The Steklov problem and Remainder Estimates for Krein Systems generated by a Muckenhoupt weight
Abstract.
We show that solutions to Krein systems, the continuous frequency analogue of orthogonal polynomials on the unit circle, generated by an weight satisfying , are uniformly bounded in for sufficiently close to . This provides a positive answer to the Steklov problem for Krein systems. Furthermore, we define a “remainder” which measures the difference between the solution to a Krein system and a polynomial-like approximant, and we estimate these remainders in for satisfying some additional conditions. Such polynomial-like approximants, and hence remainder estimates, seem unique to Krein systems, with no analogue for orthogonal polynomials on the unit circle.
1. Introduction
Let be a measure on with weight satisfying . Then one can define a family of “orthonormal continuous polynomials” with the following properties.
-
(i)
is a “continuous polynomial” in of “degree” , where : for each there exists so that
-
(ii)
is an orthonormal system with respect to , i.e. for all we have equality of the inner-products
(1) Equivalently, the generalized Fourier transform is an isometry from to , where
-
(iii)
is the solution to the Krein differential system (20), whose coefficients uniquely determine . As such is referred to as “the solution to the Krein system.”
For instance, when , then , so is simply the Fourier orthonormal system . See e.g. references [8, 3] for more information on Krein systems.
Notation. When the measure is clear from context, we simply write . And given a weight we write .
First introduced in 1954 by Krein in [8], solutions to Krein systems are the continuous frequency analogue of Orthogonal Polynomials on the Unit Circle (OPUC): given a finite measure on the unit circle with infinite support, the OPUC consist of an orthonormal sequence in , where each is a degree polynomial over with positive leading coefficient. One can obtain this sequence by, e.g., applying the Gram-Schmidt algorithm to the polynomials . See [12] for a robust reference on OPUC.
Consider the following problem for Krein systems.
Problem 1.1 (The Steklov problem for Krein systems).
If , what conditions on guarantee there exists such that is bounded in ? More precisely, when does there exist such that for each compact ,
(2) |
The Steklov problem and estimates like (2) are related to the following problem: for which measures is the maximal function
finite at almost every ? Known as nonlinear Carleson’s Theorem, this was recently proved in [11] for a conjectured class of measures.
The Steklov problem was originally posed for orthogonal polynomials: given the OPUC generated by the measure on the unit circle , what conditions on guarantee there exists for which
(3) |
Nazarov first showed that when , then (3) holds for some ([4]). Then in [5], Denisov-Rush generalized Nazarov’s result to the case when . And most recently in [1], Alexis-Aptekarev-Denisov further generalized these results to the case when . Since Krein systems are the continuous analogue of OPUC, we adapt the most recent successes in [1] to solutions of Krein systems generated by a weight with as defined below (see [13, p.194]).
Definition. Let . The weight if
where ranges over the finite intervals in .
Note (2) will immediately follow if we can show there exists such that if , then
(4) |
Whence the first main theorem of this paper below, for which we recall when .
Theorem 1.2.
Suppose that where with . If , then
-
(a)
there exists with such that for any satisfying , we have
-
(b)
for any there exists such that whenever , we have
(5)
Solutions to the Krein system generate generalized eigenfunctions for a unique Dirac operator with spectral measure . Furthermore, if , the coefficient of the Krein system (20), is real-valued and in , then when written in its canonical form, is a diagonal operator whose entries are Schrödinger operators. Thus, Theorem 1.2 allows one to obtain -information on generalized eigenfunctions for the Dirac and Schrödinger operators. See e.g. [3, Sections 14-16] for more details.
We note the following corollaries of Theorem 1.2.
Corollary 1.3.
If for , then for any , we have
where the limit is taken among weights satisfying , with an arbitrary absolute constant.
Proof.
Take in (5). ∎
Corollary 1.3 states that as becomes flatter, tends to standard exponential, demonstrating some continuity of solutions to Krein systems with respect to the measure.
Corollary 1.4.
Suppose with . If , then there exists such that
for all , .
Corollary 1.4 is non-trivial, since as we discuss in Section 3, a priori all we can say about is that it belongs in for .
The requirement that in Theorem 1.2 is sharp in the sense of Proposition 1.5 below, which morally says that the decay of cannot exceed the decay of .
Proposition 1.5.
If , then for every , there exists a weight such that , and
for every .
In the same vein as Theorem 1.2 (a), one may also estimate a “remainder” term, which we define precisely in Section 6. Let . If then the remainder function exists and satisfies
where , . Note that measures how much deviates from a polynomial-like term of “degree” . In Theorem 1.6 below, we estimate , which quantifies the decay of , providing information with no analogue for OPUC. Furthermore, the assumptions of Theorem 1.6 below imply
where each does not depend on , , and exists (see Lemma 6.2); as such, the estimate on below also qualifies the behavior of the “continuous polynomials” for large. We will need the following definitions for what follows.
-
•
Let denotes the Hardy space in the domain .
-
•
If is a Banach space with norm , is a measure space and , let
-
•
Given two Banach spaces with norms and , for each define
Theorem 1.6.
Suppose that for some and
(6) |
If , then there exists such that whenever satisfies , then
Compare Theorem 1.6 and Theorem 1.2 (a), the latter of which can be understood as stating that the -norm of th order remainder is finite.
