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

On maximizers of convolution operators in LpL_{p} spaces

Gleb Kalachev111 Moscow State University, Russia; email: gleb.kalachev@yandex.ru  and Sergey Sadov222Moscow, Russia; email: serge.sadov@gmail.com
Abstract

We consider a convolution operator in d\mathbb{R}^{d} with kernel in LqL_{q} acting from LpL_{p} to LsL_{s}, where 1/p+1/q=1+1/s1/p+1/q=1+1/s. The main theorem states that if 1<q,p,s<1<q,p,s<\infty, then there exists an LpL_{p} function of unit norm on which the ss-norm of the convolution is attained. A number of questions, related to the statement and proof of the main theorem, are discussed. Also the problem of computing best constants in the Hausdorff-Young inequality for the Laplace transform, which prompted this research, is considered.


Keywords: convolution, Young inequality, existence of extremizer, concentration compactness, tight sequence, Laplace transform, best constants.


MSC 44A35, 49J99, 44A10, 41A44

1 Introduction

Let Lp(d)L_{p}(\mathbb{R}^{d}) denote the Lebesgue space of measurable complex-valued functions with norm fp=(|f|p)1/p\|f\|_{p}=(\int|f|^{p})^{1/p}, where 1p<1\leq p<\infty, or with norm f=sup{a>0||f(x)|aa.e.}\|f\|_{\infty}=\sup\{a>0\,|\,|f(x)|\leq a\;\text{a.e.}\}. Throughout, p=(11/p)1p^{\prime}=(1-1/p)^{-1} denotes the conjugate exponent. We consider a convolution operator Kk:fkfK_{k}:\,f\mapsto k*f with kernel kLq(d)k\in L_{q}(\mathbb{R}^{d}),

Kkf(x)=k(xy)f(y)𝑑y.K_{k}f(x)=\int k(x-y)f(y)\,dy.

As long as there is no ambiguity, we use shorthand notation: LpL_{p} instead of Lp(d)L_{p}(\mathbb{R}^{d}), f\int f instead of f(x)𝑑x\int f(x)\,dx, and KK instead of KkK_{k}. (If kλ()k_{\lambda}(\cdot) is a family of kernels depending on a parameter, we write KλK_{\lambda} instead of KkλK_{k_{\lambda}}.) In formulations and proofs of statements it is assumed that the dimension dd and the kernel kk are fixed.

Let 1p,q,r1\leq p,q,r\leq\infty. If kLqk\in L_{q} and the relation

1p+1q+1r=2\frac{1}{p}+\frac{1}{q}+\frac{1}{r}=2 (1.1)

holds, then the operator KK acts boundedly from LpL_{p} to LrL_{r^{\prime}} and its norm (to be called the (p,r)(p,r)-norm333Emphasizing that Kp,r\|K\|_{p,r} is the extremum of the symmetric bilinear form Kp,r=Kr,p=sup|k(x+y)f(x)g(y)𝑑x𝑑y|;fp=gr=1\|K\|_{p,r}=\|K\|_{r,p}=\sup\big{|}\int k(x+y)f(x)g(y)\,dx\,dy\big{|};\,\left\|f\right\|_{p}=\left\|g\right\|_{r}=1. ) has an upper bound given by Young’s inequality Kp,rkq\|K\|_{p,r}\leq\|k\|_{q}. A function fLpf\in L_{p} is called a maximizer of the convolution operator KK with respect to the pair of exponents (p,r)(p,r) if fp=1\|f\|_{p}=1 and kfr=Kp,r\|k*f\|_{r^{\prime}}=\|K\|_{p,r}.

The main result of the paper is the theorem on existence of a maximizer.

Theorem 1.

Let 1<p,q,r<1<p,q,r<\infty and the relation (1.1) holds. Then for any kernel kLqk\in L_{q} there exists a maximizer of the operator KK with respect to the pair of exponents (p,r)(p,r).

In a narrow sense, the only predecessor of this result that we are aware of is the paper by M. Pearson [25]. In it, the existence of a maximizer is proved under the following assumptions: the function kk is radially symmetric, nonnegative, decreasing away from the origin; besides, an “extra room of integrability”, kLq+εLqεk\in L_{q+\varepsilon}\cap L_{q-\varepsilon}, is required.

In a wider sense, our paper’s context is related, on the one hand (from the motivation side), to the business of sharp constant in analytical inequalities, and on the other hand, to techniques of proving the existence of extremizers in variational problems with non-compact groups of symmetries.

The starting point of this work was computation of the norms of the Laplace transorm as an operator from LpL_{p} to LpL_{p^{\prime}} (1p21\leq p\leq 2) on +\mathbb{R}_{+} (that is, of sharp constants in G.H. Hardy’s [16] inequality), reported here in Section 6. An equivalent problem is to compute the corresponding norms of the convolution operators on 1\mathbb{R}^{1} with kernels hp(x)=exp(x/pex)h_{p}(x)=\exp(x/p^{\prime}-e^{x}). Hardy’s original estimate can be improved by making use of W. Beckner’s sharp form [6] of Young’s inequality(8){}^{(\mbox{\it\ref{com:Beckner}})}.444The labels (8){}^{(\mbox{\it\ref{com:Beckner}})}, (8){}^{(\mbox{\it\ref{com:Christ}})} etc. refer to the comments (8), (8) etc. in Section 8.

Kp,r(ApAqAr)dkq,whereAm=(m1/mm1/m)1/2,\|K\|_{p,r}\leq(A_{p}A_{q}A_{r})^{d}\,\|k\|_{q},\quad\text{where}\quad A_{m}=\left(\frac{m^{1/m}}{{m^{\prime}}^{1/m^{\prime}}}\right)^{1/2}, (1.2)

The so improved estimate — see (6.8) — is found in the 2005’s M. Sc. Thesis of E. Setterqvist [27, Theorem 2.2].

The equality in Beckner’s inequality (1.2) takes place only in the case of a Gaussian kernel kk. Furhter analytical enhancement of the estimate (6.8) should in principle be possible by using recent subtle results of M. Christ [11, 12](8){}^{(\mbox{\it\ref{com:Christ}})}.

Our numerical results on the norms Khpp,p\|K_{h_{p}}\|_{p,p} (lacking full justification) allow one to make judgement about comparative strength of the available analytical estimates (see Section 6). For 1<p<21<p<2, the existence of maximizers was first observed experimentally. (If p=2p=2, there is no maximizer as one can easily see.) Since the kernel hph_{p} is not symmetric, Pearson’s theorem is not applicable. The question about possible weaker conditions sufficient for the existence of a maximizer naturally presented itself. Somewhat surprisingly, it turned out that no artificial conditions are needed at all.

A difficulty in proving the existence of a maximizer of a convolution operator, as well as in similar situations, owes to the problem’s invariance with respect to a non-compact group of transformations (the additive group d\mathbb{R}^{d} of translations, in our case). A natural attempt is to begin with some normalized maximizing sequence (fn)(f_{n}), fnp=1\|f_{n}\|_{p}=1, KfnrK\|Kf_{n}\|_{r^{\prime}}\to\|K\|, and, referring to the Banach-Alaoglu theorem, to find a weakly converging subsequence (fn(m))(f_{n(m)}). There is no chance to prove strong convergence of the subsequence (fn(m))(f_{n(m)}) since it may “run away” to infinity, thus weakly converging to zero.

However one can take shifted functions f~n=fn(an)\tilde{f}_{n}=f_{n}(\cdot-a_{n}) hoping to chose the shifts ana_{n} so as to obtain a relatively compact sequence (f~n)(\tilde{f}_{n}). (In terms of inequalities it amounts to establishing suitable uniform estimates). If this idea works out, a weakly converging subsequence of the shifted sequence will converge strongly and its limit will be a maximizer.

Person’s proof exploits the fact that in the case of a radially symmetric kernel the functions f~n\tilde{f}_{n} can also be made radially symmetric. This is due to M. Riesz’s inequality for nondecreasing rearrangements [19, Theorem 3.7]. We do not know of any substitute (analog, generalization) for Riesz’s inequality in the case of a non-symmetric kernel, which would allow a generalization of Pearson’s argument. Availability of full analytical control of the kernel in all directions (like in the case of hph_{p}) does not help. Our approach is completely different. Absent a natural reference point, we develop an intrinsic way to describe localization of near-maximizers.

The scheme and some elements of our proof exhibit clear similarities with the concentration compactness method of P. L. Lions [20]. Moreover, the class of variational problems described in the introductory section of Lions’ paper [20] includes the problem of finding the (p,r)(p,r)-norm of a convolution operator as one of the simplest representatives of the class.(8){}^{(\mbox{\it\ref{com:Lions-intro-problem}})}.

For this reason one may wonder at the absence (to the best of our knowledge) of Theorem 1 in the existing literature.(8){}^{(\mbox{\it\ref{com:Tao-toy-problem}})} Note though that the proof presented here does not depend on Lions’ work, whether directly or indirectly.

Let us outline the paper’s structure. The proof of Theorem 1 is given in Sections 24. In Section 2 we introduce relevant terminology and describe the proof “in the large”. Properties required at the steps of the proof are mentioned and references to the places where they are treated in detail are given. Detailed formulations and proofs of the required properties as well as auxiliary intermediate results are contained in Sections  3 and 4. The key Lemmas 3.53.7 of Section 3 provide uniform control on the size of “near-supports” of member functions of a maximizing sequence, to exclude a possibility of diffusion. (It helps to always ask in the course of the proof, where it fails when q=1q=1, p=r=2p=r=2, in which case there is no maximizer if k(x)>0k(x)>0). The lemmas of Section 4 deal with compactness and shifts.

In Section 5, a number of diverse results related to Theorem 1 and its proof are treated. In particular, we discuss the cases of exponents 11 and \infty, the equation satisfied by a maximizer, the lower bounds for convolution operators, etc. A survey of the results is given in Subsection 5.0.

Section 6 is devoted to computation of the norms of the Laplace transform from Lp(+)L_{p}(\mathbb{R}_{+}) to Lp(+)L_{p^{\prime}}(\mathbb{R}_{+}); the equivalent problem being, as mentioned before, the calculation of the norms of convolutions with kernels hp(x)h_{p}(x) as operators from Lp()L_{p}(\mathbb{R}) to Lp()L_{p^{\prime}}(\mathbb{R}). The obtained numerical results are compared with several analytical estimates. The numerical method used is straightforward; however, its convergence remains an empirical fact.

The short Section 7 contains some open questions and conjectures that lie close to the paper’s contents.

In the final Section 8, bibliographical and terminological comments are gathered.

———————

The following equivalent forms of the relation (1.1) will be used repeatedly as convenient. Let {x,y,z}={p,q,r}\{x,y,z\}=\{p,q,r\} in any order. Then

1x+1y=1+1z,1x+1y=1z,xz.\frac{1}{x}+\frac{1}{y}=1+\frac{1}{z^{\prime}},\qquad\frac{1}{x^{\prime}}+\frac{1}{y^{\prime}}=\frac{1}{z},\qquad x^{\prime}\geq z.

Consequently, min(p,q,r)max(p,q,r),\min(p^{\prime},q^{\prime},r^{\prime})\geq\max(p,q,r), with equality if and only if 1{p,q,r}1\in\{p,q,r\}.

2 Preliminaries and the proof in the large

The exponents q,p,rq,p,r and the convolution kernel kLqk\in L_{q} are assumed fixed.

By definition of the norm of the operator KK, for any ε>0\varepsilon>0 there exists a function fεLpf_{\varepsilon}\in L_{p} such that fεp=1\left\|f_{\varepsilon}\right\|_{p}=1 and KfεrKp,r(1ε)\left\|Kf_{\varepsilon}\right\|_{r^{\prime}}\geq\left\|K\right\|_{p,r}(1-\varepsilon). Any such function will be called an ε\varepsilon-maximizer.

Definition 2.1.

A sequence (fn)(f_{n}) of norm one functions from LpL_{p} is a maximizing sequence (for the operator KK) if KfnrKp,r\left\|Kf_{n}\right\|_{r^{\prime}}\to\left\|K\right\|_{p,r} as nn\to\infty.

The proof of Theorem 1 aims, quite obviously, at constructing a maximizing sequence for the operator KK that converges in norm. The limit will be a norm one function and a maximizer. We will start out with an arbitrarily chosen maximizing sequence, apply to its members certain “improving operators” and shifts (translations), and finally select a suitable, strongly convergent subsequence.

Note a few trivial but important properties.

(1) If a function ff is an ε\varepsilon-maximizer and ε1>ε\varepsilon_{1}>\varepsilon, then ff is also an ε1\varepsilon_{1}-maximizer.

(2) The set of ε\varepsilon-maximizers is shift-invariant (since convolution operators commute with shifts).

(3) Any subsequence of a maximizing sequence is itself a maximizing sequence.

Let us introduce some notions to be used in the proof: a special ε\varepsilon-maximizer, δ\delta-near-centered function, a tight LpL_{p} sequence, and a few more.(8){}^{(\mbox{\it\ref{com:def-tight}})}

Definition 2.2.

Let fLpf\in L_{p} be an ε\varepsilon-maximizer. In Section 4.2 we define a nonlinear improving operator B:LpLpB:\,L_{p}\to L_{p}, such that Bfp=1\|Bf\|_{p}=1 and kBfrkfr\|k*Bf\|_{r^{\prime}}\geq\|k*f\|_{r^{\prime}}, so that BfBf is also an ε\varepsilon-maximizer (Lemma 4.3). The so obtained ε\varepsilon-maximizers will be called special ε\varepsilon-maximizers.

The operator BB appears naturally in the necessary condition of extremum, i.e. the equation that must be satisfied by a maximizer if one exists (Section 5.4).

A crucial property of the operator BB that justifies the qualifier “improving” (as opposed to “not-worsening”, say) is the fact that the BB-image of a weakly convergent sequence of LpL_{p} functions converges in LpL_{p} norm on bounded sets (and moreover, on any sets of finite measure) in d\mathbb{R}^{d} (Lemma 4.11).

Definition 2.3.

Given a function fLpf\in L_{p} and a unit vector555We use the standard Euclidean inner product and the Euclidean norm in d\mathbb{R}^{d}. vdv\in\mathbb{R}^{d}, the δ\delta-diameter of the function ff of order pp in the direction vv is the nonnegative number

Dδ,vp(f)=infb>a{ba|a<(x,v)<b|f|pfppδ}.D^{p}_{\delta,v}(f)=\inf_{b>a}\left\{b-a\,\left|\,\int_{a<(x,v)<b}|f|^{p}\geq\|f\|_{p}^{p}-\delta\right.\right\}.

In this formula we implicitly assume that δ<fpp\delta<\|f\|_{p}^{p}. If this is not the case (in particular, if f=0f=0) we define Dδ,vp(f)=0D^{p}_{\delta,v}(f)=0.

Suppose aa and bb are such that ba=Dδ,vp(f)b-a=D^{p}_{\delta,v}(f) and a<(x,v)<b|f|p=fppδ\int_{a<(x,v)<b}|f|^{p}=\|f\|_{p}^{p}-\delta. The existence of such aa and bb is obvious.(8){}^{(\mbox{\it\ref{com:delta-support}})} We say that the segment [a,b][a,b] is the δ\delta-near-support of the function ff of order pp in the direction vv and denote it suppδ,vp(f)\mathrm{supp}\,^{p}_{\delta,v}(f). The function ff is δ\delta-near-centered of order pp in the direction vv if a0ba\leq 0\leq b and

a<(x,v)<0|f|p=0<(x,v)<b|f|p=fppδ2.\int_{a<(x,v)<0}|f|^{p}=\int_{0<(x,v)<b}|f|^{p}=\frac{\left\|f\right\|_{p}^{p}-\delta}{2}.

Let us fix, once for all, an orthonormal basis {e1,,ed}\{e_{1},\dots,e_{d}\} in d\mathbb{R}^{d} (that is, fix the coordinate axes).

We say that the function ff is δ\delta-near-centered of order pp if ff is δ\delta-near-centered of order pp in the direction eje_{j} for j=1,,dj=1,\dots,d.

Clearly, any function can be made δ\delta-near-centered by means of a suitable shift. However, different values of δ\delta may require different shifts.

The transformation of functions corresponding to the argument shift by a vector aa will be denoted TaT_{a}; that is, Taf(x)=f(xa)T_{a}f(x)=f(x-a).

Remark 2.1.

The above defined centering can be called a mass-centering. As a natural alternative, one might propose the geometric centering, whereby the δ\delta-near-support of a δ\delta-near-centered function in the given direction would be a symmetric interval [b,b][-b,b]. However this latter approach is not suitable for our proof because Lemma 4.12 would be lost.

Definition 2.4.

A sequence of functions fnLpf_{n}\in L_{p} is relatively tight if for any δ>0\delta>0 there exists n0=n0(δ)n_{0}=n_{0}(\delta) such that

supnn0supv=1Dδ,vp(fn)<.\sup_{n\geq n_{0}}\sup_{\|v\|=1}D^{p}_{\delta,v}(f_{n})<\infty.

A sequence of functions fnLpf_{n}\in L_{p} is tight if for any δ>0\delta>0 it is δ\delta-near-finite (of order pp), which means that there exist n0n_{0} and a cube QQ in d\mathbb{R}^{d} with edges parallel to the coordinate axes, such that for nn0n\geq n_{0}

Q|fn|pfnppδ.\int_{Q}|f_{n}|^{p}\geq\left\|f_{n}\right\|_{p}^{p}-\delta.

As it turns out (see Lemma 4.15), a relatively tight sequence is tight provided the δ\delta-finiteness property holds for just one sufficiently small (depending on supnfnp\sup_{n}\left\|f_{n}\right\|_{p}) positive δ\delta.

We are ready to present a high-level structure of the proof of Theorem 1.

Proof of Theorem 1 .

Introduce the following classes of function sequences in LpL_{p} defined in terms of imposed constraints.

  1. 1.

    The class Max\mathrm{Max} comprises all maximizing sequences (for the convolution operator K:LpLrK:\,L_{p}\to L_{r^{\prime}}).

  2. 2.

    The class SMax\mathrm{SMax} comprises all special maximizing sequences, that is, maximizing sequences of the form (Bfn)(Bf_{n}), where (fn)Max(f_{n})\in\mathrm{Max}.

  3. 3.

    The class RTgt\mathrm{RTgt} comprises all relatively tight sequences.

  4. 4.

    The class Tgt\mathrm{Tgt} comprises all tight sequences.

  5. 5.

    The class WCvg\mathrm{WCvg} comprises all weakly convergent sequences.

  6. 6.

