Shift operators, Cauchy integrals and approximations
Abstract.
This article consists of two connected parts. In the first part, we study the shift invariant subspaces in certain -spaces, which are the closures of analytic polynomials in the Lebesgue spaces defined by a class of measures living on the closed unit disk . The measures which occur in our study have a part on the open disk 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 . 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 on the unit circle has a Cauchy transform with Taylor coefficients of order for some , then the set is essentially open and is locally integrable on . We establish also a simple characterization of analytic functions with the property that the de Branges-Rovnyak space 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) |
where and are the area and arc-length measures on, respectively, the unit disk and its boundary circle . The radial weight living on is defined in terms of a continuous, increasing and positive function , and the weight living on is a general Borel measurable non-negative integrable function. Given such a measure, we may construct first the Lebesgue space of (equivalence classes of) Borel measurable functions living on the carrier of , and next consider its subspace , by which we denote the smallest closed subspace of which contains the set of analytic polynomials. The space will be the setting for the first part of our study.
The shift operator , which takes a function to , is a subnormal operator, in the sense that it is the restriction of a normal operator, namely , to an invariant subspace. From the point of view of an operator theorist, the significance of the pair lies in the fact the study of subnormal operators can essentially be reduced to the study of the operator for some measure 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 is, like , a space of Borel measurable functions on the closed disk . In certain cases it is even a space of analytic functions on . In such a case, each element , a priori interpreted as a function on , has a unique restriction to the disk . The restriction 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 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 for some , and being a characteristic function of a set 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 when and is a general subset of , and showcases further applications to the theory of the Cauchy integral operator and de Branges-Rovnyak spaces.
1.2. Irreducible -spaces
Recently, the author found in [24] an exact condition for irreducibility of in the case when decays at least exponentially as , thus confirming a conjecture by Kriete and MacCluer from [20]. Roughly speaking, if is smaller than the weight for some , or more precisely if
(ExpDec) |
but large enough to satisfy
(LogLogInt) |
for some , then the space is irreducible if and only if the carrier set of the measure on can be covered by intervals satisfying the condition
(1.2) |
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 be a non-negative integrable function on . We define to be the union of all open intervals for which (1.2) holds. In other words,
(1.3) |
The set is open, and it does not depend on the particular representative of in the space of equivalence classes of functions which are Lebesgue integrable on with respect to .
Definition 1.2.
(Carrier sets) Let be a non-negative Borel measure on . A Borel subset of is a carrier for if
If is a Borel measurable function on , then we say that a set is a carrier for if it is a carrier for the Borel measure .
Carriers are obviously not unique. The set
(1.4) |
is a carrier for . If is only defined up to a set of -measure zero, then we may take as a carrier for any set differing from (1.4) by a set of -measure zero. Since , it is obvious from (1.2) that is essentially contained in any carrier of .
Irreducibility of -spaces of the form (1.1) with 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.
IrrDef For a space defined by a measure of the form (1.1), with satisfying (ExpDec) and (LogLogInt), the following three conditions are equivalent:
-
(i)
the space contains no non-trivial characteristic function of a measurable subset of : if is a Borel subset of and is not the zero element, then .
-
(ii)
the space is a space of analytic functions on in which the analytic polynomials are dense,
-
(iii)
the set is a carrier for , or in other words it coincides with (1.4), up to a set of -measure zero.
Definition 1.4.
IrrDef2(Irreducible spaces) A space is irreducible if it satisfies the three equivalent conditions stated in \threfIrrDef.
In particular, the following measures correspond to irreducible :
() |
and
() |
If the is not a carrier of , then the space will contain a full Lebesgue space , members of which live only on . Here 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 in (). Then (ExpDec) is violated, and condition in \threfIrrDef implies , 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 in (), which means that (LogInt) is violated. This case is less interesting: Volberg’s theorem in [30] implies that is then either a close cousin of the Hardy space (this happens when ) or it is not a space of analytic functions at all (if ). 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 -invariant subspaces of generated by functions in , the algebra of bounded analytic functions in . This question readily reduces to the study of invariant subspaces generated by singular inner functions
(1.5) |
where is a finite positive singular Borel measure on . For , we will denote by the smallest -invariant subspace containing . It is well-known that any singular inner function generates a non-trivial invariant subspace in the classical Hardy space of square-summable Taylor series, and it is almost as well-known that in order for to generate a non-trivial invariant subspace in the standard weighted Bergman spaces (which are -spaces of the kind (1.1) themselves, with for some , and ) we must have for some Beurling-Carleson set (see [18], [19], [27]).
Our first main result characterizes the cyclic singular inner functions in the considered class of -spaces. By cyclicity we mean that . It is not hard to see that the minimal considered rate of decay (ExpDec) of the part of living on makes every non-vanishing bounded function be cyclic in in the case that . Thus only properties of can stop from being cyclic.
Theorem A.
CyclicityMainTheorem Let be an irreducible space defined by a measure of the form (1.1). The following two statements are equivalent.
-
(i)
The singular inner function is cyclic in .
-
(ii)
The measure assigns no mass to the core of the weight :
Note that is open, and hence Borel measurable, so makes perfect sense.
Example 1.5.
PointMassExample Let be a point mass at , and
(1.6) |
Then it is easy to check that
Consequently, the singular inner function
is cyclic, in the considered class of constructed from appearing in (1.6), if and only if .
Having settled the cyclicity question, we turn our attention to the invariant subspace generated by a singular inner function corresponding to a measure which places all its mass on the core: . In other words, is a carrier for . A problem which arises in the theory of normalized Cauchy integrals and de Branges-Rovnyak spaces (to be discussed below) is to determine which functions are contained in the intersection , or sometimes in , where is the Smirnov class of the disk (see [13] for precise definitions):
In this context, we have the following result.
Theorem B.
PermanenceMainTheorem Let be a singular inner function corresponding to a measure which satisfies
In an irreducible -space defined by a measure of the form (1.1), the invariant subspace satisfies
In other words, if can be approximated by polynomial multiples of in the norm of , and places all of its mass on , then appears in the inner-outer factorization of . Under the additional assumption that is bounded, a simple argument will show that in fact . In [21] and [23], the feature of 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 -invariant subspaces generated by bounded analytic functions. Indeed, it follows that if is the inner-outer factorization of into a Blaschke product , singular inner function and outer function , then
where is the restriction of the singular measure to the set .
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 be an analytic function in . If the Taylor coefficients decay so fast that for some we have
(RSD) |
then we say that is a function of rapid spectral decay.
Trivial examples of functions satisfying (RSD) are the analytic polynomials, and functions which extend analytically to a larger disk , . In those cases, the limit in (RSD) is zero even when the term in (RSD) is replaced by for . Conversely, if has an analytic extension to a disk around the origin of radius larger than 1, then for some .
