1: \documentclass[10pt]{amsart}
2:
3: \usepackage{amsmath,amsfonts,amscd,amssymb,epsf}
4: \newtheorem{theorem}{Theorem}[section]
5: \newtheorem{proposition}[theorem]{Proposition}
6: \newtheorem{lemma}[theorem]{Lemma}
7: \newtheorem{corollary}[theorem]{Corollary}
8: \theoremstyle{definition}
9: \newtheorem{definition}[theorem]{Definition}
10: \newtheorem{example}[theorem]{Example}
11: \theoremstyle{remark} \newtheorem{remark}[theorem]{Remark}
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: \numberwithin{equation}{section}
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: \newcommand{\field}[1]{\ensuremath{\mathbb{#1}}}
16: \newcommand{\CC}{\field{C}}
17: \newcommand{\DD}{\field{D}}
18: \newcommand{\HH}{\field{H}}
19:
20: \newcommand{\RR}{\field{R}}
21: \newcommand{\TT}{\field{T}}
22: \newcommand{\ZZ}{\field{Z}}
23:
24: \newcommand{\complex}[1]{\mathsf{#1}} %for differential complexes
25: \newcommand{\EEE}{\complex{E}}
26: \newcommand{\SSS}{\complex{S}}
27: \newcommand{\BBB}{\complex{B}}
28: \newcommand{\KKK}{\complex{K}}
29: \newcommand{\AAA}{\complex{A}}
30: \newcommand{\CCC}{\complex{C}}
31: \newcommand{\OOO}{\complex{O}}
32: \newcommand{\TTT}{\complex{T}}
33: \newcommand{\diff}[1]{\mathcal{#1}} %for differential
34: %operators
35: %
36: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
37: %
38: % operator names
39: \renewcommand{\d}{\operatorname{d}}
40: \DeclareMathOperator{\id}{id} \DeclareMathOperator{\I}{I}
41: \DeclareMathOperator{\Tot}{Tot} \DeclareMathOperator{\Hom}{Hom}
42: \DeclareMathOperator{\Tor}{Tor} \DeclareMathOperator{\Area}{Area}
43: \DeclareMathOperator{\Ker}{Ker} \DeclareMathOperator{\Diff}{Diff}
44: \DeclareMathOperator{\Mob}{M\ddot{o}b} \DeclareMathOperator{\im}{Im}
45: \DeclareMathOperator{\SL}{SL} \DeclareMathOperator{\PSL}{PSL}
46: \DeclareMathOperator{\PSU}{PSU} \DeclareMathOperator{\re}{Re}
47: \DeclareMathOperator{\fixed}{fixed} \DeclareMathOperator{\Res}{Res}
48: \DeclareMathOperator{\PP}{PP}
49: % differentials
50: \newcommand{\del}{\partial}
51: \newcommand{\delb}{\bar\partial}
52: \newcommand{\delp}{\partial^\prime}
53: \newcommand{\delpp}{\partial^{\prime\prime}}
54: \newcommand{\deltap}{\d}
55: \newcommand{\deltapp}{\delta}
56:
57: \newcommand{\eqdef}{\overset{\text{def}}{=}}
58: \newcommand{\ihalf}{\frac{i}{2}}
59: \newcommand{\lo}{{\mathbb{L}}}
60: \newcommand{\Z}{{\mathbb{Z}}}
61: \newcommand{\U}{\mathbb{U}}
62: \newcommand{\mL}{\mathcal{L}}
63: \newcommand{\M}{\mathcal{M}}
64: \newcommand{\B}{\mathcal{B}}
65: \newcommand{\F}{\mathcal{F}}
66: \newcommand{\mP}{\mathcal{P}}
67: \newcommand{\C}{\mathbb{C}}
68: \newcommand{\Del}{\mathbb{D}}
69: \newcommand{\blambda}{\boldsymbol{\lambda}}
70: \newcommand{\R}{\mathbb{R}}
71: \newcommand{\bk}{\backslash}
72: \newcommand{\pa}{\partial}
73: \newcommand{\g}{\gamma}
74: \newcommand{\la}{\langle}
75: \newcommand{\ra}{\rangle}
76: \newcommand{\bu}{\bullet}
77: \newcommand{\ov}{\overline}
78: \newcommand{\ep}{\epsilon}
79: \newcommand{\vep}{\varepsilon}
80: \newcommand{\z}{\bar{z}}
81: \newcommand{\ma}[4]{(\begin{smallmatrix}
82: #1 & #2 \\ #3 & #4
83: \end{smallmatrix})}
84: \DeclareMathOperator{\ad}{ad}\DeclareMathOperator{\sgn}{sgn}
85:
86: \begin{document}
87:
88: \title{On the minima and convexity of Epstein Zeta function}
89: \author{S.C. Lim$^1$}\email{$^1$sclim@mmu.edu.my}\author{L.P.
90: Teo$^{2}$}\email{$^2$lpteo@mmu.edu.my}
91:
92: \keywords{Epstein Zeta function, convexity, minima}
93: \subjclass[2000]{Primary 11E45, 26B15 } \maketitle
94:
95: \noindent {\scriptsize \hspace{1cm}$^1$Faculty of Engineering,
96: Multimedia University, Jalan Multimedia, }
97:
98: \noindent {\scriptsize \hspace{1.1cm} Cyberjaya, 63100, Selangor
99: Darul Ehsan, Malaysia.}
100:
101: \noindent {\scriptsize \hspace{1cm} $^2$Faculty of Information
102: Technology, Multimedia University, Jalan Multimedia,}
103:
104: \noindent{\scriptsize \hspace{1.1cm} Cyberjaya, 63100, Selangor
105: Darul Ehsan, Malaysia.}
106: \begin{abstract}
107: Let $Z_n(s; a_1,\ldots, a_n)$ be the Epstein zeta function defined
108: as the meromorphic continuation of the function
109: \begin{align*}
110: \sum_{k\in\Z^n\setminus\{0\}}\left(\sum_{i=1}^n [a_i
111: k_i]^2\right)^{-s},\hspace{1cm}\text{Re}\; s>\frac{n}{2}
112: \end{align*}to the complex plane. We show that for
113: fixed $s\neq n/2$, the function $Z_n(s; a_1,\ldots, a_n)$, as a
114: function of $(a_1,\ldots, a_n)\in (\R^+)^n$ with fixed
115: $\prod_{i=1}^n a_i$, has a unique minimum at the point
116: $a_1=\ldots=a_n$. When $\sum_{i=1}^n c_i$ is fixed, the function
117: $$(c_1,\ldots, c_n)\mapsto Z_n\left(s; e^{c_1},\ldots,
118: e^{c_n}\right)$$ can be shown to be a convex function of any $(n-1)$
119: of the variables $\{c_1,\ldots,c_n\}$. These results are then
120: applied to the study of the sign of $Z_n(s; a_1,\ldots, a_n)$ when
121: $s$ is in the critical range $(0, n/2)$. It is shown that when
122: $1\leq n\leq 9$, $Z_n(s; a_1,\ldots, a_n)$ as a function of
123: $(a_1,\ldots, a_n)\in (\R^+)^n$,
124: can be both positive and negative for
125: every $s\in (0,n/2)$. When $n\geq 10$, there are some open subsets
126: $I_{n,+}$ of $s\in(0,n/2)$, where $Z_{n}(s; a_1,\ldots, a_n)$ is
127: positive for all $(a_1,\ldots, a_n)\in(\R^+)^n$. By regarding
128: $Z_n(s; a_1,\ldots, a_n)$ as a function of $s$, we find that when
129: $n\geq 10$, the generalized Riemann hypothesis is false for all
130: $(a_1,\ldots,a_n)$.
131: \end{abstract}
132:
133: \section{Introduction} In \cite{Ep1, Ep2}, Epstein introduced the
134: following two-dimensional zeta function
135: \begin{align*}
136: Z_2(A;s)=\sum_{(j,k)\in
137: \Z^2\setminus\{(0,0)\}}\left[A(j,k)\right]^{-s},\hspace{1cm}\text{Re}\;
138: s> 1,
139: \end{align*}where $$A(j,k)=aj^2+2bjk+ck^2=\begin{pmatrix} j & k\end{pmatrix}\begin{pmatrix} a & b\\b &c \end{pmatrix}
140: \begin{pmatrix} j\\k\end{pmatrix}$$ is a positive definite real quadratic form associated to the
141: positive definite symmetric matrix
142: $$A=\begin{pmatrix} a & b\\b &c \end{pmatrix}.$$Later on, some generalizations of
143: this zeta function to higher dimension were considered \cite{e4, e1,
144: e5, e2, e6}. One of the generalizations is given by
145: \begin{align}\label{eq109_3}
146: Z_n(A;s)=\sum_{k\in
147: \Z^n\setminus\{0\}}\left[k^tAk\right]^{-s},\hspace{1cm}\text{Re}\;
148: s> \frac{n}{2},
149: \end{align}where now $A$ is an $n\times n$ positive definite
150: symmetric matrix and
151: $$k^tAk=\sum_{i=1}^n\sum_{j=1}^nA_{ij}k_i k_j$$ is
152: the associated quadratic form. It was proved that the Epstein zeta
153: function $Z_{n}(A,s)$ has a meromorphic continuation to the whole
154: complex plane with a single pole at $s=n/2$, and it satisfies a
155: functional equation. When $n=1$ and $A=(1)$, $Z_{1}(A;s)$ is nothing
156: but equal to $2\zeta_R(2s)$, where $\zeta_R(s)$ is the Riemann zeta
157: function. The generalized Riemann hypothesis associated to the
158: Epstein zeta function says that aside from the trivial zeros at
159: $s=-j, j\in \mathbb{N}$, all the other zeros of $Z_{n}(A;s)$ are
160: located at the critical line $\text{Re}\;s=\frac{n}{4}$. As far back
161: as 1947, it has been known \cite{e3} that the Riemann hypothesis is
162: in general not true for the Epstein zeta function, even for $n=2$.
163:
164: There exists another interpretation of Epstein zeta function with
165: more geometric flavour. Let $v_1, \ldots, v_n$ be $n$ linearly
166: independent vectors in $\R^n$ and let $L$ be the non-degenerate
167: lattice generated by $\{v_1, \ldots, v_n\}$, i.e. $$L=\left\{
168: \sum_{i=1}^n k_i v_i\,:\, (k_1,\ldots, k_n)\in \Z^n\right\}.$$ The
169: quotient of $\R^n$ by $L$, i.e. $\R^n/L$, is an $n$-dimensional
170: torus. The function
171: \begin{align*}
172: \zeta_n(L;s)=\sum_{v\in L\setminus\{\mathbf{0}\}}\langle v,
173: v\rangle^{-s},\hspace{1cm}\text{where}\;\;\langle v,w\rangle :=
174: \sum_{i=1}^nv_iw_i,
175: \end{align*}is called the zeta function associated to the lattice
176: $L$. It is actually the same as the Epstein zeta function
177: $Z_n(A;s)$ if $A=T_L^t T_L$, and $T_L$ being the $n\times n$ matrix
178: whose columns are the generators $v_1, \ldots, v_n$ of the lattice
179: $L$, i.e.,
180: $$T_L=\begin{pmatrix} v_1 & \ldots & v_n\end{pmatrix}.$$Given a lattice $L$, its dual lattice $L^*$ is defined as the set
181: \begin{align*}
182: L^*=\left\{ w\in\R^n\,:\, \langle w, v\rangle =\sum_{i=1}^n w_i v_i
183: \in \Z \;\text{for all}\; v\in L\right\}.
184: \end{align*}$L^*$ is a lattice generated by a dual set of
185: vectors $\{w_1,\ldots, w_n\}$ with $\langle w_i, v_j\rangle
186: =\delta_{ij}$. The corresponding matrix $T_{L^*}$ is related to
187: $T_L$ by $T_{L^*}=(T_L^{-1})^t$. The zeta function $\zeta_n(L^*;s)$
188: of the dual lattice $L^*$ is related to a spectral zeta function
189: associated to the compact torus $\R^n/L$. Given a compact manifold
190: $M$ with Laplace operator $\Delta_M$, the zeta function of
191: $\Delta_M$ is defined as
192: \begin{align}\label{eq109_2}\zeta_{\Delta_M}(s)=\sum_{n=1}^{\infty} \lambda_n^{-s},\end{align} where
193: $0< \lambda_1\leq \lambda_2\leq \ldots$ are the nonzero eigenvalues
194: of $\Delta_M$. For the compact torus $\R^n/L$,
195: the Laplace operator
196: $\Delta=\frac{\pa^2}{\pa x_1^2}+\ldots+\frac{\pa^2}{\pa x_n^2}$ on
197: $\R^n$ descends to the Laplace operator on $\R^n/L$. One can easily
198: check that the associated zeta function $\zeta_{\Delta_{\R^n/L}}$ is
199: up to a constant, the zeta function of the dual lattice $L^*$, i.e.
200: \begin{align*}
201: \zeta_{\Delta_{\R^n/L}}=c\zeta_{n}(L^*;
202: s)=cZ_n\left(T_L^{-1}(T_{L}^{-1})^t;s\right).
203: \end{align*}Under this perspective, the Epstein zeta function enters the realm of
204: theoretical physics. In quantum theory, very often one needs to
205: compute the functional determinant of a positive definite
206: (pseudo)-differential operator $P$ (e.g., the Laplace operator),
207: which is defined as
208: \begin{align}\label{eq109_1}
209: \mathfrak{D}_P=\prod_{i=1}^{\infty}\lambda_i,
210: \end{align}where $\lambda_i, 1\leq i<\infty$ are the nonzero eigenvalues of
211: $P$. This functional determinant usually appears as one-loop
212: partition function of a field theory, and is closely related to
213: Casimir effect
214: as well as the one-loop effective potential of the theory.
215: In general, the infinite product in \eqref{eq109_1} is divergent and
216: regularization is required to obtain a finite quantity. One of the
217: regularization techniques, known as zeta regularization method (see
218: e.g. \cite{E1, ET,K}), uses the zeta function associated to $P$,
219: $\zeta_P(s)$, which is defined as in \eqref{eq109_2}. By proving
220: that $\zeta_P(s)$ has an analytic continuation to a neighborhood of
221: $s=0$, the regularized functional determinant $\mathfrak{D}_P$ is
222: then defined as
223: \begin{align*}
224: \mathfrak{D}_P=\exp\left(-\zeta_P'(0)\right).
