Update to v1.0:

Fixed: Missing factor of 1/2 in expansion of f.

  Fixed: mean is not necessary in Theorem 3.

  Added: Lemma on Central Limit Theorem with infinite Variance.

  Miscellaneous changes and fixes.
This commit is contained in:
Ian Jauslin 2020-11-24 23:43:37 -05:00
parent 9c9f888bb4
commit e8cb253fb2
2 changed files with 124 additions and 77 deletions

View File

@ -34,11 +34,11 @@ On the convolution inequality $f\geqslant f\star f$\par
\medskip \medskip
We consider the inequality $f \geqslant f\star f$ for real functions in $L^1(\mathbb R^d)$ where $f\star f$ denotes the convolution of $f$ with itself. We show that all such functions $f$ are non-negative, We consider the inequality $f \geqslant f\star f$ for real functions in $L^1(\mathbb R^d)$ where $f\star f$ denotes the convolution of $f$ with itself. We show that all such functions $f$ are non-negative,
which is not the case for the same inequality in $L^p$ for any $1 < p \leqslant 2$, for which the convolution is defined. We also show that all solutions in $L^1(\mathbb R^d)$ satisfy $\int_{\mathbb R^d}f(x){\rm d}x \leqslant \textstyle\frac12$. Moreover, if which is not the case for the same inequality in $L^p$ for any $1 < p \leqslant 2$, for which the convolution is defined. We also show that all solutions in $L^1(\mathbb R^d)$ satisfy $\int_{\mathbb R^d}f(x)dx \leqslant \textstyle\frac12$. Moreover, if
$\int_{\mathbb R^d}f(x){\rm d}x = \textstyle\frac12$, then $f$ must decay fairly slowly: $\int_{\mathbb R^d}|x| f(x){\rm d}x = \infty$, and this is sharp since for all $\int_{\mathbb R^d}f(x)dx = \textstyle\frac12$, then $f$ must decay fairly slowly: $\int_{\mathbb R^d}|x| f(x)dx = \infty$, and this is sharp since for all
$r< 1$, there are solutions with $\int_{\mathbb R^d}f(x){\rm d}x = \textstyle\frac12$ and $\int_{\mathbb R^d}|x|^r f(x){\rm d}x <\infty$. However, if $r< 1$, there are solutions with $\int_{\mathbb R^d}f(x)dx = \textstyle\frac12$ and $\int_{\mathbb R^d}|x|^r f(x)dx <\infty$. However, if
$\int_{\mathbb R^d}f(x){\rm d}x = : a < \textstyle\frac12$, the decay at infinity can be much more rapid: we show that for all $a<\textstyle\frac12$, there are solutions such that for some $\epsilon>0$, $\int_{\mathbb R^d}f(x)dx = : a < \textstyle\frac12$, the decay at infinity can be much more rapid: we show that for all $a<\textstyle\frac12$, there are solutions such that for some $\epsilon>0$,
$\int_{\mathbb R^d}e^{\epsilon|x|}f(x){\rm d}x < \infty$. $\int_{\mathbb R^d}e^{\epsilon|x|}f(x)dx < \infty$.
\vfill \vfill
\hfil{\footnotesize\copyright\, 2020 by the authors. This paper may be reproduced, in its entirety, for non-commercial purposes.} \hfil{\footnotesize\copyright\, 2020 by the authors. This paper may be reproduced, in its entirety, for non-commercial purposes.}
@ -55,71 +55,72 @@ $\int_{\mathbb R^d}e^{\epsilon|x|}f(x){\rm d}x < \infty$.
,\quad\forall x\in\mathbb R^d\ , ,\quad\forall x\in\mathbb R^d\ ,
\label{ineq} \label{ineq}
\end{equation} \end{equation}
where $f\star f(x)$ denotes the convolution $f\star f(x) = \int_{\mathbb R^d} f(x-y) f(y){\rm d}y$, which by Young's inequality \cite[Theorem 4.2]{LL96} is well defined as an element of where $f\star f(x)$ denotes the convolution $f\star f(x) = \int_{\mathbb R^d} f(x-y) f(y)dy$. By Young's inequality \cite[Theorem 4.2]{LL96}, for all $1\leqslant p\leqslant 2$ and all $f\in L^p(\mathbb R^d)$, $f\star f$ is well defined as an element of
$L^{p/(2-p)}(\mathbb R^d)$ for all $1 \leqslant p\leqslant 2$. Thus, one may consider this inequality in $L^p(\mathbb R^d)$ for all $1 \leqslant p \leqslant 2$, but the case $p=1$ is special: $L^{p/(2-p)}(\mathbb R^d)$. Thus, one may consider this inequality in $L^p(\mathbb R^d)$ for all $1 \leqslant p \leqslant 2$, but the case $p=1$ is special:
the solution set of (\ref{ineq}) is restricted in a number of surprising ways. Integrating both sides of (\ref{ineq}), one sees immediately that $\int_{\mathbb R^d} f(x){\rm d}x \leqslant 1$. the solution set of (\ref{ineq}) is restricted in a number of surprising ways. Integrating both sides of (\ref{ineq}), one sees immediately that $\int_{\mathbb R^d} f(x)dx \leqslant 1$.
We prove that, in fact, all integrable solutions satisfy $\int_{\mathbb R^d} f(x){\rm d}x \leqslant \textstyle\frac12$, and this upper bound is sharp. We prove that, in fact, all integrable solutions satisfy $\int_{\mathbb R^d} f(x)dx \leqslant \textstyle\frac12$, and this upper bound is sharp.
Perhaps even more surprising, we prove that all integrable solutions of (\ref{ineq}) are non-negative. This is {\em not true} for solutions in $L^p(\mathbb R^d)$, $ 1 < p \leqslant 2$. Perhaps even more surprising, we prove that all integrable solutions of (\ref{ineq}) are non-negative. This is {\em not true} for solutions in $L^p(\mathbb R^d)$, $ 1 < p \leqslant 2$.
