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

On Gaussian decay rates of harmonic oscillators and equivalences of related Fourier uncertainty principles

Aleksei Kulikov Department of Mathematical Sciences, Norwegian University of Science and Technology, NO-7491 Trondheim, Norway lyosha.kulikov@mail.ru Lucas Oliveira Departamento de Matemática
UFRGS
Porto Alegre RS 91509-900
lucas.oliveira@ufrgs.br
 and  João P. G. Ramos ETH Zürich D-MATH
Rämistrasse 101, 8092 Zürich, Switzerland
joao.ramos@math.ethz.ch
Abstract.

We make progress on a question by Vemuri on the optimal Gaussian decay of harmonic oscillators, proving the original conjecture up to an arithmetic progression of times. The techniques used are a suitable translation of the problem at hand in terms of the free Schrödinger equation, the machinery developed in the work of Cowling, Escauriaza, Kenig, Ponce and Vega [CEK+10], and a lemma which relates decay on average to pointwise decay.

Such a lemma produces many more consequences in terms of equivalences of uncertainty principles. Complementing such results, we provide endpoint results in particular classes induced by certain Laplace transforms, both to the decay Lemma and to the remaining cases of Vemuri’s conjecture, shedding light on the full endpoint question.

1. Introduction

1.1. Historical Background

Uncertainty principles have permeated Mathematics and Physics for many years, since the introduction of such a concept by Heisenberg in the context of Quantum Mechanics. For the Fourier transform

(1.1) f^(ξ)=e2πixξf(x)𝑑x,\widehat{f}(\xi)=\int_{\mathbb{R}}e^{-2\pi ix\cdot\xi}f(x)\,dx\,,

Heisenberg’s uncertainty principle can be stated simply as

|f(x)|2𝑑x4π(|x|2|f(x)|2𝑑x)1/2(|ξ|2|f^(ξ)|2𝑑ξ)1/2.\int_{\mathbb{R}}|f(x)|^{2}dx\leq 4\pi\left(\int_{\mathbb{R}}|x|^{2}|f(x)|^{2}dx\right)^{1/2}\left(\int_{\mathbb{R}}|\xi|^{2}|\widehat{f}(\xi)|^{2}d\xi\right)^{1/2}.

This inequality states essentially that we cannot concentrate in space and frequency sides too much simultaneously, and has the physical interpretation that we cannot make measurements about the position and momentum of a particle (in the probabilistic sense) with high precision for both.

Further than Heisenberg’s initial contribution, there are many other instances and kinds of uncertainty principles. Benedicks’s uncertainty principle [Ben85], for example, predicts that for fL1(d)f\in L^{1}(\mathbb{R}^{d}) the measures of the sets {xd:f(x)0}\{x\in\mathbb{R}^{d}:f(x)\neq 0\} and {xd:f^(x)0}\{x\in\mathbb{R}^{d}:\widehat{f}(x)\neq 0\} cannot be both finite, unless f0f\equiv 0. The Amrein–Berthier Uncertainty Principle [AB77] complements the previous one stating that for any fL2(d)f\in L^{2}(\mathbb{R}^{d}) and any pair of finite measure sets E,FdE,F\subseteq\mathbb{R}^{d} there is a positive constant C=C(E,F)C=C(E,F) such that

fL22=C(Ec|f(x)|2𝑑x+Fc|f^(x)|2𝑑x).\|f\|_{L^{2}}^{2}=C\left(\int_{E^{c}}|f(x)|^{2}dx+\int_{F^{c}}|\widehat{f}(x)|^{2}dx\right).

Recently, a different kind of uncertainty principle related to sign changes of the Fourier transform has attracted some attention. For example, in [BCK10] Bourgain, Clozel and Kahane had proved that ff and f^\widehat{f} cannot simultaneously concentrate negative mass on arbitrarily small neighbourhoods of the origin. For further developments in this direction, see [CG19, GOeSR21, GOeSR20, GOeSS17] and references therein.

In this work we are concerned with uncertainty principles that are, to some extent, related to the properties of the Gaussians. The first such result was obtained by Hardy [Har33] in 1933 and can be stated in the following way: if |f(x)|Aπx2|f(x)|\leq A^{-\pi x^{2}} and |f^(x)|Beπx2|\widehat{f}(x)|\leq Be^{-\pi x^{2}}, then there is a constant CC such that f(x)=f^(x)=Ceπx2f(x)=\widehat{f}(x)=Ce^{-\pi x^{2}}. In fact, Hardy proved more than this:

  • If ff and f^\widehat{f} are of order O(|x|meπx2)O(|x|^{m}e^{-\pi x^{2}}) for some mm and for large xx, then ff is a linear combination of Hermite functions;

  • If ff is O(eπx2)O(e^{-\pi x^{2}}) and f^\widehat{f} is o(eπx2)o(e^{-\pi x^{2}}) (or vice-versa), then f=0f=0.

On the other hand, it is not enough to assume that |f(x)|Aeλπx2|f(x)|\leq Ae^{-\lambda\pi x^{2}} and |f^(ξ)|Aeμπξ2|\widehat{f}(\xi)|\leq Ae^{-\mu\pi\xi^{2}} for some 0<λ,μ<10<\lambda,\mu<1 since nontrivial functions ff satisfying these conditions form an infinite dimensional space.

The techniques coming from Complex Analysis (to be precise, the Phrmagmén-Lindelöf principle) were decisive in the proof of the above result, as well as in the proof of the following extension of it obtained by Beurling in 1964 (whose proof seems to have been lost, until Hörmander [H9̈1] in 1991 provided a full proof, based on personal notes taken during a discussion of this result with Beurling himself): if fL1()f\in L^{1}(\mathbb{R}) is such that

B(f,f^):=|f(x)||f^(y)|e2π|xy|𝑑x𝑑y<,B(f,\widehat{f}):=\int_{\mathbb{R}}\int_{\mathbb{R}}|f(x)||\widehat{f}(y)|e^{2\pi|xy|}dx\,dy<\infty\,,

then f0f\equiv 0.

It is worth mentioning that interesting generalisations of this result, in an almost subcritical level, have been recently obtained by Bonami-Demange [BD06], Hedenmalm [Hed12] and by Gao [Gao16]; see also [H9̈1].

In a different direction, and also relevant to our current work, are the generalisations of Hardy’s uncertainty principle where we can combine different kinds of Gaussians and different control of LpL^{p}-norms, where Hardy’s theorem can be seen as an LL^{\infty}-norm version of a more general principle. Major contributions along these lines were obtained by Cowling and Price [CP83] and Morgan [Mor34].

In our context, Cowling–Price’s Uncertainty Principle can be stated in the following way: feax2Lp<\|fe^{ax^{2}}\|_{L^{p}}<\infty and febx2Lq<\|fe^{bx^{2}}\|_{L^{q}}<\infty imply f0f\equiv 0 when ab>π2ab>\pi^{2}. When ab<π2,ab<\pi^{2}, in the same way as in Hardy’s theorem, there are nontrivial examples of functions ff satisfying these conditions.

More recently, Hardy’s Uncertainty Principle has been shown to be related to the study of decay behaviour of evolution equations. Indeed, let us consider the question of uniqueness for solutions of Schrödinger evolutions of the kind

(1.2) iut+2ux2+F(u,u¯)=0.i\frac{\partial u}{\partial t}+\frac{\partial^{2}u}{\partial x^{2}}+F(u,\overline{u})=0.

In other terms, we are interested in determining when two solutions u1,u2u_{1},u_{2} of (1.2) coincide, given they are equal on a set S(0,+)×.S\subset(0,+\infty)\times\mathbb{R}. Escauriaza et al [EKPV06] extended such a study by observing that Hardy’s result may be reformulated as the property that, if a solution to

(1.3) iut+2ux2=0i\frac{\partial u}{\partial t}+\frac{\partial^{2}u}{\partial x^{2}}=0

has sufficient Gaussian decay at two different times, it must vanish identically. In line with this, exploring techniques and ideas based on convexity of solutions of Schrödinger equations such as (1.2) with additional Gaussian control, in [EKPV08b], Euscariaza et al observed that such solutions should satisfy a weak version of Hardy’s Uncertainty Principle. In [CEK+10], Cowling et al showed a real variable proof of Hardy’s and Cowling-Price’s Uncertainty Principles. Their result may be summarised as follows: if v(x,t)v(x,t) is a solution of the free Schrödinger equation (1.3), with initial condition v(x,0)=O(eαx2),v(x,0)=O(e^{-\alpha x^{2}}), and such that v(x,T)=O(eβx2)v(x,T)=O(e^{-\beta x^{2}}) for some T>0T>0 with Tαβ>π2T\alpha\beta>\pi^{2}, then v0v\equiv 0.

1.2. Main results

We were able to show that LpL^{p}-bounds on a function and its Fourier transform imply pointwise bounds up to an ϵ\epsilon in the exponent. For example, we prove that if

n|f(x)|2e2πα|x|2𝑑x< and n|f^(x)|2e2πα|x|2𝑑x<\int_{\mathbb{R}^{n}}|f(x)|^{2}e^{2\pi\alpha|x|^{2}}\,dx<\infty\mbox{ and }\int_{\mathbb{R}^{n}}|\widehat{f}(x)|^{2}e^{2\pi\alpha|x|^{2}}\,dx<\infty

then, for each ϵ>0\epsilon>0 there is a positive constant A=A(ϵ)A=A(\epsilon) such that

|f(x)|Aϵe(1ϵ)aπ|x|2,|f(x)|\leq A_{\epsilon}e^{-(1-\epsilon)a\pi|x|^{2}},

in particular this implies that Cowling–Price’s Uncertainty Principle follows from the Hardy’s one.

This result is inspired by an attempt to attack a conjecture of Vemuri [Vem08] about the decay of solutions of the quantum harmonic oscillator. For f,g:nf,g:\mathbb{R}^{n}\to\mathbb{C} we denote Cp(f,g)=fgLppC^{p}(f,g)=||fg||_{L^{p}}^{p} and for g(x)=e2πa|x|2g(x)=e^{2\pi a|x|^{2}} we write Cp(f,g)=Cap(f)C^{p}(f,g)=C^{p}_{a}(f). Consider the class of functions

(1.4) Eap(n)={f:n:Cap(f)< and Cap(f^)<}.E_{a}^{p}(n)=\{f:\mathbb{R}^{n}\to\mathbb{C}:\mid C^{p}_{a}(f)<\infty\mbox{ and }C^{p}_{a}(\hat{f})<\infty\}.

In terms of these spaces, the result we formulated in the beginning of this section can be restated in the following form: if fEa2(n)f\in E_{a}^{2}(n) then for all ϵ>0\epsilon>0 we have fEaϵ(n)f\in E_{a-\epsilon}^{\infty}(n).