225: \end{align*}When the underlying spacetime of the field theory is a
226: toroidal manifold $T^n\times \R^N$, where $T^n$ is an
227: $n$-dimensional torus with compactification lengths $L_1, \ldots,
228: L_n$ and $P$ is the Laplace operator, the associated zeta function
229: $\zeta_P(s)$ is, up to a constant, equal to
230: \begin{align*}
231: \sum_{k\in\Z^n\setminus\{0\}}\left(\sum_{i=1}^n
232: \left[\frac{k_i}{L_i}\right]^2\right)^{-s+\frac{N}{2}},
233: \end{align*}which is the Epstein zeta function $Z_n\left(A;
234: s-\frac{N}{2}\right)$, with $A=\text{diag}\,\{ 1/L_1^2, \ldots,
235: 1/L_n^2\}$ a diagonal matrix. We use the notation
236: \begin{align}\label{eq109_5} Z_{n}(s; a_1, \ldots,
237: a_n):=\sum_{k\in\Z^n\setminus\{0\}}\left(\sum_{i=1}^n
238: [a_ik_i]^2\right)^{-s}
239: \end{align} to denote this special subclass of the Epstein
240: zeta function corresponding to $$A=\begin{pmatrix} a_1^2 & 0 &
241: \ldots & 0\\0 & a_2^2 & \ldots &0\\
242: \vdots &\vdots & & \vdots\\
243: 0 & 0 &\ldots & a_n^2\end{pmatrix}$$ in \eqref{eq109_3}. Here it is
244: assumed that $a_1, \ldots, a_n$ are positive real numbers. Another
245: situation where $Z_n(s;a_1,\ldots, a_n)$ appears is when the
246: underlying spacetime is a rectangular cavity $\Omega:=[0,
247: L_1]\times\ldots\times[0, L_n]\times \R^{N}$. The function $Z_n(s;
248: a_1,\ldots, a_n)$ satisfies a functional equation (or called
249: reflection formula), which can be written in the following symmetric
250: form
251: \begin{align}\label{eq109_8}
252: \sqrt{\prod_{i=1}^n a_i}\pi^{-s}\Gamma(s)Z_{n}\left(s; a_1,\ldots,
253: a_n\right)=\frac{\pi^{s-\frac{n}{2}}}{\sqrt{\prod_{i=1}^n
254: a_i}}\Gamma\left(\frac{n}{2}-s\right)Z_n\left(\frac{n}{2}-s;
255: \frac{1}{a_1},\ldots,\frac{1}{a_n}\right).
256: \end{align}
257: In physics literature, the Epstein zeta function $Z_{n}(s; A)$
258: usually appears together with gamma function in the combination
259: \begin{align}\label{eq109_9} \Gamma(s)Z_{n}(s; A).
260: \end{align} For example, in \cite{AW}, it is found that the Casimir energy of a massless
261: scalar field in the rectangular cavity $\Omega=[0,
262: L_1]\times\ldots\times[0, L_n]\times \R^{N}$ under periodic boundary
263: conditions is, up to a positive constant, equal to
264: \begin{align}\label{eq109_7}
265: -\Gamma\left(-\frac{N}{2}\right)Z_{n}\left(-\frac{N}{2};
266: \frac{1}{L_1},\ldots, \frac{1}{L_n}\right).
267: \end{align}
268: For massless scalar fields under Dirichlet or Neumann boundary
269: conditions, or electromagnetic fields in cavities with perfectly
270: conducting walls or walls with infinite permeability, the
271: corresponding Casimir energy is expressible as linear
272: combinations of expressions of the form \eqref{eq109_7}. Another
273: example is the the massless scalar field theory with
274: $\lambda\varphi^4$ interaction on the toroidal manifold $T^n\times
275: \R^{4-n}$, where $n=1,3,4$. It was found that \cite{c41, c42, c43,
276: c44, EK1} the topologically generated mass of this theory, to the
277: one-loop order, is, up to a positive constant, given by :
278: \begin{align*}
279: \Gamma\left(\frac{n}{2}-1\right)Z_{n}\left(\frac{n}{2}-1;
280: \frac{1}{L_1},\ldots,\frac{1}{L_n}\right).
281: \end{align*} In view of the functional equation \eqref{eq109_8} and the importance of the combination \eqref{eq109_9},
282: we define the Xi-function as
283: \begin{align}\label{eq109_10}
284: \Xi_n(s; a_1,\ldots, a_n)=\sqrt{\prod_{i=1}^n a_i} \pi^{-s}\Gamma(s)
285: Z_n(s; a_1,\ldots, a_n).
286: \end{align}
287:
288:
289: In the study of Casimir
290: effect, the attractive or repulsive nature of the Casimir force, and
291: the geometric configuration of the spacetime that will minimize or
292: maximize the Casimir effect are important issues. In the study of
293: interacting scalar field theory, the sign of the topologically
294: generated mass determines whether symmetry breaking mechanism
295: occurs. Therefore it is important to study the sign of the
296: Xi-function \eqref{eq109_10} and determine its extrema as a function
297: of $(a_1,\ldots, a_n)$. This paper is devoted to the study of these
298: issues. In fact, the main motivation comes from our recent study
299: of $\lambda\varphi^4$ interacting fractional Klein--Gordon
300: field theory on a toroidal manifold $T^n\times \R^N$ \cite{e10}. In
301: that study, we find that to the one-loop order, the topological
302: generated mass of a $\alpha$-fractional massless Klein--Gordon field
303: is given by
304: \begin{align*}
305: m_T^2=&\frac{\lambda}{\Gamma(\alpha)}\frac{1}{2^{2\alpha+1}\pi^{2\alpha-\frac{N}{2}}\left[\prod_{i=1}^n
306: L_i\right]}\Gamma\left(\alpha-\frac{N}{2}\right)Z_{n}\left(\alpha-\frac{N}{2};
307: \frac{1}{L_1},\ldots,\frac{1}{L_n}\right)\\
308: =&\frac{\lambda}{\Gamma(\alpha)}\frac{1}{2^{2\alpha+1}\pi^{\frac{n+N}{2}}}\Gamma\left(\frac{n+N}{2}-\alpha\right)
309: Z_{n}\left( \frac{n+N}{2}-\alpha; L_1, \ldots, L_n\right)\nonumber
310: \end{align*}when $\alpha\neq N/2$ or $(n+N)/2$. We need to study the
311: sign of $m_T^2$ as a function of $\alpha, n, N$ and $L_1,\ldots,
312: L_n$ to determine whether symmetry breaking appears. To solve this
313: problem, it is necessary to determine the minimum of $\Xi_n(s;
314: L_1,\ldots, L_n)$ for fixed $\prod_{i=1}^n L_i$ and $s$.
315:
316: The study of the (local) minima of the general Epstein zeta function
317: $Z_{n}(A;s)$ \eqref{eq109_3} as a function of $A\in
318: \{\text{symmetric positive definite} \; n\times n \; \text{real
319: matrices}\}=\mathcal{P}_n$ with fixed determinant has a long history
320: [15--38]. Such investigation is important from a number of points of
321: view. One can read the introduction of the recent paper \cite{f11}
322: for an overview of this problem. The local minima of $Z_n(A;s), A\in
323: \mathcal{P}_n, \det A=1$, for $s$ in some domain of $\R$ has only
324: been determined for $n=2, 3, 4, 5, 6, 7, 8, 24$; and the problem is
325: far from being solved. In this paper, we restrict ourselves to a
326: milder problem of determining the minimum of $Z_{n}(A;s)$ in the
327: subspace of positive definite symmetric forms consists of diagonal
328: forms. We give an elementary proof to show that for any $n$ and $s$,
329: if $\prod_{i=1}^n a_i$ is fixed, the minimum of $\Xi_n(s;
330: a_1,\ldots, a_n)$ (and hence of $Z_{n}(s; a_1, \ldots, a_n)$)
331: appears at the point $a_1=\ldots=a_n$. This result implies that if
332: $\Xi_n (s)=\Xi_n(s; 1,\ldots, 1)\geq 0$, then $\Xi_n(s; a_1, \ldots,
333: a_n)\geq 0$ for all $(a_1, \ldots, a_n)$. Therefore, the problem
334: whether $\Xi_n(s; a_1, \ldots, a_n)$ will be negative for some
335: $(a_1,\ldots, a_n)$ is reduced to to the problem of whether
336: $\Xi_n(s)$ is negative. In section 3, we give a detail study of the
337: sign of the function $\Xi_n(s)$. We show that for $1\leq n\leq 9$,
338: $\Xi_n(s)$ is negative for all $s\in (0, n/2)$; whereas for $n\geq
339: 10$, there is a nonempty subset of $(0, n/2)$ where $\Xi_n(s)$ is
340: positive. In section 4, we prove that for fixed $s$, when
341: $\sum_{i=1}^n c_i$ is fixed, the function
342: \begin{align}\label{eq1017_8}(c_1,\ldots, c_n)\mapsto Z_n\left(s; e^{c_1},\ldots,
343: e^{c_n}\right)\end{align} is a convex function of any $(n-1)$ of the
344: variables $\{c_1,\ldots,c_n\}$. In fact, we prove a stronger result.
345: We show that for a function of the form
346: $$\mathcal{F}_n(x_1,\ldots, x_n)=\prod_{i=1}^n f(x_i),$$ where
347: $f:\R\rightarrow \R^+$ is a positive twice continuously
348: differentiable function, $\mathcal{F}_n$ is convex if the function
349: $\log f(x)$ is convex. Finally, the result about the convexity of
350: the function \eqref{eq1017_8} can be used to obtain conclusion on
351: the connectivity and convexity of some regions of $(c_1,\ldots,
352: c_n)$ where $Z_n(s; e^{c_1},\ldots, e^{c_n})$ is negative.
353:
354: \section{The minimum of Epstein zeta function}
355: In this section, we show that for $s\notin \{ n/2\}\cup\{ 0, -1, -2,
356: \ldots\}$, the function $Z_n(s; a_1,\ldots, a_n), (a_1, \ldots,
357: a_n)\in (\R^+)^n$ has a minimum at $a_1=\ldots=a_n$ when
358: $\prod_{i=1}^n a_i$ is fixed.
359:
360:
361: There are various ways to obtain the meromorphic continuation of the
362: Epstein zeta function $Z_n (s; a_1, \ldots, a_n)$ \eqref{eq109_5}.
363: Two of them will be useful to us. The first one has the form (see
364: e.g. \cite{f11}):
365: \begin{align}\label{eq1011_1}
366: &V\pi^{-s}\Gamma(s)Z_n(s; a_1,\ldots, a_n)=\Xi_n(s; a_1,\ldots,
367: a_n)\\\nonumber=&-\frac{V}{s}
368: -\frac{V^{-1}}{\left(\frac{n}{2}-s\right)} + V\int_1^{\infty}
369: t^{s-1}\sum_{k\in \Z^n\setminus\{0\}}\exp\left(-\pi t\sum_{i=1}^n
370: [a_i k_i]^2\right)dt\\&+V^{-1}\int_1^{\infty}
371: t^{\frac{n}{2}-s-1}\sum_{k\in \Z^n\setminus\{0\}}\exp\left(-\pi
372: t\sum_{i=1}^n \left[\frac{k_i}{a_i}\right]^2\right)dt,\nonumber
373: \end{align}
374: where $$V=\sqrt{\prod_{i=1}^n a_i}.$$This formula shows clearly the
375: Xi-function $\Xi_n(s;a_1,\ldots, a_n)$ only has simple poles at
376: $s=0$ and $s=\frac{n}{2}$. Moreover, since the gamma function
377: $\Gamma(s)$ has simple poles at $s=0,-1, -2, -3,\ldots$, the Epstein
378: zeta function $Z_n(s; a_1, \ldots, a_n)$ is identically zero at the
379: negative integer points $-1, -2, -3,\ldots$. The second formula is
380: one form of the Chowla-Selberg formula \cite{d19,
381: d20}:\begin{align}\label{eq830_9}
382: &\pi^{-s}\Gamma(s)Z_{n}(s; a_1,\ldots, a_n)\\
383: =\nonumber &2a_1^{-2s}\pi^{-s}\Gamma(s)\zeta_R(2s)
384: +2\sum_{j=1}^{n-1}\frac{\pi^{-s+\frac{j}{2}}\Gamma\left(s-\frac{j}{2}\right)}
385: {a_{j+1}^{2s-j}\prod_{l=1}^{j}a_l} \zeta_R(2s-j) +4 \sum_{j=1}^{n-1}
386: \frac{1}{\prod_{l=1}^{j}a_l}\times\\\nonumber&\sum_{\mathbf{k}\in
387: \Z^j\setminus\{\mathbf{0}\}}\sum_{p=1}^{\infty}\frac{1}{(pa_{j+1})^{s-\frac{j}{2}}}\left(\sum_{l=1}^j
388: \left[\frac{k_l}{a_l}\right]^2\right)^{\frac{s}{2}-\frac{j}{4}}K_{s-\frac{j}{2}}\left(
389: 2\pi p a_{j+1}\sqrt{\sum_{l=1}^j
390: \left[\frac{k_l}{a_l}\right]^2}\right),
391: \end{align}where $\zeta_R(s)=\sum_{k=1}^{\infty} k^{-s}$ is the
392: Riemann zeta function.
393: From \eqref{eq830_9}, it is easy to
394: verify that for any $\lambda\in\R^+$,
395: \begin{align*}
396: Z_n(s; \lambda a_1,\ldots, \lambda a_n )=\lambda^{-2s} Z_n(s; a_1,
397: \ldots, a_n).
398: \end{align*}Therefore, it only makes sense to look for the minimum of
399: $Z_n(s; a_1,\ldots, a_n )$ as a function of $(a_1,\ldots, a_n)$
400: when $V$ is fixed. Without loss of generality, it suffices to
401: consider $V=1$. We are going to make use of \eqref{eq1011_1} to
402: find the minimum of the Xi-function $\Xi_n(s; a_1, \ldots, a_n)$ as
403: a function of $(a_1,\ldots, a_n)$ for fixed $s\neq 0, n/2$ and when
404: $V=1$. Since multiplying a constant does not affect the minimum of
405: a function, this will give the minimum of the Epstein zeta function
406: $Z_n(s; a_1,\ldots, a_n )$ as a function of $(a_1,\ldots, a_n)$ for
407: $s\notin \{n/2\}\cup\{0,-1,-2, \ldots\}$ and when $V=1$. By
408: \eqref{eq1011_1}, for fixed $s$ and $V=1$, the minimum of $\Xi_n(s;
409: a_1, \ldots, a_n)$ is the same as the minimum of
410: \begin{align}\label{eq1011_2}
411: \Lambda_n(s;a_1,\ldots, a_n)=&\int_1^{\infty} t^{s-1}\sum_{k\in
412: \Z^n\setminus\{0\}}\exp\left(-\pi t\sum_{i=1}^n [a_i
413: k_i]^2\right)dt\\&+\int_1^{\infty} t^{\frac{n}{2}-s-1}\sum_{k\in
414: \Z^n\setminus\{0\}}\exp\left(-\pi t\sum_{i=1}^n
415: \left[\frac{k_i}{a_i}\right]^2\right)dt.\nonumber
416: \end{align}Define the theta function $\vartheta(t)$ by
417: \begin{align}\label{eq1023_1}
418: \vartheta(t)=\vartheta_3(0, e^{-\pi t})=\sum_{k=-\infty}^{\infty}
419: e^{-\pi t k^2}
420: \end{align}Here $\vartheta_3(z,q)$ is a Jacobi theta function
421: \cite{g1}. The theta function $\vartheta(t)$ satisfies a reflection
422: formula: \begin{align}\label{eq1011_6}
423: \vartheta(t)=\frac{1}{\sqrt{t}}\vartheta\left(\frac{1}{t}\right).
424: \end{align}The theta function \eqref{eq1023_1} can be used to reexpress \eqref{eq1011_2} as
425: \begin{align}\label{eq1011_2_1}
426: \Lambda_n(s;a_1,\ldots, a_n)=&\int_1^{\infty}
427: t^{s-1}\left(\prod_{i=1}^n
428: \vartheta\left(ta_i^2\right)-1\right)dt\\&+\int_1^{\infty}
429: t^{\frac{n}{2}-s-1}\left(\prod_{i=1}^n
430: \vartheta\left(\frac{t}{a_i^2}\right)-1\right)dt.\nonumber
431: \end{align}
432: To determine the minimum of $\Lambda_n(s; a_1,\ldots, a_n)$, we need
433: the following:
434: \begin{proposition}\label{p1}
435: The function $\log\vartheta(e^u)$, $u\in\R$ is strictly convex.
