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

Spectral clumping for functions and distributions decreasing rapidly on a half-line

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

We demonstrate a phenomenon of condensation of the Fourier transform f^\widehat{f} of a function ff defined on the real line \mathbb{R} which decreases rapidly on one half of the line. For instance, we prove that if ff is square-integrable on \mathbb{R}, then a one-sided estimate of the form

ρf(x):=x|f(t)|𝑑t=𝒪(ecx),x>0\rho_{f}(x):=\int_{x}^{\infty}|f(t)|\,dt=\mathcal{O}\big{(}e^{-c\sqrt{x}}\big{)},\quad x>0

for some c>0c>0, forces the non-zero frequencies σ(f):={ζ:|f^(ζ)|>0}\sigma(f):=\{\zeta\in\mathbb{R}:|\widehat{f}(\zeta)|>0\} to clump: this set differs from an open set UU only by a set of Lebesgue measure zero, and log|f^|\log|\widehat{f}| is locally integrable on UU. In particular, if ff is non-zero, then there exists an interval on which log|f^|\log|\widehat{f}| is integrable. The roles of ff and f^\widehat{f} 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 ff exists which is square-integrable on \mathbb{R}, for which σ(f)\sigma(f) is a subset of a compact set EE containing no intervals, and for which the estimate ρf(x)=𝒪(exa)\rho_{f}(x)=\mathcal{O}\big{(}e^{-x^{a}}\big{)}, x>0x>0, holds for every a(0,1/2)a\in(0,1/2).

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 ff forces its Fourier transform f^\widehat{f} to be, in some sense, large. Vice versa, smallness of f^\widehat{f} forces ff 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) f^(ζ):=f(x)eixζ𝑑λ(x),ζ.\widehat{f}(\zeta):=\int_{\mathbb{R}}f(x)e^{-ix\zeta}\,d\lambda(x),\quad\zeta\in\mathbb{R}.

Here dλ(x)=dx/2πd\lambda(x)=dx/\sqrt{2\pi} is a normalization of the Lebesgue measure dxdx on \mathbb{R}. Then, the inversion formula is given by

(1.2) f(x):=f^(ζ)eiζx𝑑λ(ζ),x.f(x):=\int_{\mathbb{R}}\widehat{f}(\zeta)e^{i\zeta x}\,d\lambda(\zeta),\quad x\in\mathbb{R}.

For p>0p>0, let p(,dx)\mathcal{L}^{p}(\mathbb{R},dx) be the usual Lebesgue space of functions ff for which |f|p|f|^{p} is integrable with respect to dxdx. The formula (1.1) can be interpreted literally only for f1(,dx)f\in\mathcal{L}^{1}(\mathbb{R},dx). It is interpreted in terms of Plancherel’s theorem in the case f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx), 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 σ(f)\sigma(f) of a function ff is the subset of \mathbb{R} on which f^\widehat{f} lives. Since f^2(,dx)\widehat{f}\in\mathcal{L}^{2}(\mathbb{R},dx) is defined only up to a set of Lebesgue measure zero, so is the spectrum σ(f)\sigma(f) in this case. If we accept making errors of measure zero (which we will), we may define the spectrum as

σ(f):={ζ:|f^(ζ)|>0},f1(,dx)2(,dx).\sigma(f):=\{\zeta\in\mathbb{R}:|\widehat{f}(\zeta)|>0\},\quad f\in\mathcal{L}^{1}(\mathbb{R},dx)\cup\mathcal{L}^{2}(\mathbb{R},dx).

Note specifically that our definition of σ(f)\sigma(f) might not coincide with the usual notion of closed support of the distribution f^\widehat{f}.

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 f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) is non-zero and \mathbb{R}_{-} is the negative half-axis, then we have the implication

(1.3) f(x)0 on log|f^(ζ)|1+ζ2𝑑ζ>.f(x)\equiv 0\text{ on }\mathbb{R}_{-}\quad\Rightarrow\quad\int_{\mathbb{R}}\frac{\log|\widehat{f}(\zeta)|}{1+\zeta^{2}}\,d\zeta>-\infty.

Here the extreme decay (indeed, vanishing) of ff on a half-axis implies global integrability of log|f^|\log|\widehat{f}| against the Poisson measure dζ/(1+ζ2)d\zeta/(1+\zeta^{2}). A fortiori, log|f^|\log|\widehat{f}| is integrable on every interval II of \mathbb{R}. Naturally, this is not typical. By Plancherel’s theorem, every function in 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx) is the Fourier transform of some other function in the same space. So on the other extreme, plenty of functions f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) have a Fourier transform which lives on sparse sets containing no intervals. This forces the divergence of the logarithmic integral of f^\widehat{f} 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 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx). Here we can prove a claim which symmetric in ff and f^\widehat{f}, 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.
\thlabel

CondensationTheorem If f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) satisfies the estimate

(1.4) ρf(x):=x|f(t)|𝑑t=𝒪(ecx),x>0\rho_{f}(x):=\int_{x}^{\infty}|f(t)|\,dt=\mathcal{O}\big{(}e^{-c\sqrt{x}}\big{)},\quad x>0

for some constant c>0c>0, then there exists an open set UU which coincides with σ(f)\sigma(f) up to a set of Lebesgue measure zero, and for every xUx\in U there exists an interval II containing xx such that

Ilog|f^(t)|dt>.\int_{I}\log|\widehat{f}(t)|\,dt>-\infty.

In other words, the one-sided decay condition (1.4) implies that f^\widehat{f} lives on the union of the spectral clumps of ff. Since the Fourier transform is a unitary operation on 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx), the roles of ff and f^\widehat{f} may obviously be interchanged in the statement of \threfCondensationTheorem. Thus a one-sided spectral decay condition of ff implies local integrability properties of log|f|\log|f| on the set where ff 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 ff. 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 x|f(t)|2𝑑t=𝒪(ecx)\int_{x}^{\infty}|f(t)|^{2}\,dt=\mathcal{O}\big{(}e^{-c\sqrt{x}}\big{)} in fact implies (1.4) for some slightly smaller cc.

We can prove also that the condition (1.4) on the decay of ρf\rho_{f} 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.
\thlabel

SparsenessTheorem For every b>0b>0, there exists a compact set EE\subset\mathbb{R} contained in [0,b][0,b] which contains no intervals, and a non-zero entire function f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) which satisfies

ρf(x)=𝒪(exa),x>0\rho_{f}(x)=\mathcal{O}\big{(}e^{-x^{a}}\big{)},\quad x>0

for every a(0,1/2)a\in(0,1/2), and such that σ(f)\sigma(f) is contained within EE.

After an initial reduction, the proof of this result follows ideas of Khrushchev from [8]. Note that the function ff appearing in \threfSparsenessTheorem is entire by the virtue of having a spectrum σ(f)\sigma(f) of compact support. More importantly, the condition on EE implies that IEI\setminus E has positive Lebesgue measure for every interval II, so we obtain

Ilog|f^(t)|dt=\int_{I}\log|\widehat{f}(t)|\,dt=-\infty

for every interval II\subset\mathbb{R}. This is in contrast to the conclusion of \threfCondensationTheorem. It follows that the exponent a=1/2a=1/2 in estimates of the form ρf(x)=𝒪(exa)\rho_{f}(x)=\mathcal{O}\big{(}e^{-x^{a}}\big{)} is critical for the spectral clumping phenomenon.

As mentioned above, clumping statements makes sense for objects in a class much wider than 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx). Here is our distrbutional result.

Theorem C.
\thlabel

DistributionalClumpingTheorem Let ff be a tempered distribution on \mathbb{R} which is a measurable function satisfying

|f(x)|(1+|x|)n𝑑x<\int_{\mathbb{R}}\frac{|f(x)|}{(1+|x|)^{n}}\,dx<\infty

for some n>0n>0. If the distributional Fourier transform f^\widehat{f} is an integrable function on some interval [A,)[A,\infty), and the estimate ρf^(ζ)=𝒪(ecζ)\rho_{\widehat{f}}(\zeta)=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)} holds for all sufficiently large positive ζ\zeta, then there exists an open set UU such that ff vanishes almost everywhere outside of UU, and for each xUx\in U there exists an interval II containing xx satisfying

(1.5) Ilog|f(t)|dt>.\int_{I}\log|f(t)|\,dt>-\infty.

For instance, the result shows that a function f1(,dx)f\in\mathcal{L}^{1}(\mathbb{R},dx) which lives on a sparse set containing no intervals cannot satisfy even a one-sided spectral decay condition of the form ρf^(ζ)ecζ\rho_{\widehat{f}}(\zeta)\lesssim e^{-c\sqrt{\zeta}}. Note also that in this extended form, our result includes the trivial but important examples such as f=1f=1 and f^=δ0\widehat{f}=\delta_{0} (Dirac delta), the trigonometric functions and the polynomials.

1.3. A converse result

The Beurling-Malliavin theory implies a partial converse result. If ff is a locally integrable function on \mathbb{R} which has a clump II as in (1.5), and a constant c>0c>0 is given, then a bounded multiplier mm exists for which mfmf has a Fourier transform satisfying ρmf^(ζ)=𝒪(ecζ)\rho_{\widehat{mf}}(\zeta)=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)} for ζ>0\zeta>0. To see this, recall that a smooth function gg supported in II exists which satisfies the bilateral spectral decay |g^(ζ)|ec|ζ||\widehat{g}(\zeta)|\leq e^{-c\sqrt{|\zeta|}}, ζ\zeta\in\mathbb{R} (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 gg). There exists also a bounded function h1(,dx)h\in\mathcal{L}^{1}(\mathbb{R},dx) which satisfies σ(h)(0,)\sigma(h)\subset(0,\infty) and |h(x)|=min(|f(x)|,1)|h(x)|=\min(|f(x)|,1) on II (we use the assumption that II is a clump for ff and construct hh as in (2.7) below). Then an argument similar to the one used in the proof of \threfConvolutionFourierDecayLemma below shows that the function h¯g\overline{h}g will satisfy the desired one-sided spectral decay, and clearly h¯g=mf\overline{h}g=mf for some bounded function mm supported in II.

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 𝔻:={z:|z|<1}\mathbb{D}:=\{z\in\mathbb{C}:|z|<1\}. Here we are presented with a measure

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

where dAdA and dmd\textit{m} are the area and arc-length measures on 𝔻\mathbb{D} and 𝕋:=𝔻={z:|z|=1}\mathbb{T}:=\partial\mathbb{D}=\{z\in\mathbb{C}:|z|=1\}. The functions GG and ww are non-negative weights, and one would like to understand under which conditions splitting occurs. Namely, when is the weighted space 2(𝕋,wdm)\mathcal{L}^{2}(\mathbb{T},w\,d\textit{m}) contained in the closure of analytic polynomials in the 2\mathcal{L}^{2}-norm induced by the measure μ\mu? In the case that G(1|z|)G(1-|z|) decays exponentially as |z|1|z|\to 1^{-}, the necessary and sufficient condition is that ww has no clumps, or in other words that the integral of logw\log w diverges over any arc on 𝕋\mathbb{T}. 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 (b)\mathcal{H}(b) defined by de Branges and Rovnyak, well-known to operator theorists. To any positive finite Borel measure ν\nu on 𝕋\mathbb{T} we may associate a so-called Clark operator 𝒞ν\mathcal{C}_{\nu} which takes a function g2(𝕋,dν)g\in\mathcal{L}^{2}(\mathbb{T},d\nu) to the analytic function in 𝔻\mathbb{D} given by the formula

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

The operator 𝒞ν\mathcal{C}_{\nu} maps 2(𝕋,dν)\mathcal{L}^{2}(\mathbb{T},d\nu) onto a space of analytic functions denoted by (b)\mathcal{H}(b), the symbol function b:𝔻𝔻b:\mathbb{D}\to\mathbb{D} itself being related to ν\nu by the formula

11b(z)=𝕋11x¯z𝑑ν(x),z𝔻\frac{1}{1-b(z)}=\int_{\mathbb{T}}\frac{1}{1-\overline{x}z}d\nu(x),\quad z\in\mathbb{D}

in the case that ν\nu is a probability measure, with a similar formula in the general case. For many choices of ν\nu (or equivalently, choices of bb), the space (b)\mathcal{H}(b) is somewhat mysterious, with the distinctive feature of containing very few functions extending analytically to a disk larger than 𝔻\mathbb{D}. 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 ν\nu is decisive for existence and density of functions in (b)\mathcal{H}(b) 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 ff defined on the circle 𝕋:={z:|z|=1}\mathbb{T}:=\{z\in\mathbb{C}:|z|=1\} is integrable with respect to arc-length dsds on 𝕋\mathbb{T}, and the negative portion of the Fourier series of ff vanishes, then 𝕋log|f|ds>\int_{\mathbb{T}}\log|f|ds>-\infty, unless ff 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 f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) prohibits f^\widehat{f} from living on a set SS containing no intervals. Somewhat related are uniqueness statements in which one seeks to give examples of pairs of sets (E,S)(E,S) for which the following implication is valid: if ff in a certain class lives on EE and f^\widehat{f} lives on SS, then f0f\equiv 0. One says that (E,S)(E,S) is then a uniqueness pair for the corresponding class. A famous result of Benedicks presented in [2] (see also [1]) says that (E,S)(E,S) is a uniqueness pair for integrable ff if both sets have finite Lebesgue measure, and the result holds not only for the real line \mathbb{R} but also for the dd-dimensional Euclidean space d\mathbb{R}^{d}. Hedenmalm and Montes-Rodríguez worked with the hyperbola H={(x,y)2:xy=1}H=\{(x,y)\in\mathbb{R}^{2}:xy=1\} and th class of finite Borel measures μ\mu supported on HH which are absolutely continuous with respect to arclength on HH. They proved in [7] that if μ^\widehat{\mu} vanishes on certain types of discrete sets Λ2\Lambda\subset\mathbb{R}^{2}, then μ0\mu\equiv 0, thus exhibiting interesting uniqueness pairs of the form (H,2Λ)(H,\mathbb{\mathbb{R}}^{2}\setminus\Lambda). Recent work of Radchenko and Viazovska on interpolation formulas for Schwartz functions in [13] gives examples of pairs of discrete subsets EE and SS of \mathbb{R} for which (E,S)(\mathbb{R}\setminus E,\mathbb{R}\setminus S) 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 EE\subset\mathbb{R} and a measure μ\mu defined on \mathbb{R}, the space p(E,dμ)\mathcal{L}^{p}(E,d\mu) denotes the usual Lebesgue space consisting of equivalence classes of functions living only on EE and satisfying the integrability condition E|f(x)|p𝑑μ(x)<\int_{E}|f(x)|^{p}d\mu(x)<\infty. The containment p(E,dμ)p(,dμ)\mathcal{L}^{p}(E,d\mu)\subset\mathcal{L}^{p}(\mathbb{R},d\mu) is interpreted in the natural way. The symbols such as dxdx, dtdt and dζd\zeta denote the usual Lebesgue measure of the real line, while dλ=dx/2πd\lambda=dx/\sqrt{2\pi} will be the normalized version used in formulas involving Fourier transforms. If EE is a subset of \mathbb{R}, then |E||E| denotes its usual Lebesgue measure. The positive half-axis of \mathbb{R} is denoted by +:={x:x0}\mathbb{R}_{+}:=\{x\in\mathbb{R}:x\geq 0\}, and we set also :=+\mathbb{R}_{-}:=\mathbb{R}\setminus\mathbb{R}_{+}. The notions of almost everywhere and of measure zero are always to be interpreted in the sense of Lebesgue measure on \mathbb{R}. The indicator function of a measurable set EE is denoted by 𝟙E\mathbbm{1}_{E}. Finally, we put log+(x):=max(log(x),0)\log^{+}(x):=\max(\log(x),0).