Let H:=Δ+4π2|x|2H:=-\Delta+4\pi^{2}|x|^{2} denote the (normalised) Quantum Harmonic Oscillator. Fix fL2(n).f\in L^{2}(\mathbb{R}^{n}). We define Φ(x,t)\Phi(x,t) to be the solution of the time-dependent initial value problem

(1.5) {itΦ=HΦ,for(x,t)n×;Φ(x,0)=f(x),onn.\begin{cases}i\partial_{t}\Phi=H\Phi,\,\,&\text{for}\,(x,t)\in\mathbb{R}^{n}\times\mathbb{R};\cr\Phi(x,0)=f(x),\,\,&\text{on}\,\mathbb{R}^{n}.\end{cases}

The solution Φ\Phi to this problem is intimately related to the Hermite functions when n=1n=1. Indeed, if we have

f(x)=k0akhk(x),f(x)=\sum_{k\geq 0}a_{k}h_{k}(x),

then we may write the solution above at time tt\in\mathbb{R} as

Φ(x,t)=k0e2(2k+1)πitakhk(x),\Phi(x,t)=\sum_{k\geq 0}e^{2\cdot(2k+1)\pi it}a_{k}h_{k}(x),

where we define our normalisation of the Hermite functions {hk}k0\{h_{k}\}_{k\geq 0} to be the complete orthonormal system in L2L^{2} such that (hk)=(i)khk.\mathcal{F}(h_{k})=(-i)^{k}h_{k}. This formula for the solution converges in a pointwise sense for ff in the Schwartz space 𝒮()\mathcal{S}(\mathbb{R}). From now on, we will use the notation Φf(x,t)\Phi f(x,t) to denote the solution to (1.5) with initial value ff. Whenever it is obvious from context, we shall simply write Φ(x,t)\Phi(x,t) as above.

With these definitions, Vemuri’s conjecture [Vem08] states that, if fEtanh(2α)(n),f\in E^{\infty}_{\tanh(2\alpha)}(n), then

Φ(,t)Etanh(α)(n),t>0.\Phi(\cdot,t)\in E^{\infty}_{\tanh(\alpha)}(n),\,\forall t>0.

In fact, Vemuri proved that Φ(,t)Etanh(α)ϵ(n),ϵ>0.\Phi(\cdot,t)\in E^{\infty}_{\tanh(\alpha)-\epsilon}(n),\,\forall\epsilon>0.

By relating the evolution of the Harmonic Oscillator problem to the Schrödinger equation and the optimal decay for Schrödinger evolutions as in, for instance, [CEK+10], we obtain an L2L^{2}-version of Vemuri’s conjecture: if Ca2(f),Ca2(f^)<+,C_{a}^{2}(f),C_{a}^{2}(\widehat{f})<+\infty, then

Ctanh(α)2(Φ(,t))<+,C_{\tanh(\alpha)}^{2}(\Phi(\cdot,t))<+\infty,

where a=tanh(2α).a=\tanh(2\alpha). Our first main result is, as far as we know, the first step towards settling Vemuri’s conjecture in the original LL^{\infty} case.

Theorem 1.1.

Let fEtanh(2α)(n),f\in E^{\infty}_{\tanh(2\alpha)}(n), for some α>0.\alpha>0. Then Φ(,t)Etanh(α)(n)\Phi(\cdot,t)\in E^{\infty}_{\tanh(\alpha)}(n) whenever t{116+k8,k}.t\not\in\{\frac{1}{16}+\frac{k}{8},\,k\in\mathbb{Z}\}.

We will, in fact, prove that Vemuri’s conjecture can be sharpenned in the case t{116+k8,k}.t\not\in\{\frac{1}{16}+\frac{k}{8},\,k\in\mathbb{Z}\}. That is, the largest b>0b>0 for which Φ(,t)Eb\Phi(\cdot,t)\in E^{\infty}_{b} satisfies b>tanh(α),b>\tanh(\alpha), whenever tt is not in the exceptional set above.

The techniques used in order to prove Theorem 1.1 are based on several recent results in the literature involving Gaussian decay of Schrödinger equations. Indeed, we first make use of a change of variables which takes the evolution of the harmonic oscillator into that of the free Schrödinger equation. Although we provide an alternative proof of such Lemma, we note that this kind of formulas seems to be known in the physics literature; see, for instance, [Tak91a, Tak91b]. It was pointed to us recently that such changes of variables have also been employed in a similar context by B. Cassano and L. Fanelli in [CF17] (see also [CF15, BCF22] and the references therein).

We use a change of variables which preserves the free Schrödinger equation, in the same spirit as in [CEK+10], in order to be able to use the original results by Escauriaza, Kenig, Ponce and Vega on convexity properties of Gaussian decay of Schrödinger equations. Finally, the last technique used is the mechanism described above to pass from L2L^{2} to L,L^{\infty}, and vice versa.

It is worth to mention, though, that, in order to achieve such a result in higher dimensions, we will need a version of the Gaussian observation above for all dimensions. This is achieved through the following result:

Lemma 1.2.

Suppose that w:d[1,)w:\mathbb{R}^{d}\to[1,\infty) is a measurable function and f:df:\mathbb{R}^{d}\to\mathbb{C} is a CC^{\infty} function which are related by the following assumptions:

  • (i)

    For some 1p<1\leq p<\infty we have

    (1.6) n|f(x)|pw(x)p𝑑x<;\int_{\mathbb{R}^{n}}|f(x)|^{p}w(x)^{p}\,dx<\infty\,;
  • (ii)

    The sets {x:w(x)<t}\{x:w(x)<t\} are convex for each t>1t>1;

  • (iii)

    There is 1r1\leq r\leq\infty such that for all m0m\in\mathbb{N}_{0} we have mfLr(d)\nabla^{m}f\in L^{r}(\mathbb{R}^{d}).

Then, for each ϵ>0\epsilon>0 and each m0m\in\mathbb{N}_{0}, there is a constant Am,f,ϵA_{m,f,\epsilon} such that

(1.7) |mf(x)|Am,f,ϵw(x)(1ϵ)xd.|\nabla^{m}f(x)|\leq A_{m,f,\epsilon}w(x)^{-(1-\epsilon)}\quad\forall x\in\mathbb{R}^{d}\,.

It has recently come to our attention that a version of such result is known in dimension 1 from [KZ92, Theorem 1.7]. As we could not find a suitable reference for the higher-dimensional result, we decided to include it here together with its proof, as it is also of independent interest. With such a tool at hand, we get a sharp relation (up to the endpoint) between Hardy’s, Cowling–Price’s and Morgan’s Uncertainty Principles in the sub-critical regime. As we’ve already recalled Hardy’s and Cowling–Price’s Uncertainty Principles above, we briefly recall Morgan’s Uncertainty Principle below (in a generalized version obtained by Ben Farah and Mokni [BFM03]). For that, we shall use the notation ea,b(x)=eaπ|x|b.e_{a,b}(x)=e^{a\pi|x|^{b}}.

Theorem (Morgan; Ben Farah–Mokni).

Suppose that ea,αfLp(d)e_{a,\alpha}f\in L^{p}(\mathbb{R}^{d}) and that eb,βf^Lq(d)e_{b,\beta}\widehat{f}\in L^{q}(\mathbb{R}^{d}) for 1p,q1\leq p,q\leq\infty, α>2\alpha>2 and β=α/(α1)\beta=\alpha/(\alpha-1). Then we have the following conclusions

  • If (aα)1/α(bβ)1/β>sin1/β(π2(β1))(a\alpha)^{1/\alpha}(b\beta)^{1/\beta}>\sin^{1/\beta}\left(\frac{\pi}{2}(\beta-1)\right), then f0f\equiv 0;

  • If (aα)1/α(bβ)1/β<sin1/β(π2(β1))(a\alpha)^{1/\alpha}(b\beta)^{1/\beta}<\sin^{1/\beta}\left(\frac{\pi}{2}(\beta-1)\right), then there are nontrivial functions verifying both conditions.

Observe that, in all the situations mentioned, when we are in the subcritical situation, the theorems do not provide a clear information about the behavior of the functions. Our goal is to provide better information about the structure and behavior of such functions, and additionally, to reformulate this as a kind of quantitative relation. In that regard, we have the following:

Corollary 1.3 (Subcritical estimates).

If the function f:df:\mathbb{R}^{d}\to\mathbb{C} is such that ea,αfLpe_{a,\alpha}f\in L^{p} and ea,βf^Lqe_{a,\beta}\hat{f}\in L^{q} for some a,b,α,β>0a,b,\alpha,\beta>0 and p,q1p,q\geq 1 then for all ϵ>0\epsilon>0 there exists C=C(ϵ,f)C=C(\epsilon,f) such that |f(x)|Me(1ϵ)aπ|x|α|f(x)|\leq Me^{-(1-\epsilon)a\pi|x|^{\alpha}}.

Corollary 1.3 may be then seen as a step in order to convert L2L^{2} results for Gaussian weights into LL^{\infty} ones. Indeed, in the supercritical case ab1ab\geq 1 in Cowling–Price’s UP, this result shows that the only relevant case is indeed ab=1,ab=1, as all others imply the hypotheses in Hardy’s uncertainty principle.

The last results which we prove in this paper address the question of the endpoint in both Theorem 1.1 and Corollary 1.3. Indeed, Theorem 1.1 leaves, perhaps suggestively, the sequence {(2k+1)/16}k\{(2k+1)/16\}_{k\in\mathbb{Z}} out of its statement – which contains the (dilated) version of eigenvalues of the harmonic oscillator. Furthermore, Corollary 1.3 leaves open the question of determining whether a function ff such that

|f(x)|2e2πα|x|2𝑑x< and |f^(x)|2e2πα|x|2𝑑x<\int_{\mathbb{R}}|f(x)|^{2}e^{2\pi\alpha|x|^{2}}\,dx<\infty\mbox{ and }\int_{\mathbb{R}}|\widehat{f}(x)|^{2}e^{2\pi\alpha|x|^{2}}\,dx<\infty

automatically satisfies f(x)eaπx2L.f(x)e^{a\pi x^{2}}\in L^{\infty}.

In this direction, given a finite measure μ\mu with support on the positive real line, we consider its Laplace transform

(1.8) μ(s)=0+est𝑑μ(t)\mathcal{L}\mu(s)=\int_{0}^{+\infty}e^{-st}\,d\mu(t)

and let φ(x)=μ(π|x|2).\varphi(x)=\mathcal{L}\mu(\pi|x|^{2}).

Theorem 1.4.

If φEa2,\varphi\in E^{2}_{a}, then φEa.\varphi\in E^{\infty}_{a}.

For the question on the endpoint of Theorem 1.1, we consider a slightly different class of functions: indeed, as we shall see in subsection 4.1, the endpoint version of Corollary 1.3 is much easier for Laplace transforms of measures supported on the positive real line.