436: Namely, for any $m$ distinct points $u_1, \ldots, u_m\in\R$ and $m$
437: constants $\lambda_1,\ldots, \lambda_m\in (0,1)$ such that
438: $\sum_{i=1}^m \lambda_i=1$, we have
439: \begin{align*}
440: \log\vartheta \left(\exp\left[\sum_{i=1}^m
441: \lambda_iu_i\right]\right)< \sum_{i=1}^m\lambda_i
442: \log\vartheta(e^{u_i}).
443: \end{align*}
444: \end{proposition}
445: This proposition can be used to prove the main result of this
446: section.
447: \begin{theorem}\label{th1}
448: For fixed $s\in\R\setminus\{0, n/2\}$, the function $\Xi_n(s;
449: a_1,\ldots, a_n)$, where $(a_1,\ldots, a_n)\in (\R^+)^n$ with
450: $\prod_{i=1}^n a_i=1$, has a unique minimum at $a_1=\ldots=a_n=1$.
451: Therefore for fixed $s\in\R\setminus\left(\{n/2\}\cup\{ 0, -1, -2,
452: -3, \ldots\}\right)$, the same statement holds for the function
453: $Z_{n}\left(s; a_1,\ldots, a_n\right)$.
454: \end{theorem}First we show how we prove this theorem from
455: Proposition \ref{p1}.
456: \begin{proof}
457: By the strict convexity of $\log \vartheta(e^u)$ asserted in
458: Proposition \ref{p1}, we find that if $(a_1,\ldots, a_n)\neq
459: (1,\ldots,1)$, then
460: \begin{align*}
461: &\log\prod_{i=1}^n \vartheta\left(a_i^2 t\right)= \sum_{i=1}^n \log
462: \vartheta\left(\exp\left[\log t +\log a_i^2\right]\right) \\>& n
463: \log\vartheta\left(\exp\left\{\frac{1}{n} \sum_{i=1}^n\left[\log
464: t+\log a_i^2\right]\right\}\right)= \log \vartheta(t)^n.
465: \end{align*}Therefore,
466: \begin{align}\label{eq1011_3}
467: \prod_{i=1}^n \vartheta\left(a_i^2 t\right)> \vartheta(t)^n.
468: \end{align}
469: Similarly,\begin{align}\label{eq1011_4} \prod_{i=1}^n
470: \vartheta\left(\frac{ t}{a_i^2}\right)> \vartheta(t)^n.
471: \end{align}Since $t^{\beta-1}>0$ for all $t>0$ and $\beta\in\R$ and
472: \begin{align*}
473: \prod_{i=1}^n \vartheta\left(b_i^2 t\right)-1> 0
474: \hspace{0.5cm}\text{for all}\; (b_1,\ldots, b_n)\in (\R^+)^n,
475: \end{align*}we conclude from \eqref{eq1011_2_1}, \eqref{eq1011_3} and \eqref{eq1011_4}
476: that for all $(a_1,\ldots, a_n)\in (\R^+)^n\setminus\{(1,\ldots,
477: 1)\}$ with $\prod_{i=1}^n a_i=1$, the inequality
478: \begin{align*}
479: \Lambda_n(s; a_1, \ldots, a_n)> \Lambda_n(s; 1,\ldots, 1)
480: \end{align*}holds. The assertion of the theorem follows.
481: \end{proof}
482: Now we return to the proof of Proposition \ref{p1}.
483: \begin{proof}Let
484: \begin{align}\label{eq1011_5}
485: g(u)&=\log\vartheta(u), \hspace{2.3cm} u\in \R^+,
486: \\f(u)&=g(e^u)=\log\vartheta(e^u),\hspace{1cm}u\in \R.\nonumber
487: \end{align}
488: We need to show that $f$ is strictly convex. Since $f$ is an
489: infinitely differentiable function, we need to show that
490: $f^{\prime\prime}(u)>0$ for all $u\in\R$. First, note that the
491: reflection formula \eqref{eq1011_6} implies that
492: \begin{align}\label{eq1011_7}
493: f(-u)=\frac{u}{2}+f(u).
494: \end{align} Therefore
495: $$f^{\prime\prime}(-u)=f^{\prime\prime}(u).$$ As a result, we only need to show that $f^{\prime\prime}(u)> 0$
496: for all $u\in\R^+$. By definition,
497: \begin{align*}
498: g'(u)=\frac{\vartheta'(u)}{\vartheta(u)},
499: \hspace{1cm}g^{\prime\prime}(u)=\frac{\vartheta^{\prime\prime}(u)\vartheta(u)-\vartheta'(u)^2}{\vartheta
500: (u)^2},
501: \end{align*}and\begin{align*}
502: f^{\prime\prime}(u) =& e^{2u}g^{\prime\prime}(u)+e^ug'(e^u)\\
503: =&e^{2u}\frac{
504: \vartheta^{\prime\prime}(e^u)\vartheta(e^u)-\vartheta'(e^u)^2+e^{-u}\vartheta(e^u)\vartheta'(e^u)}{\vartheta(e^u)^2}.
505: \end{align*}Hence, we have to prove that for all $v\geq 1$,
506: \begin{align*}h(v)=&\vartheta^{\prime\prime}(v)\vartheta(v)-\vartheta'(v)^2+v^{-1}\vartheta(v)\vartheta'(v)>0.
507: \end{align*}Again by definition,\begin{align}\label{eq1011_9}h(v)
508: =&\sum_{j\in\Z} \pi^2j^4 e^{-\pi v j^2}\sum_{k\in\Z}e^{-\pi v
509: k^2}-\sum_{j\in \Z}\pi j^2 e^{-\pi vj^2}\sum_{k\in\Z} \pi k^2e^{-\pi
510: vk^2} -v^{-1}\sum_{j\in\Z} e^{-\pi v j^2}\sum_{k\in\Z}\pi k^2e^{-\pi
511: vk^2}\\
512: =&\frac{\pi}{2}\sum_{(j,k)\in\Z^2}\left[\pi j^4+\pi k^4-2\pi
513: j^2k^2-\frac{1}{v}(j^2+k^2)\right] e^{-\pi v (j^2+k^2)}.\nonumber
514: \end{align}For fixed $v\geq 1$ and $(j,k)\in \Z^2$, define
515: \begin{align*}
516: C(j,k) =\pi j^4+\pi k^4-2\pi
517: j^2k^2-\frac{1}{v}(j^2+k^2)=\pi(k^2-j^2)^2-\frac{1}{v}(j^2+k^2).
518: \end{align*} If $C(j,k)>0$ for all $(j,k)\in\Z^2$, it follows directly that $h(v)>0$. However, $C(k,k)=
519: -2v^{-1}k^2\leq 0$, which renders the verification of $h(v)>0$
520: slightly more complicated. We need to compare the values of $C(j,k)$
521: for different pairs of $(j,k)$. Since
522: $C(j,k)=C(-j,k)=C(j,-k)=C(-j,-k)$ and $C(j,k)=C(k,j)$, we need only
523: to consider $C(j,k)$ as a function of $(j,k)\in \mathbb{N}_0$ with
524: $j\leq k$. Here $\mathbb{N}_0=\mathbb{N}\cup\{0\}$. We claim that
525: \begin{align*}
526: &(a) \; \text{If}\; j\neq k, C(j,k)>0,\\
527: &(b)\; \text{For any}\; k\geq 1, C(k-1,
528: k)+\frac{1}{2}C(k,k)>0.\hspace{6cm}
529: \end{align*} By rewriting $C(j,k)$ as
530: \begin{align*}
531: C(j,k)=\pi j^4-\left(2\pi k^2+\frac{1}{v}\right)j^2+\pi
532: k^4-\frac{k^2}{v},
533: \end{align*} we see that for fixed $k$, $C(j,k)$ is decreasing if
534: $j\in [0, k]$. Therefore, to prove $(a)$, it suffices to verify that
535: $C(k-1, k)>0$. Since $C(k,k)\leq 0$, this will follow if we prove
536: $(b)$. Now for $v\geq 1$, the function
537: \begin{align*}
538: q(k)=C(k-1,k)+\frac{1}{2}C(k,k)
539: =&\pi(4k^2-4k+1)-\frac{1}{v}(2k^2-2k+1)-\frac{k^2}{v}\\
540: \geq &\pi(4k^2-4k+1)-(3k^2-2k+1)\\
541: =&(4\pi -3)k^2 -2(2\pi-1)k +\pi-1
542: \end{align*}is increasing for $k\geq 1$, and $q(1)= \pi-2>0$, this
543: proves $(b)$. Returning to the proof of $h(v)>0$, we write the
544: summation over $(j,k)\in\Z^2$ in \eqref{eq1011_9} as
545: \begin{align*}
546: \frac{\pi}{2}\Biggl\{\sum_{\substack{(j,k)\in \Z^2\\|j-k|\geq
547: 2}}+\sum_{\substack{(j,k)\in \Z^2\\|j-k|=
548: 1}}+\sum_{\substack{(j,k)\in \Z^2\\j=k}}\Biggr\}C(j,k)e^{-\pi v
549: (j^2+k^2)}.
550: \end{align*}By $(a)$, the first sum is strictly positive. Now since $C(0,0)=0$ and
551: $C(\pm j,\pm k)=C(j,k)=C(k,j)$, the last two sums can be written as
552: \begin{align*}
553: 2\pi\sum_{k=1}^{\infty}C(k-1, k) e^{-\pi((k-1)^2+k^2)}
554: +\pi\sum_{k=1}^{\infty}C(k,k)e^{-2\pi k^2}.
555: \end{align*}Using the fact that $e^{-x}$ is a decreasing function,
556: we find that this term is larger than
557: \begin{align*}
558: 2\pi\sum_{k=1}^{\infty}\left(C(k-1,1)+\frac{1}{2}C(k,k)\right)
559: e^{-2\pi k^2},
560: \end{align*}which, by $(b)$, is positive. This concludes the proof
561: that $h(v)>0$, and therefore the assertion of the proposition
562: follows.
563: \end{proof}
564:
565: \section{The sign of Epstein zeta function} In this section,
566: we are going to study the sign of the Xi-function $\Xi_n(s;
567: a_1,\ldots, a_n)$ when $s\in\R \setminus \{0, n/2\}$. First, since
568: $\pi^{-s}\Gamma(s)>0$ for $s>0$ and it is obvious from the power
569: series definition of $Z_{n}(s; a_1, \ldots, a_n)$ \eqref{eq109_5}
570: that $Z_n(s; a_1,\ldots, a_n)>0$ for all $s>n/2$, we immediately
571: deduce that
572: $$\Xi_n(s; a_1, \ldots, a_n)>0 \hspace{1cm}\text{for all}\;\;
573: s>\frac{n}{2} \;\;\text{and}\;\;(a_1,\ldots, a_n)\in (\R^+)^n.$$
574: From the functional equation \eqref{eq109_8}, we then obtain
575: $$\Xi_n(s; a_1, \ldots, a_n)>0 \hspace{1cm}\text{for all}\;\;
576: s<0 \;\;\text{and}\;\;(a_1,\ldots, a_n)\in (\R^+)^n.$$In other
577: words, $\Xi_n(s; a_1, \ldots, a_n)>0 $ for all $s\in (-\infty,
578: 0)\cup(n/2, \infty)$. For the remaining case where $s\in (0,n/2)$,
579: the sign of the function $\Xi_n(s; a_1,\ldots, a_n)$ is not trivial.
580: Using the formula \eqref{eq1011_1}, we find that
581: around the two simple poles $s=0$ and $s=\frac{n}{2}$,
582: \begin{align*}
583: \Xi_n(s; a_1,\ldots, a_n)=&-\frac{V}{s}+O(1)\hspace{1cm}\text{as}\;\; s\rightarrow 0,\\
584: \Xi_n(s; a_1, \ldots,
585: a_n)=&-\frac{V^{-1}}{\frac{n}{2}-s}+O(1)\hspace{1cm}\text{as}\;\;
586: s\rightarrow \frac{n}{2}.
587: \end{align*}Therefore we can conclude that\begin{proposition}\label{p2} For any $a=(a_1,\ldots, a_n)\in(\R^+)^n$, one
588: can find a right nonempty neighbourhood $I_1(a)=(0, s_{1}(a))$ of
589: $s=0$ and a left nonempty neighbourhood $I_2(a)=\left(s_{2}(a),
590: \frac{n}{2}\right)$ of $n/2$ such that
591: \begin{align*}
592: \Xi_n(s; a_1,\ldots, a_n)<0, \hspace{1cm}\text{for all}\;\; s\in
593: I_1(a)\cup I_2(a).
594: \end{align*}\end{proposition}In fact, when $n=1$, using the fact that
595: $$\Xi_1(s;a)=\sqrt{a}
596: \pi^{-s}\Gamma(s)Z_1(s;a)=2a^{-2s+\frac{1}{2}}\pi^{-s}\Gamma(s)\zeta_R(2s),$$
597: it is easy to verify a stronger result:
598: \begin{proposition}
599: $$\Xi_1(s;a)<0,\hspace{1cm}\text{for all}\;\; s\in \left(0,
600: \frac{1}{2}\right).$$
601: \end{proposition}\begin{proof}
602: It suffices to show that $\zeta_R(s)<0$ for all $s\in (0,1)$. For
603: $s>0$, there is a well-known formula
604: \begin{align*}
605: \sum_{k=1}^{\infty}\frac{(-1)^{k-1}}{k^s}=\left(1-2^{1-s}\right)\zeta_R(s).
606: \end{align*}The alternating series on the left hand side is
607: convergent and positive for all $s>0$. However, for $s\in (0,1)$,
608: $2^{1-s}>1$. Therefore, $\zeta_R(s)<0$ for all $s\in (0,1)$. This
609: completes the proof.
610: \end{proof}In the study of the zeros of Riemann zeta function, this result asserts that $\zeta_R(s)$ has no nontrivial
611: zeros in $(0, 1)$, a necessary condition for the validity of Riemann
612: hypothesis. For $n\geq 2$, in view of Proposition \ref{p2}, we find
613: that a necessary condition for the validity of the generalized
614: Riemann hypothesis for $Z_n(s; a_1,\ldots, a_n)$ at a fixed
615: $(a_1,\ldots, a_n)\in (\R^+)^n$, is that $\Xi_n(s; a_1,\ldots,
616: a_n)<0$ for all $s\in (0, n/2)$. However, this is not true for all
617: $(a_1,\ldots, a_n)\in (\R^+)^n$. In fact, we can show that
618: \begin{proposition}\label{p4}
619: Given $n\geq 2$, $j\in \{1,\ldots, n\}$ and $s\in (0,n/2)$, if we
620: fix $a_1,\ldots, a_{j-1}$, $a_{j+1},\ldots, a_n$ and vary $a_j$,
621: then as $a_j$ is large enough, $\Xi_n(s; a_1,\ldots, a_n)$ is
622: positive.
623: \end{proposition}
624: \begin{proof}Since $\Xi_n (s; a_1,\ldots, a_n)$ is symmetric in the
625: variables $a_1,\ldots, a_n$, it is sufficient to show that given
626: $n\geq 2$ and $s\in (0,n/2)$, there exists $j\in \{1,\ldots, n\}$
627: such that if we fix $a_1,\ldots, a_{j-1}$, $a_{j+1},\ldots, a_n$ and
628: vary $a_j$, then as $a_j$ is large enough, $\Xi_n(s; a_1,\ldots,
629: a_n)$ is positive.
