1: \documentstyle{amltd}
2: \begin{document}
3: \input amssym.def
4: \input amssym.tex
5: \annalsline{151}{2000}
6: %\received{March 17, 1997}
7: %\revised{July 21, 1999}
8: \startingpage{1}
9: \def\bye{\end{document}}
10: \font\tenrm=cmr10
11:
12: %--------------- Author macros ---------------
13: \def\underset#1#2{{#2}_{#1}}
14: \def\eqref#1{(\ref{#1})}
15: \def\ritem#1{\item[{\rm #1}]}
16: \catcode`\@=11
17: \font\twelvemsb=msbm10 scaled 1100
18: \font\tenmsb=msbm10
19: %\font\ninemsb=msbm7 scaled 1100%msbm9
20: \font\ninemsb=msbm10 scaled 800
21: \newfam\msbfam
22: \textfont\msbfam=\twelvemsb \scriptfont\msbfam=\ninemsb
23: \scriptscriptfont\msbfam=\ninemsb
24: \def\msb@{\hexnumber@\msbfam}
25: \def\Bbb{\relax\ifmmode\let\next\Bbb@\else
26: \def\next{\errmessage{Use \string\Bbb\space only in math
27: mode}}\fi\next}
28: \def\Bbb@#1{{\Bbb@@{#1}}}
29: \def\Bbb@@#1{\fam\msbfam#1}
30: \catcode`\@=12
31:
32:
33: \catcode`\@=11
34: \font\twelveeuf=eufm10 scaled 1100
35: \font\teneuf=eufm10
36: \font\nineeuf=eufm7 scaled 1100%eufm9
37: \newfam\euffam
38: \textfont\euffam=\twelveeuf \scriptfont\euffam=\teneuf
39: \scriptscriptfont\euffam=\nineeuf
40: \def\euf@{\hexnumber@\euffam}
41: \def\frak{\relax\ifmmode\let\next\frak@\else
42: \def\next{\errmessage{Use \string\frak\space only in math
43: mode}}\fi\next}
44: \def\frak@#1{{\frak@@{#1}}}
45: \def\frak@@#1{\fam\euffam#1}
46: \catcode`\@=12
47:
48:
49:
50:
51: %-------------- Author entries --------------------
52:
53: \title{Some spherical uniqueness theorems\\ for
54: multiple trigonometric series}
55: \shorttitle{Spherical Uniqueness Theorems}
56:
57: \acknowledgements{The research of both authors
58: was partially supported by NSF grant DMS 9707011 and a grant from the Faculty and Development Program
59: of the College of Liberal Arts and Sciences, De Paul University.\\
60: \phantom{From*}1991 {\it Mathematics Subject Classification}.
61: Primary 42B05, 42B99; Secondary 42B08, 42B15, 42A63. }
62: \twoauthors{J.\ Marshall Ash}{Gang Wang}
63: \institutions{De Paul University, Chicago, IL\\
64: {\eightpoint {\it E-mail addresses\/}: mash@math.depaul.edu}\\
65: \hglue.97in {\eightpoint gwang@math.depaul.edu}}
66:
67: \phantom{more space}
68:
69: \bigbreak \centerline{\bf Abstract}
70: \bigbreak
71: We prove that if a multiple trigonometric series is spherically Abel
72: sum\-mable everywhere to an everywhere finite function $f(x)$ which is
73: bounded below by an integrable function, then the series is the Fourier
74: series of $f(x)$ if the coefficients of the multiple trigonometric series
75: satisfy a mild growth condition. As a consequence, we show that if a
76: multiple trigonometric series is spherically convergent everywhere to an
77: everywhere finite integrable function $f(x)$, then the series is the Fourier
78: series of $f(x)$. We also show that a singleton is a set of uniqueness. These results are generalizations of a recent theorem of
79: J. Bourgain and some results of V. Shapiro.
80:
81: \section{Introduction and summary of results}\label{s:0}
82:
83: We start with the question of spherical uniqueness of multiple
84: trigonometric series for integrable functions under Abel summability. Greek
85: letters $\xi, \eta, \cdots$ will denote points of the $d$-dimensional
86: lattice $\Bbb Z^d$, Roman letters $x, y,\cdots$ points of the
87: $d$-dimensional torus $\Bbb {T}^d = [-\pi, \pi)^d$, $\langle \cdot, \cdot
88: \rangle$ inner product, and $|\cdot|$ $d$-dimensional Euclidean norm.
89: For a multiple trigonometric series
90: $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$
91: where the coefficients $a_{\xi}$ are arbitrary complex numbers, the Abel
92: sum is defined to be the limit of the function
93: \[
94: f(x, t) ={\sum}_{\xi\in Z^d} a_{\xi}
95: e^{i\langle x, \xi \rangle-|\xi|t}
96: \]
97: as $t\to 0^+$ if such limit exists. In general, denote
98: \begin{eqnarray*}
99: \noalign{\vskip4pt}
100: f^*(x) & =& \limsup_{t\to 0^+} f(x,t)\\
101: \noalign{\vskip4pt}
102: & =& \Re f^*(x) + i \Im f^*(x)
103: \\
104: \noalign{\vskip-6pt}
105: \end{eqnarray*}
106: $f_*(x) = \Re f_*(x) + i \Im f_*(x)$ is similarly defined with $\limsup$ being
107: replaced by $\liminf$.
108:
109: It is well-known, when $d=1$, that if $\sum a_{\xi}e^{i\xi x}$ is Abel
110: summable to 0 everywhere and if $a_{\xi}=o(|\xi|)$, then all $a_{\xi}=0$.
111: See, for example, \cite{Ve1} and \cite{Ve2}. To see that this theorem is
112: sharp, look at the one dimensional series\break $\delta ^{\prime }(x)=-
113: \sum \xi\sin {\xi x}$, which may be thought of as the derivative
114: of the Dirac delta function. It is easy to check that this series is
115: Abel summable to $0$, although the growth condition is just
116: barely violated. Thinking of $\delta ^{\prime }$ as a
117: degenerate $d$ dimensional function, it is immediately clear that the
118: hypothesis of a $d$ dimensional uniqueness theorem concerning Abel
119: summability will necessarily have to carry some growth condition.
120: One generalization of this fact is due to Victor Shapiro, who extended one
121: dimensional work of Verblunsky and of Rajchman and Zygmund (\cite{Sh}).
122:
123: \phantom{hee}
124:
125: \proclaimtitle{Shapiro}
126: \proclaim{Theorem} \label{t:sh1}
127: Let $\sum a_{\xi}e^{i\langle\xi, x\rangle}$ be a multiple trigonometric
128: series{\rm .} Suppose that
129: \begin{itemize}
130: \ritem{1.} the coefficients $a_{\xi}$ satisfy the following growth rate
131: condition\/{\rm :}
132: \end{itemize}
133: \begin{equation}\label{0.1}
134: \sum_{R-1<\left| \xi\right| \leq R}\left| a_{\xi}\right|
135: =o\left(R\right) \hbox{ as } R\rightarrow \infty ,
136: \end{equation}
137: \begin{itemize}
138: \ritem{2.} $f^{*}(x)$ and $f_{*}(x)$ are finite for all $x${\rm , }
139:
140: \ritem{3.} $a_{\xi}=\overline{a_{-\xi}}$ for all $\xi${\rm ,} and
141:
142: \ritem{4.} $\min\{\Re f_{*}(x), \Im f_{*}(x)\} \geq A(x)$ where $A\left( x\right)
143: $ is in $L^{1}\left( \Bbb {T}^{d}\right) .$
144: \end{itemize}
145:
146: \noindent Then $f_{*}\left( x\right) \in L^{1}\left( \Bbb {T}^{d}\right) $ and $\sum
147: a_{\xi}e^{i\langle \xi,x\rangle}$ is the Fourier series of $f_{*}.$
148: \endproclaim
149:
150: \phantom{hee}
151:
152:
153: This theorem is sharp because the example $\delta ^{\prime }$
154: mentioned above just barely fails to meet condition \eqref{0.1}. Nevertheless
155: condition \eqref{0.1} is disappointingly strong in the sense that when Abel
156: summability is replaced by regular convergence, condition \eqref{0.1} is not a direct consequence of convergence. However, there is a known theorem concerning the coefficients' growth
157: rate for spherically convergent multiple trigonometric series. In fact, it is
158: implied by the following Cantor-Lebesgue type theorem.
159:
160: \proclaimtitle{Connes}
161: \proclaim{Theorem} \label{t:con} Let ${\cal O}\subset \Bbb T^d$
162: be a ball or a subset which has full measure and is of Baire second category
163: relative to $\Bbb {T}^d$.
164: If $\sum_{|\xi|=R} a_{\xi} e^{i\langle x, \xi \rangle}$ tends
165: to $0$ as $R\to\infty$ at every point of ${\cal O}$, then
166: \begin{equation}\label{0.2}
167: \varepsilon_R^2 = \sum_{|\xi| = R}|a_{\xi}|^2
168: = o(1) \qquad \hbox{ as } R\to\infty.
169: \end{equation}
170: \endproclaim
171:
172: Connes proved this theorem for dimension $d$ in 1976, twenty years
173: after Shapiro's Theorem~\ref{t:sh1}. Cooke \cite{C} and shortly thereafter
174: Zygmund \cite{Z} had completed the $d=2$ case five years before Connes' work.
175:
176: An easy corollary of Theorem~\ref{t:con} gives the coefficients' growth rate
177: condition for spherically convergent multiple trigonometric series.
178:
179: \proclaimtitle{Connes}
180: \proclaim{{C}orollary} \label{c:con}
181: Let ${\cal O}\subset \Bbb T^d$
182: be a ball or a subset which has full measure and is of Baire second category
183: relative to $\Bbb {T}${\rm .}
184: If\/ $\lim_{R\to\infty}\sum_{|\xi|\leq R} a_{\xi}
185: e^{i\langle x, \xi \rangle}$ exists {\rm (}\/as a finite number\/{\rm )} at each point of
186: ${\cal O}${\rm ,} then
187: \begin{equation}\label{0.3}
188: \varepsilon_R^2 = \sum_{|\xi| = R}|a_{\xi}|^2
189: = o(1) \qquad \hbox{ as } R\to\infty.
190: \end{equation}
191: \endproclaim
192:
193:
194: The coefficients' growth rate condition given by \eqref{0.3} does
195: not imply condition \eqref{0.1} when $d\geq 3$. To remedy this problem, we
196: first prove the following analogue of Theorem~\ref{t:sh1} under the condition \eqref{0.3}. We use notation $A\sim B$ to denote $B/2 \leq A < B$.
197:
198: \proclaim{Theorem} \label{t:1} Consider the multiple trigonometric series
199: $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$ where
200: the coefficients $a_{\xi}$ are arbitrary complex numbers{\rm .} Suppose that
201: \begin{itemize}
202: \ritem{1.} the coefficients of the series $a_{\xi}$ satisfy
203: \end{itemize}
204: \begin{equation}\label{e:bou}
205: \sum_{|\xi|\sim R} |a_{\xi}|^2 = \underset{R/2 \leq |\xi|< R}
206: {\sum} |a_{\xi}|^2 = o(R^2) \hbox{ as } R\to\infty,
207: \end{equation}
208: \begin{itemize}
209: \ritem{2.} $f^*(x)$ and $f_*(x)$ are finite for all $x${\rm ,} and
210: \ritem{3.} $\min\{\Re f_*(x), \Im f_*(x)\}$ is bounded below by a function $A(x)$
211: in $L^1(\Bbb T^d)${\rm .}
212: \end{itemize}
213: \noindent Then $f_*(x)$ is in $L^1(\Bbb T^d)$ and $\sum_{\xi\in Z^d}
214: a_{\xi} e^{i\langle x, \xi \rangle}$ is its Fourier series{\rm .}
215: \endproclaim
216:
217: Note that condition \eqref{0.3} implies condition \eqref{e:bou} since
218: \[
219: \sum_{|\xi|\sim R} |a_{\xi}|^2 = \sum_{k= R^2/4}^{R^2 - 1} \sum_{|\xi|^2 = k}
220: |a_{\xi}|^2 = o\left(\sum_{k= R^2/4}^{R^2 - 1} 1 \right) = o(R^2).
221: \]
222:
223: Since \eqref{0.1} implies
224: \[
225: \sum_{|\xi|\sim R} |a_{\xi}| = o(R^2),
226: \]
227: \eqref{0.1} implies \eqref{e:bou} if $a_{\xi}$ is bounded. But in
228: general, and when $d\geq 3$, \eqref{0.1} and \eqref{e:bou} do not relate to
229: each other. Notice that \eqref{e:bou} implies that
230: \[
231: \sum_{|\xi|^2\sim R}|a_{\xi}|^2 = \sum_{R/2 \leq |\xi|^2<R}|a_{\xi}|^2 \leq \sum_{|\xi|\sim \sqrt{R}}|a_{\xi}|^2 = o(R).
232: \]
233:
234: As a consequence of Theorem~\ref{t:con} and Theorem~\ref{t:1}, we obtain the following two
235: spherical uniqueness theorems for multiple trigonometric series which are convergent to a function. These theorems make no assumption whatsoever about coefficient size.
236:
237: \proclaim{Theorem}\label{t:2}
238: Let $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$ be
239: a trigonometric series which converges spherically everywhere to an
240: everywhere finite function $f(x)${\rm ;} i{\rm .}e{\rm .,}
241: \begin{equation}\label{0.4}
242: \lim_{R\to\infty}\sum_{|\xi|\leq R} a_{\xi} e^{i\langle x, \xi\rangle} = f(x)
243: \hbox{ for all } x\in \Bbb {T}^d.
244: \end{equation}
245: If $\min\{\Re f(x), \Im f(x)\}\geq g(x)$ for all $x$ and
246: $g(x) \in L^1(\Bbb {T}^d)${\rm ,} then
247: $f(x)$ is in $L^1(\Bbb {T}^d)$ and $a_{\xi} $ is the $\xi^{\rm th}$ Fourier
248: coefficient of $f(x)$ for all $\xi\in \Bbb {Z}^d${\rm . }
249: \endproclaim
250:
251: In particular,
252:
253: \proclaim{Theorem}\label{t:3}
254: Let $f(x)\in L^1(\Bbb {T}^d)$ be finite at every $x${\rm .} If
255: $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$ is
256: a trigonometric series which converges spherically to $f(x)$ at every
257: point $x${\rm ,} i{\rm .}e{\rm . }
258: \begin{equation}\label{0.5}
259: \lim_{R\to\infty}\sum_{|\xi|\leq R} a_{\xi} e^{i\langle x, \xi\rangle} = f(x)
260: \hbox{ for all } x\in \Bbb {T}^d,
261: \end{equation}
262: then $a_{\xi} $ is the $\xi^{\rm th}$ Fourier coefficient of $f(x)$ for all
263: $\xi\in \Bbb {Z}^d${\rm . }
264: \endproclaim
265:
266:
267: Special cases of Theorem~\ref{t:3} have been proved by various people.
268: When $d=1$ and $f(x) \equiv 0$, this is the original uniqueness theorem of
269: Cantor. For general $f(x)\in L^1(\Bbb T^1)$, it was first proved by de
270: la Vall\'ee-Poussin. When $d=2$, Theorem~\ref{t:sh1} combined with
271: the work of Cooke \cite{C} implies Theorem~\ref{t:1} and thus,
272: Theorem~\ref{t:2} and Theorem~\ref{t:3}. The major breakthrough came
273: when Bourgain \cite{B} proved Theorem~\ref{t:3} for the special case of
274: $f(x) \equiv 0$.
275: For a survey on the uniqueness of multiple trigonometric series under
276: various summation modes, as well as many open problems in this area, please
277: refer to Ash and Wang \cite{AW}.
278:
279: The proof of Theorem~\ref{t:1} is mainly based on Shapiro's framework
280: \cite{Sh}. To avoid assuming condition \eqref{0.1}, we exploit an idea that
281: Bourgain \cite{B}
282: used when he proved Theorem~\ref{t:3} for the special case $f(x) \equiv 0$.
283: We refer to \eqref{e:bou} hereafter as Bourgain's condition, in his honor. This condition simply asserts that Connes'
284: condition holds ``on the average."
285:
286: The detailed proof of Theorem~\ref{t:1} is given in Sections~\ref{s:1}
287: through \ref{s:4}.
288:
289: At the end of the paper, we begin the study of sets of uniqueness for spherical convergence. As a first step toward establishing this theory, we show that any singleton is a set of uniqueness.
