Ian Jauslin
summaryrefslogtreecommitdiff
blob: 34fe424240ceb2cacbd3902cf16aa7e4a3734ddd (plain)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
\documentclass[no_section_in_all]{ian}

\begin{document}

\hbox{}
\hfil{\bf\LARGE
On the convolution inequality $f\geqslant f\star f$\par
}
\vfill

\hfil{\bf\large Eric Carlen}\par
\hfil{\it Department of Mathematics, Rutgers University}\par
\hfil{\color{blue}\tt\href{mailto:carlen@rutgers.edu}{carlen@rutgers.edu}}\par
\vskip20pt

\hfil{\bf\large Ian Jauslin}\par
\hfil{\it Department of Physics, Princeton University}\par
\hfil{\color{blue}\tt\href{mailto:ijauslin@princeton.edu}{ijauslin@princeton.edu}}\par
\vskip20pt

\hfil{\bf\large Elliott H. Lieb}\par
\hfil{\it Departments of Mathematics and Physics, Princeton University}\par
\hfil{\color{blue}\tt\href{mailto:lieb@princeton.edu}{lieb@princeton.edu}}\par
\vskip20pt

\hfil{\bf\large Michael Loss}\par
\hfil{\it School of Mathematics, Georgia Institute of Technology}\par
\hfil{\color{blue}\tt\href{mailto:loss@math.gatech.edu}{loss@math.gatech.edu}}\par

\vfill


\hfil {\bf Abstract}\par
\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, 
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 
$\int_{\mathbb R^d}f(x){\rm d}x = \textstyle\frac12$, then $f$ must decay fairly slowly: $\int_{\mathbb R^d}|x| f(x){\rm d}x = \infty$, and this is sharp since for all 
$r< 1$, 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 
$\int_{\mathbb R^d}f(x){\rm d}x  = : a < \textstyle\frac12$, the decay at infinity can be much more rapid: we show that for all $a<\textstyle\frac12$, there are solutions such that for some $\epsilon>0$, 
$\int_{\mathbb R^d}e^{\epsilon|x|}f(x){\rm d}x < \infty$. 
\vfill

\hfil{\footnotesize\copyright\, 2020 by the authors. This paper may be reproduced, in its entirety, for non-commercial purposes.}

\eject

\setcounter{page}1
\pagestyle{plain}


\indent Our subject is  the set of real,  integrable solutions of the inequality 
\begin{equation}
  f(x)\geqslant f\star f(x)
  ,\quad\forall x\in\mathbb R^d\ ,
  \label{ineq}
\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 
$L^{p/(2-p)}(\mathbb R^d)$ 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:  
the solution set of (\ref{ineq}) is restricted in a number of surprising ways. Integrating both sides of (\ref{ineq}), one sees immediately that $\int_{\mathbb R^d} f(x){\rm d}x \leqslant 1$. 
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.  


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 $f = f\star f$,
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 
$f = f*f$ 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 
\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}\ ,
 \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.  


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},
$$
f_{a,t}(x) =  \int_{\mathbb R^d} e^{-i2\pi k x} g_{a,t}(k){\rm d}k = a\Gamma((d+1)/2) \pi^{-(d+1)/2} \frac{t}{(t^2 + x^2)^{(d+1)/2}}\ .
$$
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 
$$
\frac{t}{(t^2 + x^2)^{(d+1)/2}} \geqslant  \frac{2at}{(4t^2 + x^2)^{(d+1)/2}}
$$
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 
\begin{equation}\label{half}
\int_{\mathbb R^d}f(x){\rm d}x \leqslant \frac12\ ,
\end{equation}
 all of which have fairly slow decay at infinity, so that in every case, 
\begin{equation}\label{tail}
 \int_{\mathbb R^d}|x|f(x){\rm d}x =\infty \ .
 \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 
$f$  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 
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})  
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 
the proof of positivity for solutions of this partial differential equations; see \cite{CJL19}. 


\theo{Theorem}\label{theo:positivity}
If $f$ is a real valued function in $L^1(\mathbb R^d)$ such that
\begin{equation}\label{uineq}
   f(x) - f\star f(x) =: u(x) \geqslant 0
\end{equation}
for all $x$.  Then  $\int_{\mathbb R^d} f(x)\ dx\leqslant\frac12$, 
and $f$ is given by the convergent series
\begin{equation}
f(x) = \frac{1}{2} \sum_{n=1}^\infty  c_n 4^n (\star^n u)(x)
\label{fun}
\end{equation} 
where the $c_n\geqslant0$ are the Taylor coefficients in the expansion of $\sqrt{1-x}$
\begin{equation}\label{3half}
  \sqrt{1-x}=1-\sum_{n=1}^\infty c_n x^n
  ,\quad
  c_n=\frac{(2n-3)!!}{2^nn!}  \sim n^{-3/2}
\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}).
\endtheo
\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,
(\ref{uineq}) becomes
\begin{equation} \label{ft}
\widehat f(k) = \widehat f(k)^2 +\widehat u(k)\ .
\end{equation}
At $k=0$, $a^2 - a = -b$, so that $\left(a - \textstyle\frac12\right)^2  = \textstyle\frac14 - b$.  Thus $0 \leqslant b \leqslant\textstyle\frac14$.  Furthermore, since  $u \geqslant 0$,
\begin{equation}
  |\widehat u(k)|\leqslant\widehat u(0) \leqslant \textstyle\frac14
\end{equation}
and the first   inequality is strict  for $k\neq 0$. Hence for $k\neq 0$,  $\sqrt{1-4\widehat u(k)} \neq 0$.  By the Riemann-Lebesgue Theorem, $\widehat{f}(k)$ and $\widehat{u}(k)$ are both 
continuous and vanish at infinity, and hence we must have that 
\begin{equation}
  \widehat f(k)=\textstyle\frac12-\textstyle\frac12\sqrt{1-4\widehat u(k)}
  \label{hatf}
\end{equation}
for all sufficiently large $k$, and in any case $\widehat f(k)= \frac12\pm\frac12\sqrt{1-4\widehat u(k)}$.  But by continuity and the fact that  $\sqrt{1-4\widehat u(k)} \neq 0$ for any $k\neq 0$, the sign cannot switch.  
Hence (\ref{hatf}) is valid for all $k$, including $k=0$, again by continuity.  At $k=0$,  $a = \textstyle\frac12 - \sqrt{1-4b}$, which proves (\ref{half}).
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 $|4 \widehat u(k)| \leqslant 1$,
${\displaystyle 
 \sqrt{1-4 \widehat u(k)} = 1 -\sum_{n=1}^\infty c_n (4 \widehat u(k))^n}$. Inverting the Fourier transform, yields (\ref{fun}),and since $\int_{\mathbb R^d} 4^n\star^n u(x){\rm d}x \leqslant 1$, 
 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 
 valid, and then (\ref{ft}) and (\ref{uineq}) are satisfied.
\qed
\bigskip

\theo{Theorem}\label{theo:decay}
Let $f\in L^1(\mathbb R^d)$ satisfy (\ref{ineq}) and $\int_{\mathbb R^d} f(x)\ dx=\textstyle\frac12$. Then 
$\int_{\mathbb R^d}|x| f(x)\ dx=\infty$.
\endtheo
\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 
$$|m|\int_{\mathbb R^d} |x| \star^nw(x){\rm d}x \geqslant  \int_{\mathbb R^d} m\cdot x \star^nw(x){\rm d}x = n|m|^2\ .$$
 It follows that  $\int_{\mathbb R^d} |x| f(x){\rm d}x \geqslant |m|\sum_{n=1}^\infty nc_n = \infty$. Hence $m=0$. 


Suppose temporarily that in addition, $|x|^2w(x)$ is integrable. Let $\sigma^2$ be the variance of $w$; i.e., $\sigma^2 = \int_{\mathbb R^d}|x|^2w(x){\rm d}x\ .$
Define the function $\varphi(x) = \min\{1,|x|\}$.  Then 
$$
\int_{\mathbb R^d}|x| \star^n w(x){\rm d}x =  \int_{\mathbb R^d}|n^{1/2}x| \star^n w(n^{1/2}x)n^{d/2}{\rm d}x  \geqslant  n^{1/2} \int_{\mathbb R^d}\varphi(x)\star^n w(n^{1/2}x)n^{d/2}{\rm d}x.
$$
By the Central Limit Theorem,  since $\varphi$ is bounded and continuous,
\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
\end{equation}
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 $c_n\sim n^{3/2}$, $\sum_{n=1}^\infty c_n \int_{\mathbb R^d}|x| \star^n w(x){\rm d}x= \infty$. 

 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}
\lim_{n\to\infty} \int_{|x| \leqslant R}\star^n w(n^{1/2}x)n^{d/2}{\rm d}x = 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}.
\qed
\bigskip

\delimtitle{\bf Remark}
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.
\endtheo
\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}
    \int |x|^pf(x)\ dx<\infty.
  \end{equation}
\endtheo
\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 
$$
f(x) = \sum_{n=1}^\infty c_n t^n \star^n w(x)\ . 
$$
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 $0 < p < 2$,
$\int_{\mathbb R^d} |x-m|^p \star^n w(x){\rm d}x   \leqslant  (n\sigma^2)^{p/2}$.
It follows that for $0 < p < 1$,
$$
\int_{\mathbb R^d} |x-m|^p f(x){\rm d}x   \leqslant  (\sigma^2)^{p/2} \sum_{n=1}^\infty n^{p/2} c_n  < \infty\ ,
$$
again using the fact that $c_n\sim n^{-3/2}$. 
\qed




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 $\int_{\mathbb R^d}f(x){\rm d}x < \textstyle\frac12$
can decay quite rapidly, even having finite exponential moments,  as we now show.


Consider a non-negative, integrable function $u$, which integrates to $r<\frac14$, and satisfies
\begin{equation}
  \int_{\mathbb R^d} u(x)e^{\lambda|x|}{\rm d}x < \infty
\end{equation}
for some $\lambda>0$.
The Laplace transform of $u$ is
$ \widetilde u(p):=\int e^{-px}u(x)\ {\rm d} x$ which is analytic for $|p|<\lambda$, and $\widetilde u(0) < \textstyle\frac14$.
Therefore, there exists $0<\lambda_0\leqslant \lambda$ such that, for all $|p|\leqslant\lambda_0$,
  $\widetilde u(p)<\textstyle\frac14$.
By Theorem \ref{theo:positivity},
${\displaystyle 
  f(x):=\frac12\sum_{n=1}^\infty 4^nc_n(\star^n u)(x)}$
is an integrable solution of (\ref{ineq}).  For  
 $|p|\leqslant\lambda_0$, it has a well-defined Laplace transform $ \widetilde f(p)$ given by 
\begin{equation}
  \widetilde f(p)=\int e^{-px}f(x)\ dx=\frac12(1-\sqrt{1-4\widetilde u(p)})
\end{equation}
which is analytic for $|p|\leqslant \lambda_0$.
Note that
${\displaystyle e^{s|x|} \leqslant \prod_{j=1}^d e^{|sx_j|} \leqslant \frac{1}{d}\sum_{j=1}^d e^{d|sx_j|} \leqslant \frac{2}{d}\sum_{j=1}^d \cosh(dsx_j)}$.
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 
$0 < s < \delta$, $\int_{\mathbb R^d} e^{s|x|}f(x){\rm d}x < \infty$. 
\bigskip 

 
\delimtitle{\bf Remark}
One might also consider the inequality $f \leqslant f \star f$ in $L^1(\mathbb R^d)$, but it is simple to construct  solutions that  have both signs. Consider any radial Gaussian probability density $g$,  
Then $g\star g(x) \geqslant g(x)$ for all sufficiently large $|x|$, and taking  $f:= ag$ for $a$ sufficiently large, we obtain $f< f\star f$ everywhere. Now on a small neighborhood of the origin, replace the value of 
$f$ by $-1$. If the region is taken small enough, the new function $f$ will still satisfy $f < f\star f$ everywhere.  
\endtheo
\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 $\mathbb R^d$.  
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 $R$. 

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
\eject
{\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.
\vskip20pt

\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[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[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[E42]{E42} C.-G. Esseen, {\em A moment inequality with an application to the central limit theorem}. Skand. Aktuarietidskr. {\bf 39} 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[SW71]{SW71} E. Stein and G. Weiss, {\em Introduction to Fourier analysis on Euclidean spaces}, Princeton University Press, Princeton NJ, 1971.





\end{thebibliography}

\end{document}