630:
631: First we consider the case $s=\frac{m}{2}$, where $m\in
632: \{1,2,\ldots, n-1\}$. Equation \eqref{eq830_9} and the fact that
633: (see \cite{e10})
634: \begin{align*}\pi^{-s}\Gamma\left(s\right)\zeta_R(2s)
635: =& -\frac{1}{2}\left\{\frac{1}{s}+\log
636: (4\pi)+\psi(1)\right\}+O(s),\hspace{1cm}\text{as}\;\; s\rightarrow
637: 0,\\\pi^{-s}\Gamma\left(s\right)\zeta_R(2s) =&
638: \frac{1}{2}\left\{\frac{1}{s-\frac{1}{2}}-\log
639: (4\pi)-\psi(1)\right\}+O\left(s-\frac{1}{2}\right),\hspace{1cm}\text{as}\;\;
640: s\rightarrow \frac{1}{2},\nonumber
641: \end{align*}
642: give
643: \begin{align}\label{eq1023_2}
644: &V^{-1}\Xi_n\left(\frac{m}{2}; a_1,\ldots,
645: a_n\right)=\frac{1}{\prod_{l=1}^m
646: a_l}\left\{\log\frac{a_{m+1}}{a_m}-\log(4\pi)-\psi(1)\right\}\\
647: +\nonumber &2\sum_{j\in\{0,1,\ldots,n-1\}\setminus\{m-1,
648: m\}}\frac{\pi^\frac{j-m}{2}\Gamma\left(\frac{m-j}{2}\right)}
649: {a_{j+1}^{m-j}\prod_{l=1}^{j}a_l} \zeta_R(m-j) +4 \sum_{j=1}^{n-1}
650: \frac{1}{\prod_{l=1}^{j}a_l}\times\nonumber\\\nonumber&\sum_{\mathbf{k}\in
651: \Z^j\setminus\{\mathbf{0}\}}\sum_{p=1}^{\infty}\frac{1}{(pa_{j+1})^{\frac{m-j}{2}}}\left(\sum_{l=1}^j
652: \left[\frac{k_l}{a_l}\right]^2\right)^{\frac{m-j}{4}}K_{\frac{m-j}{2}}\left(
653: 2\pi p a_{j+1}\sqrt{\sum_{l=1}^j
654: \left[\frac{k_l}{a_l}\right]^2}\right).
655: \end{align}Since the function $\pi^{-s}\Gamma(s)\zeta_R(2s)$ is
656: strictly positive for $s<0$ or $s>\frac{1}{2}$ and the function
657: $K_s(z)$ is positive for all $s\in\R$ and $z\in \R^+$, the last two
658: terms of \eqref{eq1023_2} are always positive. The sign of the first
659: term depend on the ratio $a_{m+1}/a_m$. It is easy to see that if
660: $a_1,\ldots, a_m, a_{m+2},\ldots, a_n$ are kept fixed, then for
661: $a_{m+1}$ large, the first term is positive and therefore
662: $\Xi_{n}\left(\frac{m}{2}; a_1, \ldots, a_n\right)>0$. This
663: establishes the proposition when $s\in \{m/2\,:\, m=1,2,\ldots,
664: n-1\}$.
665:
666: For the general case, given $s\in (0, n/2)\setminus\{m/2\,:\,
667: m=1,2,\ldots, n-1\}$, there is a unique $m\in \{0, \ldots, n-1\}$
668: such that $0<2s-m<1$. Equation \eqref{eq830_9} gives
669: \begin{align*}
670: &V^{-1}\Xi_n(s; a_1,\ldots,
671: a_n)=\frac{\pi^{\frac{m}{2}-s}\Gamma\left(s-\frac{m}{2}\right)}
672: {a_{m+1}^{2s-m}\prod_{l=1}^{m}a_l} \zeta_R(2s-m)\\
673: \nonumber
674: &+2\sum_{j=0}^{m-1}\frac{\pi^{\frac{j}{2}-s}\Gamma\left(s-\frac{j}{2}\right)}
675: {a_{j+1}^{2s-j}\prod_{l=1}^{j}a_l} \zeta_R(2s-j)
676: +2\sum_{j=m+1}^{n-1}\frac{\pi^{\frac{j}{2}-s}\Gamma\left(s-\frac{j}{2}\right)}
677: {a_{j+1}^{2s-j}\prod_{l=1}^{j}a_l} \zeta_R(2s-j) +4 \sum_{j=1}^{n-1}
678: \frac{1}{\prod_{l=1}^{j}a_l}\times\\\nonumber&\sum_{\mathbf{k}\in
679: \Z^j\setminus\{\mathbf{0}\}}\sum_{p=1}^{\infty}\frac{1}{(pa_{j+1})^{s-\frac{j}{2}}}\left(\sum_{l=1}^j
680: \left[\frac{k_l}{a_l}\right]^2\right)^{\frac{s}{2}-\frac{j}{4}}K_{s-\frac{j}{2}}\left(
681: 2\pi p a_{j+1}\sqrt{\sum_{l=1}^j
682: \left[\frac{k_l}{a_l}\right]^2}\right).
683: \end{align*}Only the first term is negative. If $a_1,\ldots, a_{m},
684: a_{m+2},\ldots, a_n$ are fixed and $a_{m+1}\rightarrow \infty$, the
685: second term is fixed but the first term approaches zero. Therefore,
686: $\Xi_n(s; a_1,\ldots,
687: a_n)>0$ when $a_{m+1}$ is large enough, and our assertion is proved.
688:
689: We would like to remark that the case $s=\frac{m}{2}$, where $m\in
690: \{1,2,\ldots, n-1\}$ can actually be deduced from the general case
691: $s\in (0, n/2)\setminus\{m/2\,:\, m=1,2,\ldots, n-1\}$ using the
692: joint continuity of $\Xi_n(s; a_1, \ldots, a_n)$ as a function of
693: $s$ and $(a_1,\ldots, a_n)$.
694: \end{proof} This proposition shows that if $s\in (0, n/2)$, then when one of the ratios between the
695: $a_i$'s, $1\leq i\leq n$ is large, the Epstein zeta function
696: $Z_{n}(s; a_1,\ldots, a_n)$ is positive, and therefore the
697: generalized Riemann hypothesis does not hold for this zeta function.
698: It is then interesting to ask whether the Riemann hypothesis will
699: still be valid for some $(a_1,\ldots, a_n)\in (\R^+)^n$. Another
700: interesting question which is related to our problem \cite{e10} of
701: determining whether there exists symmetry breaking mechanisms in an
702: interacting fractional Klein--Gordon field is, for what values of
703: $s$, the Xi-function $\Xi_n(s; a_1,\ldots, a_n)$ will be negative
704: for some $(a_1,\ldots, a_n)$. Recall that we have proved in section
705: 2 that for any fixed $s$, the minimum of $Z_{n}(s; a_1,\ldots,
706: a_n)$, as a function of $(a_1,\ldots, a_n)\in (\R^+)^n$ with
707: $\prod_{i=1} a_i$ fixed, appears at the point $a_1=\ldots=a_n$.
708: Therefore to answer these two questions, we need to study the
709: function $\Xi_n(s)=\Xi_n(s; 1,\ldots, 1)$ first.
710:
711: From \eqref{eq1011_1}, we find that
712: \begin{align}\label{eq1015_3}
713: \Xi_n(s)=-\frac{1}{s} -\frac{1}{\frac{n}{2}-s}+\int_1^{\infty}
714: t^{s-1}\left(\vartheta(t)^n-1\right)dt+\int_1^{\infty}
715: t^{\frac{n}{2}-1-s}\left(\vartheta(t)^n-1\right)dt.
716: \end{align}
717: In order to have a more unified treatment for all $n$, we define
718: $\hat{\Xi}_n(s)=\Xi_n\left(\frac{ns}{2}\right)$. Then
719: \begin{align}\label{eq1015_10}
720: \hat{\Xi}_n(s)=-\frac{2}{n}\left(\frac{1}{s}
721: +\frac{1}{1-s}\right)+\int_1^{\infty}
722: t^{\frac{ns}{2}-1}\left(\vartheta(t)^n-1\right)dt+\int_1^{\infty}
723: t^{\frac{n}{2}(1-s)-1}\left(\vartheta(t)^n-1\right)dt.
724: \end{align} This formula shows that for all $s\in
725: (0,1)$ and $n\in \mathbb{N}$,
726: \begin{align}\label{eq1015_1}\hat{\Xi}_{n+1}(s)
727: >\hat{\Xi}_{n}(s).\end{align} Moreover, for any fixed $s$, since
728: \begin{align*}
729: \frac{2}{n}\left(\frac{1}{s}+\frac{1}{1-s}\right) \longrightarrow 0
730: \hspace{1cm}\text{as}\;\; n\rightarrow \infty,
731: \end{align*}and
732: \begin{align*}
733: &\int_1^{\infty} t^{\frac{ns}{2}-1}
734: \left(\vartheta(t)^n-1\right)dt+\int_1^{\infty}t^{\frac{n}{2}(1-s)-1}\left(\vartheta(t)^n-1\right)dt
735: \\\geq & \int_1^{\infty} t^{\frac{s}{2}-1}
736: \left(\vartheta(t)^n-1\right)dt+\int_1^{\infty}t^{\frac{1}{2}(1-s)-1}\left(\vartheta(t)^n-1\right)dt>0,
737: \end{align*}therefore $\hat{\Xi}_n(s)$ becomes positive when $n$ is
738: large enough. In fact for $s=1/2$, a lengthy analysis (31 pages) has
739: been done in \cite{g2} and it was proved that\begin{lemma}\label{l1}
740: \begin{align*}\begin{cases}\Xi_n\left(\frac{n}{4}\right)=\hat{\Xi}_n\left(\frac{1}{2}\right)<0,
741: \hspace{1cm}\text{if}\;\;& 1\leq n\leq 9;\\
742: \Xi_n\left(\frac{n}{4}\right)=\hat{\Xi}_n\left(\frac{1}{2}\right)>0,
743: \hspace{1cm}\text{if}\;\; & n\geq 10.
744: \end{cases}
745: \end{align*}\end{lemma}\begin{proof}Here we provide a much shorter proof of this result.
746: Using the inequality \eqref{eq1015_1}, we need
747: only to show that
748: \begin{align}\label{eq1015_2}
749: \Xi_9\left(\frac{9}{4}\right)<0
750: \hspace{1cm}\text{and}\hspace{1cm}\Xi_{10}\left(\frac{5}{2}\right)>0.
751: \end{align}Using the incomplete gamma function
752: \begin{align*}
753: \Gamma(\beta,x)=\int_x^{\infty} t^{\beta-1} e^{-t} dt,
754: \hspace{1cm}x>0,
755: \end{align*}we can rewrite \eqref{eq1015_3} as
756: \begin{align}\label{eq1015_4}
757: \Xi_n(s)=&-\frac{1}{s}-\frac{1}{\frac{n}{2}-s}
758: +\sum_{k\in\Z^n\setminus\{0\}} \left(\pi
759: |k|^2\right)^{-s}\Gamma\left( s, \pi|k|^2\right)
760: \\&+\sum_{k\in\Z^n\setminus\{0\}} \left(\pi
761: |k|^2\right)^{-\frac{n}{2}+s}\Gamma\left( \frac{n}{2}-s,
762: \pi|k|^2\right),\nonumber
763: \end{align}where for $k\in\Z^n$,
764: $|k|^2=\sum_{i=1}^n k_i^2$. We need to obtain upper and lower bounds
765: for $\Gamma(\beta, x)$. Using integration by parts, it is easy to
766: show that when $\beta>0$,
767: \begin{align*}
768: \Gamma(\beta, x)=x^{\beta-1}e^{-x}\sum_{j=0}^m \frac{\prod_{i=1}^j
769: (\beta-i)}{x^j}+\left[\prod_{i=1}^{m+1}(\beta-i)\right]\int_x^{\infty}
770: t^{\beta-m-2}e^{-t}dt.
771: \end{align*}From this, we find that
772: \begin{align}\label{eq1015_8}
773: \Gamma(\beta,x)\leq x^{\beta-1}e^{-x}\sum_{j=0}^{[\beta]}
774: \frac{\prod_{i=1}^j (\beta-i)}{x^j}, \hspace{1cm}\Gamma(\beta,x)\geq
775: x^{\beta-1}e^{-x}\sum_{j=0}^{[\beta]+1} \frac{\prod_{i=1}^j
776: (\beta-i)}{x^j}.
777: \end{align}Therefore from \eqref{eq1015_4}, we have
778: \begin{align*}
779: \Xi_9\left(\frac{9}{4}\right)\leq
780: -\frac{8}{9}+2\sum_{k\in\Z^9\setminus\{0\}}\left(\pi|k|^2\right)^{-1}e^{-\pi|k|^2}\sum_{j=0}^{2}
781: \frac{\prod_{i=1}^j \left(\frac{9}{4}-i\right)}{(\pi|k|^2)^j}.
782: \end{align*}For the summation over $k\in \Z^9\setminus\{0\}$, we
783: single out the 18 terms that contribute to $|k|=1$, i.e. we write
784: \begin{align}\label{eq1015_7}
785: &\sum_{k\in\Z^9\setminus\{0\}}\left(\pi|k|^2\right)^{-1}e^{-\pi|k|^2}\sum_{j=0}^{2}
786: \frac{\prod_{i=1}^j \left(\frac{9}{4}-i\right)}{(\pi|k|^2)^j}\\
787: =&\frac{18}{\pi}e^{-\pi}\sum_{j=0}^2\frac{\prod_{i=1}^j
788: \left(\frac{9}{4}-i\right)}{\pi^j}+\sum_{k\in\Z^9\setminus\{k\,:\,
789: |k|=0\;\text{or}\;
790: 1\}}\left(\pi|k|^2\right)^{-1}e^{-\pi|k|^2}\sum_{j=0}^{2}
791: \frac{\prod_{i=1}^j
792: \left(\frac{9}{4}-i\right)}{(\pi|k|^2)^j}.\nonumber
793: \end{align}The first term of \eqref{eq1015_7} gives
794: \begin{align}\label{eq1015_6}
795: \frac{18}{\pi}e^{-\pi}\sum_{j=0}^2\frac{\prod_{i=1}^j
796: \left(\frac{9}{4}-i\right)}{\pi^j}=0.3540;
797: \end{align}whereas the second term of \eqref{eq1015_7} is
798: \begin{align}\label{eq1015_5}
799: &\sum_{k\in\Z^9\setminus\{k\,:\, |k|=0\;\text{or}\;
800: 1\}}\left(\pi|k|^2\right)^{-1}e^{-\pi|k|^2}\sum_{j=0}^{2}
801: \frac{\prod_{i=1}^j \left(\frac{9}{4}-i\right)}{(\pi|k|^2)^j}\\\leq
802: & \frac{1}{2\pi}\sum_{j=0}^{2} \frac{\prod_{i=1}^j
803: \left(\frac{9}{4}-i\right)}{(2\pi
804: )^j}\sum_{k\in\Z^9\setminus\{k\,:\, |k|=0\;\text{or}\; 1\}}
805: e^{-\pi|k|^2}.\nonumber
806: \end{align}Using a simple inequality
807: \begin{align*}
808: \exp\left(-\pi \sum_{i=1}^n k_i^2\right)\leq
809: \exp\left(-\pi\sum_{i=1}^n |k_i|\right),
810: \end{align*}we find that
811: \begin{align*}
812: \sum_{k\in\Z^9\setminus\{k\,:\, |k|=0\;\text{or}\; 1\}}
813: e^{-\pi|k|^2}=&\sum_{k\in\Z^9} e^{-\pi|k|^2}-1-18e^{-\pi}\\
814: \leq & \sum_{k\in\Z^9} e^{-\pi\sum_{i=1}^9|k_i|}-1-18e^{-\pi}\\
815: =&\left(1+2\sum_{m=1}^{\infty}e^{-\pi m}\right)^9-1-18e^{-\pi}\\
816: =&\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}.