Let us assume that is a finite Borel measure for which the Cauchy integral
(1.7) |
is a function satisfying (RSD). Can we say something about the nature of the measure ? The Cauchy integral has a representation of the form
where is the sequence of Fourier coefficients of indexed by non-negative integers. The rest of the coefficients are annihilated under , and the condition (RSD) gives us no information about for . However, the following statement is a consequence of the irreducibility of spaces corresponding to measures of the form ().
Theorem C.
UncertThmRSD Let be a finite Borel measure on , and assume that the Cauchy integral , given by (1.7), satisfies (RSD). Then the measure is absolutely continuous with respect to the Lebesgue measure :
and there exists an open set which differs from
(1.8) |
only by a set of m-measure zero, with the property that to each there corresponds an interval containing for which we have
The function is, in general, not integrable on the entire open set appearing in \threfUncertThmRSD.
In a way, \threfUncertThmRSD is similar to the classical theorem of brothers Riesz on structure of measures on 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 by for any , and thus consider the weaker unilateral spectral decay condition
then a structural result for as in \threfUncertThmRSD does not hold: will still be absolutely continuous, but examples show that 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 . 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 , where is an inner function, and more generally the de Branges-Rovnyak spaces , where the symbol is any analytic self-map of the unit disk. There are several ways to define the space , the easiest perhaps being by stating that it is the Hilbert space of analytic functions on with a reproducing kernel of the form
Alternatively, we may realize it as the space of normalized Cauchy integrals of functions , given in the special case by the formula
(1.9) |
Here is the Aleksandrov-Clark measure of , these two objects being related by the formula
(1.10) |
The normalization refers to multiplication of the Cauchy integral in (1.9) by the factor , which ensures that the product lands in . It is well-known that model spaces correspond to the purely singular measures in (1.10). In fact, every positive finite Borel measure on corresponds to a function through the formula (1.10). See [5] for more details.
If is a singular inner function, then will contain no functions which extend analytically across . 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 is singular, then for any non-zero function and for any it holds that
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 of functions in , the algebra of functions analytic in with all derivatives extending continuously to , has been established [22].
The situation is more interesting, and much more difficult to handle, in the general class of -spaces. It was proved long ago by Sarason that the set of analytic polynomials is contained and dense in if and only if is a non-extreme point of the unit ball of , a condition characterized by
(1.11) |
where
In terms of core sets, this result can be stated as follows, and a proof can be found in [28].
Theorem (Sarason).
Let be an analytic function. The following three statements are equivalent.
-
(i)
The analytic polynomials are dense in .
-
(ii)
The function is a non-extreme point of the unit ball of .
-
(iii)
We have the set equality .
Since these conditions are very restrictive, it is tempting to make an effort to capture a larger class of symbols for which contains a dense subset of functions in some nice regularity class which is strictly larger than . The article [23] connects the approximation problem in with the structure of -invariant subspaces of , and [21] refines the method to prove the density of for a large class of symbols . The method from [23] is very general and applies to a wide range of approximation problems in . 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 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 by a condition of the type (RSD) but with replaced by faster increasing sequences. To this end, we define below in \threfAdmissibleSequenceDef the admissible sequences . These sequences are logarithmically convex (at least eventually, for large ) and are decreasing to zero at least as fast as , but satisfy a condition of the form
which prohibits, for instance, their exponentially fast decay.
Example 1.7.
The sequence defined by
(1.12) |
is admissible for every and , but it is not admissible for and any .
Theorem D.
MainTheoremHbExistenceESD Let be an analytic function. The following three statements are equivalent.
-
(i)
The space contains a non-zero function which satisfies (RSD).
-
(ii)
For any admissible sequence , the space contains a non-zero function which satisfies
(1.13) -
(iii)
The function vanishes at some point , or there exists an arc of positive length for which
In , the condition of vanishing of at some is the uninteresting case, sice then contains a rational function with no poles on . For such a function holds trivially.
To reach \threfMainTheoremHbExistenceESD we only really need the characterization of irreducibility of . Proof of the next theorem requires the full strength of the invariant subspace results developed in the first part of this article.
Theorem E.
MainTheoremHbDensityESD Let be an analytic function, and be the inner-outer factorization of . The following three statements are equivalent.
-
(i)
The set of functions in which satisfy (RSD) is dense in .
-
(ii)
For any admissible sequence , the set of functions in which satisfy
(1.14) is dense in .
-
(iii)
The set is a carrier for and for the singular measure .
Example 1.8.
For instance, by applying our theorem to the admissible sequence (1.12) for any , we get that the density in of functions satisfying
simultaneously for any and any , is equivalent to condition in \threfMainTheoremHbDensityESD. Roughly speaking, functions satisfying such decay a condition just barely fail to have an analytic continuation to a disk larger than .
Example 1.9.
HbEsetExample Generalizing the setting of \threfPointMassExample, we may replace a point by a general closed subset of , and define the outer function by specifying its modulus , , to satisfy the equation
for , where is the Euclidean distance from the point to the closed set , and for . We can easily check that
If is a Blaschke product and is a singuler inner function, then functions of rapid spectral decay will be dense in the space , with , if and only if .
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 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 is a logarithmically convex sequence which is not admissible according to \threfAdmissibleSequenceDef, because we have
(1.15) |
For instance, could be defined by (1.12) for . If Volberg or Kriete and MacCluer were interested in approximations in -spaces, they would have proved the following theorem by a use of their techniques in [20] and [30].
Theorem (Volberg, Kriete-MacCluer).
Let be a logarithmically convex sequence satisfying the property (1.15). The following two statements are equivalent.
-
(i)
The space contains a non-zero function which satisfies
-
(ii)
The function vanishes at some point , or 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 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 satisfies (RSD) if and only if the multiplication operator acts boundedly on for all non-extreme symbols . In particular, this means that every space defined by a non-extreme symbol contains all functions satisfying (RSD). Our \threfMainTheoremHbExistenceESD then establishes a converse statement: a characterization of for which 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 is the characteristic function of a set contained in , and there exists a function living only on such that in (1.7) has some regularity properties, then what can be said about ? Khrushchev used the phrase removal of singularities of Cauchy integrals in the context of his study of nowhere dense which support a function with a smooth Cauchy integral . Thus ”removing” the singularities of the irregular set . His solution is given in terms of Beurling-Carleson sets. The weighted version of the problem replaces by a general weight .
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 forces the normalization. Indeed, given a positive finite Borel measure on , we may ask if the space contains a dense subset of functions for which the normalized Cauchy integral in (1.9) (with replaced by ) satisfies (RSD). The condition, in terms of the associated function given by (1.10), is given in of \threfMainTheoremHbDensityESD. In this context, it would be of interest to characterize intrinsically the measures which correspond to satisfying condition of \threfMainTheoremHbDensityESD.