290:
291: \proclaim{Theorem}\label{c:0.1}
292: Let $q$ be a point on $\Bbb T^d${\rm .} Suppose that a multiple trigonometric
293: series $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$
294: spherically converges everywhere except at $q$ to a function $f(x)\in L^1 (\Bbb {T}^d)${\rm .}
295: Furthermore{\rm ,} suppose $f(x)$ is finite for all $x$ except $q${\rm .} Then
296: $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$ is
297: the Fourier series of $f(x)${\rm .}
298: \endproclaim
299:
300: It is easily deduced from Theorem~\ref{t:con} and the following fact about Abel summability,
301: which is an analogue of a theorem of Shapiro \cite[\S 6]{Sh}:
302:
303: \proclaim{Theorem}\label{t:4}
304: Consider the multiple ($d\geq 2$) trigonometric series\break
305: $\sum_{\xi\in Z^d} a_{\xi}
306: e^{i\langle x, \xi \rangle}$ where the coefficients $a_{\xi}$ are arbitrary
307: complex numbers{\rm .} Let $q$ be a point on $\Bbb T^d${\rm .} Suppose that
308: \begin{itemize}
309: \ritem{1.} $\sum_{|\xi|\sim R} |a_{\xi}|^2 = o(R^2)$ as
310: $R\to\infty${\rm ,}
311: \ritem{2.} $f^*(x)$ and $f_*(x)$ are finite for all $x$ except $q${\rm ,} and
312: \ritem{3.} $f^*(x)$ and $f_*(x)$ are functions in $L^1(\Bbb T^d)${\rm .}
313: \end{itemize}
314: \noindent Then $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$ is
315: the Fourier series of $f_*(x)${\rm .}
316: \endproclaim
317:
318: Note that the theorem is false when $d=1$ since the trigonometric series $\sum
319: e^{i\xi x}$ is Abel convergent to $0$ everywhere in $\Bbb T \setminus
320: \{0\}$.
321:
322: \section{Proof of Theorem~\ref{t:1}}\label{s:1}
323:
324: We may assume $d\ge 3$ since the cases $d=1$ and $d=2$ are known.
325:
326: We need some preliminary results and some notation before we start the
327: proof. Without loss of generality, by considering the real and imaginary
328: parts separately, we may assume that $a_{\xi} = \overline a_{-\xi}$, where
329: $\overline a$ is the conjugate of the complex number $a$. Thus $f(x, t),
330: f^*(x)$ and $f_*(x)$ are all real functions. In addition, we may assume
331: that $a_0 = 0$.
332:
333: Define
334: \begin{eqnarray}\label{1.0}
335: f_1(x,t) &=& -\underset{\xi \neq 0}{\sum} \frac{a_{\xi}}{|\xi|}
336: e^{i\langle x, \xi \rangle-|\xi|t}\\
337: \noalign{\vskip6pt}
338: f_2(x,t) &=& -\underset{\xi \neq 0}{\sum} \frac{a_{\xi}}{|\xi|^2}
339: e^{i\langle x, \xi \rangle-|\xi|t}.
340: \nonumber
341: \end{eqnarray}
342: Under the condition \eqref{e:bou}, it is easy to see that for each
343: $x\in \Bbb {T}^d$ and $t>0$, $f(x, t), f_1(x, t)$ and $f_2(x, t)$ converge
344: absolutely and hence are infinitely differentiable as functions of $t>0$.
345: Thus by the mean value theorem, for $t_1>t_2>0$, there exist $t_3, t_4\in
346: (t_2, t_1)$ such that $f_1(x, t_1)-f_1(x, t_2) = f(x, t_3)(t_1-t_2)$, and
347: $f_2(x, t_1)- f_2(x, t_2) = -f_1(x, t_4)(t_1-t_2)$. Since for each $x$,
348: $f^*(x)$ and $f_*(x)$ are finite, $f(x, t)$ is bounded for all $t>0$. The
349: bound depends on $x$ in general. Thus, for each $x$, there exist finite-valued functions $f_1(x)$ and $f_2(x)$ such that
350: \begin{equation}\label{1.001}
351: f_1(x, t)\to f_1(x) \hbox{ and } f_2(x, t)\to f_2(x) \hbox{ as } t\to 0^+.
352: \end{equation}
353:
354: On the other hand, if we define the Riemann function $F(x)$ by
355: \begin{equation}\label{1.1}
356: F(x) = -\sum_{\xi \neq 0} \frac{a_{\xi}}{|\xi|^2} e^{i\langle x, \xi\rangle},
357: \end{equation}
358: then, because of the Bourgain condition \eqref{e:bou}, $F(x)\in L^2(\Bbb T^d)$
359: and $f_2(x, t)\break \stackrel{L^2}{\rightarrow} F(x)$ as $t\to 0^{+}$. In fact,
360: observe that there is an absolute constant $C$ such that
361: \begin{eqnarray}\label{1.01}
362: \qquad ||f_2(x, t)-F(x)||_2^2 &=& \sum_{\xi\neq
363: 0}\frac{|a_{\xi}|^2}{|\xi|^4}\left(1-e^{-2|\xi|t}\right)\\
364: & \leq & C \sum_{k = 1}^{\infty} 2^{-2k}\left(1 - e^{-2^{k/2+1}t}\right)
365: \sum_{|\xi|^2 \sim 2^{k}}|a_{\xi}|^2\nonumber \\
366: &\leq& C \sum_{k=1}^{\infty} 2^{-k}\left(1- e ^{-2^{k/2+1}t}\right)\nonumber\\
367: &\to& 0 \hbox{ as } t\to 0^+. \nonumber
368: \end{eqnarray}
369: Thus,
370: \begin{equation}\label{1.2}
371: f_2(x) =F(x) \quad {\rm a.e.}
372: \end{equation}
373:
374: The key to the proof is to show that $\Delta f_2(x) = f_*(x)$ almost
375: everywhere. To this end, we need to use a generalized Laplacian.
376:
377: Let $B(x, \rho)$ be an open ball in $\Bbb {T}^d$ centered at $x\in
378: \Bbb {T}^d$
379: with radius $\rho >0$ and $m(B(x, \rho))$ the volume of $B(x,
380: \rho)$. Then $m(B(x, \rho)) = v_d \rho^d$, where $v_d$ is the volume of
381: the unit ball in $\Bbb R^d$. For any locally integrable function $g(x)$, the average
382: of $g$ over $B(x, \rho)$ is
383: \begin{eqnarray*}
384: A_\rho g(x) &= &\frac{1}{m(B(x, \rho))}\int_{B(x, \rho)} g(y) \, dy\\
385: \noalign{\vskip4pt}
386: & =& \frac{1}{v_d\rho^d}\int_{B(x, \rho)} g(y) \, dy.
387: \end{eqnarray*}
388:
389:
390: Let
391: \[
392: I(x) = \frac{I_{B(0, 1)}(x)}{m(B(0, 1))},
393: \]
394: where $I_{B(0, 1)}(x)$ is the characteristic function of the unit ball.
395: Denote
396: $\hat I(\xi)$ to be the Fourier transform of $I(x)$. Then $\hat I(\rho
397: \xi)$ satisfies the following properties:
398: \begin{equation}\label{1.3}
399: \lim_{\rho\to 0} \frac{\hat I(\rho\xi)-1}{\rho^2 |\xi|^2} = -\frac12
400: \int_{B(0, 1)} x_1^2 I(x) \, dx = c_d < 0,
401: \end{equation}
402: and for $|\xi| = 1$,
403: \begin{equation}\label{1.4}
404: \int_0^\infty \left|\partial_r\left[\frac{\hat I(\rho r \xi) -1}{\rho^2
405: r^2}\right]\right| \, dr = \int_0^\infty \left|\partial_r\left[\frac{\hat
406: I(r \xi) -1}{r^2}\right]\right| \, dr < c.
407: \end{equation}
408: Note that the constant $c$ in \eqref{1.4} is independent of $\rho$.
409:
410: The above two equalities are standard. In fact, to see \eqref{1.3}, rotate
411: (choose the first coordinate axis to be in the direction of
412: $\xi$) and use polar coordinates to get
413: \begin{eqnarray}\label{1.5}
414: \hat I(\rho \xi) -1 &= &\frac{1}{v_d} \int_{|x|\le 1}
415: (e^{i\langle x, \rho\xi\rangle}-1)\, dx \\
416: \noalign{\vskip4pt}
417: & =& \frac{1}{v_d} \int_{|x| \le 1} (e^{i
418: x_1\rho|\xi|}-1)\, dx\nonumber\\ \noalign{\vskip4pt}
419: & = &\frac{v_{d-1}}{v_d} \int_{-1}^{1} (\cos(\rho |\xi| x_1)-1)
420: (1-x_1^2)^{\frac{d-1}{2}}\, dx_1.\nonumber
421: \end{eqnarray}
422: Since for any $x\in \Bbb {T}$, $|\cos x -1| \le \frac{x^2}{2}$, and
423: $\lim_{x\to 0} \frac{\cos x -1}{x^2} = -\frac12$,
424: by the bounded convergence theorem and \eqref{1.5}, we have
425: \begin{eqnarray*}
426: \lim_{\rho\to 0} \frac{\hat I(\rho\xi)-1}{\rho^2 |\xi|^2} &=&
427: -\frac{v_{d-1}}{2v_d} \int_{-1}^{1} x_1^2 (1-x_1^2)^{\frac{d-1}{2}}\, dx_1\\ \noalign{\vskip4pt}
428: & =& -\frac{1}{2}\int_{B(0,1)} x_1^2 I(x) \, dx = c_d < 0.
429: \end{eqnarray*}
430: Observe that the above argument shows $\hat I(\xi_1) = \hat
431: I(\xi_2)$ if $|\xi_1| = |\xi_2|$. Thus, we may abuse our notation and write
432: $\hat I(\xi) = \hat I(|\xi|)$.
433:
434:
435: Inequality \eqref{1.4} also follows similarly. If $|\xi| = 1$, then
436: \[
437: \partial_r \left[\frac{\hat I(\rho r \xi) -1}{\rho^2 r^2}\right] =
438: -2\frac{v_{d-1}}{v_d}
439: \int_{-1}^{1} \frac{(1- \frac{1}{2}i r \rho x_1) e^{i r \rho x_1} -1}{r^3
440: \rho^2} (1-x_1^2)^{\frac{d-1}{2}}\, dx_1.
441: \]
442: Thus, for $c= -2v_{d-1}/v_{d}$,
443: \begin{eqnarray*}
444: &&
445: \hskip-36pt\int_0^\infty \left|\partial_r\left[\frac{\hat I(\rho r \xi) -1}{\rho^2
446: r^2}\right]\right| \, dr\\
447: &&\qquad \enspace = \ c \int_{0}^{\infty}\left| \int_{-1}^{1}
448: \frac{(1- \frac{1}{2}i r \rho x_1) e^{i r \rho x_1} -1}{r^3 \rho^2}
449: (1-x_1^2)^{\frac{d-1}{2}}\, dx_1 \right|\, dr\\ \noalign{\vskip4pt}
450: &&\qquad \enspace = \ c \int_{0}^{\infty}\left| \int_{-1}^{1}
451: \frac{(1 -\frac{1}{2} i r x_1) e^{i r x_1} -1}{r^3 }
452: (1-x_1^2)^{\frac{d-1}{2}}\, dx_1 \right|\, dr\\ \noalign{\vskip4pt}
453: &&\qquad \enspace = \ \int_0^\infty \left|\partial_r\left[\frac{\hat
454: I(r \xi) -1}{r^2}\right]\right| \, dr
455: \end{eqnarray*}
456: by the simple change of variable argument: $\rho r \to r$.
457: The above integral is finite since
458: \[
459: \left|\int_{-1}^{1} \frac{(1- \frac{1}{2}i r x_1) e^{i r x_1} -1}{r^3
460: }(1-x_1^2)^{\frac{d-1}{2}} \, dx_1 \right| \le
461: \frac{c}{r^2}
462: \hbox{ as } r \to \infty ,
463: \]
464: and
465: \[
466: (1- \frac{1}{2} i r x_1) e^{i r x_1} -1 = \frac{1}{2} i r x_1 + O(r^3)
467: \hbox{ as } r \to 0.
468: \]
469:
470: Define the generalized Laplacian operator on $g(x)\in L^1$ to be
471: \[
472: \tilde \Delta g(x) = \lim_{\rho\to 0}-\frac{1}{c_d}
473: \frac{A_\rho g(x)-g(x)}{\rho^2}
474: \]
475: if such a limit exists (not necessarily finite), where $c_d<0$ is the
476: constant given in \eqref{1.3}. We can also define the upper and lower
477: generalized Laplacians $\tilde\Delta^{*} g(x)$ and $\tilde\Delta_{*} g(x)$
478: by replacing $\lim$ by $\limsup$ and $\liminf$ respectively when the
479: function $g(x)$ is real-valued. It is clear that all three of these generalized Laplacians agree with the usual Laplacian when applied to a $C^2$ function. Recall that $a_0 = 0$. For $f_2(x, t)$ given
480: by \eqref{1.0}, we have for $x\in \Bbb {T}^d$,
481: \begin{eqnarray}\label{1.6}
482: %
483: \frac{A_{\rho} f_2(x, t) -f_2(x, t)}{\rho^2} & = & - \sum_{\xi\ne 0}
484: \frac{a_{\xi}}{|\xi|^2}
485: \frac{\hat I(\rho \xi) -1}{\rho^2} e^{i\langle x, \xi\rangle-|\xi|t}\\ \noalign{\vskip4pt}
486: & = &- \sum_{k\ge 1} \frac{\hat I(\rho \sqrt{k}) -1}{\rho^2 k}\sum_{|\xi|^2 =
487: k} a_{\xi} e^{i\langle x, \xi \rangle-|\xi|t}\nonumber\\ \noalign{\vskip4pt}
488: & =&- \sum_{k\ge 1} \left(\sum_{|\xi|^2 \le k} a_{\xi}e^{i\langle x,\xi
489: \rangle-|\xi|t}\right)\nonumber\\ \noalign{\vskip4pt}
490: && \qquad\quad\times \left(\frac{\hat I(\rho\sqrt{k}) -1}{\rho^2 k}-\frac{\hat
491: I(\rho\sqrt{k+1}) -1}{\rho^2 (k+1)}\right)\nonumber \\ \noalign{\vskip4pt}
492: & \to & - f(x, t) c_d \hbox{ as } \rho \to 0 \nonumber
493: \end{eqnarray}
494: since by the fundamental theorem of calculus and \eqref{1.4},
495: \[
496: \sum_{k\ge 1} \left|\frac{\hat I(\rho\sqrt{k}) -1}{\rho^2 k} -\frac{\hat
497: I(\rho\sqrt{k+1}) -1}{\rho^2 (k+1)}\right| \le \int_0^\infty
498: \left|\partial_r\left[\frac{\hat I(\rho r \xi) -1}{\rho^2
499: r^2}\right]\right| \, dr < b <\infty
500: \]
501: for a constant $b$ independent of $\rho$. Thus, the above argument shows
502: that
503: \begin{equation}\label{1.7}
504: \tilde \Delta f_2(x, t) = \lim_{\rho\to 0}-\frac{1}{c_d}
505: \frac{A_{\rho} f_2(x, t)-f_2(x, t)}{\rho^2} =
506: f(x, t)
507: \end{equation}
508: for $x\in \Bbb {T}^d$ and $t> 0$.
509:
510: To pass to the limit as $t\to 0^+$, we need the following lemma of Shapiro
511: (Lemma~7 of \cite{Sh2}). To see that Shapiro's lemma applies,
512: note that $F \in L^2(\Bbb {T}^d)$ implies $F \in L^1(\Bbb {T}^d)$.
513: \proclaim{Lemma}\label{l:1.0}
514: If
515: {\rm \eqref{e:bou}} holds{\rm ,} then at every point $x$ where $f_*(x)$ and $f^*(x)$ are
516: finite{\rm ,}
517: \begin{equation}\label{1.8}
518: \tilde\Delta_* f_2(x) \leq f^*(x) \quad \hbox{ and } \quad f_*(x)
519: \leq \tilde\Delta^* f_2(x).
520: \end{equation}
521: \endproclaim
522:
523: The following classical results on the Green's function $G(x)$ appear with proof as
524: Lemma 8 of Shapiro \cite{Sh2}. (Also see Theorem 6 of Bochner \cite{Bo}.)
525:
526: \proclaim{Lemma}\label{l:1.1}
527: There is a function $G(x)$ in $L^1(\Bbb {T}^d)$
528: whose Fourier series is given by $\sum_{|\xi|\ne 0} |\xi|^{-2}
529: e^{i\langle x, \xi\rangle}${\rm .} Further{\rm ,} $G(x)$ has the following properties\/{\rm :}
530: \begin{itemize}
531: \ritem{1.} $G(x)$ is in class $C^{\infty}(\Bbb {T}^d)$ away from $0$ and $\Delta G(x) = 1$ for $x\ne 0${\rm .}
532: \ritem{2.} $G(x) = \Phi (x) + H^{*}(x)$ where $H^{*}$ is continuous on
533: $\Bbb T^d$ and $\Delta H^*(x) = 1$ for $x\in\Bbb {T}^d\setminus \{0\}$ and where $\Phi(x) = C_d |x|^{-(d-2)}$
534: for $d\ge 3$ with
535: $C_d = 2^{d-1}{\pi}^{d/2}\Gamma(d/2)/(d-2)$ and $\Phi(x)=-2\pi\log |x|$ when $d=2$.
536: \ritem{3.} Let $u(x)$ be an upper semi\/{\rm -}\/continuous function on $\Bbb T^d$
537: which is also\break in $L^1(\Bbb T^d)${\rm .} Define $U(x) = (2\pi)^{-d}\int_{T^d}
538: G(x-y) u(y)\, dy$ and $u_0 =\break (2\pi)^{-d}\int_{T^d} u(y)\, dy${\rm .} Then $U(x)$
539: is upper semi\/{\rm -}\/continuous on $\Bbb {T}^d${\rm ,}\break $U(x)\in L^1(\Bbb T^d)${\rm ,} and
540: $\tilde\Delta_{*}U(x) \ge -u(x) + u_0$ for $x\in \Bbb {T}^d${\rm .} Moreover{\rm ,}
541: $\tilde\Delta^{*}U(x) = \tilde\Delta_{*}U(x) = - u(x) + u_0$ almost
542: everywhere in $\Bbb T^d${\rm .}
543: \end{itemize}
544:
545: \endproclaim
546:
547: A consequence of Lemma~\ref{l:1.1} is that for any integrable
548: function $u$, the Fourier series of $U(x) = (2\pi)^{-d}\int_{T^d} G(x-y)
549: u(y)\, dy$ is $\sum_{|\xi|\ne 0} u_{\xi} |\xi|^{-2}e^{i\langle x, \xi
550: \rangle}$, where $u_0 + \sum_{|\xi|\ne 0} u_{\xi}e^{i\langle x, \xi
551: \rangle}$ is the Fourier series of $u$.