817: \end{align*}Therefore, the term \eqref{eq1015_5} is bounded above by
818: \begin{align*}
819: \frac{1}{2\pi}\sum_{j=0}^{2} \frac{\prod_{i=1}^j
820: \left(\frac{9}{4}-i\right)}{(2\pi
821: )^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)=0.0769.
822: \end{align*}Combine with \eqref{eq1015_6}, we find that
823: \eqref{eq1015_7} is bounded above by
824: \begin{align*}
825: 0.3540+0.0769=0.4309<\frac{4}{9}.
826: \end{align*}Therefore, $\Xi_9(9/4)<0$.
827: Now for $\Xi_{10}(5/2)$, a similar argument but using the lower
828: bound in \eqref{eq1015_8} for the incomplete gamma function gives
829: \begin{align*}
830: \Xi_{10}\left(\frac{5}{2}\right)>&-\frac{4}{5}+2\sum_{k\in\Z^{10}\setminus\{0\}}
831: \left(\pi|k|^2\right)^{-1}e^{-\pi|k|^2}\sum_{j=0}^{3}
832: \frac{\prod_{i=1}^j \left(\frac{5}{2}-i\right)}{(\pi|k|^2)^j}\\
833: >&-\frac{4}{5}+\frac{40}{\pi}e^{-\pi}\sum_{j=0}^{3}
834: \frac{\prod_{i=1}^j \left(\frac{5}{2}-i\right)}{\pi ^j}=0.04808>0.
835: \end{align*}
836:
837:
838:
839:
840: As a side remark, using standard mathematics softwares such as
841: MATLAB and MATHEMATICA, it is easy to compute $\Xi_n(s)$ to any
842: desired degree of accuracy from formula \eqref{eq1015_3} or
843: \eqref{eq1015_4}. Up to $10^{-12}$, we have
844: \begin{align*}
845: \Xi_9\left(\frac{9}{4}\right)=-0.065884758538,\\
846: \Xi_{10}\left(\frac{5}{2}\right) =0.205903040487.
847: \end{align*}
848:
849: \end{proof}
850:
851: Lemma \ref{l1}, Proposition \ref{p2} and the inequality
852: \eqref{eq1015_1} immediately implies that
853: \begin{proposition}\label{p5}
854: For all $n\geq 10$, the sets
855: \begin{align}\label{eq1016_1}I_{n,+}=&\left\{ s\in(0, n/2)\,:\,
856: \Xi_{n}(s)>0\right\},\\ \label{eq1017_1} I_{n,-}=&\left\{ s\in(0,
857: n/2)\,:\, \Xi_{n}(s)<0\right\}\end{align}are open nonempty subsets
858: of $(0, n/2)$. Moreover, $n/4\in I_{n,+}$.
859: \end{proposition}In fact, from the functional equation
860: $\Xi_n(s)=\Xi_n((n/2)-s)$, one can even conclude that the intervals
861: $I_{n,+}$ and $I_{n,-}$ are symmetric with respect to the point
862: $n/4$, i.e.
863: $$\frac{n}{2}-I_{n,+}:=\left\{\frac{n}{2}-s\,:\, s\in I_{n,+}\right\}=I_{n,+}$$ and similarly for
864: $I_{n,-}$. Since $$\lim_{s\rightarrow 0^+}\Xi_n(s)=-\infty,$$ there
865: must exists an odd number of points $\gamma_{n,1}$, $\ldots$,
866: $\gamma_{n,2m_n+1}$ so that $0<\gamma_{n,1}<\gamma_{n,2}\leq
867: \gamma_{n,3}<\ldots<\gamma_{n, 2m_n}\leq \gamma_{n,2m_n+1}<n/4$ and
868: $$I_{n, +}= \bigcup_{i=1}^{m_n}\left( \gamma_{n,2i-1}, \gamma_{n,
869: 2i}\right) \bigcup \left(\gamma_{n, 2m_n+1}, \frac{n}{2}-\gamma_{n,
870: 2m_n+1}\right)\bigcup_{i=1}^{m_n}\left(\frac{n}{2}- \gamma_{n,2i},
871: \frac{n}{2}-\gamma_{n, 2i-1}\right).$$ For $n\leq 9$, Lemma \ref{l1}
872: is not sufficient to show that $\Xi_n(s)<0$ for all $s\in (0,
873: n/2)$. Nevertheless, one can employ the same method as the proof in
874: Lemma \ref{l1} to show that
875: \begin{proposition}\label{p30}
876: $\Xi_9(s)<0$ for all $s\in (0, 9/2)$.
877: \end{proposition}
878:
879:
880:
881: Then the inequality \eqref{eq1015_1} shows that
882: \begin{proposition}\label{p6}
883: For all $1\leq n\leq 9$, the function $\Xi_{n}(s)$ is negative for
884: all $s\in (0, n/2)$.
885: \end{proposition}
886:
887:
888:
889: We give the proof of Proposition \ref{p30} here.
890: \begin{proof}By the reflection formula $\Xi_9(s)=\Xi_9((9/2)-s)$, it
891: is sufficient to show that $\Xi_9(s)<0$ for all $s\in (0, 9/4]$.
892: Let
893: \begin{align*}
894: \mathcal{Z}_1(s)=-\frac{1}{s}-\frac{1}{\frac{9}{2}-s}
895: \end{align*}and
896: \begin{align*}
897: \mathcal{Z}_2(s)=\int_1^{\infty}
898: t^{s-1}\left(\vartheta(t)^9-1\right)dt+\int_1^{\infty}
899: t^{\frac{9}{2}-s-1}\left(\vartheta(t)^9-1\right)dt,
900: \end{align*}so that $\Xi_9(s)=\mathcal{Z}_1(s)+\mathcal{Z}_2(s)$. By taking derivatives with respect to $s$,
901: it is easy to verify that as functions of $s$, $\mathcal{Z}_1(s)$ is
902: increasing on $(0,9/4]$ and $\mathcal{Z}_2(s)$ is decreasing on $[0,
903: 9/4]$. Using the arguments in the proof of Lemma \ref{l1}, we find
904: that
905: \begin{align*}
906: \mathcal{Z}_2(0)\leq& \frac{18}{\pi}
907: e^{-\pi}+\frac{1}{2\pi}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)
908: \\&+\frac{18}{\pi}e^{-\pi}\sum_{j=0}^{4} \frac{\prod_{i=1}^j
909: \left(\frac{9}{2}-i\right)}{\pi^j}+\frac{1}{2\pi}\sum_{j=0}^{4}
910: \frac{\prod_{i=1}^j \left(\frac{9}{2}-i\right)}{(2\pi
911: )^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)\\=&1.2926.
912: \end{align*}Now we want to find an $x$ so that
913: $\mathcal{Z}_1(x)<-1.2926$. By try and error method, we find that
914: $x=0.95$ satisfies the required condition. In fact,
915: \begin{align*}
916: \mathcal{Z}_1(0.95)=-1.3343.
917: \end{align*}Therefore, for all $s\in (0, 0.95]$,
918: \begin{align*}
919: \Xi_9(s)=\mathcal{Z}_1(s)+\mathcal{Z}_2(s)\leq
920: \mathcal{Z}_1(0.95)+\mathcal{Z}_2(0)\leq -0.0417<0.
921: \end{align*}Repeating these steps, we have first
922: \begin{align*}
923: \mathcal{Z}_2(0.95)\leq & \frac{18}{\pi}
924: e^{-\pi}+\frac{1}{2\pi}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)
925: \\&+\frac{18}{\pi}e^{-\pi}\sum_{j=0}^{3} \frac{\prod_{i=1}^j
926: \left(3.55-i\right)}{\pi^j}+\frac{1}{2\pi}\sum_{j=0}^{3}
927: \frac{\prod_{i=1}^j \left(3.55-i\right)}{(2\pi
928: )^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)\\&=0.9728.
929: \end{align*}Then by try and error, we find that
930: \begin{align*}
931: \mathcal{Z}_1(1.55)=-0.9841.
932: \end{align*}Therefore, for all $s\in [0.95, 1.55]$,
933: \begin{align*}
934: \Xi_9(s)=\mathcal{Z}_1(s)+\mathcal{Z}_2(s)\leq
935: \mathcal{Z}_1(1.55)+\mathcal{Z}_2(0.95)\leq -0.0113<0.
936: \end{align*}Repeating again, we find that
937: \begin{align*}
938: \mathcal{Z}_2(1.55)\leq & \frac{18}{\pi}
939: e^{-\pi}\sum_{j=0}^{1}\frac{\prod_{i=1}^j
940: \left(1.55-i\right)}{\pi^j}+\frac{1}{2\pi}\sum_{j=0}^{1}\frac{\prod_{i=1}^j
941: \left(1.55-i\right)}{(2\pi)^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)
942: \\&+\frac{18}{\pi}e^{-\pi}\sum_{j=0}^{2} \frac{\prod_{i=1}^j
943: \left(2.95-i\right)}{\pi^j}+\frac{1}{2\pi}\sum_{j=0}^{2}
944: \frac{\prod_{i=1}^j \left(2.95-i\right)}{(2\pi
945: )^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)\\=&0.8943,
946: \end{align*}and\begin{align*}
947: \mathcal{Z}_1(2)=-0.9.
948: \end{align*}Therefore, for all $s\in [1.55, 2]$,
949: \begin{align*}
950: \Xi_9(s)=\mathcal{Z}_1(s)+\mathcal{Z}_2(s)\leq
951: \mathcal{Z}_1(2)+\mathcal{Z}_2(1.55)\leq -0.0057<0.
952: \end{align*}Finally,
953: \begin{align*}
954: \mathcal{Z}_2(2)\leq & \frac{18}{\pi}
955: e^{-\pi}\sum_{j=0}^{2}\frac{\prod_{i=1}^j
956: \left(2-i\right)}{\pi^j}+\frac{1}{2\pi}\sum_{j=0}^{2}\frac{\prod_{i=1}^j
957: \left(2-i\right)}{(2\pi)^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)
958: \\&+\frac{18}{\pi}e^{-\pi}\sum_{j=0}^{2} \frac{\prod_{i=1}^j
959: \left(2.5-i\right)}{\pi^j}+\frac{1}{2\pi}\sum_{j=0}^{2}
960: \frac{\prod_{i=1}^j \left(2.5-i\right)}{(2\pi
961: )^j}\left(\left(\frac{e^{\pi}+1}{e^{\pi}-1}\right)^9-1-18e^{-\pi}\right)\\
962: =&0.8649,
963: \end{align*}and
964: \begin{align*}
965: \mathcal{Z}_1(2.25)=-0.8889,
966: \end{align*}which implies that for all $s\in[2, 2.25]$,
967: \begin{align*}\Xi_9(s)=\mathcal{Z}_1(s)+\mathcal{Z}_2(s)\leq
968: \mathcal{Z}_1(2.25)+\mathcal{Z}_2(2)\leq -0.0240<0.
969: \end{align*}This completes the proof of the proposition.
970: \end{proof}
971:
972: \begin{figure}\centering \epsfxsize=.49\linewidth
973: \epsffile{graph1.eps} \epsfxsize=.49\linewidth
974: \epsffile{graph2.eps}\caption{The graphs of
975: $\hat{\Xi}_n(s)=\Xi_n\left(\frac{ns}{2}\right)$ for $s\in (0,1)$.}
976: \end{figure}
977: In Figure 1, we show the
978: graphs of $\hat{\Xi}_n(s)=\Xi_n(ns/2)$ for $s\in (0, 1)$ and $1\leq
979: n\leq 10$. From the graphs, we see that for $1\leq n\leq 10$,
980: $\hat{\Xi}_n(s)$ is increasing for $s\in (0, 1/2)$, decreasing for
981: $s\in (1/2, 1)$, and $s=1/2$ is a local maximum of $\hat{\Xi}_n(s)$.
982: One would tend to jump into the conclusion that this will be true
983: for all $n$. However, this is not the case. Since $\hat{\Xi}_n(s)$
984: satisfies the functional equation $\hat{\Xi}_n(s)=\hat{\Xi}_n(1-s)$,
985: it is easy to verify that $\hat{\Xi}_n'(1/2)=0$ and therefore
986: $s=1/2$ is indeed a local extremum of $\hat{\Xi}_n(s)$. In order to
987: determine the nature of this extremum, it is necessary to look at
988: the second derivative of $\hat{\Xi}_n(s)$. We obtain from
989: \eqref{eq1015_10} that
990: \begin{align}\label{eq1015_11}
991: \hat{\Xi}^{\prime\prime}_n(s)=&-\frac{4}{n}\left(\frac{1}{s^3}+\frac{1}{(1-s)^3}\right)
992: +\left(\frac{n}{2}\right)^2\int_1^{\infty} t^{\frac{ns}{2}-1}(\log
993: t)^2 \left(\vartheta(t)^n-1\right)dt\\&+\left(\frac{n}{2}\right)^2
994: \int_1^{\infty}t^{\frac{n}{2}(1-s)-1}\left(\vartheta(t)^n-1\right)dt.\nonumber
995: \end{align}From here, it is easy to verify that
996: \begin{align}\label{eq1015_12}\hat{\Xi}^{\prime\prime}_{n+1}(s)
997: >\hat{\Xi}^{\prime\prime}_n(s).\end{align} Using the same argument as
998: employed in showing that for any fixed $s\in (0, n/2)$,
999: $\hat{\Xi}_n(s)>0$ when $n$ is large enough, we can use
1000: \eqref{eq1015_11} to prove that for any fixed $s\in (0,n/2)$,
1001: $\Xi_n^{\prime\prime}(s)>0$ whenever $n$ is large enough. This
1002: implies that the point $n/4$ will become a local minimum of
1003: $\Xi_n(s)$ when $n$ is large enough. In fact, using numerical
1004: computation, we find that
1005: \begin{figure}\centering \epsfxsize=.49\linewidth
1006: \epsffile{graph3.eps} \epsfxsize=.49\linewidth
1007: \epsffile{graph4.eps} \caption{The graphs of
1008: $\hat{\Xi}_{11}(s)=\Xi_{11}\left(\frac{11s}{2}\right)$ and
1009: $\hat{\Xi}_{12}(s)=\Xi_{12}(6s)$.}
1010: \end{figure}
1011: \begin{align*}
1012: \Xi^{\prime\prime}_{10}\left(\frac{5}{2}\right)=-0.101080515709,\\
1013: \Xi^{\prime\prime}_{11}\left(\frac{11}{4}\right)=0.009568954836,
1014: \end{align*}which together with \eqref{eq1015_12}, imply that for $1\leq n\leq 10$,
1015: $s=n/4$ is a local maximum of $\Xi_n(s)$; whereas for $n\geq 11$,
1016: $s=n/4$ is a local minimum of $\Xi_n(s)$. In Figure 2, we show
1017: graphically that $s=1/2$ is a local minimum of $\hat{\Xi}_{11}(s)$
1018: and $\hat{\Xi}_{12}(s)$.
1019:
1020: In Table 1, we tabulate the
1021: interval $I_{n,+}, 10\leq n\leq 21$ where $\Xi_n(s)$ is positive.
1022:
1023: \vspace{0.2cm} \noindent Table 1: The interval $I_{n,+}$ for $10\leq
1024: n\leq 21$.
