Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2024, Volume 215, Issue 7, Pages 920–931
DOI: https://doi.org/10.4213/sm10006e
(Mi sm10006)
 

An extremal problem for positive definite functions with support in a ball

A. D. Manov

St Petersburg State University, St Petersburg, Russia
References:
Abstract: The extremal problem under consideration is related to the set of continuous positive definite functions on $\mathbb{R}^n$ with support in a closed ball of radius $r>0$ and fixed value at the origin (the class $\mathfrak{F}_r(\mathbb{R}^n)$).
Given $r>0$, the problem consists in finding the supremum on $\mathfrak{F}_r(\mathbb{R}^n)$ of a functional of a special form.
A general solution to this problem is obtained for $n\neq2$. As a consequence, new sharp inequalities are obtained for derivatives of entire functions of exponential spherical type $\leqslant r$.
Bibliography: 24 titles.
Keywords: positive definite functions, extremal problems, Fourier transform, entire functions of exponential spherical type.
Funding agency Grant number
Russian Science Foundation 23-11-00153
This research was supported by the Russian Science Foundation under grant no. 23-11-00153, https://rscf.ru/en/project/23-11-00153/.
Received: 03.10.2023 and 31.03.2024
Bibliographic databases:
Document Type: Article
MSC: Primary 42B10; Secondary 41A17
Language: English
Original paper language: Russian

§ 1. Introduction

We fix some notation: $|\cdot|$ is the Euclidean norm in $\mathbb{R}^n$, $\mathbb{B}_r:=\{x\in \mathbb{R}^n\colon |x|<r\}$ is the open ball of radius $r>0$ centred at the origin, $\overline{\mathbb{B}_r}$ is the closure of this ball, $\widetilde{f}(x):=\overline{f(-x)}$, $(f\ast g)(x)=\displaystyle\int_{\mathbb{R}^n}f(x-t)g(t)\,dt$, and $L_\infty^{\mathrm{loc}}(\mathbb{R}^n)$ is the space of functions locally bounded almost everywhere on $\mathbb{R}^n$.

A complex-valued function $f\colon \mathbb{R}^n\to\mathbb{C}$ is said to be positive definite on $\mathbb{R}^n$ ($f\in\Phi(\mathbb{R}^n)$) if for each $m\in\mathbb{N}$, any $\{x_i\}_{i=1}^m\subset \mathbb{R}^n$ and each set of complex numbers $\{c_i\}_{i=1}^m\subset\mathbb{C}$ we have the inequality

$$ \begin{equation} \sum_{i,j=1}^m c_i\overline{c_j}f(x_i-x_j)\geqslant0. \end{equation} \tag{1.1} $$

If $f\in\Phi(\mathbb{R}^n)$, then from (1.1), for $m=2$ we obtain $|f(x)|\leqslant f(0)$ for $x\in \mathbb{R}^n$, and $f$ is an Hermitian function, that is, $f=\widetilde{f}$.

In this paper we are interested in the following convex subset of positive definite functions. Let $r>0$. Then we denote by $\mathfrak{F}_r(\mathbb{R}^n)$ the set of functions

$$ \begin{equation*} {\varphi\in\Phi(\mathbb{R}^n)\cap C(\mathbb{R}^n) \quad \text{such that}\ \ \varphi(0)=1 \quad \text{and}\ \ \operatorname{supp}\varphi\subset\overline{\mathbb{B}_r}.} \end{equation*} \notag $$
It is obvious that the class $\mathfrak{F}_r(\mathbb{R}^n)$ is nonempty. For example, if $u\in L_2(\mathbb{R}^n)$ is a function such that $u(x)=0$ for $|x|\geqslant r/2$ and $\|u\|_2=1$, then the following function belongs to $\mathfrak{F}_r(\mathbb{R}^n)$:
$$ \begin{equation} \varphi(x)=(u\ast\widetilde{u})(x)=\int_{\mathbb{R}^n}u(x-t)\widetilde{u}(t)\,dt, \qquad x\in\mathbb{R}^n. \end{equation} \tag{1.2} $$
In fact, we have $\varphi\in C(\mathbb{R}^n)$ and $ \operatorname{supp}\varphi\subset\overline{\mathbb{B}_{r/2}}+\overline{\mathbb{B}_{r/2}} =\overline{\mathbb{B}_r}$. That $\varphi$ is positive definite is straightforward. Note that for $n=1$ it follows from the Boas–Kac–Krein theorem (for instance, see [1], Theorem 3.10.2) that each function $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$ can be represented in the form (1.2). Generally speaking, this is not true for $n\geqslant2$.

In this paper we consider the following extremal problem for positive definite functions $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$.

Problem. Let $r>0$, and let $\rho\in L_\infty^{\mathrm{loc}}(\mathbb{R}^n)$ be a radial real function. Find the quantity

$$ \begin{equation*} M(n,\rho,r):=\sup\biggl\{\biggl|\int_{\mathbb{R}^n}\varphi(x)\rho(x)\,dx\biggr|\colon \varphi\in\mathfrak{F}_r(\mathbb{R}^n)\biggr\}. \end{equation*} \notag $$

For $\rho(x)\equiv1$ the quantity $M(n,\rho,r)$ was found by Siegel [2] in 1935 and, independently, by Boas and Kac (see [3], Theorem 5) in 1945 for $n=1$. They showed that

$$ \begin{equation*} M(n,\rho,r)=\operatorname{vol}(\mathbb{B}_{r/2})=\frac{\pi^{n/2}r^n}{2^n\Gamma(n/2+1)}, \end{equation*} \notag $$
where $\operatorname{vol}(\,\cdot\,)$ is the Lebesgue measure in $\mathbb{R}^n$. In this case the extremal function is the convolution of the characteristic function of the ball $\mathbb{B}_{r/2}$ with itself:
$$ \begin{equation*} \varphi(x)=\frac1{\operatorname{vol}(\mathbb{B}_{r/2})}(\chi_{\mathbb{B}_{r/2}}\ast\chi_{\mathbb{B}_{r/2}})(x),\qquad x\in\mathbb{R}^n. \end{equation*} \notag $$

Siegel’s result was also re-discovered by Gorbachev [4] in 2001, who used other methods. It can also be mentioned that another method of the proof of Siegel’s results can be found in the recent paper [5].