552:
553: We now state the following key lemma which will be proved in
554: Section~\ref{s:2}. The function $\overline U$ will not in general be periodic, so we have to work in $\Bbb R^d$, rather than in $\Bbb T^d$.
555:
556: \proclaim{Lemma}\label{l:1.2}
557: Let $f_2(x)$ be as given in {\rm \eqref{1.001}} where $f(x, t)$ satisfies the
558: conditions in Theorem~{\rm \ref{t:1}.} Suppose that $\overline U(x)$ is an
559: upper semi\/{\rm -}\/continuous function and that it is in $L_{\rm loc}^1(\Bbb {R}^d)${\rm .}
560: Let $S(x) = f_2(x)+\overline U(x)${\rm .} If
561: $\tilde\Delta^* S(x) \geq 0${\rm ,} then $S(x)$ is subharmonic in $\Bbb R^d${\rm .}
562: \endproclaim
563:
564: \demo{{R}emark\/ {\rm 2.1}}
565: By modifying the proof of Lemma~\ref{l:1.2} in Section~\ref{s:2},
566: Lemma~\ref{l:1.2} can be shown to hold locally. Explicitly, we can replace
567: $\Bbb
568: R^d$ everywhere in Lemma~\ref{l:1.2} by any open ball $B\subset \Bbb
569: R^d$ and $L_{\rm loc}^1(\Bbb {R}^d)$ by $L^1(B)$.
570: \enddemo
571:
572:
573: We now are ready to prove Theorem~\ref{t:1}.
574:
575: Since $A(x)\in L^1(\Bbb {T}^d)$, there exists an upper
576: semi-continuous function $u(x)$ (see p.75 of \cite{S}, for example) such
577: that $u(x) \leq A(x)$. As in Lemma~\ref{l:1.1}, define
578: $U(x) = (2\pi)^{-d}\int_{T^d} G(x-y) u(y)\, dy$,
579: $u_0 = (2\pi)^{-d}\int_{T^d} u(y)\, dy$ and
580: $S(x) = f_2(x) + U(x) - u_0|x|^2/(2d)$. Then by Lemma~\ref{l:1.0},
581: $\tilde\Delta^{*} f_2(x) \geq f_*(x) \geq A(x) \geq u(x)$. Consequently,
582: by periodicity, Lemmas~\ref{l:1.1} and \ref{l:1.2}, $S(x)$ is subharmonic in
583: $\Bbb R^d$. Therefore, by Riesz's representation for subharmonic functions and a theorem of Saks \cite{S}, $\tilde\Delta^{*} S(x) = \tilde\Delta_{*} S(x)$ almost everywhere and is in $L^{1}$ locally. Since $\tilde\Delta^* U(x) = \tilde\Delta_* U(x)$ almost everywhere and is in $L^{1}$ locally, this shows that $\tilde\Delta^{*} f_2(x) = \tilde\Delta_{*} f_2(x)$ almost everywhere and is in $L^{1}$ locally. Thus by assumption and Lemma~\ref{l:1.0}, $f_*(x)$ is in $L^1$ locally.
584:
585: Let $B(x) = \min\{f^*(x), \tilde\Delta^{*} f_2(x)\}$. Then by Lemma~\ref{l:1.0}, $\tilde\Delta_* f_2(x) \leq B(x) \leq \tilde\Delta^{*}
586: f_2(x)$. Consequently, $B(x) = \tilde\Delta^{*} f_2(x)$ almost everywhere,
587: and is in $L_{\rm loc}^{1}(\Bbb R^d)$. By a theorem of Vitali-Carath\'eodory (p. 75 of \cite{S}), there exists a nondecreasing sequence of upper semi-continuous functions $\{u^k(x)\}$ on $\Bbb {R}^d$, which are also in $L_{\rm loc}^{1}(\Bbb R^d)$, such that each $u^k(x)$
588: is bounded above and $u^{k}(x) \leq B(x)$ for all $x\in \Bbb R^d$,
589: \begin{equation}\label{1.9}
590: \lim_{k\to\infty} u^k(x) = B(x) \hbox{ for almost all } x\in \Bbb {R}^d
591: \end{equation}
592: and
593: \begin{equation}\label{1.10}
594: \lim_{k\to\infty}\int_E u^{k}(y)\, dy= \int_E B(y)\,dy
595: \end{equation}
596: for any bounded set $E\subset\Bbb {R}^d$.
597: Set $U^k(x) = (2\pi)^{-d}\int_{T^d} G(x-y) u^k(y)\, dy$ and
598: $u^k_0 = (2\pi)^{-d}\int_{T^d} u^k(y)\, dy$. Then \eqref{1.10} implies that
599: $u^k_0$ is convergent to\break $b_0 = (2\pi)^{-d}\int_{T^d} B(y)\, dy$ as
600: $k \to \infty$. By Lemma~\ref{l:1.0}--Lemma~\ref{l:1.2},
601: we have $S^k(x) = f_2(x) + U^k(x) -u^k_0|x|^2/(2d)$ is subharmonic in
602: $\Bbb {R}^d$.
603:
604: Note that $0\leq B(x) - u^k(x) \leq B(x) - u^1(x)$. Since $B(x)$ and $u^1(x)$
605: are locally integrable on $\Bbb {R}^d$, by Lemma~\ref{l:1.1}, \eqref{1.9}, and
606: the dominated convergence theorem,
607: \[
608: \lim_{k\to\infty} U^{k}(x) = U(x) = (2\pi)^{-d}\int_{T^d} G(x-y) B(y)
609: \, dy \quad\hbox{ in } L_{\rm loc}^1(\Bbb R^d)
610: \]
611: and hence there exists a subsequence, still called $U^k$ for notational
612: simplicity, such that
613: \[
614: \lim_{k\to\infty} U^k(x) = U(x)\quad\hbox{ a.e.}
615: \]
616: Since for any sequence of subharmonic functions convergent in $L^1$, there is a subharmonic function which is almost everywhere the $L^1$ limit of that sequence (see p. 20 of \cite{R}); $S(x) = f_2(x) + U(x)-b_0|x|^2/(2d)$ is almost everywhere equal to a subharmonic function $S_*(x)$ in $\Bbb R^d$.
617:
618: Similarly, there exists a sequence of nonincreasing lower semi-continuous
619: functions $v^k(x)$ on $\Bbb {R}^d$, which are also
620: in $L_{\rm loc}^1(\Bbb {R}^d)$, such that each $v^{k}(x)$ is bounded below and $v^{k}(x) \geq B(x)$,
621: \[
622: \lim_{k\to\infty} v^k(x) = B(x) \hbox{ for almost all } x\in \Bbb {R}^d
623: \]
624: and
625: \[
626: \lim_{k\to\infty}\int_E v^{k}(y)\, dy= \int_E B(y)\,dy
627: \]
628: for any bounded set $E\subset \Bbb {R}^d$. Since $-v^k(x)$ is
629: nondecreasing the above arguments show that there exists a superharmonic
630: function $S^{*}(x)$, which is almost everywhere equal to $S(x)$.
631:
632: Therefore $S_*(x) = S^*(x)$ almost everywhere. The subharmonicity of $S_*$
633: and superharmonicity of $S^*$ show that at every $x$
634: \begin{equation}\label{1.11}
635: S_*(x) \leq A_1 S_*(x) = A_1 S(x) = A_1 S^*(x) \leq S^*(x).
636: \end{equation}
637: In addition, if both $S_*(x)$ and $S^*(x)$ are finite, for any $\epsilon >0$, there exists $\delta>0$, such that
638: \begin{equation}\label{1.12}
639: S_*(y) \leq S_*(x) + \epsilon \hbox{ and } S^*(y) \geq S^{*}(x) -\epsilon
640: \end{equation}
641: for all $y\in B(x, \delta)$. Thus the fact that $S_* = S^*$ almost everywhere and \eqref{1.12} imply that $ S^*(x) \leq S_*(x) + 2\epsilon$.
642: So $S^*(x) \leq S_{*}(x)$. In fact, a similar argument shows that for all $x$, $S_*(x) > -\infty$ and $S^*(x) < \infty$ since $S_*(x) < \infty$ and $S^*(x) > - \infty$ by sub- or superharmonicity. Thus $S_*(x)$ and $S^*(x)$ are finite for all $x$ and $S^*(x) \leq S_*(x)$. Consequently, by \eqref{1.11} $S^*(x) = S_*(x)$ everywhere and hence it is harmonic in $\Bbb {R}^d$.
643:
644: But then,
645: \begin{eqnarray*}
646: \noalign{\vskip4pt}
647: S^*(x) & =& \frac{1}{v_d}\int_{|y-x|\leq 1} S^*(y)\, dy\\ \noalign{\vskip4pt}
648: & =& \frac{1}{v_d}\int_{|y-x|\leq 1} S(y)\, dy\\\noalign{\vskip4pt}
649: & =& \frac{1}{v_d}\int_{|y-x|\leq 1} f_2(y)\, dy +
650: \frac{1}{v_d}\int_{|y-x|\leq 1} U(y)\, dy \\\noalign{\vskip4pt}
651: && \qquad-\frac{b_0}{2d}
652: \frac{1}{v_d}\int_{|y-x|\leq 1} |y|^2\, dy\\\noalign{\vskip4pt}
653: & = &I + II + III.
654: \end{eqnarray*}
655: Since $f_2$ and $U$ are periodic, $I$ and $II$ are bounded. Thus, $S^*(x) =
656: O(|x|^2)$. By the penultimate inequality of Section 2.13 of \cite{PW} it follows that every second order partial derivative of $S^{*}$ is a bounded harmonic function and hence constant, so that $S^{*}$ itself is a quadratic polynomial. (An alternative argument can be based on expanding $S^*(x)$ into spherical harmonics.) Thus, the periodic function $f_{2} +U$ is almost everywhere equal to a quadratic polynomial $Q(x)=c_{1,0,\cdots,0}x^2_{1}+\cdots$. A simple countability argument shows that for almost every $x\in \Bbb R^d$ we have $(f_2+U)(x+2\pi ne_{1})=Q(x+2\pi ne_{1})$ for $n=1,2,3,\cdots$, where $e_1= (1,0,\cdots,0).$ Let $n\to\infty$ to see that $c_{1,0,\cdots,0}=0$. Similar reasoning shows that $Q(x)$ reduces to a constant $K$.
657: Consequently, we have $f_2(x) = -U(x) +K$ almost everywhere. However,
658: both $U$ and $F$ are integrable over $\Bbb T^d$. The integrals of $U$ and
659: $f_2$ over $\Bbb T^d$ are both $0$ by \eqref{1.2} and Lemma~\ref{l:1.1}.
660: So $K = 0$. Hence,
661: \begin{eqnarray*}
662: \noalign{\vskip-9pt}
663: f_2(x) & = & - (2\pi)^{-d}\int_{T^d}G(x-y)B(y)\, dy\quad {\rm a.e.} \\
664: & =& -(2\pi)^{-d}\int_{T^d}G(x-y)\tilde\Delta^*f_2(y)\, dy\quad {\rm a.e.}\\
665: \noalign{\vskip-16pt}
666: \end{eqnarray*}
667: Finally, by \eqref{1.2}, we have
668: \[
669: F(x) = -(2\pi)^{-d}\int_{T^d}G(x-y)\tilde\Delta^* f_2(y)\, dy\quad {\rm a.e.}
670: \]
671: Comparing the Fourier series of both sides, we see that the $a_{\xi}$
672: are the Fourier
673: coefficients of $\tilde\Delta^* f_2(x)- K_1$ for some constant $K_1$. The
674: Fourier series of the integrable function $\tilde\Delta^* f_2(x)- K_1$, $\sum
675: a_{\xi} e^{i\langle \xi, x\rangle}$ is Abel summable to $\tilde\Delta^*
676: f_2(x)- K_1$ almost everywhere (Theorem 2 of \cite{Sh2}). Thus, from the
677: definition of $f_*(x)$, $f_*(x) = \tilde\Delta^* f_2(x)- K_1$ almost
678: everywhere. Therefore $f_*(x) \in L^1(\Bbb {T}^d)$ and $a_{\xi}$ is the
679: $\xi^{\rm th}$ Fourier coefficient of $f_*(x)$. In fact, $K_1 = 0$ by
680: Lemma~\ref{l:1.0}. This completes the proof that
681: Lemma~\ref{l:1.2} will imply Theorem~\ref{t:1}.
682:
683: We end this section with the following observation. It is well known that if $u(x)$ is an
684: upper semi-continuous function in $B= B(x_0, h_0)\subset \Bbb T^d$ and $G_B$ denotes the Green function of $B$,
685: then when $d\geq 3$, the function
686: \[
687: U'(x) = \frac{1}{\sigma_d(d-2)}\int_{B} G_B(x, y) u(y)\, dy ,
688: \]
689: where $\sigma_d$ is the surface area of the unit ball in $\Bbb R^d$,
690: satisfies
691: \[
692: \tilde\Delta_{*} U'(x) \geq - u(x), \hbox{ for all } x\in B(x_0, h_0).
693: \]
694: Replacing $U$ everywhere by $U'$ in the proof of Theorem~\ref{t:1}, we
695: have the following lemma. Notice that we include the case where $a_0$ may not be zero.
696:
697: \proclaim{Lemma}\label{l:1.3} Let
698: $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$ be a
699: multiple $(d\geq 3)$ trigonometric series with
700: $\overline a_{\xi} = a_{-\xi}${\rm .} Suppose that the coefficients $a_{\xi}$
701: satisfy condition {\rm \eqref{e:bou}; }
702: \begin{itemize}
703: \ritem{1.} $f^*(x)$ and $f_*(x)$ are finite for all $x\in B$ where $B\subset
704: \Bbb T^d$ is a ball\/{\rm ;} and
705: \ritem{2.} $f_*(x) \geq A(x)$ for almost all $x\in B${\rm ,} where
706: $A(x)$ is in $L^1(B)${\rm .}
707: \end{itemize}
708: \noindent Then for any ball $B_1\subset \overline{B_1}\subsetneqq B${\rm ,} $f_*$ is in $L^1(B_1)${\rm .} Moreover{\rm , }
709: \[
710: \overline f_2(x) = f_2(x)+
711: \frac{1}{\sigma_d(d-2)}\int_{B_1}G_{B_1}(x, y) f_*(y)\, dy
712: + a_0|x|^2/(2d)
713: \]
714: is finite everywhere and is almost everywhere equal to a function $h(x)$
715: harmonic in $B_1${\rm .} In addition{\rm ,} if $f^*(x) = f_*(x)$
716: everywhere in $B${\rm ,} is in $L^1(B)${\rm ,} and is continuous in $B${\rm ,} then
717: \[
718: f_2(x) + \frac{1}{\sigma_d(d-2)}\int_{B}G_{B}(x, y) f_*(y)\, dy
719: + a_0|x|^2/(2d)
720: \]
721: is harmonic on $B${\rm .}
722: \endproclaim
723:
724: Note that under condition \eqref{e:bou},
725: \[
726: F(x) = -\sum_{\xi\neq 0}\frac{a_{\xi}}{|\xi|^2}e^{i\langle \xi, x\rangle}
727: \]
728: is in $L^{2}(\Bbb T^d)$ and $F(x) = f_2(x)$ almost everywhere in $\Bbb
729: T^d$. Combining the above lemma with Lemma~5 of Shapiro \cite{Sh1}, we have
730: the following analogue of Lemma~3 of Shapiro \cite{Sh4}.
731:
732: Let $B^{o}$ denote the interior of $B$.
733:
734: \proclaim{Lemma}\label{l:1.4} Let
735: $\sum_{\xi\in Z^d} a_{\xi} e^{i\langle x, \xi \rangle}$
736: be a multiple trigonometric series
737: with
738: $\overline a_{\xi} = a_{-\xi}${\rm .} Suppose that the coefficients $a_{\xi}$
739: satisfy condition {\rm \eqref{e:bou},}
740: \begin{itemize}
741: \ritem{1.} $f^*(x)$ and $f_*(x)$ are finite for all $x\in B$ where $B\subset
742: \Bbb T^d$ is a ball {\rm (}\/open or closed\/{\rm ),} and
743: \ritem{2.} $f_*(x) =0$ for almost all $x\in B${\rm .}
744: \end{itemize}
745: \noindent Then for any ball $B_1 \subset \overline{B_1}\subset B^{o}$, $f(x,t)$ converges to $0$ as
746: $t
747: \to 0^+$ uniformly in $B_1${\rm .} In particular{\rm ,} $f^*(x) = f_*(x) = 0$ in
748: $B${\rm . }
749: \endproclaim
750:
751:
752:
753: \section{Proof of Lemma~\ref{l:1.2}}\label{s:2}
754:
755: The proof of Lemma~\ref{l:1.2} is so difficult that this section will be given the following preface.