1025:
1026: \vspace{0.2cm}\noindent
1027: \begin{tabular}{|c|c||c|c|}
1028: \hline
1029: \hspace{0.5cm}$n$\hspace{0.5cm} & $I_{n,+}$ & \hspace{0.5cm} $n$ \hspace{0.5cm} & $I_{n,+}$\\
1030: \hline
1031: 10 & \hspace{0.5cm}(1.0899, 3.9101)\hspace{0.5cm} & 11 & \hspace{0.5cm} (0.6401, 4.8599)\hspace{0.5cm} \\
1032: 12 & (0.3976, 5.6024) & 13 & \hspace{0.5cm}(0.2498, 6.2502)\hspace{0.5cm} \\
1033: 14 & (0.1562, 6.8438) & 15 & \hspace{0.5cm}(0.0964, 7.4036)\hspace{0.5cm} \\
1034: 16 & (0.0585, 7.9415) & 17 & \hspace{0.5cm}(0.0348, 8.4652)\hspace{0.5cm} \\
1035: 18 & (0.0202, 8.9798) & 19 & \hspace{0.5cm}(0.0115, 9.4885)\hspace{0.5cm} \\
1036: 20 & (0.0064, 9.9936) & 21 & \hspace{0.5cm}(0.0034, 10.4966)\hspace{0.5cm} \\
1037: \hline
1038: \end{tabular}
1039:
1040: \vspace{0.5cm} One notices that for $10\leq n\leq 21$, $m_n=0$,
1041: $I_{n,+}=(\gamma_{n,1}, (n/2)-\gamma_{n,1})$, $\gamma_{n,1}$ is
1042: decreasing and $I_{n,+}\subseteq I_{n+1, +}$. In fact, from the
1043: formula \eqref{eq1015_3}, it is easy to verify that for $s\in (0,
1044: n/2)$,
1045: $$\Xi_{n+1}(s)>\Xi_n(s).$$ This shows that $I_{n,+}\subseteq I_{n+1,
1046: +}$ for all $n\geq 10$. From this table, it is also natural to
1047: conjecture that $I_{n,+}$ is an open connected interval with center
1048: at $n/4$ for all $n\geq 10$. However, to prove this would require a
1049: very detailed analysis of $\Xi_n(s)$ and its higher derivatives. We
1050: do not intend to deal further into this problem here.
1051:
1052: We now return to the discussion about the sign of the general
1053: Xi-function $\Xi_n(s;a_1,\ldots, a_n)$ in the range $s\in (0, n/2)$.
1054: Since we have shown in Theorem \ref{th1} that for any fixed $s$, the
1055: point $a_1=\ldots=a_n$ is the minimum of $\Xi_n(s;a_1,\ldots, a_n),
1056: (a_1,\ldots, a_n)\in(\R^n)^+$ with $\prod_{i=1}^n a_n$ fixed, now
1057: together with Proposition \ref{p5}, Proposition \ref{p6} and
1058: Proposition \ref{p2}, we obtain
1059: \begin{proposition}
1060: For $1\leq n\leq 9$, there exists a nonempty region containing the
1061: ray $a_1=\ldots=a_n$ in $(\R^n)^+$ where $\Xi_n(s;a_1,\ldots, a_n)$
1062: is negative for all $s\in (0,n/2)$.
1063: \end{proposition}
1064:
1065: \begin{proposition}
1066: If $n\geq 10$, then for every $(a_1,\ldots, a_n)\in (\R^+)^n$, there
1067: exists a nonempty open subset of $(0, n/2)$ where $\Xi_n(s;
1068: a_1,\ldots, a_n)$ is positive and a nonempty open subset of $(0,
1069: n/2)$ where $\Xi_n(s; a_1,\ldots, a_n)$ is negative.
1070:
1071:
1072: \end{proposition}This proposition implies that
1073: generalized Riemann hypothesis is not true for all the Epstein zeta
1074: function of the form $Z_{n}(s; a_1,\ldots, a_n)$ (i.e. Epstein zeta
1075: function of positive definite diagonal quadratic forms) when $n\geq
1076: 10$.
1077:
1078: Now consider the Epstein zeta function as a function of
1079: $(a_1,\ldots, a_n)\in(\R^+)^n$ with $s$ fixed, we can conclude with
1080: the help of Propositions \ref{p4} and \ref{p6} that
1081: \begin{proposition}\label{p7}
1082: If $1\leq n\leq 9$, then for any fixed $s\in (0, n/2)$, there
1083: exists a nonempty open region $\Omega_{s,n}^+$ of $(a_1,\ldots,
1084: a_n)\in(\R^+)^n$ where $\Xi_{n}(s; a_1,\ldots, a_n)$ is positive and
1085: a nonempty open region $\Omega_{s,n}^-$ of $(a_1,\ldots,
1086: a_n)\in(\R^+)^n$ where $\Xi_{n}(s; a_1,\ldots, a_n)$ is negative.
1087: The ray $a_1=\dots=a_n$ lies inside the region $\Omega_{s,n}^-$.
1088: \end{proposition}
1089:
1090: \begin{proposition}\label{p8}
1091: If $ n\geq 10$, then if $s\in I_{n,+}$ (eq. \eqref{eq1016_1}), the
1092: function $\Xi_n(s; a_1,\ldots, a_n)$ is positive for all $
1093: (a_1,\ldots, a_n)\in (\R^+)^n$; whereas if $s\in I_{n,-}$ (eq.
1094: \eqref{eq1017_1}), there exists a nonempty open region
1095: $\Omega_{s,n}^+$ of $(a_1,\ldots, a_n)\in(\R^+)^n$ where $\Xi_{n}(s;
1096: a_1,\ldots, a_n)$ is positive and a nonempty open region
1097: $\Omega_{s,n}^-$ of $(a_1,\ldots, a_n)\in(\R^+)^n$ where $\Xi_{n}(s;
1098: a_1,\ldots, a_n)$ is negative. In the latter case, the ray
1099: $a_1=\dots=a_n$ lies inside the region $\Omega_{s,n}^-$.
1100: \end{proposition}
1101:
1102: Propositions \ref{p7} and \ref{p8} are important for
1103: our study of symmetry breaking of interacting fractional
1104: Klein--Gordon field in \cite{e10}. In the following section, it will
1105: be shown that the region $\Omega_{s,n}^-$ is connected. Moreover,
1106: under the map $(a_1,\ldots, a_n)\in (\R^+)^n\mapsto (\log
1107: a_1,\ldots, \log a_n)\in \R^n$, the projection of its image to the
1108: plane $x_1+\ldots+x_n=$ constant is a convex set.
1109: \section{Convexity of the Epstein zeta function}
1110:
1111: In this section, we show that \begin{theorem}\label{th2}For any
1112: fixed $s$, when $\sum_{i=1}^n c_i =c$ is fixed, the function
1113: \begin{align}\label{eq1017_2}(c_1,\ldots, c_n) \mapsto &\Xi_n\left(s; e^{c_1},\ldots,
1114: e^{c_n}\right)\nonumber\\&=-\frac{e^{\frac{c}{2}}}{s}-\frac{e^{-\frac{c}{2}}}{\frac{n}{2}-s}
1115: +e^{\frac{c}{2}}\int_1^{\infty}t^{s-1}\left(\prod_{i=1}^n
1116: \vartheta(e^{2c_i}t)-1\right)dt\\
1117: &+e^{-\frac{c}{2}}\int_1^{\infty}t^{\frac{n}{2}-s-1}\left(\prod_{i=1}^n
1118: \vartheta(e^{-2c_i}t)-1\right)dt,\nonumber\end{align}regarded as a
1119: function of any $(n-1)$ of the variables $\{c_1,\ldots,c_n)$, is a
1120: convex function.\end{theorem} Recall that a function $f(x)$, $x\in
1121: \R^n$ is convex if and only if for any $\lambda\in (0,1)$ and any
1122: two points $x, y\in\R^n$,
1123: \begin{align*}
1124: f\left(\lambda x+(1-\lambda)y\right)\leq \lambda
1125: f(x)+(1-\lambda)f(y).
1126: \end{align*}If $f$ has continuous second partial derivatives, then
1127: $f$ is convex if and only if the Hessian of $f$,
1128: \begin{align*}
1129: H_f(x_1,\ldots, x_n)=\left(\frac{\pa^2f}{\pa x_i\pa
1130: x_j}\right)_{\substack{1\leq i\leq n\\1\leq j\leq n}},
1131: \end{align*} is a positive semi-definite matrix. One way to prove that a Hermitian matrix is positive semi-definite is
1132: to use the Sylvester criterion, which says that an $n\times n$
1133: Hermitian matrix is positive semi-definite if and only if all its
1134: leading principal minors are nonnegative. For $1\leq i\leq n$, the
1135: $i^{\text{th}}$ leading principal minor of an $n\times n$ matrix $A$
1136: is the determinant of the $i\times i$ square matrix on the upper
1137: left corner of $A$.
1138:
1139:
1140: In the
1141: formula \eqref{eq1017_2}, the dependence on $(c_1,\ldots, c_n)$
1142: only comes from the terms $\prod_{i=1}^n\vartheta(e^{2c_i}t)$ and
1143: $\prod_{i=1}^n\vartheta(e^{-2c_i}t)$. Notice that if a function
1144: $f(x_1,\ldots, x_n)$ is convex, then under any affine change of
1145: coordinates $$x_i=\sum_{j=1}^n l_{ij} y_j+d_i, \hspace{1cm} 1\leq
1146: i\leq n,$$ the function $$\hat f(y_1,\ldots,
1147: y_n)=f\left(\sum_{j=1}^n l_{1j}y_j+d_1,\ldots,\sum_{j=1}^n
1148: l_{nj}y_j+d_n\right)$$ is also convex. Moreover, the projection of
1149: $f$ to any hyperplane $\sum_{i=1}^n b_i x_i=b$ is also convex.
1150: Therefore, if the function
1151: \begin{align*}
1152: \Theta_n(x_1,\ldots, x_n)=\prod_{i=1}^n
1153: \vartheta\left(e^{x_i}\right)=\prod_{i=1}^n \Theta_1(x_i),
1154: \hspace{1cm}(x_1,\ldots, x_n)\in\R^n
1155: \end{align*}is convex, then $\prod_{i=1}^n\vartheta(e^{2c_i}t)$ and
1156: $\prod_{i=1}^n\vartheta(e^{-2c_i}t)$, regarded as functions of
1157: $\{c_1,\ldots,c_{n-1}\}$ (so that $c_n =c-\sum_{i=1}^{n-1}c_i$),
1158: are convex. This will show that
1159: $\Xi_n(s; e^{c_1},\ldots, e^{c_n})$ is a convex function of
1160: $(c_1,\ldots,c_{n-1})$. Therefore to prove Theorem \ref{th2}, we
1161: only need to show that for all $n\geq 1$, the function
1162: $\Theta_n(x_1,\ldots, x_n)$ is convex.
1163: For $n= 1$, we use the following fact:
1164: \begin{lemma}\label{l10}
1165: If $g:\R\rightarrow \R$ is a convex function, then the function $f$
1166: defined by $$f(x)=e^{g(x)},$$is also convex.
1167: \end{lemma}\begin{proof}This is a simple consequence of the fact that
1168: the exponential function $e^x$ is a convex function, and composition
1169: of convex functions are convex.
1170: \end{proof}
1171:
1172: Using this lemma, we can immediately conclude from Proposition
1173: \ref{p1} that
1174: \begin{lemma}\label{l3}
1175: The function $$\Theta_1(x)=\vartheta\left(e^x\right),
1176: \hspace{1cm}x\in\R$$ is convex.
1177: \end{lemma}
1178:
1179: For $n\geq 2$, in view of the definition of $\Theta_n$, we are led
1180: to the study of a general function of the type
1181: \begin{align}\label{eq1017_11}
1182: \mathcal{F}_n(x_1, \ldots, x_n)=\prod_{i=1}^n f(x_i),
1183: \end{align}where $f:\R\rightarrow \R^+$ is a positive function of one variable.
1184: For such functions, its Hessian is given by
1185: \begin{align*}
1186: H_{\mathcal{F}_n}(x_1,\ldots,x_n) =\mathcal{F}_n(x_1,\ldots,
1187: x_n)^n\begin{pmatrix} \frac{f^{\prime\prime}(x_1)}{f(x_1)} &
1188: \frac{f'(x_1)}{f(x_1)}\frac{f'(x_2)}{f(x_2)}&\ldots &
1189: \frac{f'(x_1)}{f(x_1)}\frac{f'(x_n)}{f(x_n)}\\
1190: \frac{f'(x_2)}{f(x_2)}\frac{f'(x_1)}{f(x_1)} &
1191: \frac{f^{\prime\prime}(x_2)}{f(x_2)}&\ldots &
1192: \frac{f'(x_2)}{f(x_2)}\frac{f'(x_n)}{f(x_n)}\\
1193: \vdots &\vdots & & \vdots\\
1194: \frac{f'(x_n)}{f(x_n)}\frac{f'(x_1)}{f(x_1)}&\frac{f'(x_n)}{f(x_n)}\frac{f'(x_2)}{f(x_2)}
1195: &\ldots &\frac{f^{\prime\prime}(x_n)}{f(x_n)}\end{pmatrix}.
1196: \end{align*}By the Sylvester criterion, to show that $\mathcal{F}_n$
1197: is convex, it is sufficient to show that all the leading principal
1198: minors of
1199: \begin{align}\label{eq1017_12}
1200: H_n(x_1,\ldots, x_n)=\begin{pmatrix}
1201: \frac{f^{\prime\prime}(x_1)}{f(x_1)} &
1202: \frac{f'(x_1)}{f(x_1)}\frac{f'(x_2)}{f(x_2)}&\ldots &
1203: \frac{f'(x_1)}{f(x_1)}\frac{f'(x_n)}{f(x_n)}\\
1204: \frac{f'(x_2)}{f(x_2)}\frac{f'(x_1)}{f(x_1)} &
1205: \frac{f^{\prime\prime}(x_2)}{f(x_2)}&\ldots &
1206: \frac{f'(x_2)}{f(x_2)}\frac{f'(x_n)}{f(x_n)}\\
1207: \vdots &\vdots & & \vdots\\
1208: \frac{f'(x_n)}{f(x_n)}\frac{f'(x_1)}{f(x_1)}&\frac{f'(x_n)}{f(x_n)}\frac{f'(x_2)}{f(x_2)}
1209: &\ldots &\frac{f^{\prime\prime}(x_n)}{f(x_n)}\end{pmatrix}
1210: \end{align}are nonnegative. Notice that the $i^{\text{th}}$ leading
1211: principal minor of $H_n(x_1,\ldots,x_n)$ is the determinant of
1212: $H_i(x_1,\ldots, x_i)$. Therefore, one only have to show that $\det
1213: H_n\geq 0$ for all $n\geq 1$. For $f(x)=\vartheta(e^x)$, the $n=1$
1214: case is already proved in Lemma \ref{l3}. For $n\geq 2$, to further
1215: simplify our problem, notice that for fixed $(x_1,\ldots,x_n)\in
1216: \R^n$, the matrix $H_n$ has the form
1217: \begin{align*}
1218: J_n =\begin{pmatrix} z_1& w_1 w_2 & \ldots & w_1 w_n\\
1219: w_2w_1 & z_2 &\ldots & w_2w_n\\
1220: \vdots &\vdots & & \vdots\\
1221: w_n w_1 & w_n w_2 & \ldots & z_n\end{pmatrix}
1222: \end{align*}for some $2n$ real numbers $z_1,\ldots, z_n$ and
1223: $w_1,\ldots, w_n$. When $n=2$, the determinant of $J_2$ is
1224: \begin{align}\label{eq1017_6}
1225: \det J_2= z_1 z_2- w_1^2
1226: w_2^2=(z_1-w_1^2)(z_2-w_2^2)+w_1^2(z_1-w_1^2)+w_2^2(z_2-w_2^2).