Note that for $\rho(x)\equiv1$ the problem under consideration is an extremal problem of Turán type. In this class of problems one must find the supremum of the values of the integral $\displaystyle\int_{\mathbb{R}^n}\varphi(x)\,dx$ over all functions $\varphi\in\Phi(\mathbb{R}^n)\cap C(\mathbb{R}^n)$ with fixed value at the origin whose support lies in a fixed centrally symmetric convex body. To this date, apart from the ball case, solutions of the Turán problem are only known for polytopes tiling the space (see Arestov and Berdysheva [6]) and for spectral bodies (see Kolountzakis and Révész [7]).

It is also worth pointing out that similar problems arise in a natural way in various fields of mathematics, for instance, convex analysis (see [8]). Applications to function theory can be found in [9]. Also note Efimov’s paper [10], where he considered a version of Turán problem for a ball. For further information about the history, versions and applications of this type of problem the reader can consult Révész’s paper [11].

For $n=1$ and weaker assumptions about the function $\rho$ an analogue of the problem in question was considered by this author in [12]. Here we establish the following theorem, which solves the problem for $n\neq2$.

Theorem 1. Let $n\neq2$, $r>0$, and let $\rho\in L_\infty^{\mathrm{loc}}(\mathbb{R}^n)$ be a radial real function. Let $A_\rho\colon L_2(\mathbb{B}_{r/2})\to L_2(\mathbb{B}_{r/2})$ be the operator defined by

$$ \begin{equation*} (A_\rho u)(t):=\int_{\mathbb{B}_{r/2}}\rho(t-x)u(x)\,dx, \qquad u(x)\in L_2(\mathbb{B}_{r/2}). \end{equation*} \notag $$

Then $A_\rho$ is a compact selfadjoint operator in $L_2(\mathbb{B}_{r/2})$ and

$$ \begin{equation*} M(n,\rho,r)=\|A_\rho\|, \end{equation*} \notag $$
where $\|A_\rho\|$ is the norm of $A_\rho$ in $L_2(\mathbb{B}_{r/2})$.

Remark 1. It follows from Theorem 1 that the solution of the problem under consideration reduces to finding the eigenvalue of $A_\rho$ that is greatest in modulus. Note also that Theorem 1 is an analogue of Szász’s theorem for nonnegative trigonometric polynomials (see [13], Theorem IV).

Remark 2. The proof of Theorem 1 is based on the Rudin–Efimov representation (see Theorem 5 below) of functions $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$. In general, it has been established for $n\neq2$. On the other hand, if $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$ is infinitely smooth, then this representation also holds for $n=2$.

In some cases, in the above problem it is sufficient to consider infinitely smooth functions from $\mathfrak{F}_r(\mathbb{R}^n)$. For instance, this is true when $\rho$ is continuous. In this case Theorem 1 also holds for $n=2$.

If $\rho(x)$ is a polynomial, then our problem is connected with sharp bounds for entire functions of exponential spherical type $\leqslant r$. Recall that an entire function $f\colon \mathbb{C}^n\to\mathbb{C}$ is said to have an exponential spherical type $\leqslant r$ if for each $\varepsilon>0$ there exists a positive constant $A_\varepsilon$ such that

$$ \begin{equation*} |f(z)|\leqslant A_\varepsilon e^{(r+\varepsilon)|z|}, \qquad z\in\mathbb{C}^n, \quad\text{where } |z|=\biggl(\sum_{k=1}^n|z_k|^2\biggr)^{1/2}. \end{equation*} \notag $$
We let $W_{p,r}(\mathbb{R}^n)$ denote the set of entire functions of exponential spherical type $\leqslant r$ such that their restrictions to $\mathbb{R}^n$ belong to $L_p(\mathbb{R}^n)$, $p\geqslant1$, and let $W_{p,r}^{+}(\mathbb{R}^n)$ denote the subset of functions in $W_{p,r}(\mathbb{R}^n)$ that are nonnegative on $\mathbb{R}^n$. Then the following result holds.

Theorem 2. Let $n,m\in\mathbb{N}$, $r>0$, let $\Delta$ be the Laplace operator, $L$ be a linear differential operator of the form

$$ \begin{equation*} L:=\sum_{k=0}^m a_k\Delta^k, \qquad \textit{where } a_k\in\mathbb{R}, \end{equation*} \notag $$
and let
$$ \begin{equation*} \rho(x):=\sum_{k=0}^m (-1)^ka_k |x|^{2k}, \qquad x\in\mathbb{R}^n. \end{equation*} \notag $$

Then the following sharp inequality holds for $f\in W_{1,r}^{+}(\mathbb{R}^n)$:

$$ \begin{equation*} \|L f\|_\infty\leqslant\frac{M(n,\rho,r)}{(2\pi)^n}\|f\|_1. \end{equation*} \notag $$

The paper is organized as follows. In § 2 we present some auxiliary facts and statements. In §§ 3 and 4 we prove Theorems 1 and 2, respectively. In § 5 we present for example the solution of the problem under consideration for $\rho(x)\equiv1$. Also, in § 5 we obtain solutions for $\rho(x)=|x|^2$ and $n\neq2$, and for $\rho(x)=x^{2m}$ and $n=1$, $m\in\mathbb{N}$. As a consequence, we deduce sharp inequalities of Bernstein–Nikol’skii type for functions in $W_{1,r}^{+}(\mathbb{R}^n)$.