756:
757: The proof of Lemma~\ref{l:1.2} is extremely delicate, incorporating all the subtle ideas from Bourgain's landmark work \cite{B} as well as an additional Baire category argument that overcomes the unpleasant fact that an upper semi-continuous function on a compact set
758: need not be uniformly upper semi-continuous. Some of the difficulty is pushed into Lemmas~\ref{l:2.1} and
759: \ref{l:2.2}. The proof of Lemma~\ref{l:2.1} contains a great deal of hard analysis. Even after so much of the
760: work in Lemmas~\ref{l:2.1} and \ref{l:2.2} has been hidden, the reasoning involved in the proof of Lemma~\ref{l:1.2} is
761: still tortuous, and so we will provide an overview here.
762:
763: We assume that the set $W$ where $S$ fails to be upper semi-continuous is nonempty and then reason down a path which eventually divides into two paths each ending in a contradiction. First, a Baire category argument produces a nonempty portion $Z$ of $W$ ($Z=W\cap B$ for some ball $B$) such that $S$ restricted to $Z$ is ``very good," $f_2$ restricted to $Z$ is ``very good,"
764: et~cetera.
765:
766: Next, for each $\varepsilon>0$, let $W_\varepsilon$ be the points of $W$ where $S$ has a jump of at least~$\varepsilon$:
767: \[
768: \limsup_{y\to x} S(y)\geq S(x)+\varepsilon, \hbox{ for every } x\in W_\varepsilon.
769: \]
770: For each $\varepsilon>0$ and each $x\in B\setminus \overline W_{\varepsilon}$ consider the harmonic measure $\omega$ of $\partial W_\varepsilon$ with respect to $B\setminus \overline W_\varepsilon$ at $x$. Our path splits depending on whether the harmonic measure is ``thin:"
771: \begin{equation}\label{I}
772: \omega (B\setminus\overline W_\varepsilon, \partial W_\varepsilon, x) = 0
773: \hbox{ for all pairs } (\varepsilon, x) \hbox{ with } \varepsilon >0 \hbox{ and } x\in B\setminus\overline W_\varepsilon,
774: \end{equation}
775: or whether it is ``thick:"
776: \begin{equation}\label{II}
777: \omega (B\setminus\overline W_\varepsilon, \partial W_\varepsilon, x) > 0
778: \hbox{ for some } \varepsilon >0 \hbox{ and some } x\in B\setminus\overline W_\varepsilon.
779: \end{equation}
780:
781: If \eqref{I} holds, from Lemma~\ref{l:2.1} it follows that $S$ is bounded above and Lemma~\ref{l:2.2} then applies and asserts that $W\cap B=\emptyset$, a contradiction.
782:
783: On the other hand, if \eqref{II} is the case, we apply a second Baire category argument to strengthen assumption \eqref{II} by producing an $\varepsilon>0$ and a subset $Z_\varepsilon$ of $W_\varepsilon$ so that $U$ is ``uniformly $\varepsilon/40$ subharmonic" when restricted to $Z_\varepsilon$. Furthermore, the set $Z_\varepsilon$ is still ``thick:" $\omega (B\setminus\overline Z_\varepsilon, \partial Z_\varepsilon, x) >0$.
784:
785: Finally a very careful procedure involving picking balls within balls within balls
786: is used to find a point $p_1$ of $W_\varepsilon$ and a very nearby point $p_2$ of $B$ so that $S(p_2)-S(p_1)$ is small
787: relative to $\varepsilon$ because of Lemma~\ref{l:2.1}, but large relative to $\varepsilon$ because of $S$ having large
788: (relative to $\varepsilon$) jumps at each point of $W_\varepsilon$. This contradiction will complete the proof of
789: Lemma~\ref{l:1.2} which we begin here.
790:
791: Since $\tilde\Delta^* S(x) \geq 0$ and $S(x)$ is in $L^1$ locally, we have $S(x)<\infty$ for all $x\in \Bbb R^d$.
792: $S(x)\not\equiv -\infty$ since $S(x)$ is in $L^1$ locally.
793: We first show that $S(x) = f_2(x) + \overline U(x)$ is
794: upper semi-continuous in $\Bbb {R}^d$.
795:
796: Let
797: \begin{equation}\label{2.1}
798: W_{\varepsilon} = \left\{x\in \Bbb R^d: \sup_{|x-y| <\delta}
799: S(y)-S(x) > \varepsilon \hbox{
800: for all } \delta > 0\right\}.
801: \end{equation}
802: Then the set where $S(x)$ in $\Bbb {R}^d$ is not upper semi-continuous is
803: given by
804: \[
805: W =\bigcup_{\varepsilon >0} W_\varepsilon .
806: \]
807:
808: If $W = \emptyset$, then $S(x)$ is upper-semicontinuous. Now we assume
809: $W\ne \emptyset$ and construct the set $Z$. Bourgain's condition \eqref{e:bou} implies that $f(x,t)=\sum a_{\xi} e^{i\langle \xi,x\rangle-|\xi|t}$ is a uniform limit of its partial sums and hence is continuous on $\Bbb {T}^d\times [\frac1j,\infty)$ for every positive integer $j$. Taking periodicity into account, we see that for each $k$, $f(x,t)$ is uniformly continuous on $\Bbb {R}^d\times [\frac{1}{k+1},\frac1k]$. So we may partition $[\frac12,1]$ into $1=t_1>t_2>\cdots>t_r=\frac12$ so that for $i=1,2,\cdots,r-1$,
810: \begin{equation}\label{e:locdelta}
811: \sup_{ x\in R^d}\,\sup_{t_i\ge t\ge t_{i+1}} |f(x,t_i)-f(x,t)|\le 1\ .
812: \end{equation}
813: Then partition $[\frac13,\frac12]$ into $\frac12=t_r>t_{r+1}>\cdots>t_s=\frac13$ so that inequality \eqref{e:locdelta} holds for $i=r,r+1,\cdots,s-1$ and so on, thereby producing a sequence
814: ${\cal T} = \{t_{n}\}$ satisfying $1=t_1>t_2>\cdots$, $\lim_{k\to\infty} t_k =0$, and
815: \begin{equation}\label{e:locdelta1}
816: \sup_{ x\in R^d}\,\sup_{t_i\ge t\ge t_{i+1}} |f(x,t_i)-f(x,t)|\le 1
817: \end{equation}
818: holds for all $k$. Since $f(x,t)$ is bounded as $t\to 0^+$ for each $x\in\Bbb R^d$,
819: \[
820: \bigcup_{n\geq 1}\bigcap_{t\in {\cal T}}\left\{x\in\Bbb
821: R^d: |f(x, t)|\leq n \right\} =
822: \Bbb R^d.
823: \]
824: Therefore,
825: \[
826: \bigcup_{n\geq 1}\bigcap_{t\in {\cal T}} \left\{x\in
827: \overline W: |f(x, t)|\leq n \right\}
828: = \overline W.
829: \]
830: Since for each positive integer $n$ and each $t\in{\cal T}$, the set
831: \[
832: \left\{x\in \overline W:
833: |f(x, t)|\leq n \right\}
834: \]
835: is relatively closed with respect to $\overline W$, by Baire's
836: category theorem applied to the space $\overline W$ (the intersection of countably many relatively open dense sets
837: is not empty), for some $N_0\geq 1$,
838: \[
839: \bigcap_{t\in {\cal T}}\left\{x\in \overline W: |f(x, t)| \leq N_0 \right\}
840: \]
841: has a nonempty interior relative to $\overline W$. This means that there exist an
842: open ball $B(p, \rho_0), p\in W$, and a constant $N_0$ such that
843: \begin{equation}\label{2.06}
844: \sup_{t\in {\cal T}} \,\sup_{x\in B(p, \rho_0)\cap\overline W} \,
845: |f(x, t)|\leq N_0 <\infty.
846: \end{equation}
847: Bourgain's condition implies that $\sup_{x\in\Bbb R^d} |f(x,t)|\le C$, $\sup_{x\in\Bbb R^d}|f_1(x,t)|\le C$, and $\sup_{x\in\Bbb R^d}|f_2(x,t)|\le C$ whenever $t\ge 1$. Use
848: \[
849: f_1(x,t)=\int_{1}^{t} f(x,s)\, ds +f_1(x,1),
850: \]
851: \eqref{e:locdelta1} and \eqref{2.06} to see that there is a constant $N>0$ such that
852: \begin{equation}\label{l:2.01-1}
853: \sup_{x\in Z\atop t>0} |f_1 (x,t)|\le N,
854: \end{equation}
855: where $Z=B(p, \rho_0)\cap\overline W$.
856: Similarly, since
857: \[
858: f_2(x,t)=-\int_{0}^{t} f_1(x,s)\, ds +f_2(x)
859: \]
860: by \eqref{1.001},
861: \begin{equation}\label{2.3}
862: \sup_{{x\in Z\atop t>0}}|f_2(x) - f_2(x, t)|\leq N t.
863: \end{equation}
864: Therefore, $f_2(x)$ is continuous when restricted to $Z$. It follows that $S(x)$ is upper semi-continuous restricted to $Z$.
865:
866: We will show a contradiction if $W\ne \emptyset$. Once $S$
867: is everywhere upper semi-continuous, it is subharmonic since $\tilde\Delta_* S\geq 0$. For this, see p.14 of \cite{R}. This
868: will complete the proof of Lemma~\ref{l:1.2}.
869:
870: The following lemmas are needed in proving $W = \emptyset$.
871:
872: \smallbreak
873:
874: For a bounded open set $G$ and Borel measurable set $F$, we denote $\omega (G, F, x)$ to be the harmonic measure of a Borel
875: set
876: $F$ relative to $G$ at $x\in G$. Harmonic measure is closely related to Brownian motion. Let $(\{X_t\}_t, {\cal F}_t, P)$
877: be the standard Brownian motion in $\Bbb {R}^d$. For $x\in G$, let $T$ be the exiting time of $X_t$ from~$G$:
878: \[
879: T = \inf\{t\geq 0: X_t\notin G\}.
880: \]
881: Then $X_T\in \partial G$ since $X_t$ is continuous in $t$. Let $P^x$
882: denote the probability measure such that $X_0 = x$ almost everywhere.
883: Then the harmonic measure $\omega (G, F, x) = P^x(X_T\in F)$.
884:
885: The following properties of harmonic measure are well-known. We summarize
886: them as a preliminary lemma.
887:
888:
889:
890: \proclaim{Lemma}\label{l:2.0} Let $F_0\subset F_1\subset F_2$ be closed subsets of a
891: bounded open set~$G${\rm .} Then for $x\in G\setminus F_2${\rm ,}
892: \begin{eqnarray}
893: \quad \omega (G\setminus F_2, \partial F_2, x)&\geq &\omega (G\setminus F_1, \partial F_1, x)\\
894: \noalign{\vskip4pt}
895: &\geq&
896: \omega (G\setminus F_1, \partial F_0, x)\geq\omega (G\setminus F_2, \partial F_0, x).
897: \nonumber
898: \end{eqnarray}
899: \endproclaim
900:
901: To see the last inequality, let $T_i = \inf\{t\geq 0: X_t\notin G\setminus F_i\},\, i = 1, 2$. Then $T_2\leq T_1$. Note
902: that on
903: $\{ T_2< T_1\}$, we must have $X_{T_2} \in \overline G\setminus F_1$. Otherwise, $X_{T_2}\in \overline G\setminus
904: (\overline G \setminus F_1) = F_1$. Thus by definition $T_2 \geq T_1$, a contradiction. But $\{T_2< T_1\}\subset \{
905: X_{T_2}\in
906: \overline G\setminus F_1\}$ implies that $\{ X_{T_2}\in F_1\} \subset \{
907: T_1 = T_2\}$. Consequently, $\{ X_{T_2}\in \partial F_0\}\subset \{
908: X_{T_2}\in F_0\}\subset \{
909: X_{T_2}\in F_1\} \subset \{ T_1 = T_2\}$. This proves the inequality since
910: \[
911: P^x(X_{T_2}\in \partial F_0) = P^x(X_{T_2}\in \partial F_0, T_1 = T_2) \leq
912: P^x(X_{T_1}\in \partial F_0).
913: \]
914:
915: The middle inequality is simply the monotonicity of the harmonic measure.
916: To see the left inequality, observe that on $\{ X_{T_1}\in \partial F_1\}$, $X_{T_2} \in \partial F_2$. Otherwise,
917: $X_{T_2}\in
918: \partial G$ and $T_2\leq T_1$ imply that $X_{T_1}\in \partial G$, a contradiction. Consequently, $\{ X_{T_1}\in \partial
919: F_1\} \subset \{ X_{T_2}\in \partial F_2\}$. This completes the proof.
920:
921: The next three lemmas are essential to the proof of Lemma~\ref{l:1.2}.
922:
923:
924: \proclaim{Lemma}\label{l:2.1} Let $S${\rm ,} $f_2${\rm ,} and $\overline U$ be as given
925: in Lemma~{\rm \ref{l:1.2}.} Let $W$ be the set where $S$ is not upper
926: semi\/{\rm -}\/continuous{\rm .}
927: Assume that there is a open ball $B(p, \rho_0)${\rm ,} $p\in W${\rm ,}
928: such that when restricted to $Z =B(p, \rho_0)\cap\overline W${\rm ,}
929: $f_2(x)$ is continuous and {\rm \eqref{2.3}} holds{\rm .} Then{\rm ,}
930: for $p_1\in W, B(p_1,\rho_1) \subset B(p, \frac12\rho_0)$ and $p_2\in
931: B(p_1, \frac12\rho_1)${\rm ,} there exists a constant $c>0$ such that for almost all such $\rho_1${\rm ,}
932: \begin{eqnarray}\label{2.4}
933: \qquad \enspace S(p_2) - S(p_1) &\leq & c\biggl( \left[|f_2(p_1)|
934: +\rho_1^{-\frac34(d-1)}\right]\\ \noalign{\vskip4pt}
935: && \times\ [1-\omega (B(p_1, \rho_1)\setminus\overline
936: W, \partial(W\cap B(p_1, \rho_1)), p_2)]^{\frac14}\nonumber \\ \noalign{\vskip4pt}
937: && +\ \sup_{q\in B(p_1, 2\rho_1)\cap\overline W} \, |f_2(q)- f_2(p_1)|\nonumber\\ \noalign{\vskip4pt}
938: && +\ 2\sup_{q\in B(p_1, 2\rho_1)}\, \left(\overline U(q)
939: -\overline U(p_1)\right)\biggr). \nonumber
940: \end{eqnarray}
941: \endproclaim
942:
943: Lemma~\ref{l:2.1} is a one-sided version of Bourgain's key lemma in
944: \cite{B}. The proof is also similar and is given in Section~\ref{s:3}. It
945: follows from Lemma~\ref{l:2.1} that $S(x)$ is bounded from above in $B(p,
946: \frac{\rho_0}{4})$ when $p_1 =p$.
947:
948:
949:
950: \proclaim{Lemma}\label{l:2.2} Assume $\overline U$ is defined on
951: $\overline B(p, r)$ and is upper semi\/{\rm -}\/continuous on $B(p, r)${\rm .} Let $f_2$ be
952: a function in $\overline B(p, r)$ such
953: that $S(x)= f_2(x) + \overline U(x)$ is bounded from above in $\overline B(p,
954: r)${\rm ,} in $L^1(\overline B(p, r))${\rm ,} and satisfies
955: \begin{equation}\label{2.5}
956: \tilde\Delta^* S(x) \geq 0, \hbox{ and } \tilde\Delta_* f_2(x) <\infty
957: \end{equation}
958: for each $x\in B(p, r).$
959: If $S$ is upper semi\/{\rm -}\/continuous when restricted to
960: $\overline W= \{x \in B(p, r): $S(x)$ \hbox{ is not upper
961: semi\/{\rm -}\/continuous}\}${\rm ,} and for all $x\in B(p, r)\setminus\overline W_{\varepsilon}$ the harmonic measure
962: \begin{equation}\label{2.7}
963: \omega(B(p, r)\setminus\overline W_{\varepsilon}, B(p,r)\cap W_{\varepsilon}, x) = 0
964: \end{equation}
965: for all $\varepsilon>0$ where $W_{\varepsilon}$ is given by {\rm \eqref{2.1},}
966: then $W$ must be empty and $S(x)$ is subharmonic on $B(p, r)${\rm . }
967: \endproclaim
968:
969:
970:
971: The proof of Lemma~\ref{l:2.2} is given in Section~\ref{s:4}. The special
972: case when $\overline U\equiv 0$ was proved by Bourgain.
973:
974: The next lemma provides a harmonic measure version of a point density.
975:
976:
977:
978: \proclaim{Lemma}\label{l:2.3} Let $B(p_0, r)$ be a ball in $\Bbb R^d$ and
979: $F$ a closed set such that $B(p_0, r)\cap F\ne\emptyset${\rm .} Suppose for some
980: $x\in B(p_0, r)\setminus F${\rm , }
981: \[
982: \omega(B(p_0, r)\setminus F, \partial (B(p_0, r)\cap F), x) >0.
983: \]
984: Then there exists $p_1\in B(p_0, r)\cap F${\rm ,} such that
985: \begin{equation}\label{2.8}
986: \inf_{\delta_1> 0}\liminf_{\delta_2\to 0}\inf_{x\in B(p_1,
987: \delta_2)} \, \omega(B(p_1, \delta_1)\setminus F, \partial (B(p_1,
988: \delta_1)\cap F), x) = 1.
989: \end{equation}
990: \endproclaim
991:
992: The proof of Lemma~\ref{l:2.3} is outlined in \cite{B}. For a detailed proof,
993: see the proof of Theorem 3.14 in \cite{AW}.
994:
995: We now return to the proof of Lemma~\ref{l:1.2}.
996: By \eqref{1.2} and the fact that
997: $F(x) \in L^2(\Bbb T^d)$, we have $S$ is in $L^1_{\rm loc}(\Bbb {R}^d)$.
998: For the duration of this proof, we abbreviate $B(p, \rho_0/8)$ to $B$. There are two cases.
999: \medbreak
1000: Case one: for all $\epsilon > 0$,
1001: \[
1002: \omega(B\setminus \overline {W_{\epsilon}},
1003: \partial (B\cap W_{\epsilon}), x) = 0
1004: \]
1005: for all $x\in B\setminus \overline {W_{\epsilon}}$. Then by Lemma~\ref{l:2.1}, $S$ is bounded
1006: from above and by Lemma~\ref{l:1.0}, $\tilde\Delta_{*} f_2<\infty$ everywhere. Also $B(p, p_0)$ was chosen so that $S$
1007: is upper semi-continuous when restricted to $B(p,p_0)\cap \overline W$. Thus all the hypotheses of Lemma~\ref{l:2.2} are
1008: satisfied and $W\cap B = \emptyset$, which is a contradiction.
1009: \medbreak
1010: Case two: for some $\epsilon > 0$ and for some
1011: $x_0\in B\setminus \overline {W_{\epsilon}}$, we have
1012: \begin{equation}\label{2.10}
1013: \omega(B\setminus \overline {W_{\epsilon}},
1014: \partial (B\cap W_{\epsilon}), x_0) > 0.
1015: \end{equation}
1016:
1017: Even though $\overline U$ is upper semi-continuous everywhere, it may
1018: not be uniformly upper semi-continuous on $W_{\epsilon}$. This presents
1019: a problem which did not arise at the corresponding point in Bourgain's
1020: proof. To deal with this, we now introduce a subset of $W_{\epsilon}$
1021: called $Z_{\epsilon}$, on a portion of which there holds a kind of
1022: uniform upper semi-continuity.
1023:
1024: Let
1025: \[
1026: Z_\epsilon = \{y\in B\cap W_\epsilon:
1027: \omega(B\setminus \overline {W_{\epsilon}},
1028: \overline{B(y, \delta)\cap W_{\epsilon}}, x_0)>0, \hbox{ for all }
1029: \delta>0\}.
1030: \]
1031: Then,
1032: \begin{equation}\label{2.11}
1033: \omega(B\setminus \overline {W_{\epsilon}},
1034: \partial Z_\epsilon, x_0)>0.
1035: \end{equation}
1036: In fact, by definition, for each $z\in B\cap
1037: W_\epsilon\setminus \overline {Z_\epsilon}$, there exists a ball
1038: $B(z, \delta_z)$, such that
1039: \begin{equation}\label{2.12}
1040: \omega(B\setminus \overline {W_{\epsilon}},
1041: \overline{B(z, \delta_z)\cap W_{\epsilon}}, x_0)=0.
1042: \end{equation}
1043: The open cover $\{B(z, \delta_z)\}$ of $B\cap
1044: W_\epsilon\setminus \overline {Z_\epsilon}$ has a countable subcover $\{B(z_i,\delta_{z_i})\}$.
1045: Thus, \eqref{2.12} implies
1046: \[
1047: \omega(B\setminus \overline {W_{\epsilon}},
1048: B\cap W_\epsilon\setminus \overline {Z_\epsilon}, x_0)=0 .
1049: \]
1050: So \eqref{2.11} follows from \eqref{2.10} as $\omega(B\setminus \overline {W_{\epsilon}},
1051: \overline {Z_\epsilon}, x_0)=\omega(B\setminus \overline {W_{\epsilon}},
1052: \partial {Z_\epsilon}, x_0)$.
1053:
1054: Since $\overline U$ is upper semi-continuous,
1055: \[
1056: \bigcup_{m\geq 1}\left\{y\in \overline {Z_\epsilon}: \sup_{|z-y|\leq
1057: 2/m} \overline U(z) -\overline U(y) \leq \frac{\epsilon}{40}\right\} =
1058: \overline {Z_\epsilon}.
1059: \]
1060: Apply Baire's category theorem to the space $\overline {Z_\epsilon}$ to see that there exists $m\geq 1$ and an open ball
1061: $B(q, \rho)\subset B, q\in Z_\epsilon$,
1062: such that
1063: \[
1064: B(q, \rho)\cap \overline {Z_\epsilon} \subset \overline{\left\{y\in
1065: \overline {Z_\epsilon}: \sup_{|z-y|\leq 2/m}
1066: \overline U(z) -\overline U(y) \leq \frac{\epsilon}{40}\right\}}.
1067: \]
1068: Equivalently, for any fixed $y\in B(q, \rho)\cap \overline {Z_\epsilon}$,
1069: there exists a sequence $y_n\in B(q, \rho)\cap Z_\epsilon$ convergent to $y$ such that
1070: \begin{equation}\label{2.13}
1071: \sup_{|z-y_n|\leq 2/m} \overline U(z) - \overline U(y_n) \leq
1072: \frac{\epsilon}{40}.
1073: \end{equation}
1074: However, $\overline U$ is upper semi-continuous. So there exists $0< \delta <
1075: \frac{1}{m}$ such that for $|y_n-y|< \delta$,
1076: \[
1077: \overline U(y_n) - \overline U(y) < \frac{\epsilon}{40}.
1078: \]
1079: Thus, for $|z-y|< \frac{1}{m}$, since $|z-y_n|<\frac{2}{m}$ if
1080: $|y_n-y|<\delta$,
1081: \begin{equation}\label{2.14}
1082: \sup_{|z-y|<\frac{1}{m}}\overline U(z) < \overline U(y_n) +
1083: \frac{\epsilon}{40} < \overline U(y) +\frac{\epsilon}{20}.
1084: \end{equation}
1085: Without loss of generality, we assume that $\frac{1}{m} \leq
1086: \frac{\rho}{2}$.
1087:
1088: Because $q\in Z_\epsilon$, we also have
1089: \begin{equation}\label{2.15}
1090: \omega(B\setminus \overline {W_{\epsilon}},
1091: \overline {B(q, \frac{\rho}{2})\cap W_\epsilon}, x_0)=\omega(B\setminus \overline {W_{\epsilon}},
1092: \partial (B(q, \frac{\rho}{2})\cap W_\epsilon), x_0) >0.
1093: \end{equation}
1094:
1095: Set $F_\epsilon = B(q, \frac{\rho}{2})\cap W_\epsilon$. Then the rightmost inequality of
1096: Lemma~\ref{l:2.0} and \eqref{2.15} imply that
1097: \[
1098: \omega(B\setminus \overline F_{\epsilon},
1099: \partial F_\epsilon, x_0) \geq \omega(B\setminus \overline
1100: W_{\epsilon}, \partial F_\epsilon, x_0) >0.
1101: \]
1102:
1103: From Lemma~\ref{l:2.3}, there exists $p'\in \overline F_{\epsilon}$
1104: such that
1105: \begin{equation}\label{2.16}
1106: \inf_{\delta_1> 0}\underset{\delta_2\to 0}\liminf\inf_{x\in B(p',
1107: \delta_2)} \, \omega(B(p', \delta_1)\setminus\overline F_{\epsilon},
1108: \partial (B(p',\delta_1)\cap F_{\epsilon}), x) = 1.
1109: \end{equation}
1110:
1111: Notice that Lemma~\ref{l:2.3} requires the set $F$ to be closed, so we cannot be sure that $p'\in F_\varepsilon$.
1112: Although $F_\varepsilon$ may not be closed, the uniformity implied by \eqref{2.16} allows us to continue.
1113:
1114: Since $f_2$ restricted to $\overline W \cap B$
1115: is continuous, we may select $1/(8m) > \delta_1>0$ such that
1116: \begin{equation}\label{2.17}
1117: |f_2(z) - f_2(y)| \leq \frac{\varepsilon}{10}
1118: \end{equation}
1119: for all $y, z\in B(p', 8\delta_1)\cap \overline W$.
1120:
1121: Let $\eta>0$ be any positive number. From \eqref{2.16}, it follows that
1122: there exists $0< \delta_2 = \delta_2(\eta, \delta_1)< \delta_1$ such that
1123: \[
1124: \omega(B(p', \delta_1)\setminus \overline F_{\epsilon},
1125: \partial (B(p', \delta_1)\cap
1126: F_{\epsilon}), y) > 1-\eta\hbox{ for all }y\in B(p', \delta_2).
1127: \]
1128:
1129: We may also assume that $\delta_1+\delta_2 = \delta_3^{'}$ satisfies
1130: $B(p', \delta_3^{'})\subset B(p, \frac{\rho_0}{2})$. Pick any $\delta_3$ bigger than $\delta'_3$ but small enough to force $B(p', \delta_3) \subset B(p, \frac{\rho_0}{2})$. Note
1131: that
1132: $p'\in \overline F_\varepsilon$ implies that there exists
1133: $p_1\in B(p', \frac{\delta_2}{2})\cap F_\epsilon$. Since
1134: $B(p', \delta_1) \subset B(p_1, \delta_3)$,
1135: \[
1136: B(p_1, \delta_3)\setminus\overline{[F_\varepsilon\cup \{B(p_1, \delta_3)\setminus B(p', \delta_1)\}]}=B(p',\delta_1)\setminus\overline F_\varepsilon .
1137: \]
1138: So by the rightmost inequality of Lemma~\ref{l:2.0},
1139: \begin{eqnarray*}
1140: && \hskip-48pt \omega(B(p_1, \delta_3)\setminus \overline F_{\epsilon},
1141: \partial (B(p', \delta_1)\cap
1142: F_{\epsilon}), y)\\ \noalign{\vskip4pt}
1143: \quad & \geq& \omega (B(p_1, \delta_3)\setminus\overline{[F_\varepsilon\cup \{B(p_1, \delta_3)\setminus B(p',
1144: \delta_1)\}]},
1145: \partial (B(p', \delta_1)\cap F_{\epsilon}), y) \\ \noalign{\vskip4pt}
1146: \quad & =& \omega(B(p', \delta_1)\setminus \overline F_{\epsilon},
1147: \partial (B(p', \delta_1)\cap
1148: F_{\epsilon}), y)\\ \noalign{\vskip4pt}
1149: \quad &\geq& 1-\eta\hbox{ for all }y\in B(p_1, \frac{\delta_2}{2}).
1150: \end{eqnarray*}
1151: Consequently,
1152: \[
1153: \omega(B(p_1, \delta_3)\setminus \overline F_{\epsilon},
1154: \partial (B(p_1, \delta_3)\cap
1155: F_{\epsilon}), y) \geq 1-\eta\hbox{ for all }y\in B(p_1,
1156: \frac{\delta_2}{2}),
1157: \]
1158: since $B(p', \delta_1)\cap F_{\epsilon} \subset B(p_1,
1159: \delta_3)\cap F_{\epsilon}$.
1160: Finally, by the left inequality of Lemma~\ref{l:2.0}
1161: \[
1162: \omega(B(p_1, \delta_3)\setminus \overline W,
1163: \partial (B(p_1, \delta_3)\cap
1164: W), y) \geq \omega(B(p_1, \delta_3)\setminus \overline F_{\epsilon},
1165: \partial (B(p_1, \delta_3)\cap F_{\epsilon}), y).
1166: \]
1167: We therefore have
1168: \begin{equation}\label{2.18}
1169: \omega(B(p_1, \delta_3)\setminus \overline W,
1170: \partial (B(p_1, \delta_3)\cap
1171: W), y)> 1-\eta\hbox{ for all }y\in B(p_1,
1172: \frac{\delta_2}{2}).
1173: \end{equation}
1174: By definition, $p_1\in W_{\epsilon}$ implies that there exists $p_2\in
1175: B(p_1, \frac{\delta_2}{2})$ such that
1176: \[
1177: S(p_2)-S(p_1) \geq \frac{\varepsilon}{2}.
1178: \]
1179: Apply Lemma~\ref{l:2.1} at $p_1, p_2,$ and $\rho_1 = \delta_3$ where the
1180: inequality \eqref{2.4} holds for $\delta_3$. Then by \eqref{2.14},
1181: \eqref{2.17}, \eqref{2.18}, and the above inequality, we
1182: have
1183: \begin{equation}\label{2.19}
1184: \frac{\varepsilon}{2} \leq S(p_2) - S(p_1) \leq c[|f_2(p_1)| +
1185: \delta_3^{-\frac34 (d-1)}]\eta^{1/4} + \frac{\varepsilon}{5}.
1186: \end{equation}
1187: Note here that $p_1, p_2$, and $\delta_3$ depend on $\eta$. However, since $f_2$ is continuous and hence bounded on $B(p,
1188: \rho_0)\cap\overline W$ and $\delta_3$ is bounded below by $\delta_1$ as $\eta \to 0$, so \eqref{2.19} becomes a
1189: contradiction upon choosing $\eta$ sufficiently small.
1190:
1191: \section{Proof of Lemma~\ref{l:2.1}}\label{s:3}
1192:
1193: For any bounded measurable function $f(x)$ defined on $\partial G$,
1194: \begin{equation}\label{3.1}
1195: H_f(x) = \int_{\partial G} f(z)\, \omega(G, dz, x)
1196: \end{equation}
1197: is harmonic in $G$. If every point on $\partial G$ satisfies the exterior
1198: cone condition and $f$ is continuous at $x\in \partial G$, then
1199: \[
1200: \lim_{{y\rightarrow x\atop y\in G}} H_f(y)=f(x).
1201: \]
1202: Since any upper semi-continuous function is the limit of a decreasing sequence of continuous functions, so the maximum principle for subharmonic functions and \eqref{3.1} imply that
1203: \begin{equation}\label{3.1.1}
1204: f(x) \leq \int_{\partial G} f(z)\, \omega(G, dz, x)
1205: \end{equation}
1206: for any function $f$ subharmonic on an open set $\tilde G\supset \overline G \supset G$.
1207:
1208: We need only to consider $p_2\notin \overline W$. Let $\tau = \hbox{dist}(p_2, \overline W) \leq \frac12 \rho_1$. For $\kappa \ll \tau$, define
1209: \[
1210: G_\kappa = \{x\in B(p_1, \rho_1): \hbox{dist} (x, \overline W) < \kappa\}.
1211: \]
1212: Clearly $\overline W \cap B(p_1, \rho_1)\subset G_\kappa$. We know that
1213: $S$ is upper semi-continuous and $\tilde \Delta^* S(x) \geq 0$ on
1214: $B(p, \rho_0)\setminus\overline W$. This is the hypothesis of a
1215: classical theorem (see for example \cite[p.~14]{R}) which concludes that
1216: $S$ is subharmonic on $B(p, \rho_0)\setminus \overline W$. Thus,
1217: $S(x) - S(p_1)$ is subharmonic on $B(p, \rho_0)\setminus \overline
1218: W$. In particular, $S(x) - S(p_1)$ is subharmonic on an open set containing $\overline {B(p_1, \rho_1)
1219: \setminus \overline G_\kappa}$. Note that $B(p_1, \rho_1)\setminus \overline
1220: G_\kappa$ satisfies the exterior cone condition everywhere on the boundary.
1221: So by \eqref{3.1.1}, we
1222: have
1223: \begin{eqnarray*}
1224: \noalign{\vskip6pt}
1225: &&\hskip-.5in S(p_2) - S(p_1) \leq \int\limits_{\partial (B(p_1,
1226: \rho_1)\setminus \overline G_\kappa)}\,
1227: [S(x) - S(p_1)]\, \omega(B(p_1, \rho_1)\setminus
1228: \overline G_\kappa, dx, p_2)\\ \noalign{\vskip6pt}
1229: &&\quad =\ \int\limits_{\partial B(p_1, \rho_1)\setminus (B(p_1, \rho_1)\cap
1230: \partial G_{\kappa})}\,
1231: [S(x) - S(p_1)]\, \omega(B(p_1, \rho_1)\setminus
1232: \overline G_\kappa, dx, p_2)\\ \noalign{\vskip6pt}
1233: &&\qquad+ \ \int\limits_{B(p_1, \rho_1)\cap\partial G_\kappa}\,
1234: [S(x)-S(p_1)]\, \omega(B(p_1, \rho_1)\setminus
1235: \overline G_\kappa, dx, p_2)\\ \noalign{\vskip6pt}
1236: &&\quad =\ I_1 + I_2.
1237: \end{eqnarray*}
1238:
1239: We first estimate $I_1$. When $p_2\in B(p_1, \rho_1/2)$, a classical result on
1240: harmonic measure shows that $\omega =
1241: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa, dx, p_2)$ is absolutely
1242: continuous with respect to the surface Lebesgue measure $\sigma$ when
1243: restricted to the sphere $B(p_1, \rho_1)$. (See \cite{D} or
1244: (4.39) of \cite{AW}.) By \eqref{1.2}, $f_2(x) = F(x)$
1245: almost everywhere with respect to Lebesgue measure; thus, for almost every
1246: $\rho_1>0$, $f_2(x) = F(x)$ almost everywhere with respect to the surface Lebesgue measure on $B(p_1, \rho_1)$ and hence with respect to the harmonic measure $\omega$ for all $p_2\in B(p_1, \rho_1/2)$.