We can also estimate higher order remainders when the weight is very close to .
Theorem 1.7.
Suppose for some integer . If , for some , then
-
(a)
there exists , with , such that for all satisfying ,
-
(b)
for any , there exists such that for all , we have
In particular Theorem 1.7 (b) has the following corollary, which shows some continuity of with respect to the measure.
Corollary 1.8.
If , where is any fixed constant and a nonnegative integer, then for any fixed we have
While examples of weights to which we can apply Theorem 1.7 are easy to find, e.g. any weight satisfying , let’s mention some weights to which we can apply Theorems 1.2 and 1.6. Define
which is an weight when . Then finite products , where and are all distinct, satisfy the conditions of Theorems 1.2 and 1.6.
In practice, condition (6) is difficult to check, thus making it difficult to specify weights with infinitely many singularities to which Theorem 1.6 applies. However, if we require the singularities be well spaced out, one can generate other weights that satisfy the assumptions of Theorem 1.2. Consider for instance
where , and is a sequence of positive reals such that .
As far as methods are concerned, the proofs of Theorems 1.2 and 1.6 are based on [1]’s proof of a positive answer to the OPUC Steklov problem for weights: this involves the theory of weighted operators, some basic spectral theory and complex interpolation. Meanwhile Theorem 1.7 is based on the original, elegantly simply method of Nazarov used to prove [4, Theorem 2.1], which involves just basic knowledge of the Hilbert transform.
1.1. Lingering questions: some potentially open problems
Regarding lingering questions, let us first comment on our methods. Essential to our proof of Theorem 1.2 is the orthogonality
where is Fourier projection onto the frequency band . But using for instance Lemma 4.1, we end up showing
where is any nonnegative polynomial. Thus, as far as the algebra in the proof of Theorem 1.2 is concerned, one is tempted to consider the case and prove another version of this result. But issues arise: for Krein systems and their solutions to be well-defined, the weight needs to be “centered” near , with e.g. . This directly conflicts with . Indeed, heuristically means is more or less constant, which would then mean decays to , which would contradict the required decay of . A possible remedy here is to study the more general de Branges canonical systems, wherein the weight can deviate from in more exotic ways. I have not investigated this potential framing of the problem and thus so far it remains an open question.
While Proposition 1.5 shows that the requirement that is sharp in Theorem 1.2, it is unclear to me whether the requirement that is sharp in the statement of Theorem 1.6. Furthermore, while Proposition 1.5 had a nice interpretation — that cannot decay faster that — the requirement that is more difficult to interpret. in elucidating this part of Theorem 1.6.
1.2. Organization of this Paper
This paper is organized as follows. First, we outline the basic theory of Krein systems in Section 2. In Section 3, we discuss a priori which spaces belongs to. In Section 4 we prove a useful orthogonality lemma. We spend Sections 5, 6 and 7 proving the main results of this paper, i.e. Theorem 1.2 and Proposition 1.5, Theorem 1.6, and Theorem 1.7. In Sections 8 and 9 we prove the technical Lemmas 5.3 and 5.4 essential in the proofs of the results in this paper. And finally, in Appendix A, we prove Proposition 5.5.
1.3. Notation
If is a Banach space, we denote its dual by , and the space of bounded operators on by .
If and , we define . If is Lebesgue measure, we omit , i.e. . If is a bounded linear operator between two Banach spaces , , we write its norm as .
If , the dual exponent is denoted by .
For , denote the average of over the measure space by
If and is Lebesgue measure, we will simply write .
In this paper we work with spatial variable and frequency variable . Thus for Banach spaces like or , the variable of integration is understood to be , i.e. . And given a measure, weight or function over , we write them with respect to variable , i.e. .
Given a measure space , we denote the inner product by
If it is clear from context that , we will simply write .
If , then the Fourier transform , or , of a function and its inverse transform , or are given by
Given a set where , we will use notation for its complement, i.e., .
Let denote the space of smooth functions compactly supported in , an open subset of .
For two non-negative functions and , we write if there is an absolute constant such that
for all values of the arguments of and . If the constant depends on a parameter , we will write . We define similarly and write if and simultaneously.
If a constant depends on parameter , we will express this by writing or .
2. Krein Systems Basics
In this section we explain the basics of Krein systems following [3], listing most facts needed for this paper without proof; see [3] for an accessible in-depth survey on Krein systems.
Suppose is a Poisson-finite measure over , i.e.
Definition. A function is Hermitian if .
Definition. A Hermitian function is the accelerant associated to the measure if there exists a real constant such that for all ,
(9) |
Formally differentiating (9) twice yields “.” Thus intuitively the accelerant captures the moments of the signed measure .
If an accelerant exists, then for each , define the operator on by
(10) |
Of interest is when , which occurs for a large class of measures.
Lemma 2.1.
Suppose . If , then
is the accelerant associated to , and .
Proof.
Let and first define and
(11) |
Suppose we can show that
(12) |
holds for all . Then setting , we will show that is the accelerant associated to , i.e. we will show (12) holds.
We now note that : take and extend to a function over by defining for . Then Fourier inversion for distributions yields
whenever , i.e. .
It remains to show (12) holds for as above. If we can show (12) holds for and for , then by linearity (12) holds for all .
The case that : First note that both sides of (12) are well-defined functions for . Next note that the second derivative (with respect to variable ) of the left side of (12) equals the second derivative of the right side of (12). Indeed, this follows from the dominated convergence theorem and that implies .
Thus the first derivatives of each side of (12) will be identical so long as we can verify they agree at a point, say . But is defined precisely to make this occur; one can check this using the dominated convergence theorem.
Assume is an accelerant for which for each . Then the resolvent operator
(13) |
is an integral operator on with kernel possessing the following properties.
Properties of the Resolvent kernel
-
(i)
Resolvent Identity: from (13) we have
(14) -
(ii)
Symmetries: has following symmetries:
(15) -
(iii)
Regularity inherited from accelerant: if for some , then for each , . Furthermore, is continuously differentiable in and
(16) - (iv)
Using , one can define the “continuous polynomials” . These are entire functions , each of “degree” , given by
(18) |
This definition is motivated by analogy to the OPUC: if one considers the Toeplitz matrix
where are the moments of a measure on , then the orthogonal polynomial associated to is given by last row of the column-vector
And so similar to the OPUC, is an orthonormal system with respect to , i.e. (1) holds for all . Furthermore, like how each may be expressed as a sum of the OPUC , each exponential is a “continuous sum” of the polynomials , i.e. given we have
(19) |
for some Volterra operator bounded on .
Similar to the role the Verblunsky coefficients play for OPUC (see e.g. [12, Theorem 1.5.2]), there exists such that is the solution to the Krein (differential) system
(20) |
where the derivative is taken with respect to and . If is continuous, then is given by the “coefficient” of , i.e.
(21) |
3. A priori estimates: when does ?
Suppose, just within the scope of this paragraph, that is very regular, e.g. . For (4) to hold for a fixed , it is then necessary that for each to begin with. But this is not immediate for arbitrary weights. How do properties of determine which spaces belongs to?
Proposition 3.1.
If , then for , for each . In particular, for , for each .
Thus a priori, if all we can say is that for .
We spend the rest of this section proving Proposition 3.1.
Lemma 3.2.
Proof.
Lemma 3.3.
If for an integer, then has accelerant , the resolvent kernel exists, and for each .
Proof.
Proof of Proposition 3.1.
To get the weighted-norm estimate, write
where , with and . We are left with checking the sum is integrable.
We just showed is integrable for any . Then is integrable since and . To see that is integrable, apply the Cauchy-Schwarz inequality, and use the fact that and . ∎
4. Krein system solutions are orthogonal to lower Fourier frequencies
Definition. Let denote the Fourier projection onto the frequency band , i.e.
for all . Note .
Remark. Note that . Indeed, by Plancherel’s theorem, it suffices to show
which follows from the fact that
where the last boundedness property follows from the Riemann-Lebesgue lemma.
Solutions to the Krein system satisfy the following orthogonality Lemma.
Lemma 4.1.
If , then for each nonnegative integer , we have
(24) |
for all .
Furthermore, we also have
(25) | |||
(26) |
Remark. The main statement of interest in this lemma is (24). Ignoring issues of integrability, if , then this is the intuitive statement that
whose analogue for OPUC was used in [5, 1] in addressing the Steklov problem for OPUC.
For , this gives us the more surprising statement
possessing no analogue for OPUC.
Proof.
We focus first on (24). Use integration by parts to write
Since , then it follows from Lemma 2.1 and then Lemma 3.2 that the accelerant associated to is in and is continuous in both variables. This last property allows us to write as the limit of its averages in and so we may write the inner-product as
Since and , then . Apply the dominated convergence theorem, using e.g. (22) as justification (which holds thanks to Lemma 2.1), to pull outside the inner-product, yielding
(27) |
Write
which, by the change of basis formula (19), equals
where . Thus we can rewrite (27) as
By the orthogonality of Krein system solutions (1) this equals , thereby completing the proof of (24).
As for (26), use integration by parts to write
Use Fourier inversion to write the inner-product as
where the last equality follows from . ∎
The above lemma will often be paired with the one below.
Lemma 4.2.
Let . Then given , we have , which equals if and only if
for all .
Proof.
Let with . By the remark at the beginning of the section and Plancherel’s theorem, we have .
As for the if and only if, note if and only if if and only if for all , we have
which, by taking Fourier inverses of both entries in the inner product, holds if and only if
∎
In the sections that follow, we will consider linear operators which satisfy
(28) |
where the function on is continuous in for every fixed . In what follows, we do not need to know explicitly. However, is known in many applications. For example, the Hunt-Muckenhoupt-Wheeden theorem [13, p.205] shows that can be taken as a singular integral operator and recent breakthrough on domination of singular integrals by sparse operators provides the sharp dependence of on . In particular, for a large class of singular integral operators, one can take (see, e.g., [9, p.264]).
Lemma 4.3.
The Fourier projections satisfy (28) for some independent of .
Proof.
By e.g. [9, p.264], (28) is satisfied by the Hilbert transform , which has Fourier multiplier [13, p.26]. Now note each is a linear combination of modulated Hilbert transforms, i.e.
(29) |
One can check this by e.g. looking at the Fourier multipliers of all the operators involved. The triangle inequality then yields (28) for and function independent of . ∎
5. The Steklov problem for an weight: proof of Theorem 1.2 and Proposition 1.5
In this section we prove Theorem 1.2, and demonstrate its sharpness by Proposition 1.5. We need the following result, which follows from e.g. [15, Theorem 1, Corollary to Theorem 1].
Lemma 5.1 (Reverse Hölder inequality, open inclusion of weights).
Suppose , . Then there exists , satisfying , such that for all , we have
for all intervals , and .
Furthermore, there exists such that for all , we have , where and satisfy and .
Assume and define , where is as in Lemma 5.1. Then , and
Define
(30) |
Thus, if , then
which in particular implies that for all such , we have
(31) |
Note , with .
If we additionally assume , then Proposition 3.1 implies . Together, these estimates yield
(32) |
Proposition 3.1 and the fact that also imply
(33) |
Since
it suffices to estimate in , which we’ll do using functional analysis methods. Note (33) means we can do functional analysis on in the space for any when .
Lemma 5.2.
If and , then there exists , with , such that for all satisfying , we have
(34) | ||||
(35) |
in . In particular, for all such values of we have
(36) |
where
Remark. One might wonder if ; it is by taking , and both and equal in (37) of the following lemma.
Lemma 5.3 (Integrability of weights).
Let for some . Suppose that
-
•
for some .
-
•
with .
-
•
.
Then there exists , with , such that for all satisfying , we have
(37) |
If and , then there exists a small constant such that whenever
we have the perturbative estimate
(38) |
Proof of Lemma 5.2.
Take as in (30). Then note all relevant quantities are well-defined as elements of, or operators on, : by (33), and are bounded on by the Hunt-Muckenhoupt-Wheeden theorem and the fact that . And as per the previous remark, by Lemma 5.3.
We note (34) is equivalent to
Since with as in the discussion in Section 2, then this clearly holds.
Meanwhile (35) is equivalent to
As it turns out, we can invert for sufficiently close to .
Lemma 5.4.
For , consider the formal operator
If , then there exists , independent of , with , such that for all satisfying , has bounded inverse on with operator bound
(39) |
Let us briefly discuss the strategy for proving Theorem 1.2. Using Lemma 5.4 and (36), we have
(40) |
for all sufficiently close to . Then, we can estimate by estimating , , and . However, such a process will only give us a bound for when since our starting point, (36), was only valid for . We address this is by noting that all elements involving in (40) are actually analytic in variable , and the right-side of (40) is well-defined for . Hence equality must hold for . In particular, we have the following Proposition.
Proposition 5.5.
Suppose . If , then there exists , with , such that
(41) |
for all satisfying .
Proof of Theorem 1.2.
Since satisfies (28) for some independent of , we may in fact write
We now turn our attention towards Proposition 1.5. Let us first discuss its meaning: suppose for and ; we think of as measuring the decay of , with smaller ’s indicating better decay. Recall Theorem 1.2 (a), which says that if , then
whenever satisfies , where . So let be sufficiently small that . Then the requirement that in Theorem 1.2 (a) is in part saying that cannot hope to decay faster than decays. This seemingly makes sense a priori: since the accelerant satisfies , then whatever decay possesses gets “converted” into the regularity of . But then the resolvent kernel , as a rule of thumb, is at most as regular as . Since by (18) we have
i.e. is essentially the inverse Fourier transform of and so whatever regularity possesses gets “converted” into the decay of . Thus heuristically, we expect the decay of to not exceed that of , as the former “inherits” its decay from the latter.
Proof of Proposition 1.5.
By Hölder’s inequality, we can assume without loss of generality that .
Without loss of generality, assume . Let be an even, real-valued function on such that and if and only if . Then define so that and .
Without loss of generality, assume sufficiently small that , where is as in Lemma 5.4. Then by Lemma 5.4, . Thus by Proposition 5.5, we have
or rather
Estimating norms yields
Since , then and so by Lemma 4.3, we have and so
This then means
It suffices to show
Since , write , where . Since is real-valued and even, then so is . In particular this means , and combined with the fact that is real-valued, we get if .
In fact, is unbounded in . Indeed, suppose to the contrary the sequence is bounded. Then there exists some increasing sequence of integers and some such that converges to weakly in . In particular, for every Schwarz function , we have
where the last equality follows from Plancherel’s theorem. Thus , which contradicts our requirement that if and only if . This completes the proof. ∎
6. A mixed norm remainder estimate
In this section, we prove Theorem 1.6.
Definition. If , integrate (18) by parts times to yield
where
(42) |
and the remainder term is defined by
(43) |
We can also express the remainder in terms of the solution to the Krein system and the “coefficients” , i.e.
Note that .
The Steklov problem can be reformulated in terms of mixed norms: it asks when does one have the bound
(44) |
for every compact ? In Theorem 1.2, we estimated
for some close to , which implied (44). However, it also makes sense to consider other mixed norms, as we do in Theorem 1.6. To estimate in , it suffices to estimate in .
We begin with two lemmas.
Proof.
By [3, Theorem 12.14], follows if both Condition (6) holds and . Thus it suffices to check our assumptions imply .
Note that when , then . Whence
Since , then has finite Lebesgue measure. By taking and in Lemma 5.3, we get and in particular is square-integrable on . Meanwhile is integrable on since . Thus the -triangle inequality implies is integrable on : indeed, equals
And hence must be integrable on as well. ∎
Lemma 6.2.
Proof.
Since , then by Lemma 3.3 it follows that . Hence is well-defined and is given by
Now we can do functional analysis like in the proof of Theorem 1.2.
Lemma 6.3.
Suppose for some . If , then there exists , with , such that for all satisfying , we have
(46) | ||||
(47) |
in . In particular, for all such values of , we have
(48) |
where .
We also remark that the right-side of (47) is well-defined, i.e.
for , for each . Indeed this follows from Lemma 5.3, (45), and that . We require just so that we may consider , which we will do later.
Proof.
Take as in (30), which in particular implies (31) holds for all of concern, i.e. for all such that . Then note all relevant quantities are well-defined as elements of, or operators on, : and are bounded operators on by the Hunt-Muckenhoupt-Wheeden theorem and that . And as per the remark above, , and by Lemma 5.3.
Concerning (47), it is equivalent to
We note that is acting on
to see this, rewrite it as
By the previous remark, and also . If we combine these two estimates with the assumption that , then we get . Finally, combine assumption with the estimate , which follows from (42), to get .
Proof of Theorem 1.6.
It suffices to estimate ; we first estimate in terms of .
7. Higher order remainder estimates: proof of Theorem 1.7
We will spend the rest of this subsection proving Theorem 1.7. Suppose
Then by Lemma 3.3, and so is well-defined. The following Lemma shows are uniformly bounded in .
Lemma 7.1.
Suppose . If , then
In particular .
Proof.
The bound on follows from its definition in (42) and the bound on . To get this latter bound, by the resolvent symmetries (15) it suffices to estimate and all its derivatives in . The resolvent identity (14) is equivalent to
and thus
By Lemma 2.1 we have ; we in fact claim . Indeed:
Thus by geometric sum
We estimate and all its derivatives. Differentiate (10) to get
for all . Take to get
where in the last inequality we estimated
and similarly for .
Thus the sum converges absolutely and so one can use the dominated convergence theorem to show we get the desired estimate on
∎
Now we proceed with our usual functional-analytic approach.
Lemma 7.2.
Suppose that . Then
(49) |
and
(50) |
in for . If in addition , then
(51) |
where both sides are well-defined in, e.g., .
Proof.
We will also need the following Lemma to estimate the Fourier projections.
Lemma 7.3.
If , then there exists , with , such that for all satisfying we have
Proof.
By (29) and self-adjointness of , it follows that , where is the Hilbert Transform. By [10], for . Let be the unique element of that satisfies , and let .
To see that , it suffices to note that is increasing in for and has singularity at . Thus as , we have , meaning that for our definition of , we have and so . ∎
We are now in a position to prove Theorem 1.7.
8. Proof of Lemma 5.3
Before proving Lemma 5.3, we will need terminology and lemmas for a Calderón-Zygmund decomposition.
Definition. An interval has left neighbor and right neighbor .
Two intervals are almost disjoint if their intersection is empty or a single point.
Given an interval , let denote its distance from the origin.
An interval in is dyadic if it is of the form .
Two dyadic intervals and are siblings if their union is a dyadic interval of strictly larger size, which we call the parent.
So define a partial ordering on the dyadic intervals: if , we write , and say is an ancestor of . Note that two dyadic intervals are either almost disjoint, or comparable via .
See e.g. [13, Chapter 1, Section 3] or [14, Chapter 1, Theorem 4] for other variants of the Calderón-Zygmund decomposition below.
Lemma 8.1 (Calderón-Zygmund decomposition).
Let for some , and suppose . If and , then there exists a collection of dyadic intervals with the following properties:
-
(i)
.
-
(ii)
Each is maximal with respect to among those dyadic intervals satisfying .
-
(iii)
The are pairwise almost disjoint.
-
(iv)
We have the estimate
(53)
This is known as the Calderón-Zygmund decomposition of at level .
Proof.
Fix , and by substituting with , assume without loss of generality that . Since , then for any interval of size , we have , since . Thus any dyadic interval has an ancestor such that for all dyadics , we have . Meaning if we consider the set of dyadic intervals
then for each , there exists such that , and is maximal in with respect to . Define to be the set of maximal dyadic intervals in . By definition, satisfies Property (ii), which then immediately implies Property (iii).
We are left with showing Property (i), which will follow if we can show for almost every in . Consider the function . By the dyadic Lebesgue differentiation theorem, for almost every , we have , where is any sequence of dyadic intervals shrinking to . Fix such a point . Then by construction of , for every dyadic interval containing , it follows that
Letting shrink down to , the above yields
i.e. . Since arbitrary in where is a set of measure , we get Property (i). ∎
Lemma 8.2 (Calderón-Zygmund decomposition for ).
Let for some , and suppose with , , where . For each , define and , . Then , all of which are of finite -measure, and there exists a collection of dyadic intervals with the following properties:
-
(i)
.
-
(ii)
The are pairwise almost disjoint.
-
(iii)
We have the estimate
(54) -
(iv)
We have
(55)
Proof.
By the triangle inequality, . Thus , since .
For , we apply Lemma 8.1 to at level and get a collection of pairwise disjoint dyadic intervals such that
Let . Note by maximality of the ’s specified by Lemma 8.1 (ii), each is contained in at most one other . So now let be the maximal intervals among with respect to ; maximality ensures the are pairwise almost disjoint and so (ii) holds. Note maximality also yields
i.e. (i) holds. And (53) yields (54), as
Also note by our choice of as the maximal intervals among , then for any ancestor of , it follows that
The triangle inequality followed by Hölder’s inequality then imply (55), i.e.
∎
We also need the following additional lemma. Recall that for an interval , we defined .
Lemma 8.3 ( is flat away from ).
Let for some , and suppose is an interval such that , where is some constant. Then there exists sufficiently large so that if , then
(56) |
In particular,
(57) |
for all , where may equal , either of its neighbors, or its parent in the case that is dyadic.
Proof.
If , the lemma is trivial.
For , choose so large that . Thus if , then
and so . Thus on or either of its neighbors, we have . Hence for or either of its neighbors, (57) holds as
If is dyadic, then since the parent of is the union of and one of its neighbors, it follows on the parent of . Then the same computations as done previously yield (57) when is the parent of . ∎
Proof that in Lemma 5.3.
Let and . Let’s apply Lemma 8.2 with and , where , so that and . Then .
We begin by splitting
(58) |
Since , we may estimate
By the triangle inequality, is
Thus,
(59) |
Recall and so . By applying Lemma 5.1 to both of these weights, we can choose , with , so that whenever is chosen so that , we have
(61) |
In particular, this means and are locally integrable.
Let us focus on first, as will be similar. Set , then write
where is a constant which will be chosen later.
Since by (54), then the length of each interval is bounded. Since is locally integrable, then
where is some sufficiently large interval centered at . And so it suffices to show
Since , to show , it suffices to show ; similarly will follow from showing for a suitable choice of .
Since there exists some constant such that
we can take as in Lemma 8.3 so that when , we have on , its neighbors and its parent .
For the proof of (38), we need to understand how affects .
Lemma 8.4.
Suppose and let be an interval. If , then
Proof.
By Jensen’s inequality,
But also since , then
which implies
Combine the estimates of from above and below to get
∎
Next, we recall a few facts about , the space of functions with bounded mean oscillation, and how it relates to weights. Recall that if
where denotes a ball in (see, e.g., p.140 in [13]). Functions in all satisfy the John-Nirenberg estimates below.
Theorem 8.5 (John-Nirenberg, [13, p.144-146]).
Suppose . Then
-
(a)
There exist positive absolute constants such that for each and every ball ,
-
(b)
For any , and
for all balls .
-
(c)
If , then
It is well known that if , then . We also have the following well-known quantification (see, e.g., [7, Corollary 6]).
Lemma 8.6.
If for some , then . If in addition , then
Proof of (38).
By Lemma 8.6, when we have . Take small enough so that additionally .
Set and define , and as in the proof that , i.e. by applying Lemma 8.2 with and . We begin by splitting as in (58), with the same definitions of and . By (59) and our choice of , it follows that
As for estimating , we split slightly differently than in (60):
It suffices to show
(62) |
We prove this for ; the estimate for will follow similarly.
Note that (62) will follow if we can show
(63) |
for all : indeed, sum over and use (54) to get (62). As such, fix ; our goal is to show (63).
Now note
(64) |
Indeed, if is the dyadic parent of , then (55) guarantees . By Lemma 8.4, we get (64) for sufficiently small, which we will make use of repeatedly.
Split the left side of (63) into
But equals
where and we applied (64) in the inequality. Write
and apply Minkowski’s inequality, followed by the Cauchy-Schwarz inequality, to get
The John-Nirenberg Theorem 8.5 yields
for small enough, which we can arrange by taking as small as necessary and applying Lemma 8.6. Thus
Hence , which completes the proof. ∎
9. Inverting : proof of Lemma 5.4
Let us introduce some notation and definitions.
Notation. Given , we will generally write , where are real-valued.
Given , we let denote an open set containing , although we allow for the particular set to change line to line.
Definition. Given some , define
(65) |
Given , define the open strip
We define the closed strip similarly.
Given an interval, let
Lemma 9.1.
Let be an arbitrary absolute constant, and let is as in (65) for some . For , consider the operator
where is an operator satisfying (28) for some and is also self-adjoint as an operator on .
Then there exists , with , such that on , has bounded inverse on with operator bound
(66) |
and is analytic as a map .
Lemma 5.4 follows from Lemma 9.1 by taking , and applying Lemma 4.3. As such, this section is dedicated to the proof of Lemma 9.1. The proof strategy will proceed more or less as follows:
-
(1)
Show uniformly bounded on the strip.
-
(2)
Show that if we have uniform bounds on is uniformly bounded on the strip, then is weakly analytic.
-
(3)
Chop into small pieces and show we have uniform bounds for plus the small piece; using the two previous parts of the strategy, we have our function is weakly analytic. We maintain boundedness while adding the small pieces by applying analytic interpolation.
Before we begin, we will need a few preliminary lemmas.
Proposition 9.2.
Suppose is an Banach space and are linear bounded operators from to . Then,
provided the operators involved are well-defined and bounded in . Moreover, assuming , we get
(67) |
Finally, if , then
(68) |
The proof of this proposition is a straightforward calculation.
We will also need continuity of weighted operators as proved in [1, Theorem 1.2].
Theorem 9.3 ([1, Theorem 1.2]).
Definition. Suppose is a set of linear functionals acting on a vector space . If is an open set, and is contained within , then is weakly analytic with respect to if is analytic for all .
We will generally be concerned with two cases:
-
(1)
for , and , the space of simple functions with compact support, where acts on by .
-
(2)
and .
Lemma 9.4.
Let be an operator satisfying (28) that is self-adjoint on . If , then there exists , with , such that for all , we have
(69) |
and
(70) |
and is weakly analytic with respect to on .
Proof.
Initially define where is as in Lemma 5.1. Then , , and (69) holds for ; by taking a slightly smaller , still with , then (69) holds for .
If satisfies (28), then
for all satisfying . Since , we can interpolate between the extremal values of given by to in fact get
for all satisfying . In particular,
for all . If we swap the role of with , this has the net effect of swapping the role with . Thus if is self-adjoint on , then swapping as described and taking adjoints yields
for . Combining all of this together yields (70) for ; by slightly shrinking , while still requiring , we get (70) for .
By Proposition A.4, if we again shrinking we may assume without loss of generality that is analytic as a map , and hence weakly analytic with respect to on . ∎
Lemma 9.5.
Proof.
Let . We show that for each , there exists a ball on which is analytic as a map ; this will clearly imply weak analyticity on . Fix , define
and write
(71) |
We claim that there exists a ball on which , and hence , is analytic as a map . Given the claim, we can then take small enough so that . Then (71) yields
where the sum converges uniformly in . Since each partial sum is analytic in and the sum converges uniformly, we have is analytic as a map .
We must now show there exists a ball on which is analytic as a map . As per Lemma 9.4, is weakly analytic with respect to on , and so by Proposition A.1 it suffices to show there exists such that for all , we have . We show this for , as the proof for will follow similarly. Write
The last term is clearly finite. As for , by (69) we have . Thus we can apply the continuity of weighted operators Theorem 9.3 with , to argue that for sufficiently small, we have . This completes the proof. ∎
Proposition 9.6 ([1]).
Proof.
We notice that is bounded and antisymmetric operator in Hilbert space . Therefore, . By Lemma 9.5, is weakly analytic and continuous in the sense of Stein (p.209, [2]). Applying Stein’s interpolation theorem on the strips and , we get
∎
Remark. We emphasize here that positive does not depend on .
Proof of Lemma 9.1.
The reader may also consult a similar, but simpler proof in [1, Proof of Lemma 3.6].
We just need to find such that on , we have is bounded on and analytic as an -valued function.
In (65), we take parameter as follows: define by , where is as in Lemma 9.4, and set ; note then . Consider , , where is large and will be fixed later (it will depend on , , only).
We continue with an inductive argument in which the bound for provides the bound for when .
Base of induction: handling . Apply Proposition 9.6 with to get an absolute constant so that
for and . Next, we use (67) with and . This gives
(73) |
by (72).
That finishes the first step. Next, we will explain how estimates on give bounds for .
Handling . In Proposition 9.6, we now take (here is obtained at the previous step) and compute new by (65):
(74) |
Therefore, when belongs to , belongs to , and still satisfies . In this domain, the estimate (73) can be rewritten as
where is different from only by the choice of parameter in (65) and is in fact a rescaling of the original as follows from (74). From Proposition 9.6, we have
for . We use the perturbative bound (67) one more time with and to get
for .
Induction in and the bound for . Next, we take and repeat the process in which the bound
implies
Notice that each time the new is in fact a rescaling of the original by as can be seen from a calculation analogous to (74). In steps, we get
Thus recalling that , one has
Since , we get (66) with
The estimates (72) implies that we can take
Thus, we showed there exists such that for all .
In fact, we may assume without loss of generality that : indeed, it suffices to show that for any , we can choose sufficiently small so that for all satisfying . Note that if , then by geometric sum
Thus it suffices to show we can take sufficiently small so that
for . In fact, by Stein’s analytic interpolation it suffices to check this at the vertical lines such that . By duality and the triangle inequality this in turn will follow from showing
for such that . Fix one the values satisfying the last equality: if is sufficiently small, then by Lemma 8.6 we have , which can be made arbitrarily small. Apply Theorem 9.3 with to get
We can now choose sufficiently small that this is at most . Thus, we can choose so that as , we get .
As for showing is analytic as an -valued function, by Lemma 9.5 and Proposition A.1, it suffices to show is bounded on for . Fix one such ; we just showed previously that
However the proof just as easily applies to weight , operator and Hölder index (as opposed to ) and so we get
Interpolate between both estimates to get
This completes the proof. ∎
Appendix A A detour through complex analysis: proof of Proposition 5.5
We recall the following well-known lemma, which we prove for the sake of completeness.
Proposition A.1.
Let be a Banach space, let be a dense subset of the dual space and an open set. If is weakly analytic with respect to , then is analytic.
Proof.
We adapt the proof of [6, Theorem 8.20].
Let so that is analytic.
Fix and let for sufficiently small. By the Cauchy integral formula,
Then
(75) |
For with , define the second order difference quotient
It suffices to show is uniformly bounded in , for both sufficiently small. Indeed, uniform boundedness in would then yield
is Lipschitz for near , and so has a limit as .
By (75),
For , the denominator in the integral is bounded uniformly away from , and the numerator is bounded above by , which is finite by the uniform boundedness principle. Whence
where does not depend on . By duality
∎
Fix a characteristic function of some finite interval , and let , like in Section 5.
Proposition A.2.
Suppose , and is as in (65). Then is analytic as a map
Proof.
Without loss of generality, assume ; the general case follows by a rescaling argument.
We first compute
Since , then by Lemma 5.1 there exists such that for all , . Since , then and so for some by (65).
Now let us show analyticity: write
by Taylor expansion of . We simply need to check the series converges uniformly in norm for sufficiently small; by Stirling’s approximation, it suffices to show
Apply Hölder’s inequality to bound
where is the dual exponent to .
Since we already showed , then we will be done once we show . If , write
It suffices to show .
Write
The last term is . For the first term, the John-Nirenberg Theorem 8.5 yields
where function is the usual analytic extension of factorial. Finally, Stirling’s formula yields
thereby completing the proof. ∎
Remark. To avoid repeating similar arguments, from now onwards if we write that a function or operator is analytic (or weakly analytic) thanks to a “John-Nirenberg argument,” the reader should understand this as meaning that analyticity follows from an argument similar to the one above where we showed was analytic as a map .
Lemma A.3.
Suppose , and is as in (65). Then there exists , with , such that is analytic as a map .
Proof.
By a John-Nirenberg argument, and are weakly analytic with respect to on . By Proposition A.1, this means is analytic as a map . ∎
Proposition A.4.
Proof.
We will prove the portion of the proposition involving ; the result for follows by then replacing by and by .
A John-Nirenberg argument shows that is weakly analytic with respect to for all where . By Proposition A.1, is then analytic as a map if we can show on for some . To see this latter part, notice that
By Lemma 5.1, there exists such that when , we have . By (28), the operator is then indeed bounded on . ∎
Corollary A.5.
Suppose and is as in (65) for some . Then there exists , with , such that are analytic as maps .
Remark. In what follows, we will need to show some exists, such that . In our reasoning, we will consider various other such functions . We consider only finitely many such functions, so by taking the minimum of all of them we will assume without loss of generality that all the ’s are the same.
Corollary A.6.
Proof.
Proof of Proposition 5.5.
We begin with (36). By applying Lemma 4.3 and Lemma 9.1 with and e.g., we can invert to get (41) for all satisfying , where and . So for all satisfying , we have
(76) |
for all .
Remark. Both sides of (41) always makes sense as elements of for .
References
- [1] M. Alexis, A. Aptekarev, and S. Denisov. Continuity of Weighted Operators, Muckenhoupt Weights, and Steklov Problem for Orthogonal Polynomials. International Mathematics Research Notices, 2022(8):5935–5972, 10 2020.
- [2] C. Bennett and R. Sharpley. Interpolation of operators, volume 129 of Pure and Applied Mathematics. Academic Press, Inc., Boston, MA, 1988.
- [3] S. Denisov. Continuous analogs of polynomials orthogonal on the unit circle and Kreĭn systems. International Mathematics Research Surveys, 01 2006.
- [4] S. Denisov. On the growth of polynomials orthogonal on the unit circle with a weight that satisfies . Mat. Sb., 209(7):71–105, 2018.
- [5] S. Denisov and K. Rush. Orthogonal polynomials on the circle for the weight satisfying conditions . Constr. Approx., 46(2):285–303, 2017.
- [6] M. Einsiedler and T. Ward. Functional Analysis, Spectral Theory, and Applications. Graduate Texts in Mathematics. Springer International Publishing, 2017.
- [7] M. B. Korey. Ideal weights: asymptotically optimal versions of doubling, absolute continuity, and bounded mean oscillation. J. Fourier Anal. Appl., 4(4-5):491–519, 1998.
- [8] M. G. Krein. Continuous analogues of propositions on polynomials orthogonal on the unit circle. Dokl. Akad. Nauk SSSR (N.S.), 105:637–640, 1955.
- [9] A. K. Lerner and F. Nazarov. Intuitive dyadic calculus: the basics. Expo. Math., 37(3):225–265, 2019.
- [10] S. Pichorides. On the best values of the constants in the theorem of M. Riesz, Zygmund and Kolmogorov. Studia Mathematica, 44(2):165–179, 1972.
- [11] Alexei Poltoratski. Pointwise convergence of the non-linear fourier transform. arXiv:2103.13349, 2021.
- [12] B. Simon. Orthogonal Polynomials on the Unit Circle. Part 1: Classical Theory. Colloquium Publications. American Mathematical Society, 2004.
- [13] E. M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.
- [14] Elias M. Stein. Singular Integrals and Differentiability Properties of Functions (PMS-30). Princeton University Press, 2016.
- [15] V. I. Vasyunin. The exact constant in the inverse Hölder inequality for Muckenhoupt weights. Algebra i Analiz, 15(1):73–117, 2003.