    The class LCvg\mathrm{LCvg} comprises all locally convergent sequences, i.e. sequences converging in LpL_{p} norm on any bounded measurable subset of d\mathbb{R}^{d}.

  7. 7.

    The class Cvg\mathrm{Cvg} comprises all sequences converging in Lp(d)L_{p}(\mathbb{R}^{d}).

Note that a subsequence of a sequence that belongs to any of these classes also belongs to that class.

The construction.

1. We start out with an arbitrary sequence (fn)Max(f_{n})\in\mathrm{Max}.

2. Applying the operator BB to its members we get the sequence (Bfn)SMax(Bf_{n})\in\mathrm{SMax}.

3. In view of the inclusions MaxRTgt\mathrm{Max}\subset\mathrm{RTgt} (Corollary 3.10) and SMaxMax\mathrm{SMax}\subset\mathrm{Max} (Corollary 4.4), we have (Bfn)RTgt(Bf_{n})\in\mathrm{RTgt}.

4. Putting δ0=1/4\delta_{0}=1/4 (any δ0<1/3\delta_{0}<1/3 is as good), we can find vectors ana_{n} such that the shifted functions gn=Tan(Bfn)g_{n}=T_{a_{n}}(Bf_{n}) are δ0\delta_{0}-near-centered (Lemma 4.12).

5. Lemma 4.14 implies (gn)Tgt(g_{n})\in\mathrm{Tgt}.

6. The operator BB commutes with shifts. (It is a rather trivial fact, yet it is stated as Lemma 4.5). Consequently, gn=B(Tanfn)g_{n}=B(T_{a_{n}}f_{n}). The class Max\mathrm{Max} is shift-invariant, hence (gn)SMaxTgt(g_{n})\in\mathrm{SMax}\cap\mathrm{Tgt}.

7. (This is the most subtle step from the logic of proof viewpoint: we “undo” the operator BB in order to select a subsequence in the pre-image). The sequence (Tanfn)(T_{a_{n}}f_{n}) is bounded in LpL_{p}, hence it contains a weakly converging subsequence, which we denote (f^m)(\hat{f}_{m}), avoiding multilevel subscripts.

8. Put g^m=Bf^m\hat{g}_{m}=B\hat{f}_{m}. Then, on the one hand, (g^m)(\hat{g}_{m}) is a subsequence of the sequence (gn)(g_{n}) thus inheriting the class memberships of the latter. On the other hand, since (f^m)MaxWCvg(\hat{f}_{m})\in\mathrm{Max}\cap\mathrm{WCvg}, Lemma 4.11 implies (g^m)LCvg(\hat{g}_{m})\in\mathrm{LCvg}. As a result, (g^m)MaxLCvgTgt(\hat{g}_{m})\in\mathrm{Max}\cap\mathrm{LCvg}\cap\mathrm{Tgt}.

9. Applying Lemma 4.16 we conclude that (g^m)MaxCvg(\hat{g}_{m})\in\mathrm{Max}\cap\mathrm{Cvg}. Let h=limmg^mh=\lim_{m\to\infty}\hat{g}_{m}. Then hp=1\left\|h\right\|_{p}=1 and Khr=limKg^mr=Kp,r\left\|Kh\right\|_{r^{\prime}}=\lim\left\|K\hat{g}_{m}\right\|_{r^{\prime}}=\left\|K\right\|_{p,r} by continuity. The function hh is a maximizer.

Theorem 1 is proved modulo the statements referred to at the various steps of the construction. Proofs of all those statements are given in the next two sections.

The proof of Lemma 3.7, whose Corollary 4.4 is used at Step 3, is the longest. We devote the whole Section 3 to it and break the proof into short steps. The other results referred to in the above construction are proved in Section 4.

3 Estimates for δ\delta-diameters of near-maximizers

The main technical result of this Section is Lemma 3.7, while the conceptual conclusion is Corollary 3.10. We approach the proof of Lemma 3.7 through a chain of preparatory results, of which all but Lemma 3.5 are very simple.

The indicator function of a set Ω\Omega will be denoted IΩI_{\Omega}; if the set Ω\Omega is defined by means of a property (or predicate) PP, then the indicator function is written as IPI_{P}.

Lemma 3.1.

If γ>1\gamma>1, λ(0,1/2)\lambda\in(0,1/2) and u[λ,1λ]u\in[\lambda,1-\lambda], then

uγ+(1u)γ1κ,u^{\gamma}+(1-u)^{\gamma}\leq 1-\kappa,

where κ=κ(λ,γ)=2λ(121γ)>0\kappa=\kappa(\lambda,\gamma)=2\lambda\left(1-2^{1-\gamma}\right)>0.

Proof.

For γ>1\gamma>1 the function h(u)=uγ+(1u)γh(u)=u^{\gamma}+(1-u)^{\gamma} is convex and symmetric about u=1/2u=1/2. We may assume that u[λ,1/2]u\in[\lambda,1/2]. By convexity, we have the chain of inequalities

h(u)=h((12u)0+2u1/2)(12u)h(0)+2uh(1/2)==(12u)+2u2(1/2)γ=12u(121γ)12λ(121γ),h(u)=h((1-2u)\cdot 0+2u\cdot 1/2)\leq(1-2u)h(0)+2uh(1/2)=\\ =(1-2u)+2u\cdot 2(1/2)^{\gamma}=1-2u(1-2^{1-\gamma})\leq 1-2\lambda(1-2^{1-\gamma}),

which proves the Lemma.(8){}^{(\mbox{\it\ref{com:subadditivity_lemma}})}

Lemma 3.2.

Let Ω\Omega be a measure space, γ>1\gamma>1, and 0<λ<1/20<\lambda<1/2. Suppose that gL1(Ω)g\in L_{1}(\Omega) with norm g1=1\|g\|_{1}=1 is split into the sum g=g1+g2g=g_{1}+g_{2} and the summands satisfy gi1λ\|g_{i}\|_{1}\geq\lambda (i=1,2i=1,2) and g1g2=0g_{1}g_{2}=0. Then

g11γ+g21γ1κ,\|g_{1}\|_{1}^{\gamma}+\|g_{2}\|_{1}^{\gamma}\leq 1-\kappa,

with κ=κ(λ,γ)\kappa=\kappa(\lambda,\gamma), the same as in Lemma 3.1.

Proof.

Since g1g2=0g_{1}g_{2}=0, we have g11+g21=1\|g_{1}\|_{1}+\|g_{2}\|_{1}=1. It remains to apply Lemma 3.1 with u=g11u=\|g_{1}\|_{1}. ∎

Lemma 3.3.

Let R>0R>0 and fL1([R,R])f\in L_{1}([-R,R]). Then for any a<Ra<R there exists t0t_{0}, |t0|Ra|t_{0}|\leq R-a such that

12a|tt0|a|f(t)|1Rf1.\frac{1}{2a}\int_{|t-t_{0}|\leq a}|f(t)|\leq\frac{1}{R}\|f\|_{1}.
Proof.

Suppose aR/2a\leq R/2; otherwise the inequality is a tautology. The function h(t)=|xt|a|f(x)|h(t)=\int_{|x-t|\leq a}|f(x)| defined for |t|Ra|t|\leq R-a is continuous and satisfies the inequality

2(Ra)minh(t)|t|Rah(t)2af1.2(R-a)\,\min h(t)\leq\int_{|t|\leq R-a}h(t)\leq 2a\|f\|_{1}.

It suffices to choose t0t_{0} as the point of minimum, h(t0)=minh(t)h(t_{0})=\min h(t) and recall that 2(Ra)R2(R-a)\geq R.(8){}^{(\mbox{\it\ref{com:midpoint-average_lemma}})}

Lemma 3.4.

Let fLp(d)f\in L_{p}(\mathbb{R}^{d}). Given R>a>0R>a>0, a unit vector vdv\in\mathbb{R}^{d} and cc\in\mathbb{R}, there exists t0[c(Ra),c+(Ra)]t_{0}\in[c-(R-a),c+(R-a)] such that

12a|(v,x)t0|a|f(x)|p1R|(v,x)c|R|f(x)|p.\frac{1}{2a}\int_{|(v,x)-t_{0}|\leq a}|f(x)|^{p}\leq\frac{1}{R}\int_{|(v,x)-c|\leq R}|f(x)|^{p}.
Proof.

We may assume that v=(1,0,,0)v=(1,0,\dots,0) and c=0c=0. The result follows by applying Lemma 3.3 to the one-variable function x1|f(x)|p𝑑x2𝑑xdx_{1}\mapsto\int|f(x)|^{p}\,dx_{2}\dots dx_{d} considered on the interval x1[R,R]x_{1}\in[-R,R]. ∎

Definition 3.1.

Let AA be a map from LL to L~\tilde{L}, where LL and L~\tilde{L} are some spaces of measurable functions in d\mathbb{R}^{d}. Suppose a>0a>0 and a unit vector vv in d\mathbb{R}^{d} are given. We say that the map AA is an aa-expander in the direction vv if the property f(x)=0f(x)=0 a.e. for t1<(x,v)<t2t_{1}<(x,v)<t_{2}, where t1<t2+-\infty\leq t_{1}<t_{2}\leq+\infty, implies the property Af(x)=0Af(x)=0 a.e. for t1+a<(x,v)<t2at_{1}+a<(x,v)<t_{2}-a.

The next Lemma utilizes the notions and notation introduced in Definition 2.3.

Lemma 3.5.

Let AA be a linear bounded operator from Lp(d)L_{p}(\mathbb{R}^{d}) to Ls(d)L_{s}(\mathbb{R}^{d}), where 1p<s<1\leq p<s<\infty. Suppose AA is an aa-expander in the direction vv. Let fp=1\|f\|_{p}=1 and D=Dδ,vp(f)D=D_{\delta,v}^{p}(f). For any β>0\beta>0 there are two possibilities: (i) either D8βaD\leq 8\beta a or (ii) D>8βaD>8\beta a and

Afss<As(1κ+βγ),\left\|Af\right\|_{s}^{s}<\left\|A\right\|^{s}\left(1-\kappa+\beta^{-\gamma}\right), (3.1)

where γ=s/p\gamma=s/p and κ=κ(δ/2,γ)=δ(121γ)\kappa=\kappa(\delta/2,\gamma)=\delta(1-2^{1-\gamma}), consistent with notation in Lemma 3.1.

Proof.

The cases δ1\delta\geq 1 (where D=0D=0) and D8βaD\leq 8\beta a are trivial. Therefore we assume that δ<1\delta<1 and D>8βaD>8\beta a. Let cc and R2aR\geq 2a are such that

fI(v,x)>c+Rpp=fI(v,x)<cRppδ2.\left\|fI_{(v,x)>c+R}\right\|^{p}_{p}=\left\|fI_{(v,x)<c-R}\right\|^{p}_{p}\geq\frac{\delta}{2}.

Clearly, D2RD\leq 2R. By Lemma 3.4 there exists t0[c(R2a),c+(R2a)]t_{0}\in[c-(R-2a),c+(R-2a)] such that

|(v,x)t0|2a|f|p4aR|(v,x)c|R|f|p8aD<β1.\int\limits_{|(v,x)-t_{0}|\leq 2a}|f|^{p}\leq\frac{4a}{R}\int\limits_{|(v,x)-c|\leq R}|f|^{p}\leq\frac{8a}{D}<\beta^{-1}. (3.2)

Denote (see Fig. 1)

fl(x)=f(x)I(v,x)<t0,fr(x)=f(x)I(v,x)t0.f_{l}(x)=f(x)I_{(v,x)<t_{0}},\qquad f_{r}(x)=f(x)I_{(v,x)\geq t_{0}}.

Then f(x)=fl(x)+fr(x)f(x)=f_{l}(x)+f_{r}(x) and fl(x)fr(x)=0f_{l}(x)f_{r}(x)=0, flppδ/2\|f_{l}\|_{p}^{p}\geq\delta/2, frppδ/2\|f_{r}\|_{p}^{p}\geq\delta/2. Applying Lemma 3.2 to the pair of functions g1=|fl|pg_{1}=|f_{l}|^{p} and g2=|fr|pg_{2}=|f_{r}|^{p} we get(8){}^{(\mbox{\it\ref{com:prevent_splitting}})} flps+frps1κ(δ/2,γ)\|f_{l}\|_{p}^{s}+\|f_{r}\|_{p}^{s}\leq 1-\kappa(\delta/2,\gamma).

Introduce an yet another function,

fm(x)=f(x)I|(v,x)t0|<2a.f_{m}(x)=f(x)I_{|(v,x)-t_{0}|<2a}.

By (3.2), fmps<βγ\left\|f_{m}\right\|_{p}^{s}<\beta^{-\gamma}.

Refer to caption

Figure 1: Illustration of notation used in the proof of Lemma 3.5

The subsets in d\mathbb{R}^{d} defined by the inequalities

Ωl\displaystyle\Omega_{l} ={x|(v,x)<t0a},\displaystyle=\left\{x\,\left|\,(v,x)<t_{0}-a\right.\right\},
Ωm\displaystyle\Omega_{m} ={x||(v,x)t0|a},\displaystyle=\left\{x\,\left|\,|(v,x)-t_{0}|\leq a\right.\right\},
Ωr\displaystyle\Omega_{r} ={x|(v,x)>t0+a}\displaystyle=\left\{x\,\left|\,(v,x)>t_{0}+a\right.\right\}

are pairwise disjoint and ΩlΩmΩr=d\Omega_{l}\cup\Omega_{m}\cup\Omega_{r}=\mathbb{R}^{d}. We have

Afr=0 xΩl,\displaystyle Af_{r}=0\;\;\text{ }\;x\in\Omega_{l},
Afl=0 xΩr,\displaystyle Af_{l}=0\;\;\text{ }\;x\in\Omega_{r},
A(ffm)=0 xΩm.\displaystyle A(f-f_{m})=0\;\;\text{ }\;x\in\Omega_{m}.

Therefore

Afss\displaystyle\left\|Af\right\|_{s}^{s} =Ωl|Afl|s+Ωr|Afr|s+Ωm|Afm|s\displaystyle=\int_{\Omega_{l}}|Af_{l}|^{s}+\int_{\Omega_{r}}|Af_{r}|^{s}+\int_{\Omega_{m}}|Af_{m}|^{s}\leq
As(flps+frps+fmps)<\displaystyle\leq\left\|A\right\|^{s}\left(\left\|f_{l}\right\|_{p}^{s}+\left\|f_{r}\right\|_{p}^{s}+\left\|f_{m}\right\|_{p}^{s}\right)<
<As(1κ+βγ).\displaystyle<\left\|A\right\|^{s}\left(1-\kappa+\beta^{-\gamma}\right).

Q.E.D. ∎

Lemma 3.6.

Suppose the operator AA satisfies the assumptions of Lemma 3.5. Suppose also that fp=1\left\|f\right\|_{p}=1 and AfssAs(1τ)\left\|Af\right\|_{s}^{s}\geq\left\|A\right\|^{s}(1-\tau). Then for any δ>τ(121γ)1\delta>\tau(1-2^{1-\gamma})^{-1} the δ\delta-diameter D=Dδ,vp(f)D=D_{\delta,v}^{p}(f) satisfies the inequality

D8a(κτ)1/γ,κ=δ(121γ).D\leq 8a(\kappa-\tau)^{-1/\gamma},\qquad\kappa=\delta(1-2^{1-\gamma}).
Proof.

Put β=(κτ)1/γ\beta=(\kappa-\tau)^{-1/\gamma} and apply Lemma 3.5. Suppose the case (ii) takes place. Then

1τAfssAs<1κ+βγ=1τ,1-\tau\leq\frac{\left\|Af\right\|_{s}^{s}}{\left\|A\right\|^{s}}<1-\kappa+\beta^{-\gamma}=1-\tau,

a contradiction. Therefore the case (i) takes place and we are done. ∎

Lemma 3.7.

Let q>1q>1, kLq(d)k\in L_{q}(\mathbb{R}^{d}), and let K:LpLrK:\,L_{p}\to L_{r^{\prime}} be the convolution operator with kernel kk. Put γ=r/p>1\gamma=r^{\prime}/p>1. Suppose ε>0\varepsilon>0 and δ>εr(121γ)1\delta>\varepsilon r^{\prime}(1-2^{1-\gamma})^{-1} are given. If ρ>0\rho>0 is small enough, so that

ε+2ρ1/qKp,rδ121γr,\varepsilon+\frac{2\rho^{1/q}}{\left\|K\right\|_{p,r}}\leq\delta\frac{1-2^{1-\gamma}}{r^{\prime}},

then for any unit vector vdv\in\mathbb{R}^{d} and any ε\varepsilon-maximizer ff of the operator KK the inequality

Dδ,vp(f)cDρ,vq(k),D^{p}_{\delta,v}(f)\leq cD^{q}_{\rho,v}(k), (3.3)

holds with

c=4(δ(121γ)r(ε+2ρ1/qKp,r))1/γ.c=4\left(\delta(1-2^{1-\gamma})-r^{\prime}\left(\varepsilon+\frac{2\rho^{1/q}}{\left\|K\right\|_{p,r}}\right)\right)^{-1/\gamma}.
Remark 3.8.

The function ρDρ,vq(k)\rho\mapsto D^{q}_{\rho,v}(k) is nonincreasing. Hence for fixed ε\varepsilon and δ\delta, given two kernels kk and k~\tilde{k} with |k|=|k~||k|=|\tilde{k}| a.e., a weaker estimate (i.e. a smaller value of ρ\rho or larger value of cc in the r.h.s. of the inequality (3.3)) takes place for that of the two kernels with smaller norm of the corresponding convolution operator.

Proof.

Put M=Kp,rM=\left\|K\right\|_{p,r} and a=12Dρ,vq(k)a=\frac{1}{2}D_{\rho,v}^{q}(k). Without loss of generality we may assume that suppρ,vqk=[a,a]\mathrm{supp}\,_{\rho,v}^{q}k=[-a,a]. Let kρ=kI|(v,x)|ak_{\rho}=kI_{|(v,x)|\leq a} and KρK_{\rho} is the convolution operator with kernel kρk_{\rho}. We have kρkqq=ρ\left\|k_{\rho}-k\right\|_{q}^{q}=\rho and, by Young’s inequality, KρKp,rρ1/q\left\|K_{\rho}-K\right\|_{p,r}\leq\rho^{1/q}. In particular, Kρp,rM+ρ1/q\left\|K_{\rho}\right\|_{p,r}\leq M+\rho^{1/q}.

Fix an ε\varepsilon-maximizer fLpf\in L_{p} for the operator KK. We have KρfrKfrKρKp,rfpM(1ε)ρ1/q\left\|K_{\rho}f\right\|_{r^{\prime}}\geq\left\|Kf\right\|_{r^{\prime}}-\left\|K_{\rho}-K\right\|_{p,r}\left\|f\right\|_{p}\geq M(1-\varepsilon)-\rho^{1/q}. Therefore,

KρfrKρp,rM(1ε)ρ1/qM+ρ1/q>1ε2ρ1/qM.\frac{\left\|K_{\rho}f\right\|_{r^{\prime}}}{\left\|K_{\rho}\right\|_{p,r}}\geq\frac{M(1-\varepsilon)-\rho^{1/q}}{M+\rho^{1/q}}>1-\varepsilon-\frac{2\rho^{1/q}}{M}.

The operator KρK_{\rho} is an aa-expander in the direction vv. Let us apply Lemma 3.6 with A=KρA=K_{\rho} and s=rs=r^{\prime}. We have Afss=As(1τ)\left\|Af\right\|_{s}^{s}=\left\|A\right\|^{s}(1-\tau), where

1τ=(1ε2ρ1/qM)r>1r(ε+2ρ1/qM)>1δ(121γ)1-\tau=\left(1-\varepsilon-\frac{2\rho^{1/q}}{M}\right)^{r^{\prime}}>1-r^{\prime}\left(\varepsilon+\frac{2\rho^{1/q}}{M}\right)>1-\delta(1-2^{1-\gamma})

(due to the Bernoulli inequality and the inequality relating ε\varepsilon, δ\delta ρ\rho).

The estimate for DD provided by Lemma 3.6 yields (3.3). ∎

Corollary 3.9.

Let qq, kk and KK be as in Lemma 3.7. Suppose that ε\varepsilon, δ\delta, ρ\rho cc are related by the equalities

δ=4r121γε,ρ=(Kp,rε)q,c=4(εr)1/γ.\delta=\frac{4r^{\prime}}{1-2^{1-\gamma}}\varepsilon,\quad\rho=\left(\left\|K\right\|_{p,r}\varepsilon\right)^{q},\quad c=4(\varepsilon r^{\prime})^{-1/\gamma}. (3.4)

Then for any unit vector vdv\in\mathbb{R}^{d} and any ε\varepsilon-maximizer ff of the operator KK the estimate (3.3) holds.

Corollary 3.10.

Any maximizing sequence (fn)(f_{n}) of LpL_{p} functions for the convolution operator KK is relatively tight.

Indeed, let fnf_{n} be an εn\varepsilon_{n}-maximizer and εn0\varepsilon_{n}\to 0. Given a δ>0\delta>0 we define ε\varepsilon and ρ\rho by (3.4) and choose n0n_{0} in Definition 2.4 by the condition εnε\varepsilon_{n}\leq\varepsilon for nn0n\geq n_{0}.

4 Lemmas for construction of a convergent maximizing sequence

Recall that we always assume the relation 1/q+1/p+1/r=21/q+1/p+1/r=2.

Introduce the operation zzγ=z¯|z|γ1z\mapsto{z}^{\langle\gamma\rangle}=\overline{z}|z|^{\gamma-1}, where zz\in\mathbb{C}, γ\gamma\in\mathbb{R}, and the bar stands for complex conjugation. Thus, zzγ=|z|γ+1z{z}^{\langle\gamma\rangle}=|z|^{\gamma+1} and |zγ|=|z|γ|{z}^{\langle\gamma\rangle}|=|z|^{\gamma}.

4.1 Auxiliary numerical inequalities

Lemma 4.1.

For any u,vu,v\in\mathbb{C} the following inequalities hold:
(a) for 0<γ10<\gamma\leq 1,

|uγvγ|C|uv|γ,C=21γ;\left|{u}^{\langle\gamma\rangle}-{v}^{\langle\gamma\rangle}\right|\leq C|u-v|^{\gamma},\quad C=2^{1-\gamma}; (4.1)

(b) for γ>1\gamma>1,

|uγvγ|C|uv|(max(|u|,|v|))γ1,C=γ.\left|{u}^{\langle\gamma\rangle}-{v}^{\langle\gamma\rangle}\right|\leq C|u-v|\,\left(\max(|u|,|v|)\right)^{\gamma-1},\quad C={\gamma}. (4.2)

(This Lemma will be used in the proof of Lemma 4.6.)

Proof.

(a) Put u/v=reiϕu/v=re^{i\phi}. Due to the symmetry between uu and vv we may assume that r1r\leq 1. The inequality (4.1) reduces to the following:

|rγeiϕ1|C|reiϕ1|γ.\left|r^{\gamma}e^{i\phi}-1\right|\leq C\left|re^{i\phi}-1\right|^{\gamma}.

Using the Cosine Theorem and putting t=2r/(r2+1)1t=2r/(r^{2}+1)\leq 1, we can restate the required inequality in the form

r2γ+1(r2+1)γ21γtγcosϕC2(1tcosϕ)γ.\frac{r^{2\gamma}+1}{(r^{2}+1)^{\gamma}}-2^{1-\gamma}t^{\gamma}\cos\phi\leq C^{2}\left(1-t\cos\phi\right)^{\gamma}.

By concavity, (r2γ+1)/2((r2+1)/2)γ(r^{2\gamma}+1)/2\leq((r^{2}+1)/2)^{\gamma}, hence

r2γ+1(r2+1)γ21γtγcosϕ21γ(1tγcosϕ).\frac{r^{2\gamma}+1}{(r^{2}+1)^{\gamma}}-2^{1-\gamma}t^{\gamma}\cos\phi\leq 2^{1-\gamma}(1-t^{\gamma}\cos\phi).

If cosϕ0\cos\phi\geq 0, then

1tγcosϕ1tcosϕ(1tcosϕ)γ,1-t^{\gamma}\cos\phi\leq 1-t\cos\phi\leq(1-t\cos\phi)^{\gamma},

and (4.1) holds (even with a better constant), since 21γ<C22^{1-\gamma}<C^{2}.

If cosϕ<0\cos\phi<0, then, again due to concavity, we get

1tγcosϕ1+|tcosϕ|γ21γ(1+|tcosϕ|)γ=21γ(1tcosϕ)γ,1-t^{\gamma}\cos\phi\leq 1+|t\cos\phi|^{\gamma}\leq 2^{1-\gamma}(1+|t\cos\phi|)^{\gamma}=2^{1-\gamma}(1-t\cos\phi)^{\gamma},

and the proof of the inequality (4.1) is complete.

(b) Similarly, in order to prove the inequality (4.2) it suffices to show that for 0r10\leq r\leq 1 λ=cosϕ[1,1]\lambda=\cos\phi\in[-1,1]

r2γ2rγλ+1γ2(r22rλ+1).r^{2\gamma}-2r^{\gamma}\lambda+1\leq\gamma^{2}(r^{2}-2r\lambda+1).

Comparing the right-hand sides of the identities t22tλ+1=(1t)2+2(1λ)tt^{2}-2t\lambda+1=(1-t)^{2}+2(1-\lambda)t with t=rγt=r^{\gamma} and t=rt=r and using the Bernoulli inequality 1rγγ(1r)1-r^{\gamma}\leq\gamma(1-r), we get the required result. ∎

4.2 The improving operator

Denote α=p/p=p1\alpha=p^{\prime}/p=p^{\prime}-1, β=r/r=r1\beta=r^{\prime}/r=r^{\prime}-1, h~(x)=h(x)\tilde{h}(x)=h(-x). Transposition of the convolution operator amounts to changing the original kernel into the kernel with tilde, i.e.

(kf,g)=k(xy)f(y)g(x)𝑑y𝑑x=(f,k~g).\left(k*f,g\right)=\int\int k(x-y)f(y)g(x)\,dy\,dx=\left(f,\tilde{k}*g\right).

Clearly,

Kp,r=supfp=1,gr=1|(kf,g)|=supfp=1,gr=1|(f,kg)|=Kr,p.\left\|K\right\|_{p,r}=\sup_{\|f\|_{p}=1,\;\|g\|_{r}=1}\left|\left(k*f,g\right)\right|=\sup_{\|f\|_{p}=1,\;\|g\|_{r}=1}\left|\left(f,k*g\right)\right|=\left\|K\right\|_{r,p}.

Let SpS_{p} be the operator of radial projection onto the unit sphere in LpL_{p},

Spf=ffp.S_{p}f=\frac{f}{\left\|f\right\|_{p}}.

Hereinafter we assume that the function acted upon by the operator SpS_{p} is nonzero.

Suppose kLqk\in L_{q}, fLpf\in L_{p}, so that kfLrk*f\in L_{r^{\prime}} and hence (kf)βLr{(k*f)}^{\langle\beta\rangle}\in L_{r}. Introduce the operator Brp:LpLrB^{p}_{r}:L_{p}\to L_{r} by the formula

Brpf=Sr((kf~)β).B^{p}_{r}f=S_{r}\left({(\widetilde{k*f})}^{\langle\beta\rangle}\right).

(Its domain is the set {fLp|kf0}\{f\in L_{p}\,|\,k*f\neq 0\}.) Interchanging the exponents pp rr we have the operator Bpr:LrLpB^{r}_{p}:L_{r}\to L_{p}. Explicitly,

Bprg=Sp((kg~)α).B^{r}_{p}g=S_{p}\left({(\widetilde{k*g})}^{\langle\alpha\rangle}\right).

The improving operator B:LpLpB:\,L_{p}\to L_{p} is the composition

Bf=BprBrpf.Bf=B_{p}^{r}B_{r}^{p}f.
Remark 4.2.

In the ¡¡symmetric¿¿ case r=pr=p the operator B~:fBppf~\tilde{B}:\,f\mapsto B^{p}_{p}\tilde{f}, whose square is BB, is already a self-map of LpL_{p}. As such, it can be used for the purposes of the proof instead of the operator BB. With this approach, the case γ1\gamma\leq 1 in Lemma 4.6 is not needed; also the proof of Lemma 4.11 becomes one-step. One property of the operator BB that B~\tilde{B} lacks is the analog of the necessary condition of extremum Bf=fBf=f (see Proposition 5.4 in Section 5.4). One can instead propose that a maximizer in the case p=rp=r must satisfy the equation B~f=Taf\tilde{B}f=T_{a}f with TaT_{a} a shift. We do not know whether this condition is indeed necessary.

Lemma 4.3.

Let fLpf\in L_{p}, fp=1\left\|f\right\|_{p}=1, and kfr>0\left\|k*f\right\|_{r^{\prime}}>0. Then

kBrpfpkfr.\left\|k*B^{p}_{r}f\right\|_{p^{\prime}}\geq\|k*f\|_{r^{\prime}}. (4.3)
Proof.

Using the definition of the operator BrpB^{p}_{r}, we rewrite the inequality (4.3) to be proved in the form

k(kf~)βp(kf~)βrkfr=kfrr.\left\|k*{(\widetilde{k*f})}^{\langle\beta\rangle}\right\|_{p^{\prime}}\geq\left\|{(\widetilde{k*f})}^{\langle\beta\rangle}\right\|_{r}\,\left\|k*f\right\|_{r^{\prime}}=\left\|k*f\right\|_{r^{\prime}}^{r^{\prime}}.

(The identities hβr=(|h|βr)1/r=hrβ\left\|{h}^{\langle\beta\rangle}\right\|_{r}=\left(\int|h|^{\beta r}\right)^{1/r}=\left\|h\right\|_{r^{\prime}}^{\beta} and β+1=r\beta+1=r^{\prime} are used.)

Since f~p=fp=1\|\tilde{f}\|_{p}=\|f\|_{p}=1, the left-hand side is estimated as

k(kf~)βp|(k(kf~)β,f~)|=|(kf~,kf~β)|=|kf|β+1.\left\|k*{(\widetilde{k*f})}^{\langle\beta\rangle}\right\|_{p^{\prime}}\geq\left|\left(k*{(\widetilde{k*f})}^{\langle\beta\rangle},\tilde{f}\right)\right|=\left|\left(\widetilde{k*f},{\widetilde{k*f}}^{\langle\beta\rangle}\right)\right|=\int|k*f|^{\beta+1}.

The lemma is proved. ∎

Corollary 4.4.

If 0<ε<10<\varepsilon<1 and the function ff is an ε\varepsilon-maximizer for the convolution operator K:LpLrK:\,L_{p}\to L_{r^{\prime}}, then it belongs to the domain of BB, and BfBf is also an ε\varepsilon-maximizer for the operator KK.

Proof.

We have kfrKp,r(1ε)>0\left\|k*f\right\|_{r^{\prime}}\geq\|K\|_{p,r}(1-\varepsilon)>0, hence the function g=BrpfLrg=B^{p}_{r}f\in L_{r} is defined and gr=1\left\|g\right\|_{r}=1. According to (4.3),

kgpkfr>0.\left\|k*g\right\|_{p^{\prime}}\geq\left\|k*f\right\|_{r^{\prime}}>0.

Therefore the function h=Bprg=Bfh=B^{r}_{p}g=Bf, hp=1\left\|h\right\|_{p}=1 is defined and again, according to (4.3) with pp rr swapped,

khrkgp,\left\|k*h\right\|_{r^{\prime}}\geq\left\|k*g\right\|_{p^{\prime}},

whence K(Bf)rKfr\left\|K(Bf)\right\|_{r^{\prime}}\geq\left\|Kf\right\|_{r^{\prime}}, as required. ∎

Lemma 4.5.

The operator BB commutes with shifts: B(Taf)=Ta(Bf)B(T_{a}f)=T_{a}(Bf) for any ada\in\mathbb{R}^{d}. (If one side of the formula is defined, then the other is defined too.)

Proof.

We have Taf~=Taf~\widetilde{T_{a}f}=T_{-a}\tilde{f}, therefore, Ta(Brpf)=Brp(Taf)T_{-a}(B^{p}_{r}f)=B^{p}_{r}(T_{a}f), and similarly for BprB^{r}_{p}. The claimed equality follows. ∎

Lemma 4.6.

Let s>1s>1 and γs>1\gamma s>1. Then the map Q:ffγQ:\,f\mapsto{f}^{\langle\gamma\rangle} from LγsL_{\gamma s} to LsL_{s} is continuous.

Proof.

Consider two cases.

1. In the case γ1\gamma\leq 1 the continuity of QQ easily follows from the numerical inequality (4.1):

QfQgss=|fγgγ|sCs|fg|γs=Csfgγsγs.\left\|Qf-Qg\right\|_{s}^{s}=\int\left|{f}^{\langle\gamma\rangle}-{g}^{\langle\gamma\rangle}\right|^{s}\leq C^{s}\int\left|f-g\right|^{\gamma s}=C^{s}\left\|f-g\right\|_{\gamma s}^{\gamma s}.

2. In the case γ>1\gamma>1 we use the numerical inequality (4.2) and Hölder’s inequality and find

QfQgssCs|fg|s(|f|+|g|)(γ1)sCs(|fg|γs)1/γ((|f|+|g|)γ(γ1)s)1/γ.\left\|Qf-Qg\right\|_{s}^{s}\leq C^{s}\int\left|f-g\right|^{s}\,\left(|f|+|g|\right)^{(\gamma-1)s}\leq\\ \leq C^{s}\left(\int\left|f-g\right|^{\gamma s}\right)^{1/\gamma}\,\left(\int\left(|f|+|g|\right)^{\gamma^{\prime}(\gamma-1)s}\right)^{1/\gamma^{\prime}}.

Since γ(γ1)=γ\gamma^{\prime}(\gamma-1)=\gamma and (|f|+|g|)γs2γs1(|f|γs+|g|γs)(|f|+|g|)^{\gamma s}\leq 2^{\gamma s-1}(|f|^{\gamma s}+|g|^{\gamma s}) (by concavity), we get

QfQgss2γs1Csfgγss(fγsγs/γ+gγsγs/γ).\left\|Qf-Qg\right\|_{s}^{s}\leq 2^{\gamma s-1}C^{s}\,\left\|f-g\right\|_{\gamma s}^{s}\,\left(\left\|f\right\|_{\gamma s}^{\gamma s/\gamma^{\prime}}+\left\|g\right\|_{\gamma s}^{\gamma s/\gamma^{\prime}}\right).

This concludes the proof. ∎

Corollary 4.7.

The operators BrpB^{p}_{r}, BprB^{r}_{p} and BB are continuous on their domains with respect to the norm topologies in the preimage and image spaces.

Proof.

Each of these operators is a composition of continuous maps; Lemma 4.6 provides the continuity in the only place where it is not a commonly known fact. ∎

4.3 A compactness lemma

Lemma 4.8.

Let kLq(d)k\in L_{q}(\mathbb{R}^{d}) and χLrL(d)\chi\in L_{r^{\prime}}\cap L_{\infty}(\mathbb{R}^{d}). Then the integral operator with kernel χ(x)k(xy)\chi(x)k(x-y),

χK:f(x)χ(x)(kf)(x),\chi K:\;f(x)\mapsto\chi(x)(k*f)(x),

maps any weakly convergent sequence fnLpf_{n}\in L_{p} to a sequence convergent in LrL_{r^{\prime}} norm.

Proof.

Without loss of generality we may assume that fnp1\left\|f_{n}\right\|_{p}\leq 1 for all nn. Let fnff_{n}\rightharpoonup f in LpL_{p}; then fp1\left\|f\right\|_{p}\leq 1.

Consider first the case kLqLk\in L_{q}\cap L_{\infty}. Since q<p<q<p^{\prime}<\infty, we have kLpk\in L_{p^{\prime}}, hence the sequence kfnk*f_{n} converges pointwise. Besides, kfnkpfnpkp\left\|k*f_{n}\right\|_{\infty}\leq\left\|k\right\|_{p^{\prime}}\left\|f_{n}\right\|_{p}\leq\left\|k\right\|_{p^{\prime}}, therefore

|χ(x)(kfn)(x)|kp|χ(x)|.|\chi(x)\cdot(k*f_{n})(x)|\leq\left\|k\right\|_{p^{\prime}}|\chi(x)|.

The majorant in the right-hand side lies in LrL_{r^{\prime}}. By the Dominated Convergence Theorem we conclude that fnfr0\|f_{n}-f\|_{r^{\prime}}\to 0.