1227: \end{align}The reason for expressing the determinant in this form is
1228: inspired by the fact that for
1229: $z_i=\frac{f^{\prime\prime}(x_i)}{f(x_i)}$,
1230: $w_i=\frac{f'(x_i)}{f(x_i)}$,
1231: \begin{align}\label{eq1017_13}z_i-w_i^2=\frac{d^2}{dx_i^2}\log f(x_i).\end{align}Eq.
1232: \eqref{eq1017_6} implies that if $z_i-w_i^2\geq 0$, then $\det
1233: J_2\geq 0$. This gives
1234: \begin{lemma}\label{l4}
1235: Let $f:\R\rightarrow \R^+$ be a twice continuously differentiable
1236: positive function. If $\log f(x)$ is a convex function, then the
1237: function $\mathcal{F}_2(x_1, x_2)=f(x_1)f(x_2)$ is convex.
1238: \end{lemma}
1239: \begin{proof}
1240: Recall that $\mathcal{F}_2(x_1, x_2)$ is convex if and only if
1241: $$\det H_1(x_1)=\frac{f^{\prime\prime}(x_1)}{f(x_1)}\geq 0$$ and
1242: $$\det H_2(x_1, x_2)=\det
1243: J_2\left[z_1=\frac{f^{\prime\prime}(x_1)}{f(x_1)},
1244: z_2=\frac{f^{\prime\prime}(x_2)}{f(x_2)}; w_1=\frac{f^{\prime
1245: }(x_1)}{f(x_1)}, w_2=\frac{f^{\prime }(x_2)}{f(x_2)}\right]\geq 0.$$
1246: As discussed above, if $\log f(x)$ is convex, then $\det H_2(x_1,
1247: x_2)\geq 0$. On the other hand, Lemma \ref{l10} implies that $f(x)$
1248: is also convex. Therefore, $\det H_1(x_1)\geq 0$. This completes the
1249: proof.
1250: \end{proof}
1251:
1252: For the case under consideration, $f(x)=\vartheta(e^x)$, and it is
1253: shown in Proposition
1254: \ref{p1} that $\log \vartheta(e^x)$ is a
1255: convex function. Therefore, Lemma \ref{l4} implies that
1256: \begin{lemma}\label{l5}
1257: The function $$\Theta_2(x_1,
1258: x_2)=\vartheta(e^{x_1})\vartheta(e^{x_2}), \hspace{1cm}(x_1, x_2)\in
1259: \R^2$$ is convex.
1260: \end{lemma}
1261:
1262: Returning to the general case. First we claim the following, which
1263: is a generalization of \eqref{eq1017_6}.
1264: \begin{proposition}\label{p10}
1265: For $n\geq 2$,
1266: \begin{align}\label{eq1017_14}
1267: \det J_n =\det \begin{pmatrix} z_1& w_1 w_2 & \ldots & w_1 w_n\\
1268: w_2w_1 & z_2 &\ldots & w_2w_n\\
1269: \vdots &\vdots & & \vdots\\
1270: w_n w_1 & w_n w_2 & \ldots & z_n\end{pmatrix}=\prod_{i=1}^n
1271: (z_i-w_i^2)+\sum_{i=1}^n w_i^2\left[\prod_{j\neq
1272: i}\left(z_j-w_j^2\right)\right].
1273: \end{align}Therefore, if $z_i\geq w_i^2$ for all $1\leq i\leq n$, then $\det
1274: J_n\geq 0$.
1275: \end{proposition}We would like to discuss the consequence of this
1276: proposition first, and give the proof later. From the above
1277: discussion, the function $\mathcal{F}_n$ of the form
1278: \eqref{eq1017_11} is convex if and only if all the determinants of
1279: $H_i$ (eq. \eqref{eq1017_12}), $1\leq i\leq n$ are nonnegative. Now
1280: $H_n$ is a matrix of the form $J_n$, with $z_i -w_i^2$ given by
1281: \eqref{eq1017_13}. Therefore, we conclude from Proposition \ref{p10}
1282: that
1283: \begin{proposition}
1284: Let $f:\R\rightarrow \R^+$ be a twice continuously differentiable
1285: positive function. If $\log f(x)$ is a convex function, then the
1286: function $$\mathcal{F}_n(x_1,\ldots, x_n)=\prod_{i=1}^nf(x_i)$$ is
1287: convex.
1288: \end{proposition}As in Lemma \ref{l5}, this
1289: proposition implies that
1290: \begin{proposition}
1291: For any $n\geq 1$, the function $$\Theta_n(x_1,\ldots,
1292: x_n)=\prod_{i=1}^n \vartheta(e^{x_i}), \hspace{1cm} (x_1,\ldots,
1293: x_n)\in \R^n,$$ is a convex function.
1294: \end{proposition}As discussed in the beginning of this section, this
1295: concludes the proof of Theorem \ref{th2}. Therefore, what is left is
1296: the proof of Proposition \ref{p10}.
1297:
1298: \begin{proof}By factoring out $w_i$ from row $i$ and column $i$, we have
1299: \begin{align*}
1300: \det J_n =\left[\prod_{i=1}^n w_i^2\right]^2
1301: \mathcal{J}_n\left(\frac{z_1}{w_1^2},\frac{z_2}{w_2^2},\ldots,
1302: \frac{z_n}{w_n^2}\right),
1303: \end{align*}where \begin{align*}
1304: \mathcal{J}_n(\alpha_1,\ldots, \alpha_n)=\det \begin{pmatrix}\alpha_1 & 1 &1 &\ldots & 1\\
1305: 1 &\alpha_2 & 1 &\ldots & 1\\
1306: 1 & 1 & \alpha_3 & \ldots & 1\\
1307: \vdots & \vdots & \vdots & &\vdots\\
1308: 1 & 1 & 1 & \ldots & \alpha_n\end{pmatrix}.
1309: \end{align*}By elementary row operation, we find that
1310: \begin{align*}
1311: \mathcal{J}_n(\alpha_1,\ldots, \alpha_n)=&\det \begin{pmatrix}\alpha_1 & 1 &1 &\ldots & 1\\
1312: 1 &\alpha_2 & 1 &\ldots & 1\\
1313: 1 & 1 & \alpha_3 & \ldots & 1\\
1314: \vdots & \vdots & \vdots & &\vdots\\
1315: 1 & 1 & 1 & \ldots & \alpha_n\end{pmatrix}\\=&\det \begin{pmatrix}\alpha_1-1 & 1-\alpha_2 &0 &\ldots & 0\\
1316: 0 &\alpha_2-1 & 1-\alpha_3 &\ldots & 0\\
1317: 0 & 0 & \alpha_3-1 & \ldots & 0\\
1318: \vdots & \vdots & \vdots & &\vdots\\
1319: 1 & 1 & 1 & \ldots & \alpha_n\end{pmatrix}\\
1320: =&\left(\alpha_1-1\right)\mathcal{J}_{n-1}\left(\alpha_2,\ldots,
1321: \alpha_n\right)+(-1)^{n-1}\prod_{i=2}^n(1-\alpha_i).
1322: \end{align*}It then follows by induction that
1323: \begin{align*}
1324: \mathcal{J}_n(\alpha_1,\ldots, \alpha_n)=\prod_{i=1}^n (\alpha_i-1)
1325: +\sum_{i=1}^n \prod_{j\neq i}(\alpha_j-1).
1326: \end{align*}This proves the proposition.
1327: \end{proof}
1328:
1329: Next we discuss the consequences of Theorem \ref{th2}. Let $j$ be a
1330: positive integer less than or equal to $n-1$, $A$ an $n\times j$
1331: matrix such that the sum of every column vanishes, i.e.,
1332: \begin{align}\label{eq1018_1}\sum_{i=1}^n A_{ij}=0, \end{align}and let $v$ be a vector in $\R^n$.
1333: Define a $j$-variable function ${\Xi}_{n,A,v}(s; b_1,\ldots, b_j)$
1334: by
1335: \begin{align*}
1336: {\Xi}_{n,A,v}(s; b_1,\ldots, b_j) = \Xi_n\left(s;
1337: \exp\left[v_1+\sum_{l=1}^j A_{1l}b_l \right],\ldots,
1338: \exp\left[v_n+\sum_{l=1}^j A_{nl}b_l \right]\right).
1339: \end{align*}Notice that the condition \eqref{eq1018_1} implies
1340: \begin{align*}
1341: \prod_{i=1}^n \exp\left[v_i+\sum_{l=1}^j A_{il}b_l
1342: \right]=\exp\left[\sum_{i=1}^n v_i\right]=\text{constant}.
1343: \end{align*}Since affine change of coordinates does not affect the
1344: convexity of a function, we conclude from Theorem \ref{th2} that
1345: $\Xi_{n,A,v}(s;b_1,\ldots, b_j)$ is a convex function of
1346: $(b_1,\ldots, b_j)$. As an example, if $v=0$ and $A$ is the $n\times
1347: (n-1)$ matrix
1348: \begin{align}\label{eq1018_3}
1349: A=\begin{pmatrix} 1 & 0 &\ldots & 0\\
1350: 0 & 1 & \ldots & 0\\
1351: \vdots & \vdots & &\vdots\\
1352: 0 & 0 & \ldots & 1\\
1353: -1 & -1 & \ldots & -1\end{pmatrix},
1354: \end{align}then $\Xi_{n,A,v}(b_1,\ldots, b_n)$ is just the function defined in
1355: \eqref{eq1017_2} with $b_i=c_i$ for $1\leq i\leq n-1$ and
1356: $\sum_{i=1}^n c_n=0$.
1357:
1358: By the convexity of $\Xi_{n,A,v}\left(s; b_1,\ldots, b_j\right)$, we
1359: can conclude that the region $\Omega_{s, A, v}^-$ of $(b_1,\ldots,
1360: b_j)\in \R^j$ where $\Xi_{n,A,v}\left(s; b_1,\ldots, b_j\right)<0$
1361: is a convex, and therefore connected region. Using the fact that at
1362: fixed $\prod_{i=1}^na_i$, the minimum of $\Xi_n(s; a_1,\ldots, a_n)$
1363: appears at $a_1=\ldots=a_n$, one can even conclude that if
1364: $\Omega_{s, A, v}^-$ is nonempty and if $b=\hat{b}$ is a solution of
1365: the system
1366: \begin{align}\label{eq1018_4}
1367: Ab+v=\left[\frac{1}{n}\sum_{i=1}^n
1368: v_i\right]\begin{pmatrix}1\\1\\\vdots\\1\end{pmatrix},
1369: \end{align} then $\hat{b}\in
1370: \Omega_{s,A,v}^-$. When $A$ is given by \eqref{eq1018_3} and $v=0$,
1371: the system \eqref{eq1018_4} has a unique solution $b=0$. Therefore,
1372: the region of $(c_1,\ldots, c_{n-1})\in \R^n$ where the function
1373: $\Xi_n\left(s; e^{c_1}, \ldots, e^{c_{n-1}},
1374: e^{c-c_1-\ldots-c_{n-1}}\right)$ \eqref{eq1017_2} is negative is a
1375: convex connected region containing the origin
1376: $c_1=\ldots=c_{n-1}=0$.
1377:
1378:
1379: As a second example, suppose the variables $(a_1,\ldots, a_n)$ in
1380: $\Xi_n(s; a_1,\ldots, a_n)$ are such that $a_1:\ldots : a_n
1381: =1:k_2:\ldots:k_n$ and $\prod_{i=1}^n a_i=1$. A simple computation
1382: shows that under these conditions, $a_1,\ldots, a_n$ can be
1383: expressed as functions of $(k_2, \ldots, k_n)$:
1384: \begin{align*}
1385: a_1=\frac{1}{ \prod_{i=2}^n k_i^{\frac{1}{n}}},
1386: \hspace{0.5cm}a_2=\frac{k_2^{\frac{n-1}{n}}}{\prod_{i=3}^n
1387: k_i^{\frac{1}{n}}},\hspace{0.5cm}\ldots,\hspace{0.5cm}
1388: a_n=\frac{k_n^{\frac{n-1}{n}}}{\prod_{i=2}^{n-1}k_i^{\frac{1}{n}}}.
1389: \end{align*}Therefore the function $\Xi_n(s; a_1,\ldots, a_n)$ can
1390: be regarded as a function of $\log k_i$, $2\leq i\leq n$ with
1391: corresponding $A$ and $v$ given by
1392: $$A=\begin{pmatrix}-\frac{1}{n} & -\frac{1}{n} & \ldots & -\frac{1}{n}\\
1393: \frac{n-1}{n} & -\frac{1}{n} & \ldots & -\frac{1}{n}\\
1394: \vdots & \vdots & & \vdots\\
1395: -\frac{1}{n} & -\frac{1}{n} & \ldots &\frac{n-1}{n}
1396: \end{pmatrix}$$ and $v=0$. Consequently, the region $\Omega_{s,n}^-$
1397: where $\Xi_n(s; a_1,\ldots, a_n)<0$, plotted with respect to the
1398: variables $\log k_2,\ldots, \log k_n$ is a convex and connected
1399: region containing the point $\log k_2=\ldots=\log k_n=0$. Reducing
1400: the number of variables by setting some of the variables $k_i$ to be
1401: equal to a constant or setting $k_i=k_j$ for some pairs of $i\neq j$
1402: are tantamount to restricting the variables $(\log k_2, \ldots, \log
1403: k_n)$ to the intersections of hyperplanes in $\R^{n-1}$. Therefore
1404: plotting the region $\Omega_{s,n}^-$ with respect to the remaining
1405: $\log k_i$ variables, the new region is still convex and connected.
1406: In \cite{e10}, we have plotted the regions where $\Xi_3(s; a_1, a_2,
1407: a_3)<0$ with respect to the variables $\log k_2$ and $\log k_3$, and
1408: the regions where $\Xi_4(s; a_1,a_2,a_3, a_4)<0$ with respect to the
1409: variables $\log k_2, \log k_3$, with $k_4=1$ and $k_4=3$, for some
1410: values of $s$. The graphs show that these regions are indeed convex
1411: and connected. As a matter of fact, when the regions
1412: $\Omega_{s,n}^-$ are plotted using computer softwares, we can only
1413: determine the regions in finite domains of the $\log k$ variables.
1414: Our convexity and connectivity results guarantee that there does not
1415: exists region of $\Omega_{s,n}^-$ outside the finite domain we
1416: consider.
1417:
1418: Finally, notice that the map $(a_1,\ldots, a_n)\in (\R^+)^n \mapsto
1419: (\log a_1,\ldots, \log a_n)\in \R^n$ is continuous. On the other
1420: hand, we have shown that the intersection of the region
1421: $\Omega_{s,n}^-$ where $\Xi_{n}(s; a_1,\ldots, a_n)<0$ with any
1422: hypersurface in $(\R^+)^n$ of the form $\prod_{i=1}^n a_i=a$ a
1423: constant, contains the point $a_1=\ldots=a_n=a^{\frac{1}{n}}$. These
1424: allow one to conclude that the region $\Omega^-_{s,n}$, as a region
1425: of $(a_1,\ldots, a_n)\in (\R^+)^n$, is a connected region that
1426: contains the ray $a_1=\ldots=a_n$.