§ 2. Auxiliary facts and statements

We point out the following properties of functions in $\Phi(\mathbb{R}^n)$. Let $f,f_{i}\in\Phi(\mathbb{R}^n)$. Then:

(1) $|f(x+y)-f(x)|^2\leqslant 2f(0)(f(0)-\operatorname{Re}f(y))$, $x,y\in \mathbb{R}^n$;

(2) $\lambda_{1}f_{1}+\lambda_{2}f_{2}$, $\overline{f}$, $\operatorname{Re}f$, $f_1f_2\in\Phi(\mathbb{R}^n)$ for $\lambda_i\geqslant 0$;

(3) if a finite limit $\lim_{n\to\infty} f_{n}(x)=:g(x)$ exists for all $x\in \mathbb{R}$, then $g\in\Phi(\mathbb{R}^n)$.

Properties (1)–(3) are well known (for instance, see [14], [1] and [15]).

In 1932 Bochner and, independently, Khintchine proved the following criterion for a function to be positive definite.

Theorem 3 (Bochner–Khintchine theorem). A function $f$ belongs to $\Phi(\mathbb{R}^n)\cap C(\mathbb{R}^n)$ if and only if there exists a finite nonnegative Borel measure $\mu$ on $\mathbb{R}^n$ such that

$$ \begin{equation*} f(x)=\int_{\mathbb{R}^n}e^{i(x,t)}\,d\mu(t), \qquad x\in\mathbb{R}^n. \end{equation*} \notag $$

The proof can be found, for instance, in [14], [1] or [15]. As a direct consequence (for instance, see [1], Theorem 1.8.7), we obtain the following criterion of positive definiteness in terms of the Fourier transform.

Theorem 4. If $f\in C(\mathbb{R}^n)\cap L_1(\mathbb{R}^n)$, then

$$ \begin{equation*} f\in\Phi(\mathbb{R}^n) \quad\Longleftrightarrow \quad \widehat f(t)\geqslant 0, \qquad t\in\mathbb{R}^n, \end{equation*} \notag $$
where
$$ \begin{equation*} \widehat f(t):=\frac1{(2\pi)^n}\int_{\mathbb{R}^n}f(x)e^{-i(t,x)}\,dx, \qquad t\in\mathbb{R}^n, \end{equation*} \notag $$
and in this case $\widehat f\in L_1(\mathbb{R}^n)$.

The following result is important for the proof of Theorem 1.

Theorem 5 (Rudin and Efimov). Let $n\neq2$, $r>0$, and let $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$ be a radial function. Then $\varphi$ can be represented by a uniformly convergent series:

$$ \begin{equation} \varphi(x)=\sum_{k=1}^\infty (u_k\ast\widetilde{u_k})(x), \qquad x\in\mathbb{R}^n, \end{equation} \tag{2.1} $$
where $u_k\in L_2(\mathbb{R}^n)$ and $u_k(x)=0$ as $|x|\geqslant r/2$.

Remark 3. For $n=1$ Theorem 5 is a corollary to the Boas–Kac–Krein theorem. For ${n\in\mathbb{N}}$ Theorem 5 was proved by Rudin under the assumption that $\varphi$ is infinitely differentiable (see [16] and [1], Theorem 3.10.4). Ehm, Gneiting and Richards [17] noted without proof that in Rudin’s theorem it is sufficient to assume merely that $\varphi$ is continuous. In [18] Efimov proved Theorem 5 for $n\geqslant3$.

Remark 4. Representation (2.1) for radial infinitely smooth functions in $\mathfrak{F}_r(\mathbb{R}^n)$ was used by Rudin in [16] to show that each positive definite radial function in a ball in $\mathbb{R}^n$ can be extended to a positive definite function on the whole space. This generalizes a theorem of M. Krein on the extension of positive definite functions from an interval to the whole line $\mathbb{R}$ (see [19]).

Problems of the extension of positive definite radial functions to spaces of higher dimension were considered in [20].

Since a positive definite function is Hermitian, in the problem under consideration we can assume that functions $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$ are even. It follows from the lemma below that they can also be assumed to be radial.

Lemma. Let $n\neq1$ and $r>0$, let $\rho\in L_\infty^{\mathrm{loc}}(\mathbb{R}^n)$ be a radial function, let $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$, and let $\varphi_{\mathrm{rad}}$ be the radialization of $\varphi$:

$$ \begin{equation*} \varphi_{\mathrm{rad}}(x):=\int_{\operatorname{SO}(n)}\varphi(\tau x)\,d\tau, \qquad x\in\mathbb{R}^n, \end{equation*} \notag $$
where $\operatorname{SO}(n)$ is the special orthogonal group and $d\tau$ is the normalized Haar measure on $\operatorname{SO}(n)$. Then $\varphi_{\mathrm{rad}}\in\mathfrak{F}_r(\mathbb{R}^n)$ and
$$ \begin{equation} \int_{\mathbb{R}^n}\varphi_{\mathrm{rad}}(x)\rho(x)\,dx =\int_{\mathbb{R}^n}\varphi(x)\rho(x)\,dx. \end{equation} \tag{2.2} $$