Question 1.
Let be an analytic functions which satisfies the condition in \threfMainTheoremHbDensityESD. Can we describe the structure of the corresponding Aleksandrov-Clark measure of 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 plays its part in the proof of \threfAdmissibleSequenceLemma. In relation to that, we would like to answer the following question.
Question 2.
If , , is an increasing, positive and continuous function which satisfies
(1.16) |
then under what additional conditons on may we replace in (1.16) by its least concave majorant?
Any interesting condition on 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 as a sharp condition
Consider the condition
(LogInt) |
for some . The condition (ExpDec) implies that our considered functions 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 to satisfy (ExpDec) is replaced by the requirement for not to satisfy (LogInt). This condition is, in turn, equivalent to the statement that
where is defined in (5.3) and is the sequence of moments of the function . 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 with , and so in contrast to the situation dealt with in the present article.
For instance, a sharp version of the irreducibility of with of the form (1.1) would follow if we could prove the following statement.
Conjecture 1.
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 is cyclic in the space of analytic functions if and only if .
In relation to \threfUncertThmRSD, we expect the following improvement.
Conjecture 3.
The conclusion of \threfUncertThmRSD can be reached if
merely satisfies
for some (say, logarithmically convex) sequence satisfying
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 on we will denote by and its restriction to and , respectively. In some contexts we will also use the same notations and to emphasize that the considered measure lives only on or . The area measure will always be normalized by the condition , and a similar convention will be used also for the arc-length measure on the circle: . We let .
The symbol always denotes the usual -norm corresponding to the finite positive Borel measure . For a set , we sometimes use the shorter notation to denote the space of functions on which vanish outside of and are square-integrable with respect to the Lebesgue measure m. The notation denotes different kinds of duality pairings between spaces. By we will denote the standard inner product in .
The operator is the orthogonal projection onto the Hardy space . For a bounded analytic function , the notation stands for the co-analytic Toeplitz operator with symbol , this operator being defined by the formula .
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 . 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 . 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 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 denote the harmonic measure of a segment of the boundary , based at the point . Then
is a positive harmonic function in which extends continuously to the boundary except at the endpoints of . It attains the boundary value at the relative interior of , and boundary value on . Let denote the Borel -algebra on . For each fixed , the mapping
defines a positive Borel probability measure on . 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
denote the upper half-plane of . The main efforts of this section will go into estimation of the harmonic measure on a wizard hat domain . The domain is constructed from an interval and a profile function , , which by our definition is increasing, positive and continuous, smooth (say, continuously differentiable) for , and which satisfies . Given a profile function and an interval , we define the wizard hat to be the simply connected Jordan domain
(2.1) |
The boundary is a piecewise smooth curve, with three smooth parts divided by three cusps. An example of a domain , constructed from a profile function for some , 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 .
Definition 2.1.
(Majorants) \thlabelRegMajorantDef Let be some positive number. A positive function will be called a majorant if it satisfies the following two properties:
-
(i)
is a decreasing function of , and ,
-
(ii)
.
The properties of appearing in \threfRegMajorantDef are related to growth estimates on functions in the investigated class of -spaces. See \threfPointEvaluationBoundRegMajorant below.
Proposition 2.2.
WizardHatMainProposition Let be a finite interval and be a majorant in the sense of \threfRegMajorantDef. There exists a profile function , a wizard hat , and a positive harmonic function on which extends continuously to the boundary except at the two cusps of on , satisfies for in the interior of , and for .
In order to prove \threfWizardHatMainProposition, we will need to estimate the harmonic measure of the following piece of the boundary of :
(2.2) |
See Figure 1, where is marked. A result of Beurling and Ahlfors (see [14, Theorem 6.1 of Chapter IV]) can be applied to the union of , and the reflected domain to obtain a good estimate for the harmonic measure of .
[scale=1]WizardHat
Proposition 2.3.
(Beurling-Ahlfors estimate) Let be a positive continuous function defined on an interval , and let be the domain
If and is the left vertical part of the boundary of , then
In Figure 1, the symmetrized domain is bounded by the top part of the boundary of and the dotted reflection below the line . Let be the domain obtained by cutting along the cross-section and keeping the right part of the two resulting pieces. Define similarly (so that is the intersection of and ). The Beurling-Ahlfors estimate immediately implies that
where is as in Figure 1. By a comparison of the values on of the two harmonic functions and , and the maximum principle for harmonic functions, we get the inequality
(2.3) |
In particular, this holds at . We have obtained the following harmonic measure estimation.
Proposition 2.4.
Given a majorant as in \threfRegMajorantDef, we will now show how to construct a profile function and harmonic function which satisfies the properties stated in \threfWizardHatMainProposition. Without loss of generality, we may assume that . For some large integer , let
(2.4) |
We define also the sequence
(2.5) |
This sequence is positive if the integer in (2.4) is chosen large enough. Next, we make the following simple observation.
Lemma 2.5.
Proof.
Since is a majorant, by part of \threfRegMajorantDef we have
where we used that . Now, by property in \threfRegMajorantDef we have
and so our claim follows. ∎
We set to some value which ensures that
(2.6) |
or in other words, the sum is less than one quarter of the length of the interval . Further, we let be a sequence of positive numbers starting with
which tends monotonically to . We shall soon define by specifying the sequence of differences , where
The differences are positive numbers, and will be recursively defined in terms of those differences by the relations
and so on. In order for so defined sequence to converge to zero it is necessary and sufficient that
(2.7) |
a requirement which we will later ensure. Given any as above, a profile function may be readily constructed which satisfies
(2.8) |
Indeed, since the sequence is assumed to be monotonically decreasing to zero, the function can be chosen to be smooth, increasing and positive for , and satisfy . A proper choice of will produce a wizard hat with our desired properties. Assume that has been given, let be the corresponding wizard hat, and be the harmonic measure at some point which lies on the symmetry line of . Let be defined on by
(2.9) |
We will ensure that . Since is decreasing, the definition of shows that for any the values of the function on the arc are dominated by its value at the point which lies closest to the real line , i.e., at the point . In other words, we have
(2.10) |
Moreover, from positivity and monotonicity of , and from \threfBtHarmEst, we deduce the estimate
(2.11) |
which holds for .
(2.12) |
In the last step we used (2.4), (2.5) and (2.8). We may now specify the values of by setting the values of the differences:
(2.13) |
for an appropriate constant which ensures the necessary summation condition (2.7). This can be done, since
by (2.6). We obtain from (2) that
Consequently, with this definition of and a corresponding profile function , we have that .