Now let us withdraw the assumption kLk\in L_{\infty}. Let KλK_{\lambda} be the operator of convolution with truncated function kλ(x)=k(x)I|k(x)|λ(x)LqLk_{\lambda}(x)=k(x)I_{|k(x)|\leq\lambda}(x)\in L_{q}\cap L_{\infty}. As follows from the previous, χKλ(fnf)r0\left\|\chi K_{\lambda}(f_{n}-f)\right\|_{r^{\prime}}\to 0. The proof is finished by use of the ε/3\varepsilon/3 trick. Given ε>0\varepsilon>0 we find λ\lambda such that χkkλq<ε/3\left\|\chi\right\|_{\infty}\left\|k-k_{\lambda}\right\|_{q}<\varepsilon/3. Let n0n_{0} be such that χKλ(fnf)r<ε/3\left\|\chi K_{\lambda}(f_{n}-f)\right\|_{r^{\prime}}<\varepsilon/3 when nn0n\geq n_{0}. Then for nn0n\geq n_{0} we have

χK(fnf)rχKλ(fnf)r+χ(kλk)fnr+χ(kλk)fr<ε.\left\|\chi K(f_{n}-f)\right\|_{r^{\prime}}\leq\left\|\chi K_{\lambda}(f_{n}-f)\right\|_{r^{\prime}}+\left\|\chi(k_{\lambda}-k)*f_{n}\right\|_{r^{\prime}}+\left\|\chi(k_{\lambda}-k)*f\right\|_{r^{\prime}}<\varepsilon.

(The 22nd and 33rd terms in the middle are estimated by Young’s inequality.) The proof is complete.(8){}^{(\mbox{\it\ref{com:compactness-lemma}})}

Corollary 4.9.

Let kLq(d)k\in L_{q}(\mathbb{R}^{d}) and Ωd\Omega\subset\mathbb{R}^{d} be a set of finite measure. If the sequence (fn)(f_{n}) is weakly convergent in Lp(d)L_{p}(\mathbb{R}^{d}), then the sequence of convolution restrictions (kfn)|Ω\left.(k*f_{n})\right|_{\Omega} strongly converges in Lr(Ω)L_{r^{\prime}}(\Omega).

4.4 Special maximizers and strong convergence on sets of finite measure

Lemma 4.10.

Let kLq(d)k\in L_{q}(\mathbb{R}^{d}) and a weakly convergent sequence fnff_{n}\rightharpoonup f in LpL_{p} be given. Put gn=Brpfng_{n}=B^{p}_{r}f_{n} and g=mβ(kf~)βg=m^{-\beta}{(\widetilde{k*f})}^{\langle\beta\rangle}. If kfnrm>0\left\|k*f_{n}\right\|_{r^{\prime}}\to m>0 for nn\to\infty, then for any set Ωd\Omega\subset\mathbb{R}^{d} of finite measure the sequence (gn)(g_{n}) restricted onto Ω\Omega converges in Lr(Ω)L_{r}(\Omega) norm to gg,

gngLr(Ω)0.\left\|g_{n}-g\right\|_{L_{r}(\Omega)}\to 0.

Also the weak convergence gngg_{n}\rightharpoonup g holds in Lr(d)L_{r}(\mathbb{R}^{d}).

Proof.

1. Put hn=kfnh_{n}=k*f_{n}, h=kfh=k*f, and mn=hnrm_{n}=\left\|h_{n}\right\|_{r^{\prime}}, Then gn=(hn~/mn)βg_{n}={(\tilde{h_{n}}/m_{n})}^{\langle\beta\rangle}. By Corollary 4.9, hnhh_{n}\to h in Lr(Ω)L_{r^{\prime}}(\Omega). Since 1/mn1/m1/m_{n}\to 1/m, it follows that hn/mnh/mh_{n}/m_{n}\to h/m in Lr(Ω)L_{r^{\prime}}(\Omega). The tilde operation commutes with passing to the limit. Applying Lemma 4.6 we get gnβgβ{g_{n}}^{\langle\beta\rangle}\to{g}^{\langle\beta\rangle} Lr(Ω)L_{r}(\Omega).

2. Let us now prove that (gng,ψ)0(g_{n}-g,\psi)\to 0 for any ψLr\psi\in L_{r^{\prime}}. Suppose ε>0\varepsilon>0 is given. Fix a set Ω\Omega of fonite measure and such that ψLr(dΩ)ε\left\|\psi\right\|_{L_{r^{\prime}}(\mathbb{R}^{d}\setminus\Omega)}\leq\varepsilon. By part 1, there exists n0n_{0} such that gngLr(Ω)ε\left\|g_{n}-g\right\|_{L_{r}(\Omega)}\leq\varepsilon for nn0n\geq n_{0}. Then for nn0n\geq n_{0} we have

|(gng,ψ)|εgngLr(dΩ)+εψLr(Ω)ε(1+gr+ψr).\left|(g_{n}-g,\psi)\right|\leq\varepsilon\left\|g_{n}-g\right\|_{L_{r}(\mathbb{R}^{d}\setminus\Omega)}+\varepsilon\left\|\psi\right\|_{L_{r^{\prime}}(\Omega)}\leq\varepsilon(1+\left\|g\right\|_{r}+\left\|\psi\right\|_{r^{\prime}}).

It is clear now that limn(gng,ψ)=0\lim_{n\to\infty}(g_{n}-g,\psi)=0.

The lemma is proved. ∎

Lemma 4.11.

Let kLq(d)k\in L_{q}(\mathbb{R}^{d}) (fn)(f_{n}) be a maximizing sequence of LpL_{p} functions for the convolution operator K:LpLrK:\,L_{p}\to L_{r^{\prime}}. Put hn=Bfnh_{n}=Bf_{n}; according to Corollary 4.4, (hn)(h_{n}) is also a maximizing sequence for the operator KK. If the sequence (fn)(f_{n}) converges weakly in LpL_{p}, then there exists a function hLph\in L_{p} such that

(i) hnhh_{n}\rightharpoonup h in LpL_{p};

(ii) for any set Ωd\Omega\subset\mathbb{R}^{d} of finite measure, hnhLp(Ω)0\left\|h_{n}-h\right\|_{L_{p}(\Omega)}\to 0 as nn\to\infty.

Proof.

Put gn=Brpfng_{n}=B^{p}_{r}f_{n}. Lemma 4.10 is applicable with m=Kp,rm=\left\|K\right\|_{p,r} and it yields weak convergence of (gn)(g_{n}) in LrL_{r}.

Due to the equality Kr,p=Kp,r\left\|K\right\|_{r,p}=\left\|K\right\|_{p,r} and Lemma 4.3, (gn)(g_{n}) is a maximizing sequence for the operator K:LrLpK:\,L_{r}\to L_{p^{\prime}}. Applying Lemma 4.10 again, with replacements BrpBprB^{p}_{r}\mapsto B^{r}_{p}, βα\beta\mapsto\alpha, fngnf_{n}\mapsto g_{n} and gnhng_{n}\mapsto h_{n}, we obtain the function h=w-limnhnLph=\mathop{\mathrm{w\mbox{-}lim}}\limits_{n\to\infty}h_{n}\in L_{p} that possesses all the required properties. ∎

4.5 Shifts, centering, and tightness

The lemmas of this subsection are but various technical expressions of the simple idea: if a mass is concentrated near the origin, then a long distance shift is incompatible with centering.

Let us first turn to the notions introduced in Definition 2.3 and prove boundedness of the set of shift vectors that provide δ\delta-near-centering of a given function for varying but small values of δ\delta.

Lemma 4.12.

Let 1p<1\leq p<\infty. Fix fLpf\in L_{p}, δ0<13fpp\delta_{0}<\frac{1}{3}\left\|f\right\|_{p}^{p} and a unit vector vdv\in\mathbb{R}^{d}. Put D=Dδ0,vp(f)D=D^{p}_{\delta_{0},v}(f). Let a0da_{0}\in\mathbb{R}^{d} be the vector for which the function Ta0fT_{a_{0}}f is δ0\delta_{0}-near-centered in the direction vv. If a function TafT_{a}f is δ\delta-near-centered in the direction vv for some δδ0\delta\leq\delta_{0} and ada\in\mathbb{R}^{d}, then |(aa0,v)|D|(a-a_{0},v)|\leq D.

Proof.

Assume, without loss of generality, that v=(1,0,,0)v=(1,0,\dots,0). Introducing the one-variable function

f1(x1)=d1|f(x)|p𝑑x2𝑑xn,f_{1}(x_{1})=\int_{\mathbb{R}^{d-1}}|f(x)|^{p}\,dx_{2}\,\dots\,dx_{n},

we reduce the general case to the case d=1d=1, p=1p=1, f0f\geq 0 (where ff now stands for f1f_{1} from the line above.)

Now a0a_{0} and aa are scalars. Let f1=m\left\|f\right\|_{1}=m. Due to the assumed centerings we have, first,

a0Da0fmδ02,a0a0+Dfmδ02.\int_{a_{0}-D}^{a_{0}}f\geq\frac{m-\delta_{0}}{2},\qquad\int_{a_{0}}^{a_{0}+D}f\geq\frac{m-\delta_{0}}{2}.

Next,

afmδ2,afmδ2.\int_{-\infty}^{a}f\geq\frac{m-\delta}{2},\qquad\int_{a}^{\infty}f\geq\frac{m-\delta}{2}.

Suppose that a>a0+Da>a_{0}+D. Then

m3<mδ2afa0+Dfma0Da0+Dfδ0<m3,\frac{m}{3}<\frac{m-\delta}{2}\leq\int_{a}^{\infty}f\leq\int_{a_{0}+D}^{\infty}f\leq m-\int_{a_{0}-D}^{a_{0}+D}f\leq\delta_{0}<\frac{m}{3},

a contradiciton. Likewise, the assumption a<a0Da<a_{0}-D leads to a contradiciton. We conclude that |aa0|D|a-a_{0}|\leq D. The lemma is proved. ∎

Further lemmas of this subsection pertain to the notions introduced in Definition 2.4.

Lemma 4.13.

Suppose the sequence of vectors anda_{n}\in\mathbb{R}^{d} is bounded. If the sequences (fn)(f_{n}) and (f^n)(\hat{f}_{n}) in LpL_{p} are related by shifts, f^n=Tanfn\hat{f}_{n}=T_{a_{n}}f_{n}, and one of them is tight, then the other one is tight, too.

Proof.

Let anR\left\|a_{n}\right\|\leq R for all nn. For any coordinate cube QQ, the shifted cube TanQT_{a_{n}}Q is contained in the nn-independent cube QRQ_{R} concentric with QQ and with side length which exceeds that of QQ by 2R2R. Therefore for any δ>0\delta>0 the sequence (fn)(f_{n}) is δ\delta-near-finite if and only if such the same is true about the sequence (f^n)(\hat{f}_{n}). The lemma is proved. ∎

Lemma 4.14.

Suppose the sequence (fn)(f_{n}) LpL_{p} (1p<1\leq p<\infty) is relatively tight and fnp=1\left\|f_{n}\right\|_{p}=1 for all nn. Suppose further that all the functions fnf_{n} are δ0\delta_{0}-near-centered (of order pp) with some δ0<1/3\delta_{0}<1/3. Then the sequence (fn)(f_{n}) is tight.

Proof.

Without loss of generality we may assume that for δ=δ0\delta=\delta_{0} and all vectors eje_{j} of the fixed orthonormal basis in RdR^{d} the condition in the last part of Definition 2.4 holds with n0=1n_{0}=1. Thus there is D0>0D_{0}>0 such that for any n1n\geq 1,

Dδ0,ejp(fn)<D0.D^{p}_{\delta_{0},e_{j}}(f_{n})<D_{0}.

It suffices to verify the condition of δ\delta-near-finiteness for any given δ>0\delta>0.

Fix δ\delta; we may assume that δ<1/3\delta<1/3. Let us select the vectors ana_{n} so as to obtain δ\delta-near-centered functions f^n=Tanfn\hat{f}_{n}=T_{a_{n}}f_{n}. By Lemma 4.12, for all n1n\geq 1 and j=1,,dj=1,\dots,d we have |(an,ej)|D0|(a_{n},e_{j})|\leq D_{0}.

By definition of a relatively tight sequence, there exist n0n_{0} and DD such that D(δ/d),ejp(fn)DD^{p}_{(\delta/d),e_{j}}(f_{n})\leq D for all nn0n\geq n_{0} and j=1,,dj=1,\dots,d. Then

|(x,ej)|>D+D0|fn|p|(x,ej)|>D|f^n|pδd\int_{|(x,e_{j})|>D+D_{0}}|f_{n}|^{p}\leq\int_{|(x,e_{j})|>D}|\hat{f}_{n}|^{p}\leq\frac{\delta}{d}

for nn0n\geq n_{0} and j=1,,dj=1,\dots,d.

Put R=D+D0R=D+D_{0}. The complement of the cube Q=[R,R]dQ=[-R,R]^{d} is the union of the sets {x||(x,ej)|>R}\{x\,|\,|(x,e_{j})|>R\}, j=1,,dj=1,\dots,d. Therefore, Q|fn|pfnppd(δ/d)=1δ\int_{Q}|f_{n}|^{p}\geq\left\|f_{n}\right\|_{p}^{p}-d\cdot(\delta/d)=1-\delta.

The condition of δ\delta-near-finiteness is affirmed, and the Lemma is proved. ∎

The next lemma, though not used in the proof of Theorem 1, is included as it further clarifies the connection between the notions of relative tightness and tightness.

Lemma 4.15.

Suppose (fn)(f_{n}) is a relative tight sequence in LpL_{p} (1p<1\leq p<\infty) and fnp=1\left\|f_{n}\right\|_{p}=1 for all nn. If the sequence (fn)(f_{n}) is δ0\delta_{0}-near-finite for some δ0<1/3\delta_{0}<1/3, then it is tight.

Proof.

Consider the δ0\delta_{0}-near-centered sequence (f^n)(\hat{f}_{n}), obtained from (fn)(f_{n}) by means of suitable shifts, f^n=Tanfn\hat{f}_{n}=T_{a_{n}}f_{n}. Let us show that the sequence of vectors (an)(a_{n}) is bounded in d\mathbb{R}^{d}.

We may assume that the coordinate cube QQ in the definition of δ0\delta_{0}-near-finiteness has the origin as its center and is described by the inequalities |(x,ej)|R|(x,e_{j})|\leq R, j=1,,dj=1,\dots,d. If (an,ej)>R(a_{n},e_{j})>R, then

(x,ej)>0|f^n|p(x,ej)>R|f|pδ0,\int_{(x,e_{j})>0}|\hat{f}_{n}|^{p}\leq\int_{(x,e_{j})>R}|f|^{p}\leq\delta_{0},

which contradicts the function f^n\hat{f}_{n} being δ0\delta_{0}-near-centered. Therefore, (an,ej)R(a_{n},e_{j})\leq R. Similarly (an,ej)R(a_{n},e_{j})\geq-R. Thus, supnanRd\sup_{n}\|a_{n}\|\leq R\sqrt{d}.

Applying Lemma 4.14 to the sequence (f^n)(\hat{f}_{n}) and then applying Lemma 4.13 to the pair of sequences (fn)(f_{n}), (f^n)(\hat{f}_{n}), we come to the conclusion as stated. ∎

4.6 The final lemma

Lemma 4.16.

Suppose that the sequence of functions (fn)(f_{n}) in Lp(d)L_{p}(\mathbb{R}^{d}) possesses the following properties:

  • (i)

    normalization: fnp=1\left\|f_{n}\right\|_{p}=1 nn;

  • (ii)

    tightness (see Definition 2.4);

  • (iii)

    local convergence: there exists a function fLpf\in L_{p} to which fnf_{n} converges on bounded sets: IΩ(fnf)p0\left\|I_{\Omega}(f_{n}-f)\right\|_{p}\to 0 for any bounded set Ωd\Omega\subset\mathbb{R}^{d}.

Then fnff_{n}\to f in LpL_{p}. In particular, fp=1\left\|f\right\|_{p}=1.

Proof.

Let ε>0\varepsilon>0 be given. Take a bounded set UU such that IdUfp<ε/3\left\|I_{\mathbb{R}^{d}\setminus U}f\right\|_{p}<\varepsilon/3.

Due to the assumptions (i) and (ii), there are n1n_{1} and a cube QQ in d\mathbb{R}^{d} such that IdQfnp<ε/3\left\|I_{\mathbb{R}^{d}\setminus Q}f_{n}\right\|_{p}<\varepsilon/3 for nn1n\geq n_{1}.

Put Ω=UQ\Omega=U\cup Q. Due to the assumption (iii), there is n2n_{2} such that IΩ(fnf)p<ε/3\left\|I_{\Omega}\cdot(f_{n}-f)\right\|_{p}<\varepsilon/3 for nn2n\geq n_{2}. Clearly, for nmax(n1,n2)n\geq\max(n_{1},n_{2}) we have the inequalities

fnfpIΩ(fnf)p+IdΩfnp+IdΩfp<ε/3+ε/3+ε/3=ε.\left\|f_{n}-f\right\|_{p}\leq\left\|I_{\Omega}(f_{n}-f)\right\|_{p}+\left\|I_{\mathbb{R}^{d}\setminus\Omega}f_{n}\right\|_{p}+\left\|I_{\mathbb{R}^{d}\setminus\Omega}f\right\|_{p}<\varepsilon/3+\varepsilon/3+\varepsilon/3=\varepsilon.

The lemma is proved. ∎

5 Supplementary results

5.0 A survey

In this section we put together diverse, relatively simple results related to various aspects and details of formulation and proof of Theorem 1. Some other related, but unsolved questions are considered in Section 7.

Subsection 5.1. Limit cases. Theorem 1 excludes the cases where at least one of the exponents pp, qq, rr in Young’s inequality equals 11 or \infty. We analyse all such cases. A summary of the results is presented on Fig. 2.

Subsection 5.2. Convolution on compact groups. The analog of Theorem 1 for compact groups is an easy result. The groups need not be commutative.

Subsection 5.3. Counterexample: a near-convolution without a maximizer. We give a counterexample to demonstrate that the assumptions in Theorem 1 cannot be relaxed by allowing integral operators KK with non-translation-invariant kernels, even under the assumption that the kernel is pointwise dominated by the kernel k(xy)k(x-y) of an admissible convolution operator. Another possibility to relax the assumptions is to consider compact or even finite-dimensional perturbations of a convolution operator. In that case, we were unable to prove or disprove the existence of a maximizer; see Question 3 in Section 7.

Subsection 5.4. Necessary condition of extremum. First, using the standard Lagrange multipliers method, we derive a nonlinear integral equation that must be satisfied by a maximizer. Then we prove an “approximative” version of the necessary condition of extremum: if the norm of the convolution kfk*f is close to K\|K\| and f=1\|f\|=1, then ff satisfies the mentioned equation up to a small error.(8){}^{(\mbox{\it\ref{com:discrepancy}})}

Subsection 5.5. Convergence to a maximizer in the class Max\mathrm{Max} ((rather than in SMax)\mathrm{SMax}). In the course of the proof of Theorem 1 we have established that any special (that is, lying in the image of the improving operator BB) maximizing sequence becomes relatively compact after applying appropriate shifts. Here we show that the same is true for arbitrary maximizing sequences.