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 1(),2()\mathcal{H}^{1}(\mathbb{R}),\mathcal{H}^{2}(\mathbb{R}) and ()\mathcal{H}^{\infty}(\mathbb{R}) which will be important in the coming sections. We discuss also properties of the shift operators f(t)eitsf(t)f(t)\mapsto e^{its}f(t) on weighted spaces on the real line, and their invariant subspaces.

2.1. Hardy classes

For pp equal to 11 or 22, we denote by p()\mathcal{H}^{p}(\mathbb{R}) the subspace of p(,dx)\mathcal{L}^{p}(\mathbb{R},dx) consisting of those functions ff for which the Fourier transform f^\widehat{f} vanishes on the negative part of the real axis:

p():={fp(,dx):f^|0}.\mathcal{H}^{p}(\mathbb{R}):=\{f\in\mathcal{L}^{p}(\mathbb{R},dx):\widehat{f}|\mathbb{R}_{-}\equiv 0\}.

It is a well-known fact that functions in the Hardy classes 1()\mathcal{H}^{1}(\mathbb{R}) and 2()\mathcal{H}^{2}(\mathbb{R}) admit a type of analytic extension to the upper half-plane

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

We recall what exactly is meant by this extension and how it can be constructed. The Poisson kernel of the upper half-plane

𝒫(t,x+iy):=1πy(xt)2+y2,y>0,\mathcal{P}(t,x+iy):=\frac{1}{\pi}\frac{y}{(x-t)^{2}+y^{2}},\quad y>0,

admits a decomposition

(2.1) 𝒫(t,z)=Re(1πi(tz))=12πi(1tz1tz¯),z=x+iy.\mathcal{P}(t,z)=\operatorname{Re}\Bigg{(}\frac{1}{\pi i(t-z)}\Bigg{)}=\frac{1}{2\pi i}\Bigg{(}\frac{1}{t-z}-\frac{1}{t-\overline{z}}\Bigg{)},\quad z=x+iy\in\mathbb{H}.

Since

(2.2) 1tz¯=i0eiz¯seits𝑑s\frac{1}{t-\overline{z}}=-i\int_{0}^{\infty}e^{-i\overline{z}s}e^{its}\,ds

we may use Fubini’s theorem to compute, in the case f1()f\in\mathcal{H}^{1}(\mathbb{R}), that

(2.3) f(t)tz¯𝑑t=i0(f(t)eits𝑑t)eiz¯s𝑑s=0,\int_{\mathbb{R}}\frac{f(t)}{t-\overline{z}}\,dt=-i\int_{0}^{\infty}\Bigg{(}\int_{\mathbb{R}}f(t)e^{its}\,dt\Bigg{)}e^{-i\overline{z}s}ds=0,

where the vanishing of the integral follows from

f(t)eits𝑑s=f^(s)=0,s>0,\int_{\mathbb{R}}f(t)e^{its}\,ds=\widehat{f}(-s)=0,\quad s>0,

which holds for any f1()f\in\mathcal{H}^{1}(\mathbb{R}) by the definition of the class. In the case f2()f\in\mathcal{H}^{2}(\mathbb{R}) 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 fp()f\in\mathcal{H}^{p}(\mathbb{R}) for p=1,2p=1,2, the formula

f(z):=f(t)𝒫(t,z)𝑑t=f(t)tzdt2πi,zf(z):=\int_{\mathbb{R}}f(t)\mathcal{P}(t,z)\,dt=\int_{\mathbb{R}}\frac{f(t)}{t-z}\frac{dt}{2\pi i},\quad z\in\mathbb{H}

defines, by the second integral expression above, an analytic extension of ff to \mathbb{H}. By the first expression, and classical properties of the Poisson kernel (see [4, Chapter I]), this extension satisfies

(2.4) limy0+f(x+iy)=f(x), for almost every x.\lim_{y\to 0^{+}}f(x+iy)=f(x),\text{ for almost every }x\in\mathbb{R}.

Moreover, we have

(2.5) supy>0|f(x+iy)|p𝑑x<\sup_{y>0}\int_{\mathbb{R}}|f(x+iy)|^{p}\,dx<\infty

and

(2.6) limy0+|f(x+iy)f(x)|p𝑑x=0.\lim_{y\to 0^{+}}\int_{\mathbb{R}}|f(x+iy)-f(x)|^{p}\,dx=0.

if fp()f\in\mathcal{H}^{p}(\mathbb{R}). The property (2.5) follows readily from the Poisson integral formula for the extension of ff 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.
\thlabel

H1CharacterizationProp For p=1p=1 and p=2p=2, a function fp(,dx)f\in\mathcal{L}^{p}(\mathbb{R},dx) is a member of p()\mathcal{H}^{p}(\mathbb{R}) if and only if there exists an analytic extension of ff to \mathbb{H} which satisfies the three properties in (2.4), (2.5) and (2.6).

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 |f||f| of a function f1()f\in\mathcal{H}^{1}(\mathbb{R}) will be of crucial importance to us.

Proposition 2.2.
\thlabel

HadyClassLogIntProp If f1()f\in\mathcal{H}^{1}(\mathbb{R}), then

log|f(x)|1+x2𝑑x>\int_{\mathbb{R}}\frac{\log|f(x)|}{1+x^{2}}dx>-\infty

unless ff 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 \mathbb{R}, and not necessarily integrable or square-integrable on \mathbb{R}. We use directly the complex interpretation of the class. Namely, we define ()\mathcal{H}^{\infty}(\mathbb{R}) to consist of those functions f(,dx)f\in\mathcal{L}^{\infty}(\mathbb{R},dx) which can be realized as limits

limy0+f(x+iy):=f(x)\lim_{y\to 0^{+}}f(x+iy):=f(x)

for almost every xx\in\mathbb{R}, where ff is bounded and analytic in \mathbb{H}. It can be checked that such ff has a distributional spectrum which vanishes on \mathbb{R}_{-}. Another important point is that if f()f\in\mathcal{H}^{\infty}(\mathbb{R}), then

f(x)(i+x)21(),\frac{f(x)}{(i+x)^{2}}\in\mathcal{H}^{1}(\mathbb{R}),

since we may apply \threfH1CharacterizationProp to the analytic function

zf(z)(i+z)2,z.z\mapsto\frac{f(z)}{(i+z)^{2}},\quad z\in\mathbb{H}.

A function h()h\in\mathcal{H}^{\infty}(\mathbb{R}) of a given (bounded, measurable) modulus |h|=W|h|=W on \mathbb{R} may be constructed by setting

(2.7) logh(z):=1πi(1tzt1+t2)logW(t)𝑑t,z,\log h(z):=\frac{1}{\pi i}\int_{\mathbb{R}}\Big{(}\frac{1}{t-z}-\frac{t}{1+t^{2}}\Big{)}\log W(t)\,dt,\quad z\in\mathbb{H},

and h(z):=elogh(z)h(z):=e^{\log h(z)}. The integral above converges if

logW(t)1+t2𝑑t>\int_{\mathbb{R}}\frac{\log W(t)}{1+t^{2}}\,dt>-\infty

which is a necessary condition for the construction to be possible. Then

log|h(z)|=𝒫(t,z)logW(t)𝑑t,\log|h(z)|=\int\mathcal{P}(t,z)\log W(t)\,dt,

so that the equality limy0+|h(x+iy)|=|h(x)|=W(x)\lim_{y\to 0^{+}}|h(x+iy)|=|h(x)|=W(x) for almost every xx\in\mathbb{R} 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 f1()f\in\mathcal{H}^{1}(\mathbb{R}), then the values of f^(ζ)\widehat{f}(\zeta) may be computed using a formula different from (1.1). To wit, denote by f(z)f(z) the extension of ff to \mathbb{H} which was discussed in Section 2.1. The function

Gζ(z):=f(z)eizζ=f(z)eixζ+yζ,z=x+iyG_{\zeta}(z):=f(z)e^{-iz\zeta}=f(z)e^{-ix\zeta+y\zeta},\quad z=x+iy\in\mathbb{H}

is analytic in \mathbb{H}, and for this reason Cauchy’s integral theorem implies

(2.8) R(ϵ,y,a)Gζ(z)𝑑z=0,\int_{R(\epsilon,y,a)}G_{\zeta}(z)dz=0,

where dzdz denotes the complex line integral, ϵ,y,a\epsilon,y,a are all positive numbers, ϵ<y\epsilon<y, and R(ϵ,y,a)R(\epsilon,y,a) denotes the rectangular contour having as corners the four points with coordinates (a,ϵ)(-a,\epsilon), (a,ϵ)(a,\epsilon), (a,y)(a,y), (a,y)(-a,y), oriented counter-clockwise.

Fix y>0y>0 and let SyS_{y} denote the horizontal strip in \mathbb{H} consisting of all complex numbers with imaginary part between 0 and yy. Then it follows from Fubini’s theorem and (2.5) that

S|Gζ(z)|𝑑A(z)ey|ζ|0y|f(x+is)|𝑑s𝑑x<,\displaystyle\int_{S}|G_{\zeta}(z)|dA(z)\leq e^{y|\zeta|}\int_{\mathbb{R}}\int_{0}^{y}|f(x+is)|dsdx<\infty,

where dA(z)dA(z) denotes the area measure on the complex plane. This expresses the integrability on \mathbb{R} of the continuous function

x0y|Gζ(x+is)|𝑑s.x\mapsto\int_{0}^{y}|G_{\zeta}(x+is)|ds.

Hence there exists a positive sequence {an}n\{a_{n}\}_{n} which satisfies

limnan=+\lim_{n\to\infty}a_{n}=+\infty

and for which

limn0y|Gζ(an+is)|𝑑s+0y|Gζ(an+is)|𝑑s=0.\lim_{n\to\infty}\int_{0}^{y}|G_{\zeta}(a_{n}+is)|ds+\int_{0}^{y}|G_{\zeta}(-a_{n}+is)|ds=0.

This means that

0\displaystyle 0 =limnR(ϵ,y,an)Gζ(z)𝑑z\displaystyle=\lim_{n\to\infty}\int_{R(\epsilon,y,a_{n})}G_{\zeta}(z)\,dz
=Gζ(x+iy)𝑑x+Gζ(x+iϵ)𝑑x\displaystyle=-\int_{\mathbb{R}}G_{\zeta}(x+iy)dx+\int_{\mathbb{R}}G_{\zeta}(x+i\epsilon)dx

Moreover, equation (2.6) quite easily implies

limϵ0+Gζ(x+iϵ)𝑑x=f(x)eixζ𝑑x=2πf^(ζ).\lim_{\epsilon\to 0^{+}}\int_{\mathbb{R}}G_{\zeta}(x+i\epsilon)dx=\int_{\mathbb{R}}f(x)e^{-ix\zeta}dx=\sqrt{2\pi}\widehat{f}(\zeta).

We have proven the following formula by combining the above two expressions.

Proposition 2.3.
\thlabel

H1FourierTransformFormula For f1()f\in\mathcal{H}^{1}(\mathbb{R}) we may compute the Fourier transform f^(ζ)\widehat{f}(\zeta) using the formula

f^(ζ)=eyζf(x+iy)eixζ𝑑λ(x)\widehat{f}(\zeta)=e^{y\zeta}\int_{\mathbb{R}}f(x+iy)e^{-ix\zeta}\,d\lambda(x)

for any choice of y>0y>0, where f(x+iy)f(x+iy) denotes the values of the analytic extension of ff to \mathbb{H}.

This formula has the following simple corollary which will be of critical importance below.

Corollary 2.4.
\thlabel

FourierDecayFromExtensionGrowthCorollary If h()h\in\mathcal{H}^{\infty}(\mathbb{R}) has an analytic extension to \mathbb{H} which satisfies, for some constant c>0c>0, an estimate of the form

supx|h(x+iy)|ec/y, for all y>0,\sup_{x\in\mathbb{R}}\,|h(x+iy)|\leq e^{c/y},\quad\text{ for all }y>0,

then the Fourier transform h^\widehat{h_{*}} of the function

h(x):=h(x)(i+x)21()h_{*}(x):=\frac{h(x)}{(i+x)^{2}}\in\mathcal{H}^{1}(\mathbb{R})

satisfies

|h^(ζ)|π2e2cζ,ζ>0.|\widehat{h_{*}}(\zeta)|\leq\sqrt{\frac{\pi}{2}}e^{2\sqrt{c}\sqrt{\zeta}},\quad\zeta>0.
Proof.

It was mentioned in Section 2.1 that h1()h_{*}\in\mathcal{H}^{1}(\mathbb{R}). Therefore, we may use \threfH1FourierTransformFormula to estimate

|h^(ζ)|\displaystyle|\widehat{h_{*}}(\zeta)| eyζ|h(x+iy)||i+x+iy|2𝑑λ(x)\displaystyle\leq e^{y\zeta}\int_{\mathbb{R}}\frac{|h(x+iy)|}{|i+x+iy|^{2}}\,d\lambda(x)
eyζec/y1+x2𝑑λ(x)\displaystyle\leq e^{y\zeta}\int_{\mathbb{R}}\frac{e^{c/y}}{1+x^{2}}\,d\lambda(x)
=π2eyζ+c/y.\displaystyle=\sqrt{\frac{\pi}{2}}e^{y\zeta+c/y}.

Since y>0y>0 can be freely chosen, we may now set it to y=c/ζy=\sqrt{c/\zeta} to obtain the desired estimate. ∎

2.3. A semigroup of operators and its invariant subspaces

If w1(,dx)w\in\mathcal{L}^{1}(\mathbb{R},dx) and ss\in\mathbb{R}, the operator Us:2(,wdx)2(,wdx)U^{s}:\mathcal{L}^{2}(\mathbb{R},w\,dx)\to\mathcal{L}^{2}(\mathbb{R},w\,dx) given by

Usf(x):=eisxf(x)U^{s}f(x):=e^{isx}f(x)

is unitary on 2(,wdx)\mathcal{L}^{2}(\mathbb{R},w\,dx). We shall be interested in subspaces of 2(,wdx)\mathcal{L}^{2}(\mathbb{R},w\,dx) which are invariant for the operators in the semigroup {Us}s>0\{U^{s}\}_{s>0}. Given any element f2(,wdx)f\in\mathcal{L}^{2}(\mathbb{R},w\,dx), we denote by [f]w[f]_{w} the smallest closed linear subspace of 2(,wdx)\mathcal{L}^{2}(\mathbb{R},w\,dx) which contains ff and also all the functions UsfU^{s}f, s>0s>0.

Proposition 2.5.
\thlabel

UsInvSubspacesProp Let f2(,wdx)f\in\mathcal{L}^{2}(\mathbb{R},w\,dx) be a non-zero element which satisfies

log(|f(x)|2w(x))1+x2𝑑x=.\int_{\mathbb{R}}\frac{\log\big{(}|f(x)|^{2}w(x)\big{)}}{1+x^{2}}dx=-\infty.

Then the subspace [f]w[f]_{w} coincides with L2(E,wdx)L^{2}(E,w\,dx), where E={x:|f(x)|>0}E=\{x\in\mathbb{R}:|f(x)|>0\}.

Remark 2.6.

As usual, the set EE above is defined in a bit imprecise way. Since ff is, strictly speaking, merely a representative of an equivalence class of measurable functions in 2(,wdx)\mathcal{L}^{2}(\mathbb{R},w\,dx), the set EE 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 ff vanishes almost everywhere outside of the set EE, then so does UsfU^{s}f for any s>0s>0. Consequently, [f]w2(E,wdx)[f]_{w}\subset\mathcal{L}^{2}(E,w\,dx). Conversely, let us consider an element g2(E,wdx)g\in\mathcal{L}^{2}(E,w\,dx) with the property that

Usf(x)g(x)¯w(x)𝑑x=eisxf(x)g(x)¯w(x)𝑑x=0,s>0.\int_{\mathbb{R}}U_{s}f(x)\overline{g(x)}w(x)dx=\int_{\mathbb{R}}e^{isx}f(x)\overline{g(x)}w(x)dx=0,\quad s>0.

Setting h:=fg¯w1(,dx)h:=f\overline{g}w\in\mathcal{L}^{1}(\mathbb{R},dx), we note that the vanishing of the integrals above is equivalent to hh being a member of the Hardy class 1()\mathcal{H}^{1}(\mathbb{R}). We note also that

log|h(x)|1+x2𝑑x=12log(|f(x)|2w(x))1+x2𝑑x+12log(|g(x)|2w(x))1+x2𝑑x.\int_{\mathbb{R}}\frac{\log|h(x)|}{1+x^{2}}dx=\frac{1}{2}\int_{\mathbb{R}}\frac{\log\big{(}|f(x)|^{2}w(x)\big{)}}{1+x^{2}}dx+\frac{1}{2}\int_{\mathbb{R}}\frac{\log\big{(}|g(x)|^{2}w(x)\big{)}}{1+x^{2}}dx.

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 f,g2(,wdx)f,g\in\mathcal{L}^{2}(\mathbb{R},w\,dx). This implies that

log|h(x)|1+x2=.\int_{\mathbb{R}}\frac{\log|h(x)|}{1+x^{2}}=-\infty.
\thref

HadyClassLogIntProp now shows that h=fg¯wh=f\overline{g}w must be the zero function. Since |f(x)|w(x)>0|f(x)|w(x)>0 on EE and gg vanishes outside of EE, this means that g0g\equiv 0. So [f]w[f]_{w} is a closed and dense subspace of 2(E,wdx)\mathcal{L}^{2}(E,w\,dx), which means that the two spaces are equal. ∎

Corollary 2.7.
\thlabel

invSubspaceCorollary If f2(,wdx)f\in\mathcal{L}^{2}(\mathbb{R},w\,dx) is also a member of p(,wdx)\mathcal{L}^{p}(\mathbb{R},w\,dx) for some p>2p>2, and if

logw(x)1+x2𝑑x=,\int_{\mathbb{R}}\frac{\log w(x)}{1+x^{2}}\,dx=-\infty,

then [f]w[f]_{w} coincides with 2(E,wdx)\mathcal{L}^{2}(E,w\,dx), where E={x:|f(x)|>0}E=\{x\in\mathbb{R}:|f(x)|>0\}.

Proof.

To prove the corollary we need to verify the condition in \threfUsInvSubspacesProp. Note that, pointwise, we have

log(|f|2w)=(2/p)log(|f|pw)+(12/p)logw.\log\big{(}|f|^{2}w\big{)}=(2/p)\log\big{(}|f|^{p}w\big{)}+(1-2/p)\log w.

The coefficients 2/p2/p and 12/p1-2/p are positive. The inequality log(x)x\log(x)\leq x for x>0x>0 shows that

log(|f(x)|pw(x))1+x2𝑑x|f(x)|pw(x)𝑑x<+.\int_{\mathbb{R}}\frac{\log\big{(}|f(x)|^{p}w(x)\big{)}}{1+x^{2}}\,dx\leq\int_{\mathbb{R}}|f(x)|^{p}w(x)dx<+\infty.

Note that the integral on the left might very well be equal to -\infty, but that is of no concern to us: we conclude from the assumption, and the pointwise inequality above, that

log(|f(x)|2w(x))1+x2𝑑x=\int_{\mathbb{R}}\frac{\log\big{(}|f(x)|^{2}w(x)\big{)}}{1+x^{2}}\,dx=-\infty

and apply \threfUsInvSubspacesProp. ∎

3. A product space and its Hardy subspace

Let ρ:++\rho:\mathbb{R}_{+}\to\mathbb{R}_{+} be a bounded, continuous, non-negative and decreasing function, and w1(,dx)(,dx)w\in\mathcal{L}^{1}(\mathbb{R},dx)\cap\mathcal{L}^{\infty}(\mathbb{R},dx) be a non-negative function. We consider the product space 2(,wdx)2(+,ρdx)\mathcal{L}^{2}(\mathbb{R},w\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx). Inside of this space we embed the linear manifold 1()2()\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) in the following way:

Jf:=(f,f^)2(,wdx)2(+,ρdx),f1()2().Jf:=(f,\widehat{f})\in\mathcal{L}^{2}(\mathbb{R},w\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx),\quad f\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}).

The tuple JfJf is well-defined as an element of the product space, since both ff and f^\widehat{f} are members of 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx) and both ρ\rho and ww are bounded. We define the Hardy subspace (w,ρ)\mathcal{H}(w,\rho) as the norm-closure of the linear manifold

{Jf:f1()2()}\{Jf:f\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R})\}

inside of the product space 2(,wdx)2(+,ρdx)\mathcal{L}^{2}(\mathbb{R},w\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx). Thus each tuple (h,k)(w,ρ)(h,k)\in\mathcal{H}(w,\rho) has the property that there exists some sequence {fn}n\{f_{n}\}_{n} of functions in 1()2()\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) such that

h=limnfnh=\lim_{n\to\infty}f_{n}

in the space 2(,wdx)\mathcal{L}^{2}(\mathbb{R},w\,dx), and simultaneously

k=limnfn^k=\lim_{n\to\infty}\widehat{f_{n}}

in the space 2(+,ρdx)\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx).

We could have used a set of tuples JfJf with f2()f\in\mathcal{H}^{2}(\mathbb{R}) in the definition of the Hardy subspace, and arrived at the same space. Indeed, we have the following proposition.

Proposition 3.1.
\thlabel

KernelContainmentHardySubspace With ww and ρ\rho as above, the Hardy subspace (w,ρ)\mathcal{H}(w,\rho) contains all tuples of the form (f,f^)(f,\widehat{f}), f2()f\in\mathcal{H}^{2}(\mathbb{R}). Moreover, tuples JfJf where f1()2()f\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) and ff extends analytically to a half-space {z=x+iy:y>δ}\{z=x+iy\in\mathbb{C}:y>-\delta\}, δ=δ(f)>0\delta=\delta(f)>0, are norm-dense in (w,ρ)\mathcal{H}(w,\rho).

Proof.

Fix fH2()f\in H^{2}(\mathbb{R}), and consider the functions fϵ(x)f_{\epsilon}(x) defined by the formula

fϵ(x):=if(x+iϵ)ϵx+i,x,ϵ>0.f_{\epsilon}(x):=\frac{if(x+i\epsilon)}{\epsilon x+i},\quad x\in\mathbb{R},\epsilon>0.

These functions are contained in 1()2()\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) for each ϵ>0\epsilon>0, and they are analytic in a half-space larger than \mathbb{H}. Note that

|iϵx+i|1,x\Bigg{|}\frac{i}{\epsilon x+i}\Bigg{|}\leq 1,\quad x\in\mathbb{R}

and that

limϵ0+iϵx+i=1.\lim_{\epsilon\to 0^{+}}\frac{i}{\epsilon x+i}=1.

We readily see from \threfH1CharacterizationProp and the dominated convergence theorem that we have

limϵ0+|fϵf|2𝑑x=0.\lim_{\epsilon\to 0^{+}}\int_{\mathbb{R}}|f_{\epsilon}-f|^{2}dx=0.

By Plancherel’s theorem, we therefore also have

limϵ0++|fϵ^f^|2𝑑ζ=0.\lim_{\epsilon\to 0^{+}}\int_{\mathbb{R}_{+}}|\widehat{f_{\epsilon}}-\widehat{f}|^{2}d\zeta=0.

A fortiori, we have

limϵ0+|fϵf|2w𝑑x=0\lim_{\epsilon\to 0^{+}}\int_{\mathbb{R}}|f_{\epsilon}-f|^{2}w\,dx=0

and

limϵ0++|fϵ^f^|2ρ𝑑ζ=0.\lim_{\epsilon\to 0^{+}}\int_{\mathbb{R}_{+}}|\widehat{f_{\epsilon}}-\widehat{f}|^{2}\rho\,d\zeta=0.

Thus, as ϵ0\epsilon\to 0, the tuples Jfϵ(w,ρ)Jf_{\epsilon}\in\mathcal{H}(w,\rho) converge in the norm of the space to the tuple (f,f^)(f,\widehat{f}), which is therefore contained in (w,ρ)\mathcal{H}(w,\rho). This proves the first statement of the proposition. The second has the same proof, we merely start with f1()2()f\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) and run the same argument. ∎

The shift operators

Usf(x)=eixsf(x),f2(,wdx)U^{s}f(x)=e^{ixs}f(x),\quad f\in\mathcal{L}^{2}(\mathbb{R},w\,dx)

are unitary. Using the convention that g(x)0g(x)\equiv 0 for x<0x<0 and g2(+,ρdx)g\in\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx), the translation operators

Us^g(x):=g(xs),g2(+,ρdx)\widehat{U^{s}}g(x):=g(x-s),\quad g\in\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx)

are contractions on 2(+,ρdx)\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx), whenever s>0s>0. This fact is a consequence of the assumption that ρ\rho is decreasing:

+|g(xs)|2ρ(x)𝑑x\displaystyle\int_{\mathbb{R}_{+}}|g(x-s)|^{2}\rho(x)dx =s|g(xs)|2ρ(x)𝑑x\displaystyle=\int_{s}^{\infty}|g(x-s)|^{2}\rho(x)dx
s|g(xs)|2ρ(xs)𝑑x\displaystyle\leq\int_{s}^{\infty}|g(x-s)|^{2}\rho(x-s)dx
=+|g(x)|2ρ(x)𝑑x.\displaystyle=\int_{\mathbb{R}_{+}}|g(x)|^{2}\rho(x)dx.

We used that g(xs)g(x-s) vanishes for x(0,s)x\in(0,s). Consequently, the operators

Us:=UsUs^,s>0U^{s}_{*}:=U^{s}\oplus\widehat{U^{s}},\quad s>0

are bounded on the space 2(,wdx)2(+,ρdx)\mathcal{L}^{2}(\mathbb{R},w\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx). Moreover, the Hardy subspace (w,ρ)\mathcal{H}(w,\rho) is invariant for these operators. Indeed, if f1()2()f\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}), then by the well-known property of the Fourier transform Us^f^=Usf^\widehat{U^{s}}\widehat{f}=\widehat{U^{s}f}, we obtain

UsJf=(Usf,Us^f^)=(Usf,Usf^)=JUsf.U_{*}^{s}Jf=(U^{s}f,\widehat{U^{s}}\widehat{f})=(U^{s}f,\widehat{U^{s}f})=JU^{s}f.

The function UsfU^{s}f is contained in 1()2()\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}), and so the above relation shows that a dense subset of (w,ρ)\mathcal{H}(w,\rho) is mapped into (w,ρ)\mathcal{H}(w,\rho) under each of the bounded operators UsU^{s}_{*}. 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.
\thlabel

FourierTransformCauchyKernel Let

ψz(x):=i2π(xz¯),z.\psi_{z}(x):=\frac{i}{\sqrt{2\pi}(x-\overline{z})},\quad z\in\mathbb{H}.

The Fourier transform ψz^\widehat{\psi_{z}} equals

ψz^(ζ)=eiz¯ζ𝟙+(ζ),\widehat{\psi_{z}}(\zeta)=e^{-i\overline{z}\zeta}\mathbbm{1}_{\mathbb{R}_{+}}(\zeta),

and the Fourier transform ψz¯^\widehat{\overline{\psi_{z}}} of the conjugate of ψz\psi_{z} equals

ψz¯^(ζ)=eizζ𝟙(ζ).\widehat{\overline{\psi_{z}}}(\zeta)=e^{-iz\zeta}\mathbbm{1}_{\mathbb{R}_{-}}(\zeta).
Proof.

It is perhaps easiest to apply the Fourier inversion formula to the asserted formula for ψz^\widehat{\psi_{z}}. We readily compute

ψz^(ζ)eiζx𝑑λ(ζ)\displaystyle\int_{\mathbb{R}}\widehat{\psi_{z}}(\zeta)e^{i\zeta x}\,d\lambda(\zeta) =12π0e(iz¯+ix)ζ𝑑ζ\displaystyle=\frac{1}{\sqrt{2\pi}}\int_{0}^{\infty}e^{(-i\overline{z}+ix)\zeta}\,d\zeta
=ψz(x).\displaystyle=\psi_{z}(x).

The other formula follows from ψz¯^(ζ)=ψz^(ζ)¯\widehat{\overline{\psi_{z}}}(\zeta)=\overline{\widehat{\psi_{z}}(-\zeta)}, which is an easily established property of the Fourier transform. ∎

Proposition 4.2.
\thlabel

ConvolutionFourierDecayLemma Assume that f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) satisfies

ρf^(x)=x|f^(ζ)|𝑑ζ=𝒪(ecx),x>0\rho_{\widehat{f}}(x)=\int_{x}^{\infty}|\widehat{f}(\zeta)|d\zeta=\mathcal{O}\big{(}e^{-c\sqrt{x}}\big{)},\quad x>0

for some c>0c>0, and let

s(x):=ψi(x)¯=i2π(xi).s(x):=\overline{\psi_{i}(x)}=\frac{-i}{\sqrt{2\pi}(x-i)}.

Then

|fs^(ζ)|=𝒪(ecζ),ζ>0.\big{|}\widehat{fs}(\zeta)\big{|}=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)},\quad\zeta>0.
Proof.

Note that fs1(,dx)fs\in\mathcal{L}^{1}(\mathbb{R},dx), and recall that the Fourier transform fs^\widehat{fs} is thus a continuous function given by the convolution of the Fourier transforms of ff and ss. By \threfFourierTransformCauchyKernel, we obtain

s^(ζ)=eζ𝟙(ζ),\widehat{s}(\zeta)=e^{\zeta}\mathbbm{1}_{\mathbb{R}_{-}}(\zeta),

and so

fs^(ζ)=f^(x)eζx𝟙(ζx)𝑑λ(x)=ζf^(x)eζx𝑑λ(x).\widehat{fs}(\zeta)=\int_{\mathbb{R}}\widehat{f}(x)e^{\zeta-x}\mathbbm{1}_{\mathbb{R}_{-}}(\zeta-x)\,d\lambda(x)=\int_{\zeta}^{\infty}\widehat{f}(x)e^{\zeta-x}\,d\lambda(x).

The exponential term in the last integral is bounded by 11. Therefore |fs^(ζ)|ρf^(ζ)\big{|}\widehat{fs}(\zeta)\big{|}\leq\rho_{\widehat{f}}(\zeta), and the desired estimate follows from the decay assumption on ρf^\rho_{\widehat{f}}. ∎

4.2. Strategy of the proof of \threfCondensationTheorem

Given a function f2(,dx)f\in\mathcal{L}^{2}(\mathbb{R},dx) we consider the set

E:={x:|f(x)|>0},E:=\{x\in\mathbb{R}:|f(x)|>0\},

which is well-defined up to a set of Lebesgue measure zero. Let \mathcal{F} denote the family of all finite open intervals II which satisfy

Ilog|f(x)|dx>,\int_{I}\log|f(x)|\,dx>-\infty,

and set

U:=II.U:=\cup_{I\in\mathcal{F}}I.

Since log|f|\log|f|\equiv-\infty on IEI\setminus E for every interval II, it follows that if log|f|\log|f| is integrable on II, then the set difference IEI\setminus E must have measure zero. Consequently, since one can easily argue that we can express UU as a countable union of intervals II on which log|f|\log|f| is integrable, the Lebesgue measure of the set difference UEU\setminus E must be zero. However, the set difference EUE\setminus U might have positive measure. We set

res(f):=EU\text{res}(f):=E\setminus U

and call this set the residual of ff. The residual is well-defined up to a set of Lebesgue measure zero.

Claim 1.
\thlabel

claim1 Under the assumption that ρf^(ζ)=𝒪(ecζ)\rho_{\widehat{f}}(\zeta)=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)} for some c>0c>0, the set res(f)\text{res}(f) has Lebesgue measure zero.

\thref

CondensationTheorem follows immediately from the above claim. Indeed, the roles of ff and f^\widehat{f} may obviously be interchanged in the statement of \threfCondensationTheorem, and the above claim implies that the open set UU equals EE up to an error of measure zero. Local integrability of log|f|\log|f| on the set UU follows from its construction.

We set

w(x):=min(|f(x)|2,1).w(x):=\min(|f(x)|^{2},1).

Note that res(w)=res(f)\text{res}(w)=\text{res}(f) and that w1(,dx)w\in\mathcal{L}^{1}(\mathbb{R},dx). Our \threfclaim1 will follow from the next assertion.

Claim 2.
\thlabel

claim2 Let ρ:++\rho:\mathbb{R}_{+}\to\mathbb{R}_{+} be a bounded, continuous, non-negative and decreasing function which satisfies ρ(x)=𝒪(edx)\rho(x)=\mathcal{O}\big{(}e^{-d\sqrt{x}}\big{)} for some d>0d>0 and x>0x>0. Then, every tuple of the form

(h,0)2(,wdx)2(+,ρdx),(h,0)\in\mathcal{L}^{2}(\mathbb{R},w\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx),

where hh is any function in 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx) which lives only on the set res(w)\text{res}(w), is contained in the Hardy subspace (w,ρ)\mathcal{H}(w,\rho).

To prove \threfclaim1 from \threfclaim2 we will use a trick involving Plancherel’s theorem. We set ρ(x)=ecx\rho(x)=e^{-c\sqrt{x}}, where c>0c>0 is the constant appearing in \threfclaim1. Let hh be as in \threfclaim2, and

s(x):=ψi¯(x)=i2π(xi)s(x):=\overline{\psi_{i}}(x)=\frac{-i}{\sqrt{2\pi}(x-i)}

be as in \threfConvolutionFourierDecayLemma. We will show that

hfs¯𝑑x=0.\int_{\mathbb{R}}h\overline{fs}\,dx=0.

This implies, by the generality of hh, that fsfs is zero on the set res(w)=res(f)\text{res}(w)=\text{res}(f). Since ss is non-zero everywhere on \mathbb{R}, in fact ff is zero on res(f)\text{res}(f). Since res(f)E={x:|f(x)|>0}\text{res}(f)\subset E=\{x\in\mathbb{R}:|f(x)|>0\}, 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 (h,0)(w,ρ)(h,0)\in\mathcal{H}(w,\rho), there exists a sequence {gn}\{g_{n}\} of functions gn1()2()g_{n}\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) such that gnhg_{n}\to h in the norm of 2(,wdx)\mathcal{L}^{2}(\mathbb{R},w\,dx) and gn^0\widehat{g_{n}}\to 0 in the norm of 2(+,ρdx)\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx). Consider the quantities

(hgn)fs¯𝑑λ=E(hgn)wfs¯w𝑑λ.\int_{\mathbb{R}}(h-g_{n})\overline{fs}\,d\lambda=\int_{E}(h-g_{n})\sqrt{w}\frac{\overline{fs}}{\sqrt{w}}d\lambda.

We passed from domain of integration \mathbb{R} into EE, since ff vanishes outside of EE anyway (note also that w>0w>0 almost everywhere on EE). By the Cauchy-Schwarz inequality, we obtain

|(hgn)fs¯𝑑λ||hgn|2w𝑑λE|f|2|w|s|2𝑑λ.\displaystyle\Bigg{|}\int_{\mathbb{R}}(h-g_{n})\overline{fs}\,d\lambda\Bigg{|}\leq\sqrt{\int_{\mathbb{R}}|h-g_{n}|^{2}w\,d\lambda}\sqrt{\int_{E}\frac{|f|^{2}|}{w}|s|^{2}d\lambda}.

Note that the first of the factors on the right-hand side of the inequality above converges to 0. The other factor is finite. Indeed, since |f|2/w1|f|^{2}/w\equiv 1 on the set where |f|<1|f|<1, and |f|2/w=|f|2|f|^{2}/w=|f|^{2} on the set where |f|1|f|\geq 1, we obtain |f|2w|f|2+1\frac{|f|^{2}}{w}\leq|f|^{2}+1, and consequently

|f|2w|s|2|f|2|s|2+|s|21(,dx).\frac{|f|^{2}}{w}|s|^{2}\leq|f|^{2}|s|^{2}+|s|^{2}\in\mathcal{L}^{1}(\mathbb{R},dx).

This computation implies the formula

hfs¯𝑑λ=limngnfs¯𝑑λ=limn+gn^fs^¯𝑑λ,\int_{\mathbb{R}}h\overline{fs}\,d\lambda=\lim_{n\to\infty}\int_{\mathbb{R}}g_{n}\overline{fs}\,d\lambda=\lim_{n\to\infty}\int_{\mathbb{R}_{+}}\widehat{g_{n}}\overline{\widehat{fs}}\,d\lambda,

where we used Plancherel’s theorem in the last step. Now, recall that by \threfConvolutionFourierDecayLemma we may estimate

|+gn^fs^¯𝑑λ|A+|gn^(ζ)|ecζ𝑑λ(ζ)\Bigg{|}\int_{\mathbb{R}_{+}}\widehat{g_{n}}\overline{\widehat{fs}}\,d\lambda\Bigg{|}\leq A\int_{\mathbb{R}_{+}}|\widehat{g_{n}}(\zeta)|e^{-c\sqrt{\zeta}}\,d\lambda(\zeta)

for some positive constant AA. By again using Cauchy-Schwarz inequality, we obtain

|gn^fs^¯𝑑λ|A+|gn^(ζ)|2ecζ𝑑λ(ζ)+ecζ𝑑λ(ζ).\displaystyle\Bigg{|}\int_{\mathbb{R}}\widehat{g_{n}}\overline{\widehat{fs}}\,d\lambda\Bigg{|}\leq A\sqrt{\int_{\mathbb{R}_{+}}|\widehat{g_{n}}(\zeta)|^{2}e^{-c\sqrt{\zeta}}\,d\lambda(\zeta)}\sqrt{\int_{\mathbb{R}_{+}}e^{-c\sqrt{\zeta}}\,d\lambda(\zeta)}.

The second factor on the right-hand side is certainly finite. Since ρ(x)=ecx\rho(x)=e^{-c\sqrt{x}} and gn0g_{n}\to 0 in 2(+,ρdx)\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx), the first factor above converges to 0, as nn\to\infty. All in all, we have obtained that

hfs¯𝑑λ=limn+gn^fs^¯𝑑λ=0.\int_{\mathbb{R}}h\overline{fs}\,d\lambda=\lim_{n\to\infty}\int_{\mathbb{R}_{+}}\widehat{g_{n}}\overline{\widehat{fs}}\,d\lambda=0.

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.
\thlabel

claim3 There exists a compact set EE\subset\mathbb{R} of positive Lebesgue measure, and an increasing function M:++M:\mathbb{R}_{+}\to\mathbb{R}_{+} which satisfies

(4.1) limxM(x)xa=\lim_{x\to\infty}\frac{M(x)}{x^{a}}=\infty

for every a(0,1/2)a\in(0,1/2), such that if

ρ(x):=eM(x),x>0,\rho(x):=e^{-M(x)},\quad x>0,

then the Hardy subspace (𝟙E,ρ)\mathcal{H}(\mathbbm{1}_{E},\rho) is properly contained in 2(E,𝟙Edx)2(+,ρdx)\mathcal{L}^{2}(E,\mathbbm{1}_{E}\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx).

We proceed to show how one proves \threfSparsenessTheorem from this claim. Let EE and ρ=eM\rho=e^{-M} be as in \threfclaim3, and assume that the non-zero tuple (h,k)2(E,𝟙Edx)2(+,ρdx)(h,k)\in\mathcal{L}^{2}(E,\mathbbm{1}_{E}\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx) is orthogonal to (𝟙E,ρ)\mathcal{H}(\mathbbm{1}_{E},\rho). We shall soon see that hh in fact is non-zero. The function hh lives only on the set EE, and we will show that it has the required spectral decay. Recall \threfFourierTransformCauchyKernel, let ψz2()\psi_{z}\in\mathcal{H}^{2}(\mathbb{R}) be as in that lemma, and let gg be the inverse Fourier transform of kρk\rho, so that g^=kρ2(+,dx)\widehat{g}=k\rho\in\mathcal{L}^{2}(\mathbb{R}_{+},dx). In fact g2()g\in\mathcal{H}^{2}(\mathbb{R}), since its spectrum is positive. The orthogonality means that

0\displaystyle 0 =Ehψz¯𝑑λ++kψz^¯ρ𝑑λ\displaystyle=\int_{E}h\overline{\psi_{z}}\,d\lambda+\int_{\mathbb{R}_{+}}k\overline{\widehat{\psi_{z}}}\rho\,d\lambda
=Ehψz¯𝑑λ+gψz¯𝑑λ.\displaystyle=\int_{E}h\overline{\psi_{z}}\,d\lambda+\int_{\mathbb{R}}g\overline{\psi_{z}}\,d\lambda.

We used Plancherel’s theorem. The above relation shows that the function G:=h+g2(,dx)G:=h+g\in\mathcal{L}^{2}(\mathbb{R},dx) is orthogonal in 2(,dx)\mathcal{L}^{2}(\mathbb{R},dx) to each of the functions ψz\psi_{z}. Let P:2(,dx)2()P:\mathcal{L}^{2}(\mathbb{R},dx)\to\mathcal{H}^{2}(\mathbb{R}) be the orthogonal projection. In terms of Fourier transforms, we have

Ph^(ζ)=h^(ζ)𝟙+(ζ).\widehat{Ph}(\zeta)=\widehat{h}(\zeta)\mathbbm{1}_{\mathbb{R}_{+}}(\zeta).

Then

G0:=PG=Ph+gG_{0}:=PG=Ph+g

is orthogonal not only to ψz\psi_{z}, but also to ψz¯\overline{\psi_{z}}, since by \threfFourierTransformCauchyKernel the functions ψz¯\overline{\psi_{z}} have spectrum contained in \mathbb{R}_{-}. But then the decomposition formula for the Poisson kernel in (2.1) shows that

G0(t)𝒫(t,z)𝑑t=0\int_{\mathbb{R}}G_{0}(t)\mathcal{P}(t,z)\,dt=0

for each zz\in\mathbb{H}, and it is an elementary fact about the Poisson kernel that we must, in this case, have G00G_{0}\equiv 0. So Ph=gPh=-g. We can now argue that h0h\neq 0. Indeed, if h=0h=0, then g=0g=0. Since g^=kρ\widehat{g}=k\rho, that would mean k=0k=0, contradicting that the tuple (h,k)(h,k) is non-zero. Having established that h0h\neq 0, we proceed by taking Fourier transforms to obtain

h^𝟙+=Ph^=g^=kρ.\widehat{h}\mathbbm{1}_{\mathbb{R}_{+}}=\widehat{Ph}=-\widehat{g}=-k\rho.

Using Cauchy-Schwarz inequality, we may now estimate

ρh^(x)\displaystyle\rho_{\widehat{h}}(x) =x|k|ρ𝑑ζ\displaystyle=\int_{x}^{\infty}|k|\rho\,d\zeta
xρ𝑑ζx|k|2ρ𝑑ζ,x>0.\displaystyle\leq\sqrt{\int_{x}^{\infty}\rho\,d\zeta}\sqrt{\int_{x}^{\infty}|k|^{2}\rho\,d\zeta},\quad x>0.

The second factor is finite, and the growth of MM asserted in \threfclaim3 implies that for every fixed a(0,1/2)a\in(0,1/2) there exists a constant C(a)>0C(a)>0 for which we have

ρ(ζ)=eM(ζ)eC(a)ζa,ζ>0.\rho(\zeta)=e^{-M(\zeta)}\leq e^{-C(a)\zeta^{a}},\quad\zeta>0.

It follows that the integral inside the square root of the first factor above satisfies

xρ𝑑ζxeC(a)ζa𝑑ζ=𝒪(eC(a)xa),x>0.\int_{x}^{\infty}\rho\,d\zeta\leq\int_{x}^{\infty}e^{-C(a)\zeta^{a}}\,d\zeta=\mathcal{O}\big{(}e^{-C(a)x^{a}}\big{)},\quad x>0.

Since a(0,1/2)a\in(0,1/2) was arbitrary, we conclude that ρh^(x)=𝒪(exa)\rho_{\widehat{h}}(x)=\mathcal{O}\big{(}e^{-x^{a}}) for every a(0,1/2)a\in(0,1/2) and x>0x>0. 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 w(x)=min(|f(x)|2,1)1(,dx)w(x)=\min(|f(x)|^{2},1)\in\mathcal{L}^{1}(\mathbb{R},dx) and that ρ\rho has the decay ρ(ζ)=𝒪(edζ)\rho(\zeta)=\mathcal{O}\big{(}e^{-d\sqrt{\zeta}}\big{)} for ζ>0\zeta>0 and some d>0d>0. Note that we may assume throughout that

logw(x)1+x2𝑑x=.\int_{\mathbb{R}}\frac{\log w(x)}{1+x^{2}}\,dx=-\infty.

Indeed, if on the contrary this integral converges, then res(w)=res(f)\text{res}(w)=\text{res}(f) is void, and both \threfclaim2 of Section 4 and \threfCondensationTheorem (with ff and f^\widehat{f} playing opposite roles) hold trivially.

We may decompose res(w)\text{res}(w) as

res(w)=m1Fm\text{res}(w)=\cup_{m\geq 1}\,F_{m}

where

(5.1) Fm:=[m,m]{x:w(x)>1/m}res(w).F_{m}:=[-m,m]\cap\{x\in\mathbb{R}:w(x)>1/m\}\cap\text{res}(w).

The set equality above holds up to an error of measure zero. The sets FmF_{m} are bounded, and on each of them ww is bounded from below.

Proposition 5.1.
\thlabel

hnSplittingSequence In order to establish \threfclaim2, it suffices to construct, for any fixed mm\in\mathbb{N} and c>0c>0, a sequence of functions {hn}n\{h_{n}\}_{n} in ()\mathcal{H}^{\infty}(\mathbb{R}) which has the following properties.

  1. (i)

    The analytic extensions of the functions hnh_{n} to \mathbb{H} satisfy the bound |hn(x+iy)|ecy|h_{n}(x+iy)|\leq e^{\frac{c}{y}} for y>0y>0,

  2. (ii)

    limnhn(x)=0\lim_{n\to\infty}h_{n}(x)=0 for almost every xFmx\in F_{m},

  3. (iii)

    limnhn(z)=1\lim_{n\to\infty}h_{n}(z)=1 for every zz\in\mathbb{H},

  4. (iv)

    there exists an A>0A>0 and p>2p>2 such that |h(x)|pw(x)<A|h(x)|^{p}w(x)<A for almost every xx\in\mathbb{R}.

Proof.

Property (i)(i), together with \threfFourierDecayFromExtensionGrowthCorollary, implies that the functions gn(x)=hn(x)(i+x)21()2()g_{n}(x)=\frac{h_{n}(x)}{(i+x)^{2}}\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) obey the spectral bound |gn^(ζ)|π2e2cζ,ζ>0|\widehat{g_{n}}(\zeta)|\leq\sqrt{\frac{\pi}{2}}e^{2\sqrt{c}\sqrt{\zeta}},\zeta>0. Since ρ(ζ)=𝒪(edζ)\rho(\zeta)=\mathcal{O}\big{(}e^{-d\sqrt{\zeta}}\big{)} for some d>0d>0, the spectral bound implies that

supn+|gn^|2ρ𝑑ζ<\sup_{n}\int_{\mathbb{R}_{+}}|\widehat{g_{n}}|^{2}\rho\,d\zeta<\infty

if cc is small enough. Together with (iv)(iv), we see that Jgn=(gn,gn^)Jg_{n}=(g_{n},\widehat{g_{n}}) forms a bounded subset of the Hardy subspace of product space 2(,wdx)2(+,ρdx)\mathcal{L}^{2}(\mathbb{R},w\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx). We may thus assume, by passing to a subsequence, that {Jg}n\{Jg\}_{n} tends weakly in (w,ρ)\mathcal{H}(w,\rho) to some tuple (h,k)(h,k). In fact, (iv)(iv) implies that {gn}n\{g_{n}\}_{n} is a sequence bounded in p(,wdx)\mathcal{L}^{p}(\mathbb{R},w\,dx), so we have that hp(,wdx)h\in\mathcal{L}^{p}(\mathbb{R},w\,dx) for some p>2p>2. The fact that h0h\equiv 0 on FmF_{m} is a consequence of the weak convergence of gng_{n} to hh and the condition (ii)(ii), which implies limngn(x)=0\lim_{n\to\infty}g_{n}(x)=0 for almost every xFmx\in F_{m}. Moreover, by the formula in \threfH1FourierTransformFormula, we have

gn^(ζ)=eyζhn(x+iy)(i+x+iy)2eixζ𝑑λ(x)\widehat{g_{n}}(\zeta)=e^{y\zeta}\int_{\mathbb{R}}\frac{h_{n}(x+iy)}{(i+x+iy)^{2}}e^{-ix\zeta}d\lambda(x)

for any y>0y>0. The integrand converges pointwise to eixζ(i+x+iy)2\frac{e^{-ix\zeta}}{(i+x+iy)^{2}} by (iii)(iii), and it is dominated pointwise by the integrable function

xecy1+x2,x.x\mapsto\frac{e^{\frac{c}{y}}}{1+x^{2}},\quad x\in\mathbb{R}.

The dominated convergence theorem implies that

limngn^(ζ)=eyζ1i+x+iyeixζ𝑑λ(x)=1(i+x)2eixζ𝑑λ(x)=Ψ^(ζ),\lim_{n\to\infty}\widehat{g_{n}}(\zeta)=e^{y\zeta}\int_{\mathbb{R}}\frac{1}{i+x+iy}e^{-ix\zeta}\,d\lambda(x)=\int_{\mathbb{R}}\frac{1}{(i+x)^{2}}e^{-ix\zeta}\,d\lambda(x)=\widehat{\Psi}(\zeta),

where Ψ(x):=1(i+x)21()2()\Psi(x):=\frac{1}{(i+x)^{2}}\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}). In the next-to-last equality we used \threfH1FourierTransformFormula backwards. Weak and pointwise convergence implies, as previously, that k=Ψ^k=\widehat{\Psi}. Since JΨ(w,ρ)J\Psi\in\mathcal{H}(w,\rho), we have that (h,k)JΨ=(hΨ,0)(w,ρ)(h,k)-J\Psi=(h-\Psi,0)\in\mathcal{H}(w,\rho). The function hΨh-\Psi is non-zero almost everywhere on FmF_{m}. Indeed, hh vanishes on FmF_{m}, and Ψ(x)\Psi(x) is non-zero everywhere on FmF_{m}. Also, since hp(,wdx)h\in\mathcal{L}^{p}(\mathbb{R},w\,dx), we have hΨp(,wdx)h-\Psi\in\mathcal{L}^{p}(\mathbb{R},w\,dx). The conditions to apply \threfinvSubspaceCorollary are thus satisfied, and by the invariance of (w,ρ)\mathcal{H}(w,\rho) under the operators UsU^{*}_{s} defined in Section 3, we conclude that

2(E,wdx){0}(w,ρ),\mathcal{L}^{2}(E,w\,dx)\oplus\{0\}\subset\mathcal{H}(w,\rho),

where FmE:={x:|h(x)Ψ(x)|>0}F_{m}\subset E:=\{x\in\mathbb{R}:|h(x)-\Psi(x)|>0\}. Since mm is arbitrary, we conclude that

2(mFm,wdx){0}=2(res(w),wdx){0}(w,ρ).\mathcal{L}^{2}(\cup_{m}F_{m},w\,dx)\oplus\{0\}=\mathcal{L}^{2}(\text{res}(w),w\,dx)\oplus\{0\}\subset\mathcal{H}(w,\rho).

This is sufficient to conclude the validity of \threfclaim2. ∎

5.2. An estimate for Poisson integrals

Let μ\mu be a finite real-valued measure on \mathbb{R}. The Poisson integral of μ\mu is the harmonic function 𝒫μ:\mathcal{P}_{\mu}:\mathbb{H}\to\mathbb{R} which is given by the formula

𝒫μ(z):=𝒫(t,z)𝑑μ(t)=1πy(xt)2+y2𝑑μ(t),z=x+iy.\mathcal{P}_{\mu}(z):=\int_{\mathbb{R}}\mathcal{P}(t,z)\,d\mu(t)=\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{(x-t)^{2}+y^{2}}\,d\mu(t),\quad z=x+iy\in\mathbb{H}.

By the triangle inequality, and an estimation of 𝒫(t,z)\mathcal{P}(t,z) by its supremum 1πy\frac{1}{\pi y} over \mathbb{R}, we easily obtain the inequality

|𝒫μ(z)||μ|()πy,z=x+iy,\big{|}\mathcal{P}_{\mu}(z)\big{|}\leq\frac{|\mu|(\mathbb{R})}{\pi y},\quad z=x+iy\in\mathbb{H},

and where |μ||\mu| denotes the usual variation of the measure μ\mu. We obtain a much better inequality for measures μ\mu which are oscillating rapidly. The following lemma is the half-plane version of an estimate in [11, Lemma 3.2].

Lemma 5.2.
\thlabel

PsnIntegralEstimate Let μ\mu be a finite real-valued measure on \mathbb{R} which has the following structure: there exists a finite sequence of disjoint intervals {Ij}j\{I_{j}\}_{j} of \mathbb{R}, and a decomposition μ=jμj\mu=\sum_{j}\mu_{j}, where μj\mu_{j} is a real-valued measure supported inside IjI_{j}, μj(Ij)=0\mu_{j}(I_{j})=0, and |μj|(Ij)C|\mu_{j}|(I_{j})\leq C for some C>0C>0 which is independent of jj. Then

|𝒫μ(x+iy)|Cπy,x+iy.\big{|}\mathcal{P}_{\mu}(x+iy)\big{|}\leq\frac{C}{\pi y},\quad x+iy\in\mathbb{H}.
Proof.

Since μ-\mu satisfies the same conditions as μ\mu, and so does any translation of μ\mu, it suffices to prove that 𝒫μ(y)Cπy\mathcal{P}_{\mu}(y)\leq\frac{C}{\pi y} for any y>0y>0. We have

𝒫μ(y)=j1πyt2+y2𝑑μj(t).\mathcal{P}_{\mu}(y)=\sum_{j}\frac{1}{\pi}\int_{\mathbb{R}}\frac{y}{t^{2}+y^{2}}d\mu_{j}(t).

If μj=μj+μj\mu_{j}=\mu_{j}^{+}-\mu_{j}^{-} is the decomposition of μj\mu_{j} into its positive and negative parts, then we have the estimate

𝒫μ(y)\displaystyle\mathcal{P}_{\mu}(y) 1πjsuptIjyt2+y2μj+(Ij)inftIjyt2+y2μj(Ij)\displaystyle\leq\frac{1}{\pi}\sum_{j}\sup_{t\in I_{j}}\frac{y}{t^{2}+y^{2}}\cdot\mu^{+}_{j}(I_{j})-\inf_{t\in I_{j}}\frac{y}{t^{2}+y^{2}}\cdot\mu^{-}_{j}(I_{j})
C2πjsuptIjyt2+y2inftIjyt2+y2\displaystyle\leq\frac{C}{2\pi}\sum_{j}\sup_{t\in I_{j}}\frac{y}{t^{2}+y^{2}}-\inf_{t\in I_{j}}\frac{y}{t^{2}+y^{2}}
:=C2πS\displaystyle:=\frac{C}{2\pi}S

In the second step we used that μj(Ij)=μj+(Ij)μj(Ij)=0\mu_{j}(I_{j})=\mu_{j}^{+}(I_{j})-\mu_{j}^{-}(I_{j})=0 and that |μj|(Ij)=μj+(Ij)+μj(Ij)C|\mu_{j}|(I_{j})=\mu_{j}^{+}(I_{j})+\mu_{j}^{-}(I_{j})\leq C. Since the intervals {Ij}j\{I_{j}\}_{j} are disjoint, and the function tyt2+y2t\mapsto\frac{y}{t^{2}+y^{2}} is increasing for t<0t<0, decreasing for t>0t>0, and attains a maximum value of 1/y1/y at t=0t=0, the sum SS in the above estimate cannot be larger than 2y\frac{2}{y} (it is easily seen to be bounded by twice the height of the graph of tyt2+y2t\mapsto\frac{y}{t^{2}+y^{2}}). 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 F:=FmF:=F_{m} is a bounded subset of res(w){x:w(x)>δ}\text{res}(w)\cap\{x\in\mathbb{R}:w(x)>\delta\} for some δ>0\delta>0. The set FF inherits the following property from res(w)\text{res}(w): if II is an interval, and |IF|>0|I\cap F|>0, then Ilogwdx=.\int_{I}\log w\,dx=-\infty.

Lemma 5.3.
\thlabel

pieceLemma If |IF|>0|I\cap F|>0 for some finite interval II, then given any c>0c>0 and any p>0p>0, there exists D>0D>0 and a measurable subset EIIE_{I}\subset I disjoint from FF for which we have

(5.2) EImin(p1log+(1/w),D)𝑑x=c.\int_{E_{I}}\min\big{(}p^{-1}\log^{+}(1/w),D\big{)}\,dx=c.
Proof.

On the set FF, log(w)\log(w) is bounded from below by logδ\log\delta. Hence IFlogwdx=\int_{I\setminus F}\log w\,dx=-\infty. Consequently,

limD+IFmin(p1log+(1/w),D)𝑑x=IFp1log+(1/w)𝑑x=+.\lim_{D\to+\infty}\int_{I\setminus F}\min\big{(}p^{-1}\log^{+}(1/w),D\big{)}\,dx=\int_{I\setminus F}p^{-1}\log^{+}(1/w)\,dx=+\infty.

So for DD sufficiently large we will have IFmin(p1log+(1/w),D)𝑑x>c\int_{I\setminus F}\min\big{(}p^{-1}\log^{+}(1/w),D\big{)}\,dx>c, and then by the absolute continuity of the finite positive measure min(p1log+(1/w),D)dx\min\big{(}p^{-1}\log^{+}(1/w),D\big{)}\,dx we may choose a set EIIFE_{I}\subset I\setminus F for which (5.2) holds. ∎

We will now construct the sequence in \threfhnSplittingSequence. Let p>2p>2, {cn}n1\{c_{n}\}_{n\geq 1} be some sequence of positive numbers which tends to 0 slowly enough so that cn2nc_{n}2^{n} tends to ++\infty, and let FF be as above. Fix some integer n>0n>0, cover \mathbb{R} by a sequence of (say, half-open) disjoint intervals of length 2n2^{-n} and let {k}k\{\ell_{k}\}_{k} be those intervals for which |kF|>0|\ell_{k}\cap F|>0. Apply \threfpieceLemma with c=cnc=c_{n} to each of the intervals k\ell_{k} to obtain a corresponding constant Dk>0D_{k}>0 and a set Ek:=EkkE_{k}:=E_{\ell_{k}}\subset\ell_{k} for which (5.2) holds. We set dμ(t)=logW(t)dtd\mu(t)=\log W(t)\,dt, where

logW(t)=kmin(p1log+(1/w(t)),Dk)𝟙Ek(t)cn|kF|𝟙kF(t).\log W(t)=\sum_{k}\min\big{(}p^{-1}\log^{+}(1/w(t)),D_{k}\big{)}\mathbbm{1}_{E_{k}}(t)-\frac{c_{n}}{|\ell_{k}\cap F|}\mathbbm{1}_{\ell_{k}\cap F}(t).

Then μ\mu is an absolutely continuous real-valued measure with bounded density logW\log W, μ(k)=0\mu(\ell_{k})=0 and |μ|(k)=2cn|\mu|(\ell_{k})=2c_{n}. We construct hn()h_{n}\in\mathcal{H}^{\infty}(\mathbb{R}) by letting the logarithm loghn(z)\log h_{n}(z) of its extension to \mathbb{H} be given by the right-hand side of the formula (2.7). Then, by \threfPsnIntegralEstimate, we have the inequalities

|𝒫μ(x+iy)|2cnπy,z=x+iy|\mathcal{P}_{\mu}(x+iy)|\leq\frac{2c_{n}}{\pi y},\quad z=x+iy\in\mathbb{H}

and consequently, since |h(z)|=e𝒫μ(z)|h(z)|=e^{\mathcal{P}_{\mu}(z)}, we have the bounds

(5.3) e2cn/πy|hn(z)|e2cn/πy,z=x+iy.e^{-2c_{n}/\pi y}\leq|h_{n}(z)|\leq e^{2c_{n}/\pi y},\quad z=x+iy\in\mathbb{H}.

Since cn0c_{n}\to 0, by multiplying hnh_{n} by an appropriate unimodular constant, we may assume by (5.3) that

(5.4) limnhn(z)=1\lim_{n\to\infty}h_{n}(z)=1

for every zz\in\mathbb{H} (possibly after passing to a subsequence). Also, for almost every xFx\in F, we have

(5.5) |hn(x)|=ecn/|kF|ecn2n.|h_{n}(x)|=e^{-c_{n}/|\ell_{k}\cap F|}\leq e^{-c_{n}2^{n}}.

Our assumption on cnc_{n} then implies that limnhn(x)=0\lim_{n\to\infty}h_{n}(x)=0 for almost every xFx\in F. For almost every xFx\in\mathbb{R}\setminus F, we have instead

(5.6) |hn(x)|pw(x)=eplogW(x)w(x)elog+(1/w(x))w(x)1.|h_{n}(x)|^{p}w(x)=e^{p\log W(x)}w(x)\leq e^{\log^{+}(1/w(x))}w(x)\leq 1.

The equations (5.3), (5.4), (5.5) and (5.6) show that the sequence {hn}n\{h_{n}\}_{n} satisfies the conditions stated in \threfhnSplittingSequence. Thus \threfclaim2 holds, and consequently the proof of \threfCondensationTheorem is complete.

6. Proof of \threfSparsenessTheorem

6.1. A bit of concave analysis

Let M:++M:\mathbb{R}_{+}\to\mathbb{R}_{+} be an increasing and concave function which is differentiable for x>0x>0. We assume that M(0)=0M(0)=0, and that

(6.1) M(x)x,x>0.M(x)\leq\sqrt{x},\quad x>0.

For a function MM with the above properties, the integrals

(6.2) IM(y):=0eM(x)2yx𝑑xI_{M}(y):=\int_{0}^{\infty}e^{M(x)-2yx}\,dx

converge for every y>0y>0, and estimation of the growth of IM(y)I_{M}(y) as y0+y\to 0^{+} will be of importance in the proof of \threfSparsenessTheorem. To estimate IMI_{M}, we define a function MM_{*} in the following way. Since MM is increasing, concave and differentiable, the derivative M(x)M^{\prime}(x) is defined for x>0x>0, and it is a positive and decreasing function. The concavity of MM implies that

(6.3) M(x)M(x)/xM^{\prime}(x)\leq M(x)/x

from which it follows by (6.1) that

limxM(x)=0.\lim_{x\to\infty}M^{\prime}(x)=0.

It is only the asymptotic behaviour of MM as xx\to\infty that concerns us, so we will also assume for convenience that limx0+M(x)=+\lim_{x\to 0^{+}}M^{\prime}(x)=+\infty. In this case, the inverse function

(6.4) K(y):=(M)1(y),y(0,)K(y):=(M^{\prime})^{-1}(y),\quad y\in(0,\infty)

is well-defined and positive. It is decreasing, and satisfies limy0+K(y)=+\lim_{y\to 0^{+}}K(y)=+\infty. We set

(6.5) M(y):=M(K(y)),y(0,)M_{*}(y):=M(K(y)),\quad y\in(0,\infty)

The function MM_{*} is decreasing, and satisfies limy0+M(y)=+\lim_{y\to 0^{+}}M_{*}(y)=+\infty. The integrals IM(y)I_{M}(y) can be estimated in terms of MM_{*}.

Proposition 6.1.
\thlabel

IMestimate For MM as above, we have

IM(y)2eM(y)y2I_{M}(y)\leq\frac{2e^{M_{*}(y)}}{y^{2}}

for all sufficiently small y>0y>0.

Proof.

We will need the following observation. The inequality (6.3) together with (6.1) implies that M(x)1xM^{\prime}(x)\leq\frac{1}{\sqrt{x}} if x>0x>0 is sufficiently large. We want to show the inequality

K(y)1y2,K(y)\leq\frac{1}{y^{2}},

for sufficiently small y>0y>0. Set K(y)=xK(y)=x and 1y2=x\frac{1}{y^{2}}=x_{*}. Then

M(x)=y=1xM(x)M^{\prime}(x)=y=\frac{1}{\sqrt{x_{*}}}\geq M^{\prime}(x_{*})

if y>0y>0 is sufficiently small (and consequently xx_{*} is sufficiently large). Since MM^{\prime} is a decreasing function, the above inequality shows that xxx_{*}\geq x, which is the same as the desired inequality.

We split the integral (6.2) at x=K(y)x=K(y). Both pieces can be estimated very crudely. For the first piece, we have

0K(y)eM(x)2yx𝑑x\displaystyle\int_{0}^{K(y)}e^{M(x)-2yx}\,dx 0K(y)eM(K(y))𝑑x\displaystyle\leq\int_{0}^{K(y)}e^{M(K(y))}\,dx
=K(y)eM(y)\displaystyle=K(y)e^{M_{*}(y)}
eM(y)y2.\displaystyle\leq\frac{e^{M_{*}(y)}}{y^{2}}.

We used our initial observation in the last step. For the second piece, we note that

supx>0M(x)yx=M(K(y))yK(y)M(y).\sup_{x>0}\,M(x)-yx=M(K(y))-yK(y)\leq M_{*}(y).

Indeed, the supremum is attained at the point xx where M(x)=yM^{\prime}(x)=y, which by definition is x=K(y)x=K(y). Thus

K(y)eM(x)2yx𝑑x\displaystyle\int_{K(y)}^{\infty}e^{M(x)-2yx}\,dx\leq K(y)eM(y)xy𝑑x\displaystyle\int_{K(y)}^{\infty}e^{M_{*}(y)-xy}\,dx
\displaystyle\leq eM(y)0eyx𝑑x\displaystyle e^{M_{*}(y)}\int_{0}^{\infty}e^{-yx}\,dx
=\displaystyle= eM(y)y\displaystyle\frac{e^{M_{*}(y)}}{y}
\displaystyle\leq eM(y)y2.\displaystyle\frac{e^{M_{*}(y)}}{y^{2}}.

In the last inequality we require that y(0,1)y\in(0,1). 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 MM_{*} near the origin. The next proposition describes the functions MM corresponding to MM_{*} which are integrable in this way.

Proposition 6.2.
\thlabel

MstarIntegrabilityProp For MM as above, the following two statements are equivalent.

  1. (i)

    0δM(y)𝑑y<\int_{0}^{\delta}M_{*}(y)\,dy<\infty for some δ>0\delta>0.

  2. (ii)

    1(M(x))2𝑑x<\int_{1}^{\infty}(M^{\prime}(x))^{2}\,dx<\infty.

Proof.

We start with the integral in (i)(i) above and implement the change of variable y=M(x)y=M^{\prime}(x). Then dy=M′′(x)dxdy=M^{\prime\prime}(x)\,dx and M(y)=M(x)M_{*}(y)=M(x), and by next using integration by parts, we obtain

0δM(y)𝑑y\displaystyle\int_{0}^{\delta}M_{*}(y)\,dy =K(δ)M(x)M′′(x)𝑑x\displaystyle=-\int_{K(\delta)}^{\infty}M(x)M^{\prime\prime}(x)\,dx
=limR+M(R)M(R)+M(K(δ))M(K(δ))\displaystyle=\lim_{R\to+\infty}-M(R)M^{\prime}(R)+M(K(\delta))M^{\prime}(K(\delta))
+limR+K(δ)R(M(x))2𝑑x.\displaystyle+\lim_{R\to+\infty}\int_{K(\delta)}^{R}(M^{\prime}(x))^{2}\,dx.

The assumptions (6.1) and (6.3) together imply that the quantity M(R)M(R)M^{\prime}(R)M(R) stays bounded, as R+R\to+\infty. Thus the desired equivalence follows from the above computation. ∎

Remark 6.3.
\thlabel

ConcFuncRemark We should point out that a concave function indeed exists which satisfies all of our conditions. For instance, note that any concave function MM which for large positive xx coincides with

M(x):=xlogxM(x):=\frac{\sqrt{x}}{\log x}

satisfies the equivalent conditions of the above proposition. Indeed, one verifies by differentiation that the right-hand side is concave on some interval [A,)[A,\infty) and that the integral in (ii)(ii) of \threfMstarIntegrabilityProp converges. Such a function MM 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.
\thlabel

growthestimateUpperHalfplane Let f1()2()f\in\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) have a Fourier transform f^\widehat{f} satisfying

+|f^|2ρ𝑑ζC\int_{\mathbb{R}_{+}}|\widehat{f}|^{2}\rho\,d\zeta\leq C

for some constant C>0C>0 and ρ(ζ)=eM(ζ)\rho(\zeta)=e^{-M(\zeta)}. Then there exist a positive constant δ\delta such that the analytic extension of ff to \mathbb{H} satisfies the estimate

|f(x+iy)|2CeM(y)y,y(0,δ).|f(x+iy)|\leq\sqrt{2C}\frac{e^{M_{*}(y)}}{y},\quad y\in(0,\delta).
Proof.

By developments of Section 2.1, we have

f(z)=f(t)tzdt2πi=f(t)ψz(t)¯𝑑λ(t),z=x+iy,f(z)=\int_{\mathbb{R}}\frac{f(t)}{t-z}\frac{dt}{2\pi i}=\int_{\mathbb{R}}f(t)\overline{\psi_{z}(t)}d\lambda(t),\quad z=x+iy\in\mathbb{H},

and where ψz2()\psi_{z}\in\mathcal{H}^{2}(\mathbb{R}) is as in \threfFourierTransformCauchyKernel. By Plancherel’s theorem and the lemma, we obtain

f(z)\displaystyle f(z) =+f^(ζ)eizζ𝑑λ(ζ)\displaystyle=\int_{\mathbb{R}_{+}}\widehat{f}(\zeta)e^{iz\zeta}\,d\lambda(\zeta)
=+f^(ζ)ρ(ζ)eyζ+ixζρ(ζ)𝑑λ(ζ)\displaystyle=\int_{\mathbb{R}_{+}}\widehat{f}(\zeta)\sqrt{\rho(\zeta)}\frac{e^{-y\zeta+ix\zeta}}{\sqrt{\rho(\zeta)}}\,d\lambda(\zeta)

An application of Cauchy-Schwarz inequality leads to

(6.6) |f(z)|C0eM(ζ)2yζ𝑑ζ.|f(z)|\leq\sqrt{C}\sqrt{\int_{0}^{\infty}e^{M(\zeta)-2y\zeta}\,d\zeta}.

Now \threfIMestimate applies to obtain the desired estimate. ∎

6.3. Construction of the compact set

If MM satisfies the equivalent conditions of \threfMstarIntegrabilityProp, then the logarithm of the right-hand side in the inequality of \threfgrowthestimateUpperHalfplane, namely

(6.7) H(y):=log(2C)2+M(y)logy,y(0,δ),H(y):=\frac{\log(2C)}{2}+M_{*}(y)-\log y,\quad y\in(0,\delta),

is, for small enough δ>0\delta>0, positive and integrable over the interval y(0,δ)y\in(0,\delta). To HH and any A>0A>0 we will associate a Cantor-type compact set EE contained in [0,A][0,A] which contains no intervals and for which the integral

(6.8) [0,A]EH(dist(x,E))𝑑x\int_{[0,A]\setminus E}H(\text{dist}(x,E))\,dx

converges. Here dist(x,E)\text{dist}(x,E) denotes the distance from the point x[0,A]x\in[0,A] to the closed set EE. Let 𝒰={}\mathcal{U}=\{\ell\} be the system of maximal disjoint open intervals, union of which constitutes the complement of EE within (0,A)(0,A). The convergence of the integral above is easily seen to be equivalent to the convergence of the sum

(6.9) 𝒰0||H(x)𝑑x\sum_{\ell\in\mathcal{U}}\int_{0}^{|\ell|}H(x)\,dx

where |||\ell| denotes the length of the interval \ell. To construct EE, we choose a sequence of numbers {Ln}n1\{L_{n}\}_{n\geq 1} which is so quickly decreasing that

(6.10) n=12n0LnH(x)𝑑x<\sum_{n=1}^{\infty}2^{n}\int_{0}^{L_{n}}H(x)\,dx<\infty

and

(6.11) n=12nLnA/2.\sum_{n=1}^{\infty}2^{n}L_{n}\leq A/2.

From such a sequence, we construct EE as in the classical Cantor set construction. We set E0=[0,A]E_{0}=[0,A], and recursively define a compact set En+1E_{n+1} contained in EnE_{n}. The set En+1E_{n+1} consists of 2n+12^{n+1} closed intervals in [0,A][0,A] which we obtain by removing from the 2n2^{n} closed intervals {En,i}i=12n\{E_{n,i}\}_{i=1}^{2^{n}} constituting EnE_{n} an open interval of length LnL_{n} lying in the middle of En,iE_{n,i}. Thus splitting each En,iE_{n,i} into two new closed intervals. The above summation condition (6.11) ensures that |En|>A/2|E_{n}|>A/2, and so E:=n=0EnE:=\cap_{n=0}^{\infty}E_{n} has positive Lebesgue measure which is not less than A/2A/2. The integral condition (6.8) holds by its equivalence to (6.9) and by (6.10). Clearly EE contains no intervals.

6.4. Collapse of the Fourier transforms

We are now ready to prove \threfclaim3. We will do so by showing that (𝟙E,ρ)\mathcal{H}(\mathbbm{1}_{E},\rho) does not contain any non-zero tuple of the form (0,k)(0,k), k2(+,ρdx)k\in\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx), where MM is as in \threfConcFuncRemark, for instance, and where EE is as in Section 6.9. We set ρ=eM\rho=e^{-M}.

Lemma 6.5.
\thlabel

FourierCollapseLemma Let EE, MMand ρ\rho be chosen as above. Assume that {fn}n\{f_{n}\}_{n} is a sequence of functions in 1()2()\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}), each of which has an analytic extension to a half-space larger than \mathbb{H}. If limnfn=0\lim_{n\to\infty}f_{n}=0 in the norm of 2(E,𝟙Edx)\mathcal{L}^{2}(E,\mathbbm{1}_{E}\,dx) and the sequence of Fourier transforms {fn^}n\{\widehat{f_{n}}\}_{n} satisfies

supn+|fn^|2ρ𝑑ζ<,\sup_{n}\int_{\mathbb{R}_{+}}|\widehat{f_{n}}|^{2}\rho\,d\zeta<\infty,

then we have

limnfn(z)=0,z.\lim_{n\to\infty}f_{n}(z)=0,\quad z\in\mathbb{H}.

The convergence is uniform on compact subsets of \mathbb{H}.

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 𝒰={}\mathcal{U}=\{\ell\} be the collection of finite open intervals complementary to EE, and let

T:={x+iy:x,ydist(x,E)}T_{\ell}:=\{x+iy\in\mathbb{H}:x\in\ell,y\leq\text{dist}(x,E)\}

be a triangle with base at \ell. We define Ω=R(𝒰T)\Omega=R\setminus\big{(}\cup_{\ell\in\mathcal{U}}T_{\ell}\big{)} to be the bounded domain in \mathbb{H} which consists of a rectangle RR, with a base being the shortest closed interval containing the set EE, with the triangles TT_{\ell} removed from RR. See Figure 1. An observation that Khrushchev made regarding this type of domains is the following property of their harmonic measure.

Lemma 6.6.
\thlabel

KhruschevsEstimateLemma Let EE, MM and ρ\rho be chosen as above, and let HH be given by (6.7). Let Ω\Omega be the domain described above. If ωz\omega_{z} is the harmonic measure of the domain Ω\Omega at any point zΩz\in\Omega, then

ΩH(Imt)𝑑ωz(t)<.\int_{\partial\Omega\cap\mathbb{H}}H(\operatorname{Im}t)\,d\omega_{z}(t)<\infty.

We emphasize that Ω\partial\Omega\cap\mathbb{H} equals Ω\partial\Omega\setminus\mathbb{R}.

Proof.

The proof is very similar to the one given by Khrushchev in [8], only minor details differ. If =(a,b)\ell=(a,b) is one of the finite intervals complementary to EE, and TT_{\ell} is the triangle standing on top of it, then we denote by A(s)A(s) the part of the boundary of TT_{\ell} which lies above the interval (a,a+s)(a,a+s)\subset\mathbb{R}, 0<s<||/20<s<|\ell|/2. If uu is the harmonic measure in \mathbb{H} of the interval (a,a+s)(a,a+s), then it is easy to see from the explicit formula

u(x+iy)=1πaa+sy(xt)2+y2𝑑t,x+iyu(x+iy)=\frac{1}{\pi}\int_{a}^{a+s}\frac{y}{(x-t)^{2}+y^{2}}\,dt,\quad x+iy\in\mathbb{H}

that u(x+iy)12πu(x+iy)\geq\frac{1}{2\pi} for x+iyA(s)x+iy\in A(s). Since uu is harmonic and continuous in the closure of Ω\Omega except possibly at the two points aa and a+sa+s, the reproducing formula Ωu𝑑ωz=u(z)\int_{\partial\Omega}u\,d\omega_{z}=u(z) holds, and so

ωz(A(s))=A(s)𝑑ωzΩ2πu𝑑ωz=2πu(z)2sy,z=x+iy.\omega_{z}(A(s))=\int_{A(s)}d\omega_{z}\leq\int_{\partial\Omega}2\pi u\,d\omega_{z}=2\pi u(z)\leq\frac{2s}{y},\quad z=x+iy\in\mathbb{H}.

We have used the positivity of uu and ωz\omega_{z} in the first inequality, and the second one is an easy consequence of the explicit formula for uu above. Set A1:=A(||/2)A_{1}:=A(|\ell|/2), which is the left side of the boundary of TT_{\ell}, and further set An:=A(||/2n)A_{n}:=A(|\ell|/2^{n}), n1n\geq 1. Then, since HH is decreasing,

TH(Imt)𝑑ωz(t)\displaystyle\int_{\partial T_{\ell}\cap\mathbb{H}}H(\operatorname{Im}t)d\omega_{z}(t) =2A1H(Imt)𝑑ωz(t)\displaystyle=2\int_{A_{1}}H(\operatorname{Im}t)\,d\omega_{z}(t)
2n=1AnAn+1H(||/2n+1)𝑑ωz(t)\displaystyle\leq 2\sum_{n=1}^{\infty}\int_{A_{n}-A_{n+1}}H(|\ell|/2^{n+1})\,d\omega_{z}(t)
2n=1H(||/2n+1)2||2ny\displaystyle\leq 2\sum_{n=1}^{\infty}H(|\ell|/2^{n+1})\frac{2|\ell|}{2^{n}y}
16y0||H(t)𝑑t.\displaystyle\leq\frac{16}{y}\int_{0}^{|\ell|}H(t)\,dt.

Now the desired claim follows from (6.9). ∎

\includestandalone

[scale=1]OmegaDomain

Figure 1. The domain Ω\Omega in the proof of \threfFourierCollapseLemma. There is a triangular tent between Ω\Omega and each complementary interval of EE, and EE lives on \mathbb{R} inbetween the tents.
Proof of \threfFourierCollapseLemma .

Note that it is sufficient to establish the claim that the sequence {fn}n\{f_{n}\}_{n} contains a subsequence which converges pointwise in Ω\Omega to 0. Indeed, the proof of \threfgrowthestimateUpperHalfplane shows that our assumption on the Fourier transforms f^\widehat{f} implies pointwise boundedness of the sequence {fn}n\{f_{n}\}_{n} on each half-plane {x+iy:y>δ}\{x+iy\in\mathbb{H}:y>\delta\}, δ>0\delta>0. Hence the sequence {fn}n\{f_{n}\}_{n} forms a normal family on \mathbb{H}. If we establish the above claim, then every subsequences of {fn}n\{f_{n}\}_{n} contains a further subsequence convergent to 0 in \mathbb{H}. This is equivalent to convergence of the entire initial sequence {fn}n\{f_{n}\}_{n} to 0.

Fix zΩz\in\Omega. Since log|fn|\log|f_{n}| is a subharmonic function and max(N,log|fn|)\max(-N,\log|f_{n}|) is a bounded continuous function on Ω\partial\Omega, we obtain by the maximum principle for subharmonic functions that

log|fn(z)|Ωmax(N,log|fn(t)|)𝑑ωz(t).\log|f_{n}(z)|\leq\int_{\partial\Omega}\max(-N,\log|f_{n}(t)|)\,d\omega_{z}(t).

We let N+N\to+\infty and, by the monotone convergence theorem, obtain

log|fn(z)|\displaystyle\log|f_{n}(z)| Ωlog|fn(t)|dωz(t)\displaystyle\leq\int_{\partial\Omega}\log|f_{n}(t)|\,d\omega_{z}(t)
=Ωlog|fn(t)|dωz(t)+Elog|fn(t)|dωz(t)\displaystyle=\int_{\partial\Omega\cap\mathbb{H}}\log|f_{n}(t)|\,d\omega_{z}(t)+\int_{E}\log|f_{n}(t)|\,d\omega_{z}(t)

The assumption in the lemma, \threfgrowthestimateUpperHalfplane, the definition of HH in (6.7) and \threfKhruschevsEstimateLemma show that

Ωlog|fn(t)|dωz(t)ΩH(Imt)𝑑ωz(t)<A\int_{\partial\Omega\cap\mathbb{H}}\log|f_{n}(t)|\,d\omega_{z}(t)\leq\int_{\partial\Omega\cap\mathbb{H}}H(\operatorname{Im}t)d\omega_{z}(t)<A

where AA is some positive constant which is independent of nn. By Egorov’s theorem, we may pass to a subsequence (the same subsequence for each zΩ)z\in\Omega) and assume that fnf_{n} converge uniformly to 0 on some subset EE^{\prime} of EE which is of positive Lebesgue measure. On EEE\setminus E^{\prime} we have the estimate

EElog|fn(t)|dωz(t)EE|fn(t)|2𝑑ωz(t)1πImzEE|fn(t)|2𝑑x.\int_{E\setminus E^{\prime}}\log|f_{n}(t)|\,d\omega_{z}(t)\leq\int_{E\setminus E^{\prime}}|f_{n}(t)|^{2}\,d\omega_{z}(t)\leq\frac{1}{\pi\operatorname{Im}z}\int_{E\setminus E^{\prime}}|f_{n}(t)|^{2}\,dx.

The last inequality follows from monotonicity of the harmonic measure with respect to domains (see [14, Corollary 4.3.9]), which applied to Ω\Omega\subset\mathbb{H} leads to the inequality

ωz(B)B𝒫(t,z)𝑑t|B|πImz\omega_{z}(B)\leq\int_{B}\mathcal{P}(t,z)\,dt\leq\frac{|B|}{\pi\operatorname{Im}z}

for any Borel subset BB of EE. Thus dωzdxπImzd\omega_{z}\leq\frac{dx}{\pi\operatorname{Im}z}, attesting the integral inequality above. By the convergence of fnf_{n} to 0 in the norm of 2(E,𝟙Edx)\mathcal{L}^{2}(E,\mathbbm{1}_{E}\,dx), the integrals EE|fn|2𝑑x\int_{E\setminus E^{\prime}}|f_{n}|^{2}dx are uniformly bounded by some constant C>0C>0, and so the above inequalities give

log|fn(z)|A+C+Elog|fn(t)|dωz(t).\log|f_{n}(z)|\leq A+C+\int_{E^{\prime}}\log|f_{n}(t)|d\omega_{z}(t).

But ωz(E)>0\omega_{z}(E^{\prime})>0, since the harmonic measure and the arc-length measure on the rectifiable curve Ω\partial\Omega are mutualy absolutely continuous (see [5, Theorem 1.2 of Chapter VI]), so since |fn||f_{n}| converge uniformly to 0 on EE^{\prime}, the integral on the right-hand side above converges to -\infty as nn\to\infty. Thus |fn(z)|0|f_{n}(z)|\to 0, and since zΩz\in\Omega was arbitrary, the desired claim follows. ∎

The following proposition implies \threfclaim3 of Section 4, and so it also implies \threfSparsenessTheorem.

Proposition 6.7.
\thlabel

FourierCollapseProposition Let EE, MMand ρ\rho be chosen as above. If a tuple of the form (0,k)2(E,𝟙Edx)2(+,ρ)(0,k)\in\mathcal{L}^{2}(E,\mathbbm{1}_{E}\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho) is contained in the Hardy subspace (𝟙E,ρ)\mathcal{H}(\mathbbm{1}_{E},\rho), then k0k\equiv 0.

Proof.

By the containment (0,k)(𝟙E,ρ)(0,k)\in\mathcal{H}(\mathbbm{1}_{E},\rho) and \threfKernelContainmentHardySubspace, there exists a sequence {fn}n\{f_{n}\}_{n} of functions in 1()2()\mathcal{H}^{1}(\mathbb{R})\cap\mathcal{H}^{2}(\mathbb{R}) extending analytically across \mathbb{R} and for which the tuples Jfn=(fn,f^n)Jf_{n}=(f_{n},\widehat{f}_{n}) converge in the norm of the product space 2(,𝟙Edx)2(+,ρdx)\mathcal{L}^{2}(\mathbb{R},\mathbbm{1}_{E}\,dx)\oplus\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx) to (0,k)(0,k). By passing to a subsequence, we may assume that the Fourier transforms fn^\widehat{f_{n}} converge pointwise almost everywhere on +\mathbb{R}_{+} to kk. One might attempt to prove the proposition by using the formula in \threfH1FourierTransformFormula, and observing that

k(ζ)=limnfn^(ζ)=limneyζfn(x+iy)eixζ𝑑λ(x)k(\zeta)=\lim_{n\to\infty}\widehat{f_{n}}(\zeta)=\lim_{n\to\infty}e^{y\zeta}\int_{\mathbb{R}}f_{n}(x+iy)e^{-ix\zeta}\,d\lambda(x)

holds for almost every ζ+\zeta\in\mathbb{R}_{+} and for any y>0y>0. By \threfFourierCollapseLemma the integrand converges pointwise to 0. However, an appeal to the usual convergence theorems for integrals is not justified, and we have to proceed more carefully. Note that since k2(+,ρdx)k\in\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx) and ρ\rho is bounded from below on compact subsets of +\mathbb{R}_{+}, in fact kk is locally integrable on \mathbb{R}. It follows that we can interpret kk as a distribution on +\mathbb{R}_{+}. Thus to show that k0k\equiv 0, it suffices to establish that +kϕ𝑑λ=0\int_{\mathbb{R}_{+}}k\phi\,d\lambda=0 for every smooth function ϕ\phi which is compactly supported in +\mathbb{R}_{+}.

Let ϕ\phi be as above. Since we have that fn^k\widehat{f_{n}}\to k in 2(+,ρdx)\mathcal{L}^{2}(\mathbb{R}_{+},\rho\,dx) and ρ\rho is bounded from below on compact subsets of +\mathbb{R}_{+}, we obtain

(6.12) +kϕ𝑑λ=limn+fn^ϕ𝑑λ.\int_{\mathbb{R}_{+}}k\phi\,d\lambda=\lim_{n\to\infty}\int_{\mathbb{R}_{+}}\widehat{f_{n}}\phi\,d\lambda.

Fix some small y>0y>0. By \threfH1FourierTransformFormula, we get that

fn^(ζ)=eyζfn(x+iy)eixζ𝑑λ(x).\widehat{f_{n}}(\zeta)=e^{y\zeta}\int_{\mathbb{R}}f_{n}(x+iy)e^{-ix\zeta}\,d\lambda(x).

Plugging this formula into (6.12) and noting that the use of Fubini’s theorem is permitted, we obtain

+kϕ𝑑λ\displaystyle\int_{\mathbb{R}_{+}}k\phi\,d\lambda =limn(+ϕ(ζ)eyζeixζ𝑑λ(ζ))fn(x+iy)𝑑λ(x)\displaystyle=\lim_{n\to\infty}\int_{\mathbb{R}}\Big{(}\int_{\mathbb{R}_{+}}\phi(\zeta)e^{y\zeta}e^{-ix\zeta}d\lambda(\zeta)\Big{)}f_{n}(x+iy)\,d\lambda(x)
(6.13) =limnD(x)fn(x+iy)𝑑λ(x)\displaystyle=\lim_{n\to\infty}\int_{\mathbb{R}}D(x)f_{n}(x+iy)\,d\lambda(x)

where

D(x):=+ϕ(ζ)eyζeixζ𝑑λ(ζ)D(x):=\int_{\mathbb{R}_{+}}\phi(\zeta)e^{y\zeta}e^{-ix\zeta}d\lambda(\zeta)

is the Fourier transform of the compactly supported smooth function ζϕ(ζ)eyζ\zeta\mapsto\phi(\zeta)e^{y\zeta}. As such, DD is certainly integrable on \mathbb{R}. By \threfFourierCollapseLemma, we have limnfn(x+iy)=0\lim_{n\to\infty}f_{n}(x+iy)=0, and

supnsupx|fn(x+iy)|<\sup_{n}\,\sup_{x\in\mathbb{R}}|f_{n}(x+iy)|<\infty

holds by \threfgrowthestimateUpperHalfplane. Therefore, this time, the dominated convergence theorem applies to (6.13), and we conclude that

+kϕ𝑑λ=0.\int_{\mathbb{R}_{+}}k\phi\,d\lambda=0.

Thus kk is the zero distribution on +\mathbb{R}_{+}, and therefore k0k\equiv 0. ∎

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 ff be a function which satisfies

(7.1) |f(x)|(1+|x|)n𝑑x<\int_{\mathbb{R}}\frac{|f(x)|}{(1+|x|)^{n}}\,dx<\infty

for some positive integer nn. Then ff can be interpreted as a tempered distribution on \mathbb{R} in the usual way, and so ff has a distributional Fourier transform f^\widehat{f}. Our hypothesis is that f^\widehat{f} is an integrable function on some half-axis [ζ0,)[\zeta_{0},\infty) and that

(7.2) ρf^(ζ)=𝒪(ecζ),ζ>ζ0.\rho_{\widehat{f}}(\zeta)=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)},\quad\zeta>\zeta_{0}.

We may assume that ζ0=0\zeta_{0}=0. In order to prove \threfDistributionalClumpingTheorem, we will construct an appropriate multiplier m:m:\mathbb{R}\to\mathbb{C} with the property that mfmf is a function to which \threfCondensationTheorem applies. In particular, the following properties will be satisfied by mm:

  1. (i)

    m(x)m(x) is a bounded function of xx\in\mathbb{R} which is non-zero for almost every xx\in\mathbb{R}.

  2. (ii)

    mf2(,dx)mf\in\mathcal{L}^{2}(\mathbb{R},dx),

  3. (iii)

    ρmf^(ζ)=𝒪(ecζ)\rho_{\widehat{mf}}(\zeta)=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)} for some c>0c>0 and ζ>0\zeta>0,

  4. (iv)

    Ilog|m|dx>\int_{I}\log|m|\,dx>-\infty for every interval II\subset\mathbb{R}.

If we construct such a multiplier mm, then (ii)(ii), (iii)(iii) and \threfCondensationTheorem imply that log|mf|\log|mf| is locally integrable on an open set UU which coincides, up to a set of measure zero, with {x:|f(x)m(x)|>0}\{x\in\mathbb{R}:|f(x)m(x)|>0\}. By (i)(i), UU differs from {x:|f(x)|>0}\{x\in\mathbb{R}:|f(x)|>0\} at most by a set of measure zero. Moreover, the formula log|f|=log|fm|log|m|\log|f|=\log|fm|-\log|m| and (iv)(iv) show that log|f|\log|f| is locally integrable on UU. 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

Φ(x):=(i)nn!2π(xi)n\Phi(x):=\frac{(-i)^{n}n!}{\sqrt{2\pi}(x-i)^{n}}

and let hh be defined by the equation (2.7), with

log|h(x)|=logmin(1,|f(x)|1),x.\log|h(x)|=\log\min(1,|f(x)|^{-1}),\quad x\in\mathbb{R}.

The condition (7.1) ensures that hh is well-defined, and it is a member of ()\mathcal{H}^{\infty}(\mathbb{R}). We put

h(x):=h(x)(x+i)2,x.h_{*}(x):=\frac{h(x)}{(x+i)^{2}},\quad x\in\mathbb{R}.

and finally

m(x):=Φ(x)h(x)¯,x.m(x):=\Phi(x)\overline{h_{*}(x)},\quad x\in\mathbb{R}.

Clearly, mm is bounded. Since |f||f| is locally integrable on \mathbb{R}, the set {x:|f|=}\{x\in\mathbb{R}:|f|=\infty\} has measure zero. Consequently, |h|>0|h_{*}|>0 almost everywhere in \mathbb{R}, and so the desired property (i)(i) of mm holds. The choice of hh_{*} and Φ\Phi ensures that mfmf is both bounded and integrable on \mathbb{R}, implying mf2(,dx)mf\in\mathcal{L}^{2}(\mathbb{R},dx), so that (ii)(ii) above holds. Property (iv)(iv) holds by \threfHadyClassLogIntProp, since h1()h_{*}\in\mathbb{H}^{1}(\mathbb{R}). So the critical property left to be verified is the spectral estimate of mfmf in (iii)(iii) above.

Lemma 7.1.

With notation and definitions as above, the Fourier transform mf^\widehat{mf} satisfies

ρmf^(ζ)=𝒪(ecζ),ζ>0\rho_{\widehat{mf}}(\zeta)=\mathcal{O}\big{(}e^{-c\sqrt{\zeta}}\big{)},\quad\zeta>0

for some c>0c>0.

Proof.

A standard argument shows that fΦ^\widehat{f\Phi} must coincide on +\mathbb{R}_{+} with the convolution f^Φ^\widehat{f}\ast\widehat{\Phi} (which, note, is a function on +\mathbb{R}_{+}). Indeed, let ss be a Schwartz function which has a Fourier transform s^\widehat{s} supported on some compact interval [a,b][a,b], 0<a<b0<a<b. Note that the function Φs\Phi s is also of Schwartz class. It follows immediately from the integral definition of the Fourier transform (1.1) that Φ¯s^=Φ¯^s^\widehat{\overline{\Phi}s}=\widehat{\overline{\Phi}}\ast\widehat{s}, and that Φ¯s^\widehat{\overline{\Phi}s} is supported on the interval [a,)[a,\infty). Hence, by the definition of the distributional Fourier transform, we obtain

+fΦ^s^¯𝑑λ=fΦs¯𝑑λ=fΦ¯s¯𝑑λ=+f^(Φ¯^s^)¯𝑑λ.\int_{\mathbb{R}_{+}}\widehat{f\Phi}\,\overline{\widehat{s}}\,d\lambda=\int_{\mathbb{R}}f\Phi\overline{s}\,d\lambda=\int_{\mathbb{R}}f\overline{\overline{\Phi}s}\,d\lambda=\int_{\mathbb{R}_{+}}\widehat{f}\overline{(\widehat{\overline{\Phi}}\ast\widehat{s})}\,d\lambda.

Fubini’s theorem and the computational rule Φ¯^(x)=Φ^¯(x)\widehat{\overline{\Phi}}(x)=\overline{\widehat{\Phi}}(-x) shows that the last integral above equals

+(f^Φ^)s^¯𝑑λ,\int_{\mathbb{R}_{+}}(\widehat{f}\ast\widehat{\Phi})\,\overline{\widehat{s}}\,d\lambda,

proving our claim about the structure of fΦ^\widehat{f\Phi} on +\mathbb{R}_{+}.

Hence fΦ^\widehat{f\Phi} is a bounded continuous function which coincides with

fΦ^(ζ)=f^(x)Φ^(ζx)𝑑λ(x)\widehat{f\Phi}(\zeta)=\int_{\mathbb{R}}\widehat{f}(x)\widehat{\Phi}(\zeta-x)\,d\lambda(x)

for ζ>0\zeta>0. By a computation similar to the one in the proof of \threfFourierTransformCauchyKernel one sees that Φ\Phi has the Fourier transform

Φ^(ζ)=|ζ|neζ𝟙(ζ).\widehat{\Phi}(\zeta)=|\zeta|^{n}e^{\zeta}\mathbbm{1}_{\mathbb{R}_{-}}(\zeta).

For such ζ\zeta, we estimate

|fΦ^(ζ)|\displaystyle\big{|}\widehat{f\Phi}(\zeta)\big{|} |f^(x)||ζx|neζx𝟙(ζx)𝑑λ(x)\displaystyle\leq\int_{\mathbb{R}}|\widehat{f}(x)||\zeta-x|^{n}e^{\zeta-x}\mathbbm{1}_{\mathbb{R}_{-}}(\zeta-x)\,d\lambda(x)
=ζ|f^(x)||ζx|neζx𝑑λ(x)\displaystyle=\int_{\zeta}^{\infty}|\widehat{f}(x)||\zeta-x|^{n}e^{\zeta-x}\,d\lambda(x)
=k=0ζ2kζ2k+1|f^(x)||ζx|neζx𝑑λ(x).\displaystyle=\sum_{k=0}^{\infty}\int_{\zeta 2^{k}}^{\zeta 2^{k+1}}|\widehat{f}(x)||\zeta-x|^{n}e^{\zeta-x}\,d\lambda(x).

We now make the rather rough estimate

|ζx|neζxζ2(k+1)n,x[ζ2k,ζ2k+1],|\zeta-x|^{n}e^{\zeta-x}\leq\zeta^{2(k+1)n},\quad x\in[\zeta 2^{k},\zeta 2^{k+1}],

which gives

|fΦ^(ζ)|k=0ρf^(ζ2k)ζ2(k+1)nAk=0ecζ2kζ2(k+1)n\big{|}\widehat{f\Phi}(\zeta)\big{|}\leq\sum_{k=0}^{\infty}\rho_{\widehat{f}}(\zeta 2^{k})\zeta^{2(k+1)n}\leq A\sum_{k=0}^{\infty}e^{-c\sqrt{\zeta}\sqrt{2}^{k}}\zeta^{2(k+1)n}

for some A>0A>0. The above sum can be readily estimated to be of order 𝒪(edζ)\mathcal{O}\big{(}e^{-d\sqrt{\zeta}}) for some d>0d>0 slightly smaller than cc. Since fmfm is the product of two integrable functions fΦf\Phi and h¯\overline{h_{*}}, we have

fm^=fΦ^h¯^,\widehat{fm}=\widehat{f\Phi}\ast\widehat{\overline{h_{*}}},

where h¯^(ζ)=h^(ζ)¯\widehat{\overline{h_{*}}}(\zeta)=\overline{\widehat{h_{*}}(-\zeta)} is non-zero only for ζ<0\zeta<0. Note that hh_{*} is integrable on \mathbb{R}, and so h^\widehat{h_{*}} is bounded. We obtain

|fm^(ζ)|\displaystyle|\widehat{fm}(\zeta)| |fΦ^(x)h^(xζ)¯|𝑑λ(x)\displaystyle\leq\int_{\mathbb{R}}\Big{|}\widehat{f\Phi}(x)\overline{\widehat{h_{*}}(x-\zeta)}\Big{|}\,d\lambda(x)
Bζedx𝑑λ(x)=𝒪(edζ),\displaystyle\leq B\int_{\zeta}^{\infty}e^{-d\sqrt{x}}\,d\lambda(x)=\mathcal{O}\big{(}e^{-d\sqrt{\zeta}}\big{)},

where BB is some positive constant. The desired estimate on ρfm^\rho_{\widehat{fm}} 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 lpl_{p}-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.