We may now complete the proof of \threfWizardHatMainProposition. Let and be chosen as above, and be a conformal mapping which maps to . Since is a Jordan domain, extends to a homeomorphism between and . If is defined by , then a change of variables shows that
Since is continuous on all of except at the two cusps of on , the function is continuous on except at the two points which map under to the cusps. We verified above that , so we may extend to by means of its Poisson integral. This extension is continuous in except at the two points corresponding to the cusps. If we define in by , then 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 . This is done, again, only for convenience. An elementary conformal mapping argument will carry the result over to the intended domain .
In this section, the Lebesgue measure (length measure) on will be denoted by , and the -measure of a set will be denoted by , similar to lengths of sets on the circle (this should not cause confusion). The algebra of bounded analytic functions in will be denoted by . In the proofs below we shall use some basic facts regarding , 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 admits an inner-outer factorization into
(3.1) |
Here is some unimodular constant, , is a Blaschke product given by
where is a non-negative integer, is a singular inner function given by
where is a singular positive Borel measure on , and is the outer function given by
The measures and appearing in the integrals above are both finite. It follows from this factorization that we have
(3.2) | ||||
The last two terms in (3.2) represent the Poisson integrals and of the measure and of the function , respectively.
Lemma 3.1.
IntervalWeakStarConvLemma Let be a finite open interval of . With notation as above, as , the restrictions to of the measures converge weak-star to the restriction to of the measure .
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 is any smooth function which is supported on a compact subset of , then by the symmetry of the Poisson kernel, we have
(3.3) |
where is the Poisson integral of . Since is uniformly continuous on , uniformly in as . Moreover, the compact support of implies that as , uniformly in, say, . Thus expression (3.3) and the finiteness of the measure implies now that
and we have shown that converges weak-star to on the interval . By the same argument converges weak-star to on .
We consider now the measures . Jensen’s inequality for the upper-half plane (see, for instance, [15, p. 35]) implies that
Thus by letting tend to , we obtain
The last inequality is trivial, since is negative in . For a finite Blaschke product , the limit between the inequalities above is certainly equal to . Thus we obtain
Now let tend to through a sequence of finite partial products of to obtain
This says that the restriction to of converges to 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 is a sequence of non-negative absolutely continuous Borel measures on and is an interval, then we will say that the sequence is uniformly absolutely continuous on if to each there corresponds a , independent of , such that for Borel sets we have
Recall that the notion of a majorant has been introduced in \threfRegMajorantDef.
Lemma 3.3.
permanenceLemma Let be a finite interval of the real line , a singular inner function in defined by a singular measure supported in the interior of , and a sequence of functions in such that
where is a non-zero function. Assume that
-
(i)
there exists a majorant for which we have
for some constant independent of , and where is some rectangle in with base :
-
(ii)
the sequence of positive Borel measures is uniformly absolutely continuous on an interval larger than .
Then .
Proof.
The assumption implies that
So, denoting by the characteristic function of the interval and by passing to a subsequence, we can assume that the measures converge weak-star to a non-negative measure supported on . The measure must be absolutely continuous with respect to : any set of -measure zero can be covered by an open set of total length arbitrarily small, and then we can use to conclude that
for any . Consequently for some non-negative . We denote by the harmonic function in which is the Poisson extension of the measure to :
Let also denote the Poisson extension of the measure :
The assumption implies that
for some positive constant . By \threfWizardHatMainProposition, there exists a wizard hat domain and a corresponding positive harmonic function defined on which satisfies
By the assumption that the singular measure defining is supported in the interior of , it follows that is analytic and non-zero in a neighbourhood of , and so is bounded on . Therefore, by possibly replacing by a positive scalar multiple of itself, in fact we have that
(3.4) |
For the bottom side of the wizard hat, we have the non-tangential boundary value inequality
(3.5) |
for -almost every . 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) |
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
in all of , and so by letting we obtain, from (3.6) and the earlier mentioned weak-star convergence of measures (which guarantees that for ), that
(3.7) |
Let be some interval containing the support of , and which is strictly contained in . By \threfIntervalWeakStarConvLemma, as , the restrictions to of the real-valued measures converge weak-star to the restriction to of the measure . Similar claims hold for the Poisson integrals
(3.8) |
and . For sufficiently small we have that . Thus from the weak-star convergence of measures discussed above, the inequality (3.7), and the fact that on , we obtain the real-valued measure inequality
This measure inequality is to be interpreted in the following way:
is a non-negative measure on . The -singular part of this measure is , which is thus non-negative on . Since is supported inside , in fact is non-negative in all of . Now subtracting (3.8) from (3.2), using the inequality and the non-negativity of and of the Poisson kernel, we get for that
By exponentiating, we finally obtain
∎
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.
TdomLemma Let the wizard hat be as in the proof of \threfpermanenceLemma, be a bounded analytic function in and be a positive harmonic function in . Assume that both and extend continuously to and also that both have non-tangential limits almost everywhere on . If we have that for , and moreover that the non-tangential limits of on are -almost everywhere dominated by the non-tangential limits of on , then for all .
Proof.
Let be a conformal mapping. The local smoothness of the boundary of and basic conformal mapping theory ensure that is conformal at almost every point of (see [14, Chapter V.5]). This implies that the functions and , which are defined in , have non-tangential limits almost everywhere on . Let be a harmonic conjugate of in and consider the function , . By positivity of , the function is bounded in , and our assumptions leads to the conclusion that the non-tangential boundary values on of the bounded function are not larger than 1 in modulus. Thus by basic function theory in , we obtain the inequality for all . This easily translates into for . ∎
We will need \threfpermanenceLemma in the disk . Here is the precise statement which we will use. The uniform absolute continuity of sequences of Borel measures on arcs of the circle is defined analogously to how it was defined in \threfuniAbsContDef for intervals on the line.
Corollary 3.5.
PermanenceLemmaDisk Let be an arc properly contained in the circle , be a singular inner function in defined by a singular measure supported in the interior of , and be a sequence of functions in such that
where . Assume that
-
(i)
there exists a majorant for which we have
for some positive constant independent of ,
-
(ii)
the sequence of positive Borel measures is uniformly absolutely continuous on an arc larger than .
Then .
It is easy to see that \threfpermanenceLemma implies \threfPermanenceLemmaDisk. Indeed, if is a conformal map for which is a finite segment on , then the distortion of lengths and distances by the map is bounded above and below near and , since is a bi-Lipschitz bijection between some open sets containing and . For instance, the growth condition in \threfPermanenceLemmaDisk for the sequence is easily translated into a corresponding condition in \threfpermanenceLemma for the sequence , where and , by replacing with a new majorant of the form for some . Moreover, the mapping 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
PermanenceMainTheorem follows almost immediately from \threfPermanenceLemmaDisk, we just need to verify that a bounded sequence in the corresponding -space satisfies properties and in \threfPermanenceLemmaDisk. This is done in the next two lemmas.