1247: Consequently,
1248: \begin{eqnarray*}
1249: \noalign{\vskip4pt}
1250: I_1 &\leq& \int\limits_{\partial B(p_1, \rho_1)} |F(p_2) - f_2(p_1)|\,
1251: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa, dx, p_2)\\ \noalign{\vskip4pt}
1252: &&+\ \underset{q\in B(p_1, \rho_1)}\sup \overline U(q) -\overline
1253: U(p_1)\\ \noalign{\vskip4pt}
1254: & =& I_3 + I_4.
1255: \end{eqnarray*}
1256: A result of Bourgain \cite{B} (see also Lemma 4.5 of \cite{AW}) shows that
1257: \[
1258: I_3 \leq c\left( \left[|f_2(p_1)|
1259: +\rho_1^{-\frac34(d-1)}\right] [1-\omega (B(p_1, \rho_1)\setminus\overline
1260: W, \partial(W\cap B(p_1, \rho_1)), p_2)]^{\frac14}\right).
1261: \]
1262: \demo{{R}emark\/ {\rm 4.1}}
1263: In fact, Lemma 4.5 of \cite{AW} was based on Connes' condition
1264: \eqref{0.3}. But a careful reading of the proof shows the conclusion of
1265: Lemma 4.5 holds true under Bourgain's condition \eqref{e:bou} since
1266: inequalities (4.18) and (4.19) in Lemma 4.2 and Corollary 4.3
1267: respectively can be replaced by
1268: \[
1269: \sup_{k} \frac{1}{2^{k}}\sum_{|\xi|^2\sim 2^k}|c_{\xi}|^2 \leq M
1270: \]
1271: as they are used only in (4.21).
1272: \enddemo
1273:
1274: This gives the first half of \eqref{2.4}. Now we estimate $I_2$.
1275:
1276: For any $x\in B(p_1, \rho_1)\cap \partial G_{\kappa}$, there exists
1277: $\tilde x \in \overline W\cap B(p_1, 2\rho_1)$, such that $|x-\tilde x| =
1278: \kappa$. Since $S$ is subharmonic at $x$,
1279: \begin{equation}\label{3.2}
1280: S(x) \leq A_\kappa f_2(x) + A_\kappa \overline U(x).
1281: \end{equation}
1282: Since $\tilde x\in \overline W\cap B(p, \rho_0)$, by assumption,
1283: \begin{equation}\label{3.3}
1284: |f_2(\tilde x) - f_2(\tilde x, \kappa)| \leq N \kappa.
1285: \end{equation}
1286:
1287: Thus combining \eqref{3.2} and \eqref{3.3}, we have
1288: \begin{eqnarray*}
1289: S(x) - S(\tilde x) &\leq& A_{\kappa} f_2(x) -
1290: A_{\kappa} f_2(\tilde x, \kappa) + A_{\kappa} f_2(\tilde x, \kappa)-
1291: f_2(\tilde x, \kappa)\\
1292: && +\ A_{\kappa} \overline U(x)-\overline U(\tilde x) + N\kappa.
1293: \end{eqnarray*}
1294: Consequently,
1295: \begin{eqnarray*}
1296: S(x) - S(p_1) & = & S(x) - S(\tilde x) + S(\tilde x) - S(p_1)\\ \noalign{\vskip4pt}
1297: &\leq& A_\kappa f_2(x) - A_\kappa f_2(\tilde x, \kappa) +
1298: A_{\kappa} f_2(\tilde x, \kappa)- f_2(\tilde x, \kappa)\\ \noalign{\vskip4pt}
1299: && +\ f_2(\tilde x) -f_2(p_1) + A_\kappa\overline U(x) - \overline
1300: U(p_1) + N\kappa\\ \noalign{\vskip4pt}
1301: &\leq& A_\kappa f_2(x) - A_\kappa f_2(\tilde x, \kappa) + A_{\kappa}
1302: f_2(\tilde x, \kappa)- f_2(\tilde x, \kappa)\\ \noalign{\vskip4pt}
1303: && +\ \underset{q\in B(p_1, 2\rho_1)\cap \overline W}\sup |f_2(q)-
1304: f_2(p_1)|\\ \noalign{\vskip4pt}
1305: && +\ \underset{q\in B(p_1, 2\rho_1)}\sup \overline U(q)-
1306: \overline U(p_1) + N\kappa.
1307: \end{eqnarray*}
1308: From \eqref{1.2} and the definition of $A_\rho f_2(x)$, we have $A_\rho
1309: f_2(x) = A_\rho F(x)$ for all $x$. Thus
1310: \begin{eqnarray}\label{3.4}
1311: %
1312: && \\
1313: I_2 &\leq& \int\limits_{B(p_1, \rho_1)\cap \partial G_\kappa}
1314: |A_\kappa F(x) - A_\kappa f_2(\tilde x, \kappa)| \,
1315: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa,
1316: dx, p_2)\nonumber \\ \noalign{\vskip4pt}
1317: &&+ \ \int\limits_{B(p_1, \rho_1)\cap \partial G_\kappa}
1318: |A_\kappa f_2(\tilde x, \kappa) - f_2(\tilde x, \kappa)| \,
1319: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa,
1320: dx, p_2)\nonumber \\ \noalign{\vskip4pt}
1321: && +\ \underset{q\in B(p_1,
1322: 2\rho_1)\cap \overline W} \sup |f_2(q)-f_2(p_1)|\nonumber \\
1323: && +\ \underset{q\in B(p_1, 2\rho_1)}\sup
1324: \overline U(q)- \overline U(p_1) + N\kappa\nonumber\\ \noalign{\vskip4pt}
1325: & = &I_5 + I_6 + \underset{q\in B(p_1,
1326: 2\rho_1)\cap \overline W} \sup |f_2(q)-f_2(p_1)|\nonumber \\ \noalign{\vskip4pt}
1327: &&+\ \underset{q\in B(p_1, 2\rho_1)}\sup
1328: \overline U(q)- \overline U(p_1) + N\kappa. \nonumber
1329: \end{eqnarray}
1330:
1331: It is enough to show that $I_5\to 0$ and $I_6\to 0 $ as $\kappa\to 0$.
1332: Observe that
1333: \begin{eqnarray*}
1334: I_5 &\leq& \int\limits_{B(p_1, \rho_1)\cap \partial G_\kappa}
1335: |A_\kappa F(x) - A_\kappa F(\tilde x)| \,
1336: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa,
1337: dx, p_2)\\
1338: && + \ \int\limits_{B(p_1, \rho_1)\cap \partial G_\kappa}
1339: |A_\kappa F(\tilde x) - A_\kappa f_2(\tilde x, \kappa)| \,
1340: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa,
1341: dx, p_2)\\
1342: & =& I_7 +I_8.
1343: \end{eqnarray*}
1344: A result of Bourgain \cite{B} (again, see also Lemma 4.4 of \cite{AW}),
1345: shows that $I_7\to 0$ as $\kappa\to 0$. However, the same proof of
1346: Lemma 4.4 in \cite{AW} shows that if $a_{\xi}$ satisfies Bourgain's
1347: condition \eqref{e:bou}, then
1348: \begin{equation}\label{3.5}
1349: \lim_{\kappa\to 0} \kappa \int\limits_{B(p_1, \rho_1)\cap \partial G_\kappa}
1350: \left|\sum_{|\xi|\neq 0} \frac{a_{\xi}}{|\xi|}\hat I(\kappa |\xi|)
1351: e^{i\langle \overline x, \xi\rangle}\right| \,
1352: \omega(B(p_1, \rho_1)\setminus \overline G_\kappa, dx, p_2) = 0,
1353: \end{equation}
1354: where $|\overline x - x|\leq \kappa$.
1355: Note that
1356: \[
1357: A_\kappa F(\tilde x) - A_\kappa f_2(\tilde x, \kappa) = -\sum_{|\xi|\neq 0}
1358: \frac{a_{\xi}}{|\xi|^2}\hat I(\kappa |\xi|)e^{i\langle\tilde x, \xi\rangle}
1359: (1-e^{-|\xi|\kappa}),
1360: \]
1361: while by the mean value theorem, for each $\xi\neq
1362: 0$, there exists $t_{\xi} > 0$ such that
1363: \[
1364: \sum_{|\xi|\neq 0} \frac{a_{\xi}}{|\xi|^2}\hat I(\kappa |\xi|)e^{i\langle
1365: \tilde x, \xi\rangle}(1-e^{-|\xi|\kappa}) = \kappa \sum_{|\xi|\neq 0}\frac{a_{\xi}
1366: e^{-|\xi|t_{\xi}}}{|\xi|}\hat I(\kappa |\xi|)
1367: e^{i\langle\tilde x, \xi\rangle}.
1368: \]
1369: Since $e^{-|\xi|t_{\xi}} < 1$, $\{a_{\xi}e^{-|\xi|t_{\xi}}\}$ satisfies
1370: Bourgain's condition \eqref{e:bou} as $\{a_{\xi}\}$ does. Thus by
1371: \eqref{3.5}, $I_8\to 0$ as $\kappa\to 0$. This shows that $I_5\to 0$ as
1372: $\kappa \to 0$.
1373:
1374: The method that Bourgain used to prove that $I_7\to 0$ as $\kappa\to 0$
1375: can also be used to prove $I_6\to 0$ as $\kappa\to0$. To establish this, we will use the following lemma of Bourgain \cite {B}. (See also the proof of Corollary~4.3 of \cite{AW}.)
1376: \proclaim{Lemma}\label{l:3.1}
1377: Let $k\geq 1, \gamma>0, \eta\leq 2^{-k}${\rm .} Let $E_{k, \gamma, \eta}$ be
1378: a set of $\eta$\/{\rm -}\/separated points $x\in B(p, q)\subset \Bbb R^d$
1379: satisfying
1380: \[
1381: \left|\sum_{|\xi|\sim 2^k} \frac{b_{\xi}}{|\xi|^2} e^{i\langle x,
1382: \xi\rangle}\right|\geq \gamma.
1383: \]
1384: Then{\rm ,} the cardinality of $E_{k,\gamma, \eta}$ satisfies
1385: \[
1386: |E_{k,\gamma,\eta}|\leq c \gamma^{-2}\eta^{-d}2^{-2k}\nu_k^2,
1387: \]
1388: where $c$ is an absolute constant and
1389: \[
1390: \nu_k^2 = 2^{-2k}\sum_{|\xi|\sim 2^{k}} |b_{\xi}|^2.
1391: \]
1392: \endproclaim
1393: Let
1394: \[
1395: \alpha_{k} = \left\{ \begin{array}{ll} c\alpha [\log(1+2^k \kappa)]^{-2}, &
1396: \hbox{ for } 2^k\geq \kappa^{-1}\\
1397: c\alpha [\log(1+2^{-k} \kappa^{-1})]^{-2}, &
1398: \hbox{ for } 2^k < \kappa^{-1} .\end{array} \right.
1399: \]
1400: The positive constant $c$ is chosen so that $\sum_{k\geq 1}\alpha_{k}
1401: \leq 2c\alpha\sum_{n\geq 0}(\log(1+2^n))^{-2}\break = \alpha$ for all $\alpha> 0$. Clearly, $c$ is an absolute constant.
1402: For $\alpha >0$, let
1403: \[
1404: S_{\kappa, \alpha} = \{ x\in B(p_1, \rho_1)\cap \partial G_{\kappa}:
1405: |A_{\kappa}f_2(\tilde x, \kappa) - f_2(\tilde x, \kappa)| > \alpha\}.
1406: \]
1407: Then
1408: \begin{equation}\label{3.6}
1409: I_6 = \int_{0}^{\infty}\omega(B(p_1, \rho_1)\setminus\overline
1410: G_{\kappa}, S_{\kappa, \alpha}, p_2)\, d\alpha.
1411: \end{equation}
1412:
1413: Let
1414: \begin{eqnarray*}
1415: S_{\kappa, k,\alpha_{k}} & = & \{x\in B(p_1, \rho_1)\cap \partial G_{\kappa}:
1416: |A_{\kappa}f_{2, k}(\tilde x, \kappa) - f_{2, k}(\tilde x, \kappa)| >
1417: \alpha_{k}\},\\ \noalign{\vskip4pt}
1418: S'_{\kappa, k, \alpha_{k}} & =& \{x\in B(p_1, \rho_1):
1419: |A_{\kappa}f_{2, k}(x, \kappa) - f_{2, k}(x, \kappa)| >
1420: \alpha_{k}\},
1421: \end{eqnarray*}
1422: where
1423: \begin{eqnarray*}
1424: f_{2, k}(x, \kappa) & = & \sum_{|\xi|\sim 2^k} \frac{a_{\xi}}{|\xi|^2}
1425: e^{i\langle x, \xi\rangle-\kappa|\xi|}, \hbox{ and} \\ \noalign{\vskip4pt}
1426: A_{\kappa}f_{2, k}(x, \kappa)& = & \sum_{|\xi|\sim 2^k} \frac{a_{\xi}}{|\xi|^2}\hat I(\kappa|\xi|)
1427: e^{i\langle x, \xi\rangle-\kappa|\xi|}.
1428: \end{eqnarray*}
1429: Then
1430: \begin{equation}\label{3.7}
1431: S_{\kappa, \alpha} \subset \bigcup_{k\geq 1} S_{\kappa, k, \alpha_k}.
1432: \end{equation}
1433: Since $|x-\tilde x|= \kappa$, observe that a collection of balls of radius
1434: $\eta\leq 2^{-k}$ centered at points in $S'_{\kappa, k, \alpha_k}$ covering
1435: $S'_{\kappa, k, \alpha_k}$ will cover $S_{\kappa, k, \alpha_k}$ if the
1436: radius of each ball is enlarged by $\kappa$.
1437:
1438: Bourgain's condition \eqref{e:bou} may be restated as $\delta_k\to 0$, where
1439: \[
1440: \delta^2_k=2^{-2k}\sum_{|\xi|\sim 2^k}|a^2_{\xi}|.
1441: \]
1442: In particular, $\delta^2_k$ is bounded for all $k$. Now apply Lemma~\ref{l:3.1} and use the fact that $\hat I(|\xi|) = O(|\xi|^{-(d+1)/2})$ as $|\xi|\to \infty$ and
1443: also use \eqref{1.3}. We find that the number of balls of radius $\eta\leq 2^{-k}$ centered at $S'_{\kappa, k, \alpha_k}$
1444: covering
1445: $S'_{\kappa, k,\alpha_k}$ is at most
1446: \begin{eqnarray}\label{3.8}
1447: %
1448: |E_{k, \alpha_k, \eta}|
1449: &\leq &c\alpha_k^{-2}\eta^{-d}2^{-2k} e^{-\kappa 2^k}\delta_k^2\sup_{2^{k-1}\leq j <2^{k}}|\hat I(\kappa j) -1|^2\\
1450: & \leq & \left\{
1451: \begin{array}{ll} c\alpha_k^{-2} \delta_k^2 \eta^{-d} 2^{-2k} e^{-
1452: \kappa 2^k}, & \quad \kappa 2^k \geq 1\nonumber \\ \noalign{\vskip4pt}
1453: c\alpha_k^{-2} \delta_k^2 \eta^{-d} 2^{-2k} (\kappa 2^{k})^4, & \quad
1454: \kappa 2^k < 1. \end{array} \right.
1455: \end{eqnarray}
1456:
1457: We estimate $\omega(B(p_1, \rho_1)\setminus \overline G_{\kappa},
1458: S_{\kappa, k,\alpha_k}, p_2)$ according to the size of $k$.
1459:
1460: \demo{Case {\rm (i):} $\kappa 2^k \geq 1$}
1461: By \eqref{3.8} with $\eta= 2^{-k}$ and the observation made after
1462: \eqref{3.7}, the number of balls of radius $2\kappa$ covering $S_{\kappa, k, \alpha_k}$ is at most
1463: $M = c\alpha_k^{-2} \delta_k^2 2^{(d-2)k}e^{-\kappa 2^{k}}$. Let $\{B_i\}_{1\le i\le M_1}, M_1\le M$, denote these balls.