Nevertheless, one may still wonder whether this example may be suitably tweaked in order to obtain a class of functions for which Vemuri’s conjecture is indeed sharp. In fact, Vemuri himself obtained that, if

𝒢a(x):=eπ(a+i1a2)|x|2,\mathcal{G}_{a}(x):=e^{-\pi(a+i\sqrt{1-a^{2}})|x|^{2}},

with a=tanh(2α)(0,1),a=\tanh(2\alpha)\in(0,1), then |Φ𝒢a(y,1/16)|=Ceπtanh(α)|y|2.|\Phi\mathcal{G}_{a}(y,-1/16)|=Ce^{-\pi\tanh(\alpha)|y|^{2}}. Inspired by this observation, we prove that the full version of Vemuri’s conjecture, as well as the endpoint version of our main result, hold and are sharp for a class of transforms based on the functions 𝒢a\mathcal{G}_{a} above.

Theorem 1.5.

Let a=tanh(2α)(0,1),a=\tanh(2\alpha)\in(0,1), and

φ(x)=01𝒢r(x)𝑑μ(r)\varphi(x)=\int_{0}^{1}\mathcal{G}_{r}(x)\,d\mu(r)

for some finite measure μ.\mu. Then:

  1. (1)

    If φEa2,\varphi\in E^{2}_{a}, then φEa;\varphi\in E^{\infty}_{a};

  2. (2)

    If φEa,\varphi\in E^{\infty}_{a}, then for all β,\beta\in\mathbb{R}, we have Φφ(,β)Etanh(α).\Phi\varphi(\cdot,\beta)\in E^{\infty}_{\tanh(\alpha)}.

The structure of the article is as follows:

  • In Section 2, we will prove Lemma 1.2, as well as Lemma 2.1 which shows that strong enough LpL^{p} bounds on ff and f^\hat{f} imply that f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}).

  • In Section 3, we will prove Corollary 1.3 and Theorem 1.1.

  • Finally, in Section 4, we will prove Theorems 1.4 and 1.5, which introduce a large class of examples that verify the conclusion of Theorem 1.1 and Corollary 1.3 without the ϵ\epsilon loss for the case of Gaussian type weights. In this part, besides our main results and techniques, we shall resort to complex analysis methods as well.

2. Main Lemmas

The proof of Lemma 1.2 is based on the following higher-dimensional version of the Kolmogorov–Landau inequality. For the reader’s convenience we provide a short proof of it.

Lemma 2.1.

Let Ω={x=(x1,,xn)n:x1>0}\Omega=\{x=(x_{1},\ldots,x_{n})\in\mathbb{R}^{n}:\,x_{1}>0\} and let fLp(Ω)Cm(Ω¯)f\in L^{p}(\Omega)\cap C^{m}(\bar{\Omega}) for some 1p1\leq p\leq\infty, m>nm>n. For each 0kmn0\leq k\leq m-n there exists C=Ck,m,nC=C_{k,m,n} such that for all 1r1\leq r\leq\infty we have

(2.1) |kf(0)|CfLp(Ω)αmfLr(Ω)1α,|\nabla^{k}f(0)|\leq C||f||_{L^{p}(\Omega)}^{\alpha}||\nabla^{m}f||_{L^{r}(\Omega)}^{1-\alpha},

where α=α(k,n,m,p,r)=mknrm+npnr\alpha=\alpha(k,n,m,p,r)=\frac{m-k-\frac{n}{r}}{m+\frac{n}{p}-\frac{n}{r}}.

Note that by applying an orthogonal transformation this lemma can be applied to any half-space in place of Ω\Omega and any point pp on its boundary in place of 0.

Proof.

First, we show that the right-hand side of (2.1) is positive unless ff is identically zero. Indeed, if fLp=0||f||_{L^{p}}=0 then ff is zero almost everywhere, hence zero identically since fCm(Ω¯)f\in C^{m}(\bar{\Omega}). Similarly, if mfLr=0||\nabla^{m}f||_{L^{r}}=0 then mf\nabla^{m}f is identically zero, hence ff is a polynomial of degree at most m1m-1. But then it is not in Lp(Ω)L^{p}(\Omega) unless it is identically zero.

So, we can assume that both fLp||f||_{L^{p}} and mfLr||\nabla^{m}f||_{L^{r}} are strictly positive. If one of them is infinite then there is nothing to prove. Let us consider the function g(x)=af(bx)g(x)=af(bx) for some a,b>0a,b>0. Observe that estimates (2.1) for ff and for gg are equivalent due to our choice of α\alpha since both sides are multiplied by the same amount. By choosing appropriate numbers aa and bb we can without loss of generality assume that gLp=mgLr=1||g||_{L^{p}}=||\nabla^{m}g||_{L^{r}}=1 and we have to show that |kg(0)|C|\nabla^{k}g(0)|\leq C.

Let U=[0,1]×[12,12]n1U=[0,1]\times[-\frac{1}{2},\frac{1}{2}]^{n-1}. Since p,r1p,r\geq 1 and the measure of UU is 11 we have gL1(U)1||g||_{L^{1}(U)}\leq 1 and mgL1(U)1||\nabla^{m}g||_{L^{1}(U)}\leq 1 by Hölder’s inequality. It remains to use two well-known facts from the theory of Sobolev spaces:

  1. (i)

    the space W1m(U)W^{m}_{1}(U) – functions having weak derivatives up to order mm in L1(U)L^{1}(U) on the bounded Lipschitz domain UU – continuously embeds into Cnm(U¯)C^{n-m}(\bar{U});

  2. (ii)

    for such spaces, considering only the norm of the function and its mm-th derivative yields an equivalent norm.

Since 0U¯,0\in\bar{U}, we get the desired result. ∎

Proof of Lemma 1.2.

Let us fix x0dx_{0}\in\mathbb{R}^{d}. Since the set V={xdw(x)<w(x0)}V=\{x\in\mathbb{R}^{d}\mid w(x)<w(x_{0})\} is convex and x0Vx_{0}\notin V we can find a half-space Ω\Omega such that x0x_{0} is on its boundary and for all xΩx\in\Omega we have w(x)w(x0)w(x)\geq w(x_{0}). We have

Ω|f(x)|p𝑑xw(x0)pΩ|f(x)|pw(x)p𝑑xw(x0)pn|f(x)|pw(x)p𝑑x=w(x0)pCf\int_{\Omega}|f(x)|^{p}dx\leq w(x_{0})^{-p}\int_{\Omega}|f(x)|^{p}w(x)^{p}dx\leq w(x_{0})^{-p}\int_{\mathbb{R}^{n}}|f(x)|^{p}w(x)^{p}dx=w(x_{0})^{-p}C_{f}

and

mfLr(Ω)mfLr(n)=Cf,m.||\nabla^{m}f||_{L^{r}(\Omega)}\leq||\nabla^{m}f||_{L^{r}(\mathbb{R}^{n})}=C_{f,m}.

Applying Lemma 2.1 to ff we get for m>n+km>n+k

|kf(x0)|Ck,n,mw(x0)αCfαCf,m1α.|\nabla^{k}f(x_{0})|\leq C_{k,n,m}w(x_{0})^{-\alpha}C_{f}^{\alpha}C_{f,m}^{1-\alpha}.

Observe that for fixed k,n,p,rk,n,p,r given ϵ>0\epsilon>0 for big enough mm we have α>1ϵ\alpha>1-\epsilon. Choosing such an mm gives us the desired estimate. ∎

To verify condition (iii) of Lemma 1.2 we will use the following lemma which says that if ff and f^\hat{f} decay faster than any polynomial on average then f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}).

Lemma 2.2.

Let f:nf:\mathbb{R}^{n}\to\mathbb{C} and let 1p,q<1\leq p,q<\infty. If for all m0m\in\mathbb{N}_{0} we have

d|f(x)|p(1+|x|)pm<\int_{\mathbb{R}^{d}}|f(x)|^{p}(1+|x|)^{pm}<\infty

and

d|f^(x)|q(1+|x|)qm<\int_{\mathbb{R}^{d}}|\hat{f}(x)|^{q}(1+|x|)^{qm}<\infty

then f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}).

Proof.

First, we show that mf\nabla^{m}f is bounded and continuous for all m0m\in\mathbb{N}_{0}. For a multi-index β0n\beta\in\mathbb{N}_{0}^{n} we have βf^(x)=(2πi)|β|xβf^(x)\widehat{\partial^{\beta}f}(x)=(2\pi i)^{|\beta|}x^{\beta}\hat{f}(x). Thus, if xβf^(x)L1(n)x^{\beta}\hat{f}(x)\in L^{1}(\mathbb{R}^{n}) then βf\partial^{\beta}f is bounded and continuous. We have

xβf^(x)L1(d)(1+|x|)|β|f^(x)L1(d)(1+|x|)mf^(x)Lq(d)(1+|x|)|β|mLqq1(d).||x^{\beta}\hat{f}(x)||_{L^{1}(\mathbb{R}^{d})}\leq||(1+|x|)^{|\beta|}\hat{f}(x)||_{L^{1}(\mathbb{R}^{d})}\leq||(1+|x|)^{m}\hat{f}(x)||_{L^{q}(\mathbb{R}^{d})}||(1+|x|)^{|\beta|-m}||_{L^{\frac{q}{q-1}}(\mathbb{R}^{d})}.

If mm is chosen bigger than |β|+n|\beta|+n then this quantity is finite and thus xβf^(x)L1(n)x^{\beta}\hat{f}(x)\in L^{1}(\mathbb{R}^{n}).

In particular, we get that fC(d)f\in C^{\infty}(\mathbb{R}^{d}). To finish the proof of the lemma, we are going to apply Lemma 1.2 to ff. Consider the weight w(x)=(1+|x|)mw(x)=(1+|x|)^{m}. Functions ff and ww satisfy the assumptions of Lemma 1.2 with p=pp=p, r=r=\infty. Thus, for all ϵ>0\epsilon>0, in particular ϵ=12\epsilon=\frac{1}{2}, we have

|kf(x)|C(1+|x|)(ϵ1)m=C(1+|x|)m/2.|\nabla^{k}f(x)|\leq C(1+|x|)^{(\epsilon-1)m}=C(1+|x|)^{-m/2}.

Since mm and kk are arbitrary, we get that f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}). ∎

3. Proof of Corollary 1.3 and Theorem 1.1

Proof of Corllary 1.3.

With the tools we have at our disposal, Corollary 1.3 becomes a trivial consequence. Indeed, when we are treating the situation in Cowling-Price’s uncertainty principle, since the estimates |f(x)|eαπ|x|2Lp|f(x)|e^{\alpha\pi|x|^{2}}\in L^{p} and |f^(x)|eβπ|x|2Lq|\widehat{f}(x)|e^{\beta\pi|x|^{2}}\in L^{q} imply, by Lemma 2.2, that f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}), we are in position to apply Lemma 1.2 and obtain for each ϵ>0\epsilon>0 the existence of a constant Aϵ>0A_{\epsilon}>0 such that

|f(x)|Aϵeπα(1ϵ)|x|2,|f(x)|\leq A_{\epsilon}e^{-\pi\alpha(1-\epsilon)|x|^{2}},

and the analogous estimate holds for the Fourier transform. The case of the Morgan uncertainty principle is entirely analogous, and thus we are done. ∎

We now move on to the proof of our main theorem.

Proof of Theorem 1.1.

The proof will be divided into several steps.

Step 1. Translating between the Quantum Harmonic Oscillator and the linear Schrödinger equation. As we saw in the introduction, there is a simple way to write solutions to write solutions to (1.5) in terms of the Hermite basis. We will use this connection, and the action of the Schrödinger evolution, to provide a simple proof of the link between (1.5) and the Schrödinger equation.

Before continuing, we introduce some notation for Hermite eigenfunctions of the Fourier transform in higher dimensions. For a multi-index (α1,α2,,αn)=α0n(\alpha_{1},\alpha_{2},\dots,\alpha_{n})=\alpha\in\mathbb{N}_{0}^{n}, we define the Hermite function of order α\alpha as

𝐡α(x)=Πi=1nhαi(xi).\mathbf{h}_{\alpha}(x)=\Pi_{i=1}^{n}h_{\alpha_{i}}(x_{i}).

We know from [Gon19, Lemma 11] that

eitΔ(𝐡α)(x)=\displaystyle e^{it\Delta}(\mathbf{h}_{\alpha})(x)=
(1+4πit)n/2exp[4π2it1+16π2t2|x|2](14πit1+4πit)|α|𝐡α(x1+16π2t2).\displaystyle(1+4\pi it)^{-n/2}\exp\left[\frac{4\pi^{2}it}{1+16\pi^{2}t^{2}}|x|^{2}\right]\cdot\left(\sqrt{\frac{1-4\pi it}{1+4\pi it}}\right)^{|\alpha|}\mathbf{h}_{\alpha}\left(\frac{x}{\sqrt{1+16\pi^{2}t^{2}}}\right).

Thus, we may write, whenever f𝒮(n),f\in\mathcal{S}(\mathbb{R}^{n}), f(x)=αnaα𝐡α(x)f(x)=\sum_{\alpha\in\mathbb{N}^{n}}a_{\alpha}\mathbf{h}_{\alpha}(x),

(3.1) eitΔf(x)\displaystyle e^{it\Delta}f(x) =(1+4πit)n/2exp[4π2it1+16π2t2|x|2]\displaystyle=(1+4\pi it)^{-n/2}\exp\left[\frac{4\pi^{2}it}{1+16\pi^{2}t^{2}}|x|^{2}\right]
(3.2) ×αneiarctan(4πt)|α|aα𝐡α(x1+16π2t2)\displaystyle\times\sum_{\alpha\in\mathbb{N}^{n}}e^{i\arctan(-4\pi t)|\alpha|}\cdot a_{\alpha}\cdot\mathbf{h}_{\alpha}\left(\frac{x}{\sqrt{1+16\pi^{2}t^{2}}}\right)
(3.3) =(1+16π2t2)n/4exp[4π2it1+16π2t2|x|2]Φ(x1+16π2t2,arctan(4πt)4π).\displaystyle=(1+16\pi^{2}t^{2})^{-n/4}\exp\left[\frac{4\pi^{2}it}{1+16\pi^{2}t^{2}}|x|^{2}\right]\cdot\Phi\left(\frac{x}{\sqrt{1+16\pi^{2}t^{2}}},\frac{\arctan(-4\pi t)}{4\pi}\right).

The correspondence established in (3.1) above will be crucial for the next step.

Step 2. Using the estimates by Escauriaza–Kenig–Ponce–Vega in order to deduce decay for the solution of the Schrödinger equation. We make use of the translation from the previous step to establish the decay. We follow the overall approach of [EKPV08a, EKPV08b, EKPV06, CEK+10, EKPV16]. In particular, the proof of Theorem 1 in [CEK+10] yields as a by-product that, if uu is a solution to

{itu=Δu in n×,u(x,0)=g(x) on n,\displaystyle\begin{cases}i\partial_{t}u=-\Delta u&\text{ in }\,\mathbb{R}^{n}\times\mathbb{R},\cr u(x,0)=g(x)&\text{ on }\mathbb{R}^{n},\cr\end{cases}

then the function

v(x,t)=(it)n/2e|x|24itu¯(x/t,1/t1)v(x,t)=(it)^{-n/2}e^{-\frac{|x|^{2}}{4it}}\overline{u}(x/t,1/t-1)

satisfies

{itv=Δv in n×(0,+),v(x,0)=(4π)n/2ei|x|24g^¯(x/4π) on n,v(x,1)=in/2e|x|2/4ig¯(x) on n.\displaystyle\begin{cases}i\partial_{t}v=-\Delta v&\text{ in }\,\mathbb{R}^{n}\times(0,+\infty),\cr v(x,0)=(4\pi)^{-n/2}e^{\frac{-i|x|^{2}}{4}}\overline{\widehat{g}}(x/4\pi)&\text{ on }\mathbb{R}^{n},\cr v(x,1)=i^{-n/2}e^{-|x|^{2}/4i}\overline{g}(x)&\text{ on }\mathbb{R}^{n}.\cr\end{cases}

We shall use this fact with gg being a suitable dilation of f.f.

Indeed, let g(x)=f(x2π).g(x)=f\left(\frac{x}{2\sqrt{\pi}}\right). Then we know that |g(x)|ea|x|24,|g^(ξ/4π)|ea|x|24,|g(x)|\lesssim e^{-\frac{a|x|^{2}}{4}},\,|\widehat{g}(\xi/4\pi)|\lesssim e^{-\frac{a|x|^{2}}{4}}, where we put a=tanh(2α).a=\tanh(2\alpha).

For such g,g, we have that the associated solution vv above satisfies v(x,0),v(x,1)v(x,0),v(x,1) L2(eaϵ4|x|2dx),\,\,\,\in L^{2}(e^{\frac{a-\epsilon}{4}|x|^{2}}\,dx), for all ϵ>0.\epsilon>0. We may now invoke the following result, which first appears in the works of Escauriaza–Kenig–Ponce–Vega [EKPV10, Theorem 3].

Lemma 3.1.

Assume that wC([0,1],L2(n))w\in C([0,1],L^{2}(\mathbb{R}^{n})) satisfies

itw+Δw=0,inn×[0,1].i\partial_{t}w+\Delta w=0,\,in\,\,\,\mathbb{R}^{n}\times[0,1].

Then

supt[0,1]eA(t)|x|2w(t)2eA|x|2w(0)2+eA|x|2w(1)2,\sup_{t\in[0,1]}\|e^{A(t)|x|^{2}}w(t)\|_{2}\lesssim\|e^{A|x|^{2}}w(0)\|_{2}+\|e^{A|x|^{2}}w(1)\|_{2},

where

A(t)=R2(1+R2(2t1)2),A=R2(1+R2),A(t)=\frac{R}{2(1+R^{2}(2t-1)^{2})},\,\,\,\,A=\frac{R}{2(1+R^{2})},

and 0<R<10<R<1.

We then use Lemma 3.1 with A=aϵ4A=\frac{a-\epsilon}{4}. Let Ra,ϵR_{a,\epsilon} be the unique number between 0 and 11 such that aϵ4=Ra,ϵ2(1+Ra,ϵ2)\frac{a-\epsilon}{4}=\frac{R_{a,\epsilon}}{2(1+R_{a,\epsilon}^{2})}, and denote the function A(t)A(t) thus obtained by Aa,ϵ(t)A_{a,\epsilon}(t). We have then

supt[0,1]eAa,ϵ(t)|x|2v(x,t)L2(dx)<+.\sup_{t\in[0,1]}\|e^{A_{a,\epsilon}(t)|x|^{2}}v(x,t)\|_{L^{2}(dx)}<+\infty.

Reverting back to u,u, we find out that

eBa,ϵ(s)|x|2u(x,s)2<+,s>0,\|e^{B_{a,\epsilon}(s)|x|^{2}}u(x,s)\|_{2}<+\infty,\,\forall\,\,s>0,

where Ba,ϵ(s)=Aa,ϵ(1/(s+1))(s+1)2.B_{a,\epsilon}(s)=\frac{A_{a,\epsilon}(1/(s+1))}{(s+1)^{2}}. Observe that eitΔf(x)=u(2πx,4πt)e^{it\Delta}f(x)=u(2\sqrt{\pi}x,4\pi t), and thus we have proven that

e4πBa,ϵ(4πt)|x|2eitΔf(x)2<+,t>0.\|e^{4\pi B_{a,\epsilon}(4\pi t)|x|^{2}}e^{it\Delta}f(x)\|_{2}<+\infty,\,\,\forall\,\,t>0.

Step 3. Translating back. Using the correspondence (3.1) between solutions of the quantum harmonic oscillator and Schrödinger’s equation, we see that

exp((1+16π2t2)4πBa,ϵ(4πt)|x|2)Φ(y,arctan(4πt)4π)2<+,t>0.\left\|\exp((1+16\pi^{2}t^{2})4\pi B_{a,\epsilon}(4\pi t)|x|^{2})\Phi\left(y,\frac{\arctan(-4\pi t)}{4\pi}\right)\right\|_{2}<+\infty,\,\forall\,\,t>0.

Let

Ωa,ϵ(t):=(1+16π2t2)4πBa,ϵ(4πt)=(1+16π2t2)4πRa,ϵ2[(4πt+1)2+Ra,ϵ2(4πt1)2].\displaystyle\Omega_{a,\epsilon}(t):=(1+16\pi^{2}t^{2})4\pi\cdot B_{a,\epsilon}(4\pi t)=\frac{(1+16\pi^{2}t^{2})\cdot 4\pi\cdot R_{a,\epsilon}}{2[(4\pi t+1)^{2}+R_{a,\epsilon}^{2}(4\pi t-1)^{2}]}.

Notice however that, as Ra,ϵ<1,R_{a,\epsilon}<1, this function has exactly one minimum point, which happens at t=14π,t=\frac{1}{4\pi}, as

Ωa,ϵ(t)πRa,ϵ=πRa,ϵ(1Ra,ϵ2)(4πt1)2[(4πt+1)2+Ra,ϵ2(4πt1)2]0.\Omega_{a,\epsilon}(t)-\pi R_{a,\epsilon}=\frac{\pi R_{a,\epsilon}(1-R_{a,\epsilon}^{2})(4\pi t-1)^{2}}{[(4\pi t+1)^{2}+R_{a,\epsilon}^{2}(4\pi t-1)^{2}]}\geq 0.

At t=1/4π,t=1/4\pi, we have

Ωa,ϵ(1/4π)=πRa,ϵ.\Omega_{a,\epsilon}(1/4\pi)=\pi R_{a,\epsilon}.
aπa\cdot\pi\cdotaπa\cdot\pi\cdot\cdotRaπR_{a}\cdot\pi 01/4π1/4\pi1/2π1/2\pitt
Figure 1. The limiting function Ωa\Omega_{a} is bounded from below by the curve in orange above. The dashed line represents the predicted gain of decay in Vemuri’s conjecture.

As ϵ0,\epsilon\to 0, we readily see that Ra,ϵtanh(α).R_{a,\epsilon}\to\tanh(\alpha). Let then Ωa=limϵ0Ωa,ϵ.\Omega_{a}=\lim_{\epsilon\to 0}\Omega_{a,\epsilon}. We observe that, for s(1/16,0)s\in(-1/16,0), we have

Φ(y,s)eb|y|22<+,b<Ωa(tan(4πs)).\|\Phi(y,s)\cdot e^{b|y|^{2}}\|_{2}<+\infty,\,\forall b<\Omega_{a}(-\tan(4\pi s)).

As Ωa(tan(4πs))>πtanh(α)\Omega_{a}(-\tan(4\pi s))>\pi\tanh(\alpha) for s(1/16,0),s\in(-1/16,0), there is b(s)>πtanh(α)b(s)>\pi\tanh(\alpha) so that

(3.5) Φ(y,s)eb(s)|y|22<+.\|\Phi(y,s)\cdot e^{b(s)|y|^{2}}\|_{2}<+\infty.

In order to extend this analysis to the rest of the claimed set, we notice that |Φ(y,1/8)|=|f(y)|,|Φ(y,1/8)|=|1f(y)|,|\Phi(y,-1/8)|=|\mathcal{F}f(y)|,|\Phi(y,1/8)|=|\mathcal{F}^{-1}f(y)|, and so on, so that Φ(y,k/8)exp(aπ|y|2)L\Phi(y,k/8)\cdot\exp(a\pi|y|^{2})\in L^{\infty} whenever k.k\in\mathbb{Z}. Moreover, if we let Ψ\Psi be a solution of (1.5) with the initial condition k(f),\mathcal{F}^{k}(f), then we have

|Ψ(y,t)|=|Φ(y,tk/8)|.|\Psi(y,t)|=|\Phi(y,t-k/8)|.

Using these observations, together with the fact that (1.5) is time-reversible, we are able to conclude that, for all s{116+k8,k},s\in\mathbb{R}\setminus\{\frac{1}{16}+\frac{k}{8},\,k\in\mathbb{Z}\}, there is b(s)>πtanh(α)b(s)>\pi\tanh(\alpha) so that (3.5) holds.

Step 4. Conclusion. Finally, we use our main result in order to conclude. Indeed, by the explicit formula for the solution of (1.5), we have that |yΦ(y,t)|=|Φ(y,t1/8)|.|\mathcal{F}_{y}\Phi(y,t)|=|\Phi(y,t-1/8)|. If t{116+k8,k},t\not\in\{\frac{1}{16}+\frac{k}{8},\,k\in\mathbb{Z}\}, then t1/8{116+k8,k},t-1/8\not\in\{\frac{1}{16}+\frac{k}{8},\,k\in\mathbb{Z}\}, and so, for c(t)=min{b(t),b(t1/8)},c(t)=\min\{b(t),b(t-1/8)\}, we have

Φ(y,t)ec(t)|y|2,(yΦ(y,t))ec(t)|y|2L2(n).\Phi(y,t)e^{c(t)|y|^{2}},(\mathcal{F}_{y}\Phi(y,t))\cdot e^{c(t)|y|^{2}}\in L^{2}(\mathbb{R}^{n}).

From the main theorem, we have Φ(y,t)e(c(t)ϵ)|y|2L\Phi(y,t)e^{(c(t)-\epsilon)|y|^{2}}\in L^{\infty} for any ϵ>0.\epsilon>0. Taking ϵ>0\epsilon>0 sufficiently small shows that

Φ(y,t)eπtanh(α)|y|2L(n),t{116+k8,k}.\Phi(y,t)e^{\pi\tanh(\alpha)|y|^{2}}\in L^{\infty}(\mathbb{R}^{n}),\,\forall\,t\not\in\{\frac{1}{16}+\frac{k}{8},\,k\in\mathbb{Z}\}.

This finishes the proof of Theorem 1.1. ∎

Remark 3.2.

Observe that the combination of the above lemmas in Section 2 is quite powerful, but in order to use the Lemma 1.2 to generate pointwise control, we do not need to impose that the function is controlled in space and frequency: much weaker estimates are more than enough to ensure the control that we need. This opens the door to understand other kinds of uncertainty principles in a broad range of situations, which we plan to do in future work.

Remark 3.3.

The proof of Theorem 1.1 highlights that the original conjecture by Vemuri, albeit sharp if one considers the set of all times t,t\in\mathbb{R}, is almost-never sharp for a given time t.t\in\mathbb{R}. Indeed, we can ‘upgrade’ Vemuri’s original conjecture to the following version:

Conjecture 3.4.

Let fEtanh(2α)(n),f\in E_{\tanh(2\alpha)}^{\infty}(n), for some α>0.\alpha>0. Then

Φ(,t)EΩα(t)(n) for all t,\Phi(\cdot,t)\in E_{\Omega_{\alpha}(t)}^{\infty}(n)\text{ for all }t\in\mathbb{R},

where we let Ωα(t)=(1+16π2s2)2tanh(α)[(4πs+1)2+tanh(α)2(4πs1)2],\Omega_{\alpha}(t)=\frac{(1+16\pi^{2}s^{2})\cdot 2\cdot\tanh(\alpha)}{[(4\pi s+1)^{2}+\tanh(\alpha)^{2}(4\pi s-1)^{2}]}, with s=tan(4πt).s=-\tan(4\pi t).

As we will see in the next section, in the particular cases of Theorem 1.5, we are also able to settle this conjecture. This is a strong reason why we believe such a conjecture should be true.

4. On the endpoint versions of Corollary 1.3 and Theorem 1.1

4.1. Proof of Theorem 1.4

In order to prove such a result, we start by noticing that, from Corollary 1.3, we have that φEaε,\varphi\in E^{\infty}_{a-\varepsilon}, for any ε>0.\varepsilon>0. We then have the following Lemma on decay of Laplace transforms:

Lemma 4.1.

Let μ\mu be a finite measure supported on the positive real line. Suppose that its Laplace transform satisfies |μ(s)|Cec0s,s>0,|\mathcal{L}\mu(s)|\leq Ce^{-c_{0}s},\,\forall s>0, for some C>0,C>0, c0+c_{0}\in\mathbb{R}_{+}. Then supp(μ)[c0,+).\text{supp}(\mu)\subset[c_{0},+\infty).

Proof.

To prove this lemma, we define the function

F(z)=eic0zμ(iz).F(z)=e^{-ic_{0}z}\mathcal{L}\mu(-iz).

Note that, by the definition of the Laplace transform, FF is a holomorphic function in the upper half plane .\mathbb{H}. Moreover, it has the following properties:

  1. (1)

    FF is bounded on the real line. This follows from the fact that μ(it)\mathcal{L}\mu(-it) is just a (rescaled) Fourer transform of the measure μ.\mu. As μ\mu is finite, its Fourier transform is bounded, and the modulation factor eic0te^{-ic_{0}t} has absolute value one.

  2. (2)

    |F(is)|C,s>0.|F(is)|\leq C,\,\forall s>0. This follows directly from our decay assumption.

  3. (3)

    |F(z)|C~ec0|z|,|F(z)|\leq\tilde{C}e^{c_{0}|z|}, for some C~>0.\tilde{C}>0. This follows again by the fact that μ\mathcal{L}\mu is uniformly bounded on the upper half space.

With these properties at hand, we are able to use the Phragmén–Lindelöf principle in the first and second quadrants separately. This implies that FF is bounded and continuous in .\mathbb{H}.

Thus, FF may be written as a Poisson integral of its boundary values. In particular, by the Young’s convolution inequality, we have

F(+iy)L2(dx)=(F|)Py(x)L2(dx)F|2.\displaystyle\|F(\cdot+iy)\|_{L^{2}(dx)}=\|(F|_{\mathbb{R}})*P_{y}(x)\|_{L^{2}(dx)}\leq\|F|_{\mathbb{R}}\|_{2}.

This inequality only holds, of course, if F|L2.F|_{\mathbb{R}}\in L^{2}. For now, let us assume that dμ=f(x)dx,d\mu=f(x)\,dx, with fL1L2.f\in L^{1}\cap L^{2}. Then the computation above shows us that FH2()F\in H^{2}(\mathbb{H}) (the Hardy space on the upper half space). In particular, by the Paley–Wiener theorem, we must have that F|=h^,F|_{\mathbb{R}}=\widehat{h}, where supp(h)(0,+).\text{supp}(h)\subset(0,+\infty).

On the other hand, we see that F|F|_{\mathbb{R}} may be written as a (rescaled) Fourier transform of f(x+c0)1x+c0>0.f(x+c_{0})1_{x+c_{0}>0}. Thus, f(x+c0)0f(x+c_{0})\neq 0 only if x0,x\geq 0, and so f(y)0f(y)\neq 0 only if yc0.y\geq c_{0}. This concludes the proof in the case where dμ=f(x)dx,fL1L2.d\mu=f(x)dx,\,f\in L^{1}\cap L^{2}.

For the general case, consider a smooth, positive compactly supported function ϕ\phi so that supp(ϕ)\text{supp}(\phi) is contained in the positive real line (0,+),(0,+\infty), and ϕ=1.\int\phi=1. Let then μϵ(x)=(dμ)ϕϵ(x),\mu_{\epsilon}(x)=(d\mu)*\phi_{\epsilon}(x), where we define ϕϵ(y)=1ϵϕ(yϵ).\phi_{\epsilon}(y)=\frac{1}{\epsilon}\phi\left(\frac{y}{\epsilon}\right).

By the Young’s inequality, we have μϵL1L2().\mu_{\epsilon}\in L^{1}\cap L^{2}(\mathbb{R}). Moreover, μϵ=(μ)(ϕϵ)\mathcal{L}\mu_{\epsilon}=(\mathcal{L}\mu)\cdot(\mathcal{L}\phi_{\epsilon}) by the definition of ϕ.\phi. As |f(s)|f1|\mathcal{L}f(s)|\leq\|f\|_{1} uniformly on s{z:Re(z)0},s\in\{z\in\mathbb{C}\colon\text{Re}(z)\geq 0\}, we have that

|μϵ(s)|Cec0s,|\mathcal{L}\mu_{\epsilon}(s)|\leq Ce^{-c_{0}s},

uniformly on ϵ>0.\epsilon>0. Thus, supp(μϵ)[c0,+).\text{supp}(\mu_{\epsilon})\subset[c_{0},+\infty). As μϵdμ\mu_{\epsilon}\stackrel{{\scriptstyle\ast}}{{\rightharpoonup}}d\mu in the space of finite measures on the real line, we see that supp(μ)[c0,+),\text{supp}(\mu)\subset[c_{0},+\infty), as desired. ∎

We are now ready to finish the proof of Theorem 1.4.

Proof of Theorem 1.4.

First we notice that, by our main result, φEaϵ,ϵ>0.\varphi\in E^{\infty}_{a-\epsilon},\,\forall\epsilon>0. This implies that |μ(s)|ϵe(aϵ)s,s>0.|\mathcal{L}\mu(s)|\lesssim_{\epsilon}e^{-(a-\epsilon)s},\,\forall s>0. By Lemma 4.1, we conclude that supp(μ)[aϵ,+),ϵ>0.\text{supp}(\mu)\subset[a-\epsilon,+\infty),\,\,\forall\epsilon>0. This plainly implies that supp(μ)[a,+),\text{supp}(\mu)\subset[a,+\infty), which in turn implies that |φ(x)|eπa|x|2.|\varphi(x)|\lesssim e^{-\pi a|x|^{2}}. By observing that the Fourier transform

φ^(ξ)=01t1/2eπt|ξ|2dμ(t)=:0erπ|ξ|2dν(r)\widehat{\varphi}(\xi)=\int_{0}^{\infty}\frac{1}{t^{1/2}}e^{-\frac{\pi}{t}|\xi|^{2}}\,\,d\mu(t)=:\int_{0}^{\infty}e^{-r\pi|\xi|^{2}}d\nu(r)

also satisfies that |ν|(+)a1t1/2|dμ|(t)<+,|\nu|(\mathbb{R}_{+})\leq\int_{a}^{\infty}\frac{1}{t^{1/2}}\,|d\mu|(t)<+\infty, we may employ the same reasoning to conclude that also |φ^(ξ)|eaπ|ξ|2.|\widehat{\varphi}(\xi)|\lesssim e^{-a\pi|\xi|^{2}}. This finishes the proof. ∎

The above results lead us to the following question: if fEa2f\in E^{2}_{a}, does it then follow that fEaf\in E^{\infty}_{a}? We are led to speculate that such a question has an affirmative answer based on the previous theorem, and thus we believe that we should have a control in the sub-critical regime of the uncertainty principles without the ϵ\epsilon-loss that is present in the Theorem 1.3.

In view of the considerations above, one may may wonder whether the class of Laplace transforms presented in Subsection 4.1 represents the almost sharp rate of decay obtained in Theorem 1.1. In order to analyse that, we need to introduce the following concept.

For β,\beta\in\mathbb{R}, we define the fractional Fourier transform of order β\beta to act on the Hermite functions as

βhk=eikβhk,\mathcal{F}_{\beta}h_{k}=e^{-ik\beta}h_{k},

and extended it to L2L^{2} in the canonical way. By the properties of the Hermite polynomials and the Mehler kernel [Bec75], one is led to deduce that these transforms have the following representation as integral transforms:

(4.1) βf(x)=ei(θ(β)π/2β/2)|sin(β)|eiπx2cot(β)e2πi(xycsc(β)y2cot(β)/2)f(y)𝑑y,\mathcal{F}_{\beta}f(x)=\frac{e^{i(\theta(\beta)\pi/2-\beta/2)}}{\sqrt{|\sin(\beta)|}}e^{i\pi x^{2}\cot(\beta)}\int_{\mathbb{R}}e^{-2\pi i(xy\csc(\beta)-y^{2}\cot(\beta)/2)}f(y)\,dy,

where θ(β)=sign(sin(β)).\theta(\beta)=\text{sign}(\sin(\beta)). The relationship between fractional Fourier transforms and the evolution of the quantum harmonic oscillator is evident from the definition. Indeed, we may write

Φ(y,t)=e2πit(4πtf)(y).\Phi(y,t)=e^{2\pi it}(\mathcal{F}_{-4\pi t}f)(y).

The key feature of this definition is the relationship with (4.1), which allows us to compute easily fractional Fourier transforms of Gaussians and related functions.

As a first observation, notice that for f(x)=eλπ|x|2,f(x)=e^{-\lambda\pi|x|^{2}}, the integral in (4.1) is just the Fourier transform of eπy2(λ+icot(β))e^{-\pi y^{2}(\lambda+i\cot(\beta))} evaluated at the point xcsc(β).x\csc(\beta). This in turn evaluates directly to

(4.2) 1(λ+icot(β))1/2eπcsc2(β)|x|2λ+icot(β)=1(λ+icot(β))1/2eπ|x|2λcsc2(β)icsc2(β)cot(β)λ2+cot2(β).\frac{1}{(\lambda+i\cot(\beta))^{1/2}}e^{-\frac{\pi\csc^{2}(\beta)|x|^{2}}{\lambda+i\cot(\beta)}}=\frac{1}{(\lambda+i\cot(\beta))^{1/2}}e^{-\pi|x|^{2}\frac{\lambda\csc^{2}(\beta)-i\csc^{2}(\beta)\cot(\beta)}{\lambda^{2}+\cot^{2}(\beta)}}.

Now let φ(x)=μ(π|x|2),\varphi(x)=\mathcal{L}\mu(\pi|x|^{2}), with dμd\mu a finite measure. If we have φEa,\varphi\in E^{\infty}_{a}, then Theorem 1.4 tells us that supp(μ)[a,1/a].\text{supp}(\mu)\subset[a,1/a]. The computation of the fractional Fourier transform of the Gaussian above shows that

(4.3) |βφ(x)|\displaystyle|\mathcal{F}_{\beta}\varphi(x)| a1/aeπ|x|2λcsc2(β)λ2+cot2(β)|dμ|(λ)\displaystyle\lesssim\int_{a}^{1/a}e^{-\pi|x|^{2}\frac{\lambda\csc^{2}(\beta)}{\lambda^{2}+\cot^{2}(\beta)}}\,|d\mu|(\lambda)
(4.4) μTVmax{eπ|x|2acsc2(β)a2+cot2(β),eπ|x|2acsc2(β)1+acot2(β)}.\displaystyle\lesssim\|\mu\|_{TV}\max\{e^{-\pi|x|^{2}\frac{a\csc^{2}(\beta)}{a^{2}+\cot^{2}(\beta)}},e^{-\pi|x|^{2}\frac{a\csc^{2}(\beta)}{1+a\cot^{2}(\beta)}}\}.

For a<1,a<1, we can see that

min{acsc2(β)a2+cot2(β),acsc2(β)1+acot2(β)}a,\min\left\{\frac{a\csc^{2}(\beta)}{a^{2}+\cot^{2}(\beta)},\frac{a\csc^{2}(\beta)}{1+a\cot^{2}(\beta)}\right\}\geq a,

with equality if and only if β=kπ/2,k.\beta=k\pi/2,\,\,k\in\mathbb{Z}. Thus, (4.3) implies that there is ϑ(β)>a\vartheta(\beta)>a whenever βkπ/2,k,\beta\neq k\pi/2,\,\,k\in\mathbb{Z}, so that

|βφ(x)|μTVeπϑ(β)|x|2.|\mathcal{F}_{\beta}\varphi(x)|\lesssim\|\mu\|_{TV}e^{-\pi\vartheta(\beta)|x|^{2}}.

Thus, by relating the fractional Fourier transform to the quantum harmonic oscillator, we obtain much stronger version of Vemuri’s conjecture when φ(x)=μ(π|x|2).\varphi(x)=\mathcal{L}\mu(\pi|x|^{2}).

4.2. Proof of Theorem 1.5

The proof of Theorem 1.5 is based on the following result on support of measures on the circle with rapidly decaying Laplace transform. This, on the other hand, is an analogue of Lemma 4.1 on the circle.

Lemma 4.2.

Let ν\nu be a finite measure on the circle 𝕊1.\mathbb{S}^{1}\subset\mathbb{C}. Suppose that the Laplace transform

ν(t)=𝕊1ezt𝑑ν(z)\mathcal{L}\nu(t)=\int_{\mathbb{S}^{1}}e^{-zt}\,d\nu(z)

satisfies |ν(t)|Cec0t|\mathcal{L}\nu(t)|\leq Ce^{-c_{0}t} for some C>0C>0, c0+.c_{0}\in\mathbb{R}_{+}. Suppose additionally that there is δ>0\delta>0 so that supp(ν)𝕊1{z:Re(z)>1+δ}\text{supp}(\nu)\subset\mathbb{S}^{1}\cap\{z\colon\text{Re}(z)>-1+\delta\} (or, equivalently, 1supp(ν)-1\notin\text{supp}(\nu)). Then

supp(ν)𝕊1{z:Re(z)c0}.\text{supp}(\nu)\subset\mathbb{S}^{1}\cap\{z\colon\text{Re}(z)\geq c_{0}\}.
Proof.

As ν\mathcal{L}\nu is defined and decays as ec0te^{-c_{0}t} on the positive half-line, we may take its (real line) Laplace transform, which we will denote by

F(s)=0estν(t)𝑑t.F(s)=\int_{0}^{\infty}e^{st}\mathcal{L}\nu(t)\,dt.

By the decay of ν,\mathcal{L}\nu, FF is well-defined and holomorphic on the half-space {s:Re(s)<c0}.\{s\colon\text{Re}(s)<c_{0}\}. Moreover, FF obeys the bound

(4.6) |F(s)|C|c0Re(s)|1, whenever Re(s)<c0.|F(s)|\leq C|c_{0}-\text{Re}(s)|^{-1},\,\,\text{ whenever }\text{Re}(s)<c_{0}.

On the other hand, by Fubini’s Theorem, we have the representation

(4.7) F(s)=𝕊11zs𝑑ν(z), whenever |s|<1.F(s)=\int_{\mathbb{S}^{1}}\frac{1}{z-s}\,d\nu(z),\,\,\text{ whenever }|s|<1.

The right-hand side of (4.7) can be further extended as an analytic function whenever ssupp(ν)s\not\in\text{supp}(\nu), as the set supp(ν)\mathbb{C}\setminus\text{supp}(\nu) is connected thanks to the additional hypothesis on the support of ν\nu. Thus, 𝕊11zs𝑑ν(z)\int_{\mathbb{S}^{1}}\frac{1}{z-s}\,d\nu(z) must agree with F(s)F(s) on the intersection between supp(ν)\mathbb{C}\setminus\text{supp}(\nu) and {Re(s)<c0}.\{\text{Re}(s)<c_{0}\}. We will also denote by F(s)F(s) the analytic function that continues over the union of both sets above.

Notice that, by this definition, we also have

(4.8) |F(s)|dist(s,supp(ν))1.|F(s)|\lesssim\text{dist}(s,\text{supp}(\nu))^{-1}.

In order to finish, we observe that we may replace the measure ν\nu by ν~=ν|A,\tilde{\nu}=\nu|_{A}, where A={z𝕊1:Re(z)c0},A=\{z\in\mathbb{S}^{1}\colon\text{Re}(z)\leq c_{0}\}, in each of the steps above. Let Fν~F_{\tilde{\nu}} be the function constructed in association with it. Then:

  1. (1)

    Fν~F_{\tilde{\nu}} is well-defined and holomorphic on {c0±i1c02};\mathbb{C}\setminus\left\{c_{0}\pm i\sqrt{1-c_{0}^{2}}\right\};

  2. (2)

    |Fν~(s)||s(c0+i1c02)|1+|s(c0i1c02)|1,s.|F_{\tilde{\nu}}(s)|\lesssim\left|s-(c_{0}+i\sqrt{1-c_{0}^{2}})\right|^{-1}+\left|s-(c_{0}-i\sqrt{1-c_{0}^{2}})\right|^{-1},\,\forall s\in\mathbb{C}.

    Indeed, If Re(s)0,\text{Re}(s)\leq 0, then the claim is trivial in light of (4.6). More generally, the claim follows by either (4.6), (4.7) or (4.8) whenever

    dist(s,supp(ν~))B,\text{dist}(s,\text{supp}({\tilde{\nu}}))\geq B,

    with B>0B>0 an absolute constant to be determined later. Thus, we may restrict ourselves to Re(s)>0,dist(s,suppν~)<B.\text{Re}(s)>0,\text{dist}(s,\text{supp}{\tilde{\nu}})<B. Let c0+i1c02=z0,c_{0}+i\sqrt{1-c_{0}^{2}}=z_{0}, for shortness.

    Consider first the region R1={s=z0+w,Re(s)>0,Im(w)>(c0+1)2|w|}.R_{1}=\{s=z_{0}+w,\text{Re}(s)>0,\text{Im}(w)>\frac{(c_{0}+1)}{2}|w|\}. In that region, the angle between ww and z0z_{0} is always strictly less than π/2,\pi/2, and thus we have

    |s|2=1+|w|2+2z0,w1+C(c0)|w|,|s|^{2}=1+|w|^{2}+2\langle z_{0},w\rangle\geq 1+C(c_{0})|w|,

    where we may write, in more explicit terms,

    C(c0)=1c02c0+12(1(c0+1)24)1/2c01c021c02.C(c_{0})=\sqrt{1-c_{0}^{2}}\frac{c_{0}+1}{2}-\left(1-\frac{(c_{0}+1)^{2}}{4}\right)^{1/2}c_{0}\geq\sqrt{1-c_{0}^{2}}\frac{1-c_{0}}{2}.

    Therefore, |s|1|w|=|sz0|,|s|-1\gtrsim|w|=|s-z_{0}|, and as |s|1=dist(s,𝕊1)|s|-1=\text{dist}(s,\mathbb{S}^{1}) for sR1,s\in R_{1}, we have the claim in that region from (4.8). Analogously, if we consider the region

    R2={s=z0+w,Re(s)>0,Im(w)<c0+12|w|,|w|1},R_{2}=\left\{s=z_{0}+w,\text{Re}(s)>0,\text{Im}(w)<-\frac{c_{0}+1}{2}|w|,|w|\ll 1\right\},

    we see that |s|21κ(c0)|w|,|s|^{2}\leq 1-\kappa(c_{0})|w|, and thus dist(s,𝕊1)=1|s||w|,\text{dist}(s,\mathbb{S}^{1})=1-|s|\gtrsim|w|, and the conclusion follows in the same manner.

    Now, if we let

    R3={s=z0+w,c0>Re(s)>0,Im(w)(c0+12|w|,c0+12|w|)},R_{3}=\left\{s=z_{0}+w,c_{0}>\text{Re}(s)>0,\text{Im}(w)\in\left(-\frac{c_{0}+1}{2}|w|,\frac{c_{0}+1}{2}|w|\right)\right\},

    we have |Re(w)|>(1(c0+1)24)1/2|w|.|\text{Re}(w)|>\left(1-\frac{(c_{0}+1)^{2}}{4}\right)^{1/2}|w|. In particular, |Re(s)c0|=|Re(w)||w|,|\text{Re}(s)-c_{0}|=|\text{Re}(w)|\gtrsim|w|, and (4.6) gives us the result once again. On the other hand, the estimate in the region R4={s=z0+w,Re(s)>c0,Im(w)(c0+12|w|,c0+12|w|)}R_{4}=\left\{s=z_{0}+w,\text{Re}(s)>c_{0},\text{Im}(w)\in\left(-\frac{c_{0}+1}{2}|w|,\frac{c_{0}+1}{2}|w|\right)\right\} follows directly from (4.8) and the fact that supp(ν~)𝕊1{Re(s)c0}.\text{supp}(\tilde{\nu})\subset\mathbb{S}^{1}\cap\{\text{Re}(s)\leq c_{0}\}.

    By repeating the same process above, but reflected, to the point z0¯\overline{z_{0}} shows the result in a neighbourhood of size δ(c0)>0\delta(c_{0})>0 of supp(ν~).\text{supp}(\tilde{\nu}). Let then B=δ(c0)B=\delta(c_{0}) in the beginning. This proves the claim.

    \cdotz0z_{0}\cdotz0¯\overline{z_{0}}R4R_{4}R1R_{1}R3R_{3}R2R_{2}
    Figure 2. The regions R1,R2,R3,R4R_{1},R_{2},R_{3},R_{4} as described in the proof above, with c0=0.7.c_{0}=0.7. Notice that we dropped the condition that dist(s,supp(ν~))\text{dist}(s,\text{supp}(\tilde{\nu})) is small for a clearer visualisation.
  3. (3)

    |sFν~(s)||sF_{\tilde{\nu}}(s)| is bounded as s,s\to\infty, which follows from (4.7).

From these properties, we are led to consider the function H(s)=(sz0)(sz0¯)Fν~(s).H(s)=(s-z_{0})(s-\overline{z_{0}})F_{\tilde{\nu}}(s). By the considerations above, HH is an entire function, bounded by a polynomial of degree 1. Thus, HH is itself a polynomial of degree 1, say, H(s)=αs+β.H(s)=\alpha s+\beta. Then

Fν~(s)=αs+β(sz0)(sz0¯)=γz0s+σz0¯s,F_{\tilde{\nu}}(s)=\frac{\alpha s+\beta}{(s-z_{0})(s-\overline{z_{0}})}=\frac{\gamma}{z_{0}-s}+\frac{\sigma}{\overline{z_{0}}-s},

for some γ,σ.\gamma,\sigma\in\mathbb{C}. In order to finish, we employ the a theorem of F. Riesz and M. Riesz [Gar07] saying that for two measures μ1\mu_{1} and μ2\mu_{2} on the unit circle we have

(4.9) 𝕊11zsd(μ1μ2)(z)=0s𝔻dμ1dμ2=ϕ(s)dm(s),\int_{\mathbb{S}^{1}}\frac{1}{z-s}\,d(\mu_{1}-\mu_{2})(z)=0\,\,\forall\,\,s\in\mathbb{D}\iff d\mu_{1}-d\mu_{2}=\phi(s)dm(s),

where ϕH1(𝔻),\phi\in H^{1}(\partial\mathbb{D}), and dmdm denotes the arclength measure on the circle 𝕊1.\mathbb{S}^{1}. Applied to our case, we obtain

ν~=γδz0+σδz0¯+ϕ(s)dm(s),\tilde{\nu}=\gamma\cdot\delta_{z_{0}}+\sigma\cdot\delta_{\overline{z_{0}}}+\phi(s)dm(s),

for some ϕH1(𝔻).\phi\in H^{1}(\partial\mathbb{D}). But supp(ν~γδz0σδz0¯)𝕊1{Re(z)c0}.\text{supp}(\tilde{\nu}-\gamma\cdot\delta_{z_{0}}-\sigma\cdot\delta_{\overline{z_{0}}})\subset\mathbb{S}^{1}\cap\{\text{Re}(z)\leq c_{0}\}. This shows that ϕ\phi vanishes on the arc joining z0z_{0} and z0¯.\overline{z_{0}}. Classical uniqueness result for functions from the Hardy space implies that ϕ0.\phi\equiv 0. This implies that

supp(ν)𝕊1{Re(z)c0},\text{supp}(\nu)\subset\mathbb{S}^{1}\cap\{\text{Re}(z)\geq c_{0}\},

which finishes the proof of Lemma 4.2. ∎

With Lemma 4.2 at our disposal, we may proceed to the proof of Theorem 1.5.

Proof of Theorem 1.5.

We first rewrite what we wish to prove in terms of Laplace transforms of measures on the circle, as done above.

For a finite measure μ\mu on the interval [0,1],[0,1], let π1:𝕊1[1,1]\pi_{1}:\mathbb{S}^{1}\to[-1,1] denote the projection onto the first coordinate, and let p:[0,1]A={z=eiθ,θ[0,π/2]}p:[0,1]\to A=\{z=e^{i\theta},\,\theta\in[0,\pi/2]\} denote the inverse of this map restricted to the set A.A. We then let

ν=p(μ)\nu=p_{\ast}(\mu)

be the pushforward measure of μ\mu to the circle through p.p. We readily see that this measure is finite, and supp(ν)A𝕊1{Re(s)0}.\text{supp}(\nu)\subset A\subset\mathbb{S}^{1}\cap\{\text{Re}(s)\geq 0\}. Finally, from the definition of ν\nu and φ,\varphi, we may write

φ(x)=𝕊1ezπ|x|2𝑑ν(z)=ν(π|x|2).\varphi(x)=\int_{\mathbb{S}^{1}}e^{-z\pi|x|^{2}}\,d\nu(z)=\mathcal{L}\nu(\pi|x|^{2}).

Proof of Part (1). As φEa2,\varphi\in E^{2}_{a}, the main result shows that φEaϵ,ϵ>0.\varphi\in E^{\infty}_{a-\epsilon},\,\forall\,\epsilon>0. But, by the correspondence above, we have |ν(t)|Ce(aϵ)t,|\mathcal{L}\nu(t)|\leq Ce^{-(a-\epsilon)t}, for any ϵ>0.\epsilon>0. By Lemma 4.2, supp(ν)𝕊1{Re(s)aϵ},ϵ>0.\text{supp}(\nu)\subset\mathbb{S}^{1}\cap\{\text{Re}(s)\geq a-\epsilon\},\,\forall\,\,\epsilon>0. Thus, supp(ν)𝕊1{Re(s)a}.\text{supp}(\nu)\subset\mathbb{S}^{1}\cap\{\text{Re}(s)\geq a\}. But this is equivalent to supp(μ)[a,1].\text{supp}(\mu)\subset[a,1]. This plainly implies that |φ(x)|eaπ|x|2L().|\varphi(x)|e^{a\pi|x|^{2}}\in L^{\infty}(\mathbb{R}).

On the other hand, we see that φ^\widehat{\varphi} may be written as

φ^(ξ)=01𝒢r(ξ)¯1(r+i1r2)1/2dμ(r)=:01𝒢r(ξ)¯dμ~(r).\widehat{\varphi}(\xi)=\int_{0}^{1}\overline{\mathcal{G}_{r}(\xi)}\cdot\frac{1}{(r+i\sqrt{1-r^{2}})^{1/2}}\,d\mu(r)=:\int_{0}^{1}\overline{\mathcal{G}_{r}(\xi)}\,d\tilde{\mu}(r).