Lemma 3.6.
Proof.
Fix , and let denote the ball around of radius . By subharmonicity of the function and the Cauchy-Schwarz inequality, we have
Since is assumed to be an increasing function, we may estimate the integral inside the square root by
Putting this into the previous estimate, we obtain
Now set
By the above estimate, the norm of the evaluation functional is bounded by
Moreover, is a decreasing function, and by virtue of satisfying (LogLogInt), also certainly satisfies
if is some small number. Thus is a majorant in the sense of \threfRegMajorantDef. ∎
Lemma 3.7.
UniAbsContLemma Assume that the weight satisfies
for some arc . If are positive functions such that
for some constant and some , then the sequence is uniformly absolutely continuous on .
Proof.
Note that
where it follows from the assumption that are positive functions which form a bounded subset of (say) , and . Clearly, if is a Borel subset of , then by Cauchy-Schwarz inequality we obtain
so that the family is uniformly absolutely continuous on . Then the above inequalities imply that is a uniformly absolutely continuous sequence on . ∎
Proof of \threfPermanenceMainTheorem.
Let . Since , there exists a sequence of polynomials such that converges to in the norm of . Multiplying by a suitable bounded outer function we can ensure that , and that converges to (see the discussion following \threfProdOfCyclicLemma below). Let be an increasing sequence of compact sets which are finite unions of intervals and such that . By \threfPermanenceLemmaDisk, \threfPointEvaluationBoundRegMajorant and \threfUniAbsContLemma, whenever is the restriction of to the compact subset , we have that with the bound . The assumption that means that the restrictions converge weak-star to the measure . Thus
In particular, since is outer, it follows that divides the inner factor of . Thus . ∎
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 of the singular inner function will follow from the existence of a sequence of non-negative bounded functions for which the measures converge weak-star to . In our context the functions will have to satisfy a severe restriction on their size, namely
(4.1) |
In case , the right-hand side is to be interpreted as (i.e., no size restriction on at the point ). Essentially, the obstacle (4.1) prohibits the existence of an approximating sequence if some part of the mass of is located in ”wrong” places on . However, if is carried outside of the core set of , then such a sequence exists. This is the content of the next lemma.
Lemma 4.1.
WeakStarSeqObstacleLemma Let be a positive singular Borel measure on which satisfies
Then there exists a sequence of non-negative bounded functions satisfying the following three properties:
-
(i)
the non-negative measures converge weak-star to ,
-
(ii)
,
-
(iii)
the functions obey the bound (4.1).
Proof.
Let us first suppose that assigns no mass to any singletons, so that whenever . For any positive integer , we let be the family of disjoint open dyadic intervals, each of length , such that their union covers the circle up to finitely many points, and such that the system possesses the usual dyadic nesting property: each is contained in a unique . Fixing an integer , we will specify how to define on each of the intervals , , in such a way that the above three properties hold.
If , then we simply set on . Conversely, if , then since , it must be that . It follows that is non-empty. Pick some point . For any open interval which contains in its interior we have (else would have been a member of ). Pick such an interval which is contained within . If there exists a subset satisfying on which is identically zero, then we may set
where is the characteristic function of . In case that such a set does not exist, then almost everywhere on , and we must have
so that
for some small . By absolute continuity of the finite measure
there must then exist a set for which we have precisely
We pick such and define
Note that on . For definiteness, we can set to be equal to on the finitely many points outside of . One way or the other, we have defined as a bounded function, and we have
By summing over all the intervals , we see that property in the statement of the lemma is satisfied (since assigns no mass to the finitely many points outside the union of the open intervals ). Property is satisfied by the construction. Property also holds. Indeed, if is the characteristic function of one of the dyadic intervals from some stage of our construction, then the nesting property of the dyadic system and the additivity of ensure that
The above equalities hold also for functions 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 , and since we have the uniform variation bound in , we conclude that the sequence { converges weak-star to . The proof is complete in the case that assigns no mass to singletons.
In the contrary case we have that
is a countable linear combination of unit masses at the sequence of points in . Our assumption implies that for all . Thus each is the midpoint of an interval which can be chosen to have arbitrarily small length and for which we have . We can then proceed in an analogous way to the above, and produce at each stage of the construction a disjoint finite sequence of intervals each covering a different point for . We then define a positive function which carries appropriate amount of mass on each of the intervals 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 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.
ProdOfCyclicLemma Assume that is a Banach space of analytic functions in which contains and with the property that for all functions in the operator is bounded on . Then the product of two cyclic bounded functions is cyclic.
By cyclicity of we mean, of course, that there exists a sequence of analytic polynomials such that converges to in the norm of the space.
Proof.
If and are two cyclic bounded functions, then for any polynomials we have the inequality
where denotes the operator norm of the multiplication operator , and denotes the norm in . We use cyclicity of to choose the polynomial to make the first term on the right arbitrarily small, and next we use cyclicity of to choose to make the second term on the right arbitrarily small. It follows that the product of two bounded cyclic functions is a cyclic function. ∎
ProdOfCyclicLemma applies to any irreducible space of the form considered here, since indeed the multiplication by any function in 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 , , of the bounded function . In particular, whenever the latter is irreducible. For future reference, note that as a subspace of (with as in (1.1)), each function is defined also on , the part of living on the circle . It is not hard to see that the values of with respect to coincide with the non-tangential boundary function of on . If is bounded, then the same conclusions hold also for any .
Proof of \threfCyclicityMainTheorem.
Note first that in \threfCyclicityMainTheorem, since the condition implies that a factor in satisfies the permanence property exhibited in \threfPermanenceMainTheorem, and so cannot by cyclic. Thus it suffices for us to show the implication . The norms induced by measures satisfying (ExpDec) are largest if the measure has the form in () defined in Section 1, with . If is cyclic in defined by any measure this form, then it is cyclic in any -space considered in this article. Thus it suffices to prove the theorem in the case of being of the form () with , and any .
Let us then assume that . The formula (1.5) shows that
for any positive integer . Then by replacing by for sufficiently large, and by \threfProdOfCyclicLemma, we may assume that . Let be a sequence of positive bounded functions given by \threfWeakStarSeqObstacleLemma for which the measures converge weak-star to , which satisfy , and for which the bound (4.1) holds. Construct the outer functions
where
is the usual Herglotz integral of . Then, since , we obtain
and from property in \threfWeakStarSeqObstacleLemma and basic properties of Herglotz integrals, we have the non-tangential boundary value estimate
for almost every with respect to . It follows from these inequalities and the definition of the norm in that the family forms a bounded subset of the Hilbert space . Moreover, by the weak-star convergence of to we have that
But this means that is a member of , since we can identify it as a weak cluster point of some subsequence of . Thus there must also exist a sequence of polynomials tending to in the norm of . Consequently, since the multiplication operator is a bounded on our space, we have that in the norm of . That is, 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 -theory to Cauchy integrals, model spaces and the de Branges-Rovnyak spaces . In order to do so, we will need to analyze the moments of the functions appearing in (1.1). This entire section is concerned with this analysis.
5.1. Admissible sequences and their properties
If is a function satisfying (ExpDec) and (LogLogInt), then the sequence of moments of , 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 with
will be called admissible if it satisfies the following three conditions:
-
(i)
the sequence is eventually convex, in the sense that
for all sufficiently large ,
-
(ii)
there exists such that
for all sufficiently large ,
-
(iii)
the summability condition
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.
PowerAdmSeq If is an admissible sequence, then so is
for any .
The proposition follows immediately from \threfAdmissibleSequenceDef
5.2. Legendre envelopes
Roughly speaking, admissible sequences are in a correspondence with moments of functions 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 be a positive and continuous function defined for , which is decreasing and satisfies
In our application, will be of the form (for small ). The lower Legendre envelope is defined as
(5.1) |
Being an infimum of concave (actually affine) and increasing functions, is itself concave and increasing, and it is easy to see that
Remark 5.3.
remarkLowerLegEnvelope Assume that we modify the function above for larger than , so that we end up with a different function which satisfies for , but the values of the two functions might differ for . Then it is not hard to see from the definition in (5.1) that for all sufficiently large . Indeed, if , then we have by positivity of that
the second inequality holding if is sufficiently large. For such , the candidate is always better than any candidate 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.
LogIntLogSumEquivalenceProp Let be a positive, continuous and decreasing function of which satisfies The following two statements are equivalent:
-
(i)
there exists a such that
-
(ii)
we have
We refer the reader to [4] for a proof of \threfLogIntLogSumEquivalenceProp.
Let be a positive concave function of which is increasing and satisfies
We will consider its upper Legendre envelope defined as
(5.2) |
Then it is easy to see that is a convex and decreasing function, and
We have the following inversion formula, which is well-known (see [15, p. 224-225]).
Lemma 5.5.
LegendreInversionFormula Let be a positive concave function of which is increasing and satisfies . Then
5.3. A characterization of admissible sequences
We will be interested in the sequence of moments of the parts of our measures living on :
(5.3) |
We define the moment function of by
(5.4) |
The next lemma gives an estimate on . 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.
MomentGGrowthLogLogInt Let , , be an increasing continuous function satisfying , and put
(5.5) |
Then, for sufficiently large , we have the estimates
where is the moment function of defined in (5.4).
Lemma 5.7.
MomentEstProp For and , let be the sequence of moments given by
(5.6) |
Put
For sufficiently large positive , we have the estimates
(5.7) |
where is comparable to if remains fixed.
Proof.
In the notation of \threfMomentGGrowthLogLogInt, and with
we have
and we need to compute the corresponding Legendre envelope defined in (5.1). Having fixed some number , we use elementary calculus to show that
is attained at the point
from which it follows that
Since
we obtain from \threfMomentGGrowthLogLogInt the inequalities
(5.8) |
which hold for all sufficiently large . Our result follows easily from this. ∎
We can now prove the main result of the section, which connects our considered class of functions with the admissible sequences appearing in \threfAdmissibleSequenceDef.
Proposition 5.8.
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
Thus is a convex sequence. The inequality for some and all sufficiently large follows readily from (ExpDec) and an application of the upper estimate \threfMomentEstProp with (and consequently ). Let be as in \threfMomentGGrowthLogLogInt. By the lower estimate in that lemma, we have
The assumption that satisfies (LogLogInt) implies that , and so from \threfLogIntLogSumEquivalenceProp we deduce that the last sum above is convergent. Consequently, is an admissible sequence.
Conversely, assume that is an admissible sequence. Since the sequence tends to zero, we may without loss of generality assume that . From property in \threfAdmissibleSequenceDef we obtain the inequality
This means that the slopes of the line segments between each consecutive pair of points in the sequence
(5.10) |
are decreasing, which means that if we define the function , , as the piecewise linear interpolant of the data (5.10), then is concave, continuous, positive and increasing, and satisfies
It also satisfies , and property in \threfAdmissibleSequenceDef easily implies that
(5.11) |
Let be the upper Legendre envelope of defined in (5.2), set
(5.12) |
and . Then is a continuous and increasing function. Define as in (5.4). By \threfremarkLowerLegEnvelope, (5.12), inversion formula in \threfLegendreInversionFormula and \threfMomentGGrowthLogLogInt, we have the estimate
for all sufficiently large . Consequently,
if is large, since interpolates the data (5.10). \threfLogIntLogSumEquivalenceProp, \threfLegendreInversionFormula and (5.11) imply that
Thus satisfies (LogLogInt). It remains to check that also satisfies (ExpDec). Note that property in \threfAdmissibleSequenceDef of the admissible sequence implies easily that satisfies a lower bound of the form
for some constant . But then, by (5.2), we have
The last equality can be derived by elementary calculus techniques. Consequently
and so satisfies (ExpDec). The proof is complete. ∎
5.4. Some auxiliary spaces of Taylor series
If is an analytic function and
(5.13) |
then we have the norm equality
(5.14) |
where is the sequence of Taylor coefficients of , and is given by (5.3). The above equality gives us an isometric isomorphism between and a space of Taylor series.
For a decreasing sequence of positive numbers we define to be the Hilbert space of analytic functions in consisting of which satisfy
(5.15) |
In our development, the sequences will be the admissible sequences studied in Section 5. Such sequences have the property that
a condition which ensures that the spaces , and their duals, are genuine spaces of analytic functions on . The dual space is to consist of power series which satisfy
(5.16) |
Since is assumed to be decreasing, the space is contained in the Hardy space . In fact, if is an admissible sequence, then consists of functions satisfying the condition (RSD) of Section 1. The duality between and is realized by the usual Cauchy pairing
(5.17) |
where the sequential definition above makes sense whenever , , and the integral definition holds only in special cases, for instance when . An application of the Cauchy-Schwarz inequality to the limit in (5.17) shows that
We introduce also the space which consists of power series satisfying
(5.18) |
Recall from \threfPowerAdmSeq that the family of admissible sequences introduced in \threfAdmissibleSequenceDef is invariant under taking powers. For this reason, the spaces and which appear in our study are very similar.
Lemma 5.9.
H1starH2starEmbeddingLemma Let be an admissible sequence, and consider the sequences
For , we have the continuous embeddings
Proof.