This simple result is perhaps of minor significance, but we included it due to an authoritative motivation(8){}^{(\mbox{\it\ref{com:Lions-convergence}})}.

Subsection 5.6. Kernel approximation and convergence of maximizers. Given a sequence of convolution kernels knk_{n} converging in LqL_{q} to a kernel kk, is it true that a maximizer for the operator KkK_{k} can be obtained as a limit (in LpL_{p}) of maximizers for the operators KknK_{k_{n}}? Proposition 5.7 answers this question in the affirmative. The result can be of use, for example, when one has to compute a maximizer for convolution with non-compact and, possibly, weakly singular kernel: the kernel can be approximated by bounded and finitely supported truncations.

Subsection 5.7. On boundedness and integrability of maximizers

If one has “a spare room of integrability”, kLqεLq+εk\in L_{q-\varepsilon}\cap L_{q+\varepsilon} (as in Pearson’s theorem), then a maximizer belongs to Lp#LL_{p_{\#}}\cap L_{\infty}, where p#<pp_{\#}<p does not depend on ε\varepsilon. See also Question 5 in Section 7.

Subsection 5.8. On the lower bound of convolution operators’ norms. The estimate in Lemma 3.7 becomes less efficient as the norm of the operator KK decreases. (Cf. Remark 3.8). This fact has no adverse consequences for the proof of Theorem 1, but one should keep it in mind if the results of Section 3 are to be used for obtaining uniform estimates (over some family of kernels kk). In particular, suppose that the absolute value |k(x)||k(x)| of the kernel fixed; then how small can the norm Kkp,r\|K_{k}\|_{p,r} be? Proposition 5.10 states that infKkp,r=0\inf\|K_{k}\|_{p,r}=0.(8){}^{(\mbox{\it\ref{com:convolution-lower-bound}})}

5.1 Limit cases

For 1p,q1\leq p,q\leq\infty the relation (1.1) defines the exponent r[1,]r\in[1,\infty] if and only if 1/p+1/q11/p+1/q\geq 1, equivalently, if qpq\leq p^{\prime}. If q=pq=p^{\prime}, then r=r^{\prime}=\infty. Consider the coordinate (u,v)(u,v)-plane with u=1/qu=1/q, v=1/pv=1/p^{\prime}. The domain corresponding to admissible pairs (q,p)(q,p) in Young’s inequality is the triangle formed by the lines (I) u=1u=1 (i.e. q=1q=1), (II) v=0v=0 (i.e. p=1p=1) and (III) u=vu=v (i.e. r=r^{\prime}=\infty). Correspondingly we have three limit cases and subcases corresponding to the vertices of the triangle. The results are summarized on Fig. 2.

In the conditions considered below we always assume that pointwise equalities and inequalities are fulfilled a.e.

Refer to caption1/q1/q1/p1/p^{\prime}I(A)I(B)I(C)II(A)II(B)III11ExistsDoes not exist Both possibilitiesNonexistence is possibleExistence — ?


Figure 2: The cases of existence/nonexistence of a maximizer

Case I. q=1q=1.

Subcase I(A). q=1q=1, p=r=1p=r^{\prime}=1.

I(A)1\mathrm{I(A)_{1}}. If k0k\geq 0, then any function f0f\geq 0 with f=1\int f=1 is a maximizer. Obviuosly, the same holds true for functions of the form k=ck+k=ck_{+}, where k+0k_{+}\geq 0, c=constc=\mathrm{const}.

I(A)2\mathrm{I(A)_{2}}. If kk is not a function with constant complex argument (in the real case — a sign-changing function), then a maximizer does not exist. Indeed, one can choose a maximizing sequence to be a δ\delta-sequence, so Kk1,=k1\|K_{k}\|_{1,\infty}=\|k\|_{1}; but the equality in the integral Minkowski inequality

|k(y)f(xy)𝑑y|𝑑x|k(y)|𝑑y|f(x)|𝑑x\int\left|\int k(y)f(x-y)\,dy\right|\,dx\leq\int|k(y)|\,dy\;\int|f(x)|\,dx

is impossible (if f0f\neq 0).

Subcase I(B). q=1q=1, p=r(1,)p=r^{\prime}\in(1,\infty).

I(B)1\mathrm{I(B)_{1}}. Let us show that if k0k\geq 0, then there is no maximizer.

In this case Kkp,p=k1\|K_{k}\|_{p,p^{\prime}}=\|k\|_{1}. Indeed, the sequence of pairs {fn,gn}\{f_{n},g_{n}\} with

fn(x)=n1/pI[0,n](x),gn(x)=n1/pI[0,n](x)f_{n}(x)=n^{-1/p}I_{[0,n]}(x),\qquad g_{n}(x)=n^{-1/p^{\prime}}I_{[0,n]}(x)

is a maximizing sequence for the bilinear form (kf,g)(k*f,g). Then fnp=gnp=1\|f_{n}\|_{p}=\|g_{n}\|_{p^{\prime}}=1 and

(kfn,gn)=1n0nxnxk(t)𝑑t𝑑x=nnk(t)(1|t|n)𝑑tk(t)𝑑t.(k*f_{n},g_{n})=\frac{1}{n}\int_{0}^{n}\int_{x-n}^{x}k(t)\,dt\,dx=\int_{-n}^{n}k(t)\left(1-\frac{|t|}{n}\right)\,dt\to\int k(t)\,dt.

A hypothetical maximizer ff would satisfy the equality

k(y)Tyf()𝑑yp=k1fp,\left\|\int k(y)T_{y}f(\cdot)\,dy\right\|_{p}=\|k\|_{1}\,\|f\|_{p},

which is the case of equality in the Minkowski integral inequality. This, in turn, would imply the existence of a function λ(y)0\lambda(y)\geq 0 such that Tyf(x)=λ(y)f(x)T_{y}f(x)=\lambda(y)f(x) for almost all xx, yy. But this is clearly impossible unless f=0f=0.

I(B)2\mathrm{I(B)_{2}}. Let us show that generally (for not-constant-sign functions) a maximizer can exist. Let p=r=2p=r^{\prime}=2. The operator KkK_{k} acts in the Hilbert space L2L_{2} and is unitary equivalent to multiplication by the continuous function w=kw=\mathcal{F}k, where k\mathcal{F}k is the Fourier transform with unitary normalization. Let m=wm=\|w\|_{\infty}. A maximizer exists if and only if the set {ξ:|w(ξ)|=m}\{\xi:\,|w(\xi)|=m\} has positive measure. This is possible. For example, let ww be a “hat” function: wC0w\in C_{0}^{\infty}, 0w(ξ)10\leq w(\xi)\leq 1 everywhere, and w(ξ)=1w(\xi)=1 in some neighborhood of zero, UU. Then k=1wL1k=\mathcal{F}^{-1}w\in L_{1} and k1IU2=wIU2=IU2Kk2,2\left\|k*\mathcal{F}^{-1}I_{U}\right\|_{2}=\left\|w\cdot I_{U}\right\|_{2}=\left\|I_{U}\right\|_{2}\left\|K_{k}\right\|_{2,2}.

The question as to whether a maximizer can exist in the case p2p\neq 2 is left open. (See Question 2 in Section 7).

Subcase I(C). q=1q=1, p=r=p=r^{\prime}=\infty. A maximizer exists: for instance, f(x)=k(x)¯/|k(x)|f(x)=\overline{k(-x)}/{|k(-x)|}. Indeed, f=1\|f\|_{\infty}=1 and

Kk,1k1=(kf)(0)kf,\|K_{k}\|_{\infty,1}\leq\|k\|_{1}=(k*f)(0)\leq\|k*f\|_{\infty},

whence kf=Kk,1\|k*f\|_{\infty}=\|K_{k}\|_{\infty,1}.

Case II. p=1p=1. The operator KkK_{k} acts from L1L_{1} to LqL_{q}. We assume that q>1q>1, since the subcase q=1q=1 is explored earlier, in I(A).

Subcase II(A). p=1p=1, 1<q=r<1<q=r^{\prime}<\infty. A maximizer does not exist. Indeed, a δ\delta-sequence is a maximizing sequence: kfnkk*f_{n}\to k in LqL_{q}, so that Kk1,q=kq\|K_{k}\|_{1,q^{\prime}}=\|k\|_{q}. The situation is similar to the one we have encountered in I(B)1\mathrm{I(B)_{1}}, with functions kk and ff interchanged. A hypothetical maximizer 0fL10\neq f\in L_{1} would realize the case of equality in Minkowski’s inequality

f(y)Tyk()𝑑yq=f1kq,\left\|\int f(y)T_{y}k(\cdot)\,dy\right\|_{q}=\|f\|_{1}\,\|k\|_{q},

but this is impossible.

Subcase II(B). p=1p=1, q=r=q=r^{\prime}=\infty. Put m=km=\|k\|_{\infty}. One readily sees that Kk1,1=m\|K_{k}\|_{1,1}=m. If k1(m)k^{-1}(m) is a set of positive measure, then a maximizer trivially exists. If k1(m)k^{-1}(m) is a set of measure zero, then both existence and non-existence of a maximizer are possible. We give a partial criterion of existence in the case of a nonnegative kernel kk.

Proposition 5.1.

Suppose that kLk\in L_{\infty}, k0k\geq 0, and |k1(m)|=0|k^{-1}(m)|=0. Put Ua={x|k(x)ma}U_{a}=\{x\,|\,k(x)\geq m-a\}. In order for a maximizer of the convolution operator Kk:L1LK_{k}:\,L_{1}\to L_{\infty} to exist it is necessary that |Ua|=|U_{a}|=\infty for all a>0a>0, and sufficient that there are vectors vnv_{n} such that the union

U^=n=1Tvn(U1/n)\hat{U}=\bigcap_{n=1}^{\infty}T_{v_{n}}\left(U_{1/n}\right)

has positive measure.

Proof.

1) Necessity. Suppose infa|Ua|<\inf_{a}|U_{a}|<\infty. Then lima0|Ua|=|k1(m)|=0\lim_{a\to 0}|U_{a}|=|k^{-1}(m)|=0. Let f0f\geq 0 and f=1\int f=1. We will prove that ff is not a maximizer. Take ε(0,1)\varepsilon\in(0,1). Due to absolute continuity of the Lebesgue integral (see e.g. [7, v. 1, Theorem 2.5.7]) there exists δ>0\delta>0 such that Ω|f|<ε\int_{\Omega}|f|<\varepsilon for any set Ω\Omega of measure |Ω|<δ|\Omega|<\delta. Let aa be such that |Ua|<δ|U_{a}|<\delta. Then for any xx we have A(x)=Uaf(xy)𝑑y<εA(x)=\int_{U_{a}}f(x-y)\,dy<\varepsilon. Hence

kf(x)=Uak(y)f(xy)𝑑y+dUak(y)f(xy)𝑑ymA(x)+(ma)(1A(x))ma+aε.\begin{array}[]{rcl}k*f(x)&=&\displaystyle\int_{U_{a}}k(y)f(x-y)\,dy+\int_{\mathbb{R}^{d}\setminus U_{a}}k(y)f(x-y)\,dy\;\leq\\[12.91663pt] &\leq&\displaystyle mA(x)+(m-a)(1-A(x))\leq m-a+a\varepsilon.\end{array}

Therefore kfma(1ε)<m\|k*f\|_{\infty}\leq m-a(1-\varepsilon)<m, as claimed.

2) Sufficiency. Let ΩU^\Omega\subset\hat{U} be a set of finite positive measure. We will show that f=|Ω|1IΩ~f=|\Omega|^{-1}\widetilde{I_{\Omega}} is a maximizer. Indeed, we have

kf(vn)=|Ω|1k(y)IΩ(y+vn)𝑑y=|Ω|1Tvn(Ω)k(y).k*f(-v_{n})=|\Omega|^{-1}\int k(y)I_{\Omega}(y+v_{n})\,dy=|\Omega|^{-1}\int_{T_{-v_{n}}(\Omega)}k(y).

But Tvn(Ω)Tvn(U^)U1/nT_{-v_{n}}(\Omega)\subset T_{-v_{n}}(\hat{U})\subset U_{1/n}. Hence k(y)m1/nk(y)\geq m-1/n whenever yTvn(Ω)y\in T_{-v_{n}}(\Omega) and kf(vn)m1/nk*f(-v_{n})\geq m-1/n. Due to L1L_{1}-continuity of the shift operator, in some neighbourhood of the point vn-v_{n} we have kf(x)m2/nk*f(x)\geq m-2/n. We conclude that kfm2/n\|k*f\|_{\infty}\geq m-2/n. Since nn is arbitrary, kf=m\|k*f\|_{\infty}=m.

The Proposition is proved. ∎

For example (in the one-dimensional case), for the kernels k(x)=e|x|k(x)=e^{-|x|} or k(x)=|sinx|k(x)=|\sin x| there is no maximizer, while for the kernel k(x)=1+tanhxk(x)=1+\tanh x a maximizer exists.

Case III. r=r^{\prime}=\infty, q=pq=p^{\prime}. It suffices to assume that 1<q<1<q<\infty, since the subcases q=1q=1 and q=q=\infty have been already covered — see. I(C) and II(B).

The present case is simple; a maximizer does exist. Given kLqk\in L_{q}, put (using notation introduced in Section 4.2) f(x)=k~(x)q/qf(x)={\tilde{k}(x)}^{\langle q/q^{\prime}\rangle}. Then the case of equality in Hölder’s inequality is realized:

kf(0)=(k,f~)=|k(x)|1+q/q=kqq=kqfq,k*f(0)=(k,\tilde{f})=\int|k(x)|^{1+q/q^{\prime}}=\|k\|_{q}^{q}=\|k\|_{q}\|f\|_{q^{\prime}},

and hence, too, the case of equality in Young’s inequality:

kf=kqfq,\|k*f\|_{\infty}=\|k\|_{q}\|f\|_{q^{\prime}},

so that ff is a maximizer.

It is instructive to compare this case with II(A), since the two cases deal with operators which are the transposes of each other. The relevant bilinear form in both cases is formally the same, however the conclusions about the existence of a maximizer are opposite.

Let kLqk\in L_{q}. We fix the symbol KK to mean the operator of convolution with kernel kk acting from LqL_{q^{\prime}} to LL_{\infty}. The transposed operator, acting from L1L_{1} to LqL_{q}, as in II(A), will be denoted KK^{\prime}. We have

K=K=supf,g:fq=g1=1(Kf,g)=supf,g:fq=g1=1(f,Kg).\|K\|=\|K^{\prime}\|=\sup_{f,g\,:\|f\|_{q^{\prime}}=\|g\|_{1}=1}(Kf,g)=\sup_{f,g\,:\|f\|_{q^{\prime}}=\|g\|_{1}=1}(f,K^{\prime}g).

Our conclusions on (non-)existence of a maximizer can be expressed by means of the formula

K=K=supg1=1maxfq=1(Kf,g),\|K\|=\|K^{\prime}\|=\sup_{\|g\|_{1}=1}\max_{\|f\|_{q^{\prime}}=1}(Kf,g),

where sup\sup cannot be replaced by max\max. The underlying cause of the difference is of course the non-reflexivity of LL_{\infty}. If we allow gLg\in L_{\infty}^{*}, then sup\sup becomes attainable. More precisely, by the Hahn-Banach theorem there exists γL\gamma\in L_{\infty}^{*}, γL=1\|\gamma\|_{L_{\infty}^{*}}=1 such that

(Kf0,γ)=Kf0=K,(Kf_{0},\gamma)=\|Kf_{0}\|_{\infty}=\|K\|,

where f0Lqf_{0}\in L_{q^{\prime}} is a maximizer for the operator KK. In order to describe the matters explicitly, let us note that the image of the operator KK lies in the closed subspace C0LCC_{0}\subset L_{\infty}\cap C of continuous functions vanishing at infinity. The space C0C_{0}^{*} is the space of finite Borel measures. The element γC0\gamma\in C_{0}^{*} realizing the equality

(Kf0,γ)=Kf0=K(Kf_{0},\gamma)=\|Kf_{0}\|_{\infty}=\|K\|

is the measure δx0\delta_{x_{0}}, where x0x_{0} is a point of maximum of Kf0(x)Kf_{0}(x).

5.2 Convolution on compact groups

Proposition 5.2.

Let GG be a compact topological group with Haar measure dμd\mu, the spaces Lp(G)L_{p}(G) defined with respect to this measure. Let kLq(G)k\in L_{q}(G). Then the convolution operator Kk:f(x)Gk(xy1)f(y)𝑑μ(y)K_{k}:\,f(x)\mapsto\int_{G}k(xy^{-1})f(y)\,d\mu(y) acts boundedly from LpL_{p} to LrL_{r^{\prime}}, where, as everywhere in this paper, 1/p+1/q+1/r=21/p+1/q+1/r=2, and there exists a maximizer FLp(G)F\in L_{p}(G):

Fp=1,KkFr=Kp,r.\|F\|_{p}=1,\qquad\|K_{k}F\|_{r^{\prime}}=\|K\|_{p,r}.
Proof.

Boundedness of the operator KkK_{k} (Young’s inequality) is a well-known fact. (Sufficient assumption is that the group GG is locally compact and unimodular, see e.g. [18, (20.18), (20.19)].) Now, take any maximizing sequence (fn)(f_{n}) and select a weakly convergent subsequence. The improving operator BB maps it to a strongly convergent one; a proof of the required analog of Lemma 4.8 is even easier here: we do not need a “truncation in the horizontal direction” to obtain a compaclty supported function. The limit is a maximizer. ∎

5.3 Counterexample: a near-convolution without a maximizer

It is natural to ask about possible relaxation of conditions of Theorem 1 and to try to exhibit sufficient conditions that the kernel K(x,y)K(x,y) of the integral operator K~:f(x)K(x,y)f(y)𝑑y\tilde{K}:\;f(x)\mapsto\int K(x,y)f(y)\,dy should satisfy, not necessarily being translation-invariant, in order to guarantee the existence of a maximizer. As the example below demonstrates, conditions of such a sort, if possible at all, cannot be stated in terms of integral and pointwise inequalities only: here, there is no maximizer, although we have a pointwise majorization 0<K(x,y)<k(xy)0<K(x,y)<k(x-y) with kLqk\in L_{q}.

Proposition 5.3.

Let p,q,rp,q,r be as in Theorem 1 and d=1d=1. Let kLq(1)k\in L_{q}(\mathbb{R}^{1}) and k(x)>0k(x)>0 everywhere. Consider the integral operator K~\tilde{K} with kernel K~(x,y)=k(xy)u(y)\tilde{K}(x,y)=k(x-y)u(y), where u(y)u(y) is monotone increasing and limyu(y)=0\lim_{y\to-\infty}u(y)=0, limy+u(y)=1\lim_{y\to+\infty}u(y)=1. The operator K~:LpLr\tilde{K}:\,L_{p}\to L_{r^{\prime}} is continuous. A maximizer for the operator K~\tilde{K} does not exist.

Proof.

One readily sees that K~p,r=Kkp,r\|\tilde{K}\|_{p,r}=\|K_{k}\|_{p,r}. If fLpf\in L_{p} (fp=1\|f\|_{p}=1) is a maximizer for the operator K~\tilde{K}, then the function f(x)u(x)f(x)u(x) must be a maximizer for the convolution operator KkK_{k}. But it is clear that fup<1\|fu\|_{p}<1, a contradiction. ∎

5.4 Necessary condition of extremum

The notation from Section 4.2 will be used. The next Proposition does not refer to the existence of maximizer result, so we allow the case q=1q=1.

Proposition 5.4.

Suppose that 1q<1\leq q<\infty and 1<p,r<1<p,r<\infty. (The relation (1.1) is assumed as always.) A maximizer of the convolution operator KK, if it exists, satisfies the equation f=Bff=Bf.

Proof.

We have the optimization problem (with given function kk and unknown ff and gg):

Rek(xy)f(x)g(y)𝑑x𝑑ymax\operatorname{Re}\iint k(x-y)f(x)g(y)\,dx\,dy\;\to\;\max

under the constraints

|f(x)|p𝑑x=1,|g(y)|r𝑑y=1.\int|f(x)|^{p}\,dx=1,\qquad\int|g(y)|^{r}\,dy=1.

Let us use the Lagrange multipliers method to derive the system of equations to be satisfied by the extremal pair of functions (f,g)(f,g). The relevant Lagrange functional can be taken in the form

(f,g)=Rek~(x+y)f(x)g~(y)𝑑x𝑑yλ|f(x)|p𝑑xμ|g(y)|r𝑑y.\mathcal{L}(f,g)=\operatorname{Re}\iint\tilde{k}(x+y)f(x)\tilde{g}(y)\,dx\,dy-\lambda\int|f(x)|^{p}\,dx-\mu\int|g(y)|^{r}\,dy.

Computing the partial variation with respect to f=f1+if2f=f_{1}+if_{2}, we get

δ=(G1λp|f|p2f1)δf1(G2+λp|f|p2f2)δf2,\delta\mathcal{L}=\int\left(G_{1}-\lambda p|f|^{p-2}f_{1}\right)\cdot\delta f_{1}-\int\left(G_{2}+\lambda p|f|^{p-2}f_{2}\right)\cdot\delta f_{2},

where G=G1+iG2=k~gG=G_{1}+iG_{2}=\tilde{k}*g. Therefore the pair (f,g)(f,g) that yields an extremum of the functional \mathcal{L} must satisfy the equation

G¯λp|f|p2f=0.\bar{G}-\lambda p|f|^{p-2}f=0.

Similarly, equating the partial variation δ/δg\delta\mathcal{L}/\delta g to zero, we come to the equation

F¯μr|g|r2g~=0,\bar{F}-\mu r|g|^{r-2}\tilde{g}=0,

where F=k~f~=kf~F=\tilde{k}*\tilde{f}=\widetilde{k*f}.

Taking into account the normalization of ff and gg and the identities r1=r/rr-1=r/r^{\prime}, p1=p/pp-1=p/p^{\prime}, the obtained system of equations can be written as

g~=Brpf,f=Bprg~.\tilde{g}=B^{p}_{r}f,\qquad f=B^{r}_{p}\tilde{g}.

Elimination of g~\tilde{g} results in the equation Bf=fBf=f. ∎

In Subsection 5.5 we will need an approximative version of the necessary condition of extremum.

Proposition 5.5.

For any ε>0\varepsilon>0 there exists δ>0\delta>0 (depending on q,pq,p and the convolution kernel kk) such that if fp=1\|f\|_{p}=1 and kfr>Kp,r(1δ)\|k*f\|_{r^{\prime}}>\|K\|_{p,r}(1-\delta), then Bffp<ε\|Bf-f\|_{p}<\varepsilon.

Proof.

We make use of the approximative version of Hölder’s inequality due to H. Hanche-Olsen [15, Lemma 2]:(8){}^{(\mbox{\it\ref{com:rev-Holder}})}

For any ε>0\varepsilon>0 there exists η>0\eta>0 such that if Fp=Gp=1\|F\|_{p}=\|G\|_{p^{\prime}}=1 and Re(F,G)>1η\operatorname{Re}(F,G)>1-\eta, then FGαp<ε\|F-{G}^{\langle\alpha\rangle}\|_{p}<\varepsilon.

Consider an improvement of the estimate in the proof of Lemma 4.3. Using the notation of Section 4.2 (in particular, recall: SpS_{p^{\prime}} is the radial projection onto the unit sphere in LpL_{p^{\prime}}, and β=r/r\beta=r^{\prime}/r), put G=Sp(k(kf~)β)G=S_{p^{\prime}}\left(k*{(\widetilde{k*f})}^{\langle\beta\rangle}\right) F=f~F=\tilde{f}. Note that G=(Bf~)1/αG={(\widetilde{Bf})}^{\langle 1/\alpha\rangle}, hence FGαp=fBfp\|F-{G}^{\langle\alpha\rangle}\|_{p}=\|f-Bf\|_{p}.

Further, denote g=(kf~)βg={(\widetilde{k*f})}^{\langle\beta\rangle} and M=Kp,r=Kr,pM=\|K\|_{p,r}=\|K\|_{r,p}.

The calculation in the proof of Lemma 4.3 implies that kfrr=(kg,f~)=(G,F)kgp\|k*f\|_{r^{\prime}}^{r^{\prime}}=(k*g,\tilde{f})=(G,F)\|k*g\|_{p^{\prime}}. In particular, (F,G)>0(F,G)>0.

We have kgpMgr=Mkfrr/rMβ+1fpβ=Mr\|k*g\|_{p^{\prime}}\leq M\|g\|_{r}=M\|k*f\|_{r^{\prime}}^{r^{\prime}/r}\leq M^{\beta+1}\|f\|_{p}^{\beta}=M^{r^{\prime}}. Assuming that kfr>M(1δ)\|k*f\|_{r^{\prime}}>M(1-\delta), we get

Mr(1δ)r<kfrr(F,G)Mr.M^{r^{\prime}}(1-\delta)^{r^{\prime}}<\|k*f\|_{r^{\prime}}^{r^{\prime}}\leq(F,G)M^{r^{\prime}}.

Therefore, (F,G)>(1δ)r(F,G)>(1-\delta)^{r^{\prime}}.

Let ε>0\varepsilon>0 be given. Find the corresponding η\eta as in Hanche-Olsen’s lemma. Define δ\delta by the equation 1η=(1δ)r1-\eta=(1-\delta)^{r^{\prime}}. According to the above, we have the inequality fBfp<ε\|f-Bf\|_{p}<\varepsilon, as required. ∎

5.5 Convergence to a maximizer in the class Max\mathrm{Max} (rather than in SMax\mathrm{SMax})

Proposition 5.6.

Let (fn)(f_{n}) be a maximizing sequence for the convolution operator K:LpLrK:\,L_{p}\to L_{r^{\prime}} with kernel kLq(n)k\in L_{q}(\mathbb{R}^{n}). There exists a subsequence (fnk)(f_{n_{k}}) and shift vectors aka_{k} such that the sequence TakfnkT_{a_{k}}f_{n_{k}} converges in LpL_{p} as kk\to\infty (its limit automatically being a maximizer for the operator KK).

Proof.

In the proof of Theorem 1 (see § 2) we found that the sequence (Bfn)(Bf_{n}) has a subsequence convergent after appropriate shifts. We may assume that the sequence (Bfn)(Bf_{n}) itself is convergent. Proposition 5.5 implies that fnBfnp0\left\|f_{n}-Bf_{n}\right\|_{p}\to 0 (since KfnrKp,r\|Kf_{n}\|_{r^{\prime}}\to\|K\|_{p,r}.) Therefore limfn=limBfn\lim f_{n}=\lim Bf_{n} does exist. ∎

5.6 Kernel approximation and convergence of maximizers

Proposition 5.7.

Suppose a sequence of function knLqk_{n}\in L_{q} converges (strongly) to a nonzero kLqk\in L_{q}. Then there exists a sequence of maximizers fnLpf_{n}\in L_{p} for the convolution operators Kn=Kkn:LpLrK_{n}=K_{k_{n}}:\,L_{p}\to L_{r^{\prime}} that converges strongly to a function fLpf\in L_{p}. The function ff is a maximizer for the convolution operator K=KkK=K_{k}.

Proof.

An arbitrary sequence (fn)(f_{n}) of maximizers for the operators KnK_{n} is obviously a maximizing sequencefor the operator KK. Applying Proposition 5.6, we obtain the claim as stated. ∎

5.7 On boundedness and integrability of maximizers

Proposition 5.8.

Let kLqk\in L_{q} and ff be a maximizer for the convolution operator KkK_{k} from LpL_{p} to LrL_{r^{\prime}}. (We assume that neither of pp, qq and rr is 0 or \infty.)

(a) If kLq+εk\in L_{q+\varepsilon} for some ε>0\varepsilon>0, then fLpLf\in L_{p}\cap L_{\infty}.

(b) If kLqεk\in L_{q-\varepsilon} for some ε>0\varepsilon>0, then fLpLp#f\in L_{p}\cap L_{p_{\#}}, where

p#={1,r/pq,p(1+p/r)1,r/p>qp_{\#}=\left\{\begin{array}[]{ll}1,&\;r^{\prime}/p^{\prime}\leq q,\\ p(1+p^{\prime}/r^{\prime})^{-1},&\;r^{\prime}/p^{\prime}>q\end{array}\right.

(1<p#<p1<p_{\#}<p for r/p>qr^{\prime}/p^{\prime}>q).

Proof.

The only information we need is that ff satisfies the equation Bf=fBf=f (see Subsection 5.4). Put g=Brpfg=B^{p}_{r}f (in notation of Subsection 4.2).

(a) Suppose that fLsf\in L_{s} for 1/pμ1/s1/p1/p-\mu\leq 1/s\leq 1/p with some μ[0,1/p]\mu\in[0,1/p]. The identity

1s+1q1=1r+(1s1p)\frac{1}{s}+\frac{1}{q}-1=\frac{1}{r^{\prime}}+\left(\frac{1}{s}-\frac{1}{p}\right)

shows that kfLuk*f\in L_{u} if max(0, 1/rμ)1/u1/r\max(0,\,1/r^{\prime}-\mu)\leq 1/u\leq 1/r^{\prime}. Therefore, gLtg\in L_{t} if max(0, 1/rν)1/t1/r\max(0,\,1/r-\nu)\leq 1/t\leq 1/r, where ν=(r/r)μ\nu=(r^{\prime}/r)\mu.

Note that μ=1/p\mu=1/p implies ν=(r/p)/r>1/r\nu=(r^{\prime}/p)/r>1/r, so that fLf\in L_{\infty} implies gLg\in L_{\infty}.

Since f=Bf=Bprgf=Bf=B^{r}_{p}g, we have similarly: if gLtg\in L_{t} for 1/rν1/t1/r1/r-\nu\leq 1/t\leq 1/r, ν[0,1/r]\nu\in[0,1/r], then fLsf\in L_{s} for max(0, 1/pκ)1/s1/p\max(0,\,1/p-\kappa)\leq 1/s\leq 1/p, where κ=(p/p)ν\kappa=(p^{\prime}/p)\nu. Also gLg\in L_{\infty} implies fLf\in L_{\infty}.

Combining the above said, we conclude: if fLpLPf\in L_{p}\cap L_{P}, P>pP>p, and μ=1/p1/P\mu=1/p-1/P, then fLpLP~f\in L_{p}\cap L_{\tilde{P}}, where either P~=\tilde{P}=\infty, or (r/r)pμ<1(r^{\prime}/r)p^{\prime}\mu<1 and 1/p1/P~=Mμ1/p-1/\tilde{P}=M\mu, where M=(r/r)(p/p)M=(r^{\prime}/r)(p^{\prime}/p). Since r/p>1r^{\prime}/p>1 and p/r>1p^{\prime}/r>1, we have M>1M>1. Iterating, we get fLf\in L_{\infty} in a finite number of steps.

The conclusion fLf\in L_{\infty} is obtained under the assumpiton that ff lies in LPL_{P} with some P>pP>p, and in the derivation we used just the inclusion kLqk\in L_{q}. Let us now make use of the condition kLq+εk\in L_{q+\varepsilon}, assuming only that fLpf\in L_{p}. Put δ=1/q1/(q+ε)\delta=1/q-1/(q+\varepsilon). Interchanging the roles of ff kk at the first half-step of the iteration (where we estimate the exponent of the space containing kfk*f), we conclude that f=BfLPf=Bf\in L_{P}, where 1/P=max(0, 1/pMδ)1/P=\max(0,\,1/p-M\delta). If PP\neq\infty, we apply the above described iteration with initial value of parameter μ=Mδ\mu=M\delta.

(b) Repeating the argument of part (a), we obtain: if fLsf\in L_{s} for 1/p1/s1/p+μ1/p\leq 1/s\leq 1/p+\mu, then gLtg\in L_{t} for 1/r1/tmin(1, 1/r+ν)1/r\leq 1/t\leq\min(1,\,1/r+\nu), ν=(r/r)μ\nu=(r^{\prime}/r)\mu. Symmetrically, if gLtg\in L_{t} for 1/r1/t1/t+ν1/r\leq 1/t\leq 1/t+\nu, then fLsf\in L_{s} for 1/p1/smin(1, 1/p+κ)1/p\leq 1/s\leq\min(1,\,1/p+\kappa), κ=(p/p)ν\kappa=(p^{\prime}/p)\nu.

The essential difference with part (a) is that the conditions fL1f\in L_{1} are gL1g\in L_{1} no longer equivalent. For instance, gL1g\in L_{1} means that ν=1/r\nu=1/r^{\prime}. The value 1/p+κ=1/p+p/(pr)1/p+\kappa=1/p+p^{\prime}/(pr^{\prime}) can happen to be less than 11.

With this remark in mind, we parallel the proof of part (a). The condition kLqεk\in L_{q-\varepsilon} implies fLPf\in L_{P} with some P<pP<p. Putting ν=1/P1/p\nu=1/P-1/p and M=(pr)/(pr)>1M=(p^{\prime}r^{\prime})/(pr)>1, we obtain at an iteration step: either (i) gL1g\in L_{1}, or (ii) fL1f\in L_{1}, or (iii) fLP~f\in L_{\tilde{P}}, where 1/P~=1/p+Mμ<11/\tilde{P}=1/p+M\mu<1. In the case (iii) we continue to iterate. Eventually, in a finite number of steps one of the cases (i) or (ii) occurs.

The exponent p#p_{\#} in the terminal case (i) is determined above: 1/p#=min(1, 1/p+p/(pr))1/p_{\#}=\min(1,\;1/p+p^{\prime}/(pr^{\prime})). The calculation

1(1p+ppr)=1p+1rpr=1qpr1-\left(\frac{1}{p}+\frac{p^{\prime}}{pr^{\prime}}\right)=\frac{1}{p^{\prime}}+\frac{1}{r^{\prime}}-\frac{p^{\prime}}{r^{\prime}}=\frac{1}{q}-\frac{p^{\prime}}{r^{\prime}}

shows that the condition p#>1p_{\#}>1 is equivalent to the inequality r/p>qr^{\prime}/p^{\prime}>q. ∎

Remark 5.9.

The asymmetry of the result (fLf\in L_{\infty} being a “more common” property than fL1f\in L_{1}) is ultimately due to the fact that convolution inherits best local properties of the two its operands, but worst global properties.

5.8 On the lower bound of convolution operators’ norms

Proposition 5.10.

Let q1q\geq 1, 1<p,r<1<p,r<\infty and 1/p+1/q+1/r=21/p+1/q+1/r=2. Let kLq(d)k\in L_{q}(\mathbb{R}^{d}) be a nonnegative function. The operator of convolution with complex-valued kernel k(x)eiϕ(x)k(x)e^{i\phi(x)} acting from LpL_{p} to LrL_{r^{\prime}} can have arbitrarily small norm. Specifically, Kk(x)exp(iλx2)p,r0\|K_{k(x)\exp(i\lambda\|x\|^{2})}\|_{p,r}\to 0 as λ\lambda\to\infty.

Proof.

It is easy to see that the set of functions kLqk\in L_{q} for which the statement if true is closed in LqL_{q}. Therefore without loss of generality we may assume that kL1Lk\in L_{1}\cap L_{\infty}.

Denote KλK_{\lambda} the operator of convolution with function kλ(x)=k(x)exp(iλx2)k_{\lambda}(x)=k(x)\exp(i\lambda\|x\|^{2}).

In the case q=1q=1, p=r=2p=r=2 the validity of the claim of the Proposition is established below, in Lemma 5.11. The general case follows from this particular one by an interpolation argument as follows.

Suppose prp\geq r (otherwise one considers the transposed operator). On the coordinate plane, let us pass a line through the points A=(1/2,1/2)A=(1/2,1/2) B=(1/p,1/r)B=(1/p,1/r^{\prime}). Let C=(ξ,0)C=(\xi,0) be the point where it meets the horizontal axis. Due to the inequalities 1/r<1/p11/r1/r^{\prime}<1/p\leq 1-1/r^{\prime} we have 0<ξ10<\xi\leq 1. Put ξ=1/s\xi=1/s, s1s\geq 1. The fact that B[AC)B\in[AC) can be written as

1p=1θs+θ2,1r=1θ+θ2,\frac{1}{p}=\frac{1-\theta}{s}+\frac{\theta}{2},\qquad\frac{1}{r^{\prime}}=\frac{1-\theta}{\infty}+\frac{\theta}{2},

where 0<θ10<\theta\leq 1.

Given ε>0\varepsilon>0, Lemma 5.11 tells us that Kλ2,2ε\|K_{\lambda}\|_{2,2}\leq\varepsilon for a large enough λ\lambda. On the other hand, due to the assumption we made at the beginning of the proof, we have kLsk\in L_{s^{\prime}}. By Hölder’s inequality,

Kλffsks.\|K_{\lambda}f\|_{\infty}\leq\|f\|_{s}\|k\|_{s^{\prime}}.