For $f\in L^p(\mathbb R^d)$, $1\leqslant p \leqslant 2$, the Fourier transform $\widehat{f}(k) = \int_{\mathbb R^d}e^{-i2\pi k\cdot x} f(x){\rm d}x$ is well defined as an element of $L^{p/(p-1)}(\mathbb R^d)$. If $f$ solves the equation $f = f\star f$, For $f\in L^p(\mathbb R^d)$, $1\leqslant p \leqslant 2$, the Fourier transform $\widehat{f}(k) = \int_{\mathbb R^d}e^{-i2\pi k\cdot x} f(x)dx$ is well defined as an element of $L^{p/(p-1)}(\mathbb R^d)$. If $f$ solves the equation $f = f\star f$,
then $\widehat{f} = \widehat{f}^2$, and hence $\widehat{f}$ is the indicator function of a measurable set. By the Riemann-Lebesgue Theorem, if $f\in L^1(\mathbb R^d)$, then then $\widehat{f} = \widehat{f}^2$, and hence $\widehat{f}$ is the indicator function of a measurable set. By the Riemann-Lebesgue Theorem, if $f\in L^1(\mathbb R^d)$, then
$\widehat{f}$ is continuous and vanishes at infinity, and the only such indicator function is the indicator function of the empty set. Hence the only integrable solution of $\widehat{f}$ is continuous and vanishes at infinity, and the only such indicator function is the indicator function of the empty set. Hence the only integrable solution of
$f = f*f$ is the trivial solution $f= 0$. However, for $1 < p \leqslant 2$, solutions abound: take $d=1$ and define $g$ to be the indicator function of the interval $[-a,a]$. Define $f = f\star f$ is the trivial solution $f= 0$. However, for $1 < p \leqslant 2$, solutions abound: take $d=1$ and define $g$ to be the indicator function of the interval $[-a,a]$. Define
\begin{equation}\label{exam} \begin{equation}\label{exam}
f(x) = \int_{\mathbb R} e^{-i2\pi k x} g(k){\rm d}k = \frac{ \sin 2\pi xa}{\pi x}\ , f(x) = \int_{\mathbb R} e^{-i2\pi k x} g(k)dk = \frac{ \sin 2\pi xa}{\pi x}\ ,
\end{equation} \end{equation}
which is not integrable, but which belongs to $L^p(\mathbb R)$ for all $p> 1$. By the Fourier Inversion Theorem $\widehat{f} = g$. Taking products, one gets examples in any dimension. which is not integrable, but which belongs to $L^p(\mathbb R)$ for all $p> 1$. By the Fourier Inversion Theorem $\widehat{f} = g$. Taking products, one gets examples in any dimension.
To construct a family of solutions to (\ref{ineq}), fix $a,t> 0$, and define $g_{a,t}(k) = a e^{-2\pi |k| t}$. By \cite[Theorem 1.14]{SW71}, To construct a family of solutions to (\ref{ineq}), fix $a,t> 0$, and define $g_{a,t}(k) = a e^{-2\pi |k| t}$. By \cite[Theorem 1.14]{SW71},
$$ $$
f_{a,t}(x) = \int_{\mathbb R^d} e^{-i2\pi k x} g_{a,t}(k){\rm d}k = a\Gamma((d+1)/2) \pi^{-(d+1)/2} \frac{t}{(t^2 + x^2)^{(d+1)/2}}\ . f_{a,t}(x) = \int_{\mathbb R^d} e^{-i2\pi k x} g_{a,t}(k)dk = a\Gamma((d+1)/2) \pi^{-(d+1)/2} \frac{t}{(t^2 + x^2)^{(d+1)/2}}\ .
$$ $$
Since $g_{a,t}^2(k) = g_{a^2,2t}$, $f_{a,t}\star f_{a,t} = f_{a^2,2t}$, Thus, $f_{a,t} \geqslant f_{a,t}\star f_{a,t}$ reduces to Since $g_{a,t}^2(k) = g_{a^2,2t}$, $f_{a,t}\star f_{a,t} = f_{a^2,2t}$, Thus, $f_{a,t} \geqslant f_{a,t}\star f_{a,t}$ reduces to
$$ $$
\frac{t}{(t^2 + x^2)^{(d+1)/2}} \geqslant \frac{2at}{(4t^2 + x^2)^{(d+1)/2}} \frac{t}{(t^2 + x^2)^{(d+1)/2}} \geqslant \frac{2at}{(4t^2 + x^2)^{(d+1)/2}}
$$ $$
which is satisfied for all $a \leqslant 1/2$. Since $\int_{\mathbb R^d}f_{a,t}(x){\rm d}x =a$, this provides a class of solutions of (\ref{ineq}) that are non-negative and satisfy which is satisfied for all $a \leqslant 1/2$. Since $\int_{\mathbb R^d}f_{a,t}(x)dx =a$, this provides a class of solutions of (\ref{ineq}) that are non-negative and satisfy
\begin{equation}\label{half} \begin{equation}\label{half}
\int_{\mathbb R^d}f(x){\rm d}x \leqslant \frac12\ , \int_{\mathbb R^d}f(x)dx \leqslant \frac12\ ,
\end{equation} \end{equation}
all of which have fairly slow decay at infinity, so that in every case, all of which have fairly slow decay at infinity, so that in every case,
\begin{equation}\label{tail} \begin{equation}\label{tail}
\int_{\mathbb R^d}|x|f(x){\rm d}x =\infty \ . \int_{\mathbb R^d}|x|f(x)dx =\infty \ .