Proof. The fact that $\varphi_{\mathrm{rad}}$ belongs to $\mathfrak{F}_r(\mathbb{R}^n)$ is verified directly. We prove (2.2):
$$ \begin{equation*} \begin{aligned} \, \int_{\mathbb{R}^n}\varphi_{\mathrm{rad}}(x)\rho(x)\,dx &=\int_{\mathbb{R}^n}\biggl(\int_{\operatorname{SO}(n)}\varphi(\tau x)\,d\tau\biggr)\rho(x)\,dx \\ &=\int_{\operatorname{SO}(n)}\biggl(\int_{\mathbb{R}^n}\varphi(\tau x)\rho(x)\,dx\biggr)d\tau \\ &=\int_{\operatorname{SO}(n)}\biggl(\int_{\mathbb{R}^n}\varphi(x)\rho(\tau^{-1} x)\,dx\biggr)d\tau \\ &=\int_{\operatorname{SO}(n)}\biggl(\int_{\mathbb{R}^n}\varphi(x)\rho(x)\,dx\biggr)d\tau \\ &=\int_{\operatorname{SO}(n)}d\tau\cdot\int_{\mathbb{R}^n}\varphi(x)\rho(x)\,dx =\int_{\mathbb{R}^n}\varphi(x)\rho(x)\,dx. \end{aligned} \end{equation*} \notag $$

The lemma is proved.

§ 3. Proof of Theorem 1

Step 1. The operator $A_\rho$ is bounded and compact in $L_2(\mathbb{B}_{r/2})$ because ${\rho\in L_\infty^{\mathrm{loc}}(\mathbb{R}^n)}$. It is selfadjoint because $\rho$ is real and radial.

For convenience, below we identify $L_2(\mathbb{B}_{r/2})$ with the subspace of functions ${u\in L_2(\mathbb{R}^n)}$ such that $u(x)=0$ almost everywhere for $|x|\geqslant r/2$.

Let $u\in L_2(\mathbb{R}^n)$ and $u(x)=0$ for $|x|\geqslant r/2$. Then

$$ \begin{equation*} \int_{\mathbb{R}^n}(u\ast\widetilde{u})(x)\rho(x)\,dx=(A_\rho u,u), \end{equation*} \notag $$
where $(\,\cdot\,{,}\,\cdot\,)$ is the inner product in $L_2(\mathbb{B}_{r/2})$. In fact,
$$ \begin{equation*} \begin{aligned} \, \int_{\mathbb{R}^n}(u\ast\widetilde{u})(x)\rho(x)\,dx &=\int_{\mathbb{R}^n}\biggl(\int_{\mathbb{R}^n}u(x-t)\widetilde{u}(t)\,dt\biggr)\rho(x)\,dx \\ &=\int_{\mathbb{R}^n}\biggl(\int_{\mathbb{R}^n}u(x+t)\overline{u(t)}\,dt\biggr)\rho(x)\,dx \\ &=\int_{\mathbb{R}^n}\biggl(\int_{\mathbb{R}^n}u(x+t)\rho(x)\,dx\biggr)\overline{u(t)}\,dt \\ &=\int_{\mathbb{R}^n}\biggl(\int_{\mathbb{R}^n}\rho(x-t)u(x)\,dx\biggr)\overline{u(t)}\,dt \\ &=\int_{\mathbb{B}_{r/2}}\biggl(\int_{\mathbb{B}_{r/2}}\rho(t-x)u(x)\,dx\biggr)\overline{u(t)}\,dt =(A_\rho u,u). \end{aligned} \end{equation*} \notag $$

Step 2. We show that $M(n,\rho,r)\leqslant \|A_\rho\|$. It follows from the above lemma that we can assume in the problem under consideration that the functions $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$ are radial.

Let $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$, and let $\varphi$ be a radial function. It follows from Theorem 5 that

$$ \begin{equation*} \varphi(x)=\sum_{k=1}^\infty (u_k\ast\widetilde{u_k})(x), \qquad x\in\mathbb{R}^n, \end{equation*} \notag $$
where the series is uniformly convergent, $u_k\in L_2(\mathbb{R}^n)$ and $u_k(x)=0$ for $|x|\geqslant r/2$. In addition, since $\varphi(0)=1$, we have $\sum_{k=1}^\infty\|u_k\|_2^2=1$ and therefore
$$ \begin{equation*} \begin{aligned} \, \biggl|\int_{\mathbb{R}^n}\varphi(x)\rho(x)\,dx\biggr| &=\biggl|\int_{\mathbb{R}^n}\sum_{k=1}^\infty (u_k\ast\widetilde{u_k})(x)\rho(x)\,dx\biggr| =\biggl|\sum_{k=1}^\infty\int_{\mathbb{R}^n} (u_k\ast\widetilde{u_k})(x)\rho(x)\,dx\biggr| \\ &=\biggl|\sum_{k=1}^\infty(A_\rho u_k,u_k)\biggr| \leqslant \sum_{k=1}^\infty|(A_\rho u_k,u_k)| \leqslant \|A_\rho\|\sum_{k=1}^\infty|(u_k,u_k)| \\ &=\|A_\rho\|\sum_{k=1}^\infty\|u_k\|_2^2=\|A_\rho\|. \end{aligned} \end{equation*} \notag $$
Hence
$$ \begin{equation*} M(n,\rho,r)\leqslant\|A_\rho\|. \end{equation*} \notag $$