If , then for any we have that
so clearly . If for some , then we may use that satisfies property of admissible sequences in \threfAdmissibleSequenceDef to obtain
Thus . ∎
The following corollary will be used several times below.
Corollary 5.10.
MhstarContainmentFromRSD If an analytic function in satisfies the condition (RSD), then there exists and a measure
with sequence of moments such that .
Proof.
The condition (RSD) and \threfH1starH2starEmbeddingLemma imply that , where for some positive constant . Now \threfMomentEstProp, with , shows that can be chosen so that for sufficiently large . Then . ∎
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 being decreasing, it is not hard to see that the spaces are invariant under the multiplication operator , and that this operator is a contraction. Then Von Neumann’s inequality ([1, p. 159]) or the Sz.-Nagy Foias -functional calculus ([29, Chapter 3]) shows that in fact every function defines a bounded multiplication operator . The adjoint operator is easily indentified with the usual Toeplitz operator with the co-analytic symbol , i.e., is the orthogonal projection to the Hardy space of the function .
For later reference, we record these observations in a proposition.
Proposition 5.11.
HToeplitzInvariance Let be an admissible sequence.
-
(i)
The space is invariant for the analytic multiplication operators
with symbols .
-
(ii)
The space is invariant for the co-analytic Toeplitz operators
with symbols .
Corollary 5.12.
ToeplitzInvRSD If an analytic function satisfies the condition (RSD), then so does for any .
Proof.
We use \threfMhstarContainmentFromRSD and \threfHToeplitzInvariance to see that the function is contained in a space , where is admissible. \threfH1starH2starEmbeddingLemma shows that , so satisfies (RSD). ∎
6. Existence in 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 -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.
AlphaKthetaBreakpointProp If is a singular inner function, then the model space contains no non-zero function which satisfies (RSD).
Proof.
We will show that any which satisfies (RSD) satisfies also . Since is invariant for the backward shift
and by \threfToeplitzInvRSD the function satisfies (RSD), the same argument will show that for . Thus will follow.
We apply \threfMhstarContainmentFromRSD to and obtain a measure with moment sequence such that . The measure is of the form () (see Section 1) for and . By \threfCyclicityMainTheorem, the singular inner function is trivially cyclic in , since . Thus there exists a sequence of analytic polynomials such that in the norm of . Using the duality pairing (5.17) and the membership of in , the following computation is justified:
The last equality holds due to being a member of . Thus , and the proof is complete by the initial remarks. ∎
6.2. Some -theory
The following description of -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 -spaces. We will employ it in a similar way below. Recall that the symbol denotes the orthogonal projection operator , and denotes the subspace of those which live only on the measurable subset .
Proposition 6.2.
normformula Let be an extreme point of the unit ball of ,
(6.1) |
and be the carrier set of :
Then is a member of if and only if the equation
(6.2) |
has a solution . The solution is unique, and the map defined by
is an isometry. Moreover,
(6.3) |
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.
CauchyTransformsinHb Let and be as in \threfnormformula. For any , the function
is a member of and, in the notation of \threfnormformula, we have
Moreover, if is extreme and is non-zero, then is non-zero.
Proof.
We compute
and so (6.2) holds for the pair . If is extreme, then , and it follows readily that also . A function is conjugate-analytic, and so if . So if is non-zero, and it follows that is non-zero. ∎
Corollary 6.4.
TconjbInvariance The Toeplitz operator acts boundedly on . If and , then
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 and consider the carrier set
We define to be the set
where is the set appearing in \threfCoreDef.
Since might only be defined up to a set of -measure zero, the same is true for the residual of any weight . 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].
6.4. Proof of \threfMainTheoremHbExistenceESD
In \threfMainTheoremHbExistenceESD, it is obvious that . We can thus prove the theorem by showing validity of the implications and .
Let us first show that , and so we assume that is non-zero and that it satisfies (RSD). We may assume that does not vanish at any point in , else certainly holds. Similarly to as it was done in the proof of \threfAlphaKthetaBreakpointProp, we use \threfMhstarContainmentFromRSD to obtain a measure
(6.4) | ||||
and a sequence such that the identity map between and is an isometry, and . By \threfHToeplitzInvariance, the space is invariant under Toeplitz operators with co-analytic symbols, and consequently we also have . By \threfTconjbInvariance and (6.2) we have and
(6.5) |
for some . The kernel of the operator is the model space , where is the inner factor of . Since does not vanish in , it follows that is a purely singular inner function. Every function in satisfies (RSD), so \threfAlphaKthetaBreakpointProp implies that . Consequently , so , and by (6.5). If
is the Taylor expansion of , then a consequence of the membership is that the function
is a member of . The function lives on , the function lives on , and hence defines a function on . The condition means simply that is square-integrable with respect to the part of in (6.4) which lives on . The containment is ensured by the boundedness of and the containment . Thus . The representation (6.5) tells us that the positive Fourier coefficients of and of coincide. Our definitions then imply that the function is orthogonal to the analytic polynomials in . Since , the function is a non-zero element of . The conclusion is that there exists an element (namely ) inside which is orthogonal to and which does not vanish identically on the circle . If there existed no interval on which was integrable, then , and so \threfResSetMainLemma would imply that the entire space is contained in . Clearly that would be a contradiction to being orthogonal to . Thus such an interval exists, and we have proved that .
The implication is easier. Let be an admissible sequence. We must show that contains a function in . If vanishes at some point of , then the implication is trivial. Assume therefore that is integrable on some (say, open) interval which is not all of , and let be the restriction of to the interval . By \threfPowerAdmSeq and \threfAdmissibleSequenceLemma there exists a function which satisfies (ExpDec), (LogLogInt), with corresponding moment sequence
satisfying
for large . If
then the space is irreducible by \threfIrrDef, since coincides with , which is a carrier of . By irreducibility we have that . So there must exist a non-zero element , with being an analytic function on and living on , which is orthogonal to in . We can’t have , for then the Taylor coefficients of would all vanish by the orthogonality to analytic monomials, and consequently would reduce to the zero element. The orthogonality means that
where are the Taylor coefficients of and are the non-negative Fourier coefficients of . For large , we have the estimate
Thus is a member of . Since lives on and , we have by \threfCauchyTransformsinHb that . This function is non-zero since by the choice of . Thus , and we have completed our proof of \threfMainTheoremHbExistenceESD.
7. Density in of functions with rapid spectral decay
The main result of [23] characterizes the density of the functions in which have Taylor series satisfying , for positive . The characterization is in terms of the structure of -invariant subspaces of with of form (1.1) and , . 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 lie outside of the closure of functions satisfying spectral decay properties as in (RSD).