1464: Then
1465: \begin{eqnarray}\label{3.9}
1466: && \nonumber \\
1467: \noalign{\vskip-24pt}
1468: &&\\
1469: \omega(B(p_1, \rho_1)\setminus \overline G_{\kappa}, S_{\kappa, k,
1470: \alpha_k}, p_2) &\leq &\omega(B(p_1, \rho_1)\setminus S_{\kappa, k,
1471: \alpha_k}, S_{\kappa, k,
1472: \alpha_k}, p_2)\nonumber \\ \noalign{\vskip5pt}
1473: &\leq&\omega(B(p_1, \rho_1)\setminus \overline{\cup B_i}, \partial(\cup B_i), p_2)\nonumber \\ \noalign{\vskip5pt}
1474: & = & \sum_{i=1}^{M_1} \omega(B(p_1, \rho_1)\setminus \overline{\cup B_i}, \partial{B_i}, p_2)\nonumber \\ \noalign{\vskip5pt}
1475: &\leq& \sum_{i=1}^{M_1} \omega(B(p_1, \rho_1)\setminus \overline{B_i}, \partial{B_i}, p_2)\nonumber \\
1476: \noalign{\vskip5pt}
1477: &\leq& c\alpha_k^{-2} \delta_k^2 2^{(d-2)k}e^{-\kappa
1478: 2^{k}}\left(\frac{\kappa}{\tau}\right)^{d-2}\nonumber \\ \noalign{\vskip5pt}
1479: & =& \frac{c}{\tau^{d-2}}\alpha_k^{-2} \delta_k^2 (\kappa 2^{k})^{d-2}
1480: e^{-\kappa 2^k}\nonumber \\ \noalign{\vskip5pt}
1481: & \leq& \frac{c}{\tau^{d-2}}\alpha_k^{-2} \delta_k^2 (\kappa 2^{k})^{-1}, \nonumber
1482: \end{eqnarray}
1483: where $c$ is a constant which may vary from line to line. The first line
1484: of \eqref{3.9} follows from the rightmost inequality of Lemma~\ref{l:2.0},
1485: since $B(p_1, \rho_1)\setminus \overline G_{\kappa}$ has been relaced by
1486: $B(p_1, \rho_1)\setminus S_{\kappa, k,
1487: \alpha_k}$. The second line follows from the left-most inequality of
1488: Lemma~\ref{l:2.0}, since both occurrences of $S_{\kappa, k, \alpha_k}$ have
1489: been replaced by the union of balls $\bigcup_{i=1}^{M_1} B_i$ of radius
1490: $2\kappa$ covering it. The third line follows from the subadditivity of
1491: harmonic measure in the second coordinate. The fourth line follows from the
1492: right-most inequality of Lemma~\ref{l:2.0}, since $\bigcup_{i=1}^{M_1} B_i$
1493: has been replaced by $B_i$ in each term. To see the next line, write $B_i$
1494: as $B(q_i, \rho_i)$; use the explicit formula for the Poisson integral to
1495: estimate each term $\omega(B(p_1, \rho_1)\setminus \overline{B_i},
1496: \partial{B_i}, p_2)$ by $\frac{\rho_i^{d-2}}{|p_2-q_i|^{d-2}}$, where
1497: $\rho_i=2\kappa$ and $|p_2 -q_i|\geq \tau-3\kappa>\tau/2$; and finally use
1498: the first line of \eqref{3.8} to estimate the number of terms.
1499: \enddemo
1500:
1501: \vglue6pt
1502: \demo{Case {\rm (ii):} $\kappa^{-1/2}\leq 2^{k} < \kappa^{-1}$}
1503: By \eqref{3.8} with $\eta = 2^{-k}$, the number of balls of radius $2\cdot 2^{-k}$ covering $S_{\kappa, k, \alpha_k}$ is at
1504: most $c\alpha_k^{-2} \delta_k^2 2^{(d-2)k} (\kappa 2^{k})^{4}$. So as shown in Case (i),
1505: \begin{eqnarray}\label{3.10}
1506: && \nonumber \\
1507: \noalign{\vskip-24pt}
1508: &&\\
1509: \noalign{\vskip-6pt}
1510: \omega(B(p_1, \rho_1)\setminus \overline G_{\kappa}, S_{\kappa, k,
1511: \alpha_k}, p_2) &\leq& c\alpha_k^{-2} \delta_k^2 2^{(d-2)k}(\kappa
1512: 2^{k})^4\left(\frac{1}{2^{k}\tau}\right)^{d-2}\nonumber \\ \noalign{\vskip5pt}
1513: & \leq&\frac{c}{\tau^{d-2}}\alpha_k^{-2} \delta_k^2 (\kappa 2^k)^{2}. \nonumber
1514: \end{eqnarray}
1515: \enddemo
1516: \demo{Case {\rm (iii):} $2^{k}< \kappa^{-1/2}$}
1517: By \eqref{3.8} with $\eta = \sqrt{\kappa}$, the number of balls of radius $2\sqrt{\kappa}$ covering $S_{\kappa, k,
1518: \alpha_k}$ is at most $\alpha_k^{-2} \delta_k^2 2^{-2k} \kappa^{-d/2}(\kappa 2^{k})^{4}$. So,
1519: \begin{eqnarray}\label{3.11}
1520: %
1521: &&\\
1522: \noalign{\vskip-6pt}
1523: \omega(B(p_1, \rho_1)\setminus \overline G_{\kappa}, S_{\kappa, k,
1524: \alpha_k}, p_2) &\leq& c\alpha_k^{-2} \delta_k^2 2^{-2k}\kappa^{-d/2}(\kappa
1525: 2^{k})^4\left(\frac{\sqrt{\kappa}}{\tau}\right)^{d-2}\nonumber \\ \noalign{\vskip4pt}
1526: &\leq& c\alpha_k^{-2} \delta_k^2 2^{-2k}\kappa^{-d/2}(\kappa
1527: 2^{k})^2\left(\frac{\sqrt{\kappa}}{\tau}\right)^{d-2}\nonumber \\ \noalign{\vskip4pt}
1528: &\leq& \frac{c}{\tau^{d-2}}\alpha_k^{-2} \delta_k^2 \kappa. \nonumber
1529: \end{eqnarray}
1530: When \eqref{3.9}, \eqref{3.10}, and \eqref{3.11} are combined, it follows from
1531: \eqref{3.7} and the definitions of $\alpha_k$ that
1532: \begin{eqnarray}\label{3.12}
1533: %
1534: &&\\
1535: \noalign{\vskip-6pt}
1536: \omega(B(p_1, \rho_1)\setminus \overline G_{\kappa}, S_{\kappa, \alpha},
1537: p_2) &\leq &\frac{c}{\tau^{d-2}}\left\{\sum_{2^k \geq \kappa^{-1}}
1538: \alpha_k^{-2}\delta_k^2(2^k\kappa)^{-1} + \right. \nonumber\\
1539: \noalign{\vskip4pt}
1540: &&
1541: \sum_{\kappa^{-1/2}\leq 2^k <\kappa^{-1}} \left.\alpha_k^{- 2}\delta_k^2(2^k\kappa)^{2}+
1542: \sum_{2^k<\kappa^{-1/2}}\alpha_k^{- 2}\delta_k^2\kappa\right\}\nonumber \\ \noalign{\vskip4pt}
1543: &\leq& \frac{c}{\tau^{d-2}}\alpha^{-2} \left\{\max_{2^k>\kappa^{-
1544: 1/2}}\delta_k^2 + \kappa |\log\kappa|^5\right\}. \nonumber
1545: \end{eqnarray}
1546: Choose
1547: \[
1548: \beta_{\kappa} = \left\{\max_{2^k>\kappa^{-1/2}}\delta_k^2 +
1549: \kappa |\log\kappa|^5\right\}^{1/2}.
1550: \]
1551: Then by \eqref{3.6},
1552: \[
1553: I_6 \leq \int_{0}^{\beta_\kappa}\, d\alpha + \frac{c}{\tau^{d-
1554: 2}}\beta_\kappa^2\int_{\beta_\kappa}^{\infty} \alpha^{-2}\, d\alpha
1555: = \left(\frac{c}{\tau^{d-2}}+1\right)\beta_{\kappa}\to 0
1556: \]
1557: as $\kappa\to 0$. This completes the proof.
1558:
1559: \section{Proof of Lemma~\ref{l:2.2}}\label{s:4}
1560:
1561: Let the average of $H$ on the surface of $B(x, \rho)$ be denoted by
1562: \[
1563: D_\rho H(x) = \frac{1}{\sigma_d \rho^{d-1}} \int_{\partial B(x,
1564: \rho)} H(z)\, d\sigma(z),
1565: \]
1566: where $\sigma$ is the surface measure, and $\sigma_d = d v_d$ is the surface
1567: area of the unit ball in $\Bbb {R}^d$. Then,
1568:
1569: \begin{eqnarray}
1570: &&\nonumber \\
1571: \noalign{\vskip-32pt}
1572: A_\rho H(x) & = & \frac{1}{v_d \rho^d}\int_{B(x,\rho)} H(z)\, dz\label{4.1}\\
1573: \noalign{\vskip4pt}
1574: & = &\frac{1}{v_d\rho^d}\int_{0}^{\rho} \int_{\partial B(x, \beta)} H(z)\,
1575: d\sigma(z)\, d\beta\nonumber \\ \noalign{\vskip4pt}
1576: & =& \frac{\sigma_d}{v_d\rho^d}\int_{0}^{\rho} D_\beta H(x)\beta^{d-1}\,
1577: d\beta \nonumber \\ \noalign{\vskip4pt}
1578: & = &\frac{d}{\rho^d}\int_{0}^{\rho} D_\beta H(x)\beta^{d-1}\,
1579: d\beta . \nonumber
1580: \end{eqnarray}
1581: For any $\eta >0$ and $x\in B(p, r)$, by \eqref{2.5} there
1582: exist two sequences $\rho_{i,n} = \rho_{i, x, \eta, n} \downarrow 0$
1583: (with $\rho_{i,n} < r - |x-p|$), $i=1, 2$, such that for all $n\geq 1$,
1584: \[
1585: A_{\rho_{1,n}} S(x)- S(x) \geq -\eta \rho_{1,n}^2 \hbox{ and } A_{\rho_{2,n}}f_2(x) - f_2(x) \leq c_x \rho_{2,n}^2,
1586: \]
1587: where $c_x$ is a positive constant independent of $\rho_{2,n}$.
1588: Thus, by \eqref{4.1}, the above inequalities imply for all $n \geq 1$,
1589: \begin{eqnarray}
1590: %
1591: \int_{0}^{\rho_{1,n}}[ D_\beta S(x) - S(x) + a\eta \beta^2] \beta^{d-1}\, d\beta
1592: &\geq& 0\\ \noalign{\vskip4pt}
1593: \int_{0}^{\rho_{2,n}}[ D_\beta f_2(x) - f_2(x) - ac_x \beta^2] \beta^{d-1}\, d\beta
1594: &\leq& 0 \nonumber
1595: \end{eqnarray}
1596: where $a=\frac{d+2}{d}$. So there exist $\beta_n =
1597: \beta_{x, \eta, n}\leq \rho_{1,n}, \beta_n \downarrow 0$, and
1598: $r_n = r_{x, \eta, n}\break\leq \rho_{2,n}, r_n \downarrow 0$, such that
1599: \begin{equation}\label{4.2}
1600: D_{\beta_n}S(x) - S(x) \geq -a\eta \beta_n^2\hbox{ and } D_{r_n}f_2(x) - f_2(x) \leq ac_x r_n^2.
1601: \end{equation}
1602: Let $B(q, \rho_1) \subset B(p, r)$. We show, for $y\in B(q,
1603: \rho_1)\setminus\overline W$, that
1604: \begin{eqnarray}\label{4.4}
1605: %
1606: S(y) & \leq& \int_{\partial B(q, \rho_1)} S(z) \omega(B(q, \rho_1), dz, y)\\ \noalign{\vskip4pt}
1607: & = &\frac{1}{\sigma_d \rho_1} \int_{\partial B(q, \rho_1)}
1608: \frac{\rho_1^2-|q-y|^2}{|z-y|^d} S(z)\, d\sigma (z). \nonumber
1609: \end{eqnarray}
1610:
1611: If \eqref{4.4} holds, then for $q\in W$, by \eqref{4.2}, there exists a
1612: decreasing sequence $r_n$ of positive numbers going to $0$ such that for
1613: each $n$
1614: \begin{equation}\label{4.5}
1615: f_2(q) - D_{r_n}f_2(q) \geq - ac_q r_n^2.
1616: \end{equation}
1617: For any given $\epsilon>0$, using upper semi-continuity of $\overline U$ at
1618: $q$, we have for large~$n$,
1619: \[
1620: \overline U(q) \geq \sup_{|y-q|\leq r_n}\overline U(y) - \epsilon.
1621: \]
1622: Thus, for large $n$,
1623: \[
1624: \overline U(q) \geq D_{r_n}\overline U(q) -\epsilon.
1625: \]
1626: Consequently
1627: \begin{equation}\label{4.6}
1628: S(q) \geq D_{r_n} S(q)-ac_q r_n^2 -\epsilon.
1629: \end{equation}
1630: Note that for each $r>0$, by the mean value theorem, there exists a
1631: constant~$c$ such that
1632: \[
1633: \left|\frac{1}{r^{d-2}} -\frac{r^2 - |q-y|^2}{|z-y|^{d}}\right|\leq
1634: c\frac{|q - y|}{r^{d-1}},
1635: \]
1636: if $|y-q|< \frac12 |z-q|=\frac12r$. Therefore, for $|y-q|<\frac12r$,
1637: \begin{eqnarray}\label{4.7}
1638: %
1639: && \hskip-48pt \left|\frac{1}{\sigma_d r} \int_{\partial B(q, r)}
1640: \frac{r^2-|q-y|^2}{|z-y|^d} S(z)\, d\sigma (z) - D_r S(q)\right|\\ \noalign{\vskip4pt}
1641: &\leq &
1642: \frac{1}{\sigma_d r}\int_{\partial B(q, r)}
1643: \left|\frac{1}{r^{d-2}} -\frac{r^2 - |q-y|^2}{|z-y|^{d}}\right|\,
1644: |S(z)| d\sigma(z)\nonumber \\ \noalign{\vskip4pt}
1645: &\leq& c\frac{|q-y|}{r} D_r|S(q)|. \nonumber
1646: \end{eqnarray}
1647: Combining \eqref{4.4}--\eqref{4.7}, we have for any given $\epsilon$, for
1648: $n$ large,
1649: \[
1650: S(y) - S(q)\leq a c_q r_n^2 +c\frac{|q-y|}{r_n} D_{r_n} |S(q)| + \epsilon,
1651: \]
1652: if $|y-q|\leq \frac12 r_n$ and $y\in B(q, r_n)\setminus \overline W$.
1653: Letting $y\to q$, then $n \to\infty$, and then $\epsilon \to 0$, we have
1654: \[
1655: \limsup_{{y\to q\atop y \in B(q, \rho_1)\setminus\overline W}}
1656: S(y) \leq S(q).
1657: \]
1658: Thus, $S$ is upper semi-continuous at $q$ since $S$ is upper
1659: semi-continuous when restricted to $B(p, r)\cap \overline W$.
1660: Consequently, $W$ must be the empty set. So $S$ is upper semi-continuous
1661: in $B(p, r)$. Inequality \eqref{4.4} also implies that $S(q) \leq A_\rho
1662: S(q)$ for all $B(q, \rho) \subset B(p, r)$. Thus $S$ is subharmonic in
1663: $B(p, r)$ since it is also in $L^1$.
1664:
1665: It only remains to prove \eqref{4.4}. Let $\{X_t\}_{t\geq 0}$ be the
1666: standard Brownian motion starting from a fixed point $y\in B(q, \rho_1)\setminus\overline W$
1667: in the probability space $(\Omega, {\cal F}, P^{y})$. Define
1668: \[
1669: T = \inf\{t\geq 0: X_t\in \partial B(q, \rho_1)\}
1670: \]
1671: to be the exit time of $X_t$ from $B(q, \rho_1)$. Then by \eqref{3.1},
1672: inequality \eqref{4.4} is equivalent to
1673: \begin{equation}\label{4.8}
1674: S(y) \leq E^{y} [S(X_T)].
1675: \end{equation}
1676:
1677: We first show that for any stopping time $S\leq T$,
1678: \begin{equation}\label{4.9}
1679: P^y(X_S\in W) = 0.
1680: \end{equation}
1681: This is implied by
1682: \begin{equation}\label{4.91}
1683: P^y(X_S\in \overline W_{\varepsilon})=0
1684: \end{equation}
1685: as $W\subset \bigcup\limits_{\varepsilon>0} W_{\varepsilon}$.
1686: Let $R$ be the hitting time of $X_t$ with $\partial (B(q,
1687: \rho_1)\setminus\overline W_{\varepsilon})$:
1688: \[
1689: R = \inf\{t\geq 0: X_t \in \partial (B(q, \rho_1)\setminus\overline W_{\varepsilon})\}.
1690: \]
1691: Then $R \leq T$. Since $y\in B(q, \rho_1)\setminus\overline W\subset B(q, \rho_1)\setminus\overline W_{\varepsilon}$
1692: and, by assumption,
1693: \begin{eqnarray}\label{4.11}
1694: %
1695: 0 & = & \omega(B(q, \rho_1)\setminus\overline W_{\varepsilon},
1696: \partial (B(q, \rho_1)\cap\overline W_{\varepsilon}), y)\\ \noalign{\vskip4pt}
1697: & =&P^y(X_R\in \partial (B(q, \rho_1)\cap\overline W_{\varepsilon}))\nonumber \\ \noalign{\vskip4pt}
1698: & =& P^y(X_R\in\overline W_{\varepsilon}), \nonumber
1699: \end{eqnarray}
1700: we see that
1701: \begin{equation}\label{4.10}
1702: P^y(X_R\in\overline W_{\varepsilon}) = 0.
1703: \end{equation}
1704: Next, by definition,
1705: \[
1706: \{R<T\} \subset \{X_R\in \partial W_{\varepsilon}\}.
1707: \]
1708: So
1709: \begin{equation}\label{4.12}
1710: P^y(R <T) \leq P^y(X_R\in \partial W_{\varepsilon}) = 0.