The measure μ~\tilde{\mu} is again a finite measure, and one can repeat the argument above, now using the “conjugate” map p¯:[0,1]A¯={z=eiθ,θ[π/2,0]}\overline{p}:[0,1]\to\overline{A}=\{z=e^{i\theta},\,\theta\in[-\pi/2,0]\} to define the pushforward measure. This directly implies that |φ^(ξ)|eaπ|ξ|2L(),|\widehat{\varphi}(\xi)|e^{a\pi|\xi|^{2}}\in L^{\infty}(\mathbb{R}), which finally shows that φEa,\varphi\in E^{\infty}_{a}, as desired.

Proof of Part (2). We first notice that for β{π/4kπ/2,k},\beta\not\in\{-\pi/4-k\pi/2,k\in\mathbb{Z}\}, Theorem 1.1 covers this part. Thus, we may suppose without loss of generality that β=π4kπ2,k.\beta=-\frac{\pi}{4}-\frac{k\pi}{2},\,k\in\mathbb{Z}.

If φEa,\varphi\in E^{\infty}_{a}, then either Theorem 1.1 – or even Vemuri’s Theorem 3.1 – shows that π/4kπ/2φEtanh(α)ϵ,ϵ>0,k.\mathcal{F}_{-\pi/4-k\pi/2}\varphi\in E^{\infty}_{\tanh(\alpha)-\epsilon},\,\forall\,\,\epsilon>0,\,\forall\,\,k\in\mathbb{Z}. Now, the computation on (4.2) with λ=r+i1r2,β=π/4+kπ/2\lambda=r+i\sqrt{1-r^{2}},\,\beta=-\pi/4+k\pi/2 shows that

(4.10) |π/4kπ/2𝒢r(x)|exp(rπ|x|21(1)k1r2).|\mathcal{F}_{-\pi/4-k\pi/2}\mathcal{G}_{r}(x)|\lesssim\exp\left(-\frac{r\pi|x|^{2}}{1-(-1)^{k}\sqrt{1-r^{2}}}\right).

On the other hand, as φEtanh(2α),\varphi\in E^{\infty}_{\tanh(2\alpha)}, using the same argument as in the proof of Part (1) above, we have that supp(μ)[tanh(2α),1].\text{supp}(\mu)\subset[\tanh(2\alpha),1]. Thus, as we know that the functions of r[0,1]r\in[0,1]

rr11r2,rr1+1r2r\mapsto\frac{r}{1-\sqrt{1-r^{2}}},\,\,\,\,r\mapsto\frac{r}{1+\sqrt{1-r^{2}}}

are, respectively, decreasing and increasing, (4.10) implies that

|π/4kπ/2φ(x)|\displaystyle|\mathcal{F}_{-\pi/4-k\pi/2}\varphi(x)| μTVmax{exp(aπ|x|21+1a2),exp(π|x|2)}\displaystyle\lesssim\|\mu\|_{TV}\max\left\{\exp\left(-\frac{a\pi|x|^{2}}{1+\sqrt{1-a^{2}}}\right),\exp(-\pi|x|^{2})\right\}
=μTVmax{exp(tanh(α)π|x|2),exp(π|x|2)},\displaystyle=\|\mu\|_{TV}\max\left\{\exp(-\tanh(\alpha)\pi|x|^{2}),\exp(-\pi|x|^{2})\right\},

as a=tanh(2α).a=\tanh(2\alpha). This concludes the proof of Part (2), and thus also that of Theorem 1.5. ∎

Remark 4.3.

Using the notation employed in the statement of Conjecture 3.4, we have, for given tt\in\mathbb{R} and ε>0,\varepsilon>0, that 4πtfEΩα(t)ε\mathcal{F}_{-4\pi t}f\in E_{\Omega_{\alpha}(t)-\varepsilon}^{\infty} for any t.t\in\mathbb{R}. A calculation using (4.2) and the same strategy as in Part (2) of the proof of Theorem 1.5 above shows that, in fact, Φ(,t)EΩα(t).\Phi(\cdot,t)\in E_{\Omega_{\alpha}(t)}^{\infty}. This shows the validity of Conjecture 3.4 for the class of Laplace transforms with support on the circle discussed above.

Acknowledgements

We would like to express our gratitude towards Prof. Dimitar K. Dimitrov, who made several comments and remarks which helped shape the manuscript, to Prof. Sundaram Thangavelu, who pointed to us some results related to the work of Vemuri which led to its higher dimensional case, and to Biagio Cassano, for directing our attention to closely related work. We would also like to thank Danylo Radchenko and Prof. Christoph M. Thiele for valuable discussions regarding the applications to different contexts and the endpoint result.

A.K. was supported by Grant 275113 of the Research Council of Norway. J.P.G.R. acknowledges financial support by the European Research Council under the Grant Agreement No. 721675 “Regularity and Stability in Partial Differential Equations (RSPDE)”.

References

  • [AB77] W. O. Amrein and A. M. Berthier. On support properties of LpL^{p}-functions and their Fourier transforms. J. Functional Analysis, 24(3):258–267, 1977.
  • [BCF22] Juan Antonio Barceló, Biagio Cassano, and Luca Fanelli. Mass propagation for electromagnetic schrödinger evolutions. Nonlinear Analysis, 217:112734, 2022.
  • [BCK10] Jean Bourgain, Laurent Clozel, and Jean-Pierre Kahane. Principe d’Heisenberg et fonctions positives. Ann. Inst. Fourier (Grenoble), 60(4):1215–1232, 2010.
  • [BD06] Aline Bonami and Bruno Demange. A survey on uncertainty principles related to quadratic forms. Collect. Math., (Vol. Extra):1–36, 2006.
  • [Bec75] William Beckner. Inequalities in Fourier analysis. Ann. of Math. (2), 102(1):159–182, 1975.
  • [Ben85] Michael Benedicks. On Fourier transforms of functions supported on sets of finite Lebesgue measure. J. Math. Anal. Appl., 106(1):180–183, 1985.
  • [BFM03] S. Ben Farah and K. Mokni. Uncertainty principle and the LpL^{p}-LqL^{q}-version of Morgan’s theorem on some groups. Russ. J. Math. Phys., 10(3):245–260, 2003.
  • [CEK+10] M. Cowling, L. Escauriaza, C. E. Kenig, G. Ponce, and L. Vega. The Hardy uncertainty principle revisited. Indiana Univ. Math. J., 59(6):2007–2025, 2010.
  • [CF15] B. Cassano and L. Fanelli. Sharp Hardy uncertainty principle and Gaussian profiles of covariant Schrödinger evolutions. Trans. Amer. Math. Soc., 367(3):2213–2233, 2015.
  • [CF17] B. Cassano and L. Fanelli. Gaussian decay of harmonic oscillators and related models. Journal of Mathematical Analysis and Applications, 456(1):214–228, 2017.
  • [CG19] Henry Cohn and Felipe Gonçalves. An optimal uncertainty principle in twelve dimensions via modular forms. Invent. Math., 217(3):799–831, 2019.
  • [CP83] Michael Cowling and John F. Price. Generalisations of Heisenberg’s inequality. In Harmonic analysis (Cortona, 1982), volume 992 of Lecture Notes in Math., pages 443–449. Springer, Berlin, 1983.
  • [EKPV06] L. Escauriaza, C. E. Kenig, G. Ponce, and L. Vega. On uniqueness properties of solutions of Schrödinger equations. Comm. Partial Differential Equations, 31(10-12):1811–1823, 2006.
  • [EKPV08a] L. Escauriaza, C. E. Kenig, G. Ponce, and L. Vega. Convexity properties of solutions to the free Schrödinger equation with Gaussian decay. Math. Res. Lett., 15(5):957–971, 2008.
  • [EKPV08b] L. Escauriaza, C. E. Kenig, G. Ponce, and L. Vega. Hardy’s uncertainty principle, convexity and Schrödinger evolutions. J. Eur. Math. Soc. (JEMS), 10(4):883–907, 2008.
  • [EKPV10] Luis Escauriaza, Carlos E Kenig, Gustavo Ponce, and Luis Vega. The sharp hardy uncertainty principle for schrödinger evolutions. Duke Mathematical Journal, 155(1):163–187, 2010.
  • [EKPV16] L. Escauriaza, C. E. Kenig, G. Ponce, and L. Vega. Hardy uncertainty principle, convexity and parabolic evolutions. Comm. Math. Phys., 346(2):667–678, 2016.
  • [Gao16] Xin Gao. On Beurling’s uncertainty principle. Bull. Lond. Math. Soc., 48(2):341–348, 2016.
  • [Gar07] John B. Garnett. Bounded analytic functions, volume 236 of Graduate Texts in Mathematics. Springer, New York, first edition, 2007.
  • [GOeSR20] Felipe Gonçalves, Diogo Oliveira e Silva, and João P. G. Ramos. New sign uncertainty principles. arxiv preprint available at arXiv:2003.10771v3, 2020.
  • [GOeSR21] Felipe Gonçalves, Diogo Oliveira e Silva, and João P. G. Ramos. On regularity and mass concentration phenomena for the sign uncertainty principle. J. Geom. Anal., 31(6):6080–6101, 2021.
  • [GOeSS17] Felipe Gonçalves, Diogo Oliveira e Silva, and Stefan Steinerberger. Hermite polynomials, linear flows on the torus, and an uncertainty principle for roots. J. Math. Anal. Appl., 451(2):678–711, 2017.
  • [Gon19] Felipe Gonçalves. Orthogonal polynomials and sharp estimates for the Schrödinger equation. Int. Math. Res. Not. IMRN, (8):2356–2383, 2019.
  • [H9̈1] Lars Hörmander. A uniqueness theorem of Beurling for Fourier transform pairs. Ark. Mat., 29(2):237–240, 1991.
  • [Har33] G. H. Hardy. A Theorem Concerning Fourier Transforms. J. London Math. Soc., 8(3):227–231, 1933.
  • [Hed12] Haakan Hedenmalm. Heisenberg’s uncertainty principle in the sense of Beurling. J. Anal. Math., 118(2):691–702, 2012.
  • [KZ92] M.K. Kwong and A. Zettl. Norm Inequalities for Derivatives and Differences. Number Nr. 1536 in Lecture notes in mathematics. Springer-Verlag, 1992.
  • [Mor34] G. W. Morgan. A Note on Fourier Transforms. J. London Math. Soc., 9(3):187–192, 1934.
  • [Tak91a] Shin Takagi. Quantum Dynamics and Non-Inertial Frames of Reference. I: Generality. Progress of Theoretical Physics, 85(3):463–479, 03 1991.
  • [Tak91b] Shin Takagi. Quantum Dynamics and Non-Inertial Frames of References. II: Harmonic Oscillators. Progress of Theoretical Physics, 85(4):723–742, 04 1991.
  • [Vem08] M. K. Vemuri. Hermite expansions and hardy’s theorem. arxiv preprint available at arXiv:0801.2234v1, 2008.