Applying now the Riesz-Thorin theorem, we conclude that

Kλp,rks1θεθ.\|K_{\lambda}\|_{p,r}\leq\|k\|_{s^{\prime}}^{1-\theta}\varepsilon^{\theta}.

The proposition is proved. ∎

Lemma 5.11.

Let kL1(d)k\in L_{1}(\mathbb{R}^{d}). Denote kλ(x)=k(x)exp(iλx2)k_{\lambda}(x)=k(x)\exp(i\lambda\|x\|^{2}) and

k^λ(ξ)=kλ(x)ei(x,ξ)𝑑x,\hat{k}_{\lambda}(\xi)=\int k_{\lambda}(x)e^{-i(x,\xi)}\,dx,

the Fourier transform of kλk_{\lambda}. Then k^λ0\|\hat{k}_{\lambda}\|_{\infty}\to 0 as |λ||\lambda|\to\infty.

Consequently, the norm of the convolution with kλk_{\lambda} as an operator in L2(d)L_{2}(\mathbb{R}^{d}) tends to 0 as |λ||\lambda|\to\infty.

Proof.

By a density argument, it suffices to prove the Lemma under the assumption k(x)C0()k(x)\in C_{0}^{\infty}(\mathbb{R}).

For Rez>0\operatorname{Re}z>0 we have the Plancherel identity

ϕ(x)ezx2𝑑x=Czd/2ϕ^(ξ)eξ2/(4z)𝑑ξ.\int\phi(x)e^{-z\|x\|^{2}}\,dx=Cz^{-d/2}\int\hat{\phi}(\xi)e^{-\|\xi\|^{2}/(4z)}\,d\xi.

Both sides are defined and continuous in the region Rez0\operatorname{Re}z\geq 0, z0z\neq 0. Therefore the equality extends to the boundary z=iλz=-i\lambda, λ{0}\lambda\in\mathbb{R}\setminus\{0\}. Thus,

|ϕ(x)eiλx2|C|λ|d/2ϕ^1.\left|\int\phi(x)e^{i\lambda\|x\|^{2}}\right|\leq C|\lambda|^{-d/2}\|\hat{\phi}\|_{1}.

Putting ϕ(x)=k(x)ei(x,η)\phi(x)=k(x)e^{-i(x,\eta)}, we obtain|k^λ(η)||\hat{k}_{\lambda}(\eta)| in the left-hand side of the latter inequality, while ϕ^(ξ)=k^(ξ+η)\hat{\phi}(\xi)=\hat{k}(\xi+\eta), so that ϕ^1=k^1\|\hat{\phi}\|_{1}=\|\hat{k}\|_{1}. The esimate

k^λC|λ|d/2k^1\|\hat{k}_{\lambda}\|_{\infty}\leq C|\lambda|^{-d/2}\|\hat{k}\|_{1}

follows and the proof is complete. ∎

6 Best constants in the Hausdorff-Young inequality for the Laplace transform on (0,+)(0,+\infty)

Denote by \mathcal{L} the Laplace transform on +\mathbb{R}_{+},

ff(x)=0extf(t)𝑑t,f\;\mapsto\;\mathcal{L}f(x)=\int_{0}^{\infty}e^{-xt}f(t)\,dt,

and by \mathcal{F} the Fourier transform on \mathbb{R},

ff(x)=eixtf(t)𝑑t.f\;\mapsto\;\mathcal{F}f(x)=\int_{-\infty}^{\infty}e^{-ixt}f(t)\,dt.

For 1p21\leq p\leq 2, the Hausdorff-Young (HY) inequalities

fpCpfp\|\mathcal{F}f\|_{p^{\prime}}\leq C^{\mathcal{F}}_{p}\|f\|_{p}

and

fpCpfp\|\mathcal{L}f\|_{p^{\prime}}\leq C^{\mathcal{L}}_{p}\|f\|_{p}

hold. They are first established under the assumption fL1Lf\in L_{1}\cap L_{\infty}, when the integral definitions of f\mathcal{L}f and f\mathcal{F}f have direct meaning, and then they are used to extend \mathcal{L} and \mathcal{F} by continuity to the operators acting from LpL_{p} to LpL_{p^{\prime}}.

The exponent pp^{\prime} in the left-hand sides of the HY inequalities cannot be replaced by any other number. This follows from “dimensional analysis”, that is, changing f(t)f(t) into the function fa(t)=a1/pf(at)f_{a}(t)=a^{1/p}f(at) with the same LpL_{p} norm, where a>0a>0 is an arbitrary scaling parameter. It is also known that inequalities of this type do not hold when p>2p>2. In the case of Fourier transform, an explicit argument to that effect can be found, e.g., in Titchmarsch’s monograph [31, § 4.11].

The optimal values of the constants CpC^{\mathcal{F}}_{p}, that is, the operator norms pp\|\mathcal{F}\|_{p\to p^{\prime}}, have been found by W. Beckner [6] (and earlier by K.I. Babenko [4] in the case p/2p^{\prime}/2\in\mathbb{Z}):

pp=(2π)1/pAp,\|\mathcal{F}\|_{p\to p^{\prime}}=(2\pi)^{1/p^{\prime}}A_{p}, (6.1)

where the constant ApA_{p} is defined in (1.2).

Analytical expressions for the optimal values of the constants CpC^{\mathcal{L}}_{p}, that is, the operator norms N(p)=ppN(p)=\|\mathcal{L}\|_{p\to p^{\prime}}, are unknown. The problem of determining N(p)N(p) is equivalent to the problem of determining the norm of the convolution operator with kernel hp()h_{p}(\cdot), see (6.6) below, acting from Lp()L_{p}(\mathbb{R}) to Lp()L_{p^{\prime}}(\mathbb{R}).

In Figure 3 and in Table 1 we present the numerical values of N(p)N(p). In order to mark the distiction between the true value of N(p)N(p) and the computed approximation to it, we designate the latter as N(p)N^{\circ}(p). The numerical method used is briefly outlined at the end of this Section.

The minimum of the norm occurs at p1.1307p\approx 1.1307,

minN(p)0.881970846.\min N(p)\approx 0.881970846. (6.2)

Refer to caption


Figure 3: The norms N(p)ppN^{\circ}(p)\approx\|\mathcal{L}\|_{p\to p^{\prime}} found numerically in comparison with the analytical estimates (6.3), (6.5), (6.7) (6.8)

pp 1.051.05 1.11.1 1.21.2 1.31.3 1.41.4 1.51.5 1.61.6 1.71.7 1.81.8 1.91.9 N(p)N^{\circ}(p) 0.908350.90835 0.884950.88495 0.893060.89306 0.935620.93562 0.998330.99833 1.076521.07652 1.168901.16890 1.276311.27631 1.401931.40193 1.553901.55390 CSB(p)C_{SB}(p) 0.914590.91459 0.896400.89640 0.912960.91296 0.961690.96169 1.028301.02830 1.108031.10803 1.199531.19953 1.303541.30354 1.423101.42310 1.566161.56616

Table 1: The norms N(p)ppN^{\circ}(p)\approx\|\mathcal{L}\|_{p\to p^{\prime}} found numerically in comparison with Setterqvist’s estimate (6.8)

The curves in Fig. 3 present the numerically evaluated norms p,p\|\mathcal{L}\|_{p,p^{\prime}} and several analytical estimates for the norms, which we describe below.

1. The simplest estimate is obtained by interpolation. The equality N(1)=1N(1)=1 is immediate and the equality N(2)=πN(2)=\sqrt{\pi} is readily obtained as the supremum of the spectrum of the self-adjoint operator \mathcal{L} in L2(+)L^{2}(\mathbb{R}_{+}). The Riesz-Thorin interpolation theorem yields the estimate666The constants are subscripted in accordance with: RT=Riesz-Thorin, F=via Fourier norm, H=Hardy, S=Setterqvist.

N(p)CRT(p)=π1/p,1p2.N(p)\leq C_{RT}(p)=\pi^{1/p^{\prime}},\quad 1\leq p\leq 2. (6.3)

2. One can show that

N(p)21/ppp.N(p)\leq 2^{-1/p^{\prime}}\|\mathcal{F}\|_{p\to p^{\prime}}. (6.4)

Using the Hausdorff-Young estimate pp(2π)1/p\|\mathcal{F}\|_{p\to p^{\prime}}\leq(2\pi)^{1/p^{\prime}} in the right-hand side we come again to the estimate (6.3), but one can instead use Beckner’s sharp constant. As a result, one gets a better estimate,

N(p)CF=π1/pAp.N(p)\leq C_{F}=\pi^{1/p^{\prime}}A_{p}. (6.5)

Let us comment on the inequality (6.4). Consider the family of operators Tz:Lp(+)Lp(+)T_{z}:\;L_{p}(\mathbb{R}_{+})\to L_{p^{\prime}}(\mathbb{R}_{+}), depending on complex parameter zz,

Tz:fTzf(x)=0eΦ(z)xyf(y)dy,Φ(z)=eizπ2.T_{z}:\;f\mapsto T_{z}f(x)=\int_{0}^{\infty}e^{-\Phi(z)xy}\,f(y)\,dy,\qquad\Phi(z)=e^{iz\frac{\pi}{2}}.

The analytic operator-valued function zTzz\mapsto T_{z} is defined in the strip |Rez|1|\operatorname{Re}z|\leq 1 and its values at z=±1z=\pm 1 are the composition of the Fourier transform with restrictions onto the negative, resp., positive real half-line. The value TzT_{z} at z=0z=0 is but the Laplace transform. The inequality in question follows by applying Stein’s interpolation theorem [28, Ch. 5, Theorem 4.2]. Using this approach, the second author and A.E. Merzon have obtained a variant of the HY inequality for the Laplace transform with variable yy on a ray in the complex plane (unpublished).

3. The substitutions

x=ey,t=es,F(y)=f(ey)ey/p,G(s)=(f)(es)es/p,x=e^{y},\qquad t=e^{s},\qquad F(y)=f(e^{y})e^{y/p},\qquad G(s)=(\mathcal{L}f)(e^{s})\,e^{s/p^{\prime}},

reduce the Laplace transform to the convolution operator

F(y)G(s)=hp(y+s)F(y)𝑑y=(hpF~)(s),F(y)\;\mapsto\;G(s)=\int_{\mathbb{R}}h_{p}(y+s)\,F(y)\,dy=(h_{p}*\tilde{F})(s),

where

hp(y)=ey/pey.h_{p}(y)=e^{y/p^{\prime}-e^{y}}. (6.6)

It is easy to see that the LpL_{p}-norms of the functions ff (defined on +\mathbb{R}_{+}) and FF (defined on \mathbb{R}) coincide; the same is true for the LpL_{p^{\prime}}-norms of the functions f\mathcal{L}f and GG. Therefore, N(p)N(p) is the norm of the convolution operator with kernel hp(y)h_{p}(y) acting from Lp()L_{p}(\mathbb{R}) to Lp()L_{p^{\prime}}(\mathbb{R}). Since hpqq=2π/p\|h_{p}\|_{q}^{q}=\sqrt{2\pi/p^{\prime}} (here q=p/2q=p^{\prime}/2), the Young inequality yields the estimate

N(p)CH(p)=(2πp)1/p.N(p)\leq C_{H}(p)=\left(\frac{2\pi}{p^{\prime}}\right)^{1/p^{\prime}}. (6.7)

G.H. Hardy [16] was the first to derive this estimate in 1933, using the method just outlined.

4. Combining Hardy’s reduction with case d=1d=1 of Beckner’s sharp form (1.2) of Young’s inequality, Setterqvist [27, Theorem 2.2] obtained the estimate

N(p)CS(p)=(π(p1))1/p(p(2p))1/p1/2=CH(p)Ap2Aq.N(p)\leq C_{S}(p)=\left(\pi(p-1)\right)^{1/p^{\prime}}\,\left(p(2-p)\right)^{1/p-1/2}=C_{H}(p)\cdot A_{p}^{2}\cdot A_{q}. (6.8)

The maximum relative error of the estimate (6.8) is about 3%3\%. The following empirical approximation has absolute error less 10310^{-3}:

CS(p)N(p)(p1)(2p)8.C_{S}(p)-N(p)\approx\frac{(p-1)(2-p)}{8}.

The numerical method

In notation of Section 4.2, we have the equation Bppf~=fB_{p}^{p}\tilde{f}=f to solve. Its solution (if it exists) is also a solution of the equation Bf=fBf=f, which any maximizer must satisfy. In practice, we solve the equation Bppf(x)=f(bx)B_{p}^{p}f(x)=f(b-x), where both the function ff and the shift parameter bb are to be determined. We employ the direct iteration method defined by

fn(x)=(Bppfn1)(bnx).f_{n}(x)=(B^{p}_{p}f_{n-1})(b_{n}-x). (6.9)

The shift parameter bnb_{n} at each step is determined by the condition

maxxfn(x)=fn(0)=1.\max_{x}f_{n}(x)=f_{n}(0)=1.

The recurrence (6.9) implies that fnp=1\|f_{n}\|_{p}=1 for all n>1n>1. Due to Lemma 4.3, the sequence of norms kfnp\|k*f_{n}\|_{p^{\prime}} is nondecreasing; due to the Young inequality, it is bounded; hence a limit exists. The computation is stopped when kfnppkfn1pp<ε\|k*f_{n}\|_{p^{\prime}}^{p^{\prime}}-\|k*f_{n-1}\|_{p^{\prime}}^{p^{\prime}}<\varepsilon, where ε\varepsilon is the given tolerance. We chose this criterion because we are not concerned (here) with computation of the solution ff.

The error of the numerical method has two sources besides the machine arithmetics and finiteness of the number of iterations.

(I) Domain compactification: the line \mathbb{R} is replaced by a finite interval I=[a,a]I=[-a,a] and the convolution on \mathbb{R} is replaced by the cyclic convolution on II.

(II) Discretization: the functions of continuous variable are replaced by the functions of a discrete parameter. We use the uniform grid with N=2kN=2^{k} nodes.

As we have noted in Remark 4.2, the existence of solution of the equation Bppf~=fB_{p}^{p}\tilde{f}=f has not been proved. This is not an important issue though: one can follow even-numbered iterations, since B~2=B\tilde{B}^{2}=B and the existence of solution of the equation Bf=fBf=f is known.

The essential gaps in the justification of our numerical method are the following:

(a) a proof of convergence of the iterations (6.9) Lp()L_{p}(\mathbb{R}) (even of the even-numbered iterations) is lacking;

(b) there is no result on uniqueness (up to a shift) of solution of the equation Bf=fBf=f, which means that the limit limnkfnp\lim_{n\to\infty}\|k*f_{n}\|_{p^{\prime}} may in principle depend on the initial condition.

In practice, for a fixed compactification we observed a geometric convergence of the norms kfnp\|k*f_{n}\|_{p^{\prime}} (and, moreover, a geometric convergence of (fn)(f_{n}) in norm), the faster the closer pp is to 11. The limit function appears to be the same for different initial conditions. (We tried the initial conditions being either the Gaussians with various dispersions or the indicator functions of intervals.) In order to control the accuracy of the results, we performed computations with doubling of the number of nodes until stabilization. (In most cases, N=512N=512 nodes were sufficient.)

As regards the compactification, the Young inequality and the triangle inequality provide an upper bound for the error of the computed norm Kkp,p\|K_{k}\|_{p,p} when the support of the kernel kk gets truncated. It is also easy to estimate the error due to the use of the cyclic convolution instead of the convolution on \mathbb{R}. Contrary to the situation with convergence of iterations, the compactification appears to run into trouble as pp approaches 1, as the convolution kernel (6.6) becomes slowly decreasing in the negative directon. For instance, h1.05(y)e|y|/20h_{1.05}(y)\approx e^{-|y|/20} for y<0y<0. However, when truncating the support of the function hph_{p}, we are concerned not with absolute values of the cut-out, but with its LqL_{q}-norm. Since q=p/2q=p^{\prime}/2 and hp(y)q=ey/2(p/2)expyh_{p}(y)^{q}=e^{y/2-(p^{\prime}/2)\exp y}, the truncation parameter can be set uniformly in pp. Computations support these considerations. For the purpose of control we used the doubling of the support [L,L][-L,L] of the truncation of hph_{p} (and the corresponding doubling of the length of the circumference obtained by identifying the ends of the interval).

All significant digits of the numerical data presented in Table 1 and in the formula (6.2) are found to be stable with respect to the described operations of parameters doubling.

7 Open questions

Question 1.

Let kLq(d)k\in L_{q}(\mathbb{R}^{d}) and f1f_{1}, f2f_{2} be two maximizers for the operator KkK_{k} from LpL_{p} to LrL_{r^{\prime}}. Is it true that there exist θ\theta\in\mathbb{R} and a vector ada\in\mathbb{R}^{d} such that f2(x)=eiθf1(xa)f_{2}(x)=e^{i\theta}f_{1}(x-a)?

Having posed (and solved) the question of the existence of a minimizer, it is natural to ask about its uniqueness up to the trivial transformations. We suppose that in general there is no uniqueness. It looks probable however that one can formulate conditions sufficient for uniqueness and embracing some narrow but meaningful class of functions kk (positive? unimodular?…). Exploring a finite-dimensional analog — the convolution on /m\mathbb{Z}/m\mathbb{Z} — might help to understand what effects one should anticipate.

In the case non-uniqueness is revealed, a number of further questions can be asked, concerning non-maximizer solutions of the equation Bf=fBf=f, bifurcation phenomena, Morse indices etc.

Question 2.

Let 1<p<1<p<\infty, p2p\neq 2. Does there exist a nonzero kernel kL1k\in L_{1}, for which the convolution operator Kk:LpLpK_{k}:\,L_{p}\to L_{p} possesses a maximizer?

The affirmative answer in the case p=2p=2 is given in Subsection 5.1 (subcase I(B)2\mathrm{I(B)_{2}}).

Question 3.

Let kLqk\in L_{q} and VV be a compact operator from LpL_{p} to LrL_{r^{\prime}}. Is it true that a maximizer for the operator Kk+VK_{k}+V exists? Is this true in the particular case when VV is a rank one operator?

An answer to this question would yield either an extension of the class of admissible integral kernels in Theorem 1 or an yet another counterexample, in addition to the one given in Subsection 5.3, stressing the role of translation-invariance towards the existence of a maximizer.

Question 4.

Generalize Theorem 1 to embrace a certain class of locally compact groups (in particular, a (sub)class of discrete finitely generated groups).