Step 3. We show that $M(n,\rho,r)\geqslant \|A_\rho\|$. Let $u\in L_2(\mathbb{R}^n)$, $u(x)=0$ for $|x|\geqslant r/2$ and $\|u\|_2=1$. Then $u\ast\widetilde{u}\in\mathfrak{F}_r(\mathbb{R}^n)$ and we have

$$ \begin{equation*} |(A_\rho u,u)| =\biggl|\int_{\mathbb{R}^n}(u\ast\widetilde{u})(x)\rho(x)\,dx\biggr|\leqslant M(n,\rho,r). \end{equation*} \notag $$
As $A_\rho$ is selfadjoint, it follows that
$$ \begin{equation*} \|A_\rho\|=\sup_{\|u\|_2=1}|(A_\rho u,u)|\leqslant M(n,\rho,r). \end{equation*} \notag $$

The proof is complete.

§ 4. Proof of Theorem 2

First we note that by multiplying each function $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$ by $e^{i(t,\cdot)}$, $t\in\mathbb{R}^n$, we map $\mathfrak{F}_r(\mathbb{R}^n)$ to itself bijectively, so that

$$ \begin{equation} M(n,\rho,r)=\sup\biggl\{\biggl|\int_{\mathbb{R}^n}\varphi(x)e^{i(t,x)}\rho(x)\,dx\biggr|\colon \varphi\in\mathfrak{F}_r(\mathbb{R}^n)\biggr\} \quad\text{for all } t\in\mathbb{R}^n. \end{equation} \tag{4.1} $$

Let $f\in W_{1,r}^{+}(\mathbb{R}^n)$ and $f(x)\not\equiv0$. Since $L$ is a linear operator, we can assume without loss of generality that $\|f\|_1=(2\pi)^n$.

By the Paley–Wiener theorem (for instance, see [15], § 3.4.9, or [21], § 3.2.6) and Theorem 4 the Fourier transform establishes a bijection between $W_{1,r}^{+}(\mathbb{R}^n)$ and the set of functions $\varphi\in\Phi(\mathbb{R}^n)\cap C(\mathbb{R}^n)$ such that $\operatorname{supp}\varphi\subset\overline{\mathbb{B}_{r}}$. Thus,

$$ \begin{equation} f(t)=\int_{\mathbb{R}^n}\varphi(x)e^{i(t,x)}\,dx, \qquad t\in\mathbb{R}^n, \end{equation} \tag{4.2} $$
where $\varphi\in\Phi(\mathbb{R}^n)\cap C(\mathbb{R}^n)$ and $ \operatorname{supp}\varphi\subset\overline{\mathbb{B}_{r}}$. By the formula for the inverse Fourier transform
$$ \begin{equation*} \varphi(x)=\frac1{(2\pi)^n}\int_{\mathbb{R}^n}f(t)e^{-i(t,x)}\,dt, \qquad x\in\mathbb{R}^n, \end{equation*} \notag $$
so that $\varphi(0)=1$.

Thus, $\varphi\in\mathfrak{F}_r(\mathbb{R}^n)$. Applying the differential operator $L$ to both sides of (4.2) we obtain

$$ \begin{equation*} Lf(t)=\int_{\mathbb{R}^n} \varphi(x)\rho(x)e^{i(t,x)}\,dx, \qquad t\in\mathbb{R}^n. \end{equation*} \notag $$
It follows from (4.1) that
$$ \begin{equation*} \|Lf\|_\infty\leqslant M(n,\rho,r). \end{equation*} \notag $$

Theorem 2 is proved.

§ 5. Some examples

5.1. Example 1

Let $n\neq2$, $r>0$ and $\rho(x)\equiv1$. Then the operator $A_\rho$ is finite-dimensional and has the form

$$ \begin{equation*} (A_\rho u)(t)=\int_{\mathbb{B}_{r/2}}u(x)\,dx, \qquad t\in\mathbb{B}_{r/2}. \end{equation*} \notag $$

We find all $\lambda\neq0$ for which the equation $A_\rho u=\lambda u$ has nontrivial solutions. Let ${u(t)=a}$ for $t\in\mathbb{B}_{r/2}$, where $a\in\mathbb{C}\setminus\{0\}$. Then $a\cdot\operatorname{vol}(\mathbb{B}_{r/2})=\lambda a$, and therefore $\lambda=\operatorname{vol}(\mathbb{B}_{r/2})$. Thus,

$$ \begin{equation*} M(n,\rho,r)=\operatorname{vol}(\mathbb{B}_{r/2})=\frac{\pi^{n/2}r^n}{2^n\Gamma(n/2+1)}. \end{equation*} \notag $$

5.2. Example 2

Let $n\neq2$, $r>0$ and $\rho(x)=|x|^2$, $x\in\mathbb{R}^n$. It is easy to verify that in this case

$$ \begin{equation} M(n,\rho,r_1)=\biggl(\frac{r_1}{r_2}\biggr)^{n+2}M(n,\rho,r_2), \qquad r_1,r_2>0. \end{equation} \tag{5.1} $$