As before, for , and we let
be the inner-outer factorization of , with a Blaschke product, a singular inner function, and an outer function. We denote by the inner factor of .
Lemma 7.1.
ResVanishingLemma Let be non-negative, and assume that for some the function satisfies (RSD). Then vanishes on .
Proof.
We use \threfMhstarContainmentFromRSD to obtain a measure as in () of Section 1, and with the parameters and chosen so that if is the sequence of moments corresponding to , then . Let be a bounded function living on , and be a sequence of analytic polynomials which converges to in the norm of . This is possible by \threfResSetMainLemma. In particular, this convergence implies that in , or in other words, in . Simultaneously, we have that in . Using the duality pairing (5.17), we obtain
Since is an arbitrary bounded function living on , it follows that on . ∎
Proposition 7.2.
Prop1TheoremD
Assume that the set has positive -measure, and let be a non-zero function which vanishes outside of . Then the non-zero function
lies outside of the norm-closure in of functions satisfying (RSD).
Proof.
Seeking a contradiction, assume that is a sequence of functions in which satisfy (RSD) and which converge in the norm of to the given . In the notation of \threfnormformula, we consider and , where according to \threfCauchyTransformsinHb. By \threfTconjbInvariance, converges to in the norm of , and since the embedding of \threfnormformula is an isometry, \threfTconjbInvariance moreover implies that converges to in . In particular, this implies that cannot all simultaneously vanish on , since lives only on that set. But satisfies (RSD) (since does), and by \threfTconjbInvariance. Thus by \threfResVanishingLemma, the functions must vanish on , and consequently must vanish on , since is non-zero -almost everywhere on that set. This is the desired contradiction. ∎
We have now proved that it is necessary for to be a carrier for if functions satisfying (RSD) are to be dense in . In the next proposition, we assume that is a carrier for , and show that if is the singular inner factor of and places some portion of its mass outside of the core of , then again functions satisfying (RSD) are not dense in . And again, we do it by exhibiting explicit functions in which cannot be approximated in this way.
Proposition 7.3.
Prop2TheoremD Assume that is a carrier for and that
where is the singular inner factor of . Decompose the measure as
where is the restriction of to the set , and is the restriction of to the set . Then all functions in the subspace
are orthogonal in to functions satisfying (RSD), being the model space generated by the singular inner function .
Proof.
Take a function , where , and satisfying (RSD). In the notation of \threfnormformula, a computation shows that , where
Let and be as in the proof of \threfResVanishingLemma, with and the sequence being chosen so that . This time the space is irreducible, and by \threfCyclicityMainTheorem the singular inner function is cyclic in . Hence there exists a sequence of analytic polynomials such that converges to the function in the norm of , and in particular in the norm of . Multiplying this sequence by , it follows from \threfHToeplitzInvariance that converges to in . Simultaneously, the -convergence implies that converge to in , and since is unimodular on , in fact we have that converge to in . Let be the corresponding pair for . We can use the above claims to compute
In the last step we used condition (6.2) for the pair . Since the embedding in \threfnormformula is an isometry, it follows that is orthogonal to in . ∎
Proof of \threfMainTheoremHbDensityESD.
We see from \threfProp1TheoremD and \threfProp2TheoremD above that condition in \threfMainTheoremHbDensityESD is necessary in order for to hold. Since implies , it suffices thus to show that implies . 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 we use \threfPowerAdmSeq and \threfAdmissibleSequenceLemma to obtain satisfying (ExpDec), (LogLogInt), with moment sequence satisfying for large . We must show that is dense in . By \threfH1starH2starEmbeddingLemma it will suffice to show that is dense in .
The space constructed from the measure
is irreducible by \threfIrrDef. Let us assume that is orthogonal to . We will show that , which will prove \threfMainTheoremHbDensityESD. Because the mapping in \threfnormformula is an isometry, it follows that is orthogonal to . Note that is a subset of , and under the duality pairing (5.17) between and , we have
(7.1) |
where is the functional on which acts by the formula
This follows readily from \threfnormformula (see, for instance, the argument in [23]). The fact that annihilates and that (7.1) holds implies that is contained in the weak-star closure of the linear manifold . Since the pairing between and is reflexive and is a convex set, basic functional analysis says that, in fact, is contained in the norm-closure of . Thus there exists a sequence with such that
(7.2) |
in the norm of . Multiply the second coordinate by to obtain
(7.3) |
But the inequalities imply that converges to also in the space , and so in fact (7.3) tells us that is a Cauchy sequence in , to which \threfPermanenceMainTheorem applies. The critical conclusion is that in the irreducible . If is the inner factor of , then \threfPermanenceMainTheorem implies that , and by the irreducibility of the sequence on must converge to the boundary function of on . Thus by (7.3), and . By \threfnormformula we get that
(7.4) |
From the above computation we infer that, in terms of boundary values, we have , and consequently has square-integrable boundary values. Since , it follows from the classical Smirnov maximum principle that . Then is an analytic function which projects to under , which implies that , and consequently . ∎
8. Proof of \threfUncertThmRSD
A proof of \threfUncertThmRSD relies on a judicious application of \threfResVanishingLemma.
Proof of \threfUncertThmRSD.
If satisfies (RSD), then the function , , is certainly smooth on and it has an analytic extension to . Since the Cauchy transform of the measure vanishes in , this measure must be absolutely continuous with respect to by the classical theorem of brothers Riesz. Hence is also absolutely continuous. Let be its Radon-Nikodym derivative, so that . Set , which by our assumption is a function satisfying (RSD). Unfortunately, we cannot directly apply \threfResVanishingLemma since we do not necessarily have that . We must take care of this slight inconvenience to prove the theorem.
ToeplitzInvRSD says that satisfies (RSD), where is any co-analytic Toeplitz operator with bounded symbol . Moreover, has the representation
The above formula can be derived by first showing through simple algebraic manipulations that it holds for , then for analytic monomials by iteration, and thus for analytic polynomials by linearity. Finally, fix a uniformly bounded sequence of analytic polynomials which converges to pointwise -almost everywhere on (the polynomials could be taken to be the Cesàro means of the partial sums of the Taylor series of ). For such a sequence we readily see from the dominated convergence theorem that
Since , in particular we have that , and this means that an outer function exists which satisfies the boundary value equation
for m-almost every . Set also
Now, we can write
with
It is easily checked that satisfies for m-almost every for which . Then
and \threfResVanishingLemma can be applied to conclude that vanishes on . Since is unimodular, it follows that in fact vanishes on , and consequently the set
coincides with , up to a set of m-measure zero. For any interval contained in it follows from the pointwise inequality and the definition of that
Thus 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 spaces. Vol. 1, volume 20 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2016.
- [12] E. Fricain and J. Mashreghi. The theory of 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 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 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 , . 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.