The proof of Theorem 1 goes through in d\mathbb{Z}^{d} with trivial modifications.777The situation in the limit cases differs between d\mathbb{R}^{d} and d\mathbb{Z}^{d}, cf. end of comment (8). One can try to get a clue about the case of discrete non-commutative groups by studying convolution on the free groups with two generators. One should be aware of the fact that the condition (1.1) on the exponents in the Young inequality is, for a general locally-compact group, not necessary, cf. [26].

Question 5.

Investigate the local and global properties of a maximizer as depending on the properties of the kernel kk.

We have stated this question in a broad and imprecise form. Here are more specific sample questions, which we would be interested to have answered.

Question 5A. What is a guaranteed rate of decay of a maximizer provided kLqk\in L_{q} has finite support?

Question 5B. What condition on kk (“a room of integrability”) beyond the assumed kLqk\in L_{q} is sufficient to guarantee that a maximizer lies in LL_{\infty}?

According to Proposition 5.8, it is sufficient that kLq+εk\in L_{q+\varepsilon} with arbitrarily small ε>0\varepsilon>0. Isn’t an “inner room of integrability” of the kernel kk ( with respect to LqL_{q}) and/or a maximizer (with respect to LpL_{p}) already sufficient? By an “inner room of integrability” we mean that, for example, kk belongs to some Orlicz space properly contained in LqL_{q}. In the proof of Proposition 5.8, one can replace the reference to Young’s inequality by the reference to O’Neil’s inequality [24] (concerning convoulution in Orlicz spaces), but it is unclear how far one can get with this approach.

Question 6.

Conjecture. For any ε>0\varepsilon>0 there exists δ>0\delta>0 (depending on q,pq,p and the convolution kernel kk) such that if fp=1\|f\|_{p}=1 and kfr>Kp,r(1δ)\|k*f\|_{r^{\prime}}>\|K\|_{p,r}(1-\delta), then infg𝔐kfgp<ε\inf_{g\in\mathfrak{M}_{k}}\|f-g\|_{p}<\varepsilon, where 𝔐k\mathfrak{M}_{k} is the set of all maximizers for the operator KK.

The formulation of the Conjecture parallels that of Proposition 5.5, cf. comment (8).

Question 7.

Find a lower bound for the LpL_{p}-distance from the function hph_{p} defined by (6.6) to the set of Gaussians, that is, estimate from below the quantity

δp=infr,b>0;mhp(x)reb(xm)2p.\delta_{p}=\inf_{r,b>0;m\in\mathbb{R}}\left\|h_{p}(x)-re^{-b(x-m)^{2}}\right\|_{p}.

This question is a step to the analytical improvement of the inequality (6.8):

N(p)CS(p)cδphpq3=CS(p)cδpCH(p)3,N(p)\leq C_{S}(p)-c\delta_{p}\|h_{p}\|_{q}^{3}=C_{S}(p)-c\delta_{p}C_{H}(p)^{3},

where the constants CS(p)C_{S}(p) and CH(p)C_{H}(p) are from (6.8) and (6.7), respectively, and cc is the constant in the (one-dimensional) Christ inequality, see comment (8) in Section 8.

Question 8.

Prove the convergence of the iterations (6.9).

This question allows a broad interpretation (convergence of the iterations fn+1=Bfnf_{n+1}=Bf_{n} under some general assumptions), as well as a narrow interpretation: explain analytically why the iterations converge in the concrete situation of Section 6.

A potential non-uniqueness of solution of the equation Bf=fBf=f (cf. Question 1) may call for certain adjustments of the question in its broad interpretation.

Note that in order to compute just the norm of the operator KkK_{k} (and not a maximizer) all that matters is not the convergence of iterations but exactly the absence of an extraneous solution fwrongf_{wrong} with a small norm.

8 Comments

Section 1

(8) The inequality (1.2) has been proved independently and almost at the same time in [6] and in [8]. See also the textbook [19, Theorem 4.2]. We note a simple proof given in [5] (essintially based but on Hölder’s inequality) and a particularly elegant proog in [10] (exploiting monotonicity of the trilinear form (f1f2,f3)(f_{1}*f_{2},f_{3}) under heat equation evolution of the functions fif_{i}).

A discussion of the Young inequality on locally compact groups with emphasis on admissible exponents and sharpness of the constants can be found in [14, 26].

(8) For review-style expositions of the results and methods of Christ’s work [11, 12] we refer to [13], [32].

The result particularly relevant to a possible improvement of Setterqvist’s estimate (6.8) is [12, Corollary 1.5]:

Let fjLj(d)f_{j}\in L_{j}(\mathbb{R}^{d}) (j=1,2,3j=1,2,3) and f1p1=f2p2=f3p3=1\|f_{1}\|_{p_{1}}=\|f_{2}\|_{p_{2}}=\|f_{3}\|_{p_{3}}=1. Put C=(Ap1Ap2Ap3)dC=(A_{p_{1}}A_{p_{2}}A_{p_{3}})^{d} (which is Beckner’s constant in d\mathbb{R}^{d}) and denote 𝔊\mathfrak{G} the set of all Gaussian functions,

𝔊={ϕ(x)=re(Bx,x)+(a,x)},\mathfrak{G}=\left\{\phi(x)=re^{-(Bx,x)+(a,x)}\right\},

where r>0r>0, ada\in\mathbb{R}^{d} and BB is a positive definite quadratic form. There exists a constant c>0c>0 (which depends on the dimension dd) such that

|(f1f2,f3)|Cε(δ),ε(δ)=cδ4,|(f_{1}*f_{2},f_{3})|\leq C-\varepsilon(\delta),\qquad\varepsilon(\delta)=c\delta^{4},

where

δ=infg1,g2,g3𝔊maxj{1,2,3}fjgjpj.\delta=\inf_{g_{1},g_{2},g_{3}\in\mathfrak{G}}\max_{j\in\{1,2,3\}}\|f_{j}-g_{j}\|_{p_{j}}.

In order to use the stated result for improvement of the estimate (6.8), one needs the numerical value of the constant cc (for d=1d=1), which is not given in [12], as well as a lower estimate for the LpL_{p}-distance from the kernel hph_{p} to the set of Gaussians. We offer the latter calcluation as an open question, see Question7 in Section 7.

(8) In [20, § I.1] the minimization problem for the functional of the form

(u)=ne(x,Au(x))𝑑x\mathcal{E}(u)=\int_{\mathbb{R}^{n}}e(x,Au(x))\,dx

under the constraint 𝒥(u)=λ\mathcal{J}(u)=\lambda is considered. Here

𝒥(u)=nj(x,Bu(x))𝑑x;\mathcal{J}(u)=\int_{\mathbb{R}^{n}}j(x,Bu(x))\,dx;

e(,)e(\cdot,\cdot), j(,)j(\cdot,\cdot) are given real-valued functions, j0j\geq 0; u()u(\cdot) are elements of a given function space on n\mathbb{R}^{n}. Denote

Iλ=inf𝒥(u)=λ(u).I_{\lambda}=\inf_{\mathcal{J}(u)=\lambda}\mathcal{E}(u).

A particular case with xx-independent functions ee, jj is referred to, in the general context, as “problems at infinity”.

Our problem concerning the norm of the operator KkK_{k} corresponds to

(u)=ukrr,𝒥(u)=upp,\mathcal{E}(u)=-\|u*k\|_{r^{\prime}}^{r^{\prime}},\qquad\mathcal{J}(u)=\|u\|_{p}^{p},

i.e. e(u)=|u|re(u)=|u|^{r^{\prime}}, Au=ukAu=u*k, Bu=uBu=u, j(u)=|u|pj(u)=|u|^{p}. The lower bound then is

Iλ=Cλγ,γ=rp,C=Kkp,rr.I_{\lambda}=-C\lambda^{\gamma},\quad\gamma=\frac{r^{\prime}}{p},\quad C=\|K_{k}\|_{p,r}^{r^{\prime}}. (8.1)

What Lions’ method provides is not a single general theorem but a general approach to proving the existence of extermizers in a broad class of variational problems of analysis and mathematical physics. It contains heuristic elements, so that details may vary and require a concrete, problem-specific approach.

The monograph [30] treats many aspects of the concentration compactness method, with emphasis on convergence in Hilbert (Sobolev) spaces. A Russian-language reader may find Ch. 5 in the textbook [21] as a useful reference concerning Lions’ method.

By all indications, it should be possible to prove Theorem 1 in the framework of Lions’ method; however, this would be a separate and not quite trivial project. Note that our proof neither refers to the Concentration Compactness Lemma [20, Lemma I.1], [21, Lemma 5.1] nor contains its close analog; the variants of “vanishing” and “dichotomy” are implicitly eliminated by other means.

(8) In T. Tao’s methodical article [29, § 1.6], a technique of “profile decomposition” is discussed: a “profile” (a function sequence) is decomposed into a sum of shifts of fixed functions and a relatively compact sequence (cf. [30, § 3.3, Theorem 3.1]). As an application, a “toy theorem” is proved, asserting that the discrete convolution operator acting from 1=L1()\ell^{1}=L_{1}(\mathbb{Z}) to p\ell^{p} by the formula (Kf)n=fnfn1(Kf)_{n}=f_{n}-f_{n-1} has a maximizer. Note that in the corresponding case II(A) of Subsection 5.1 a maximizer in \mathbb{R} does not exist.

Section 2

(8) Some notions (the “δ\delta-near” ones) introduced in Definitions 2.32.4 are there just to suit our local purposes, while other notions, with their origins in Probability Theory, have been used in different contexts. Among the latter, the term tight is standard, cf. e.g. [7, v. 2, § 8.6]. Let us comment on the remaining two.

1. The δ\delta-diameter introduced in Definition 2.3 is Lévy’s dispersion function [17, Section 1.1, Supplement 4] in disguise. Specifically, for a fixed fLp(d)f\in L_{p}(\mathbb{R}^{d}) with unit pp-norm and a unit vector vdv\in\mathbb{R}^{d}, we have the distribution function in the sense of probability theory

F(t)=(v,x)<t|f(x)|p𝑑x.F(t)=\int_{(v,x)<t}|f(x)|^{p}\,dx.

The corresponding Lévy concentration function [17] is

QF(λ)=supt(F(t+λ)F(t)).Q_{F}(\lambda)=\sup_{t\in\mathbb{R}}(F(t+\lambda)-F(t)).

The inverse function is known as the dispersion function for the measure dFdF; in our notation it is

Dδ,vp(f)=infQF(λ)1δλ.D^{p}_{\delta,v}(f)=\inf_{Q_{F}(\lambda)\geq 1-\delta}\lambda.

2. We thought it useful to have a shorter name for the property of a function sequence to be tight up to translations; we call such a sequence relatively tight. P. L. Lions, in the formulation of his Concentration Compactness Lemma [20, Lemma I.1], chose to characterize the said property as the “case of compactness” rather than to devise a descriptive adjective.

(8) We do not claim uniqueness of the δ\delta-near-support, cf. [17, § 1.1.2].

Section 3

(8) The inequality of Lemma 3.1 is interpreted in Lions’ theory as the subadditivity property (of crucial importance) of the fucntion IλI_{\lambda}, cf. (8.1). Note that the exponent γ\gamma in the final application of Lemma 3.1 (see Lemma 3.7) coincides with that in (8.1).

(8) The function in the left-hand side of the inequality of Lemma 3.3 is known as Steklov’s averaging of the function |f||f|; the lemma states one of its most elementary properties. The quantifiers can be swapped (“there exists t0t_{0} such that for any aa…”) at the expense of putting an appropriate constant in the numerator of the right-hand side; this follows from the Hardy-Littlewood maximal inequality.

(8) Perhaps, this place in our proof — the reference to Lemma 3.2 in the proof of Lemma 3.5 — most closely corresponds to Lions’ thesis “prevent the possible splitting of minimizing sequences by keeping them concentrated” [20, p.114], and also reflects the “asymptotic orthogonality” phenomenon [29].

Subsection 4.3

(8) For general integral operators in LpL_{p} spaces, sufficient conditions for compactness usually require some spare room in the space exponents as compared with sufficient conditions for boundedness, cf. e.g. [22, Theorem 7.1].888Note that the usual notation LpL_{p} (or LpL^{p}) corresponds to L1/pL_{1/p} in [22]. As it is readily seen, there is no such “spare room” in the conditions of Lemma 4.8.

A very general study of compositions of convolution and multiplication operators in Lebesgue spaces is found in the paper [9]. Our Lemma 4.8 is a particular case of Theorem 6.4 of [9]; however, it seems easier to give an independent proof, as we did, than to scrutinize involved notation and conditions.

Subsection 5.0

(8) It is harder (likely, much harder) to prove, under the assumptions of Proposition 5.5, the approximative property in the spirit of M. Christ’s results mentioned in the comment (8). Cf. Question 6 in Section 7.

(8) One of general heuristic principles stated by P. L. Lions reads “All minimizing sequences are relatively compact up to a translation iff [a certain] strict subadditivity inequality holds”. [20, p. 114] We took it as a hint that what is now Proposition 5.6 should be valid, although we did not need it in the proof of Theorem 1.

Adapting a notion of shift-compactness [17, Section 5.1.1] to our situation, the short summary can be stated: under the assumptions of Theorem 1, every maximizing sequence of the convolution operator is shift-compact.

(8) The result of Subsection 5.8 is complementary to the results of the paper [23], in which the norms Kkp,r\|K_{k}\|_{p,r} are estimated from below in terms of (absolute values of) the integrals of the kernel over certain families of sets, which are different for positive kernels and general real-valued kernels.

Subsection 5.4

(8) Another approximative version of Hölder’s inequality, with explicit constants, is found in [3].

References

  • [3] J. M. Aldaz, “A stability version of Hölder’s inequality”, J. Math. Anal. Appl. 43, 842–852 (2008).
  • [4] K. I. Babenko, “An inequality in the theory of Fourier integrals”, Amer. Math. Soc. Transl. (II), 44, 115–128 (1965).
  • [5] F. Barthe, “Optimal Young s inequality and its converse: a simple proof”, GAFA, Geom. func. anal. 8, 234–242; https://arxiv.org/pdf/math/9704210.pdf (1998).
  • [6] W. Beckner, “Inequalities in Fourier analysis”, Ann. of Math. 102, 159–182 (1975).
  • [7] V. I. Bogachev, Measure Theory (vol. 1, 2), Springer, 2007.
  • [8] H. J. Brascamp, E. H. Lieb, “Best Constants in Young’s Inequality, Its Converse, and Its Generalization to More than Three Functions”, Adv. in Math. 20, 151–173 (1976).
  • [9] R.C. Busby, H.A. Smith, “Product-convolution operators and mixed-norm spaces”, Trans. AMS, 263 (2), 309–341 (1981).
  • [10] E. A. Carlen, E. H. Lieb, M. Loss “A sharp analog of Young’s inequality on SNS^{N} and related entropy inequalities”, J. Geom. Anal. 14 (3), 487–520 (2004).
  • [11] M. Christ, Near-extremizers of Young’s Inequality for d\mathbb{R}^{d}, https://arxiv.org/pdf/1112.4875.pdf (2011).
  • [12] M. Christ, A sharpened Hausdorff-Young inequality, https://arXiv.org/pdf/1406.1210.pdf (2014).
  • [13] A. Culiuc, “A sharpened Hausdorff-Young inequality”, in: Summer School “Sharp Inequalities in Harmonic Analysis” (Kopp, 2015), R. Frank, D. Oliviera e Silva, C. Thiele (eds.), 11–15 (2015).
  • [14] J. Fournier, “Sharpness in Young s inequality for convolution”, Pacific J. Math. 72 (3), 383–397 (1977).
  • [15] H. Hanche-Olsen, “On the uniform convexity of LpL^{p}”, Proc. Amer. Math. Soc. 134 (8), 2359–2362 (2006).
  • [16] G. H. Hardy, “The constants of certain inequalities”, J. London Math. Soc. 8, 114–119 (1933).
  • [17] W. Hengartner, R. Theodorescu, Concentration functions, Academic Press, New York-London, 1973.
  • [18] E. Hewitt, K. A. Ross, Abstract harmonic analysis, vol. 1, Springer-Verlag, Berlin, 1963.
  • [19] E. H. Lieb, M. Loss, Analysis (2ed.), AMS, 2001.
  • [20] P. L. Lions, “The concentration-compactness principle in the calculus of variations. The locally compact case, part 1”, Ann. de l’I.H.P. Section C,  1, 109–145 (1984).
  • [21] M. O. Korpusov, A. G. Sveshnikov, Nonlinear functional analysis and mathematical modeling in physics. Methods of analysis for nonlinear operators, KRASAND, Moscow, 2011. (In Russian.)
  • [22] M. A. Krasnoselskii, P. P. Zabreiko, E. I. Pustylnik, P. E. Sobolevskii, Integral operators in spaces of summable functions, Noordhoff Int. Publ., Leyden, 1976.
  • [23] E. D. Nursultanov, K. S. Saidahmetov, “On lower bound of the norm of integral convolution operator”, Fundam. Prikl. Mat. 8 (1), 141–150 (2002). (In Russian.)
  • [24] R. O’Neil, “Fractional Integration in Orlicz Spaces. I”, Trans. AMS, 115, 300–328 (1965).
  • [25] M. Pearson, “Extremals for a class of convolution operators”, Houston J. Math. 25, 43–54 (1999).
  • [26] T. S. Quek, L. Y. H. Yap, “Sharpness of Young s inequality for convolution”, Math. Scand. 53, 221–237 (1983).
  • [27] E. Setterqvist, Unitary equivalence: a new approach to the Laplace operator and the Hardy operator, M. Sc. thesis, Luleå Univ. of Technology,, http://ltu.diva-portal.org/smash/get/diva2:1032398/FULLTEXT01.pdf (2005).
  • [28] E. M. Stein, G. Weiss, Introduction to Fourier Analysis on Euclidean Spaces, Princeton Univ. Press, 1971.
  • [29] T. Tao, “Concentration compactness and the profile decompositition”, in: Poincaré’s legacies: pages from year two of a mathematical blog, Part II, , AMS, 2009. http://terrytao.wordpress.com/2008/11/05
  • [30] K. Tintarev, K.-H. Fieseler, Concentration compactness. Functional-analytic grounds and applications, London, Imperial College Press, 2007.
  • [31] E. T. Titchmarsch, Introduction to the Theory of Fourier Integrals, Oxford Univ. Press, 1948.
  • [32] M. Vitturi, Fine structure of some classical affine-invariant inequalities, additive combinatorics and near extremizers (account of a talk by Michael Christ); http://www.maths.ed.ac.uk/~s1251909/christ_talk.pdf (2014).