Let $r=2$. Then $A_\rho$ is a finite-dimensional operator of the following form:

$$ \begin{equation*} \begin{aligned} \, &(A_\rho u)(t)=\int_{\mathbb{B}_1}|t-x|^2u(x)\,dx \\ &\ =\biggl(\sum_{k=1}^n t_k^2\biggr)\int_{\mathbb{B}_1}u(x)\,dx -2\sum_{k=1}^n t_k\int_{\mathbb{B}_1}x_k u(x)\,dx +\sum_{k=1}^n \int_{\mathbb{B}_1}x_k^2 u(x)\,dx, \qquad t\in\mathbb{B}_1. \end{aligned} \end{equation*} \notag $$
We find all $\lambda\neq0$ for which the equation
$$ \begin{equation} (A_\rho u)(t)=\lambda u(t), \qquad t\in\mathbb{B}_1, \end{equation} \tag{5.2} $$
has nontrivial solutions. Let $u(x)=a|x|^2+(b,x)+c$ for $x\in\mathbb{B}_1$, where $a,c\in\mathbb{C}$ and $b=(b_1,\dots,b_n)\in\mathbb{C}^n$. Here $(\,\cdot\,{,}\,\cdot\,)$ is the scalar product in $\mathbb{C}^n$. Then we have the equalities
$$ \begin{equation*} \int_{\mathbb{B}_1}u(x)\,dx=a\int_{\mathbb{B}_1}|x|^2\,dx+c\int_{\mathbb{B}_1}dx, \qquad \int_{\mathbb{B}_1}x_k u(x)\,dx=b_k\int_{\mathbb{B}_1}x_1^2\,dx \end{equation*} \notag $$
and
$$ \begin{equation*} \sum_{k=1}^n\int_{\mathbb{B}_1}x_k^2 u(x)\,dx=a\int_{\mathbb{B}_1}|x|^4\,dx+c\int_{\mathbb{B}_1}|x|^2\,dx. \end{equation*} \notag $$

Now for convenience we introduce the following notation:

$$ \begin{equation*} \begin{gathered} \, \xi:=\int_{\mathbb{B}_1}|x|^2\,dx=\frac{2\pi^{n/2}}{(n+2)\Gamma(n/2)}, \qquad \eta:=\int_{\mathbb{B}_1}dx=\frac{2\pi^{n/2}}{n\Gamma(n/2)}, \\ \tau:=\int_{\mathbb{B}_1}|x|^4\,dx=\frac{2\pi^{n/2}}{(n+4)\Gamma(n/2)}, \qquad \theta:=-2\int_{\mathbb{B}_1}x_1^2\,dx=\frac{-4\pi^{n/2}}{n(n+2)\Gamma(n/2)}. \end{gathered} \end{equation*} \notag $$

Equating the coefficients in (5.2) we obtain

$$ \begin{equation*} \begin{pmatrix} \xi & 0 & \dots & 0 & \eta \\ 0 & \theta & \dots & 0 & 0 \\ \vdots & \vdots & \ddots & \vdots & \vdots \\ 0 & 0 & \dots & \theta & 0 \\ \tau & 0 & \dots & 0 & \xi \end{pmatrix} \begin{pmatrix} a \\ b_1 \\ \vdots \\ b_n \\ c \end{pmatrix} =\lambda \begin{pmatrix} a \\ b_1 \\ \vdots \\ b_n \\ c \end{pmatrix}. \end{equation*} \notag $$
Solving the equation
$$ \begin{equation*} \begin{vmatrix} \xi-\lambda & 0 & \dots & 0 & \eta \\ 0 & \theta-\lambda & \dots & 0 & 0 \\ \vdots & \vdots & \ddots & \vdots & \vdots \\ 0 & 0 & \dots & \theta-\lambda & 0 \\ \tau & 0 & \dots & 0 & \xi-\lambda \end{vmatrix} =(\theta-\lambda)^n(\lambda^2-2\xi\lambda+\xi^2-\tau\eta)=0 \end{equation*} \notag $$
we obtain
$$ \begin{equation*} \begin{gathered} \, \lambda_1=\theta=\frac{-4\pi^{n/2}}{n(n+2)\Gamma(n/2)}, \\ \lambda_2=\xi-\sqrt{\tau\eta}=\frac{2\pi^{n/2}}{\Gamma(n/2)} \biggl(\frac1{n+2}-\frac{\sqrt{n(n+4)}}{n(n+4)}\biggr) \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \lambda_3=\xi+\sqrt{\tau\eta}=\frac{2\pi^{n/2}}{\Gamma(n/2)} \biggl(\frac1{n+2}+\frac{\sqrt{n(n+4)}}{n(n+4)}\biggr). \end{equation*} \notag $$
Clearly, $|\lambda_3|>|\lambda_2|$. For $n=1$ we have $|\lambda_3|-|\lambda_1|=(-10+6\sqrt{5})/15>0$. If $n\geqslant2$, then
$$ \begin{equation*} |\lambda_3|-|\lambda_1|=\frac{2\pi^{n/2}}{\Gamma(n/2)}\biggl(\frac{n-2}{n(n+2)}+\frac{\sqrt{n(n+4)}}{n(n+4)}\biggr)>0. \end{equation*} \notag $$
Thus, $M(n,\rho,2)=\lambda_3$ for $n\neq2$. One eigenfunction corresponding to $\lambda_3$ is, for example,
$$ \begin{equation} u(t)=|t|^2+\sqrt{\frac{\tau}{\eta}}=|t|^2+\sqrt{\frac{n}{n+4}}, \qquad t\in\mathbb{B}^n. \end{equation} \tag{5.3} $$

In the general case it follows from (5.1) that

$$ \begin{equation} M(n,\rho,r)=\frac{r^{n+2}\pi^{n/2}}{2^{n+1}\Gamma(n/2)} \biggl(\frac1{n+2}+\frac{\sqrt{n(n+4)}}{n(n+4)}\biggr), \qquad n\neq2, \quad r>0. \end{equation} \tag{5.4} $$

5.3. Example 3

Let $n=1$, $r>0$ and $\rho(x)=x^{2m}$, where $m\in\mathbb{N}$. It is easy to verify that then

$$ \begin{equation} M(n,\rho,r_1)=\biggl(\frac{r_1}{r_2}\biggr)^{2m+1}M(n,\rho,r_2), \qquad r_1,r_2>0. \end{equation} \tag{5.5} $$
Let $r=2$. Then $A_\rho$ is a finite-dimensional operator of the following form:
$$ \begin{equation*} (A_\rho u)(t)=\int_{-1}^1(t-x)^{2m}u(x)\,dx =\sum_{i=1}^{2m+1}(-1)^{i+1}C_{2m}^{i-1}t^{2m+1-i}s_{i-1}, \end{equation*} \notag $$
where
$$ \begin{equation*} s_k=\int_{-1}^1 x^ku(x)\,dx, \qquad k=0,\dots, 2m. \end{equation*} \notag $$
We find all $\lambda\neq0$ for which the equation
$$ \begin{equation} (A_\rho u)(t)=\lambda u(t), \qquad -1\leqslant t\leqslant1, \end{equation} \tag{5.6} $$
has nontrivial solutions. We seek $u(x)$ in the following form:
$$ \begin{equation*} u(x)=\sum_{p=0}^{2m} a_p x^p, \qquad a_p\in\mathbb{C}. \end{equation*} \notag $$
In this case
$$ \begin{equation*} s_k=\sum_{j=0}^{2m}\frac{1+(-1)^{k-j}}{2m+k-j+1}a_{2m-j}, \qquad k=0,\dots,2m. \end{equation*} \notag $$
Equating the coefficients in (5.6) we obtain the system $A_m\cdot a=\lambda a$, where $a:=(a_0,\dots,a_{2m})$ and
$$ \begin{equation} A_m:=\biggl(\frac{(-1)^{i+1}C_{2m}^{i-1}(1+(-1)^{i-j})}{2m+i-j+1}\biggr)_{i,j=1}^{2m+1}. \end{equation} \tag{5.7} $$
Thus,
$$ \begin{equation} M(1,\rho,r)=\biggl(\frac{r}{2}\biggr)^{2m+1}|\lambda_{\max}|, \end{equation} \tag{5.8} $$
where $\lambda_{\mathrm{max}}$ is the eigenvalue of the matrix (5.7) with the greatest modulus. In particular, for $m=1$ the matrix $A_m$ has the following form:
$$ \begin{equation*} A_1= \begin{pmatrix} \dfrac23 & 0 & 2 \\ 0 & -\dfrac43 & 0 \\ \dfrac25 & 0 & \dfrac23 \end{pmatrix}. \end{equation*} \notag $$
The eigenvalues of $A_1$ are
$$ \begin{equation*} \lambda_1=-\frac43, \quad \lambda_2=\frac{10-6\sqrt5}{15}\quad\text{and} \quad \lambda_3=\frac{10+6\sqrt5}{15}, \end{equation*} \notag $$
which corresponds to Example 2.

5.4. Remarks

Remark 5. It follows from Theorem 2 and Example 2 that for $n\neq2$, for each function $f\in W_{1,r}^{+}(\mathbb{R}^n)$ we have the sharp inequality

$$ \begin{equation} \|\Delta f\|_\infty\leqslant\frac{M(n,\rho,r)}{(2\pi)^n}\|f\|_1, \end{equation} \tag{5.9} $$
where $M(n,\rho,r)$ is defined by (5.4). Equality is attained on the Fourier transform of the function $u\ast\widetilde{u}$, where $u$ is defined by (5.3).

Remark 6. It follows from Theorem 2 and Example 3 that for $m\in\mathbb{N}$ and each $f\in W_{1,r}^{+}(\mathbb{R})$ we have the sharp inequality

$$ \begin{equation} \|f^{(2m)}\|_\infty\leqslant\frac{M(1,\rho,r)}{2\pi}\|f\|_1, \end{equation} \tag{5.10} $$
where $M(1,\rho,r)$ is defined by (5.8).

Remark 7. This author proved in [12] that the following sharp inequality holds for ${f\in W_{1,r}^{+}(\mathbb{R})}$:

$$ \begin{equation} -\frac{r^3}{12\pi}\|f\|_1\leqslant -f''(t)\leqslant\frac{r^3(5+3\sqrt{5})}{120\pi}\|f\|_1, \qquad t\in\mathbb{R}. \end{equation} \tag{5.11} $$
It is obvious that (5.11) yields inequality (5.9) for $n=1$ and inequality (5.10) for ${m=1}$.

Remark 8. Ibragimov proved in 1959 (see [22], Corollary 2) that for (not necessarily nonnegative) functions $f\in W_{1,r}(\mathbb{R})$ we have the inequalities

$$ \begin{equation} \|f^{(m)}\|_\infty\leqslant\frac{r^{m+1}}{\pi(m+1)}\|f\|_1, \qquad m\in\mathbb{N}. \end{equation} \tag{5.12} $$
For $m=1$ inequality (5.12) with constant $r^2/\pi$ was established by Korevaar [23] in 1949.

Inequalities of type (5.9), (5.10) and (5.12) belong to the class of Bernstein–Nikol’skii type inequalities. For more information about inequalities of this type, see [24] by Gorbachev.


Bibliography

1. Z. Sasvári, Multivariate characteristic and correlation functions, De Gruyter Stud. Math., 50, Walter de Gruyter & Co., Berlin, 2013, x+366 pp.  crossref  mathscinet  zmath
2. C. L. Siegel, “Über Gitterpunkte in convexen Körpern und ein damit zusammenhängendes Extremalproblem”, Acta Math., 65:1 (1935), 307–323  crossref  mathscinet  zmath
3. R. P. Boas, Jr., and M. Kac, “Inequalities for Fourier transforms of positive functions”, Duke Math. J., 12:1 (1945), 189–206  crossref  mathscinet  zmath
4. D. V. Gorbachev, “Extremum problem for periodic functions supported in a ball”, Math. Notes, 69:3 (2001), 313–319  mathnet  crossref  mathscinet  zmath
5. J.-P. Gabardo, “The Turán problem and its dual for positive definite functions supported on a ball in $\mathbb R^d$”, J. Fourier Anal. Appl., 30:1 (2024), 11, 31 pp.  crossref  mathscinet  zmath
6. A. A. Arestov and E. E. Berdysheva, “The Turán problem for a class of polytopes”, East J. Approx., 8:3 (2002), 381–388  mathscinet  zmath
7. M. Kolountzakis and S. G. Révész, “On a problem of Turán about positive definite functions”, Proc. Amer. Math. Soc., 131:11 (2003), 3423–3430  crossref  mathscinet  zmath
8. G. Bianchi and M. Kelly, “A Fourier analytic proof of the Blaschke–Santaló inequality”, Proc. Amer. Math. Soc., 143:11 (2015), 4901–4912  crossref  mathscinet  zmath
9. D. V. Gorbachev and S. Yu. Tikhonov, “Wiener's problem for positive definite functions”, Math. Z., 289:3–4 (2018), 859–874  crossref  mathscinet  zmath
10. A. V. Efimov, “A version of the Turán's problem for positive definite functions of several variables”, Proc. Steklov Inst. Math. (Suppl.), 277, suppl. 1 (2012), 93–112  mathnet  crossref  mathscinet  zmath
11. S. G. Révész, “Turán's extremal problem on locally compact abelian groups”, Anal. Math., 37:1 (2011), 15–50  crossref  mathscinet  zmath
12. A. D. Manov, “On an extremal problem for positive definite functions”, Chebyshevskii Sb., 22:5 (2021), 161–171 (Russian)  mathnet  crossref  mathscinet  zmath
13. O. Szász, “Über harmonische Funktionen und $L$-Formen”, Math. Z., 1:2–3 (1918), 149–162  crossref  mathscinet  zmath
14. Z. Sasvári, Positive definite and definitizable functions, Math. Top., 2, Akademie Verlag, Berlin, 1994, 208 pp.  mathscinet  zmath
15. R. M. Trigub and E. S. Bellinsky, Fourier analysis and approximation of functions, Kluwer Acad. Publ., Dordrecht, 2004, xiv+585 pp.  crossref  mathscinet  zmath
16. W. Rudin, “An extension theorem for positive-definite functions”, Duke Math. J., 37 (1970), 49–53  crossref  mathscinet  zmath
17. W. Ehm, T. Gneiting and D. Richards, “Convolution roots of radial positive definite functions with compact support”, Trans. Amer. Math. Soc., 356:11 (2004), 4655–4685  crossref  mathscinet  zmath
18. A. V. Efimov, “An analog of Rudin's theorem for continuous radial positive definite functions of several variables”, Proc. Steklov Inst. Math. (Suppl.), 284, suppl. 1 (2014), 79–86  mathnet  crossref  mathscinet  zmath
19. M. G. Krein, “The problem of extension of Hermitian-positive continuous functions”, Dokl. Akad. Nauk SSSR, 26:1 (1940), 17–22 (Russian)  mathscinet  zmath
20. L. Golinskii, M. Malamud and L. Oridoroga, “Radial positive definite functions and Schoenberg matrices with negative eigenvalues”, Trans. Amer. Math. Soc., 370:1 (2018), 1–25  crossref  mathscinet  zmath
21. S. M. Nikol'skiĭ, Approximation of functions of several variables and imbedding theorems, Grundlehren Math. Wiss., 205, Springer-Verlag, New York–Heidelberg, 1975, viii+418 pp.  crossref  mathscinet  zmath
22. I. I. Ibragimov, “Extremal problems in the class of entire functions of finite degree”, Izv. Akad. Nauk SSSR Ser. Mat., 23:2 (1959), 243–256 (Russian)  mathnet  mathscinet  zmath
23. J. Korevaar, “An inequality for entire functions of exponential type”, Nieuw Arch. Wiskunde (2), 23:2 (1949), 55–62  mathscinet  zmath
24. D. V. Gorbachev, “Sharp Bernstein–Nikol'skii inequalities for polynomials and entire functions of exponential type”, Chebyshevskii Sb., 22:5 (2021), 58–110  mathnet  crossref  mathscinet  zmath

Citation: A. D. Manov, “An extremal problem for positive definite functions with support in a ball”, Sb. Math., 215:7 (2024), 920–931
Citation in format AMSBIB
\Bibitem{Man24}
\by A.~D.~Manov
\paper An extremal problem for positive definite functions with support in a~ball
\jour Sb. Math.
\yr 2024
\vol 215
\issue 7
\pages 920--931
\mathnet{http://mi.mathnet.ru//eng/sm10006}
\crossref{https://doi.org/10.4213/sm10006e}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4813934}
\zmath{https://zbmath.org/?q=an:07945702}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024SbMat.215..920M}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001346292600004}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85208420043}
Linking options:
  • https://www.mathnet.ru/eng/sm10006
  • https://doi.org/10.4213/sm10006e
  • https://www.mathnet.ru/eng/sm/v215/i7/p61
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024