1711: \end{equation}
1712: Thus, by \eqref{4.11} and \eqref{4.12} we have
1713: \begin{eqnarray}\label{4.13}
1714: %
1715: \qquad \enspace P^y(X_T\in\overline W_{\varepsilon}) & = & P^y(X_T \in\overline W_{\varepsilon}, R =T) + P^y
1716: (X_T\in\overline W_{\varepsilon}, R < T)\\ \noalign{\vskip4pt}
1717: & \leq& P^y(X_R \in\overline W_{\varepsilon} ) + P^y (R< T)\nonumber \\ \noalign{\vskip4pt}
1718: & =& 0. \nonumber
1719: \end{eqnarray}
1720: To show \eqref{4.91} for a general stopping time $S$, note that for any
1721: $\tau>0$, there exists an open set $G$ such that $\overline {W_{\varepsilon}} \subset G$
1722: and
1723: \begin{equation}\label{4.14}
1724: \omega(B(q, \rho_1)\setminus\overline G, \partial (B(q, \rho_1)\cap G), y)
1725: <\tau.
1726: \end{equation}
1727: Define a function $u$ on $\overline B(q, \rho_1)$ as follows:
1728: \[
1729: u(x)= \left\{ \begin{array}{ll} \omega(B(q, \rho_1)\setminus\overline G, \partial (B(q,
1730: \rho_1)\cap G), x)
1731: & \hbox{ on } x\in B(q, \rho_1)\setminus\overline G,\\
1732: 1 &\hbox{ on } \overline{B(q, \rho_1)\cap G},\\
1733: 0 &\hbox{ on } \partial B(q, \rho_1) \setminus \partial (B(q, \rho_1)\cap G).
1734: \end{array} \right.
1735: \]
1736: Then $u$ is superharmonic on $B(q, \rho_1)$.
1737: Let $\tilde r_n$ be an increasing sequence going up to $\rho_1$ and $y\in B(q,
1738: \tilde r_1)$. Denote $T_n$ to be the exit time of $X_t$ from $B(q, \tilde
1739: r_n)$. Clearly $T_n$ is increasing and convergent to $T$. Since Brownian
1740: motion is continuous, we have
1741: \[
1742: \{X_S\in\overline W_{\varepsilon}, S < T\}\subset \underset{n\geq 1}\bigcup
1743: \{X_{S\wedge T_n}\in\overline W_{\varepsilon}, S < T\},
1744: \]
1745: where $S\wedge T_n = \min\{S, T_n\}$. So \eqref{4.91} is implied by the
1746: following:
1747: \begin{equation}\label{4.15}
1748: P^{y}(X_{S\wedge T_n}\in\overline W_{\varepsilon}) = 0, \hbox{ for each } n,
1749: \end{equation}
1750: since by \eqref{4.13}
1751: \begin{eqnarray*}
1752: P^y(X_S\in\overline W_{\varepsilon}) & = & P^y(X_S\in\overline W_{\varepsilon}, S=T) +
1753: P^y(X_S\in\overline W_{\varepsilon}, S< T)\\
1754: & \leq& P^y(X_T\in\overline W_{\varepsilon}) +P^y(X_S\in\overline W_{\varepsilon}, S<T)\\
1755: &\leq & \lim_{n\to\infty} P^y(X_{S\wedge T_n}\in\overline W_{\varepsilon}, S< T).
1756: \end{eqnarray*}
1757: For a superharmonic function $u$ and for each $n \geq 1$, there exists a
1758: sequence of increasing superharmonic functions $\{u_j\}$ such that $u_j\in C^2$ and
1759: \begin{equation}\label{4.16}
1760: \lim_{j\to\infty} u_j = u \hbox{ on }\overline B(q, \tilde r_n)
1761: \end{equation}
1762: (see, for example, Theorem 4.20 of \cite{H}). Applying It\^o's formula to
1763: $u_j(X_{S\wedge T_n})$, we have
1764: \[
1765: E^y [u_j(X_{S\wedge T_n})] \leq u_j(y).
1766: \]
1767: Let $j$ go to infinity and apply \eqref{4.16} to see that
1768: \[
1769: E^y [u(X_{S\wedge T_n})] \leq u(y).
1770: \]
1771: Consequently
1772: \[
1773: \tau \geq u(y) \geq E^y[u(X_{S\wedge T_n})] \geq P^y(X_{S\wedge T_n}\in
1774: \overline {B(q, \rho_1)\cap G})\geq P^y(X_{S\wedge T_n}\in\overline W_{\varepsilon}).
1775: \]
1776: Letting $\tau \to 0$ proves \eqref{4.15}.
1777:
1778: As a consequence of \eqref{4.9}, since $S$ is upper semi-continuous on
1779: $\overline B(p, r)\setminus W$, we have almost everywhere with respect to the
1780: probability measure $P^y$
1781: \begin{equation}\label{4.17}
1782: \limsup_{n\to\infty}S(X_{S_n}) \leq S(X_{S_\infty}) \hbox{ if }
1783: S_n\uparrow S_{\infty}\leq T.
1784: \end{equation}
1785:
1786: Let $\eta>0$. Then by \eqref{4.2}, for any $y\in B(q, \rho_1)$, there exists
1787: $0<\beta = \beta_{y, \eta} < \rho_1-|y-q|$, such that
1788: \begin{equation}\label{4.18}
1789: S(y) - D_{\beta} S(y) \leq a\eta \beta^2.
1790: \end{equation}
1791: Consider a family of stopping times
1792: \[
1793: {\frak S} = \{S\leq T: S(y) - E^y S(X_S) \leq a\eta E^y|y - X_S|^2\}.
1794: \]
1795: Define
1796: \[
1797: S_0 = \inf\{ t\geq 0: |X_t- y|\geq \beta_{y,\eta}\}.
1798: \]
1799: Then $X_{S_0}$ is uniformly distributed on $\partial B(y,
1800: \beta_{y,\eta})$. So by \eqref{4.18}, we have
1801: \[
1802: S(y) - E^y S(X_{S_0}) \leq a\eta E^y|y - X_{S_0}|^2.
1803: \]
1804: Thus $S_0\in {\frak S}$ and hence ${\frak S}$ is not empty.
1805:
1806: For a sequence of increasing stopping times $S_n$ in ${\frak S}$, let
1807: $S_\infty = \underset{n\geq 1}\lim S_n$. Then by \eqref{4.17} and
1808: Fatou's lemma, we have
1809: \[
1810: S(y) - E^y S(X_{S_\infty}) \leq a\eta E^y |y - X_{S_\infty}|^2.
1811: \]
1812: So $S_\infty\in {\frak S}$. Thus by an argument given in Halmoe [Ha] on page 121\footnote{See [A] for the details.}, there
1813: exists
1814: $S^{*}\in {\frak S}$ such that $S^{*}$ is a maximum of ${\frak S}$. We show that
1815: $S^{*} = T$ almost everywhere with respect to $P^y$. In fact, if $S^{*} < T$ with positive $P^y$ probability, then
1816: \[
1817: S^{*}_1 = \inf\{T\geq t\geq S^{*}: |X_t - X_{S^{*}}| \geq \beta_{X_{S^{*}},
1818: \eta}\},
1819: \]
1820: where for $x\in \partial B(q, \rho_1)$, we define $\beta_{x, \eta}$ to be $0$.
1821: Then, clearly, $S^{*}_1 \geq S^{*}$ with strict inequality on
1822: $\{S^{*} < T\}$. On the other hand, conditional on $X_{S^{*}}$,
1823: $X_{S^{*}_1}$ is uniformly distributed on the surface of $B(X_{S^{*}},
1824: \beta_{X_{S^{*}}, \eta})$, if $S^{*} < S^{*}_1$. So by~\eqref{4.2}
1825: \begin{eqnarray}\label{4.19}
1826: %
1827: S(X_{S^{*}}) - E^{X_{S^{*}}} S(X_{S^{*}_1}) & =& S(X_{S^{*}}) -
1828: D_{\beta_{X_{S^{*}}, \eta}}S(X_{S^{*}}) \\ \noalign{\vskip4pt}
1829: & \leq & a\eta |\beta_{X_{S^{*}}, \eta}|^2\nonumber \\ \noalign{\vskip4pt}
1830: & =& a\eta E^{X_{S^*}}|X_{S^*} - X_{S^*_1}|^2. \nonumber
1831: \end{eqnarray}
1832: Hence, by \eqref{4.19}, the strong Markovian property, orthogonality between\break
1833: $X_{S^*_1} - X_{S^*}$ and $X_{S^*} - y$, and $S^*\in {\frak S}$, we have
1834: \begin{eqnarray*}
1835: S(y)- E^y S(X_{S^*_1}) & =& S(y) - E^y S(X_{S^*}) + E^y[S(X_{S^*}) -
1836: S(X_{S^*_1})]\\ \noalign{\vskip4pt}
1837: & \leq& a\eta E^y |y - X_{S^*}|^2 + E^y[S(X_{S^*}) - S(X_{S^*_1}), S^* <
1838: S^*_1]\nonumber\\ \noalign{\vskip4pt}
1839: & =& a\eta E^y |y - X_{S^*}|^2%\nonumber \\ \noalign{\vskip4pt}&&
1840: +\ E^y[S(X_{S^*}) - E^{X_{S^*}}S(X_{S^*_1}),S^* < S^*_1]\nonumber \\ \noalign{\vskip4pt}
1841: & \leq & a\eta E^y |y - X_{S^*}|^2 + a\eta E^y[E^{X_{S^*}}|X_{S^*} -
1842: X_{S^*_1}|^2, S^* < S^*_1]\nonumber\\ \noalign{\vskip4pt}
1843: & = &a\eta E^y |y - X_{S^*}|^2 + a\eta E^y |X_{S^*} - X_{S^*_1}|^2\nonumber\\ \noalign{\vskip4pt}
1844: & = & a\eta E^y|y - X_{S^*_1}|^2.\nonumber
1845: \end{eqnarray*}
1846: Thus $S^*_1 \in {\frak S}$. This contradicts the maximality of $S^*$ in
1847: ${\frak S}$ since $S^*_1 \geq S^*$ and $S^{*}_1 \ne S^{*}$. Thus we
1848: have shown that $T\in {\frak S}$. So
1849: \begin{eqnarray*}
1850: S(y) - \int_{\partial B(q, \rho_1)} S(z)\, \omega(B(q, \rho_1), dz, y) & = & S(y) -
1851: E^y S(X_T) \\ \noalign{\vskip4pt}
1852: &\leq& a\eta E^y |X_T -y|^2 \leq 4a\eta \rho_1^2.
1853: \end{eqnarray*}
1854: This implies \eqref{4.4} by letting $\eta \to 0$. We have finished the proof.
1855:
1856: \section{Proof of Theorem~\ref{t:4}}\label{s:5}
1857:
1858: Without loss of generality, we assume that $q = 0$, the origin. As in the
1859: proof of Theorem~\ref{t:1}, we have that $\tilde \Delta^{*} f_2(x) =
1860: \tilde\Delta_* f_2(x)$ almost everywhere in $\Bbb {T}^d
1861: \setminus\{0\}$, and that $\tilde\Delta^* f_2(x)$ is in
1862: $L^1(\Bbb {T}^d\setminus B(0, r))$ for any $r>0$. Consequently,
1863: $\tilde\Delta^*f_2(x) = \tilde\Delta_* f_2(x)$ almost everywhere in
1864: $\Bbb T^d$.
1865:
1866: As in Section~\ref{s:1}, let $B(x) = \min\{f^*(x), \tilde\Delta^*
1867: f_2(x)\}$. Then by Lemma~\ref{l:1.0}, on
1868: $\Bbb {T}^d\setminus \{0\}$, $f_*(x) \leq B(x)\leq f^*(x)$ and
1869: $\tilde\Delta_{*} f_2(x)\leq B(x)\leq \tilde\Delta^{*}f_2(x)$. Thus,
1870: $B(x)\in L^{1}(\Bbb T^d)$. Consequently, when we proceed as in
1871: Section~\ref{s:1}, there exists a function $S^*(x)$, which is harmonic
1872: on $\Bbb {R}^d\setminus M$ and almost everywhere equals
1873: \[
1874: S(x) = f_2(x) + (2\pi)^{-d}\int_{T^d} G(x-y)B(y)\, dy - b_0|x|^2/(2d),
1875: \]
1876: where $M= \{2\mu\pi: \mu\in \Bbb Z^d\}$. The rest of the proof is
1877: identical to that of Theorem~2 of Shapiro ([Sh, p.479]).
1878:
1879: \medbreak {\it Acknowledgment}.
1880: We are grateful to Jean Bourgain, Robert Kaufman, and Victor Shapiro for helpful conversations.
1881:
1882: \vglue-6pt
1883: \AuthorRefNames [AFR]
1884:
1885: \begin{references}
1886:
1887: \bibitem{A}
1888: \name{J. M. Ash}, Uniqueness for spherically convergent multiple trigonometric series, in {\it Handbook on
1889: Analytic-Computational Methods in Applied Mathematics} (G. Anastassiou, ed.), CRC Press, Boca Raton, 2000.
1890:
1891: \bibitem{AW}
1892: \name{J.\ M.\ Ash} and\name{ G.\ Wang}, A survey of uniqueness
1893: questions in multiple trigonometric series, in {\it Harmonic Analysis
1894: and Nonlinear Differential Equations\/} (Riverdale, CA 1995),
1895: Contemp.\ Math.\ {\bf 208} (1997), 35--71.
1896:
1897: \bibitem{Bo}
1898: \name{S.\ Bochner}, Zeta functions and Green's
1899: functions for linear partial differential operators of elliptic type
1900: with constant coefficients, {\it Ann.\ of Math.}\ {\bf 57} (1953), 32--56.
1901:
1902: \bibitem{B} \name{J.\ Bourgain}, Spherical summation and uniqueness of
1903: multiple trigonometric series, {\it IMRN}, no.\ 3 (1996), 93--107.
1904:
1905: \bibitem{Co} \name{B.\ Connes}, Sur les coefficients des s\'eries
1906: trigonom\'etriques convergentes sph\'eriquement, {\it CR Acad.\ Sci.\ Paris
1907: S\'er.}\ A {\bf 283} (1976), 159--161.
1908:
1909: \bibitem{C} \name{R.\ Cooke}, A Cantor-Lebesgue theorem in two
1910: dimensions, {\it Proc.\ A.M.S.} {\bf 30} (1971), 547--550.
1911:
1912: \bibitem{D} \name{B.\ Dahlberg}, Estimates of harmonic measure, {\it Arch.\ Rat.\
1913: Mech.\ Anal.}\ {\bf 65} (1977), 275--288.
1914:
1915: \bibitem{H} \name{L.\ Helms}, {\it Introduction to Potential Theory}, John
1916: Wiley \& Sons, New York, 1969.
1917:
1918: \bibitem{Ha}
1919: \name{P. R. Halmos}, {\it Measure Theory}, D. Van Nostrand, Princeton, 1956.
1920:
1921: \bibitem{K} \name{O.\ D.\ Kellogg}, {\it Foundations of Potential Theory},
1922: Springer-Verlag, Berlin, 1929.
1923:
1924: \bibitem{PW} \name{M.\ H.\ Protter} and \name{H.\ F.\ Weinberger},
1925: {\it Maximum Principles in Differential Equations},
1926: Prentice-Hall, Inc., Englewood Cliffs, NJ, 1967.
1927:
1928: \bibitem{R} \name{T.\ Rado}, {\it Subharmonic Functions}, Chelsea, New York,
1929: 1949.
1930:
1931: \bibitem{S} \name{S.\ Saks}, {\it Theory of the Integral}, Dover, New York, 1964.
1932:
1933: \bibitem{Sh} \name{V.\ L.\ Shapiro}, Uniqueness of multiple trigonometric
1934: series, {\it Ann.\ of Math.}\ {\bf 66} (1957), 467--480.
1935:
1936: \bibitem{Sh1} \bibline, The divergence theorem for
1937: discontinuous vector fields, {\it Ann.\ of Math.}\ {\bf 68} (1958), 604--624.
1938:
1939: \bibitem{Sh2} \bibline, Fourier series in several
1940: variables, {\it Bull.\ A.M.S}. {\bf 70} (1964), 48--93.
1941:
1942: \bibitem{Sh3} \bibline, Removable sets for pointwise
1943: solutions of the generalized Cauchy-Riemann equations, {\it Ann.\ of Math.}\
1944: {\bf 92} (1970), 82--101.
1945:
1946: \bibitem{Sh4} \bibline, Sets of uniqueness on the
1947: $2$-torus, {\it Trans.\ A.M.S.} {\bf 165} (1972), 127--147.
1948:
1949: \bibitem{Ve1} \name{S.\ Verblunsky}, On the theory of trigonometric series,
1950: (I), {\it Proc.\ London Math.\ Soc.}\ {\bf 34} (1932), 441--456.
1951:
1952: \bibitem{Ve2} \name{S.\ Verblunsky}, On the theory of trigonometric series,
1953: (II), {\it Proc.\ London Math.\ Soc}.\ {\bf 34} (1932), 457--491.
1954:
1955: \bibitem{Z} \name{A.\ Zygmund}, A Cantor-Lebesgue theorem for double
1956: trigonometric series, {\it Studia Math.}\ {\bf 43} (1972), 173--178.
1957: \end{references}
1958:
1959: \vglue-16pt
1960: \centerline{\ninerm (Received March 17, 1997)}
1961: \centerline{\ninerm (Revised July 21, 1999)}
1962:
1963: \end{document}
1964:
1965: