Spectral clumping for functions and distributions decreasing rapidly on a half-line
Abstract.
We demonstrate a phenomenon of condensation of the Fourier transform of a function defined on the real line which decreases rapidly on one half of the line. For instance, we prove that if is square-integrable on , then a one-sided estimate of the form
for some , forces the non-zero frequencies to clump: this set differs from an open set only by a set of Lebesgue measure zero, and is locally integrable on . In particular, if is non-zero, then there exists an interval on which is integrable. The roles of and above may be interchanged, and the result extends also to a large class of tempered distributions. We show that the above decay condition is close to optimal, in the following sense: a non-zero entire function exists which is square-integrable on , for which is a subset of a compact set containing no intervals, and for which the estimate , , holds for every .
1. Introduction
1.1. Fourier transform, its support and size
This note studies a certain manifestation of the uncertainty principle in Fourier analysis, where a smallness condition on a function forces its Fourier transform to be, in some sense, large. Vice versa, smallness of forces to be large. In our context, the smallness is defined in terms of a one-sided decay condition, and the largeness in terms of the existence of a clump. This will be our moniker for an interval on which the function has an integrable logarithm. We emphasize that our results concern functions with a spectrum which might vanish on an interval (commonly referred to as functions with a spectral gap), but for which the spectrum should be large on some other interval.
We will use the following definition of the transform:
(1.1) |
Here is a normalization of the Lebesgue measure on . Then, the inversion formula is given by
(1.2) |
For , let be the usual Lebesgue space of functions for which is integrable with respect to . The formula (1.1) can be interpreted literally only for . It is interpreted in terms of Plancherel’s theorem in the case , and in order to state our most general results we will later need to interpret the transform in the sense of distribution theory. The spectrum of a function is the subset of on which lives. Since is defined only up to a set of Lebesgue measure zero, so is the spectrum in this case. If we accept making errors of measure zero (which we will), we may define the spectrum as
Note specifically that our definition of might not coincide with the usual notion of closed support of the distribution .
The uncertainty principle in Fourier analysis presents itself in plenty of ways, and the excellent monograph [6] of Havin and Jöricke describes many of its most interesting interpretations. One of them is the following statement, well-known to function theorists. If is non-zero and is the negative half-axis, then we have the implication
(1.3) |
Here the extreme decay (indeed, vanishing) of on a half-axis implies global integrability of against the Poisson measure . A fortiori, is integrable on every interval of . Naturally, this is not typical. By Plancherel’s theorem, every function in is the Fourier transform of some other function in the same space. So on the other extreme, plenty of functions have a Fourier transform which lives on sparse sets containing no intervals. This forces the divergence of the logarithmic integral of over any interval. In other words, plenty of functions admit no spectral clumps. The results of this note give conditions under which such clumps form.
1.2. Condensation and sparseness of spectra and supports
We shall introduce our results at first in the context of the Hilbert space . Here we can prove a claim which symmetric in and , and also we can argue for near-optimality of the result. This is the content of \threfCondensationTheorem and \threfSparsenessTheorem. The more general distributional clumping result is presented in \threfDistributionalClumpingTheorem.
Theorem A.
CondensationTheorem If satisfies the estimate
(1.4) |
for some constant , then there exists an open set which coincides with up to a set of Lebesgue measure zero, and for every there exists an interval containing such that
In other words, the one-sided decay condition (1.4) implies that lives on the union of the spectral clumps of . Since the Fourier transform is a unitary operation on , the roles of and may obviously be interchanged in the statement of \threfCondensationTheorem. Thus a one-sided spectral decay condition of implies local integrability properties of on the set where lives. In this form, the result encourages us to extend it to tempered distributions. We shall do so in a moment.
The integrand in (1.4) may seem a bit unnatural in the context of square-integrable functions . It is more natural in the context of functions of tempered growth appearing in \threfDistributionalClumpingTheorem. Anyhow, we note that one can prove that an estimate of the form in fact implies (1.4) for some slightly smaller .
We can prove also that the condition (1.4) on the decay of appearing in \threfCondensationTheorem is close to optimal. We do so by exhibiting a non-zero function with rapid one-sided decay but sparse spectrum.
Theorem B.
SparsenessTheorem For every , there exists a compact set contained in which contains no intervals, and a non-zero entire function which satisfies
for every , and such that is contained within .
After an initial reduction, the proof of this result follows ideas of Khrushchev from [8]. Note that the function appearing in \threfSparsenessTheorem is entire by the virtue of having a spectrum of compact support. More importantly, the condition on implies that has positive Lebesgue measure for every interval , so we obtain
for every interval . This is in contrast to the conclusion of \threfCondensationTheorem. It follows that the exponent in estimates of the form is critical for the spectral clumping phenomenon.
As mentioned above, clumping statements makes sense for objects in a class much wider than . Here is our distrbutional result.
Theorem C.
DistributionalClumpingTheorem Let be a tempered distribution on which is a measurable function satisfying
for some . If the distributional Fourier transform is an integrable function on some interval , and the estimate holds for all sufficiently large positive , then there exists an open set such that vanishes almost everywhere outside of , and for each there exists an interval containing satisfying
(1.5) |
For instance, the result shows that a function which lives on a sparse set containing no intervals cannot satisfy even a one-sided spectral decay condition of the form . Note also that in this extended form, our result includes the trivial but important examples such as and (Dirac delta), the trigonometric functions and the polynomials.
1.3. A converse result
The Beurling-Malliavin theory implies a partial converse result. If is a locally integrable function on which has a clump as in (1.5), and a constant is given, then a bounded multiplier exists for which has a Fourier transform satisfying for . To see this, recall that a smooth function supported in exists which satisfies the bilateral spectral decay , (this simpler version of the famous Beurling-Malliavin theorem is proved in [6, p. 276-277], and in fact we may ensure an even faster bilateral spectral decay of ). There exists also a bounded function which satisfies and on (we use the assumption that is a clump for and construct as in (2.7) below). Then an argument similar to the one used in the proof of \threfConvolutionFourierDecayLemma below shows that the function will satisfy the desired one-sided spectral decay, and clearly for some bounded function supported in .
1.4. Clumping in other parts of analysis
The motivation for the research presented in this note was a desire to produce a self-contained exposition of the clumping phenomenon which was observed in two other contexts, both somewhat more esoteric than Fourier analysis on the real line.
The first of these is a polynomial approximation problem in the unit disk . Here we are presented with a measure
where and are the area and arc-length measures on and . The functions and are non-negative weights, and one would like to understand under which conditions splitting occurs. Namely, when is the weighted space contained in the closure of analytic polynomials in the -norm induced by the measure ? In the case that decays exponentially as , the necessary and sufficient condition is that has no clumps, or in other words that the integral of diverges over any arc on . The lack-of-clumping condition was conjectured by Kriete and MacCluer in [9] and confirmed in [11]. Some of the techniques used in the proofs of the results in the present note are adaptations of the ideas from [11].
The other context is a circle of ideas surrounding the Aleksandrov-Clark measures appearing in spectral theory, and spaces defined by de Branges and Rovnyak, well-known to operator theorists. To any positive finite Borel measure on we may associate a so-called Clark operator which takes a function to the analytic function in given by the formula
The operator maps onto a space of analytic functions denoted by , the symbol function itself being related to by the formula
in the case that is a probability measure, with a similar formula in the general case. For many choices of (or equivalently, choices of ), the space is somewhat mysterious, with the distinctive feature of containing very few functions extending analytically to a disk larger than . This extension property is characterized by the exponential decay of the Taylor series of the function, and the clumping of the absolutely continuous part of is decisive for existence and density of functions in which have a Taylor series decaying just a bit slower than exponentially. Results of this nature are contained in [12]. In fact, a Fourier series version of \threfCondensationTheorem is a consequence of the results in [12].
1.5. Other forms of the uncertainty principle
The implication (1.3) has a well-known Fourier series version. If a function defined on the circle is integrable with respect to arc-length on , and the negative portion of the Fourier series of vanishes, then , unless is the zero function.. Volberg derived the same conclusion from the weaker hypothesis of nearly-exponentially decaying negative portion of the Fourier series (see [15] and the exposition in [16]). Work of Borichev and Volberg [3] contains related results.
The decay condition (1.4) on prohibits from living on a set containing no intervals. Somewhat related are uniqueness statements in which one seeks to give examples of pairs of sets for which the following implication is valid: if in a certain class lives on and lives on , then . One says that is then a uniqueness pair for the corresponding class. A famous result of Benedicks presented in [2] (see also [1]) says that is a uniqueness pair for integrable if both sets have finite Lebesgue measure, and the result holds not only for the real line but also for the -dimensional Euclidean space . Hedenmalm and Montes-Rodríguez worked with the hyperbola and th class of finite Borel measures supported on which are absolutely continuous with respect to arclength on . They proved in [7] that if vanishes on certain types of discrete sets , then , thus exhibiting interesting uniqueness pairs of the form . Recent work of Radchenko and Viazovska on interpolation formulas for Schwartz functions in [13] gives examples of pairs of discrete subsets and of for which is a uniqueness pair for functions in the Schwartz class. Kulikov, Nazarov and Sodin exhibit similar interpolation formulas, and consequently new uniqueness pairs, in their recent work in [10].
1.6. Notation
For a set and a measure defined on , the space denotes the usual Lebesgue space consisting of equivalence classes of functions living only on and satisfying the integrability condition . The containment is interpreted in the natural way. The symbols such as , and denote the usual Lebesgue measure of the real line, while will be the normalized version used in formulas involving Fourier transforms. If is a subset of , then denotes its usual Lebesgue measure. The positive half-axis of is denoted by , and we set also . The notions of almost everywhere and of measure zero are always to be interpreted in the sense of Lebesgue measure on . The indicator function of a measurable set is denoted by . Finally, we put .
2. Preliminaries
Our proofs will use Hilbert space techniques and the complex method. In particular, we will use the complex interpretation of the Hardy classes of functions on the line with positive spectrum. In this section, we recall those basic facts of the theory of the Hardy classes and which will be important in the coming sections. We discuss also properties of the shift operators on weighted spaces on the real line, and their invariant subspaces.
2.1. Hardy classes
For equal to or , we denote by the subspace of consisting of those functions for which the Fourier transform vanishes on the negative part of the real axis:
It is a well-known fact that functions in the Hardy classes and admit a type of analytic extension to the upper half-plane
We recall what exactly is meant by this extension and how it can be constructed. The Poisson kernel of the upper half-plane
admits a decomposition
(2.1) |
Since
(2.2) |
we may use Fubini’s theorem to compute, in the case , that
(2.3) |
where the vanishing of the integral follows from
which holds for any by the definition of the class. In the case this argument does not work, but what instead works is an application of Plancherel’s theorem and \threfFourierTransformCauchyKernel below to the first integral in (2.3), which again shows that this integral vanishes. Consequently, whenever for , the formula
defines, by the second integral expression above, an analytic extension of to . By the first expression, and classical properties of the Poisson kernel (see [4, Chapter I]), this extension satisfies
(2.4) |
Moreover, we have
(2.5) |
and
(2.6) |
if . The property (2.5) follows readily from the Poisson integral formula for the extension of and Fubini’s theorem. The property (2.6) is a bit tricky to establish, and is proved in [4, Chapter I, Theorem 3.1]. In fact, the above listed properties characterize the functions in the Hardy classes.
Proposition 2.1.
The proposition is not hard to derive from \threfH1FourierTransformFormula below. Anyway, a careful proof can be found in [6, p. 172].
The following restriction on smallness of the modulus of a function will be of crucial importance to us.
Proposition 2.2.
HadyClassLogIntProp If , then
unless is the zero function.
A proof of the proposition can be found in [6, p. 35].
We shall also need to use the corresponding Hardy class of functions which are merely bounded on , and not necessarily integrable or square-integrable on . We use directly the complex interpretation of the class. Namely, we define to consist of those functions which can be realized as limits
for almost every , where is bounded and analytic in . It can be checked that such has a distributional spectrum which vanishes on . Another important point is that if , then
since we may apply \threfH1CharacterizationProp to the analytic function
A function of a given (bounded, measurable) modulus on may be constructed by setting
(2.7) |
and . The integral above converges if
which is a necessary condition for the construction to be possible. Then
so that the equality for almost every is a consequence of the well-known properties of the Poisson kernel.
2.2. A formula and an estimate for the Fourier transform of a Hardy class function
If , then the values of may be computed using a formula different from (1.1). To wit, denote by the extension of to which was discussed in Section 2.1. The function
is analytic in , and for this reason Cauchy’s integral theorem implies
(2.8) |
where denotes the complex line integral, are all positive numbers, , and denotes the rectangular contour having as corners the four points with coordinates , , , , oriented counter-clockwise.
Fix and let denote the horizontal strip in consisting of all complex numbers with imaginary part between and . Then it follows from Fubini’s theorem and (2.5) that
where denotes the area measure on the complex plane. This expresses the integrability on of the continuous function
Hence there exists a positive sequence which satisfies
and for which
This means that
Moreover, equation (2.6) quite easily implies
We have proven the following formula by combining the above two expressions.
Proposition 2.3.
H1FourierTransformFormula For we may compute the Fourier transform using the formula
for any choice of , where denotes the values of the analytic extension of to .
This formula has the following simple corollary which will be of critical importance below.
Corollary 2.4.
FourierDecayFromExtensionGrowthCorollary If has an analytic extension to which satisfies, for some constant , an estimate of the form
then the Fourier transform of the function
satisfies
Proof.
It was mentioned in Section 2.1 that . Therefore, we may use \threfH1FourierTransformFormula to estimate
Since can be freely chosen, we may now set it to to obtain the desired estimate. ∎
2.3. A semigroup of operators and its invariant subspaces
If and , the operator given by
is unitary on . We shall be interested in subspaces of which are invariant for the operators in the semigroup . Given any element , we denote by the smallest closed linear subspace of which contains and also all the functions , .
Proposition 2.5.
UsInvSubspacesProp Let be a non-zero element which satisfies
Then the subspace coincides with , where .
Remark 2.6.
As usual, the set above is defined in a bit imprecise way. Since is, strictly speaking, merely a representative of an equivalence class of measurable functions in , the set is not well-defined pointwise. However, it is well-defined up to a set of Lebesgue measure zero, and so the initial choice of the representative is unimportant.
Proof of \threfUsInvSubspacesProp.
Since the function vanishes almost everywhere outside of the set , then so does for any . Consequently, . Conversely, let us consider an element with the property that
Setting , we note that the vanishing of the integrals above is equivalent to being a member of the Hardy class . We note also that
The above equality is to be interpreted in a generalized sense: the first integral on the right-hand side is divergent by our assumption, and so may the second, but their positive parts are certainly finite by the assumption that . This implies that
HadyClassLogIntProp now shows that must be the zero function. Since on and vanishes outside of , this means that . So is a closed and dense subspace of , which means that the two spaces are equal. ∎
Corollary 2.7.
invSubspaceCorollary If is also a member of for some , and if
then coincides with , where .
Proof.
To prove the corollary we need to verify the condition in \threfUsInvSubspacesProp. Note that, pointwise, we have
The coefficients and are positive. The inequality for shows that
Note that the integral on the left might very well be equal to , but that is of no concern to us: we conclude from the assumption, and the pointwise inequality above, that
and apply \threfUsInvSubspacesProp. ∎
3. A product space and its Hardy subspace
Let be a bounded, continuous, non-negative and decreasing function, and be a non-negative function. We consider the product space . Inside of this space we embed the linear manifold in the following way:
The tuple is well-defined as an element of the product space, since both and are members of and both and are bounded. We define the Hardy subspace as the norm-closure of the linear manifold
inside of the product space . Thus each tuple has the property that there exists some sequence of functions in such that
in the space , and simultaneously
in the space .
We could have used a set of tuples with in the definition of the Hardy subspace, and arrived at the same space. Indeed, we have the following proposition.
Proposition 3.1.
KernelContainmentHardySubspace With and as above, the Hardy subspace contains all tuples of the form , . Moreover, tuples where and extends analytically to a half-space , , are norm-dense in .
Proof.
Fix , and consider the functions defined by the formula
These functions are contained in for each , and they are analytic in a half-space larger than . Note that
and that
We readily see from \threfH1CharacterizationProp and the dominated convergence theorem that we have
By Plancherel’s theorem, we therefore also have
A fortiori, we have
and
Thus, as , the tuples converge in the norm of the space to the tuple , which is therefore contained in . This proves the first statement of the proposition. The second has the same proof, we merely start with and run the same argument. ∎
The shift operators
are unitary. Using the convention that for and , the translation operators
are contractions on , whenever . This fact is a consequence of the assumption that is decreasing:
We used that vanishes for . Consequently, the operators
are bounded on the space . Moreover, the Hardy subspace is invariant for these operators. Indeed, if , then by the well-known property of the Fourier transform , we obtain
The function is contained in , and so the above relation shows that a dense subset of is mapped into under each of the bounded operators . The mentioned invariance follows.
4. Strategy of the proofs
This section outlines the strategy of the proofs of \threfCondensationTheorem and \threfSparsenessTheorem.
4.1. Two easy computations
In our strategy, we will need to use the results of the following two computations.
Lemma 4.1.
FourierTransformCauchyKernel Let
The Fourier transform equals
and the Fourier transform of the conjugate of equals
Proof.
It is perhaps easiest to apply the Fourier inversion formula to the asserted formula for . We readily compute
The other formula follows from , which is an easily established property of the Fourier transform. ∎
Proposition 4.2.
ConvolutionFourierDecayLemma Assume that satisfies
for some , and let
Then
Proof.
Note that , and recall that the Fourier transform is thus a continuous function given by the convolution of the Fourier transforms of and . By \threfFourierTransformCauchyKernel, we obtain
and so
The exponential term in the last integral is bounded by . Therefore , and the desired estimate follows from the decay assumption on . ∎
4.2. Strategy of the proof of \threfCondensationTheorem
Given a function we consider the set
which is well-defined up to a set of Lebesgue measure zero. Let denote the family of all finite open intervals which satisfy
and set
Since on for every interval , it follows that if is integrable on , then the set difference must have measure zero. Consequently, since one can easily argue that we can express as a countable union of intervals on which is integrable, the Lebesgue measure of the set difference must be zero. However, the set difference might have positive measure. We set
and call this set the residual of . The residual is well-defined up to a set of Lebesgue measure zero.
Claim 1.
claim1 Under the assumption that for some , the set has Lebesgue measure zero.
CondensationTheorem follows immediately from the above claim. Indeed, the roles of and may obviously be interchanged in the statement of \threfCondensationTheorem, and the above claim implies that the open set equals up to an error of measure zero. Local integrability of on the set follows from its construction.
We set
Note that and that . Our \threfclaim1 will follow from the next assertion.
Claim 2.
claim2 Let be a bounded, continuous, non-negative and decreasing function which satisfies for some and . Then, every tuple of the form
where is any function in which lives only on the set , is contained in the Hardy subspace .
To prove \threfclaim1 from \threfclaim2 we will use a trick involving Plancherel’s theorem. We set , where is the constant appearing in \threfclaim1. Let be as in \threfclaim2, and
be as in \threfConvolutionFourierDecayLemma. We will show that
This implies, by the generality of , that is zero on the set . Since is non-zero everywhere on , in fact is zero on . Since , it follows that the residual has Lebesgue measure zero. Thus establishing the vanishing of the above integral is sufficient to prove \threfclaim1 from \threfclaim2. We do so next.
Because , there exists a sequence of functions such that in the norm of and in the norm of . Consider the quantities
We passed from domain of integration into , since vanishes outside of anyway (note also that almost everywhere on ). By the Cauchy-Schwarz inequality, we obtain
Note that the first of the factors on the right-hand side of the inequality above converges to . The other factor is finite. Indeed, since on the set where , and on the set where , we obtain , and consequently
This computation implies the formula
where we used Plancherel’s theorem in the last step. Now, recall that by \threfConvolutionFourierDecayLemma we may estimate
for some positive constant . By again using Cauchy-Schwarz inequality, we obtain
The second factor on the right-hand side is certainly finite. Since and in , the first factor above converges to , as . All in all, we have obtained that
By the earlier discussion, this is sufficient to establish \threfclaim1 from \threfclaim2. We need to prove \threfclaim2 in order to prove \threfCondensationTheorem. We will do so in the coming sections.
4.3. Strategy of the proof of \threfSparsenessTheorem
We will derive \threfSparsenessTheorem from the following claim.
Claim 3.
claim3 There exists a compact set of positive Lebesgue measure, and an increasing function which satisfies
(4.1) |
for every , such that if
then the Hardy subspace is properly contained in .
We proceed to show how one proves \threfSparsenessTheorem from this claim. Let and be as in \threfclaim3, and assume that the non-zero tuple is orthogonal to . We shall soon see that in fact is non-zero. The function lives only on the set , and we will show that it has the required spectral decay. Recall \threfFourierTransformCauchyKernel, let be as in that lemma, and let be the inverse Fourier transform of , so that . In fact , since its spectrum is positive. The orthogonality means that
We used Plancherel’s theorem. The above relation shows that the function is orthogonal in to each of the functions . Let be the orthogonal projection. In terms of Fourier transforms, we have
Then
is orthogonal not only to , but also to , since by \threfFourierTransformCauchyKernel the functions have spectrum contained in . But then the decomposition formula for the Poisson kernel in (2.1) shows that
for each , and it is an elementary fact about the Poisson kernel that we must, in this case, have . So . We can now argue that . Indeed, if , then . Since , that would mean , contradicting that the tuple is non-zero. Having established that , we proceed by taking Fourier transforms to obtain
Using Cauchy-Schwarz inequality, we may now estimate
The second factor is finite, and the growth of asserted in \threfclaim3 implies that for every fixed there exists a constant for which we have
It follows that the integral inside the square root of the first factor above satisfies
Since was arbitrary, we conclude that for every and . This easily implies \threfSparsenessTheorem.
It follows that \threfSparsenessTheorem is implied by \threfclaim3.
5. Proof of \threfCondensationTheorem
Our proof is an adaptation to the half-plane setting of the authors’ technique from [11], and in fact the two proofs are very similar. The problem studied in [11] is different, but in both the present work and in the reference, the main trick consists of constructing a highly oscillating sequence of functions which simultaneously obey appropriate spectral bounds.
5.1. A sufficient construction
We start by reducing our task to a construction of a certain sequence of bounded functions. Recall that and that has the decay for and some . Note that we may assume throughout that
Indeed, if on the contrary this integral converges, then is void, and both \threfclaim2 of Section 4 and \threfCondensationTheorem (with and playing opposite roles) hold trivially.
We may decompose as
where
(5.1) |
The set equality above holds up to an error of measure zero. The sets are bounded, and on each of them is bounded from below.
Proposition 5.1.
hnSplittingSequence In order to establish \threfclaim2, it suffices to construct, for any fixed and , a sequence of functions in which has the following properties.
-
(i)
The analytic extensions of the functions to satisfy the bound for ,
-
(ii)
for almost every ,
-
(iii)
for every ,
-
(iv)
there exists an and such that for almost every .
Proof.
Property , together with \threfFourierDecayFromExtensionGrowthCorollary, implies that the functions obey the spectral bound . Since for some , the spectral bound implies that
if is small enough. Together with , we see that forms a bounded subset of the Hardy subspace of product space . We may thus assume, by passing to a subsequence, that tends weakly in to some tuple . In fact, implies that is a sequence bounded in , so we have that for some . The fact that on is a consequence of the weak convergence of to and the condition , which implies for almost every . Moreover, by the formula in \threfH1FourierTransformFormula, we have
for any . The integrand converges pointwise to by , and it is dominated pointwise by the integrable function
The dominated convergence theorem implies that
where . In the next-to-last equality we used \threfH1FourierTransformFormula backwards. Weak and pointwise convergence implies, as previously, that . Since , we have that . The function is non-zero almost everywhere on . Indeed, vanishes on , and is non-zero everywhere on . Also, since , we have . The conditions to apply \threfinvSubspaceCorollary are thus satisfied, and by the invariance of under the operators defined in Section 3, we conclude that
where . Since is arbitrary, we conclude that
This is sufficient to conclude the validity of \threfclaim2. ∎
5.2. An estimate for Poisson integrals
Let be a finite real-valued measure on . The Poisson integral of is the harmonic function which is given by the formula
By the triangle inequality, and an estimation of by its supremum over , we easily obtain the inequality
and where denotes the usual variation of the measure . We obtain a much better inequality for measures which are oscillating rapidly. The following lemma is the half-plane version of an estimate in [11, Lemma 3.2].
Lemma 5.2.
PsnIntegralEstimate Let be a finite real-valued measure on which has the following structure: there exists a finite sequence of disjoint intervals of , and a decomposition , where is a real-valued measure supported inside , , and for some which is independent of . Then
Proof.
Since satisfies the same conditions as , and so does any translation of , it suffices to prove that for any . We have
If is the decomposition of into its positive and negative parts, then we have the estimate
In the second step we used that and that . Since the intervals are disjoint, and the function is increasing for , decreasing for , and attains a maximum value of at , the sum in the above estimate cannot be larger than (it is easily seen to be bounded by twice the height of the graph of ). The estimate follows. ∎
5.3. The construction
In accordance with the earlier discussion in Section 5.1, we recall the decomposition (5.1) and assume below that is a bounded subset of for some . The set inherits the following property from : if is an interval, and , then
Lemma 5.3.
pieceLemma If for some finite interval , then given any and any , there exists and a measurable subset disjoint from for which we have
(5.2) |
Proof.
On the set , is bounded from below by . Hence . Consequently,
So for sufficiently large we will have , and then by the absolute continuity of the finite positive measure we may choose a set for which (5.2) holds. ∎
We will now construct the sequence in \threfhnSplittingSequence. Let , be some sequence of positive numbers which tends to slowly enough so that tends to , and let be as above. Fix some integer , cover by a sequence of (say, half-open) disjoint intervals of length and let be those intervals for which . Apply \threfpieceLemma with to each of the intervals to obtain a corresponding constant and a set for which (5.2) holds. We set , where
Then is an absolutely continuous real-valued measure with bounded density , and . We construct by letting the logarithm of its extension to be given by the right-hand side of the formula (2.7). Then, by \threfPsnIntegralEstimate, we have the inequalities
and consequently, since , we have the bounds
(5.3) |
Since , by multiplying by an appropriate unimodular constant, we may assume by (5.3) that
(5.4) |
for every (possibly after passing to a subsequence). Also, for almost every , we have
(5.5) |
Our assumption on then implies that for almost every . For almost every , we have instead
(5.6) |
6. Proof of \threfSparsenessTheorem
6.1. A bit of concave analysis
Let be an increasing and concave function which is differentiable for . We assume that , and that
(6.1) |
For a function with the above properties, the integrals
(6.2) |
converge for every , and estimation of the growth of as will be of importance in the proof of \threfSparsenessTheorem. To estimate , we define a function in the following way. Since is increasing, concave and differentiable, the derivative is defined for , and it is a positive and decreasing function. The concavity of implies that
(6.3) |
from which it follows by (6.1) that
It is only the asymptotic behaviour of as that concerns us, so we will also assume for convenience that . In this case, the inverse function
(6.4) |
is well-defined and positive. It is decreasing, and satisfies . We set
(6.5) |
The function is decreasing, and satisfies . The integrals can be estimated in terms of .
Proposition 6.1.
IMestimate For as above, we have
for all sufficiently small .
Proof.
We will need the following observation. The inequality (6.3) together with (6.1) implies that if is sufficiently large. We want to show the inequality
for sufficiently small . Set and . Then
if is sufficiently small (and consequently is sufficiently large). Since is a decreasing function, the above inequality shows that , which is the same as the desired inequality.
We split the integral (6.2) at . Both pieces can be estimated very crudely. For the first piece, we have
We used our initial observation in the last step. For the second piece, we note that
Indeed, the supremum is attained at the point where , which by definition is . Thus
In the last inequality we require that . We obtain the desired estimate by combining the estimates for the two pieces of the integral. ∎
In the proof of \threfSparsenessTheorem, a point will come up where we will need integrability of near the origin. The next proposition describes the functions corresponding to which are integrable in this way.
Proposition 6.2.
MstarIntegrabilityProp For as above, the following two statements are equivalent.
-
(i)
for some .
-
(ii)
.
Proof.
Remark 6.3.
ConcFuncRemark We should point out that a concave function indeed exists which satisfies all of our conditions. For instance, note that any concave function which for large positive coincides with
satisfies the equivalent conditions of the above proposition. Indeed, one verifies by differentiation that the right-hand side is concave on some interval and that the integral in of \threfMstarIntegrabilityProp converges. Such a function can be easily chosen to satisfy all assumptions made in this section. Moreover, clearly it satisfies (4.1).
6.2. A growth estimate for Hardy class functions
Proposition 6.4.
growthestimateUpperHalfplane Let have a Fourier transform satisfying
for some constant and . Then there exist a positive constant such that the analytic extension of to satisfies the estimate
Proof.
By developments of Section 2.1, we have
and where is as in \threfFourierTransformCauchyKernel. By Plancherel’s theorem and the lemma, we obtain
An application of Cauchy-Schwarz inequality leads to
(6.6) |
Now \threfIMestimate applies to obtain the desired estimate. ∎
6.3. Construction of the compact set
If satisfies the equivalent conditions of \threfMstarIntegrabilityProp, then the logarithm of the right-hand side in the inequality of \threfgrowthestimateUpperHalfplane, namely
(6.7) |
is, for small enough , positive and integrable over the interval . To and any we will associate a Cantor-type compact set contained in which contains no intervals and for which the integral
(6.8) |
converges. Here denotes the distance from the point to the closed set . Let be the system of maximal disjoint open intervals, union of which constitutes the complement of within . The convergence of the integral above is easily seen to be equivalent to the convergence of the sum
(6.9) |
where denotes the length of the interval . To construct , we choose a sequence of numbers which is so quickly decreasing that
(6.10) |
and
(6.11) |
From such a sequence, we construct as in the classical Cantor set construction. We set , and recursively define a compact set contained in . The set consists of closed intervals in which we obtain by removing from the closed intervals constituting an open interval of length lying in the middle of . Thus splitting each into two new closed intervals. The above summation condition (6.11) ensures that , and so has positive Lebesgue measure which is not less than . The integral condition (6.8) holds by its equivalence to (6.9) and by (6.10). Clearly contains no intervals.
6.4. Collapse of the Fourier transforms
We are now ready to prove \threfclaim3. We will do so by showing that does not contain any non-zero tuple of the form , , where is as in \threfConcFuncRemark, for instance, and where is as in Section 6.9. We set .
Lemma 6.5.
FourierCollapseLemma Let , and be chosen as above. Assume that is a sequence of functions in , each of which has an analytic extension to a half-space larger than . If in the norm of and the sequence of Fourier transforms satisfies
then we have
The convergence is uniform on compact subsets of .
In the proof of \threfFourierCollapseLemma given below we will use a technique of Khrushchev from [8] for estimating harmonic measures on certain domains. For general background on the theory of harmonic measures, see [5] or [14].
Let be the collection of finite open intervals complementary to , and let
be a triangle with base at . We define to be the bounded domain in which consists of a rectangle , with a base being the shortest closed interval containing the set , with the triangles removed from . See Figure 1. An observation that Khrushchev made regarding this type of domains is the following property of their harmonic measure.
Lemma 6.6.
KhruschevsEstimateLemma Let , and be chosen as above, and let be given by (6.7). Let be the domain described above. If is the harmonic measure of the domain at any point , then
We emphasize that equals .
Proof.
The proof is very similar to the one given by Khrushchev in [8], only minor details differ. If is one of the finite intervals complementary to , and is the triangle standing on top of it, then we denote by the part of the boundary of which lies above the interval , . If is the harmonic measure in of the interval , then it is easy to see from the explicit formula
that for . Since is harmonic and continuous in the closure of except possibly at the two points and , the reproducing formula holds, and so
We have used the positivity of and in the first inequality, and the second one is an easy consequence of the explicit formula for above. Set , which is the left side of the boundary of , and further set , . Then, since is decreasing,
Now the desired claim follows from (6.9). ∎
[scale=1]OmegaDomain
Proof of \threfFourierCollapseLemma .
Note that it is sufficient to establish the claim that the sequence contains a subsequence which converges pointwise in to . Indeed, the proof of \threfgrowthestimateUpperHalfplane shows that our assumption on the Fourier transforms implies pointwise boundedness of the sequence on each half-plane , . Hence the sequence forms a normal family on . If we establish the above claim, then every subsequences of contains a further subsequence convergent to in . This is equivalent to convergence of the entire initial sequence to .
Fix . Since is a subharmonic function and is a bounded continuous function on , we obtain by the maximum principle for subharmonic functions that
We let and, by the monotone convergence theorem, obtain
The assumption in the lemma, \threfgrowthestimateUpperHalfplane, the definition of in (6.7) and \threfKhruschevsEstimateLemma show that
where is some positive constant which is independent of . By Egorov’s theorem, we may pass to a subsequence (the same subsequence for each and assume that converge uniformly to on some subset of which is of positive Lebesgue measure. On we have the estimate
The last inequality follows from monotonicity of the harmonic measure with respect to domains (see [14, Corollary 4.3.9]), which applied to leads to the inequality
for any Borel subset of . Thus , attesting the integral inequality above. By the convergence of to in the norm of , the integrals are uniformly bounded by some constant , and so the above inequalities give
But , since the harmonic measure and the arc-length measure on the rectifiable curve are mutualy absolutely continuous (see [5, Theorem 1.2 of Chapter VI]), so since converge uniformly to on , the integral on the right-hand side above converges to as . Thus , and since was arbitrary, the desired claim follows. ∎
The following proposition implies \threfclaim3 of Section 4, and so it also implies \threfSparsenessTheorem.
Proposition 6.7.
FourierCollapseProposition Let , and be chosen as above. If a tuple of the form is contained in the Hardy subspace , then .
Proof.
By the containment and \threfKernelContainmentHardySubspace, there exists a sequence of functions in extending analytically across and for which the tuples converge in the norm of the product space to . By passing to a subsequence, we may assume that the Fourier transforms converge pointwise almost everywhere on to . One might attempt to prove the proposition by using the formula in \threfH1FourierTransformFormula, and observing that
holds for almost every and for any . By \threfFourierCollapseLemma the integrand converges pointwise to . However, an appeal to the usual convergence theorems for integrals is not justified, and we have to proceed more carefully. Note that since and is bounded from below on compact subsets of , in fact is locally integrable on . It follows that we can interpret as a distribution on . Thus to show that , it suffices to establish that for every smooth function which is compactly supported in .
Let be as above. Since we have that in and is bounded from below on compact subsets of , we obtain
(6.12) |
Fix some small . By \threfH1FourierTransformFormula, we get that
Plugging this formula into (6.12) and noting that the use of Fubini’s theorem is permitted, we obtain
(6.13) |
where
is the Fourier transform of the compactly supported smooth function . As such, is certainly integrable on . By \threfFourierCollapseLemma, we have , and
holds by \threfgrowthestimateUpperHalfplane. Therefore, this time, the dominated convergence theorem applies to (6.13), and we conclude that
Thus is the zero distribution on , and therefore . ∎
7. Clumping for tempered distributions
In this last section, we indicate how one can derive \threfDistributionalClumpingTheorem from \threfCondensationTheorem. We will skip most of the details of the necessary computations, which are in any case standard.
Let be a function which satisfies
(7.1) |
for some positive integer . Then can be interpreted as a tempered distribution on in the usual way, and so has a distributional Fourier transform . Our hypothesis is that is an integrable function on some half-axis and that
(7.2) |
We may assume that . In order to prove \threfDistributionalClumpingTheorem, we will construct an appropriate multiplier with the property that is a function to which \threfCondensationTheorem applies. In particular, the following properties will be satisfied by :
-
(i)
is a bounded function of which is non-zero for almost every .
-
(ii)
,
-
(iii)
for some and ,
-
(iv)
for every interval .
If we construct such a multiplier , then , and \threfCondensationTheorem imply that is locally integrable on an open set which coincides, up to a set of measure zero, with . By , differs from at most by a set of measure zero. Moreover, the formula and show that is locally integrable on . This proves \threfDistributionalClumpingTheorem, as a consequence of existence of a multiplier satisfying the above conditions. We now show how to construct such a multiplier.
We set
and let be defined by the equation (2.7), with
The condition (7.1) ensures that is well-defined, and it is a member of . We put
and finally
Clearly, is bounded. Since is locally integrable on , the set has measure zero. Consequently, almost everywhere in , and so the desired property of holds. The choice of and ensures that is both bounded and integrable on , implying , so that above holds. Property holds by \threfHadyClassLogIntProp, since . So the critical property left to be verified is the spectral estimate of in above.
Lemma 7.1.
With notation and definitions as above, the Fourier transform satisfies
for some .
Proof.
A standard argument shows that must coincide on with the convolution (which, note, is a function on ). Indeed, let be a Schwartz function which has a Fourier transform supported on some compact interval , . Note that the function is also of Schwartz class. It follows immediately from the integral definition of the Fourier transform (1.1) that , and that is supported on the interval . Hence, by the definition of the distributional Fourier transform, we obtain
Fubini’s theorem and the computational rule shows that the last integral above equals
proving our claim about the structure of on .
Hence is a bounded continuous function which coincides with
for . By a computation similar to the one in the proof of \threfFourierTransformCauchyKernel one sees that has the Fourier transform
For such , we estimate
We now make the rather rough estimate
which gives
for some . The above sum can be readily estimated to be of order for some slightly smaller than . Since is the product of two integrable functions and , we have
where is non-zero only for . Note that is integrable on , and so is bounded. We obtain
where is some positive constant. The desired estimate on follows readily from this estimate. ∎
By the above discussion, we have proved \threfDistributionalClumpingTheorem.
References
- [1] W. O. Amrein and A. M. Berthier. On support properties of -functions and their Fourier transforms. Journal of Functional Analysis, 24(3):258–267, 1977.
- [2] M. Benedicks. Fourier transforms of functions supported on sets of finite Lebesgue measure. Journal of Mathematical Analysis and Applications, 106(1):180–183, 1985.
- [3] A. Borichev and A. Volberg. Uniqueness theorems for almost analytic functions. Leningrad Mathematical Journal, 1:157–190, 1990.
- [4] J. Garnett. Bounded analytic functions, volume 236. Springer Science & Business Media, 2007.
- [5] J. Garnett and D. Marshall. Harmonic measure, volume 2. Cambridge University Press, 2005.
- [6] V. P. Havin and B. Jöricke. The uncertainty principle in harmonic analysis, volume 72 of Encyclopaedia Math. Sci. Springer, Berlin, 1995.
- [7] H. Hedenmalm and A. Montes-Rodríguez. Heisenberg uniqueness pairs and the Klein-Gordon equation. Annals of Mathematics, pages 1507–1527, 2011.
- [8] 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.
- [9] 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.
- [10] A. Kulikov, F. Nazarov, and M. Sodin. Fourier uniqueness and non-uniqueness pairs. arXiv preprint arXiv:2306.14013, 2023.
- [11] B. Malman. Revisiting mean-square approximation by polynomials in the unit disk. arXiv preprint arXiv:2304.01400, 2023.
- [12] B. Malman. Shift operators, Cauchy integrals and approximations. arXiv preprint arXiv:2308.06495, 2023.
- [13] D. Radchenko and M. Viazovska. Fourier interpolation on the real line. Publications mathématiques de l’IHÉS, 129:51–81, 2019.
- [14] T. Ransford. Potential theory in the complex plane. Number 28 in London Mathematical Society Student Texts. Cambridge university press, 1995.
- [15] A. Volberg. The logarithm of an almost analytic function is summable. In Doklady Akademii Nauk, volume 265, pages 1297–1302. Russian Academy of Sciences, 1982.
- [16] 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.