\end{equation} \end{equation}
Our results show that this class of examples of integrable solutions of (\ref{ineq}) is surprisingly typical of {\em all} integrable solutions: every real integrable solution Our results show that this class of examples of integrable solutions of (\ref{ineq}) is surprisingly typical of {\em all} integrable solutions: every real integrable solution
$f$ of (\ref{ineq}) is positive, satisfies (\ref{half}), $f$ of (\ref{ineq}) is positive, satisfies (\ref{half}),
and if there is equality in (\ref{half}), $f$ also satisfies (\ref{tail}). The positivity of all real solutions of (\ref{ineq}) in $L^1(\mathbb R^d)$ may be considered surprising since it is and if there is equality in (\ref{half}), $f$ also satisfies (\ref{tail}). The positivity of all real solutions of (\ref{ineq}) in $L^1(\mathbb R^d)$ may be considered surprising since it is
false in $L^p(\mathbb R^d)$ for all $p > 1$, as the example (\ref{exam}) shows. We also show that when strict inequality holds in (\ref{half}) for a solution $f$ of (\ref{ineq}), it is possible for false in $L^p(\mathbb R^d)$ for all $p > 1$, as the example (\ref{exam}) shows. We also show that when strict inequality holds in (\ref{half}) for a solution $f$ of (\ref{ineq}), it is possible for
$f$ to have rather fast decay; we construct examples such that $\int_{\mathbb R^d}e^{\epsilon|x|}f(x){\rm d}x < \infty$ for some $\epsilon> 0$. The conjecture that integrable solutions of (\ref{ineq}) $f$ to have rather fast decay; we construct examples such that $\int_{\mathbb R^d}e^{\epsilon|x|}f(x)dx < \infty$ for some $\epsilon> 0$. The conjecture that integrable solutions of (\ref{ineq})
are necessarily positive was motivated by recent work \cite{CJL19} on a partial differential equation involving a quadratic nonlinearity of $f\star f$ type, and the result proved here is the key to are necessarily positive was motivated by recent work \cite{CJL20,CJL20b} on a partial differential equation involving a quadratic nonlinearity of $f\star f$ type, and the result proved here is the key to
the proof of positivity for solutions of this partial differential equations; see \cite{CJL19}. the proof of positivity for solutions of this partial differential equations; see \cite{CJL20}. Autoconvolutions $f\star f$ have been studied extensively; see \cite{MV10} and the work quoted there. However, the questions investigated by these authors are quite different from those considered here.
\theo{Theorem}\label{theo:positivity} \theo{Theorem}\label{theo:positivity}
If $f$ is a real valued function in $L^1(\mathbb R^d)$ such that Let $f$ be a real valued function in $L^1(\mathbb R^d)$ such that
\begin{equation}\label{uineq} \begin{equation}\label{uineq}
f(x) - f\star f(x) =: u(x) \geqslant 0 f(x) - f\star f(x) =: u(x) \geqslant 0
\end{equation} \end{equation}
for all $x$. Then $\int_{\mathbb R^d} f(x)\ dx\leqslant\frac12$, for all $x$. Then $\int_{\mathbb R^d} f(x)\ dx\leqslant\frac12$,
and $f$ is given by the convergent series and $f$ is given by the series
\begin{equation} \begin{equation}
f(x) = \frac{1}{2} \sum_{n=1}^\infty c_n 4^n (\star^n u)(x) f(x) = \frac{1}{2} \sum_{n=1}^\infty c_n 4^n (\star^n u)(x)
\label{fun} \label{fun}
\end{equation} \end{equation}
which converges in $L^1(\mathbb R^d)$, and
where the $c_n\geqslant0$ are the Taylor coefficients in the expansion of $\sqrt{1-x}$ where the $c_n\geqslant0$ are the Taylor coefficients in the expansion of $\sqrt{1-x}$
\begin{equation}\label{3half} \begin{equation}\label{3half}
\sqrt{1-x}=1-\sum_{n=1}^\infty c_n x^n \sqrt{1-x}=1-\sum_{n=1}^\infty c_n x^n
,\quad ,\quad
c_n=\frac{(2n-3)!!}{2^nn!} \sim n^{-3/2} c_n=\frac{(2n-3)!!}{2^nn!} \sim n^{-3/2}
\end{equation} \end{equation}
In particular, $f$ is positive. Moreover, if $u\geqslant 0$ is any integrable function with $\int_{\mathbb R^d}u(x){\rm d}x \leqslant \textstyle\frac14$, then the sum on the right in (\ref{fun}) defines an integrable function $f$ that satisfies (\ref{uineq}). In particular, $f$ is positive. Moreover, if $u\geqslant 0$ is any integrable function with $\int_{\mathbb R^d}u(x)dx \leqslant \textstyle\frac14$, then the sum on the right in (\ref{fun}) defines an integrable function $f$ that satisfies (\ref{uineq}), and $\int_{\mathbb R^d}f(x) dx = \frac12$ if and only if $\int_{\mathbb R^d}u(x)dx= \frac14$.
\endtheo \endtheo
\bigskip \bigskip
\indent\underline{Proof}: Note that $u$ is integrable. Let $a := \int_{\mathbb R^d}f(x){\rm d}x$ and $b := \int_{\mathbb R^d}u(x){\rm d}x \geqslant 0$. Fourier transforming, \indent\underline{Proof}: Note that $u$ is integrable. Let $a := \int_{\mathbb R^d}f(x)dx$ and $b := \int_{\mathbb R^d}u(x)dx \geqslant 0$. Fourier transforming,
(\ref{uineq}) becomes (\ref{uineq}) becomes
\begin{equation} \label{ft} \begin{equation} \label{ft}
\widehat f(k) = \widehat f(k)^2 +\widehat u(k)\ . \widehat f(k) = \widehat f(k)^2 +\widehat u(k)\ .
@ -139,8 +140,8 @@ Hence (\ref{hatf}) is valid for all $k$, including $k=0$, again by continuity.
The fact that $c_n$ as specified in (\ref{3half}) satisfies $c_n \sim n^{-3/2}$ is a simple application of Stirling's formula, and it shows that the power series for $\sqrt{1-z}$ converges absolutely and The fact that $c_n$ as specified in (\ref{3half}) satisfies $c_n \sim n^{-3/2}$ is a simple application of Stirling's formula, and it shows that the power series for $\sqrt{1-z}$ converges absolutely and
uniformly everywhere on the closed unit disc. Since $|4 \widehat u(k)| \leqslant 1$, uniformly everywhere on the closed unit disc. Since $|4 \widehat u(k)| \leqslant 1$,
${\displaystyle ${\displaystyle
\sqrt{1-4 \widehat u(k)} = 1 -\sum_{n=1}^\infty c_n (4 \widehat u(k))^n}$. Inverting the Fourier transform, yields (\ref{fun}),and since $\int_{\mathbb R^d} 4^n\star^n u(x){\rm d}x \leqslant 1$, \sqrt{1-4 \widehat u(k)} = 1 -\sum_{n=1}^\infty c_n (4 \widehat u(k))^n}$. Inverting the Fourier transform, yields (\ref{fun}), and since $\int_{\mathbb R^d} 4^n\star^n u(x)dx \leqslant 1$,
the convergence of the sum in $L^1(\mathbb R^d)$ follows from the convergence of $\sum_{n=1}^\infty c_n$. The final statement follows from the fact that if $f$ is defined in terms of $u$ in this manner, (\ref{hatf}) is the convergence of the sum in $L^1(\mathbb R^d)$ follows from the convergence of $\sum_{n=1}^\infty c_n$. The final statement follows from the fact that if $f$ is defined in terms of $u$ in this manner, then (\ref{hatf}) is
valid, and then (\ref{ft}) and (\ref{uineq}) are satisfied. valid, and then (\ref{ft}) and (\ref{uineq}) are satisfied.
\qed \qed
\bigskip \bigskip
@ -152,76 +153,75 @@ $\int_{\mathbb R^d}|x| f(x)\ dx=\infty$.
\bigskip \bigskip
\indent\underline{Proof}: If $\int_{\mathbb R^d} f(x)\ dx=\textstyle\frac12$, $\int_{\mathbb R^d} 4u(x)\ dx=1$, then $w(x) = 4u(x)$ is a probability density, and we can write $f(x) = \sum_{n=0}^\infty \star^n w$. Suppose that $|x|f(x)$ is integrable. Then $|x|w(x)$ is integrable. Let $m:= \int_{\mathbb R^d}xw(x){\rm d} x$. Since first moments add under convolution, the trivial inequality $|m||x| \geqslant m\cdot x$ yields \indent\underline{Proof}: If $\int_{\mathbb R^d} f(x)\ dx=\textstyle\frac12$, $\int_{\mathbb R^d} 4u(x)\ dx=1$, then $w(x) = 4u(x)$ is a probability density, and we can write $f(x) = \frac12\sum_{n=0}^\infty \star^n w$. Aiming for a contradiction, suppose that $|x|f(x)$ is integrable. Then $|x|w(x)$ is integrable. Let $m:= \int_{\mathbb R^d}xw(x)d x$. Since first moments add under convolution, the trivial inequality $|m||x| \geqslant m\cdot x$ yields
$$|m|\int_{\mathbb R^d} |x| \star^nw(x){\rm d}x \geqslant \int_{\mathbb R^d} m\cdot x \star^nw(x){\rm d}x = n|m|^2\ .$$ $$|m|\int_{\mathbb R^d} |x| \star^nw(x)dx \geqslant \int_{\mathbb R^d} m\cdot x \star^nw(x)dx = n|m|^2\ .$$
It follows that $\int_{\mathbb R^d} |x| f(x){\rm d}x \geqslant |m|\sum_{n=1}^\infty nc_n = \infty$. Hence $m=0$. It follows that $\int_{\mathbb R^d} |x| f(x)dx \geqslant \frac{|m|}2\sum_{n=1}^\infty nc_n = \infty$. Hence $m=0$.
Suppose temporarily that in addition, $|x|^2w(x)$ is integrable. Let $\sigma^2$ be the variance of $w$; i.e., $\sigma^2 = \int_{\mathbb R^d}|x|^2w(x){\rm d}x\ .$ Suppose temporarily that in addition, $|x|^2w(x)$ is integrable. Let $\sigma^2$ be the variance of $w$; i.e., $\sigma^2 = \int_{\mathbb R^d}|x|^2w(x)dx$
Define the function $\varphi(x) = \min\{1,|x|\}$. Then Define the function $\varphi(x) = \min\{1,|x|\}$. Then
$$ $$
\int_{\mathbb R^d}|x| \star^n w(x){\rm d}x = \int_{\mathbb R^d}|n^{1/2}x| \star^n w(n^{1/2}x)n^{d/2}{\rm d}x \geqslant n^{1/2} \int_{\mathbb R^d}\varphi(x)\star^n w(n^{1/2}x)n^{d/2}{\rm d}x. \int_{\mathbb R^d}|x| \star^n w(x)dx = \int_{\mathbb R^d}|n^{1/2}x| \star^n w(n^{1/2}x)n^{d/2}dx \geqslant n^{1/2} \int_{\mathbb R^d}\varphi(x)\star^n w(n^{1/2}x)n^{d/2}dx.
$$ $$
By the Central Limit Theorem, since $\varphi$ is bounded and continuous, By the Central Limit Theorem, since $\varphi$ is bounded and continuous,
\begin{equation}\label{CLT} \begin{equation}\label{CLT}
\lim_{n\to\infty} \int_{\mathbb R^d}\varphi(x)\star^n w(n^{1/2}x)n^{d/2}{\rm d}x = \int_{\mathbb R^d}\varphi(x) \gamma(x){\rm d}x =: C > 0 \lim_{n\to\infty} \int_{\mathbb R^d}\varphi(x)\star^n w(n^{1/2}x)n^{d/2}dx = \int_{\mathbb R^d}\varphi(x) \gamma(x)dx =: C > 0
\end{equation} \end{equation}
where $\gamma(x)$ is a centered Gaussian probability density with variance $\sigma^2$. where $\gamma(x)$ is a centered Gaussian probability density with variance $\sigma^2$.
This shows that there is a $\delta> 0$ for all sufficiently large $n$, $\int_{\mathbb R^d}|x| \star^n w(x){\rm d}x \geqslant \sqrt{n}\delta$, and then since $c_n\sim n^{3/2}$, $\sum_{n=1}^\infty c_n \int_{\mathbb R^d}|x| \star^n w(x){\rm d}x= \infty$. This shows that there is a $\delta> 0$ such that for all sufficiently large $n$, $\int_{\mathbb R^d}|x| \star^n w(x)dx \geqslant \sqrt{n}\delta$, and then since $c_n\sim n^{3/2}$, $\sum_{n=1}^\infty c_n \int_{\mathbb R^d}|x| \star^n w(x)dx= \infty$.
To remove the hypothesis that $w$ has finite variance, note that if $w$ is a probability density with zero mean and infinite variance, $\star^n w(n^{1/2}x)n^{d/2}$ is ``trying'' to converge to a Gaussian of infinite variance. In particular, one would expect that for all $R>0$, To remove the hypothesis that $w$ has finite variance, note that if $w$ is a probability density with zero mean and infinite variance, $\star^n w(n^{1/2}x)n^{d/2}$ is ``trying'' to converge to a ``Gaussian of infinite variance''. In particular, one would expect that for all $R>0$,
\begin{equation}\label{CLT2} \begin{equation}\label{CLT2}
\lim_{n\to\infty} \int_{|x| \leqslant R}\star^n w(n^{1/2}x)n^{d/2}{\rm d}x = 0\ , \lim_{n\to\infty} \int_{|x| \leqslant R}\star^n w(n^{1/2}x)n^{d/2}dx = 0\ ,
\end{equation} so that the limit in (\ref{CLT}) has the value $1$. The proof then proceeds as above. The fact that (\ref{CLT2}) is valid is proved in \cite[Corollary 1]{CGR08}. \end{equation} so that the limit in (\ref{CLT}) has the value $1$. The proof then proceeds as above. The fact that (\ref{CLT2}) is valid is a consequence of Lemma~\ref{CLTL} below, which is closely based on the proof of \cite[Corollary 1]{CGR08}.
\qed \qed
\bigskip \bigskip
\delimtitle{\bf Remark} \theo{Theorem}\label{theo:decay3}
For the convenience of the reader, we sketch, at the end of this paper, the part of of the argument in \cite{CGR08} that proves (\ref{CLT2}), since we know of no reference for this simple statement, and the proof in \cite{CGR08} deals with a more complicated variant of this problem. Let $f\in L^1(\mathbb R^d)$ satisfy\-~(\ref{ineq}), $\int_{\mathbf R^d} xu(x) dx=0$, and
\endtheo $\int_{\mathbb R^d} |x|^2u(x)\ dx<\infty$, then, for all $0\leqslant p<1$,
\bigskip
\theo{Theorem}
If $f$ satisfies\-~(\ref{ineq}), and
$\int |x|^2[f(x)-f\star f(x)]\ dx<\infty$, then, for all $0\leqslant p<1$,
\begin{equation} \begin{equation}
\int |x|^pf(x)\ dx<\infty. \int_{\mathbb R^d} |x|^pf(x)\ dx<\infty.
\end{equation} \end{equation}
\endtheo \endtheo
\bigskip \bigskip
\indent\underline{Proof}: We may suppose that $f$ is not identically $0$. Let $u = f - f\star f$ as above. Let $t := 4\int_{\mathbb R^d}u(x){\rm d}x \leqslant 1$. Then $t> 0$. Define $w := t^{-1}4u$; $w$ is a probability density and \indent\underline{Proof}: We may suppose that $f$ is not identically $0$. Let $t := 4\int_{\mathbb R^d}u(x)dx \leqslant 1$. Then $t> 0$. Define $w := t^{-1}4u$; $w$ is a probability density and
\begin{equation}\label{tfor}
f(x) = \frac12\sum_{n=1}^\infty c_n t^n \star^n w(x)\ .
\end{equation}
By hypothesis, $w$ has a zero mean and variance $\sigma^2 = \int_{\mathbb R^d} |x|^2 w(x)dx < \infty$. Since variance is additive under convolution,
$$ $$
f(x) = \sum_{n=1}^\infty c_n t^n \star^n w(x)\ . \int_{\mathbb R^d} |x|^2 \star^n w(x)dx = n\sigma^2\ .
$$
By hypothesis, $w$ has a mean $m := \int_{\mathbb R^d} x w(x){\rm d}x$ and variance $\sigma^2 = \int_{\mathbb R^d} |x-m|^2 w(x){\rm d}x < \infty$. Since variance is additive under convolution,
$$
\int_{\mathbb R^d} |x-m|^2 \star^n w(x){\rm d}x = n\sigma^2\ .
$$ $$
By H\"older's inequality, for all $0 < p < 2$, By H\"older's inequality, for all $0 < p < 2$,
$\int_{\mathbb R^d} |x-m|^p \star^n w(x){\rm d}x \leqslant (n\sigma^2)^{p/2}$. $\int_{\mathbb R^d} |x|^p \star^n w(x)dx \leqslant (n\sigma^2)^{p/2}$.
It follows that for $0 < p < 1$, It follows that for $0 < p < 1$,
$$ $$
\int_{\mathbb R^d} |x-m|^p f(x){\rm d}x \leqslant (\sigma^2)^{p/2} \sum_{n=1}^\infty n^{p/2} c_n < \infty\ , \int_{\mathbb R^d} |x|^p f(x)dx \leqslant \frac12(\sigma^2)^{p/2} \sum_{n=1}^\infty n^{p/2} c_n < \infty\ ,
$$ $$
again using the fact that $c_n\sim n^{-3/2}$. again using the fact that $c_n\sim n^{-3/2}$.
\qed \qed
\delimtitle{\bf Remark}
In the subcritical case $\int_{\mathbb R^d}f(x)dx < \frac12$, the hypothesis that $\int_{\mathbb R^d} x u(x) dx = 0$ is superfluous, and one can conclude more. In this case the quantity $t$ in (\ref{tfor}) satisfies $0 < t < 1$, and if we let $m$ denote the mean of $w$,
$\int_{\mathbb R^d} |x|^2 \star^n w(x)dx =n^2|m|^2+ n\sigma^2$. For $0<t<1$, $\sum_{n=1}^\infty n^2 c_n t^n < \infty$ and we conclude that $\int_{\mathbb R^d} |x|^2 f(x)dx < \infty$. Finally, the final statement of Theorem~\ref{theo:positivity} shows that critical case functions $f$ satisfying the hypotheses of Theorem~\ref{theo:decay} are readily constructed.
\enddelim
Theorem \ref{theo:decay} implies that when $\int f=\frac12$, $f$ cannot decay faster than $|x|^{-(d+1)}$. However, integrable solutions $f$ of (\ref{ineq}) such that $\int_{\mathbb R^d}f(x){\rm d}x < \textstyle\frac12$ Theorem \ref{theo:decay} implies that when $\int f=\frac12$, $f$ cannot decay faster than $|x|^{-(d+1)}$. However, integrable solutions $f$ of (\ref{ineq}) such that $\int_{\mathbb R^d}f(x)dx < \textstyle\frac12$
can decay quite rapidly, even having finite exponential moments, as we now show. can decay more rapidly, as indicated in the previous remark. In fact, they may even have finite exponential moments, as we now show.
Consider a non-negative, integrable function $u$, which integrates to $r<\frac14$, and satisfies Consider a non-negative, integrable function $u$, which integrates to $r<\frac14$, and satisfies
\begin{equation} \begin{equation}
\int_{\mathbb R^d} u(x)e^{\lambda|x|}{\rm d}x < \infty \int_{\mathbb R^d} u(x)e^{\lambda|x|}dx < \infty
\end{equation} \end{equation}
for some $\lambda>0$. for some $\lambda>0$.
The Laplace transform of $u$ is The Laplace transform of $u$ is
$ \widetilde u(p):=\int e^{-px}u(x)\ {\rm d} x$ which is analytic for $|p|<\lambda$, and $\widetilde u(0) < \textstyle\frac14$. $ \widetilde u(p):=\int e^{-px}u(x)\ d x$ which is analytic for $|p|<\lambda$, and $\widetilde u(0) < \textstyle\frac14$.
Therefore, there exists $0<\lambda_0\leqslant \lambda$ such that, for all $|p|\leqslant\lambda_0$, Therefore, there exists $0<\lambda_0\leqslant \lambda$ such that, for all $|p|\leqslant\lambda_0$,
$\widetilde u(p)<\textstyle\frac14$. $\widetilde u(p)<\textstyle\frac14$.
By Theorem \ref{theo:positivity}, By Theorem \ref{theo:positivity},
@ -235,8 +235,10 @@ is an integrable solution of (\ref{ineq}). For
which is analytic for $|p|\leqslant \lambda_0$. which is analytic for $|p|\leqslant \lambda_0$.
Note that Note that
${\displaystyle e^{s|x|} \leqslant \prod_{j=1}^d e^{|sx_j|} \leqslant \frac{1}{d}\sum_{j=1}^d e^{d|sx_j|} \leqslant \frac{2}{d}\sum_{j=1}^d \cosh(dsx_j)}$. ${\displaystyle e^{s|x|} \leqslant \prod_{j=1}^d e^{|sx_j|} \leqslant \frac{1}{d}\sum_{j=1}^d e^{d|sx_j|} \leqslant \frac{2}{d}\sum_{j=1}^d \cosh(dsx_j)}$.
What has just been shown, yields a $\delta>0$ so that for $|s|< \delta$, $\int_{\mathbb R^d} \cosh(dsx_j)f(x){\rm d}x < \infty$ for each $j$. It follows that for Thus, for $|s|< \delta := \lambda_0/d$, $\int_{\mathbb R^d} \cosh(dsx_j)f(x)dx < \infty$ for each $j$, and hence
$0 < s < \delta$, $\int_{\mathbb R^d} e^{s|x|}f(x){\rm d}x < \infty$. $|s| < \delta$, $\int_{\mathbb R^d} e^{s|x|}f(x)dx < \infty$.
However, there are no integrable solutions of (\ref{ineq}) that have compact support: We have seen that all solutions of (\ref{ineq}) are non-negative, and if $A$ is the support of a non-negative integrable function, the Minkowski sum $A+A$ is the support of $f\star f$.
\bigskip \bigskip
@ -247,42 +249,77 @@ $f$ by $-1$. If the region is taken small enough, the new function $f$ will stil
\endtheo \endtheo
\bigskip \bigskip
We close by sketching a proof of (\ref{CLT2}) using the construction in \cite{CGR08}. Let $w$ be a mean zero, infinite variance probability density on $\mathbb R^d$. We close with a lemma validating (\ref{CLT2}) that is closely based on a construction in \cite{CGR08}.
Pick $\epsilon>0$, and choose a large value $\sigma$ such that $(2\pi \sigma^2)^{-d/2}R^d|B| < \epsilon/2$, where $|B|$ denotes the volume of the ball. The point of this is that if
$G$ is a centered Gaussian random variable with variance $\sigma^2$, the probability is no more than $\epsilon/2$ that $G$ lies in {\em any} particular translate $B_R+y$ of the ball of radius $R$. \theo{Lemma}\label{CLTL}
Let $w$ be a mean zero, infinite variance probability density on $\mathbb R^d$. Then for all $R>0$, (\ref{CLT2}) is valid.
\endtheo
\indent\underline{Proof}: Let $X_1,\dots,X_n$ be $n$ independent samples from the density $w$, and let $B_R$ denote the centered ball of radius $R$. The quantity in (\ref{CLT2}) is $p_{n,R} := \mathbb{P}(n^{-1/2}\sum_{j=1}^n X_j\in B_R)$.
Let $\widetilde X_1,\dots,\widetilde X_n$ be another $n$ independent samples from the density $w$, independent of the first $n$. Then also $p_{n,R} := \mathbb{P}(-n^{-1/2}\sum_{j=1}^n \widetilde X_j\in B_R)$. By the independence and the triangle inequality,
$$
p_{n,R}^2 \leqslant \mathbb{P}(n^{-1/2}\sum_{j=1}^n (X_j -\widetilde X_j)\in B_{2R})\ .
$$
The random variable $X_1 - \widetilde X_1$ has zero mean and infinite variance and an even density. Therefore, without loss of generality, we may assume that $w(x) = w(-x)$ for all $x$.
Pick $\epsilon>0$, and choose a large value $\sigma_0$ such that $(2\pi \sigma_0^2)^{-d/2}R^d|B| < \epsilon/3$, where $|B|$ denotes the volume of the unit ball $B$. The point of this is that if
$G$ is a centered Gaussian random variable with variance {\em at least} $\sigma_0^2$, the probability that $G$ lies in {\em any} particular translate $B_R+y$ of the ball of radius $R$ is no more than $\epsilon/3$. Let $A\subset \mathbb R^d$ be a centered cube such that
$$
\int_{A}|x|^2w(x)dx =: \sigma^2 \geqslant 2\sigma_0^2 \quad{\rm and}\quad \int_{A}w(x)dx > \frac34\ ,
$$
and note that since $A$ and $w$ are even, ${\displaystyle \int_{A} x w(x)dx = 0}$.
It is then easy to find mutually independent random variables $X$, $Y$ and $\alpha$ such that
$X$ takes values in $A$ and, has zero mean and variance $\sigma^2$, $\alpha$ is a Bernoulli variable with success probability $\int_{A}w(x)dx$, and finally such that $\alpha X + (1-\alpha)Y $ has the probability density $w$. Taking independent identically distributed (i.i.d.) sequences of such random variables, $w(n^{1/2}x)n^{d/2}$ is the probability density of
${\displaystyle W_n := n^{-1/2}\sum_{j=1}^n \alpha_j X_j + n^{-1/2} \sum_{j=1}^n(1-\alpha_j)Y_j}$, and we seek to estimate
the expectation of $1_{B_R}(W_n)$. We first take the conditional expectation, given the values of the $\alpha$'s and the $Y$'s, and we define $\hat{n} = \sum_{j=1}^n\alpha_j$. These conditional expectations have the form
${\mathbb E}\left[ 1_{B_R + y}\left(\sum_{j=1}^n n^{-1/2}\alpha_j X_j \right)\right]$
for some translate $B_R +y$ of $B_R$, the ball of radius $R$. The sum $n^{-1/2}\sum_{j=1}^n \alpha_j X_j$ is actually the sum of $\hat{n}$ i.i.d. random variables with mean zero and variance $\sigma^2/n$. The probability that $\hat{n}$ is significantly less than $\frac34 n$ is negligible for large $n$; by classical estimates associated with the Law of Large Numbers, for all $n$ large enough, the probability that $\hat{n} < n/2$ is no more than $\epsilon/3$. Now let $Z$ be a Gaussian random variable with mean zero and variance
$\sigma^2\hat{n}/n$ which is at least $\sigma^2_0$ when $\hat{n} \geqslant n/2$. Then by the multivariate version \cite{R19} of the Berry-Esseen Theorem \cite{B41,E42}, a version of the Central Limit Theorem
with rate information, there is a constant $K_d$ depending only on $d$ such that
$${\textstyle
\left|{\mathbb E}\left[ 1_{B_R + y}\left(\sum_{j=1}^n n^{-1/2}\alpha_j X_j \right)\right] - {\mathbb P}\{Z \in B_R + y\}\right| \leqslant K_d \hat{n} \frac{{\mathbb E}|X_1|^3}{n^{3/2}} \leqslant K_d \frac{{\mathbb E}|X_1|^3}{n^{1/2}}
\ .}
$$
Since $A$ is bounded, ${\mathbb E}|X_1|^3 < \infty$, and hence for all sufficiently large $n$, when $\hat{n} \geqslant n/2$.
$${\textstyle
{\mathbb E}\left[ 1_{B_R + y}\left(\sum_{j=1}^n n^{-1/2}\alpha_j X_j \right)\right] \leqslant \frac23 \epsilon
\ .}
$$
Since this is uniform in $y$, we finally obtain ${\mathbb P}(W_n \in B_R) \leqslant \epsilon$ for all sufficiently large $n$. Since $\epsilon>0$ is arbitrary, (\ref{CLT2}) is proved.
\qed
We close by thanking the anonymous referee for useful suggestions.
Let $A \subset \mathbb R^d$ be a bounded set so that $\int_{A}xw(x){\rm d}x = 0$ and $\int_{A}|x|^2w(x){\rm d}x = \sigma^2$. It is then easy to find mutually independent random variables $X$, $Y$ and $\alpha$ such that
$X$ takes values in $A$ and, has zero mean and variance greater than $\sigma^2$, $\alpha$ is a Bernoulli variable with success probability $\int_{A}w(x){\rm d}x$, which we can take arbitrarily close to $1$ by
increasing the size of $A$, and finally such that $\alpha X + (1-\alpha)Y $ has the probability density $w$. Taking independent i.i.d. sequences of such random variables, $w(n^{1/2}x)n^{d/2}$ is the probability density of
${\displaystyle W_n := \frac{1}{\sqrt{n}}\sum_{j=1}^n \alpha_j X_j + \frac{1}{\sqrt{n}} \sum_{j=1}^n(1-\alpha_j)Y_j}$.
We are interested in estimating the expectation of $1_{B_R}(W_n)$. We first take the conditional expectation, given the values of the $\alpha$'s and the $Y$'s. These conditional expectations have the form
${\mathbb E}\left[ 1_{B_R + y}\left( \frac{1}{\sqrt{n}}\sum_{j=1}^n \alpha_j X_j \right)\right]$
for some translate $B_R +y$ of the unit Ball. Since $A$ is bounded, the $X_j$'s all have the same finite third moment, and now the Berry-Esseen Theorem \cite{B41,E42}, a version of the Central Limit Theorem
with rate information, allows us to control the error in approximating this expectation by
$\mathbb{E}(1_{B_R +y}(G))$ where $G$ is centered Gaussian with variance $\sigma^2$. By the choice of $\sigma$, this is no greater than $\epsilon/2$, independent of $y$. For $n$ large enough,
the remaining errors -- coming from the small probability that $\sum_{n=1}^n \alpha_j$ is significantly less $n$, and the error bound provided by the Berry-Esseen Theorem, are readily absorbed into the remaining
$\epsilon/2$ for large $n$. Thus for all sufficiently large $n$, the integral in (\ref{CLT2}) is no more than $\epsilon$.
\vfill \vfill
\eject \eject
{\bf Acknowledgements}: {\bf Acknowledgements}:
U.S.~National Science Foundation grants DMS-1764254 (E.A.C.), DMS-1802170 (I.J.) and NSF grant DMS-1856645 (M.P.L) are gratefully acknowledged. U.S.~National Science Foundation grants DMS-1764254 (E.A.C.), DMS-1802170 (I.J.) and DMS-1856645 (M.P.L) are gratefully acknowledged.
\vskip20pt \vskip20pt
\begin{thebibliography}{WWW99} \begin{thebibliography}{WWW99}
\bibitem[B41]{B41} A. Berry, {\em The Accuracy of the Gaussian Approximation to the Sum of Independent Variates}. Trans. of the A.M.S. {\bf 49} (1941),122-136. \bibitem[B41]{B41} A. Berry, {\em The Accuracy of the Gaussian Approximation to the Sum of Independent Variates}. Trans. of the A.M.S. {\bf 49} (1941),122--136.
\bibitem[CGR08]{CGR08} E.A. Carlen, E. Gabetta and E. Regazzini, {\it Probabilistic investigation on explosion of solutions of the Kac equation with infintte initial energy}, J. Appl. Prob. {\bf 45} (2008), 95-106 \bibitem[CGR08]{CGR08} E.A. Carlen, E. Gabetta and E. Regazzini, {\it Probabilistic investigation on explosion of solutions of the Kac equation with infinite initial energy}, J. Appl. Prob. {\bf 45} (2008), 95--106
\bibitem[CJL19]{CJL19} E.A. Carlen, I. Jauslin and E.H. Lieb, {Analysis of a simple equation for the ground state energy of the Bose gas}, arXiv preprint arXiv:1912.04987. \bibitem[CJL20]{CJL20} E.A. Carlen, I. Jauslin and E.H. Lieb, {Analysis of a simple equation for the ground state energy of the Bose gas}, Pure and Applied Analysis, 2020, in press, arXiv preprint arXiv:1912.04987.
\bibitem[E42]{E42} C.-G. Esseen, {\em A moment inequality with an application to the central limit theorem}. Skand. Aktuarietidskr. {\bf 39} 160-170. \bibitem[CJL20b]{CJL20b} E.A. Carlen, I. Jauslin and E.H. Lieb, {Analysis of a simple equation for the ground state of the Bose gas II: Monotonicity, Convexity and Condensate Fraction}, arXiv preprint arXiv:2010.13882.
\bibitem[E42]{E42} C.-G. Esseen, {\em A moment inequality with an application to the central limit theorem}. Skand. Aktuarietidskr. {\bf 39} (1942) 160--170.
\bibitem[LL96]{LL96} E.H. Lieb and M. Loss, {\em Analysis}, Graduate Studies in Mathematics {\bf 14}, A.M.S., Providence RI, 1996. \bibitem[LL96]{LL96} E.H. Lieb and M. Loss, {\em Analysis}, Graduate Studies in Mathematics {\bf 14}, A.M.S., Providence RI, 1996.
\bibitem[MV10]{MV10} M. Matolcsi, C. Vinuesa, {\it Improved bounds on the supremum of autoconvolutions}, J. Math. Anal. Appl. {\bf 372} (2010) 439--447.
\bibitem[R19]{R19} M. Rai\v c, {\em A multivariate Berry-Esseen Theorem with explicit constants}, Bernoulli {\bf 25} (2019) 2824--2853
\bibitem[SW71]{SW71} E. Stein and G. Weiss, {\em Introduction to Fourier analysis on Euclidean spaces}, Princeton University Press, Princeton NJ, 1971. \bibitem[SW71]{SW71} E. Stein and G. Weiss, {\em Introduction to Fourier analysis on Euclidean spaces}, Princeton University Press, Princeton NJ, 1971.

10
Changelog Normal file
View File

@ -0,0 +1,10 @@
v1.0:
* Fixed: Missing factor of 1/2 in expansion of f.
* Fixed: mean is not necessary in Theorem 3.
* Added: Lemma on Central Limit Theorem with infinite Variance.
* Miscellaneous changes and fixes.