcal{P}^{2}(\mu) in the case that w=0w=0. Thus only properties of ww can stop SνS_{\nu} from being cyclic.

Theorem A.
\thlabel

CyclicityMainTheorem Let 𝒫2(μ)\mathcal{P}^{2}(\mu) be an irreducible space defined by a measure μ\mu of the form (1.1). The following two statements are equivalent.

  1. (i)

    The singular inner function SνS_{\nu} is cyclic in 𝒫2(μ)\mathcal{P}^{2}(\mu).

  2. (ii)

    The measure ν\nu assigns no mass to the core of the weight ww:

    ν(core(w))=0.\nu\big{(}\text{core}(w)\big{)}=0.

Note that core(w)\text{core}(w) is open, and hence Borel measurable, so ν(core(w))\nu\big{(}\text{core}(w)\big{)} makes perfect sense.

Example 1.5.
\thlabel

PointMassExample Let δa\delta_{a} be a point mass at a𝕋a\in\mathbb{T}, and

(1.6) w(x)=exp(1|x1|),x𝕋.w(x)=\exp\Bigg{(}-\frac{1}{|x-1|}\Bigg{)},\quad x\in\mathbb{T}.

Then it is easy to check that

core(w)=𝕋{1}.\text{core}(w)=\mathbb{T}\setminus\{1\}.

Consequently, the singular inner function

Sδa(z)=exp(a+zaz),z𝔻S_{\delta_{a}}(z)=\exp\Bigg{(}-\frac{a+z}{a-z}\Bigg{)},\quad z\in\mathbb{D}

is cyclic, in the considered class of 𝒫2(μ)\mathcal{P}^{2}(\mu) constructed from ww appearing in (1.6), if and only if a=1a=1.

Having settled the cyclicity question, we turn our attention to the invariant subspace [Sν][S_{\nu}] generated by a singular inner function corresponding to a measure ν\nu which places all its mass on the core: ν(𝕋)=ν(core(w))\nu(\mathbb{T})=\nu\big{(}\text{core}(w)\big{)}. In other words, core(w)\text{core}(w) is a carrier for ν\nu. A problem which arises in the theory of normalized Cauchy integrals and de Branges-Rovnyak spaces (b)\mathcal{H}(b) (to be discussed below) is to determine which functions are contained in the intersection H2[Sν]H^{2}\cap[S_{\nu}], or sometimes in 𝒩+[Sν]\mathcal{N}^{+}\cap[S_{\nu}], where 𝒩+\mathcal{N}^{+} is the Smirnov class of the disk 𝔻\mathbb{D} (see [13] for precise definitions):

𝒩+={u/v:u,vH,v outer}\mathcal{N}^{+}=\{u/v:u,v\in H^{\infty},\,v\text{ outer}\}

In this context, we have the following result.

Theorem B.
\thlabel

PermanenceMainTheorem Let SνS_{\nu} be a singular inner function corresponding to a measure ν\nu which satisfies

ν(𝕋)=ν(core(w)).\nu(\mathbb{T})=\nu\big{(}\text{core}(w)\big{)}.

In an irreducible 𝒫2(μ)\mathcal{P}^{2}(\mu)-space defined by a measure μ\mu of the form (1.1), the invariant subspace [Sν][S_{\nu}] satisfies

[Sν]𝒩+Sν𝒩+.[S_{\nu}]\cap\mathcal{N}^{+}\subset S_{\nu}\mathcal{N}^{+}.

In other words, if f𝒩+f\in\mathcal{N}^{+} can be approximated by polynomial multiples of SνS_{\nu} in the norm of 𝒫2(μ)\mathcal{P}^{2}(\mu), and ν\nu places all of its mass on core(w)\text{core}(w), then SνS_{\nu} appears in the inner-outer factorization of ff. Under the additional assumption that ww is bounded, a simple argument will show that in fact [Sν]H2=SνH2[S_{\nu}]\cap H^{2}=S_{\nu}H^{2}. In [21] and [23], the feature of SνS_{\nu} appearing in \threfPermanenceMainTheorem is called its permanence property. It is obvious that a singular inner function satisfying the permanence property cannot be cyclic.

For the considered class of spaces, \threfCyclicityMainTheorem and \threfPermanenceMainTheorem completely determine the structure of MzM_{z}-invariant subspaces generated by bounded analytic functions. Indeed, it follows that if h=BSνUHh=BS_{\nu}U\in H^{\infty} is the inner-outer factorization of hh into a Blaschke product BB, singular inner function SνS_{\nu} and outer function UU, then

[h]=[Sνw],[h]=[S_{\nu_{w}}],

where νw\nu_{w} is the restriction of the singular measure ν\nu to the set core(w)\text{core}(w).

1.4. Functions of rapid spectral decay and Cauchy integrals

Irreducible spaces find applications in the theory of Cauchy integrals.

Definition 1.6.

(Functions of rapid spectral decay) Let f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} be an analytic function in 𝔻\mathbb{D}. If the Taylor coefficients {fn}n0\{f_{n}\}_{n\geq 0} decay so fast that for some c>0c>0 we have

(RSD) supn0|fn|exp(cn)<,\sup_{n\geq 0}\,|f_{n}|\exp\big{(}c\sqrt{n})<\infty,

then we say that ff is a function of rapid spectral decay.

Trivial examples of functions ff satisfying (RSD) are the analytic polynomials, and functions which extend analytically to a larger disk r𝔻={z:|z|<r}r\mathbb{D}=\{z\in\mathbb{C}:|z|<r\}, r>1r>1. In those cases, the limit in (RSD) is zero even when the term exp(cn)\exp\big{(}c\sqrt{n}\big{)} in (RSD) is replaced by exp(cnα)\exp\big{(}cn^{\alpha}\big{)} for α<1\alpha<1. Conversely, if ff has an analytic extension to a disk around the origin of radius larger than 1, then |fn|=𝒪(exp(cn))|f_{n}|=\mathcal{O}\big{(}\exp(-cn)\big{)} for some c>0c>0.

Let us assume that ν\nu is a finite Borel measure for which the Cauchy integral

(1.7) 𝒞ν(z):=𝕋11x¯z𝑑ν(x),z𝔻\mathcal{C}_{\nu}(z):=\int_{\mathbb{T}}\frac{1}{1-\overline{x}z}d\nu(x),\quad z\in\mathbb{D}

is a function satisfying (RSD). Can we say something about the nature of the measure ν\nu? The Cauchy integral 𝒞ν\mathcal{C}_{\nu} has a representation of the form

𝒞ν(z)=n0νnzn,z𝔻\mathcal{C}_{\nu}(z)=\sum_{n\geq 0}\nu_{n}z^{n},\quad z\in\mathbb{D}

where {νn}n0\{\nu_{n}\}_{n\geq 0} is the sequence of Fourier coefficients of ν\nu indexed by non-negative integers. The rest of the coefficients are annihilated under 𝒞\mathcal{C}, and the condition (RSD) gives us no information about νn\nu_{n} for n<0n<0. However, the following statement is a consequence of the irreducibility of spaces corresponding to measures of the form (T1T1).

Theorem C.
\thlabel

UncertThmRSD Let ν\nu be a finite Borel measure on 𝕋\mathbb{T}, and assume that the Cauchy integral 𝒞ν\mathcal{C}_{\nu}, given by (1.7), satisfies (RSD). Then the measure ν\nu is absolutely continuous with respect to the Lebesgue measure dmd\textit{m}:

dν=gdm,g1(𝕋),d\nu=g\,d\textit{m},\quad g\in\mathcal{L}^{1}(\mathbb{T}),

and there exists an open set UU which differs from

(1.8) {x𝕋:|g(x)|>0}\{x\in\mathbb{T}:|g(x)|>0\}

only by a set of m-measure zero, with the property that to each xUx\in U there corresponds an interval IUI\subset U containing xx for which we have

Ilog|g(x)|dm(x)>.\int_{I}\log|g(x)|\,d\textit{m}(x)>-\infty.

The function log|g|\log|g| is, in general, not integrable on the entire open set UU appearing in \threfUncertThmRSD.

In a way, \threfUncertThmRSD is similar to the classical theorem of brothers Riesz on structure of measures ν\nu on 𝕋\mathbb{T} with vanishing positive Fourier coefficients. In our setting, the vanishing of the coefficients is replaced by a weaker condition of their rapid decay forced by the condition (RSD). It should be noted that if we were to replace in (RSD) the term exp(cn)\exp\big{(}c\sqrt{n}\big{)} by exp(cnα)\exp\big{(}cn^{\alpha}\big{)} for any α<1/2\alpha<1/2, and thus consider the weaker unilateral spectral decay condition

supn0|νn|exp(cnα)<,\sup_{n\geq 0}\,|\nu_{n}|\exp\big{(}cn^{\alpha}\big{)}<\infty,

then a structural result for ν\nu as in \threfUncertThmRSD does not hold: dν=gdmd\nu=g\,d\textit{m} will still be absolutely continuous, but examples show that gg can be chosen so that the set in (1.8) is closed and contains no interval. This follows from a related work of Khrushchev in [16]. There should be room for a slight improvement of the result (see the discussion in Section 1.6.4 below). We ought to mention also that Volberg in [30] found spectral decay conditions making the set in (1.8) fill up the whole circle 𝕋\mathbb{T}. We will return to both these works below.

1.5. Condition (RSD) in de Branges-Rovnyak spaces

In most classical Hilbert spaces of analytic functions in the unit disk, the family of functions which extend analytically to a larger disk forms a dense subset of the space. This is not the case in Hilbert spaces of normalized Cauchy integrals. These are the so-called model spaces KθK_{\theta}, where θ\theta is an inner function, and more generally the de Branges-Rovnyak spaces (b)\mathcal{H}(b), where the symbol bb is any analytic self-map of the unit disk. There are several ways to define the space (b)\mathcal{H}(b), the easiest perhaps being by stating that it is the Hilbert space of analytic functions on 𝔻\mathbb{D} with a reproducing kernel of the form

kb(λ,z)=1b(λ)¯b(z)1λ¯z,λ,z𝔻.k_{b}(\lambda,z)=\frac{1-\overline{b(\lambda)}b(z)}{1-\overline{\lambda}z},\quad\lambda,z\in\mathbb{D}.

Alternatively, we may realize it as the space of normalized Cauchy integrals of functions f~2(νb)\tilde{f}\in\mathcal{L}^{2}(\nu_{b}), given in the special case b(0)=0b(0)=0 by the formula

(1.9) f(z)=(1b(z))𝕋f~(x)1x¯z𝑑νb(x),z𝔻.f(z)=\big{(}1-b(z)\big{)}\int_{\mathbb{T}}\frac{\tilde{f}(x)}{1-\overline{x}z}d\nu_{b}(x),\quad z\in\mathbb{D}.

Here νb\nu_{b} is the Aleksandrov-Clark measure of bb, these two objects being related by the formula

(1.10) Re(1+b(z)1b(z))=𝕋1|z|2|xz|2𝑑νb(x),z𝔻.\operatorname{Re}\Bigg{(}\frac{1+b(z)}{1-b(z)}\Bigg{)}=\int_{\mathbb{T}}\frac{1-|z|^{2}}{|x-z|^{2}}d\nu_{b}(x),\quad z\in\mathbb{D}.

The normalization refers to multiplication of the Cauchy integral in (1.9) by the factor 1b(z)1-b(z), which ensures that the product lands in H2H^{2}. It is well-known that model spaces Kθ=(θ)K_{\theta}=\mathcal{H}(\theta) correspond to the purely singular measures νθ\nu_{\theta} in (1.10). In fact, every positive finite Borel measure ν\nu on 𝕋\mathbb{T} corresponds to a function b=bνb=b_{\nu} through the formula (1.10). See [5] for more details.

If θ\theta is a singular inner function, then 𝒦θ\mathcal{K}_{\theta} will contain no functions which extend analytically across 𝕋\mathbb{T}. Moreover, it is a consequence of deep results on cyclicity of singular inner functions of Beurling from [3], and also of more recent results of El-Fallah, Kellay and Seip from [10], that in fact if θ\theta is singular, then for any non-zero function f(z)=n0fnzn𝒦θf(z)=\sum_{n\geq 0}f_{n}z^{n}\in\mathcal{K}_{\theta} and for any c>0c>0 it holds that

supn|fn|exp(cn)=.\sup_{n\to\infty}|f_{n}|\exp\big{(}c\sqrt{n})=\infty.

This fact is not as deep as the two results cited above which imply it, but it is needed in the proof of one of our main results. For this reason, we give an elementary proof in Section 6.1. We mention also that a characterization of density in 𝒦θ\mathcal{K}_{\theta} of functions in 𝒜\mathcal{A}^{\infty}, the algebra of functions analytic in 𝔻\mathbb{D} with all derivatives extending continuously to 𝔻¯\overline{\mathbb{D}}, has been established [22].

The situation is more interesting, and much more difficult to handle, in the general class of (b)\mathcal{H}(b)-spaces. It was proved long ago by Sarason that the set 𝒫\mathcal{P} of analytic polynomials is contained and dense in (b)\mathcal{H}(b) if and only if bb is a non-extreme point of the unit ball of HH^{\infty}, a condition characterized by

(1.11) 𝕋log(Δb)𝑑m>,\int_{\mathbb{T}}\log(\Delta_{b})\,d\textit{m}>-\infty,

where

Δb:=1|b|2.\Delta_{b}:=\sqrt{1-|b|^{2}}.

In terms of core sets, this result can be stated as follows, and a proof can be found in [28].

Theorem (Sarason).

Let b:𝔻𝔻b:\mathbb{D}\to\mathbb{D} be an analytic function. The following three statements are equivalent.

  1. (i)

    The analytic polynomials are dense in (b)\mathcal{H}(b).

  2. (ii)

    The function bb is a non-extreme point of the unit ball of HH^{\infty}.

  3. (iii)

    We have the set equality core(Δb)=𝕋\text{core}(\Delta_{b})=\mathbb{T}.

Since these conditions are very restrictive, it is tempting to make an effort to capture a larger class of symbols bb for which (b)\mathcal{H}(b) contains a dense subset of functions in some nice regularity class which is strictly larger than 𝒫\mathcal{P}. The article [23] connects the approximation problem in (b)\mathcal{H}(b) with the structure of MzM_{z}-invariant subspaces of 𝒫2(μ)\mathcal{P}^{2}(\mu), and [21] refines the method to prove the density of 𝒜(b)\mathcal{A}^{\infty}\cap\mathcal{H}(b) for a large class of symbols bb. The method from [23] is very general and applies to a wide range of approximation problems in (b)\mathcal{H}(b). In particular, it applies to approximations by functions in the class (RSD). Since our structural results in \threfCyclicityMainTheorem and \threfPermanenceMainTheorem are definitive, we can prove also a definitive result on existence and density of functions f(b)f\in\mathcal{H}(b) which satisfy (RSD). In fact, we will prove a much stronger (and optimal) result.

In order to state our result, we will need to quantify the spectral decay of a function ff by a condition of the type (RSD) but with exp(cn)\exp\big{(}c\sqrt{n}\big{)} replaced by faster increasing sequences. To this end, we define below in \threfAdmissibleSequenceDef the admissible sequences M={Mn}n0M=\{M_{n}\}_{n\geq 0}. These sequences are logarithmically convex (at least eventually, for large nn) and are decreasing to zero at least as fast as exp(cn)\exp\big{(}-c\sqrt{n}\big{)}, but satisfy a condition of the form

n0log1/Mn1+n2<\sum_{n\geq 0}\frac{\log 1/M_{n}}{1+n^{2}}<\infty

which prohibits, for instance, their exponentially fast decay.

Example 1.7.

The sequence defined by

(1.12) Mn=exp(cn(log(n)+1)p),n1M_{n}=\exp\Bigg{(}-c\frac{n}{(\log(n)+1)^{p}}\Bigg{)},\quad n\geq 1

is admissible for every p>1p>1 and c>0c>0, but it is not admissible for p=1p=1 and any c>0c>0.

Theorem D.
\thlabel

MainTheoremHbExistenceESD Let b:𝔻𝔻b:\mathbb{D}\to\mathbb{D} be an analytic function. The following three statements are equivalent.

  1. (i)

    The space (b)\mathcal{H}(b) contains a non-zero function ff which satisfies (RSD).

  2. (ii)

    For any admissible sequence {Mn}n0\{M_{n}\}_{n\geq 0}, the space (b)\mathcal{H}(b) contains a non-zero function f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} which satisfies

    (1.13) supn0|fn|Mn<\sup_{n\geq 0}\,\frac{|f_{n}|}{M_{n}}<\infty
  3. (iii)

    The function bb vanishes at some point λ𝔻\lambda\in\mathbb{D}, or there exists an arc I𝕋I\subset\mathbb{T} of positive length for which

    IlogΔbdm>.\int_{I}\log\Delta_{b}\,d\textit{m}>-\infty.

In (iii)(iii), the condition of vanishing of bb at some λ𝔻\lambda\in\mathbb{D} is the uninteresting case, sice then (b)\mathcal{H}(b) contains a rational function with no poles on 𝔻¯\overline{\mathbb{D}}. For such a function (ii)(ii) holds trivially.

To reach \threfMainTheoremHbExistenceESD we only really need the characterization of irreducibility of 𝒫2(μ)\mathcal{P}^{2}(\mu). Proof of the next theorem requires the full strength of the invariant subspace results developed in the first part of this article.

Theorem E.
\thlabel

MainTheoremHbDensityESD Let b:𝔻𝔻b:\mathbb{D}\to\mathbb{D} be an analytic function, and b=BSνUb=BS_{\nu}U be the inner-outer factorization of bb. The following three statements are equivalent.

  1. (i)

    The set of functions ff in (b)\mathcal{H}(b) which satisfy (RSD) is dense in (b)\mathcal{H}(b).

  2. (ii)

    For any admissible sequence {Mn}n0\{M_{n}\}_{n\geq 0}, the set of functions ff in (b)\mathcal{H}(b) which satisfy

    (1.14) supn0|fn|Mn<\sup_{n\geq 0}\,\frac{|f_{n}|}{M_{n}}<\infty

    is dense in (b)\mathcal{H}(b).

  3. (iii)

    The set core(Δb)\text{core}(\Delta_{b}) is a carrier for Δb\Delta_{b} and for the singular measure ν\nu.

Example 1.8.

For instance, by applying our theorem to the admissible sequence (1.12) for any p>1p>1, we get that the density in (b)\mathcal{H}(b) of functions f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} satisfying

limn0|fn|exp(cnα)=0\lim_{n\geq 0}\,|f_{n}|\exp(cn^{\alpha})=0

simultaneously for any c>0c>0 and any α(0,1)\alpha\in(0,1), is equivalent to condition (iii)(iii) in \threfMainTheoremHbDensityESD. Roughly speaking, functions satisfying such decay a condition just barely fail to have an analytic continuation to a disk larger than 𝔻\mathbb{D}.

Example 1.9.
\thlabel

HbEsetExample Generalizing the setting of \threfPointMassExample, we may replace a point by a general closed subset EE of 𝕋\mathbb{T}, and define the outer function b0:𝔻𝔻b_{0}:\mathbb{D}\to\mathbb{D} by specifying its modulus |b0(x)||b_{0}(x)|, x𝕋x\in\mathbb{T}, to satisfy the equation

1|b0(x)|2=Δb0(x):=12exp(1dist(x,E))\sqrt{1-|b_{0}(x)|^{2}}=\Delta_{b_{0}}(x):=\frac{1}{2}\exp\Bigg{(}-\frac{1}{\text{dist}(x,E)}\Bigg{)}

for x𝕋Ex\in\mathbb{T}\setminus E, where dist(x,E)\text{dist}(x,E) is the Euclidean distance from the point xx to the closed set EE, and |b0(x)|=1|b_{0}(x)|=1 for xEx\in E. We can easily check that

core(Δb0)=𝕋E.\text{core}(\Delta_{b_{0}})=\mathbb{T}\setminus E.

If BB is a Blaschke product and SνS_{\nu} is a singuler inner function, then functions of rapid spectral decay will be dense in the space (b)\mathcal{H}(b), with b:=BSνb0b:=BS_{\nu}b_{0}, if and only if ν(E)=0\nu(E)=0.

Our proof of \threfMainTheoremHbDensityESD depends crucially on \threfCyclicityMainTheorem and \threfPermanenceMainTheorem, but is otherwise similar to the proofs in [21] and [23]. However, in the present work we obtain new information on which functions in (b)\mathcal{H}(b) fail to be approximable by classes appearing in \threfMainTheoremHbDensityESD. These results are presented in Section 7.

We mentioned earlier that our result is optimal. This is morally true, in the following sense. Assume that M={Mn}n0M=\{M_{n}\}_{n\geq 0} is a logarithmically convex sequence which is not admissible according to \threfAdmissibleSequenceDef, because we have

(1.15) n0logMn1+n2=.\sum_{n\geq 0}\frac{\log M_{n}}{1+n^{2}}=-\infty.

For instance, MM could be defined by (1.12) for p=1p=1. If Volberg or Kriete and MacCluer were interested in approximations in (b)\mathcal{H}(b)-spaces, they would have proved the following theorem by a use of their techniques in [20] and [30].

Theorem (Volberg, Kriete-MacCluer).

Let M={Mn}n0M=\{M_{n}\}_{n\geq 0} be a logarithmically convex sequence satisfying the property (1.15). The following two statements are equivalent.

  1. (i)

    The space (b)\mathcal{H}(b) contains a non-zero function ff which satisfies

    supn0|fn|Mn<.\sup_{n\geq 0}\,\frac{|f_{n}|}{M_{n}}<\infty.
  2. (ii)

    The function bb vanishes at some point λ𝔻\lambda\in\mathbb{D}, or bb is non-extreme.

We do not prove the above theorem in the present article. Its proof is completely analogous to the proof of \threfMainTheoremHbExistenceESD. The difference consists merely of a use of theorems and observations of Volberg and Kriete-MacCluer from the above mentioned papers, instead of main theorem of [24] as we do here in the proof of \threfMainTheoremHbExistenceESD.

It follows that the investigation of existence and approximability properties in (b)\mathcal{H}(b) of functions with spectral decay satisfying at least (RSD) is essentially completed in \threfMainTheoremHbExistenceESD and \threfMainTheoremHbDensityESD.

1.6. Additional comments, questions and conjectures

1.6.1. Work of McCarthy and Davis

The class of functions satisfying (RSD) has already appeared in the theory of de Branges-Rovnyak spaces. In [7], McCarthy and Davis showed that a function hh satisfies (RSD) if and only if the multiplication operator MhM_{h} acts boundedly on (b)\mathcal{H}(b) for all non-extreme symbols bb. In particular, this means that every space (b)\mathcal{H}(b) defined by a non-extreme symbol bb contains all functions satisfying (RSD). Our \threfMainTheoremHbExistenceESD then establishes a converse statement: a characterization of bb for which (b)\mathcal{H}(b) contains no non-zero such functions.

1.6.2. Relation to Khrushchev’s results

Khrushchev in [16] studied a problem similar to one appearing in \threfUncertThmRSD. If 1E1_{E} is the characteristic function of a set EE contained in 𝕋\mathbb{T}, and there exists a function gg living only on EE such that 𝒞g\mathcal{C}_{g} in (1.7) has some regularity properties, then what can be said about EE? Khrushchev used the phrase removal of singularities of Cauchy integrals in the context of his study of nowhere dense E𝕋E\subset\mathbb{T} which support a function gg with a smooth Cauchy integral 𝒞g\mathcal{C}_{g}. Thus ”removing” the singularities of the irregular set EE. His solution is given in terms of Beurling-Carleson sets. The weighted version of the problem replaces 1E1_{E} by a general weight ww.

In turn, \threfMainTheoremHbDensityESD can be seen as a solution to the problem of removal of singularities of normalized Cauchy integrals in context of the class (RSD), where the possible existence of a singular part of the measure ν\nu forces the normalization. Indeed, given a positive finite Borel measure ν\nu on 𝕋\mathbb{T}, we may ask if the space L2(ν)L^{2}(\nu) contains a dense subset of functions f~\tilde{f} for which the normalized Cauchy integral in (1.9) (with νb\nu_{b} replaced by ν\nu) satisfies (RSD). The condition, in terms of the associated function b=bνb=b_{\nu} given by (1.10), is given in (iii)(iii) of \threfMainTheoremHbDensityESD. In this context, it would be of interest to characterize intrinsically the measures ν\nu which correspond to bb satisfying condition (iii)(iii) of \threfMainTheoremHbDensityESD.

Question 1.

Let b:𝔻𝔻b:\mathbb{D}\to\mathbb{D} be an analytic functions which satisfies the condition (iii)(iii) in \threfMainTheoremHbDensityESD. Can we describe the structure of the corresponding Aleksandrov-Clark measure νb\nu_{b} of bb appearing in formula (1.10) ?

1.6.3. Logarithmic convexity of admissible sequences

In spite of some efforts, the author has not been able to remove the assumption of logarithmic convexity in \threfAdmissibleSequenceDef. Surely the most interesting admissible sequences, such as (1.12), do satisfy such a conditon, but ideally one would like to remove this assumption. Logarithmic convexity of {Mn}n0\{M_{n}\}_{n\geq 0} plays its part in the proof of \threfAdmissibleSequenceLemma. In relation to that, we would like to answer the following question.

Question 2.

If c(x)c(x), x>0x>0, is an increasing, positive and continuous function which satisfies

(1.16) 1c(x)x2𝑑x<,\int_{1}^{\infty}\frac{c(x)}{x^{2}}\,dx<\infty,

then under what additional conditons on cc may we replace c(x)c(x) in (1.16) by its least concave majorant?

Any interesting condition on cc which guarantees the above integrability property of its concave majorant will lead to slighly improved versions of our theorems.

1.6.4. Non-integrability of logG\log G as a sharp condition

Consider the condition

(LogInt) 0dlog(1/G(x))𝑑x<\int_{0}^{d}\log(1/G(x))\,dx<\infty

for some d>0d>0. The condition (ExpDec) implies that our considered functions G(x)G(x) will always fail to satisfy (LogInt). In fact, it is (at least in the mind of the author) reasonable to conjecture that several of the results of this article should have sharp improvements in which the requirement for GG to satisfy (ExpDec) is replaced by the requirement for GG not to satisfy (LogInt). This condition is, in turn, equivalent to the statement that

n0(log1/Mn(G))21+n2=,\sum_{n\geq 0}\frac{\big{(}\log 1/M_{n}(G)\big{)}^{2}}{1+n^{2}}=\infty,

where {Mn(G)}\{M_{n}(G)\} is defined in (5.3) and is the sequence of moments of the function GG. This equivalence can be deduced using techniques appearing in Section 5 below. The above condition appears in [10] as a necessary and sufficient condition for all singular inner functions to be cyclic in a space 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}) with dμ(z)=G(1|z|)dA(z)d\mu(z)=G(1-|z|)dA(z), and so w=0w=0 in contrast to the situation dealt with in the present article.

For instance, a sharp version of the irreducibility of 𝒫2(μ)\mathcal{P}^{2}(\mu) with μ\mu of the form (1.1) would follow if we could prove the following statement.

Conjecture 1.
\thlabel

conj1 Assume that GG fails to satisfy (LogInt) and w1(𝕋)w\in\mathcal{L}^{1}(\mathbb{T}) is a non-negative weight on 𝕋\mathbb{T}. If 𝒫2(μ)\mathcal{P}^{2}(\mu) of the form (1.1) is a space of analytic functions on 𝔻\mathbb{D}, then the set core(w)\text{core}(w) of the weight w1(𝕋)w\in\mathcal{L}^{1}(\mathbb{T}) is a carrier for ww.

Given this result, one could attempt to combine our techniques appearing in Section 4 and those of El-Fallah, Kellay and Seip from [10] to prove the following strong version of both their result and our \threfCyclicityMainTheorem.

Conjecture 2.

In the setting of \threfconj1, a singular inner function SνS_{\nu} is cyclic in the space of analytic functions 𝒫2(μ)\mathcal{P}^{2}(\mu) if and only if ν(core(w))=0\nu\big{(}\text{core}(w)\big{)}=0.

In relation to \threfUncertThmRSD, we expect the following improvement.

Conjecture 3.

The conclusion of \threfUncertThmRSD can be reached if

𝒞ν(z)=n0νnzn\mathcal{C}_{\nu}(z)=\sum_{n\geq 0}\nu_{n}z^{n}

merely satisfies

supn0|νn|Mn<\sup_{n\geq 0}\,\frac{|\nu_{n}|}{M_{n}}<\infty

for some (say, logarithmically convex) sequence {Mn}n0\{M_{n}\}_{n\geq 0} satisfying

n0(log1/Mn)21+n2=.\sum_{n\geq 0}\frac{\big{(}\log 1/M_{n}\big{)}^{2}}{1+n^{2}}=\infty.

One can show, by considerations of examples, that all of the above conjectures imply sharp results.

1.7. Outline of the rest of the article

Section 2 deals with construction of special domains which look like wizard hats and which support very large positive harmonic functions. We prove \threfPermanenceMainTheorem and \threfCyclicityMainTheorem in Sections 3 and 4, respectively. Proof of \threfPermanenceMainTheorem relies heavily on results of Section 2. The second part of the article starts in Section 5. There we deal with some preparatory estimates on moments sequences which are needed later. \threfMainTheoremHbExistenceESD and \threfMainTheoremHbDensityESD are proved in Sections 6 and 7. The techniques used in these sections come from [23], but we refine some of the methods and prove auxilliary results of hopefully independent interest. Finally, in Section 8, we prove \threfUncertThmRSD.

1.8. Some notation

For a measure μ\mu on 𝔻¯\overline{\mathbb{D}} we will denote by μ𝔻\mu_{\mathbb{D}} and μ𝕋\mu_{\mathbb{T}} its restriction to 𝔻\mathbb{D} and 𝕋\mathbb{T}, respectively. In some contexts we will also use the same notations μ𝔻\mu_{\mathbb{D}} and μ𝕋\mu_{\mathbb{T}} to emphasize that the considered measure lives only on 𝔻\mathbb{D} or 𝕋\mathbb{T}. The area measure dAdA will always be normalized by the condition A(𝔻)=1A(\mathbb{D})=1, and a similar convention will be used also for the arc-length measure on the circle: m(𝕋)=1m(\mathbb{T})=1. We let log+(x)=max(0,logx)\log^{+}(x)=\max\big{(}0,\log x\big{)}.

The symbol μ\|\cdot\|_{\mu} always denotes the usual 2(μ)\mathcal{L}^{2}(\mu)-norm corresponding to the finite positive Borel measure μ\mu. For a set E𝕋E\subset\mathbb{T}, we sometimes use the shorter notation 2(E)\mathcal{L}^{2}(E) to denote the space of functions on 𝕋\mathbb{T} which vanish outside of EE and are square-integrable with respect to the Lebesgue measure m. The notation ,\big{\langle}\cdot,\cdot\big{\rangle} denotes different kinds of duality pairings between spaces. By ,2\big{\langle}\cdot,\cdot\big{\rangle}_{\mathcal{L}^{2}} we will denote the standard inner product in L2(𝕋)L^{2}(\mathbb{T}).

The operator P+:2(𝕋)H2P_{+}:\mathcal{L}^{2}(\mathbb{T})\to H^{2} is the orthogonal projection onto the Hardy space H2H^{2}. For a bounded analytic function hh, the notation Th¯:H2H2T_{\overline{h}}:H^{2}\to H^{2} stands for the co-analytic Toeplitz operator with symbol hh, this operator being defined by the formula Th¯f=P+h¯fT_{\overline{h}}f=P_{+}\overline{h}f.

2. Wizard hats and their harmonic measures

The proof of \threfPermanenceMainTheorem relies on a technique of restriction of a convergent sequence of analytic functions to a certain subdomain of 𝔻\mathbb{D}. It is easier to construct the corresponding domain in the setting of a half-plane, and later use a conformal mapping argument. We will work in the upper half-plane \mathbb{H}. There, our domain looks like a wizard’s hat (see Figure 1).

Harmonic measures will play an important role in our discussion, so we start by recalling some basic related notions, and set some further notations. Let Ω\Omega be a Jordan domain in the plane. The domains which will appear in our context have a boundary consisting of a finite union of smooth curves. Let ω(z,E,Ω)\omega(z,E,\Omega) denote the harmonic measure of a segment EE of the boundary Ω\partial\Omega, based at the point zΩz\in\Omega. Then

zω(z,E,Ω),zΩz\mapsto\omega(z,E,\Omega),\quad z\in\Omega

is a positive harmonic function in Ω\Omega which extends continuously to the boundary Ω\partial\Omega except at the endpoints of EE. It attains the boundary value 11 at the relative interior of EE, and boundary value 0 on ΩE\partial\Omega\setminus E. Let (Ω)\mathcal{B}(\partial\Omega) denote the Borel σ\sigma-algebra on Ω\partial\Omega. For each fixed z0Ωz_{0}\in\Omega, the mapping

Aω(z0,A,Ω),A(Ω)A\mapsto\omega(z_{0},A,\Omega),\quad A\in\mathcal{B}(\partial\Omega)

defines a positive Borel probability measure on Ω\partial\Omega. The reader can consult the excellent books by Garnett and Marshall [14] and by Ransford [26] for more background and other basic facts about harmonic measures which are used in this section.

Let

={z=x+iy:y>0}\mathbb{H}=\{z=x+iy\in\mathbb{C}:y>0\}

denote the upper half-plane of \mathbb{C}. The main efforts of this section will go into estimation of the harmonic measure on a wizard hat domain WW. The domain is constructed from an interval II\subset\mathbb{R} and a profile function p(x)p(x), x0x\geq 0, which by our definition is increasing, positive and continuous, smooth (say, continuously differentiable) for x>0x>0, and which satisfies p(0)=0p(0)=0. Given a profile function pp and an interval I=(a,b)I=(a,b), we define the wizard hat WW to be the simply connected Jordan domain

(2.1) W=W(p,I):={z=x+iy:xI,y<min[p(xa),p(bx)]}.W=W(p,I):=\Big{\{}z=x+iy\in\mathbb{H}:x\in I,y<\min\big{[}p(x-a),p(b-x)\big{]}\Big{\}}.

The boundary W\partial W is a piecewise smooth curve, with three smooth parts divided by three cusps. An example of a domain WW, constructed from a profile function p(x)=xqp(x)=x^{q} for some q>1q>1, is marked by the shaded area in Figure 1. Our goal is to prove a result regarding existence of harmonic functions which grow rapidly along W=W\partial W\cap\mathbb{H}=\partial W\setminus\mathbb{R}.

Definition 2.1.

(Majorants) \thlabelRegMajorantDef Let d>0d>0 be some positive number. A positive function F:(0,d)[0,)F:(0,d)\to[0,\infty) will be called a majorant if it satisfies the following two properties:

  1. (i)

    F(t)F(t) is a decreasing function of t>0t>0, and limt0+F(t)=+\lim_{t\to 0^{+}}F(t)=+\infty,

  2. (ii)

    0dlogF(t)𝑑t<\int_{0}^{d}\log F(t)\,dt<\infty.

The properties of FF appearing in \threfRegMajorantDef are related to growth estimates on functions in the investigated class of 𝒫2(μ)\mathcal{P}^{2}(\mu)-spaces. See \threfPointEvaluationBoundRegMajorant below.

Proposition 2.2.
\thlabel

WizardHatMainProposition Let II\subset\mathbb{R} be a finite interval and FF be a majorant in the sense of \threfRegMajorantDef. There exists a profile function pp, a wizard hat W=W(p,I)W=W(p,I), and a positive harmonic function uu on WW which extends continuously to the boundary W\partial W except at the two cusps of W\partial W on \mathbb{R}, satisfies u(x)=0u(x)=0 for xx in the interior of II, and u(z)=F(Imz)u(z)=F(\operatorname{Im}z) for zWz\in\partial W\cap\mathbb{H}.

In order to prove \threfWizardHatMainProposition, we will need to estimate the harmonic measure ω(z0,Bt,W)\omega(z_{0},B_{t},W) of the following piece BtB_{t} of the boundary of WW:

(2.2) Bt={z=x+iyW:0<y,a<x<t}.B_{t}=\Big{\{}z=x+iy\in\partial W:0<y,\,a<x<t\Big{\}}.

See Figure 1, where BtB_{t} is marked. A result of Beurling and Ahlfors (see [14, Theorem 6.1 of Chapter IV]) can be applied to the union of WW, II and the reflected domain W¯={z¯:zW}\overline{W}=\{\overline{z}:z\in W\} to obtain a good estimate for the harmonic measure of BtB_{t}.

\includestandalone

[scale=1]WizardHat

Figure 1. The wizard hat WW and a piece BtB_{t} of its boundary.
Proposition 2.3.

(Beurling-Ahlfors estimate) Let θ\theta be a positive continuous function defined on an interval (a,b)(a,b)\subset\mathbb{R}, and let Ω\Omega be the domain

Ω={z=x+iy:|y|<θ(x),a<x<b}.\Omega=\{z=x+iy:|y|<\theta(x),\,a<x<b\Big{\}}.

If z0Ωz_{0}\in\Omega and Sa={zΩ:Rez=a}S_{a}=\{z\in\partial\Omega:\operatorname{Re}z=a\} is the left vertical part of the boundary of Ω\Omega, then

ω(z0,Sa,Ω)8πexp(2πaRez0dxθ(x)).\omega(z_{0},S_{a},\Omega)\leq\frac{8}{\pi}\exp\Bigg{(}-2\pi\int_{a}^{\operatorname{Re}z_{0}}\frac{dx}{\theta(x)}\Bigg{)}.

In Figure 1, the symmetrized domain W~:=WIW¯\widetilde{W}:=W\cup I\cup\overline{W} is bounded by the top part of the boundary of WW and the dotted reflection below the line \mathbb{R}. Let W~t\widetilde{W}_{t} be the domain obtained by cutting W~\widetilde{W} along the cross-section St={zW~:Rez=t}S_{t}=\{z\in\widetilde{W}:\operatorname{Re}z=t\} and keeping the right part of the two resulting pieces. Define WtW_{t} similarly (so that WtW_{t} is the intersection of W~t\widetilde{W}_{t} and \mathbb{H}). The Beurling-Ahlfors estimate immediately implies that

ω(z0,St,W~t)8πexp(2πtRez01p(xa)𝑑x),\omega(z_{0},S_{t},\widetilde{W}_{t})\leq\frac{8}{\pi}\exp\Bigg{(}-2\pi\int_{t}^{\operatorname{Re}z_{0}}\frac{1}{p(x-a)}\,dx\Bigg{)},

where z0Wtz_{0}\in W_{t} is as in Figure 1. By a comparison of the values on Wt\partial W_{t} of the two harmonic functions ω(z,Bt,W)\omega(z,B_{t},W) and ω(z,St,W~t)\omega(z,S_{t},\widetilde{W}_{t}), and the maximum principle for harmonic functions, we get the inequality

(2.3) ω(z,Bt,W)ω(z,St,W~t),zWt.\omega(z,B_{t},W)\leq\omega(z,S_{t},\widetilde{W}_{t}),\quad z\in W_{t}.

In particular, this holds at z0z_{0}. We have obtained the following harmonic measure estimation.

Proposition 2.4.
\thlabel

BtHarmEst Let W=W(p,I)W=W(p,I) be the wizard hat given by (2.1), BtB_{t} the piece of its boundary given by (2.2) and z0Wz_{0}\in W. Then

ω(z0,Bt,W)8πexp(2πtRez01p(xa)𝑑x)\omega(z_{0},B_{t},W)\leq\frac{8}{\pi}\exp\Bigg{(}-2\pi\int_{t}^{\operatorname{Re}z_{0}}\frac{1}{p(x-a)}\,dx\Bigg{)}

whenever a<t<Rez0a<t<\operatorname{Re}z_{0}.

Given a majorant FF as in \threfRegMajorantDef, we will now show how to construct a profile function pp and harmonic function uu which satisfies the properties stated in \threfWizardHatMainProposition. Without loss of generality, we may assume that I=(0,2)I=(0,2). For some large integer n0>0n_{0}>0, let

(2.4) αn:=2nn0,n1.\alpha_{n}:=2^{-n-n_{0}},\quad n\geq 1.

We define also the sequence

(2.5) γn:=αnlogF(αn),n1\gamma_{n}:=\alpha_{n}\log F(\alpha_{n}),\quad n\geq 1

This sequence is positive if the integer n0n_{0} in (2.4) is chosen large enough. Next, we make the following simple observation.

Lemma 2.5.
\thlabel

GammaSeqLemma For any ϵ>0\epsilon>0, there exists an integer n0>0n_{0}>0 such that, with {αn}n1\{\alpha_{n}\}_{n\geq 1} defined by (2.4) and {γn}n1\{\gamma_{n}\}_{n\geq 1} defined by (2.5), we have

n=1γn<ϵ.\sum_{n=1}^{\infty}\gamma_{n}<\epsilon.
Proof.

Since F(t)F(t) is a majorant, by part (i)(i) of \threfRegMajorantDef we have

0α1logF(t)𝑑t\displaystyle\int_{0}^{\alpha_{1}}\log F(t)\,dt =n=1αn+1αnlogF(t)𝑑t\displaystyle=\sum_{n=1}^{\infty}\int_{\alpha_{n+1}}^{\alpha_{n}}\log F(t)\,dt
n=1logF(αn)(αnαn+1)\displaystyle\geq\sum_{n=1}^{\infty}\log F(\alpha_{n})(\alpha_{n}-\alpha_{n+1})
=n=1logF(αn)αn+1\displaystyle=\sum_{n=1}^{\infty}\log F(\alpha_{n})\alpha_{n+1}
=12n=1γn,\displaystyle=\frac{1}{2}\sum_{n=1}^{\infty}\gamma_{n},

where we used that αnαn+1=αn+1=αn/2\alpha_{n}-\alpha_{n+1}=\alpha_{n+1}=\alpha_{n}/2. Now, by property (ii)(ii) in \threfRegMajorantDef we have

limα10+0α1logF(t)𝑑t=0\lim_{\alpha_{1}\to 0^{+}}\int_{0}^{\alpha_{1}}\log F(t)\,dt=0

and so our claim follows. ∎

We set n0n_{0} to some value which ensures that

(2.6) n=1γn<1/2,\sum_{n=1}^{\infty}\gamma_{n}<1/2,

or in other words, the sum n=1γn\sum_{n=1}^{\infty}\gamma_{n} is less than one quarter of the length of the interval I=(0,2)I=(0,2). Further, we let {tn}n1\{t_{n}\}_{n\geq 1} be a sequence of positive numbers starting with

t1=1,t_{1}=1,

which tends monotonically to 0. We shall soon define {tn}n1\{t_{n}\}_{n\geq 1} by specifying the sequence of differences {Δtn}n1\{\Delta t_{n}\}_{n\geq 1}, where

Δtn:=tntn+1,n1.\Delta t_{n}:=t_{n}-t_{n+1},\quad n\geq 1.

The differences Δtn\Delta t_{n} are positive numbers, and t2,t3,t_{2},t_{3},\ldots will be recursively defined in terms of those differences by the relations

t2=t1Δt1,t3=t2Δt2,t_{2}=t_{1}-\Delta t_{1},\,t_{3}=t_{2}-\Delta t_{2},

and so on. In order for so defined sequence {tn}n1\{t_{n}\}_{n\geq 1} to converge to zero it is necessary and sufficient that

(2.7) n=1Δtn=1,\sum_{n=1}^{\infty}\Delta t_{n}=1,

a requirement which we will later ensure. Given any {tn}n1\{t_{n}\}_{n\geq 1} as above, a profile function pp may be readily constructed which satisfies

(2.8) p(tn)=αn,n1.p(t_{n})=\alpha_{n},\quad n\geq 1.

Indeed, since the sequence {tn}n1\{t_{n}\}_{n\geq 1} is assumed to be monotonically decreasing to zero, the function pp can be chosen to be smooth, increasing and positive for t>0t>0, and satisfy p(0)=0p(0)=0. A proper choice of {tn}n1\{t_{n}\}_{n\geq 1} will produce a wizard hat with our desired properties. Assume that {tn}n1\{t_{n}\}_{n\geq 1} has been given, let W=W(p,I)W=W(p,I) be the corresponding wizard hat, and ω()=ω(z0,,W)\omega(\cdot)=\omega(z_{0},\cdot,\partial W) be the harmonic measure at some point z0=1+y0iWz_{0}=1+y_{0}i\in W which lies on the symmetry line of WW. Let u~(z)\tilde{u}(z) be defined on W\partial W by

(2.9) u~(z)={F(Imz),zW,0,zW.\tilde{u}(z)=\begin{cases}F(\operatorname{Im}z),&z\in\partial W\cap\mathbb{H},\\ 0,&z\in\partial W\cap\mathbb{R}.\end{cases}

We will ensure that u~1(ω)\tilde{u}\in\mathcal{L}^{1}(\omega). Since FF is decreasing, the definition of WW shows that for any n1n\geq 1 the values of the function u~\tilde{u} on the arc BtnBtn+1B_{t_{n}}\setminus B_{t_{n+1}} are dominated by its value at the point zBtnBtn+1z\in B_{t_{n}}\setminus B_{t_{n+1}} which lies closest to the real line \mathbb{R}, i.e., at the point z=tn+1+ip(tn+1)z=t_{n+1}+ip(t_{n+1}). In other words, we have

(2.10) supzBtnBtn+1u~(z)=F(p(tn+1)).\sup_{z\in B_{t_{n}}\setminus B_{t_{n+1}}}\tilde{u}(z)=F\big{(}p(t_{n+1})\big{)}.

Moreover, from positivity and monotonicity of pp, and from \threfBtHarmEst, we deduce the estimate

ω(BtnBtn+1)ω(Btn)\displaystyle\omega\big{(}B_{t_{n}}\setminus B_{t_{n+1}}\big{)}\leq\omega(B_{t_{n}}) 8πexp(2πtntn11p(x)𝑑x)\displaystyle\leq\frac{8}{\pi}\exp\Bigg{(}-2\pi\int_{t_{n}}^{t_{n-1}}\frac{1}{p(x)}dx\Bigg{)}
(2.11) 8πexp(2πΔtn1p(tn1))\displaystyle\leq\frac{8}{\pi}\exp\Bigg{(}-2\pi\frac{\Delta t_{n-1}}{p(t_{n-1})}\Bigg{)}

which holds for n2n\geq 2.

In order to have u~1(ω)\tilde{u}\in\mathcal{L}^{1}(\omega) it is sufficient to ensure that Bt2u~(z)𝑑ω(z)<\int_{B_{t_{2}}}\tilde{u}(z)d\omega(z)<\infty. To this end, we use (2.10) and (2) to estimate

Bt2u~(z)𝑑ω(z)\displaystyle\int_{B_{t_{2}}}\tilde{u}(z)d\omega(z) =n=2BtnBtn+1u~(z)𝑑ω(z)\displaystyle=\sum_{n=2}^{\infty}\int_{B_{t_{n}}\setminus B_{t_{n+1}}}\tilde{u}(z)d\omega(z)
8πn=2F(p(tn+1))exp(2πΔtn1p(tn1))\displaystyle\leq\frac{8}{\pi}\sum_{n=2}^{\infty}F\big{(}p(t_{n+1})\big{)}\exp\Bigg{(}-2\pi\frac{\Delta t_{n-1}}{p(t_{n-1})}\Bigg{)}
=8πn=2exp(logF(p(tn+1))2πΔtn1p(tn1))\displaystyle=\frac{8}{\pi}\sum_{n=2}^{\infty}\exp\Bigg{(}\log F\big{(}p(t_{n+1})\big{)}-2\pi\frac{\Delta t_{n-1}}{p(t_{n-1})}\Bigg{)}
=8πn=2exp(1p(tn+1)(logF(p(tn+1))p(tn+1)2πΔtn1p(tn+1)p(tn1)))\displaystyle=\frac{8}{\pi}\sum_{n=2}^{\infty}\exp\Bigg{(}\frac{1}{p(t_{n+1})}\Big{(}\log F\big{(}p(t_{n+1})\big{)}p(t_{n+1})-2\pi\Delta t_{n-1}\frac{p(t_{n+1})}{p(t_{n-1})}\Big{)}\Bigg{)}
(2.12) =8πn=2exp(2n+1+n0(γn+1πΔtn12)).\displaystyle=\frac{8}{\pi}\sum_{n=2}^{\infty}\exp\Bigg{(}2^{n+1+n_{0}}\Big{(}\gamma_{n+1}-\frac{\pi\Delta t_{n-1}}{2}\Big{)}\Bigg{)}.

In the last step we used (2.4), (2.5) and (2.8). We may now specify the values of {tn}n1\{t_{n}\}_{n\geq 1} by setting the values of the differences:

(2.13) Δtn1=An2+2πγn+1,n2.\Delta t_{n-1}=\frac{A}{n^{2}}+\frac{2}{\pi}\gamma_{n+1},\quad n\geq 2.

for an appropriate constant A>0A>0 which ensures the necessary summation condition (2.7). This can be done, since

n=22πγn+1<1/π<1\sum_{n=2}\frac{2}{\pi}\gamma_{n+1}<1/\pi<1

by (2.6). We obtain from (2) that

Bt2u~(z)𝑑ω(z)8πn=2exp(Aπ22n+1+n0n2)<.\int_{B_{t_{2}}}\tilde{u}(z)d\omega(z)\leq\frac{8}{\pi}\sum_{n=2}^{\infty}\exp\Big{(}-\frac{A\pi}{2}\frac{2^{n+1+n_{0}}}{n^{2}}\Big{)}<\infty.

Consequently, with this definition of {tn}n1\{t_{n}\}_{n\geq 1} and a corresponding profile function pp, we have that u~1(ω)\tilde{u}\in\mathcal{L}^{1}(\omega).

We may now complete the proof of \threfWizardHatMainProposition. Let WW and pp be chosen as above, and ϕ:𝔻W\phi:\mathbb{D}\to W be a conformal mapping which maps 0𝔻0\in\mathbb{D} to z0Wz_{0}\in W. Since WW is a Jordan domain, ϕ\phi extends to a homeomorphism between 𝔻𝕋\mathbb{D}\cup\mathbb{T} and WWW\cup\partial W. If v:𝕋[0,)v:\mathbb{T}\to[0,\infty) is defined by v=u~ϕv=\tilde{u}\circ\phi, then a change of variables shows that

𝕋v𝑑m=Wu~𝑑ω<.\int_{\mathbb{T}}v\,d\textit{m}=\int_{\partial W}\tilde{u}\,d\omega<\infty.

Since u~\tilde{u} is continuous on all of W\partial W except at the two cusps of W\partial W on \mathbb{R}, the function vv is continuous on 𝕋\mathbb{T} except at the two points which map under ϕ\phi to the cusps. We verified above that v1(𝕋)v\in\mathcal{L}^{1}(\mathbb{T}), so we may extend vv to 𝔻\mathbb{D} by means of its Poisson integral. This extension is continuous in 𝔻𝕋\mathbb{D}\cup\mathbb{T} except at the two points corresponding to the cusps. If we define uu in WW by u=vϕ1u=v\circ\phi^{-1}, then uu is the harmonic function sought in \threfWizardHatMainProposition.

3. Proper invariant subspaces generated by singular inner functions

The goal of this section is to prove \threfPermanenceMainTheorem.

3.1. Technical lemmas

Similarly to Section 2, we prove the next lemma in the upper half-plane ={z:Imz>0}\mathbb{H}=\{z\in\mathbb{C}:\operatorname{Im}z>0\}. This is done, again, only for convenience. An elementary conformal mapping argument will carry the result over to the intended domain 𝔻\mathbb{D}.

In this section, the Lebesgue measure (length measure) on \mathbb{R} will be denoted by dtdt, and the dtdt-measure of a set AA will be denoted by |A||A|, similar to lengths of sets on the circle 𝕋\mathbb{T} (this should not cause confusion). The algebra of bounded analytic functions in \mathbb{H} will be denoted by H()H^{\infty}(\mathbb{H}). In the proofs below we shall use some basic facts regarding H()H^{\infty}(\mathbb{H}), and in particular some factorization results. An exposition of the relevant background can be found in [9, Chapter 11], [13, Chapter II] or [17, Chapter VI].

Every function hH()h\in H^{\infty}(\mathbb{H}) admits an inner-outer factorization into

(3.1) h(z)=ceiazB(z)Sν(z)U(z),z.h(z)=ce^{iaz}B(z)S_{\nu}(z)U(z),\quad z\in\mathbb{H}.

Here cc is some unimodular constant, a0a\geq 0, BB is a Blaschke product given by

B(z)=(ziz+i)mh(α)=0,αiiα¯iαzαzα¯,zB(z)=\Big{(}\frac{z-i}{z+i}\Big{)}^{m}\prod_{\begin{subarray}{c}h(\alpha)=0,\\ \alpha\neq i\end{subarray}}\frac{i-\overline{\alpha}}{i-\alpha}\cdot\frac{z-\alpha}{z-\overline{\alpha}},\quad z\in\mathbb{H}

where mm is a non-negative integer, SνhS_{\nu_{h}} is a singular inner function given by

Sνh(z)=exp(1iπ(1+tz)(tz)dνh(t)(1+t2)),zS_{\nu_{h}}(z)=\exp\Bigg{(}-\frac{1}{i\pi}\int_{\mathbb{R}}\frac{(1+tz)}{(t-z)}\frac{d\nu_{h}(t)}{(1+t^{2})}\Bigg{)},\quad z\in\mathbb{H}

where νh\nu_{h} is a singular positive Borel measure on \mathbb{R}, and UU is the outer function given by

U(z)=exp(1iπ(1+tz)log|h(t)|(tz)(1+t2)𝑑t),z.U(z)=\exp\Bigg{(}\frac{1}{i\pi}\int_{\mathbb{R}}\frac{(1+tz)\log|h(t)|}{(t-z)(1+t^{2})}\,dt\Bigg{)},\quad z\in\mathbb{H}.

The measures (1+t2)1dνh(t)(1+t^{2})^{-1}d\nu_{h}(t) and (1+t2)1log|h(t)|dt(1+t^{2})^{-1}\log|h(t)|\,dt appearing in the integrals above are both finite. It follows from this factorization that we have

(3.2) log|h(z)|\displaystyle\log|h(z)| =αy+log|B(z)|\displaystyle=-\alpha y+\log|B(z)|
1πy(xt)2+y2𝑑νh(t)\displaystyle-\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{(x-t)^{2}+y^{2}}d\nu_{h}(t)
+1πy(xt)2+y2log|h(t)|dt,z=x+iy\displaystyle+\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{(x-t)^{2}+y^{2}}\log|h(t)|dt,\quad z=x+iy\in\mathbb{H}

The last two terms in (3.2) represent the Poisson integrals 𝒫νh\mathcal{P}_{\nu_{h}} and 𝒫log|h|\mathcal{P}_{\log|h|} of the measure νh\nu_{h} and of the function log|h|\log|h|, respectively.

Lemma 3.1.
\thlabel

IntervalWeakStarConvLemma Let JJ be a finite open interval of \mathbb{R}. With notation as above, as y0+y\to 0^{+}, the restrictions to JJ of the measures log|h(t+iy)|dt\log|h(t+iy)|dt converge weak-star to the restriction to JJ of the measure log|h(t)|dtdνh(t)\log|h(t)|dt-d\nu_{h}(t).

The lemma follows easily from results presented in de Branges’ book [8, Theorem 3 and Problem 26]. We sketch an argument for the reader’s convenience.

Proof of \threfIntervalWeakStarConvLemma.

If ϕ\phi is any smooth function which is supported on a compact subset of JJ, then by the symmetry of the Poisson kernel, we have

(3.3) ϕ(t)𝒫νh(t+iy)𝑑t=𝒫ϕ(t+iy)𝑑νh(t),\int_{\mathbb{R}}\phi(t)\mathcal{P}_{\nu_{h}}(t+iy)dt=\int_{\mathbb{R}}\mathcal{P}_{\phi}(t+iy)d\nu_{h}(t),

where 𝒫ϕ\mathcal{P}_{\phi} is the Poisson integral of ϕ\phi. Since ϕ\phi is uniformly continuous on \mathbb{R}, 𝒫ϕ(t+iy)ϕ(t)\mathcal{P}_{\phi}(t+iy)\to\phi(t) uniformly in tt as y0+y\to 0^{+}. Moreover, the compact support of ϕ\phi implies that |𝒫ϕ(t+iy)|=𝒪(1/t2)|\mathcal{P}_{\phi}(t+iy)|=\mathcal{O}(1/t^{2}) as |t||t|\to\infty, uniformly in, say, y(0,1)y\in(0,1). Thus expression (3.3) and the finiteness of the measure (1+t2)1dνh(t)(1+t^{2})^{-1}d\nu_{h}(t) implies now that

limy0+ϕ(t)𝒫νh(t+iy)𝑑t=ϕ(t)𝑑νh(t),\lim_{y\to 0^{+}}\int_{\mathbb{R}}\phi(t)\mathcal{P}_{\nu_{h}}(t+iy)dt=\int_{\mathbb{R}}\phi(t)d\nu_{h}(t),

and we have shown that 𝒫νh(t+iy)dt\mathcal{P}_{\nu_{h}}(t+iy)dt converges weak-star to νh\nu_{h} on the interval JJ. By the same argument 𝒫log|h|(t+iy)dt\mathcal{P}_{\log|h|}(t+iy)dt converges weak-star to log|h(t)|dt\log|h(t)|dt on JJ.

We consider now the measures log|B(t+iy)|dt\log|B(t+iy)|dt. Jensen’s inequality for the upper-half plane (see, for instance, [15, p. 35]) implies that

log|B(i+iy)|log|B(t+iy)|π(1+t2)𝑑t.\log|B(i+iy)|\leq\int_{\mathbb{R}}\frac{\log|B(t+iy)|}{\pi(1+t^{2})}dt.

Thus by letting yy tend to 0+0^{+}, we obtain

log|B(i)|lim infy0+log|B(t+iy)|π(1+t2)𝑑t0.\log|B(i)|\leq\liminf_{y\to 0^{+}}\int_{\mathbb{R}}\frac{\log|B(t+iy)|}{\pi(1+t^{2})}dt\leq 0.

The last inequality is trivial, since log|B|\log|B| is negative in \mathbb{H}. For a finite Blaschke product B0B_{0}, the limit between the inequalities above is certainly equal to 0. Thus we obtain

log|B(i)|log|B0(i)|lim infy0+log|B(t+iy)|π(1+t2)𝑑t0.\log|B(i)|-\log|B_{0}(i)|\leq\liminf_{y\to 0^{+}}\int_{\mathbb{R}}\frac{\log|B(t+iy)|}{\pi(1+t^{2})}dt\leq 0.

Now let B0B_{0} tend to BB through a sequence of finite partial products of BB to obtain

limy0+log|B(t+iy)|π(1+t2)𝑑t=0.\lim_{y\to 0^{+}}\int_{\mathbb{R}}\frac{\log|B(t+iy)|}{\pi(1+t^{2})}dt=0.

This says that the restriction to JJ of log|B(t+iy)|dt\log|B(t+iy)|dt converges to 0 even in variation norm.

The expression (3.2) now implies the weak-star convergence result we are seeking.

Definition 3.2.

(Uniform absolute continuity) \thlabeluniAbsContDef If {fndt}n1\{f_{n}\,dt\}_{n\geq 1} is a sequence of non-negative absolutely continuous Borel measures on \mathbb{R} and II\subset\mathbb{R} is an interval, then we will say that the sequence {fndt}n1\{f_{n}\,dt\}_{n\geq 1} is uniformly absolutely continuous on II if to each ϵ>0\epsilon>0 there corresponds a δ>0\delta>0, independent of nn, such that for Borel sets AA we have

AI,|A|<δAfn𝑑t<ϵ.A\subset I,\,|A|<\delta\implies\int_{A}f_{n}\,dt<\epsilon.

Recall that the notion of a majorant has been introduced in \threfRegMajorantDef.

Lemma 3.3.
\thlabel

permanenceLemma Let II be a finite interval of the real line \mathbb{R}, θ=Sν\theta=S_{\nu} a singular inner function in \mathbb{H} defined by a singular measure ν\nu supported in the interior of II, and {hn}n1\{h_{n}\}_{n\geq 1} a sequence of functions in H()H^{\infty}(\mathbb{H}) such that

limnθ(z)hn(z)=h(z),z,\lim_{n\to\infty}\,\theta(z)h_{n}(z)=h(z),\quad z\in\mathbb{H},

where hH()h\in H^{\infty}(\mathbb{H}) is a non-zero function. Assume that

  1. (i)

    there exists a majorant FF for which we have

    supz=x+iyR|θ(z)hn(z)|exp(F(y))<C\sup_{z=x+iy\in R}\,|\theta(z)h_{n}(z)|\exp\big{(}-F(y)\big{)}<C

    for some constant C>0C>0 independent of nn, and where RR is some rectangle in \mathbb{H} with base II:

    R=R(I,d):={z=x+iy:xI,y<d},R=R(I,d):=\{z=x+iy\in\mathbb{H}:x\in I,y<d\},
  2. (ii)

    the sequence of positive Borel measures {log+|hn|dt}n1\{\log^{+}|h_{n}|dt\}_{n\geq 1} is uniformly absolutely continuous on an interval larger than II.

Then h/θH()h/\theta\in H^{\infty}(\mathbb{H}).

Proof.

The assumption (ii)(ii) implies that

supnIlog+|hn|dt<.\sup_{n}\int_{I}\log^{+}|h_{n}|\,dt<\infty.

So, denoting by 1I1_{I} the characteristic function of the interval II and by passing to a subsequence, we can assume that the measures 1Ilog+|hn|dt1_{I}\log^{+}|h_{n}|dt converge weak-star to a non-negative measure ν0\nu_{0} supported on II. The measure ν0\nu_{0} must be absolutely continuous with respect to dtdt: any set NIN\subset I of dtdt-measure zero can be covered by an open set UU of total length arbitrarily small, and then we can use (ii)(ii) to conclude that

ν0(N)ν0(U)lim infnUlog+|hn|dt<ϵ\nu_{0}(N)\leq\nu_{0}(U)\leq\liminf_{n\to\infty}\int_{U}\log^{+}|h_{n}|dt<\epsilon

for any ϵ>0\epsilon>0. Consequently dν0=wdtd\nu_{0}=w\,dt for some non-negative w1(I)w\in\mathcal{L}^{1}(I). We denote by uIu_{I} the harmonic function in \mathbb{H} which is the Poisson extension of the measure dν0=wdtd\nu_{0}=w\,dt to \mathbb{H}:

uI(z)=1πIy(xt)2+y2w(t)𝑑t,z=x+iy.u_{I}(z)=\frac{1}{\pi}\int_{I}\frac{y}{(x-t)^{2}+y^{2}}w(t)dt,\quad z=x+iy\in\mathbb{H}.

Let also unu_{n} denote the Poisson extension of the measure 1Ilog+|hn|dt1_{I}\log^{+}|h_{n}|dt:

un(z)=1πIy(xt)2+y2log+|hn(t)|dt,z=x+iy.u_{n}(z)=\frac{1}{\pi}\int_{I}\frac{y}{(x-t)^{2}+y^{2}}\log^{+}|h_{n}(t)|\,dt,\quad z=x+iy\in\mathbb{H}.

The assumption (i)(i) implies that

log|θ(z)|+log|hn(z)|c+F(y),z=x+iyR\log|\theta(z)|+\log|h_{n}(z)|\leq c+F(y),\quad z=x+iy\in R

for some positive constant c>0c>0. By \threfWizardHatMainProposition, there exists a wizard hat domain W=W(p,I)W=W(p,I) and a corresponding positive harmonic function uu defined on WW which satisfies

log|θ(z)|+log|hn(z)|u(z),zWR.\log|\theta(z)|+\log|h_{n}(z)|\leq u(z),\quad z\in\partial W\cap R.

By the assumption that the singular measure ν\nu defining θ\theta is supported in the interior of II, it follows that θ\theta is analytic and non-zero in a neighbourhood of WR\partial W\cap R, and so log|θ(z)|\log|\theta(z)| is bounded on WR\partial W\cap R. Therefore, by possibly replacing uu by a positive scalar multiple of itself, in fact we have that

(3.4) log|hn(z)|u(z),zWR.\log|h_{n}(z)|\leq u(z),\quad z\in\partial W\cap R.

For the bottom side II of the wizard hat, we have the non-tangential boundary value inequality

(3.5) log|hn(x)|un(x)\log|h_{n}(x)|\leq u_{n}(x)

for dtdt-almost every xIx\in I. This follows immediately from elementary boundary behaviour properties of Poisson integrals. We would like to conclude from the two inequalities (3.4) and (3.5) that

(3.6) log|hn(z)|u(z)+un(z),zW.\log|h_{n}(z)|\leq u(z)+u_{n}(z),\quad z\in W.

Indeed such a generalization of the maximum principle holds, and we will carefully verify this claim in \threfTdomLemma below. Assuming the claim, we recall that

hn(z)h(z)/θ(z),n+h_{n}(z)\to h(z)/\theta(z),\quad n\to+\infty

in all of \mathbb{H}, and so by letting n+n\to+\infty we obtain, from (3.6) and the earlier mentioned weak-star convergence of measures (which guarantees that un(z)uI(z)u_{n}(z)\to u_{I}(z) for zz\in\mathbb{H}), that

(3.7) log|h(z)|log|θ(z)|u(z)+uI(z),zW.\log|h(z)|-\log|\theta(z)|\leq u(z)+u_{I}(z),\quad z\in W.

Let JJ be some interval containing the support of ν\nu, and which is strictly contained in II. By \threfIntervalWeakStarConvLemma, as y0+y\to 0^{+}, the restrictions to JJ of the real-valued measures log|h(t+iy)|dt\log|h(t+iy)|dt converge weak-star to the restriction to JJ of the measure log|h(t)|dtdνh(t)\log|h(t)|dt-d\nu_{h}(t). Similar claims hold for the Poisson integrals

(3.8) 𝒫ν(z)=log|θ(z)|=1πy(xt)2+y2𝑑ν(t),z=x+iy-\mathcal{P}_{\nu}(z)=\log|\theta(z)|=-\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{(x-t)^{2}+y^{2}}d\nu(t),\quad z=x+iy\in\mathbb{H}

and 𝒫w=uI\mathcal{P}_{w}=u_{I}. For sufficiently small y>0y>0 we have that J+iy:={x+iy:xJ}WJ+iy:=\{x+iy:x\in J\}\subset W. Thus from the weak-star convergence of measures discussed above, the inequality (3.7), and the fact that u0u\equiv 0 on JJ, we obtain the real-valued measure inequality

log|h(t)|dtdνh(t)+dν(t)w(t)dt on J.\log|h(t)|dt-d\nu_{h}(t)+d\nu(t)\leq w(t)dt\quad\text{ on }J.

This measure inequality is to be interpreted in the following way:

w(t)dtlog|h(t)|dt+dνh(t)dν(t)w(t)dt-\log|h(t)|dt+d\nu_{h}(t)-d\nu(t)

is a non-negative measure on JJ. The dtdt-singular part of this measure is dνhdνd\nu_{h}-d\nu, which is thus non-negative on JJ. Since dνd\nu is supported inside JJ, in fact dνhdνd\nu_{h}-d\nu is non-negative in all of \mathbb{R}. Now subtracting (3.8) from (3.2), using the inequality αy+log|B(z)|0-\alpha y+\log|B(z)|\leq 0 and the non-negativity of dνhdνd\nu_{h}-d\nu and of the Poisson kernel, we get for z=x+iyz=x+iy\in\mathbb{H} that

log|h(z)|log|θ(z)|\displaystyle\log|h(z)|-\log|\theta(z)|\leq 1πy(xt)2+y2log|h(t)|dt\displaystyle\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{(x-t)^{2}+y^{2}}\log|h(t)|dt
\displaystyle\leq 1πy(xt)2+y2log(h)𝑑t\displaystyle\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{(x-t)^{2}+y^{2}}\log(\|h\|_{\infty})dt
=\displaystyle= log(h).\displaystyle\log(\|h\|_{\infty}).

By exponentiating, we finally obtain

|h(z)/θ(z)|hH(),z.|h(z)/\theta(z)|\leq\|h\|_{H^{\infty}(\mathbb{H})},\quad z\in\mathbb{H}.

We need to verify the claim made in the course of the proof of \threfpermanenceLemma which lead to the fundamental inequality (3.6).

Lemma 3.4.
\thlabel

TdomLemma Let the wizard hat W=W(p,I)W=W(p,I) be as in the proof of \threfpermanenceLemma, ff be a bounded analytic function in WW and uu be a positive harmonic function in WW. Assume that both ff and uu extend continuously to W\partial W\cap\mathbb{H} and also that both have non-tangential limits almost everywhere on II. If we have that log|f(z)|u(z)\log|f(z)|\leq u(z) for zWz\in\partial W\cap\mathbb{H}, and moreover that the non-tangential limits of log|f|\log|f| on II are dtdt-almost everywhere dominated by the non-tangential limits of uu on II, then log|f(z)|u(z)\log|f(z)|\leq u(z) for all zWz\in W.

Proof.

Let ϕ:𝔻W\phi:\mathbb{D}\to W be a conformal mapping. The local smoothness of the boundary of WW and basic conformal mapping theory ensure that ϕ\phi is conformal at almost every point of 𝕋\mathbb{T} (see [14, Chapter V.5]). This implies that the functions log|fϕ|\log|f\circ\phi| and uϕu\circ\phi, which are defined in 𝔻\mathbb{D}, have non-tangential limits almost everywhere on 𝕋\mathbb{T}. Let uϕ~\widetilde{u\circ\phi} be a harmonic conjugate of uϕu\circ\phi in 𝔻\mathbb{D} and consider the function H(z)=exp(uϕ(z)iuϕ~(z))H(z)=\exp\Big{(}-u\circ\phi(z)-i\widetilde{u\circ\phi}(z)\Big{)}, z𝔻z\in\mathbb{D}. By positivity of uu, the function HH is bounded in 𝔻\mathbb{D}, and our assumptions leads to the conclusion that the non-tangential boundary values on 𝕋\mathbb{T} of the bounded function (fϕ)(z)H(z)H(𝔻)(f\circ\phi)(z)H(z)\in H^{\infty}(\mathbb{D}) are not larger than 1 in modulus. Thus by basic function theory in 𝔻\mathbb{D}, we obtain the inequality |(fϕ)(z)H(z)|1|(f\circ\phi)(z)H(z)|\leq 1 for all z𝔻z\in\mathbb{D}. This easily translates into log|f(z)|u(z)\log|f(z)|\leq u(z) for zWz\in W. ∎

We will need \threfpermanenceLemma in the disk 𝔻\mathbb{D}. Here is the precise statement which we will use. The uniform absolute continuity of sequences of Borel measures on arcs II of the circle 𝕋\mathbb{T} is defined analogously to how it was defined in \threfuniAbsContDef for intervals on the line.

Corollary 3.5.
\thlabel

PermanenceLemmaDisk Let II be an arc properly contained in the circle 𝕋\mathbb{T}, θ=Sν\theta=S_{\nu} be a singular inner function in 𝔻\mathbb{D} defined by a singular measure ν\nu supported in the interior of II, and {hn}n1\{h_{n}\}_{n\geq 1} be a sequence of functions in HH^{\infty} such that

limn+θ(z)hn(z)=h(z),z𝔻\lim_{n\to+\infty}\,\theta(z)h_{n}(z)=h(z),\quad z\in\mathbb{D}

where hHh\in H^{\infty}. Assume that

  1. (i)

    there exists a majorant FF for which we have

    supz𝔻|θ(z)hn(z)|exp(F(1|z|))<C\sup_{z\in\mathbb{D}}\,|\theta(z)h_{n}(z)|\exp\Big{(}-F\big{(}1-|z|\big{)}\Big{)}<C

    for some positive constant C>0C>0 independent of nn,

  2. (ii)

    the sequence of positive Borel measures {log+|hn|dm}n1\{\log^{+}|h_{n}|d\textit{m}\}_{n\geq 1} is uniformly absolutely continuous on an arc larger than II.

Then h/θHh/\theta\in H^{\infty}.

It is easy to see that \threfpermanenceLemma implies \threfPermanenceLemmaDisk. Indeed, if ϕ:𝔻\phi:\mathbb{H}\to\mathbb{D} is a conformal map for which ϕ1(I)\phi^{-1}(I) is a finite segment on \mathbb{R}, then the distortion of lengths and distances by the map ϕ\phi is bounded above and below near ϕ1(I)\phi^{-1}(I) and II, since ϕ\phi is a bi-Lipschitz bijection between some open sets containing ϕ1(I)\phi^{-1}(I) and II. For instance, the growth condition (i)(i) in \threfPermanenceLemmaDisk for the sequence {Sνhn}n1\{S_{\nu}h_{n}\}_{n\geq 1} is easily translated into a corresponding condition (i)(i) in \threfpermanenceLemma for the sequence {Sν~h~n}n1\{S_{\tilde{\nu}}\tilde{h}_{n}\}_{n\geq 1}, where Sν~:=SνϕS_{\tilde{\nu}}:=S_{\nu}\circ\phi and h~n:=hnϕ\tilde{h}_{n}:=h_{n}\circ\phi, by replacing F(t)F(t) with a new majorant of the form F~(t):=F(at)\tilde{F}(t):=F(at) for some a>0a>0. Moreover, the mapping ϕ\phi will preserve the uniform absolute continuity properties of the corresponding measures. Thus \threfPermanenceLemmaDisk can readily be deduced from \threfpermanenceLemma and a change of variables argument.

3.2. Proof of \threfPermanenceMainTheorem

\thref

PermanenceMainTheorem follows almost immediately from \threfPermanenceLemmaDisk, we just need to verify that a bounded sequence in the corresponding 𝒫2(μ)\mathcal{P}^{2}(\mu)-space satisfies properties (i)(i) and (ii)(ii) in \threfPermanenceLemmaDisk. This is done in the next two lemmas.

Lemma 3.6.
\thlabel

PointEvaluationBoundRegMajorant Let

dμ𝔻(z)=G(1|z|)dA(z),d\mu_{\mathbb{D}}(z)=G(1-|z|)dA(z),

where GG satisfies the condition (LogLogInt) appearing in Section 1. For z𝔻z\in\mathbb{D}, denote by

Ez:=supf𝒫,fμ𝔻=1|f(z)|E_{z}:=\sup_{\begin{subarray}{c}f\in\mathcal{P},\\ \|f\|_{\mu_{\mathbb{D}}}=1\end{subarray}}|f(z)|

the norm of the evaluation functional zf(z)z\mapsto f(z) on 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}). There exists a majorant FF such that

Ezexp(F(1|z|)),z𝔻.E_{z}\leq\exp\big{(}F(1-|z|)\big{)},\quad z\in\mathbb{D}.
Proof.

Fix z𝔻z\in\mathbb{D}, δ=(1|z|)/2\delta=(1-|z|)/2 and let B(z,δ)B(z,\delta) denote the ball around zz of radius δ\delta. By subharmonicity of the function z|f(z)|z\mapsto|f(z)| and the Cauchy-Schwarz inequality, we have

|f(z)|\displaystyle|f(z)| 1δ2B(z,δ)|f(z)|𝑑A(z)\displaystyle\leq\frac{1}{\delta^{2}}\int_{B(z,\delta)}|f(z)|dA(z)
1δ2fμ𝔻B(z,δ)1G(1|z|)𝑑A(z).\displaystyle\leq\frac{1}{\delta^{2}}\|f\|_{\mu_{\mathbb{D}}}\sqrt{\int_{B(z,\delta)}\frac{1}{G(1-|z|)}dA(z)}.

Since GG is assumed to be an increasing function, we may estimate the integral inside the square root by

B(z,δ)1G(1|z|)𝑑A(z)1δ21G((1|z|)/2).\int_{B(z,\delta)}\frac{1}{G(1-|z|)}dA(z)\leq\frac{1}{\delta^{2}}\frac{1}{G\big{(}(1-|z|)/2\big{)}}.

Putting this into the previous estimate, we obtain

|f(z)|fμ𝔻δ31G((1|z|)/2)=8fμ𝔻(1|z|)31G((1|z|)/2)|f(z)|\leq\frac{\|f\|_{\mu_{\mathbb{D}}}}{\delta^{3}}\sqrt{\frac{1}{G\big{(}(1-|z|)/2\big{)}}}=\frac{8\|f\|_{\mu_{\mathbb{D}}}}{(1-|z|)^{3}}\sqrt{\frac{1}{G\big{(}(1-|z|)/2\big{)}}}

Now set

F(t):=log(8t3)+12log(1G(t/2)),t(0,1].F(t):=\log\Big{(}\frac{8}{t^{3}}\Big{)}+\frac{1}{2}\log\Big{(}\frac{1}{G(t/2)}\Big{)},\quad t\in(0,1].

By the above estimate, the norm EzE_{z} of the evaluation functional is bounded by

Ezexp(F(1|z|)),z𝔻.E_{z}\leq\exp\big{(}F(1-|z|)\big{)},\quad z\in\mathbb{D}.

Moreover, FF is a decreasing function, and by virtue of GG satisfying (LogLogInt), FF also certainly satisfies

0dlogF(t)𝑑t<\int_{0}^{d}\log F(t)\,dt<\infty

if d>0d>0 is some small number. Thus FF is a majorant in the sense of \threfRegMajorantDef. ∎

Lemma 3.7.
\thlabel

UniAbsContLemma Assume that the weight ww satisfies

Ilogwdm>\int_{I}\log w\,d\textit{m}>-\infty

for some arc I𝕋I\subset\mathbb{T}. If {fn}n1\{f_{n}\}_{n\geq 1} are positive functions such that

Ifnpw𝑑m<C,n1\int_{I}f_{n}^{p}w\,d\textit{m}<C,\quad n\geq 1

for some constant C>0C>0 and some p>0p>0, then the sequence {log+fndm}n1\{\log^{+}f_{n}\,d\textit{m}\}_{n\geq 1} is uniformly absolutely continuous on II.

Proof.

Note that

log+fn\displaystyle\log^{+}f_{n} log+(fnw1/p)+log+(w1/p)\displaystyle\leq\log^{+}(f_{n}w^{1/p})+\log^{+}(w^{-1/p})
1plog+(fnpw)+1plog+(1/w)\displaystyle\leq\frac{1}{p}\log^{+}(f_{n}^{p}w)+\frac{1}{p}\log^{+}(1/w)
:=gn+g,\displaystyle:=g_{n}+g,

where it follows from the assumption that gng_{n} are positive functions which form a bounded subset of (say) 2(I)\mathcal{L}^{2}(I), and g1(I)g\in\mathcal{L}^{1}(I). Clearly, if AA is a Borel subset of II, then by Cauchy-Schwarz inequality we obtain

Agndm|A|gn2(I),\int_{A}g_{n}\,d\textit{m}\leq\sqrt{|A}|\cdot\|g_{n}\|_{\mathcal{L}^{2}(I)},

so that the family {gndm}n1\{g_{n}d\textit{m}\}_{n\geq 1} is uniformly absolutely continuous on II. Then the above inequalities imply that {log+fndm}n1\{\log^{+}f_{n}\,d\textit{m}\}_{n\geq 1} is a uniformly absolutely continuous sequence on II. ∎

Proof of \threfPermanenceMainTheorem.

Let h[Sν]𝒩+h\in[S_{\nu}]\cap\mathcal{N}^{+}. Since h[Sν]h\in[S_{\nu}], there exists a sequence of polynomials {pn}n1\{p_{n}\}_{n\geq 1} such that SνpnS_{\nu}p_{n} converges to hh in the norm of 𝒫2(μ)\mathcal{P}^{2}(\mu). Multiplying hh by a suitable bounded outer function uu we can ensure that huHhu\in H^{\infty}, and that SνpnuS_{\nu}p_{n}u converges to huhu (see the discussion following \threfProdOfCyclicLemma below). Let {Kj}j\{K_{j}\}_{j} be an increasing sequence of compact sets which are finite unions of intervals and such that jKj=core(w)\cup_{j}K_{j}=\text{core}(w). By \threfPermanenceLemmaDisk, \threfPointEvaluationBoundRegMajorant and \threfUniAbsContLemma, whenever νj\nu_{j} is the restriction of ν\nu to the compact subset KjK_{j}, we have that hu/SνjHhu/S_{\nu_{j}}\in H^{\infty} with the bound hu/Sνjhu\|hu/S_{\nu_{j}}\|_{\infty}\leq\|hu\|_{\infty}. The assumption that ν(𝕋)=ν(core(w))\nu(\mathbb{T})=\nu\big{(}\text{core}(w)\big{)} means that the restrictions νj\nu_{j} converge weak-star to the measure ν\nu. Thus

|h(z)u(z)/Sν(z)|=limj|h(z)u(z)/Sνj(z)|hu,z𝔻|h(z)u(z)/S_{\nu}(z)|=\lim_{j\to\infty}|h(z)u(z)/S_{\nu_{j}}(z)|\leq\|hu\|_{\infty},\quad z\in\mathbb{D}

In particular, since uu is outer, it follows that SνS_{\nu} divides the inner factor of hh. Thus h/Sν𝒩+h/S_{\nu}\in\mathcal{N}^{+}. ∎

4. Cyclic singular inner functions

In this section we will study the cyclicity of singular inner functions, and prove \threfCyclicityMainTheorem.

4.1. Weak-star approximation of singular measures, with obstacles

The cyclicity in 𝒫2(μ)\mathcal{P}^{2}(\mu) of the singular inner function SνS_{\nu} will follow from the existence of a sequence of non-negative bounded functions {fn}n1\{f_{n}\}_{n\geq 1} for which the measures {fndm}n1\{f_{n}\,d\textit{m}\}_{n\geq 1} converge weak-star to ν\nu. In our context the functions fnf_{n} will have to satisfy a severe restriction on their size, namely

(4.1) 0fn(x)log+(1/w(x)),x𝕋.0\leq f_{n}(x)\leq\log^{+}(1/w(x)),\quad x\in\mathbb{T}.

In case w(x)=0w(x)=0, the right-hand side is to be interpreted as ++\infty (i.e., no size restriction on fnf_{n} at the point xx). Essentially, the obstacle (4.1) prohibits the existence of an approximating sequence {fndm}n1\{f_{n}d\textit{m}\}_{n\geq 1} if some part of the mass of ν\nu is located in ”wrong” places on 𝕋\mathbb{T}. However, if ν\nu is carried outside of the core set of ww, then such a sequence exists. This is the content of the next lemma.

Lemma 4.1.
\thlabel

WeakStarSeqObstacleLemma Let ν\nu be a positive singular Borel measure on 𝕋\mathbb{T} which satisfies

ν(core(w))=0.\nu\big{(}\text{core}(w)\big{)}=0.

Then there exists a sequence of non-negative bounded functions {fn}n1\{f_{n}\}_{n\geq 1} satisfying the following three properties:

  1. (i)

    the non-negative measures {fndm}n1\{f_{n}\,d\textit{m}\}_{n\geq 1} converge weak-star to ν\nu,

  2. (ii)

    𝕋fn𝑑m=ν(𝕋)\int_{\mathbb{T}}f_{n}\,d\textit{m}=\nu(\mathbb{T}),

  3. (iii)

    the functions fnf_{n} obey the bound (4.1).

Proof.

Let us first suppose that ν\nu assigns no mass to any singletons, so that ν({x})=0\nu(\{x\})=0 whenever x𝕋x\in\mathbb{T}. For any positive integer nn, we let DnD_{n} be the family of 2n2^{n} disjoint open dyadic intervals, each of length 2π2n2\pi\cdot 2^{-n}, such that their union covers the circle 𝕋\mathbb{T} up to finitely many points, and such that the system n1Dn\cup_{n\geq 1}D_{n} possesses the usual dyadic nesting property: each dDnd\in D_{n} is contained in a unique dDn1d^{\prime}\in D_{n-1}. Fixing an integer n1n\geq 1, we will specify how to define fnf_{n} on each of the intervals djDnd_{j}\in D_{n}, 1j2n1\leq j\leq 2^{n}, in such a way that the above three properties hold.

If ν(dj)=0\nu(d_{j})=0, then we simply set fn0f_{n}\equiv 0 on djd_{j}. Conversely, if ν(dj)>0\nu(d_{j})>0, then since ν(core(w))=0\nu\big{(}\text{core}(w)\big{)}=0, it must be that ν(dj)=ν(djcore(w))>0\nu(d_{j})=\nu(d_{j}\setminus\text{core}(w))>0. It follows that djcore(w)d_{j}\setminus\text{core}(w) is non-empty. Pick some point xdjcore(w)x\in d_{j}\setminus\text{core}(w). For any open interval II which contains xx in its interior we have Ilog+(1/w)𝑑m=+\int_{I}\log^{+}(1/w)\,d\textit{m}=+\infty (else xx would have been a member of core(w)\text{core}(w)). Pick such an interval II which is contained within djd_{j}. If there exists a subset AIA\subset I satisfying m(A)=|A|>0m(A)=|A|>0 on which ww is identically zero, then we may set

fn(x)=ν(dj)|A|11A(x),xdjf_{n}(x)=\nu(d_{j})|A|^{-1}1_{A}(x),\quad x\in d_{j}

where 1A1_{A} is the characteristic function of AA. In case that such a set does not exist, then w>0w>0 almost everywhere on II, and we must have

+=Ilog+(1/w)𝑑m=limc0+I{w>c}log+(1/w)𝑑m+\infty=\int_{I}\log^{+}(1/w)\,d\textit{m}=\lim_{c\to 0^{+}}\int_{I\cap\{w>c\}}\log^{+}(1/w)\,d\textit{m}

so that

ν(dj)<I{w>c}log+(1/w)𝑑m<+\nu(d_{j})<\int_{I\cap\{w>c\}}\log^{+}(1/w)\,d\textit{m}<+\infty

for some small c>0c>0. By absolute continuity of the finite measure

log+(1/w)1I{w>c}dm\log^{+}(1/w)1_{I\cap\{w>c\}}d\textit{m}

there must then exist a set BI{w>c}B\subset I\cap\{w>c\} for which we have precisely

ν(dj)=Blog+(1/w)𝑑m\nu(d_{j})=\int_{B}\log^{+}(1/w)\,d\textit{m}

We pick such BB and define

fn(x)=log+(1/w(x))1B(x),xdj.f_{n}(x)=\log^{+}(1/w(x))1_{B}(x),\quad x\in d_{j}.

Note that fn(x)log+(1/c)f_{n}(x)\leq\log^{+}(1/c) on djd_{j}. For definiteness, we can set fnf_{n} to be equal to 0 on the finitely many points outside of j=12ndj\cup_{j=1}^{2^{n}}d_{j}. One way or the other, we have defined fnf_{n} as a bounded function, and we have

djfn𝑑m=ν(dj).\int_{d_{j}}f_{n}\,d\textit{m}=\nu(d_{j}).

By summing over all the 2n2^{n} intervals djd_{j}, we see that property (ii)(ii) in the statement of the lemma is satisfied (since ν\nu assigns no mass to the finitely many points outside the union of the open intervals djd_{j}). Property (iii)(iii) is satisfied by the construction. Property (i)(i) also holds. Indeed, if gg is the characteristic function of one of the dyadic intervals djd_{j} from some stage of our construction, then the nesting property of the dyadic system and the additivity of ν\nu ensure that

limn𝕋gfn𝑑m=ν(dj)=𝕋g𝑑ν.\lim_{n\to\infty}\int_{\mathbb{T}}gf_{n}\,d\textit{m}=\nu(d_{j})=\int_{\mathbb{T}}g\,d\nu.

The above equalities hold also for functions gg which are finite linear combinations of characteristic functions of dyadic intervals. Since such linear combinations can be used to uniformly approximate any continuous function on 𝕋\mathbb{T}, and since we have the uniform variation bound in (ii)(ii), we conclude that the sequence {fndm}nf_{n}\,d\textit{m}\}_{n} converges weak-star to ν\nu. The proof is complete in the case that ν\nu assigns no mass to singletons.

In the contrary case we have that

ν=jcjδxj,cj>0\nu=\sum_{j}c_{j}\delta_{x_{j}},\quad c_{j}>0

is a countable linear combination of unit masses δxj\delta_{x_{j}} at the sequence of points {xj}j\{x_{j}\}_{j} in 𝕋\mathbb{T}. Our assumption implies that xjcore(w)x_{j}\not\in\text{core}(w) for all jj. Thus each xjx_{j} is the midpoint of an interval II which can be chosen to have arbitrarily small length and for which we have Ilog+(1/w)𝑑m=+\int_{I}\log^{+}(1/w)\,d\textit{m}=+\infty. We can then proceed in an analogous way to the above, and produce at each stage nn of the construction a disjoint finite sequence of intervals {In,j}j=1n\{I_{n,j}\}_{j=1}^{n} each covering a different point xjx_{j} for j=1,,nj=1,\ldots,n. We then define a positive function fnf_{n} which carries appropriate amount of mass on each of the intervals In,jI_{n,j} and satisfies the other needed properties. We skip laying out the straight-forward details of this adaptation of the previous argument.

The general case follows by decomposing a measure ν\nu into one measure which is a sum of point masses and one measure which assigns no mass to singletons. ∎

4.2. Proof of \threfCyclicityMainTheorem

We will need one more elementary lemma. It appears in [10] and many other works.

Lemma 4.2.
\thlabel

ProdOfCyclicLemma Assume that HH is a Banach space of analytic functions in 𝔻\mathbb{D} which contains HH^{\infty} and with the property that for all functions hh in HH^{\infty} the operator Mhf:=hfM_{h}f:=hf is bounded on HH. Then the product uvuv of two cyclic bounded functions u,vHu,v\in H^{\infty} is cyclic.

By cyclicity of uu we mean, of course, that there exists a sequence of analytic polynomials {pn}n1\{p_{n}\}_{n\geq 1} such that pnup_{n}u converges to 1H1\in H in the norm of the space.

Proof.

If uu and vv are two cyclic bounded functions, then for any polynomials p,qp,q we have the inequality

1puvH1qvH+MvqpuH,\|1-puv\|_{H}\leq\|1-qv\|_{H}+\|M_{v}\|\|q-pu\|_{H},

where Mv\|M_{v}\| denotes the operator norm of the multiplication operator MvM_{v}, and H\|\cdot\|_{H} denotes the norm in HH. We use cyclicity of vv to choose the polynomial qq to make the first term on the right arbitrarily small, and next we use cyclicity of uu to choose pp to make the second term on the right arbitrarily small. It follows that the product uvuv of two bounded cyclic functions is a cyclic function. ∎

\thref

ProdOfCyclicLemma applies to any irreducible space 𝒫2(μ)\mathcal{P}^{2}(\mu) of the form considered here, since indeed the multiplication by any function in HH^{\infty} induces a bounded operator on these spaces. We skip the straight-forward proof, which can for instance be based on simple analysis of the dilations hr(z):=h(rz)h_{r}(z):=h(rz), r(0,1)r\in(0,1), of the bounded function hh. In particular, H𝒫2(μ)H^{\infty}\subset\mathcal{P}^{2}(\mu) whenever the latter is irreducible. For future reference, note that as a subspace of 2(μ)\mathcal{L}^{2}(\mu) (with μ\mu as in (1.1)), each function hH𝒫2(μ)h\in H^{\infty}\subset\mathcal{P}^{2}(\mu) is defined also on μ𝕋:=wdm\mu_{\mathbb{T}}:=w\,d\textit{m}, the part of μ\mu living on the circle 𝕋\mathbb{T}. It is not hard to see that the values of hh with respect to μ𝕋\mu_{\mathbb{T}} coincide with the non-tangential boundary function of hh on 𝕋\mathbb{T}. If ww is bounded, then the same conclusions hold also for any hH2𝒫2(μ)h\in H^{2}\subset\mathcal{P}^{2}(\mu).

Proof of \threfCyclicityMainTheorem.

Note first that (i)(ii)(i)\Rightarrow(ii) in \threfCyclicityMainTheorem, since the condition ν(core(w))>0\nu\big{(}\text{core}(w)\big{)}>0 implies that a factor in SνS_{\nu} satisfies the permanence property exhibited in \threfPermanenceMainTheorem, and so SνS_{\nu} cannot by cyclic. Thus it suffices for us to show the implication (ii)(i)(ii)\Rightarrow(i). The norms induced by measures μ\mu satisfying (ExpDec) are largest if the measure μ\mu has the form in (T1T1) defined in Section 1, with β=1\beta=1. If SνS_{\nu} is cyclic in 𝒫2(μ)\mathcal{P}^{2}(\mu) defined by any measure this form, then it is cyclic in any 𝒫2(μ)\mathcal{P}^{2}(\mu)-space considered in this article. Thus it suffices to prove the theorem in the case of μ\mu being of the form (T1T1) with β=1\beta=1, and any c>0c>0.

Let us then assume that ν(core(w))=0\nu\big{(}\text{core}(w)\big{)}=0. The formula (1.5) shows that

Sν(z)=i=1NSν/N(z),z𝔻S_{\nu}(z)=\prod_{i=1}^{N}S_{\nu/N}(z),\quad z\in\mathbb{D}

for any positive integer N1N\geq 1. Then by replacing ν\nu by ν/N\nu/N for NN sufficiently large, and by \threfProdOfCyclicLemma, we may assume that ν(𝕋)<c/10\nu(\mathbb{T})<c/10. Let {fn}n1\{f_{n}\}_{n\geq 1} be a sequence of positive bounded functions given by \threfWeakStarSeqObstacleLemma for which the measures {fndm}n1\{f_{n}\,d\textit{m}\}_{n\geq 1} converge weak-star to 2ν2\nu, which satisfy 𝕋fn𝑑m=2ν(𝕋)\int_{\mathbb{T}}f_{n}\,d\textit{m}=2\nu(\mathbb{T}), and for which the bound (4.1) holds. Construct the outer functions

hn(z):=exp(Hfn(z)/2),z𝔻,h_{n}(z):=\exp\Big{(}H_{f_{n}}(z)/2\Big{)},\quad z\in\mathbb{D},

where

Hfn(z):=𝕋x+zxzfn(x)𝑑m(x),z𝔻H_{f_{n}}(z):=\int_{\mathbb{T}}\frac{x+z}{x-z}f_{n}(x)\,d\textit{m}(x),\quad z\in\mathbb{D}

is the usual Herglotz integral of fnf_{n}. Then, since |Hfn(z)|4ν(𝕋)1|z||H_{f_{n}}(z)|\leq\frac{4\nu(\mathbb{T})}{1-|z|}, we obtain

|hn(z)|exp(2ν(𝕋)1|z|)exp(c5(1|z|)),z𝔻,|h_{n}(z)|\leq\exp\Big{(}\frac{2\nu(\mathbb{T})}{1-|z|}\Big{)}\leq\exp\Big{(}\frac{c}{5(1-|z|)}\Big{)},\quad z\in\mathbb{D},

and from property (iii)(iii) in \threfWeakStarSeqObstacleLemma and basic properties of Herglotz integrals, we have the non-tangential boundary value estimate

|hn(x)|=exp(fn(x)/2)max[1,1/w(x)]|h_{n}(x)|=\exp\Big{(}f_{n}(x)/2\Big{)}\leq\sqrt{\max[1,1/w(x)]}

for almost every x𝕋x\in\mathbb{T} with respect to mm. It follows from these inequalities and the definition of the norm in 𝒫2(μ)\mathcal{P}^{2}(\mu) that the family {hn}n1H\{h_{n}\}_{n\geq 1}\subset H^{\infty} forms a bounded subset of the Hilbert space 𝒫2(μ)\mathcal{P}^{2}(\mu). Moreover, by the weak-star convergence of {fndm}n1\{f_{n}\,d\textit{m}\}_{n\geq 1} to 2ν2\nu we have that

limnhn(z)=1Sν(z),z𝔻.\lim_{n\to\infty}h_{n}(z)=\frac{1}{S_{\nu}(z)},\quad z\in\mathbb{D}.

But this means that 1/Sν1/S_{\nu} is a member of 𝒫2(μ)\mathcal{P}^{2}(\mu), since we can identify it as a weak cluster point of some subsequence of {hn}n1\{h_{n}\}_{n\geq 1}. Thus there must also exist a sequence of polynomials {pn}n\{p_{n}\}_{n} tending to 1/Sν1/S_{\nu} in the norm of 𝒫2(μ)\mathcal{P}^{2}(\mu). Consequently, since the multiplication operator MSνM_{S_{\nu}} is a bounded on our space, we have that Sνpn1S_{\nu}p_{n}\to 1 in the norm of 𝒫2(μ)\mathcal{P}^{2}(\mu). That is, SνS_{\nu} is cyclic. ∎

5. Moment functions, admissible sequences and spaces of Taylor series

This section initiates the second part of the article. In this part, we will apply our previous results in 𝒫2(μ)\mathcal{P}^{2}(\mu)-theory to Cauchy integrals, model spaces and the de Branges-Rovnyak spaces (b)\mathcal{H}(b). In order to do so, we will need to analyze the moments of the functions GG appearing in (1.1). This entire section is concerned with this analysis.

5.1. Admissible sequences and their properties

If GG is a function satisfying (ExpDec) and (LogLogInt), then the sequence of moments of GG, defined below in (5.3), will be shown to satisfy the following basic properties.

Definition 5.1.

(Admissible sequences) \thlabelAdmissibleSequenceDef A decreasing sequence of positive numbers {Mn}n0\{M_{n}\}_{n\geq 0} with

limnMn=0\lim_{n\to\infty}M_{n}=0

will be called admissible if it satisfies the following three conditions:

  1. (i)

    the sequence {logMn}n0\{\log M_{n}\}_{n\geq 0} is eventually convex, in the sense that

    2logMnlogMn+1+logMn12\log M_{n}\leq\log M_{n+1}+\log M_{n-1}

    for all sufficiently large n0n\geq 0,

  2. (ii)

    there exists d>0d>0 such that

    Mnexp(dn)M_{n}\leq\exp(-d\sqrt{n})

    for all sufficiently large n0n\geq 0,

  3. (iii)

    the summability condition

    n0log(1/Mn)1+n2<\sum_{n\geq 0}\frac{\log(1/M_{n})}{1+n^{2}}<\infty

    is satisfied.

With later applications in mind, it will be useful to single out the following simple preservation property of admissible sequences under taking powers.

Proposition 5.2.
\thlabel

PowerAdmSeq If M={Mn}n0M=\{M_{n}\}_{n\geq 0} is an admissible sequence, then so is

Mp:={Mnp}n0,M^{p}:=\{M_{n}^{p}\}_{n\geq 0},

for any p>0p>0.

The proposition follows immediately from \threfAdmissibleSequenceDef

5.2. Legendre envelopes

Roughly speaking, admissible sequences {Mn}n0\{M_{n}\}_{n\geq 0} are in a correspondence with moments of functions GG satisfying (ExpDec) and (LogLogInt), and we shall now proceed to make this statement more precise. In order to do so, we will need to recall some basic concepts from convex analysis. In particular we will use the notion of Legendre envelopes and their properties. In parts of our exposition we will follow Beurling in [4] and Havin-Jöricke in [15], and we will refer to those works for most of the proofs of the following claims.

Let m(x)m(x) be a positive and continuous function defined for x>0x>0, which is decreasing and satisfies

limx0+m(x)=+.\lim_{x\to 0^{+}}m(x)=+\infty.

In our application, mm will be of the form m(x)=log1/G(x)m(x)=\log 1/G(x) (for small xx). The lower Legendre envelope mm_{*} is defined as

(5.1) m(x):=infy>0m(y)+xy,x>0.m_{*}(x):=\inf_{y>0}\,m(y)+xy,\quad x>0.

Being an infimum of concave (actually affine) and increasing functions, mm_{*} is itself concave and increasing, and it is easy to see that

limx+m(x)=+.\lim_{x\to+\infty}m_{*}(x)=+\infty.
Remark 5.3.
\thlabel

remarkLowerLegEnvelope Assume that we modify the function mm above for xx larger than 11, so that we end up with a different function m~\widetilde{m} which satisfies m~(x)=m(x)\widetilde{m}(x)=m(x) for x<1x<1, but the values of the two functions might differ for x1x\geq 1. Then it is not hard to see from the definition in (5.1) that m~(x)=m(x)\widetilde{m}_{*}(x)=m_{*}(x) for all sufficiently large xx. Indeed, if y1y\geq 1, then we have by positivity of mm that

m(y)+xyxm(1/2)+x/2,m(y)+xy\geq x\geq m(1/2)+x/2,

the second inequality holding if xx is sufficiently large. For such xx, the candidate y=1/2y=1/2 is always better than any candidate y1y\geq 1 in the infimum in (5.1), and our claim follows.

In [4, Lemma 1], Beurling proves the following statement which will be used below.

Proposition 5.4.
\thlabel

LogIntLogSumEquivalenceProp Let m(x)m(x) be a positive, continuous and decreasing function of x>0x>0 which satisfies limx0+m(x)=+.\lim_{x\to 0^{+}}m(x)=+\infty. The following two statements are equivalent:

  1. (i)

    there exists a δ>0\delta>0 such that

    0δlogm(x)𝑑x<,\int_{0}^{\delta}\log m(x)\,dx<\infty,
  2. (ii)

    we have

    1m(x)1+x2𝑑x<.\int_{1}^{\infty}\frac{m_{*}(x)}{1+x^{2}}dx<\infty.

We refer the reader to [4] for a proof of \threfLogIntLogSumEquivalenceProp.

Let k(x)k(x) be a positive concave function of x>0x>0 which is increasing and satisfies

limx+k(x)=+.\lim_{x\to+\infty}k(x)=+\infty.

We will consider its upper Legendre envelope defined as

(5.2) k(x):=supy>0k(y)xy.k^{*}(x):=\sup_{y>0}\,k(y)-xy.

Then it is easy to see that kk^{*} is a convex and decreasing function, and

limx0+k(x)=+.\lim_{x\to 0^{+}}k^{*}(x)=+\infty.

We have the following inversion formula, which is well-known (see [15, p. 224-225]).

Lemma 5.5.
\thlabel

LegendreInversionFormula Let kk be a positive concave function of x>0x>0 which is increasing and satisfies limxk(x)=+\lim_{x\to\infty}k(x)=+\infty. Then

(k)(x)=k(x).(k^{*})_{*}(x)=k(x).

5.3. A characterization of admissible sequences

We will be interested in the sequence {Mn}n0\{M_{n}\}_{n\geq 0} of moments of the parts of our measures μ\mu living on 𝔻\mathbb{D}:

Mn=Mn(G)\displaystyle M_{n}=M_{n}(G) :=𝔻G(1|z|)|z|2n𝑑A(z)\displaystyle:=\int_{\mathbb{D}}G(1-|z|)|z|^{2n}dA(z)
(5.3) =201G(1r)r2n+1𝑑r.\displaystyle=2\int_{0}^{1}G(1-r)r^{2n+1}dr.

We define the moment function of GG by

(5.4) PG(x):=01G(1r)rx𝑑x,x>0.P_{G}(x):=\int_{0}^{1}G(1-r)r^{x}\,dx,\quad x>0.

The next lemma gives an estimate on PGP_{G}. We skip the proof, which is essentially the same as the one given in [15, p. 229] (see also the proof of [24, Lemma 4.3]).

Lemma 5.6.
\thlabel

MomentGGrowthLogLogInt Let G(x)G(x), x[0,1]x\in[0,1], be an increasing continuous function satisfying G(0)=0G(0)=0, and put

(5.5) m(x):={log1/G(x),x(0,1]log1/G(1),x>1.m(x):=\begin{cases}\log 1/G(x),&x\in(0,1]\\ \log 1/G(1),&x>1.\end{cases}

Then, for sufficiently large x>0x>0, we have the estimates

exp(m(2x))4xPG(x)exp(m(x)),\frac{\exp\big{(}-m_{*}(2x)\big{)}}{4x}\leq P_{G}(x)\leq\exp\big{(}-m_{*}(x)\big{)},

where PGP_{G} is the moment function of GG defined in (5.4).

Lemma 5.7.
\thlabel

MomentEstProp For c>0c>0 and β>0\beta>0, let {Mn(β,c)}n0\{M_{n}(\beta,c)\}_{n\geq 0} be the sequence of moments given by

Mn(β,c)\displaystyle M_{n}(\beta,c) :=𝔻exp(c(1|z|)β)|z|2n𝑑A(z)\displaystyle:=\int_{\mathbb{D}}\exp\Big{(}-c(1-|z|)^{-\beta}\Big{)}|z|^{2n}\,dA(z)
(5.6) =201exp(c(1r)β)r2n+1𝑑r,n0.\displaystyle=2\int_{0}^{1}\exp\Big{(}-c(1-r)^{-\beta}\Big{)}r^{2n+1}\,dr,\quad n\geq 0.

Put

β~:=ββ+1.\tilde{\beta}:=\frac{\beta}{\beta+1}.

For sufficiently large positive nn, we have the estimates

(5.7) exp(2dnβ~)Mn(β,c)exp(dnβ~)\exp\Big{(}-2dn^{\tilde{\beta}}\Big{)}\leq M_{n}(\beta,c)\leq\exp\Big{(}-dn^{\tilde{\beta}}\Big{)}

where d=d(β,c)d=d(\beta,c) is comparable to c1/(β+1)c^{1/(\beta+1)} if β\beta remains fixed.

Proof.

In the notation of \threfMomentGGrowthLogLogInt, and with

G(x)=exp(cxβ),x(0,1)G(x)=\exp\Big{(}-\frac{c}{x^{\beta}}\Big{)},\quad x\in(0,1)

we have

m(x)=cxβ,x(0,1)m(x)=\frac{c}{x^{\beta}},\quad x\in(0,1)

and we need to compute the corresponding Legendre envelope mm_{*} defined in (5.1). Having fixed some number x>0x>0, we use elementary calculus to show that

infy>0cyβ+xy\inf_{y>0}\,\frac{c}{y^{\beta}}+xy

is attained at the point

y:=(cβx)1β+1y_{*}:=\Bigg{(}\frac{c\beta}{x}\Bigg{)}^{\frac{1}{\beta+1}}

from which it follows that

m(x)\displaystyle m_{*}(x) =cyβ+xy\displaystyle=cy_{*}^{-\beta}+xy_{*}
=c1/(β+1)(ββ/(β+1)+β1/(β+1))xβ^\displaystyle=c^{1/(\beta+1)}(\beta^{-\beta/(\beta+1)}+\beta^{1/(\beta+1)})x^{\widehat{\beta}}
:=d(β,c)xβ^.\displaystyle:=d(\beta,c)x^{\widehat{\beta}}.

Since

Mn(β,c)=2PG(2n+1)M_{n}(\beta,c)=2P_{G}(2n+1)

we obtain from \threfMomentGGrowthLogLogInt the inequalities

(5.8) 2exp(d(β,c)(4n+2)β~log(8n+4))Mn(β,c)2exp(d(β,c)(2n+1)β~)2\exp\big{(}-d(\beta,c)(4n+2)^{\tilde{\beta}}-\log(8n+4)\big{)}\leq M_{n}(\beta,c)\leq 2\exp\big{(}-d(\beta,c)(2n+1)^{\tilde{\beta}}\big{)}

which hold for all sufficiently large nn. Our result follows easily from this. ∎

We can now prove the main result of the section, which connects our considered class of functions GG with the admissible sequences appearing in \threfAdmissibleSequenceDef.

Proposition 5.8.
\thlabel

AdmissibleSequenceLemma If GG satisfies (ExpDec) and (LogLogInt), then {Mn}n0\{M_{n}\}_{n\geq 0} defined by

(5.9) Mn:=201G(1r)r2n+1𝑑rM_{n}:=2\int_{0}^{1}G(1-r)r^{2n+1}\,dr

is an admissible sequence.

Conversely, if {Mn}n0\{M_{n}\}_{n\geq 0} is an admissible sequence, then there exists a continuous and increasing function GG satisfying (ExpDec), (LogLogInt), G(0)=0G(0)=0, and for which the inequality

PG(2n+1)MnP_{G}(2n+1)\leq M_{n}

holds for all sufficiently large n0n\geq 0.

Proof.

We start by proving that the sequence in (5.9) is admissible by verifying the three conditions in \threfAdmissibleSequenceDef. By the Cauchy-Schwarz inequality, we have

Mn\displaystyle M_{n} =201G(1r)r(n1)+1/2r(n+1)+1/2𝑑r\displaystyle=2\int_{0}^{1}G(1-r)r^{(n-1)+1/2}r^{(n+1)+1/2}\,dr
Mn1Mn+1\displaystyle\leq\sqrt{M_{n-1}}\sqrt{M_{n+1}}

Thus {logMn}n0\{\log M_{n}\}_{n\geq 0} is a convex sequence. The inequality Mnexp(cn)M_{n}\leq\exp(-c\sqrt{n}) for some c>0c>0 and all sufficiently large n0n\geq 0 follows readily from (ExpDec) and an application of the upper estimate \threfMomentEstProp with β=1\beta=1 (and consequently β~=1/2\tilde{\beta}=1/2). Let mm be as in \threfMomentGGrowthLogLogInt. By the lower estimate in that lemma, we have

n0log1/Mn1+n2\displaystyle\sum_{n\geq 0}\frac{\log 1/M_{n}}{1+n^{2}} =n0log2logPG(2n+1)1+n2\displaystyle=\sum_{n\geq 0}\frac{-\log 2-\log P_{G}(2n+1)}{1+n^{2}}
n0log2+m(4n+2)+log(8n+4)1+n2.\displaystyle\leq\sum_{n\geq 0}\frac{-\log 2+m_{*}(4n+2)+\log(8n+4)}{1+n^{2}}.

The assumption that GG satisfies (LogLogInt) implies that 01logm(x)𝑑x<\int_{0}^{1}\log m(x)\,dx<\infty, and so from \threfLogIntLogSumEquivalenceProp we deduce that the last sum above is convergent. Consequently, {Mn}n0\{M_{n}\}_{n\geq 0} is an admissible sequence.

Conversely, assume that {Mn}n0\{M_{n}\}_{n\geq 0} is an admissible sequence. Since the sequence tends to zero, we may without loss of generality assume that M0<1M_{0}<1. From property (i)(i) in \threfAdmissibleSequenceDef we obtain the inequality

log1/Mn+1log1/Mnlog1/Mnlog1/Mn1,n0.\log 1/M_{n+1}-\log 1/M_{n}\leq\log 1/M_{n}-\log 1/M_{n-1},\quad n\geq 0.

This means that the slopes of the line segments between each consecutive pair of points in the sequence

(5.10) (2n+1,log1/Mn),n0(2n+1,\log 1/M_{n}),\quad n\geq 0

are decreasing, which means that if we define the function k(x)k(x), x>0x>0, as the piecewise linear interpolant of the data (5.10), then kk is concave, continuous, positive and increasing, and satisfies

k(2n+1)=log1/Mn,n0.k(2n+1)=\log 1/M_{n},\quad n\geq 0.

It also satisfies limxk(x)=+\lim_{x\to\infty}k(x)=+\infty, and property (iii)(iii) in \threfAdmissibleSequenceDef easily implies that

(5.11) 1k(x)1+x2𝑑x<.\int_{1}^{\infty}\frac{k(x)}{1+x^{2}}\,dx<\infty.

Let kk^{*} be the upper Legendre envelope of kk defined in (5.2), set

(5.12) G(x):=exp(k(x)),x(0,1],G(x):=\exp\big{(}-k^{*}(x)\big{)},\quad x\in(0,1],

and G(0)=0G(0)=0. Then GG is a continuous and increasing function. Define PGP_{G} as in (5.4). By \threfremarkLowerLegEnvelope, (5.12), inversion formula in \threfLegendreInversionFormula and \threfMomentGGrowthLogLogInt, we have the estimate

PG(x)exp(k(x))P_{G}(x)\leq\exp\big{(}-k(x)\big{)}

for all sufficiently large xx. Consequently,

PG(2n+1)MnP_{G}(2n+1)\leq M_{n}

if nn is large, since kk interpolates the data (5.10). \threfLogIntLogSumEquivalenceProp, \threfLegendreInversionFormula and (5.11) imply that

01loglog1/G(x)𝑑x=01logk(x)𝑑x<.\int_{0}^{1}\log\log 1/G(x)\,dx=\int_{0}^{1}\log k^{*}(x)\,dx<\infty.

Thus GG satisfies (LogLogInt). It remains to check that GG also satisfies (ExpDec). Note that property (ii)(ii) in \threfAdmissibleSequenceDef of the admissible sequence {Mn}n0\{M_{n}\}_{n\geq 0} implies easily that kk satisfies a lower bound of the form

k(x)dx,x0,k(x)\geq d\sqrt{x},\quad x\geq 0,

for some constant d>0d>0. But then, by (5.2), we have

k(x)\displaystyle k^{*}(x) =supy0k(y)xy\displaystyle=\sup_{y\geq 0}\,k(y)-xy
supy>0dyxy\displaystyle\geq\sup_{y>0}\,d\sqrt{y}-xy
=d24x.\displaystyle=\frac{d^{2}}{4x}.

The last equality can be derived by elementary calculus techniques. Consequently

lim infx0+xlog1/G(x)d24>0,\liminf_{x\to 0^{+}}x\log 1/G(x)\geq\frac{d^{2}}{4}>0,

and so GG satisfies (ExpDec). The proof is complete. ∎

5.4. Some auxiliary spaces of Taylor series

If f:𝔻f:\mathbb{D}\to\mathbb{C} is an analytic function and

(5.13) dμ𝔻(z)=G(1|z|)dA(z)d\mu_{\mathbb{D}}(z)=G(1-|z|)dA(z)

then we have the norm equality

(5.14) fμ𝔻2=𝔻|f(z)|2𝑑μ𝔻(z)=n0Mn(G)|fn|2\|f\|^{2}_{\mu_{\mathbb{D}}}=\int_{\mathbb{D}}|f(z)|^{2}d\mu_{\mathbb{D}}(z)=\sum_{n\geq 0}M_{n}(G)|f_{n}|^{2}

where {fn}n0\{f_{n}\}_{n\geq 0} is the sequence of Taylor coefficients of ff, and M={Mn(G)}n0M=\{M_{n}(G)\}_{n\geq 0} is given by (5.3). The above equality gives us an isometric isomorphism between 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}) and a space of Taylor series.

For a decreasing sequence M={Mn}n0M=\{M_{n}\}_{n\geq 0} of positive numbers we define H2(M)H_{2}(M) to be the Hilbert space of analytic functions in 𝔻\mathbb{D} consisting of f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} which satisfy

(5.15) fH2(M)2:=n0Mn|fn|2<.\|f\|^{2}_{H_{2}(M)}:=\sum_{n\geq 0}M_{n}|f_{n}|^{2}<\infty.

In our development, the sequences MM will be the admissible sequences studied in Section 5. Such sequences have the property that

limnMn1/n=1,\lim_{n\to\infty}M_{n}^{1/n}=1,

a condition which ensures that the spaces H2(M)H_{2}(M), and their duals, are genuine spaces of analytic functions on 𝔻\mathbb{D}. The dual space H2(M)H_{2}^{*}(M) is to consist of power series which satisfy

(5.16) fH2(M)2:=n0|fn|2Mn<.\|f\|^{2}_{H_{2}^{*}(M)}:=\sum_{n\geq 0}\frac{|f_{n}|^{2}}{M_{n}}<\infty.

Since M={Mn}n0M=\{M_{n}\}_{n\geq 0} is assumed to be decreasing, the space H2(M)H_{2}^{*}(M) is contained in the Hardy space H2H^{2}. In fact, if MM is an admissible sequence, then H2(M)H^{*}_{2}(M) consists of functions satisfying the condition (RSD) of Section 1. The duality between H2(M)H_{2}(M) and H2(M)H_{2}^{*}(M) is realized by the usual Cauchy pairing

(5.17) f,g:=limr1n0r2nfngn¯=𝕋fg¯𝑑m=f,g2\big{\langle}f,g\big{\rangle}:=\lim_{r\to 1^{-}}\sum_{n\geq 0}r^{2n}f_{n}\overline{g_{n}}=\int_{\mathbb{T}}f\overline{g}\,d\textit{m}=\big{\langle}f,g\big{\rangle}_{\mathcal{L}^{2}}

where the sequential definition above makes sense whenever fH2(M)f\in H_{2}(M), gH2(M)g\in H_{2}^{*}(M), and the integral definition holds only in special cases, for instance when f,gH2f,g\in H^{2}. An application of the Cauchy-Schwarz inequality to the limit in (5.17) shows that

|f,g|fH2(M)gH2(M).|\big{\langle}f,g\big{\rangle}|\leq\|f\|_{H_{2}(M)}\|g\|_{H^{*}_{2}(M)}.

We introduce also the space H1(M)H_{1}^{*}(M) which consists of power series f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} satisfying

(5.18) fH1(M):=supn0|fn|Mn<.\|f\|_{H^{*}_{1}(M)}:=\sup_{n\geq 0}\,\frac{|f_{n}|}{M_{n}}<\infty.

Recall from \threfPowerAdmSeq that the family of admissible sequences introduced in \threfAdmissibleSequenceDef is invariant under taking powers. For this reason, the spaces H2(M)H^{*}_{2}(M) and H1(M)H^{*}_{1}(M) which appear in our study are very similar.

Lemma 5.9.
\thlabel

H1starH2starEmbeddingLemma Let M={Mn}n0M=\{M_{n}\}_{n\geq 0} be an admissible sequence, and consider the sequences

Mp:={Mnp}n0.M^{p}:=\{M^{p}_{n}\}_{n\geq 0}.

For p>1/2p>1/2, we have the continuous embeddings

H1(Mp)H2(M)H1(M1/2).H^{*}_{1}(M^{p})\subset H^{*}_{2}(M)\subset H^{*}_{1}(M^{1/2}).
Proof.

If fH2(M)f\in H^{*}_{2}(M), then for any n0n\geq 0 we have that

|fn|2MnfH2(M)2,\frac{|f_{n}|^{2}}{M_{n}}\leq\|f\|^{2}_{H^{*}_{2}(M)},

so clearly fH1(M1/2)f\in H^{*}_{1}(M^{1/2}). If fH1(Mp)f\in H^{*}_{1}(M^{p}) for some p>1/2p>1/2, then we may use that MM satisfies property (ii)(ii) of admissible sequences in \threfAdmissibleSequenceDef to obtain

n0|fn|2Mn\displaystyle\sum_{n\geq 0}\frac{|f_{n}|^{2}}{M_{n}} fH1(Mp)2n0Mn2p1\displaystyle\leq\|f\|^{2}_{H^{*}_{1}(M^{p})}\sum_{n\geq 0}M_{n}^{2p-1}
n0exp(d(2p1)n)\displaystyle\leq\sum_{n\geq 0}\exp\big{(}-d(2p-1)\sqrt{n}\big{)}
<.\displaystyle<\infty.

Thus fH2(M)f\in H^{*}_{2}(M). ∎

The following corollary will be used several times below.

Corollary 5.10.
\thlabel

MhstarContainmentFromRSD If an analytic function ff in 𝔻\mathbb{D} satisfies the condition (RSD), then there exists c>0c^{\prime}>0 and a measure

dμ𝔻(z)=exp(c(1|z|))dA(z)=G(1|z|)dA(z)d\mu_{\mathbb{D}}(z)=\exp\Big{(}-\frac{c^{\prime}}{(1-|z|)}\Big{)}dA(z)=G(1-|z|)dA(z)

with sequence of moments M={Mn(G)}n0M=\{M_{n}(G)\}_{n\geq 0} such that fH2(M)f\in H^{*}_{2}(M).

Proof.

The condition (RSD) and \threfH1starH2starEmbeddingLemma imply that fH2(M~)f\in H^{*}_{2}(\widetilde{M}), where M~n=exp(c0n)\widetilde{M}_{n}=\exp(-c_{0}\sqrt{n}) for some positive constant c0c_{0}. Now \threfMomentEstProp, with β=1\beta=1, shows that c>0c^{\prime}>0 can be chosen so that Mn(G)=M(β,c)M~nM_{n}(G)=M(\beta,c^{\prime})\geq\widetilde{M}_{n} for sufficiently large nn. Then fH2(M)f\in H^{*}_{2}(M). ∎

We end the section with a few words about operators acting on the introduced class of spaces. From their definition, and in particular from the assumption on MM being decreasing, it is not hard to see that the spaces H2(M)H_{2}(M) are invariant under the multiplication operator MzM_{z}, and that this operator is a contraction. Then Von Neumann’s inequality ([1, p. 159]) or the Sz.-Nagy Foias HH^{\infty}-functional calculus ([29, Chapter 3]) shows that in fact every function hHh\in H^{\infty} defines a bounded multiplication operator Mh:H2(M)H2(M)M_{h}:H_{2}(M)\to H_{2}(M). The adjoint operator Mh:H2(M)H2(M)M_{h}^{*}:H_{2}^{*}(M)\to H_{2}^{*}(M) is easily indentified with the usual Toeplitz operator Th¯T_{\overline{h}} with the co-analytic symbol h¯\overline{h}, i.e., Th¯fT_{\overline{h}}f is the orthogonal projection to the Hardy space H2H^{2} of the function h¯f2(𝕋)\overline{h}f\in\mathcal{L}^{2}(\mathbb{T}).

For later reference, we record these observations in a proposition.

Proposition 5.11.
\thlabel

HToeplitzInvariance Let M={Mn}n0M=\{M_{n}\}_{n\geq 0} be an admissible sequence.

  1. (i)

    The space H2(M)H_{2}(M) is invariant for the analytic multiplication operators

    Mhf=h(z)f(z)hf,fH2(M),M_{h}f=h(z)f(z)hf,\quad f\in H_{2}(M),

    with symbols hHh\in H^{\infty}.

  2. (ii)

    The space H2(M)H^{*}_{2}(M) is invariant for the co-analytic Toeplitz operators

    Th¯f=P+h¯f,fH2(M)T_{\overline{h}}f=P_{+}\overline{h}f,\quad f\in H^{*}_{2}(M)

    with symbols hHh\in H^{\infty}.

Corollary 5.12.
\thlabel

ToeplitzInvRSD If an analytic function f:𝔻f:\mathbb{D}\to\mathbb{C} satisfies the condition (RSD), then so does Th¯fT_{\overline{h}}f for any hHh\in H^{\infty}.

Proof.

We use \threfMhstarContainmentFromRSD and \threfHToeplitzInvariance to see that the function Th¯fT_{\overline{h}}f is contained in a space H2(M)H^{*}_{2}(M), where MM is admissible. \threfH1starH2starEmbeddingLemma shows that Th¯fH1(M)T_{\overline{h}}f\in H^{*}_{1}(\sqrt{M}), so Th¯fT_{\overline{h}}f satisfies (RSD). ∎

6. Existence in (b)\mathcal{H}(b) of functions with rapid spectral decay

This section deals with proving \threfMainTheoremHbExistenceESD. In the proof, we will need a similar result in the context of model spaces, which we establish first. Next, we present some background theory of (b)\mathcal{H}(b)-spaces which will be needed in the proof of \threfMainTheoremHbExistenceESD, and also in the proof of \threfMainTheoremHbDensityESD given in the next section.

6.1. Corresponding result in model spaces

The following \threfAlphaKthetaBreakpointProp needed in the proof of \threfMainTheoremHbExistenceESD is known, and follows for instance from the work of Beurling in [3], or from a result of El-Fallah, Kellay and Seip in [10]. The mentioned results are much stronger than \threfAlphaKthetaBreakpointProp. Because the result is important for our further purposes, we shall use the estimates from Section 5 to give a simple proof of our version of the result.

Proposition 6.1.
\thlabel

AlphaKthetaBreakpointProp If θ\theta is a singular inner function, then the model space 𝒦θ\mathcal{K}_{\theta} contains no non-zero function f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} which satisfies (RSD).

Proof.

We will show that any fKθf\in K_{\theta} which satisfies (RSD) satisfies also f(0)=f0=0f(0)=f_{0}=0. Since 𝒦θ\mathcal{K}_{\theta} is invariant for the backward shift

Lf(z):=f(z)f(0)z,z𝔻,Lf(z):=\frac{f(z)-f(0)}{z},\quad z\in\mathbb{D},

and by \threfToeplitzInvRSD the function Lf=Tz¯fLf=T_{\overline{z}}f satisfies (RSD), the same argument will show that fn=Lnf(0)=0f_{n}=L^{n}f(0)=0 for n0n\geq 0. Thus f0f\equiv 0 will follow.

We apply \threfMhstarContainmentFromRSD to ff and obtain a measure μ𝔻\mu_{\mathbb{D}} with moment sequence MM such that fH2(M)f\in H^{*}_{2}(M). The measure μ𝔻\mu_{\mathbb{D}} is of the form (T1T1) (see Section 1) for β=1\beta=1 and w0w\equiv 0. By \threfCyclicityMainTheorem, the singular inner function θ\theta is trivially cyclic in 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}), since core(w)=core(0)=\text{core}(w)=\text{core}(0)=\varnothing. Thus there exists a sequence of analytic polynomials {pn}n0\{p_{n}\}_{n\geq 0} such that θpn1\theta p_{n}\to 1 in the norm of 𝒫2(μ)=H2(M)\mathcal{P}^{2}(\mu)=H_{2}(M). Using the duality pairing (5.17) and the membership of ff in H2(M)KθH^{*}_{2}(M)\cap K_{\theta}, the following computation is justified:

f(0)¯\displaystyle\overline{f(0)} =1,f\displaystyle=\big{\langle}1,f\big{\rangle}
=limnθpn,f\displaystyle=\lim_{n\to\infty}\big{\langle}\theta p_{n},f\big{\rangle}
=limn𝕋θpnf¯𝑑m\displaystyle=\lim_{n\to\infty}\int_{\mathbb{T}}\theta p_{n}\overline{f}\,d\textit{m}
=0.\displaystyle=0.

The last equality holds due to ff being a member of Kθ=(θH2)K_{\theta}=(\theta H^{2})^{\perp}. Thus f(0)=0f(0)=0, and the proof is complete by the initial remarks. ∎

6.2. Some (b)\mathcal{H}(b)-theory

The following description of (b)\mathcal{H}(b)-spaces is very convenient in connection with various functional-analytic arguments. It has been introduced in [2], and was later used in [21] and [23], to prove approximation results in classes of (b)\mathcal{H}(b)-spaces. We will employ it in a similar way below. Recall that the symbol P+P_{+} denotes the orthogonal projection operator P+:2(𝕋)H2P_{+}:\mathcal{L}^{2}(\mathbb{T})\to H^{2}, and 2(E)\mathcal{L}^{2}(E) denotes the subspace of those g2(𝕋)g\in\mathcal{L}^{2}(\mathbb{T}) which live only on the measurable subset E𝕋E\subset\mathbb{T}.

Proposition 6.2.
\thlabel

normformula Let bb be an extreme point of the unit ball of HH^{\infty},

(6.1) Δb(x)=1|b(x)|2,x𝕋,\Delta_{b}(x)=\sqrt{1-|b(x)|^{2}},\quad x\in\mathbb{T},

and EE be the carrier set of Δb\Delta_{b}:

E={x𝕋:Δb(x)>0}.E=\{x\in\mathbb{T}:\Delta_{b}(x)>0\}.

Then fH2f\in H^{2} is a member of (b)\mathcal{H}(b) if and only if the equation

(6.2) P+b¯f=P+ΔbgP_{+}\overline{b}f=-P_{+}\Delta_{b}g

has a solution g2(E)g\in\mathcal{L}^{2}(E). The solution is unique, and the map J:(b)H22(E)J:\mathcal{H}(b)\to H^{2}\oplus\mathcal{L}^{2}(E) defined by

Jf=(f,g),Jf=(f,g),

is an isometry. Moreover,

(6.3) J((b))={(bh,Δbh):hH2}.J(\mathcal{H}(b))^{\perp}=\Big{\{}(bh,\Delta_{b}h):h\in H^{2}\Big{\}}.

Next comes a very useful corollary which is well-known and can be proved by other means (see [11], [12] for other derivations).

Corollary 6.3.
\thlabel

CauchyTransformsinHb Let EE and Δb\Delta_{b} be as in \threfnormformula. For any s2(E)s\in\mathcal{L}^{2}(E), the function

f=P+Δbsf=P_{+}\Delta_{b}s

is a member of (b)\mathcal{H}(b) and, in the notation of \threfnormformula, we have

Jf=(f,b¯s).Jf=(f,-\overline{b}s).

Moreover, if bb is extreme and ss is non-zero, then ff is non-zero.

Proof.

We compute

P+b¯f=P+b¯P+Δbs=P+b¯Δbs=P+Δbb¯s,P_{+}\overline{b}f=P_{+}\overline{b}P_{+}\Delta_{b}s=P_{+}\overline{b}\Delta_{b}s=P_{+}\Delta_{b}\overline{b}s,

and so (6.2) holds for the pair (f,g):=(P+Δbs,b¯s)(f,g):=(P_{+}\Delta_{b}s,-\overline{b}s). If bb is extreme, then logΔb1(𝕋)\log\Delta_{b}\not\in\mathcal{L}^{1}(\mathbb{T}), and it follows readily that also log(Δb|s|)1(𝕋)\log(\Delta_{b}|s|)\not\in\mathcal{L}^{1}(\mathbb{T}). A function hkerP+h\in\ker P_{+} is conjugate-analytic, and so log|h|1(𝕋)\log|h|\in\mathcal{L}^{1}(\mathbb{T}) if h0h\neq 0. So ΔbskerP+\Delta_{b}s\not\in\ker P_{+} if ss is non-zero, and it follows that ff is non-zero. ∎

Corollary 6.4.
\thlabel

TconjbInvariance The Toeplitz operator Tb¯T_{\overline{b}} acts boundedly on (b)\mathcal{H}(b). If f(b)f\in\mathcal{H}(b) and Jf=(f,g)Jf=(f,g), then

Tb¯f=(Tb¯f,b¯g).T_{\overline{b}}f=(T_{\overline{b}}f,\overline{b}g).
Proof.

Again, we only need to verify that (6.2) holds for the given pairs. This follows easily by applying the operator Tb¯T_{\overline{b}} to both sides of the equation (6.2) and computing as in the proof of \threfCauchyTransformsinHb. ∎

6.3. Main tool in the proof of \threfMainTheoremHbExistenceESD: residuals

We will now need to introduce the notion of residual sets.

Definition 6.5.

(Residual sets of weights) \thlabelResDef Let w1(𝕋)w\in\mathcal{L}^{1}(\mathbb{T}) and consider the carrier set

E={x𝕋:w(x)>0}.E=\{x\in\mathbb{T}:w(x)>0\}.

We define res(w)\text{res}(w) to be the set

res(w)=Ecore(w),\text{res}(w)=E\setminus\text{core}(w),

where core(w)\text{core}(w) is the set appearing in \threfCoreDef.

Since EE might only be defined up to a set of mm-measure zero, the same is true for the residual res(w)\text{res}(w) of any weight ww. This will not cause us any problems.

We have introduced the residuals because of their crucial role in the following special case of [24, Theorem A].

Lemma 6.6.
\thlabel

ResSetMainLemma Assume that w1(𝕋)w\in\mathcal{L}^{1}(\mathbb{T}) is a weight for which res(w)\text{res}(w) has positive mm-measure. Let wr=w|res(w)w_{r}=w|\text{res}(w) be the restriction of the weight ww to the set res(w)\text{res}(w). Then we have the containment

2(wrdm)𝒫2(μ)\mathcal{L}^{2}(w_{r}d\textit{m})\subset\mathcal{P}^{2}(\mu)

whenever μ\mu is of the form (1.1) with GG satisfying (ExpDec).

6.4. Proof of \threfMainTheoremHbExistenceESD

In \threfMainTheoremHbExistenceESD, it is obvious that (ii)(i)(ii)\Rightarrow(i). We can thus prove the theorem by showing validity of the implications (i)(iii)(i)\Rightarrow(iii) and (iii)(ii)(iii)\Rightarrow(ii).

Let us first show that (i)(iii)(i)\Rightarrow(iii), and so we assume that f(z)=n0fnzn(b)f(z)=\sum_{n\geq 0}f_{n}z^{n}\in\mathcal{H}(b) is non-zero and that it satisfies (RSD). We may assume that bb does not vanish at any point in 𝔻\mathbb{D}, else (iii)(iii) certainly holds. Similarly to as it was done in the proof of \threfAlphaKthetaBreakpointProp, we use \threfMhstarContainmentFromRSD to obtain a measure

(6.4) dμ\displaystyle d\mu =dμ𝔻+dμ𝕋\displaystyle=d\mu_{\mathbb{D}}+d\mu_{\mathbb{T}}
=exp(c(1|z|))dA(z)+Δbdm,\displaystyle=\exp\Big{(}-\frac{c^{\prime}}{(1-|z|)}\Big{)}dA(z)+\Delta_{b}\,d\textit{m},

and a sequence MM such that the identity map between 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}) and H2(M)H_{2}(M) is an isometry, and fH2(M)f\in H^{*}_{2}(M). By \threfHToeplitzInvariance, the space H2(M)H^{*}_{2}(M) is invariant under Toeplitz operators with co-analytic symbols, and consequently we also have Tb¯fH2(M)T_{\overline{b}}f\in H^{*}_{2}(M). By \threfTconjbInvariance and (6.2) we have Tb¯f=P+b¯f(b)H2(M)T_{\overline{b}}f=P_{+}\overline{b}f\in\mathcal{H}(b)\cap H^{*}_{2}(M) and

(6.5) Tb¯f=P+ΔbgT_{\overline{b}}f=P_{+}\Delta_{b}g

for some g2(E)g\in\mathcal{L}^{2}(E). The kernel of the operator Tb¯T_{\overline{b}} is the model space 𝒦Ib\mathcal{K}_{I_{b}}, where IbI_{b} is the inner factor of bb. Since bb does not vanish in 𝔻\mathbb{D}, it follows that IbI_{b} is a purely singular inner function. Every function in H2(M)H^{*}_{2}(M) satisfies (RSD), so \threfAlphaKthetaBreakpointProp implies that 𝒦IbH2(M)={0}\mathcal{K}_{I_{b}}\cap H^{*}_{2}(M)=\{0\}. Consequently f𝒦Ibf\not\in\mathcal{K}_{I_{b}}, so Tb¯f0T_{\overline{b}}f\neq 0, and Δbg0\Delta_{b}g\neq 0 by (6.5). If

Tb¯f(z)=n0cnznT_{\overline{b}}f(z)=\sum_{n\geq 0}c_{n}z^{n}

is the Taylor expansion of Tb¯f(z)T_{\overline{b}}f(z), then a consequence of the membership Tb¯fH2(M)T_{\overline{b}}f\in H^{*}_{2}(M) is that the function

F(z):=n0cnMnzn,z𝔻F(z):=\sum_{n\geq 0}\frac{c_{n}}{M_{n}}z^{n},\quad z\in\mathbb{D}

is a member of H2(M)H_{2}(M). The function FF lives on 𝔻\mathbb{D}, the function gg lives on 𝕋\mathbb{T}, and hence FgF-g defines a function on 𝔻¯\overline{\mathbb{D}}. The condition FH2(M)F\in H_{2}(M) means simply that FF is square-integrable with respect to the part μ𝔻\mu_{\mathbb{D}} of μ\mu in (6.4) which lives on 𝔻\mathbb{D}. The containment g2(Δbdm)=2(μ𝕋)g\in\mathcal{L}^{2}(\Delta_{b}d\textit{m})=\mathcal{L}^{2}(\mu_{\mathbb{T}}) is ensured by the boundedness of Δb\Delta_{b} and the containment gL2(E)g\in L^{2}(E). Thus Fg2(μ)F-g\in\mathcal{L}^{2}(\mu). The representation (6.5) tells us that the positive Fourier coefficients {cn}n0\{c_{n}\}_{n\geq 0} of Tb¯fT_{\overline{b}}f and of Δbg\Delta_{b}g coincide. Our definitions then imply that the function FgF-g is orthogonal to the analytic polynomials in 2(μ)\mathcal{L}^{2}(\mu). Since Δbg0\Delta_{b}g\neq 0, the function gg is a non-zero element of 2(μ𝕋)\mathcal{L}^{2}(\mu_{\mathbb{T}}). The conclusion is that there exists an element (namely FgF-g) inside 2(μ)\mathcal{L}^{2}(\mu) which is orthogonal to 𝒫2(μ)\mathcal{P}^{2}(\mu) and which does not vanish identically on the circle 𝕋\mathbb{T}. If there existed no interval on which logΔb\log\Delta_{b} was integrable, then core(Δb)=\text{core}(\Delta_{b})=\varnothing, and so \threfResSetMainLemma would imply that the entire space 2(Δbdm)\mathcal{L}^{2}(\Delta_{b}d\textit{m}) is contained in 𝒫2(μ)\mathcal{P}^{2}(\mu). Clearly that would be a contradiction to FgF-g being orthogonal to 𝒫2(μ)\mathcal{P}^{2}(\mu). Thus such an interval exists, and we have proved that (i)(iii)(i)\Rightarrow(iii).

The implication (iii)(ii)(iii)\Rightarrow(ii) is easier. Let M={Mn}n0M=\{M_{n}\}_{n\geq 0} be an admissible sequence. We must show that (b)\mathcal{H}(b) contains a function in H1(M)H^{*}_{1}(M). If bb vanishes at some point of 𝔻\mathbb{D}, then the implication is trivial. Assume therefore that logΔb\log\Delta_{b} is integrable on some (say, open) interval II which is not all of 𝕋\mathbb{T}, and let w=Δb2|Iw=\Delta^{2}_{b}|I be the restriction of Δb2\Delta^{2}_{b} to the interval II. By \threfPowerAdmSeq and \threfAdmissibleSequenceLemma there exists a function GG which satisfies (ExpDec), (LogLogInt), with corresponding moment sequence

M~={M~n}n0={Mn(G)}n0\widetilde{M}=\{\widetilde{M}_{n}\}_{n\geq 0}=\{M_{n}(G)\}_{n\geq 0}

satisfying

M~nMn2\widetilde{M}_{n}\leq M^{2}_{n}

for large nn. If

dμ(z)=G(1|z|)dA(z)+w(z)dm(z),d\mu(z)=G(1-|z|)dA(z)+w(z)d\textit{m}(z),

then the space 𝒫2(μ)\mathcal{P}^{2}(\mu) is irreducible by \threfIrrDef, since core(w)\text{core}(w) coincides with II, which is a carrier of ww. By irreducibility we have that 2(wdm)𝒫2(μ)\mathcal{L}^{2}(w\,d\textit{m})\not\subset\mathcal{P}^{2}(\mu). So there must exist a non-zero element Fg2(μ)F-g\in\mathcal{L}^{2}(\mu), with FF being an analytic function on 𝔻\mathbb{D} and gg living on I𝕋I\subset\mathbb{T}, which is orthogonal to 𝒫2(μ)\mathcal{P}^{2}(\mu) in 2(μ)\mathcal{L}^{2}(\mu). We can’t have g0g\equiv 0, for then the Taylor coefficients of FF would all vanish by the orthogonality to analytic monomials, and consequently FgF-g would reduce to the zero element. The orthogonality means that

FnM~n=(wg)n,n0F_{n}\widetilde{M}_{n}=(wg)_{n},\quad n\geq 0

where {Fn}n0\{F_{n}\}_{n\geq 0} are the Taylor coefficients of FF and (wg)n(wg)_{n} are the non-negative Fourier coefficients of wgwg. For large nn, we have the estimate

|(wg)n|2\displaystyle|(wg)_{n}|^{2} =|FnM~n|2\displaystyle=|F_{n}\widetilde{M}_{n}|^{2}
M~nm0|Fm|2M~m\displaystyle\leq\widetilde{M}_{n}\sum_{m\geq 0}|F_{m}|^{2}\widetilde{M}_{m}
=M~nFμ𝔻2\displaystyle=\widetilde{M}_{n}\|F\|^{2}_{\mu_{\mathbb{D}}}
Mn2Fμ𝔻2.\displaystyle\leq M_{n}^{2}\|F\|^{2}_{\mu_{\mathbb{D}}}.

Thus P+wgP_{+}wg is a member of H1(M)H^{*}_{1}(M). Since gg lives on II and gΔb2(I)g\Delta_{b}\in\mathcal{L}^{2}(I), we have by \threfCauchyTransformsinHb that P+wg=P+ΔbΔbg(b)P_{+}wg=P_{+}\Delta_{b}\Delta_{b}g\in\mathcal{H}(b). This function is non-zero since log(w|g|)1(𝕋)\log(w|g|)\not\in\mathcal{L}^{1}(\mathbb{T}) by the choice of I𝕋I\neq\mathbb{T}. Thus (iii)(ii)(iii)\Rightarrow(ii), and we have completed our proof of \threfMainTheoremHbExistenceESD.

7. Density in (b)\mathcal{H}(b) of functions with rapid spectral decay

The main result of [23] characterizes the density of the functions in (b)\mathcal{H}(b) which have Taylor series f(z)=n0fnznf(z)=\sum_{n\geq 0}f_{n}z^{n} satisfying |fn|=𝒪(1/nk)|f_{n}|=\mathcal{O}(1/n^{k}), for positive kk. The characterization is in terms of the structure of MzM_{z}-invariant subspaces of 𝒫2(μ)\mathcal{P}^{2}(\mu) with μ\mu of form (1.1) and G(t)=tkG(t)=t^{k}, k0k\geq 0. The proofs in [23] in fact carry over more-or-less verbatim from the case considered there to many other function classes defined by their spectral size, with the family of functions defined by conditions such as (1.13) being no exception. Thus, in fact, \threfMainTheoremHbDensityESD is more or less a direct consequence of \threfIrrDef, \threfCyclicityMainTheorem and \threfPermanenceMainTheorem. For reasons of completeness of the present work, we outline an argument which is in parts new, leads to a proof of \threfMainTheoremHbDensityESD, but also gives additional bits of information regarding which functions in (b)\mathcal{H}(b) lie outside of the closure of functions satisfying spectral decay properties as in (RSD).

As before, Δb(x)=1|b(x)|\Delta_{b}(x)=\sqrt{1-|b(x)|} for x𝕋x\in\mathbb{T}, and we let

b=BSνb0b=BS_{\nu}b_{0}

be the inner-outer factorization of bb, with BB a Blaschke product, SνS_{\nu} a singular inner function, and b0b_{0} an outer function. We denote by Ib=BSνI_{b}=BS_{\nu} the inner factor of bb.

Lemma 7.1.
\thlabel

ResVanishingLemma Let w1(𝕋)w\in\mathcal{L}^{1}(\mathbb{T}) be non-negative, and assume that for some g2(wdm)g\in\mathcal{L}^{2}(w\,d\textit{m}) the function P+wgP_{+}wg satisfies (RSD). Then gwgw vanishes on res(w)\text{res}(w).

Proof.

We use \threfMhstarContainmentFromRSD to obtain a measure μ\mu as in (T1T1) of Section 1, and with the parameters β=1\beta=1 and c>0c>0 chosen so that if M={Mn}n0M=\{M_{n}\}_{n\geq 0} is the sequence of moments corresponding to μ𝔻\mu_{\mathbb{D}}, then P+wgH2(M)P_{+}wg\in H^{*}_{2}(M). Let hh be a bounded function living on res(w)\text{res}(w), and {pn}n0\{p_{n}\}_{n\geq 0} be a sequence of analytic polynomials which converges to hh in the norm of 𝒫2(μ)\mathcal{P}^{2}(\mu). This is possible by \threfResSetMainLemma. In particular, this convergence implies that pn0p_{n}\to 0 in 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}), or in other words, pn0p_{n}\to 0 in H2(M)H_{2}(M). Simultaneously, we have that pnhp_{n}\to h in 2(wdm)\mathcal{L}^{2}(w\,d\textit{m}). Using the duality pairing (5.17), we obtain

0\displaystyle 0 =0,P+wg\displaystyle=\big{\langle}0,P_{+}wg\big{\rangle}
=limnpn,P+wg\displaystyle=\lim_{n\to\infty}\big{\langle}p_{n},P_{+}wg\big{\rangle}
=limnpn,P+wg2\displaystyle=\lim_{n\to\infty}\big{\langle}p_{n},P_{+}wg\big{\rangle}_{\mathcal{L}^{2}}
=limnpn,wg2\displaystyle=\lim_{n\to\infty}\big{\langle}p_{n},wg\big{\rangle}_{\mathcal{L}^{2}}
=𝕋hg¯w𝑑m.\displaystyle=\int_{\mathbb{T}}h\overline{g}\,w\,d\textit{m}.

Since hh is an arbitrary bounded function living on res(w)\text{res}(w), it follows that gw0gw\equiv 0 on res(w)\text{res}(w). ∎

Proposition 7.2.
\thlabel

Prop1TheoremD

Assume that the set res(Δb)\text{res}(\Delta_{b}) has positive mm-measure, and let s2(𝕋)s\in\mathcal{L}^{2}(\mathbb{T}) be a non-zero function which vanishes outside of res(Δb)\text{res}(\Delta_{b}). Then the non-zero function

f=P+Δbs(b)f=P_{+}\Delta_{b}s\in\mathcal{H}(b)

lies outside of the norm-closure in (b)\mathcal{H}(b) of functions satisfying (RSD).

Proof.

Seeking a contradiction, assume that {hn}n\{h_{n}\}_{n} is a sequence of functions in (b)\mathcal{H}(b) which satisfy (RSD) and which converge in the norm of (b)\mathcal{H}(b) to the given ff. In the notation of \threfnormformula, we consider Jhn=(hn,kn)H22(E)Jh_{n}=(h_{n},k_{n})\in H^{2}\oplus\mathcal{L}^{2}(E) and Jf=(f,g)H22(E)Jf=(f,g)\in H^{2}\oplus\mathcal{L}^{2}(E), where g=b¯sg=-\overline{b}s according to \threfCauchyTransformsinHb. By \threfTconjbInvariance, Tb¯hnT_{\overline{b}}h_{n} converges to Tb¯fT_{\overline{b}}f in the norm of (b)\mathcal{H}(b), and since the embedding JJ of \threfnormformula is an isometry, \threfTconjbInvariance moreover implies that b¯kn\overline{b}k_{n} converges to b¯g=b2¯s\overline{b}g=-\overline{b^{2}}s in 2(𝕋)\mathcal{L}^{2}(\mathbb{T}). In particular, this implies that knk_{n} cannot all simultaneously vanish on res(Δb)\text{res}(\Delta_{b}), since ss lives only on that set. But Tb¯hnT_{\overline{b}}h_{n} satisfies (RSD) (since hnh_{n} does), and Tb¯hn=P+b¯hn=P+Δbb¯knT_{\overline{b}}h_{n}=P_{+}\overline{b}h_{n}=P_{+}\Delta_{b}\overline{b}k_{n} by \threfTconjbInvariance. Thus by \threfResVanishingLemma, the functions Δbb¯kn\Delta_{b}\overline{b}k_{n} must vanish on res(Δb)\text{res}(\Delta_{b}), and consequently knk_{n} must vanish on res(Δb)\text{res}(\Delta_{b}), since b¯Δb\overline{b}\Delta_{b} is non-zero mm-almost everywhere on that set. This is the desired contradiction. ∎

We have now proved that it is necessary for core(Δb)\text{core}(\Delta_{b}) to be a carrier for Δb\Delta_{b} if functions satisfying (RSD) are to be dense in (b)\mathcal{H}(b). In the next proposition, we assume that core(Δb)\text{core}(\Delta_{b}) is a carrier for Δb\Delta_{b}, and show that if SνS_{\nu} is the singular inner factor of bb and ν\nu places some portion of its mass outside of the core of Δb\Delta_{b}, then again functions satisfying (RSD) are not dense in (b)\mathcal{H}(b). And again, we do it by exhibiting explicit functions in (b)\mathcal{H}(b) which cannot be approximated in this way.

Proposition 7.3.
\thlabel

Prop2TheoremD Assume that core(Δb)\text{core}(\Delta_{b}) is a carrier for Δb\Delta_{b} and that

ν(𝕋core(Δb))>0,\nu\big{(}\mathbb{T}\setminus\text{core}(\Delta_{b})\big{)}>0,

where SνS_{\nu} is the singular inner factor of bb. Decompose the measure ν\nu as

ν=νr+νc,\nu=\nu_{r}+\nu_{c},

where νr\nu_{r} is the restriction of ν\nu to the set 𝕋core(Δb)\mathbb{T}\setminus\text{core}(\Delta_{b}), and νc\nu_{c} is the restriction of ν\nu to the set core(Δb)\text{core}(\Delta_{b}). Then all functions in the subspace

(b/Sνr)𝒦Sνr=BSνcb0𝒦Sνr(b)(b/S_{\nu_{r}})\mathcal{K}_{S_{\nu_{r}}}=BS_{\nu_{c}}b_{0}\mathcal{K}_{S_{\nu_{r}}}\subset\mathcal{H}(b)

are orthogonal in (b)\mathcal{H}(b) to functions satisfying (RSD), 𝒦Sνr\mathcal{K}_{S_{\nu_{r}}} being the model space generated by the singular inner function SνrS_{\nu_{r}}.

Proof.

Take a function f=BSνcb0sf=BS_{\nu_{c}}b_{0}s, where s𝒦Sνrs\in\mathcal{K}_{S_{\nu_{r}}}, and hh satisfying (RSD). In the notation of \threfnormformula, a computation shows that Jf=(f,g)Jf=(f,g), where

g=ΔbSνr¯s.g=\Delta_{b}\overline{S_{\nu_{r}}}s.

Let 𝒫2(μ)\mathcal{P}^{2}(\mu) and H2(M)=𝒫2(μ𝔻)H_{2}(M)=\mathcal{P}^{2}(\mu_{\mathbb{D}}) be as in the proof of \threfResVanishingLemma, with w=Δbw=\Delta_{b} and the sequence MM being chosen so that hH2(M)h\in H^{*}_{2}(M). This time the space 𝒫2(μ)\mathcal{P}^{2}(\mu) is irreducible, and by \threfCyclicityMainTheorem the singular inner function SνrS_{\nu_{r}} is cyclic in 𝒫2(μ)\mathcal{P}^{2}(\mu). Hence there exists a sequence of analytic polynomials {pn}n\{p_{n}\}_{n} such that SνrpnS_{\nu_{r}}p_{n} converges to the function sH2s\in H^{2} in the norm of 𝒫2(μ)\mathcal{P}^{2}(\mu), and in particular in the norm of H2(M)H_{2}(M). Multiplying this sequence by BSνcb0BS_{\nu_{c}}b_{0}, it follows from \threfHToeplitzInvariance that bpnbp_{n} converges to ff in H2(M)H_{2}(M). Simultaneously, the 𝒫2(μ)\mathcal{P}^{2}(\mu)-convergence implies that SνrpnS_{\nu_{r}}p_{n} converge to ss in 2(Δbdm)\mathcal{L}^{2}(\Delta_{b}d\textit{m}), and since SνrS_{\nu_{r}} is unimodular on 𝕋\mathbb{T}, in fact we have that pnp_{n} converge to Sνr¯s\overline{S_{\nu_{r}}}s in 2(Δbdm)\mathcal{L}^{2}(\Delta_{b}d\textit{m}). Let Jh=(h,k)Jh=(h,k) be the corresponding pair for hh. We can use the above claims to compute

h,f2+k,g2\displaystyle\big{\langle}h,f\big{\rangle}_{\mathcal{L}^{2}}+\big{\langle}k,g\big{\rangle}_{\mathcal{L}^{2}} =h,f+k,ΔbSνr¯s2\displaystyle=\big{\langle}h,f\big{\rangle}+\big{\langle}k,\Delta_{b}\overline{S_{\nu_{r}}}s\big{\rangle}_{\mathcal{L}^{2}}
=limnh,bpn+k,Δbpn2\displaystyle=\lim_{n\to\infty}\big{\langle}h,bp_{n}\big{\rangle}+\big{\langle}k,\Delta_{b}p_{n}\big{\rangle}_{\mathcal{L}^{2}}
=limnh,bpn2+k,Δbpn2\displaystyle=\lim_{n\to\infty}\big{\langle}h,bp_{n}\big{\rangle}_{\mathcal{L}^{2}}+\big{\langle}k,\Delta_{b}p_{n}\big{\rangle}_{\mathcal{L}^{2}}
=limnP+(b¯h+Δbk),pn2\displaystyle=\lim_{n\to\infty}\big{\langle}P_{+}(\overline{b}h+\Delta_{b}k),p_{n}\big{\rangle}_{\mathcal{L}^{2}}
=limn0,pn2\displaystyle=\lim_{n\to\infty}\big{\langle}0,p_{n}\big{\rangle}_{\mathcal{L}^{2}}
=0.\displaystyle=0.

In the last step we used condition (6.2) for the pair (h,k)(h,k). Since the embedding JJ in \threfnormformula is an isometry, it follows that ff is orthogonal to hh in (b)\mathcal{H}(b). ∎

Proof of \threfMainTheoremHbDensityESD.

We see from \threfProp1TheoremD and \threfProp2TheoremD above that condition (iii)(iii) in \threfMainTheoremHbDensityESD is necessary in order for (i)(i) to hold. Since (ii)(ii) implies (i)(i), it suffices thus to show that (iii)(iii) implies (ii)(ii). The argument is essentially same as the one appearing in [23] and [21], we include it only for completeness.

Just as in the proof of \threfMainTheoremHbExistenceESD, given an admissible sequence M={Mn}n0M=\{M_{n}\}_{n\geq 0} we use \threfPowerAdmSeq and \threfAdmissibleSequenceLemma to obtain GG satisfying (ExpDec), (LogLogInt), with moment sequence M~={M~n}n0\widetilde{M}=\{\widetilde{M}_{n}\}_{n\geq 0} satisfying M~nMn2\widetilde{M}_{n}\leq M^{2}_{n} for large nn. We must show that (b)H1(M)\mathcal{H}(b)\cap H^{*}_{1}(M) is dense in (b)\mathcal{H}(b). By \threfH1starH2starEmbeddingLemma it will suffice to show that H2(M2)(b)H^{*}_{2}(M^{2})\cap\mathcal{H}(b) is dense in (b)\mathcal{H}(b).

The space 𝒫2(μ)\mathcal{P}^{2}(\mu) constructed from the measure

dμ(z)=G(1|z|)dA(z)+Δb2(z)dm(z)d\mu(z)=G(1-|z|)dA(z)+\Delta_{b}^{2}(z)d\textit{m}(z)

is irreducible by \threfIrrDef. Let us assume that f(b)f\in\mathcal{H}(b) is orthogonal to H2(M2)(b)H^{*}_{2}(M^{2})\cap\mathcal{H}(b). We will show that f=0f=0, which will prove \threfMainTheoremHbDensityESD. Because the mapping JJ in \threfnormformula is an isometry, it follows that JfJf is orthogonal to J(H2(M2)(b))J(H^{*}_{2}(M^{2})\cap\mathcal{H}(b)). Note that J(H2(M2)(b))J(H^{*}_{2}(M^{2})\cap\mathcal{H}(b)) is a subset of H2(M2)2(E)H^{*}_{2}(M^{2})\oplus\mathcal{L}^{2}(E), and under the duality pairing (5.17) between H2(M2)H_{2}(M^{2}) and H2(M2)H^{*}_{2}(M^{2}), we have

(7.1) J(H2(M2)(b))=hH2kerlh,J(H^{*}_{2}(M^{2})\cap\mathcal{H}(b))=\cap_{h\in H^{2}}\ker l_{h},

where lhl_{h} is the functional on H2(M2)2(E)H^{*}_{2}(M^{2})\oplus\mathcal{L}^{2}(E) which acts by the formula

lh(f,g):=f,bh+g,Δbh2.l_{h}(f,g):=\big{\langle}f,bh\big{\rangle}+\big{\langle}g,\Delta_{b}h\big{\rangle}_{\mathcal{L}^{2}}.

This follows readily from \threfnormformula (see, for instance, the argument in [23]). The fact that JfJf annihilates J(H2(M2)(b))J(H^{*}_{2}(M^{2})\cap\mathcal{H}(b)) and that (7.1) holds implies that JfJf is contained in the weak-star closure of the linear manifold {lh}H2H2(M2)2(E)\{l_{h}\}_{H^{2}}\subseteq H_{2}(M^{2})\oplus\mathcal{L}^{2}(E). Since the pairing between H2(M2)H_{2}(M^{2}) and H2(M2)H^{*}_{2}(M^{2}) is reflexive and {lh}hH2\{l_{h}\}_{h\in H^{2}} is a convex set, basic functional analysis says that, in fact, JfJf is contained in the norm-closure of {lh}hH2\{l_{h}\}_{h\in H^{2}}. Thus there exists a sequence {hk}k1\{h_{k}\}_{k\geq 1} with hkH2h_{k}\in H^{2} such that

(7.2) (bhk,Δbhk)Jf:=(f,g)(bh_{k},\Delta_{b}h_{k})\to Jf:=(f,g)

in the norm of H2(M2)2(E)H_{2}(M^{2})\oplus\mathcal{L}^{2}(E). Multiply the second coordinate by bb to obtain

(7.3) (bhk,Δbbhk)(f,bg).(bh_{k},\Delta_{b}bh_{k})\to(f,bg).

But the inequalities M~nMn2\widetilde{M}_{n}\leq M^{2}_{n} imply that bhkbh_{k} converges to ff also in the space 𝒫2(μ𝔻)\mathcal{P}^{2}(\mu_{\mathbb{D}}), and so in fact (7.3) tells us that {bhk}n\{bh_{k}\}_{n} is a Cauchy sequence in 𝒫2(μ)\mathcal{P}^{2}(\mu), to which \threfPermanenceMainTheorem applies. The critical conclusion is that bhkfbh_{k}\to f in the irreducible 𝒫2(μ)\mathcal{P}^{2}(\mu). If IbI_{b} is the inner factor of bb, then \threfPermanenceMainTheorem implies that f/IbH2f/I_{b}\in H^{2}, and by the irreducibility of 𝒫2(μ)\mathcal{P}^{2}(\mu) the sequence bhkbh_{k} on 𝕋\mathbb{T} must converge to the boundary function of ff on 𝕋\mathbb{T}. Thus g=Δbf/bg=\Delta_{b}f/b by (7.3), and Jf=(f,Δbf/b)Jf=(f,\Delta_{b}f/b). By \threfnormformula we get that

(7.4) 0=P+(b¯f+Δbg)=P+(b¯f+Δb2f/b)=P+(|b|2f/b+Δb2f/b)=P+(f/b).0=P_{+}(\overline{b}f+\Delta_{b}g)=P_{+}(\overline{b}f+\Delta_{b}^{2}f/b)=P_{+}(|b|^{2}f/b+\Delta_{b}^{2}f/b)=P_{+}(f/b).

From the above computation we infer that, in terms of boundary values, we have f/b=b¯f+Δbg2(𝕋)f/b=\overline{b}f+\Delta_{b}g\in\mathcal{L}^{2}(\mathbb{T}), and consequently f/bf/b has square-integrable boundary values. Since f/IbH2f/I_{b}\in H^{2}, it follows from the classical Smirnov maximum principle that f/bH2f/b\in H^{2}. Then f/bf/b is an analytic function which projects to 0 under P+P_{+}, which implies that f/b=0f/b=0, and consequently f=0f=0. ∎

8. Proof of \threfUncertThmRSD

A proof of \threfUncertThmRSD relies on a judicious application of \threfResVanishingLemma.

Proof of \threfUncertThmRSD.

If CνC_{\nu} satisfies (RSD), then the function f(z)=n0νnznf(z)=\sum_{n\geq 0}\nu_{n}z^{n}, z𝕋z\in\mathbb{T}, is certainly smooth on 𝕋\mathbb{T} and it has an analytic extension to 𝔻\mathbb{D}. Since the Cauchy transform of the measure dνfdmd\nu-f\,d\textit{m} vanishes in 𝔻\mathbb{D}, this measure must be absolutely continuous with respect to mm by the classical theorem of brothers Riesz. Hence dνd\nu is also absolutely continuous. Let g1(𝕋)g\in\mathcal{L}^{1}(\mathbb{T}) be its Radon-Nikodym derivative, so that dν=gdmd\nu=g\,d\textit{m}. Set f=𝒞ν=𝒞gf=\mathcal{C}_{\nu}=\mathcal{C}_{g}, which by our assumption is a function satisfying (RSD). Unfortunately, we cannot directly apply \threfResVanishingLemma since we do not necessarily have that g2(𝕋)g\in\mathcal{L}^{2}(\mathbb{T}). We must take care of this slight inconvenience to prove the theorem.

\thref

ToeplitzInvRSD says that Th¯fT_{\overline{h}}f satisfies (RSD), where Th¯T_{\overline{h}} is any co-analytic Toeplitz operator with bounded symbol hHh\in H^{\infty}. Moreover, Th¯fT_{\overline{h}}f has the representation

Th¯f(z)=𝒞h¯g(z),z𝔻.T_{\overline{h}}f(z)=\mathcal{C}_{\overline{h}g}(z),\quad z\in\mathbb{D}.

The above formula can be derived by first showing through simple algebraic manipulations that it holds for h(z):=zh(z):=z, then for analytic monomials by iteration, and thus for analytic polynomials by linearity. Finally, fix a uniformly bounded sequence of analytic polynomials {pn}n1\{p_{n}\}_{n\geq 1} which converges to hh pointwise mm-almost everywhere on 𝕋\mathbb{T} (the polynomials pnp_{n} could be taken to be the Cesàro means of the partial sums of the Taylor series of hh). For such a sequence we readily see from the dominated convergence theorem that

Th¯f(z)=limnTpn¯f(z)=limn𝒞pn¯g(z)=𝒞h¯g(z),z𝔻.T_{\overline{h}}f(z)=\lim_{n\to\infty}T_{\overline{p_{n}}}f(z)=\lim_{n\to\infty}\mathcal{C}_{\overline{p_{n}}g}(z)=\mathcal{C}_{\overline{h}g}(z),\quad z\in\mathbb{D}.

Since gL1(𝕋)g\in L^{1}(\mathbb{T}), in particular we have that log+|g|L1(𝕋)\log^{+}|g|\in L^{1}(\mathbb{T}), and this means that an outer function hHh\in H^{\infty} exists which satisfies the boundary value equation

|h(x)|=min(1,1/|g(x)|)|h(x)|=\min\Big{(}1,1/|g(x)|\Big{)}

for m-almost every x𝕋x\in\mathbb{T}. Set also

w(x)=min(1,|g(x)|).w(x)=\min\Big{(}1,|g(x)|\Big{)}.

Now, we can write

h¯g=h¯gww=uw\overline{h}g=\overline{h}\frac{g}{w}w=uw

with

u:=h¯gw.u:=\overline{h}\frac{g}{w}.

It is easily checked that uu satisfies |u(x)|=1|u(x)|=1 for m-almost every xx for which |g(x)|>0|g(x)|>0. Then

Th¯f(z)=𝒞h¯g(z)=P+uw(z),z𝔻T_{\overline{h}}f(z)=\mathcal{C}_{\overline{h}g}(z)=P_{+}uw(z),\quad z\in\mathbb{D}

and \threfResVanishingLemma can be applied to conclude that uwuw vanishes on res(w)\text{res}(w). Since uu is unimodular, it follows that in fact ww vanishes on res(w)\text{res}(w), and consequently the set

{x𝕋:w(x)>0}={x𝕋:|g(x)|>0}\{x\in\mathbb{T}:w(x)>0\}=\{x\in\mathbb{T}:|g(x)|>0\}

coincides with core(w)\text{core}(w), up to a set of m-measure zero. For any interval II contained in core(w)\text{core}(w) it follows from the pointwise inequality |g|w|g|\geq w and the definition of core(w)\text{core}(w) that

Ilog|g|dmIlogwdm>.\int_{I}\log|g|\,dm\geq\int_{I}\log w\,d\textit{m}>-\infty.

Thus gg has structure as claimed in the statement of \threfUncertThmRSD, and the proof is complete. ∎

References

  • [1] J. Agler and J. E. McCarthy. Pick Interpolation and Hilbert Function Spaces, volume 44 of Graduate Studies in Mathematics. American Mathematical Society, 2002.
  • [2] A. Aleman and B. Malman. Density of disk algebra functions in de Branges–Rovnyak spaces. C. R. Math. Acad. Sci. Paris, 355(8):871–875, 2017.
  • [3] A. Beurling. A critical topology in harmonic analysis on semigroups. Acta Mathematica, 112(1):215–228, 1964.
  • [4] A. Beurling. Analytic continuation across a linear boundary. Acta Mathematica, 128(1):153–182, 1972.
  • [5] J. Cima, A. Matheson, and W. Ross. The Cauchy transform, volume 125 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2006.
  • [6] J. B. Conway. The theory of subnormal operators. American Mathematical Soc., 1991.
  • [7] B. M. Davis and J. E. McCarthy. Multipliers of de Branges spaces. Michigan Math. J, 38(2):225–240, 1991.
  • [8] L. De Branges. Hilbert spaces of entire functions. Prentice-Hall, 1968.
  • [9] P. Duren. Theory of Hp Spaces. Academic press, 1970.
  • [10] O. El-Fallah, K. Kellay, and K. Seip. Cyclicity of singular inner functions from the corona theorem. Journal of the Institute of Mathematics of Jussieu, 11(4):815–824, 2012.
  • [11] E. Fricain and J. Mashreghi. The theory of (b)\mathcal{H}(b) spaces. Vol. 1, volume 20 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2016.
  • [12] E. Fricain and J. Mashreghi. The theory of (b)\mathcal{H}(b) spaces. Vol. 2, volume 21 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2016.
  • [13] J. Garnett. Bounded analytic functions, volume 236. Springer Science & Business Media, 2007.
  • [14] J. Garnett and D. Marshall. Harmonic measure, volume 2. Cambridge University Press, 2005.
  • [15] V. P. Havin and B. Jöricke. The uncertainty principle in harmonic analysis, volume 72 of Encyclopaedia Math. Sci. Springer, Berlin, 1995.
  • [16] S. V. Khrushchev. The problem of simultaneous approximation and of removal of the singularities of Cauchy type integrals. Trudy Matematicheskogo Instituta imeni VA Steklova, 130:124–195, 1978.
  • [17] P. Koosis. Introduction to HpH_{p} spaces, volume 115 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, second edition, 1998.
  • [18] B. Korenblum. An extension of the Nevanlinna theory. Acta Mathematica, 135:187–219, 1975.
  • [19] B. Korenblum. A Beurling-type theorem. Acta Mathematica, 138(1):265–293, 1977.
  • [20] T. L. Kriete and B. D. MacCluer. Mean-square approximation by polynomials on the unit disk. Transactions of the American Mathematical Society, 322(1):1–34, 1990.
  • [21] A. Limani and B. Malman. Constructions of some families of smooth Cauchy transforms. Canadian Journal of Mathematics, 2023.
  • [22] A. Limani and B. Malman. On model spaces and density of functions smooth on the boundary. Revista Matemática Iberoamericana, 39(3):1059–1071, 2023.
  • [23] A. Limani and B. Malman. On the problem of smooth approximations in (b)\mathcal{H}(b) and connections to subnormal operators. Journal of Functional Analysis, 284(5):109803, 2023.
  • [24] B. Malman. Revisiting mean-square approximation by polynomials in the unit disk. arXiv preprint arXiv:2304.01400, 2023.
  • [25] B. Malman. Thomson decompositions of measures in the disk. Transactions of the American Mathematical Society, 2023.
  • [26] T. Ransford. Potential theory in the complex plane. Number 28 in London Mathematical Society Student Texts. Cambridge university press, 1995.
  • [27] J. Roberts. Cyclic inner functions in the Bergman spaces and weak outer functions in HpH^{p}, 0<p<10<p<1. Illinois Journal of Mathematics, 29(1):25–38, 1985.
  • [28] D. Sarason. Sub-Hardy Hilbert spaces in the unit disk, volume 10 of University of Arkansas Lecture Notes in the Mathematical Sciences. John Wiley & Sons, Inc., New York, 1994.
  • [29] B. Sz.-Nagy, C. Foias, H. Bercovici, and L Kérchy. Harmonic analysis of operators on Hilbert space. Universitext. Springer, New York, second edition, 2010.
  • [30] A. Volberg and B. Jöricke. Summability of the logarithm of an almost analytic function and a generalization of the Levinson-Cartwright theorem. Mathematics of the USSR-Sbornik, 58(2):337, 1987.