1427:
1428: \section{Concluding remarks}
1429:
1430: This paper is motivated by our recent work \cite{e10}, in which we
1431: need to determine the conditions for which symmetry breaking occurs
1432: in the $\lambda\varphi^4$--interacting fractional Klein--Gordon
1433: field theory on a toroidal spacetime $T^n\times \R^{N}$. The results
1434: obtained in this work are used in \cite{e10}. Since the Epstein zeta
1435: function $Z_n(s; a_1,\ldots, a_n)$ always appears when one studies
1436: field theories on toroidal manifolds or rectangular cavities, the
1437: results presented here may have potential applications for other
1438: works along these directions. For physics applications, one is
1439: usually interested in simple toroidal manifolds of the form
1440: $T^n\times \R^N$, which can be considered as quotients of $\R^n$ by
1441: rectangular lattices. This explains why we consider Epstein zeta
1442: function of the form $Z_n(s; a_1,\ldots, a_n), a\in (\R^+)^n$
1443: instead of the general Epstein zeta function $Z_n(A;s)$ with $A$ any
1444: positive definitive symmetric matrix. An advantage of this
1445: simplification is that we can determine the minimum of $Z_{n}(s;
1446: a_1,\ldots, a_n), a\in (\R^+)^n, \prod_{i=1}^n a_i=1$ for all
1447: $s\in\R\setminus \{0,n/2\}$ and all $n\in \mathbb{N}$. In contrast,
1448: for general Epstein zeta function $Z_n(A;s)$ with $\det A=1$, only
1449: some local minima have been determined for $n=2,3,4,5,6,7,8,24$ with
1450: $s$ in some range of $\R$ [15--38]. In fact, for all these known
1451: cases, the minimum of $Z_n(s; a_1, \ldots, a_n), a\in (\R^+)^n,
1452: \prod_{i=1}^n a_i=1$ is no longer a local minimum in the larger
1453: domain where $A\in \left\{\text{positive definite
1454: symmetric}\;n\times n\;\text{matrices}\right\}$. In the attempt to
1455: search for the correct spacetime model, it might turn out that the
1456: general toroidal manifold $\R^n/L$, where $L$ is any lattice in
1457: $\R^n$, will be of importance to physics. In that case, we need to
1458: extend the work of this paper to general Epstein zeta function
1459: $Z_n(A;s)$. It is hoped that the methods and results in this paper
1460: will be useful for the study of this general problem.
1461:
1462: Our result about the convexity of the Epstein zeta function $Z_n(s;
1463: a_1,\ldots, a_n)$, as a function of $\log a_1,\ldots, \log a_n$ with
1464: $\log a_1+\ldots+\log a_n$ fixed, seems to be new. This result is
1465: important for determining the regions where $Z_n(s; a_1,\ldots,
1466: a_n)$ is positive or negative, as discussed at the end of section 4.
1467: As a matter of fact, we have obtained a more general result
1468: regarding the convexity of functions $\mathcal{F}(x_1,\ldots, x_n)$
1469: of the form $\mathcal{F}(x_1,\ldots, x_n)=\prod_{i=1}^n f(x_i)$,
1470: where $f(x)$ is a positive function that has continuous second
1471: derivative. This result may have applications in other areas of
1472: mathematics.
1473:
1474:
1475: \vspace{1cm} \noindent \textbf{Acknowledgement}\;
1476: The authors would like to thank Malaysian Academy of Sciences,
1477: Ministry of Science, Technology and Innovation for funding this
1478: project under the Scientific Advancement Fund Allocation (SAGA) Ref.
1479: No P96c.
1480:
1481: \begin{thebibliography}{10}
1482: \bibitem{Ep1} P. Epstein, \emph{Zur Theorie allgemeiner Zetafunktionen}, Math. Ann. \textbf{56}, 615--644 (1903).
1483:
1484: \bibitem{Ep2}P. Epstein, \emph{Zur Theorie allgemeiner Zetafunktionen II}, Math. Ann. \textbf{65},
1485: 205--216 (1907).
1486:
1487: \bibitem{e4}
1488: C.~L. Siegel, \emph{Contributions to the theory of the {D}irichlet
1489: {$L$}-series and the {E}pstein zeta-functions}, Ann. of Math. (2) \textbf{44}, 143--172 (1943).
1490:
1491:
1492:
1493: \bibitem{e1}
1494: M.~Koecher, \emph{\"Uber Dirichlet-Reihen mit Funktionalgleichung},
1495: J. Reine Angew. Math. \textbf{192}, 1–-23 (1953).
1496:
1497: \bibitem{e5}
1498: C.~L. Siegel, \emph{A generalization of the {E}pstein zeta
1499: function}, J. Indian
1500: Math. Soc. (N.S.) \textbf{20}, 1--10 (1956).
1501:
1502: \bibitem{e2}
1503: A.~Selberg, \emph{A new type of zeta function connected with
1504: quadratic forms}, Report of the Inst. in Theory of Numbers, U. of
1505: Colorado, Boulder, Colo., 207--210 (1959).
1506:
1507:
1508:
1509: \bibitem{e6}
1510: A. Terras, \emph{A generalization of {E}pstein's zeta function},
1511: Nagoya
1512: Math. J. \textbf{42}, 173--188 (1971).
1513:
1514: \bibitem{e3}
1515: S.~Chowla, \emph{On an unsuspected real zero of {E}pstein's zeta
1516: function},
1517: Proc. Nat. Inst. Sci. India \textbf{13}, no.~4, 1 (1947).
1518:
1519: \bibitem{E1}%5
1520: E.~Elizalde, S.~D. Odintsov, A.~Romeo, A.~A. Bytsenko, and
1521: S.~Zerbini,
1522: \emph{Zeta regularization techniques with applications}, (World Scientific
1523: Publishing Co. Inc., River Edge, NJ, 1994).
1524:
1525: \bibitem{ET}
1526: Emilio Elizalde, \emph{Ten physical applications of spectral zeta
1527: functions},
1528: (Springer-Verlag,
1529: Berlin, 1995).
1530:
1531:
1532: \bibitem{K}%7
1533: K.~Kirsten, \emph{Spectral functions in mathematics and physics},
1534: (Chapman \&
1535: Hall/ CRC, Boca Raton, FL, 2002).
1536:
1537:
1538: \bibitem{AW}%3
1539: Jan Ambj{\o}rn and S.~Wolfram, \emph{Properties of the vacuum. {I}.
1540: {M}echanical and thermodynamic}, Ann. Physics \textbf{147}, 1--32 (1983).
1541:
1542: \bibitem{c41} D.~J. Toms, \emph{Symmetry breaking and mass generation by space-time
1543: topology}, Phys. Rev. D \textbf{21}, 2805--2817 (1980).
1544:
1545: \bibitem{c42} G. Denardo and E. Spallucci, \emph{Dynamical mass generation in $S^1\times R^3$},
1546: Nucl. Phys. B \textbf{169}, 514--526 (1980).
1547:
1548:
1549: \bibitem{c43} A. Actor, \emph{Topological generation of gauge field mass by toroidal
1550: spacetime}, Class. Quantum Grav. \textbf{7}, 663--683 (1980).
1551:
1552: \bibitem{c44} K. Kirsten, \emph{Topological gauge field mass generation by toroidal
1553: spacetime}, J. Phys. A: Math. Gen. \textbf{26}, 2421--2435 (1993).
1554:
1555: \bibitem{EK1} E. Elizalde and K. Kirsten, \emph{Topological symmetry breaking in
1556: self-interacting theories on toroidal space-time}, J. Math. Phys.
1557: \textbf{35}, 1260--1273 (1994).
1558:
1559: \bibitem{e10} S.~C. Lim and L.~P. Teo, \emph{Topological symmetry
1560: breaking of self--interacting fractional Klein--Gordon field on
1561: toroidal spacetime}, submitted, can be downloaded from
1562: http://pesona.mmu.edu.my/~lpteo.
1563:
1564:
1565:
1566: \bibitem{f9}
1567: R.~A. Rankin, \emph{A minimum problem for the {E}pstein
1568: zeta-function}, Proc.
1569: Glasgow Math. Assoc. \textbf{1}, 149--158 (1953).
1570:
1571: \bibitem{f23}
1572: J.~W.~S. Cassels, \emph{On a problem of Rankin about Epstein zeta
1573: function}, Proc. Glasg. Math. Assoc. \textbf{4}, 73--80 (1959);
1574: \textbf{6}, 116 (1963).
1575:
1576: \bibitem{f3}
1577: V.~Ennola, \emph{On a problem about the {E}pstein zeta-function},
1578: Proc.
1579: Cambridge Philos. Soc. \textbf{60}, 855--875 (1964).
1580:
1581: \bibitem{f24}
1582: P.~H. Diananda, \emph{Notes on two lemmas concerning the Epstein
1583: zeta function}, Proc. Glasg. Math. Assoc. \textbf{6}, 202--204
1584: (1964).
1585:
1586: \bibitem{f10}
1587: S.~S. Ry{\v{s}}kov, \emph{On the question of the final {$\zeta
1588: $}-optimality of
1589: lattices that yield the densest packing of {$n$}-dimensional balls}, Sibirsk.
1590: Mat. \v Z. \textbf{14}, 1065--1075 (1973).
1591:
1592: \bibitem{f19}
1593: S.~{\v{S}}. {\v{S}}u{\v{s}}baev, \emph{An estimate of the origin of
1594: the ray of
1595: extremality of the principal perfect form in four variables}, Dokl. Akad.
1596: Nauk UzSSR, no.~9, 15--16 (1976).
1597:
1598: \bibitem{f20}
1599: S.~{\v{S}}. {\v{S}}u{\v{s}}baev, \emph{An estimate of the origin of
1600: the ray of extremality of the
1601: principal perfect form in five variables}, Voprosy Vy\v cisl. i Prikl. Mat.
1602: (Tashkent), no. 46, 3--8 (1977).
1603:
1604: \bibitem{f21}
1605: S.~{\v{S}}. {\v{S}}u{\v{s}}baev, \emph{The {R}ankin-{S}obolev
1606: problem on extrema of {E}pstein's zeta
1607: function. {T}he origin of the ray of extremality of the principal perfect
1608: form}, Izv. Akad. Nauk UzSSR Ser. Fiz.-Mat. Nauk, no.~1, 33--39 (1978).
1609:
1610:
1611:
1612:
1613: \bibitem{f2}
1614: K.~M. {\`E}ndibaev, \emph{The extremal of {E}pstein's zeta function
1615: from four
1616: variables}, Voprosy Vychisl. i Prikl. Mat. (Tashkent), no.~51,
1617: 178--190 (1978).
1618:
1619: \bibitem{f22}
1620: A.~Terras, \emph{The minima of quadratic forms and the behavior of
1621: {E}pstein
1622: and {D}edekind zeta functions}, J. Number Theory \textbf{12}, no.~2,
1623: 258--272 (1980).
1624:
1625: \bibitem{f1}
1626: K.~M. {\`E}ndibaev, \emph{An extremal problem for the
1627: four-dimensional {E}pstein {$\zeta
1628: $}-function}, Dokl. Akad. Nauk UzSSR, no.~4, 18--21 (1978).
1629:
1630:
1631:
1632: \bibitem{f4}
1633: K.~L. Fields, \emph{Locally minimal {E}pstein zeta functions},
1634: Mathematika
1635: \textbf{27}, no.~1, 17--24 (1980).
1636:
1637: \bibitem{f6}
1638: Z.~D. Lomakina and S.~S. Ryshkov, \emph{One-dimensional and
1639: two-dimensional
1640: faces of the polyhedron {$\mu(5)$}}, Zap. Nauchn. Sem. Leningrad. Otdel. Mat.
1641: Inst. Steklov. (LOMI) \textbf{151}, no.~9,
1642: 95--103 (1986).
1643:
1644: \bibitem{f12}
1645: S.~Sh. Shushbaev, \emph{On the {R}ankin-{S}obolev problem on extrema
1646: of
1647: {E}pstein zeta-functions and on a transcendental equation}, Voprosy Vychisl.
1648: i Prikl. Mat. (Tashkent), no.~82, 108--113 (1987).
1649:
1650: \bibitem{f13}
1651: S.~Sh. Shushbaev, \emph{Local minima of the {E}pstein zeta
1652: function}, Mat. Zametki
1653: \textbf{45}, no.~1, 123--131 (1989).
1654:
1655: \bibitem{f15}
1656: S.~Sh. Shushbaev, \emph{On the {R}ankin-{S}obolev problem on extrema
1657: of the {E}pstein
1658: zeta-function. {A}n estimate for the origin of a ray of extremality of the
1659: second perfect {V}orono\u\i\ form}, Sibirsk. Mat. Zh. \textbf{31},
1660: no.~5, 157--163 (1990).
1661:
1662: \bibitem{f14}
1663: S.~Sh. Shushbaev, \emph{Two local minima of the six-dimensional
1664: {E}pstein
1665: zeta-function}, Izv. Vyssh. Uchebn. Zaved. Mat., no.~5, 78--83 (1990).
1666:
1667: \bibitem{f16}
1668: S.~Sh. Shushbaev, \emph{Rays of extremality: eventually extremal
1669: forms in seven
1670: variables}, Sibirsk. Mat. Zh. \textbf{35}, no.~6, 1397--1400 (1994).
1671:
1672: \bibitem{f7}
1673: E.~V. Orlovskaya, \emph{The minimum for a theta function of three
1674: variables and
1675: the solution of the {R}ankin-{S}obolev problem in three-dimensional space},
1676: Vestnik S.-Peterburg. Univ. Mat. Mekh. Astronom. \textbf{149}, no. 3, 23--27
1677: (1994).
1678:
1679: \bibitem{f8}
1680: E.~V. Orlovskaya, \emph{The minimum for the theta-function of three
1681: variables and the
1682: solution of the {R}ankin-{S}obolev problem in a three-dimensional space},
1683: Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI)
1684: \textbf{211}, no. 3, 150--157 (1994).
1685:
1686: \bibitem{f5}
1687: W.~Jenkner, \emph{Sur les fonctions z\^eta attach\'ees aux classes
1688: de rayon},
1689: J. Th\'eor. Nombres Bordeaux \textbf{7}, no.~1, 1--14 (1995).
1690:
1691:
1692:
1693: \bibitem{f17}
1694: S.~Sh. Shushbaev and G.~A. Kalybaeva, \emph{On a perfect form that
1695: is not a
1696: final-extremal form}, Uzbek. Mat. Zh., no.~2, 80--86 (2002).
1697:
1698: \bibitem{f18}
1699: J.~Steuding, \emph{On the zero-distribution of {E}pstein
1700: zeta-functions}, Math.
1701: Ann. \textbf{333}, no.~3, 689--697 (2005).
1702:
1703:
1704:
1705:
1706:
1707:
1708:
1709: \bibitem{f11}
1710: P.~Sarnak and A.~Str{\"o}mbergsson, \emph{Minima of {E}pstein's zeta
1711: function
1712: and heights of flat tori}, Invent. Math. \textbf{165}, no.~1,
1713: 115--151 (2006).
1714:
1715: \bibitem{d19}
1716: S.~Chowla and A.~Selberg, \emph{On {E}pstein's zeta function. {I}},
1717: Proc. Nat.
1718: Acad. Sci. U. S. A. \textbf{35}, 371--374 (1949).
1719:
1720: \bibitem{d20}
1721: A. Selberg and S.~Chowla, \emph{On {E}pstein's zeta-function}, J.
1722: Reine
1723: Angew. Math. \textbf{227}, 86--110 (1967).
1724:
1725: \bibitem{g1}
1726: I.~S. Gradshteyn and I.~M. Ryzhik, \emph{Table of integrals, series,
1727: and
1728: products}, (Academic Press Inc., San Diego, CA, 2000).
1729:
1730: \bibitem{g2}
1731: A. Fujii, \emph{A note on the values of the Epstein zeta functions
1732: at the critical points}, Comment. Math. Univ. St. Pauli
1733: \textbf{49}, no.~2, 195--225 (2000).
1734:
1735:
1736: \end{thebibliography}
1737:
1738: \end{document}
1739: