1: \documentclass[11pt]{amsart}
2: \setlength{\textwidth}{6.5in}
3: \setlength{\textheight}{9in}
4: \setlength{\evensidemargin}{0in}
5: \setlength{\oddsidemargin}{0in}
6: \setlength{\topmargin}{-.25in}
7: \usepackage{epsfig}
8: \usepackage{amsthm}
9: \usepackage{amssymb}
10: \usepackage{amsmath}
11: \usepackage{epic}
12: \usepackage{eepic}
13:
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: % Work mode:
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17: %\newcommand {\comment}[1] {{\ \scriptsize{#1}\ }}
18: %\newcommand{\mute}[2] {{\scriptsize \ #1\ }}
19: %\newcommand{\printname}[1]
20: %{\smash{\makebox[0pt]{\hspace{-1.0in}\raisebox{8pt}{\tiny #1}}}}
21: %
22: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
23: % Clean mode:
24: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
25: \newcommand {\comment}[1] {}
26: \newcommand{\mute}[2] {}
27: \newcommand {\printname}[1] {}
28: %
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: \newcommand{\labell}[1] {\label{#1}\printname{#1}}
31:
32:
33: \DeclareMathOperator {\Orth}{Orth}
34:
35: \newtheorem {Theorem} {Theorem}
36: %\numberwithin{Theorem}{section}
37: \newtheorem {Lemma}[equation] {Lemma}
38: \newtheorem {Claim}[equation] {Claim}
39: \newtheorem {Proposition}[equation]{Proposition}
40: \theoremstyle{definition}
41: \newtheorem{Definition}[equation]{Definition}
42: \newtheorem {Corollary}[equation]{Corollary}
43: \theoremstyle{remark}
44: \newtheorem{Remark}[equation]{Remark}
45: \newtheorem{Example}[equation]{Example}
46: \newtheorem {Exercise}[equation]{Exercise}
47: %\newtheorem {Corollary}[equation]{Corollary}
48: \newtheorem {Question}[equation]{Question}
49:
50:
51: %\def \valpha {{\check{\alpha}}}
52: \def \valpha {\alpha}
53: %\def \vV {{\check{V}}}
54: \def \vV {{V}}
55: %\def \vLambda {{\check{\Lambda}}}
56: \def \vGamma {{\check{\Gamma}}}
57:
58: \def \calF {{\mathcal F}}
59: \def \calP {{\mathcal P}}
60: \def \calD {{\mathcal D}}
61:
62: \def \bfC {{\mathbf C}}
63: \def \bfS {{\mathbf S}}
64: \def \bfO {{\mathbf O}}
65: \def \bfM {{\mathbf M}}
66:
67: \def \inv {^{-1}}
68: \def \ol {\overline}
69: \def \ssminus {{\smallsetminus}}
70: \newcommand {\brac}[1] {\left< #1 \right> }
71: \def \l< {\left< }
72: \def \r> {\right> }
73: \def \del {\partial}
74: \newcommand {\deldel}[1] {\frac{\partial}{\partial #1}}
75:
76: \def \pos {{>0}}
77: \def \nonneg {{\geq 0}}
78:
79: \def \shift {{\text{shift}}}
80: \def \codim {{\operatorname{codim}}}
81: \def \red {{\operatorname{red}}}
82: \def \cut {{\operatorname{cut}}}
83: \def \Aut {{\operatorname{Aut}}}
84: \def \deg {{\operatorname{deg}}}
85: \def \id {{\operatorname{id}}}
86: \def \Hom {{\operatorname{Hom}}}
87: \def \span {{\operatorname{span}}}
88: \def \SL {{\operatorname{SL}}}
89: \def \Td {{\operatorname{Td}}}
90: \def \interior {{\operatorname{interior}}}
91: \def \relint {{\operatorname{rel-int}}}
92: \def \Cinf {C^\infty}
93: \def \half {{\frac{1}{2}}}
94:
95: \def \t {{\mathfrak t}}
96: \def \k {{\mathfrak k}}
97: \def \Z {{\mathbb Z}}
98: \def \R {{\mathbb R}}
99: \def \C {{\mathbb C}}
100: \def \L{{\bf L}}
101: \def \Ltk{{\bf L}^{2k}}
102: \def \interior {{\operatorname{interior}}}
103: \def \relint {{\operatorname{rel-int}}}
104: \def \Cinf {C^\infty}
105: \def \half {{\frac{1}{2}}}
106: \def \shift {{\text{shift}}}
107: \def \st {{\text{st}}}
108: \def \codim {{\operatorname{codim}}}
109: \def \red {{\operatorname{red}}}
110: \def \cut {{\operatorname{cut}}}
111:
112:
113:
114:
115: \def \eps {{\epsilon}}
116:
117:
118: \begin{document}
119:
120: \title[Euler Maclaurin for a simple integral polytope]
121: {Euler Maclaurin with remainder for a simple integral polytope}
122:
123: \author[Y.\ Karshon]{Yael Karshon}
124: \address{Institute of Mathematics, The Hebrew University of Jerusalem,
125: Israel, and: Department of Mathematics, The University of Toronto,
126: Toronto, Ontario M5S 3G3, Canada}
127: \email{karshon@math.toronto.edu}
128:
129: \author[S.\ Sternberg]{Shlomo Sternberg}
130: \address{Department of Mathematics, Harvard University,
131: Cambridge, MA 02138, USA}
132: \email{shlomo@math.harvard.edu}
133:
134: \author[J.\ Weitsman]{Jonathan Weitsman}
135: \address{Department of Mathematics, University of California,
136: Santa Cruz, CA 95064, USA}
137: \email{weitsman@math.UCSC.EDU}
138: \thanks{2000 \emph{Mathematics Subject Classification.}
139: Primary 65B15, 52B20.}
140:
141: \thanks{This work was partially supported by
142: United States -- Israel Binational Science Foundation
143: grant number 2000352 (to Y.K. and J.W.), by the Connaught Fund (to Y.K.),
144: and by National Science Foundation Grant DMS 99/71914 (to J.W.).}
145:
146: \begin{abstract}
147: We give an Euler Maclaurin formula with remainder for the sum of the
148: values of a smooth function on the integral points in a simple integral
149: polytope. We prove this formula by elementary methods.
150: \end{abstract}
151: \maketitle
152: %\tableofcontents
153:
154:
155: % -------------------------------------------------------------------------
156: \section{Introduction}
157: % -------------------------------------------------------------------------
158: \labell{sec:intro}
159:
160: The Euler Maclaurin formula computes the sum of the values
161: of a function $f$ over the integer points in an interval
162: in terms of the integral of $f$ over variations of that interval.
163: A version of this classical formula is this:
164:
165: For any function $f(x)$ on the real line and any integers $a<b$,
166: we will consider the weighted sum
167: \begin{equation} \labell{wted sum}
168: \sum_{[a,b]}{'} f := \half f(a) + f(a+1) + \ldots + f(b-1) + \half f(b) .
169: \end{equation}
170: If $f$ is ``nice enough", for instance, a polynomial, then
171: \begin{equation} \labell{EM exact weighted}
172: \sum_{[a,b]}{'} f =
173: \left. \L(\deldel{h_1}) \L(\deldel{h_2}) \int_{a-h_1}^{b+h_2} f(x) dx
174: \right|_{h_1 = h_2 = 0},
175: \end{equation}
176: where
177: \begin{equation} \labell{LS}
178: \L(S) = \frac{S/2}{\tanh(S/2)}
179: = 1 + \sum_{k=1}^\infty \frac{1}{(2k)!} b_{2k} S^{2k} .
180: \end{equation}
181: Because $\int_{a-h_1}^{b+h_2} f(x) dx$ is a polynomial in $h_1$ and $h_2$
182: if $f$ is a polynomial in $x$,
183: applying the infinite order differential operator $\L(\deldel{h_i})$
184: then yields a finite sum, so
185: the right hand side of \eqref{EM exact weighted} is well defined
186: when $f$ is a polynomial.
187:
188: A polytope in $\R^n$ is called \emph{integral}, or a \emph{lattice polytope},
189: if its vertices
190: are in the lattice $\Z^n$;
191: it is called \emph{simple} if exactly $n$ edges emanate from each vertex;
192: it is called \emph{regular} if, additionally,
193: the edges emanating from each vertex lie along lines which are generated
194: by a $\Z$-basis of the lattice $\Z^n$.
195:
196: Khovanskii and Pukhlikov \cite{KP1,KP2}, following
197: Khovanskii \cite{Kh1,Kh2},
198: generalized the classical Euler Maclaurin formula to give
199: a formula for the sums of the values of polynomial or exponential
200: functions on the lattice points in higher dimensional
201: convex polytopes $\Delta$ which are integral and regular.
202: This formula was generalized to \emph{simple} integral polytopes
203: by Cappell and Shaneson \cite{CS:bulletin,CS:EM,CS:private,S},
204: and subsequently by Guillemin \cite{Gu} and by Brion-Vergne \cite{BV}.
205: All of these generalizations involve ``corrections" to the
206: Khovanskii-Pukhlikov formula when the simple polytope is not regular.
207: When applied to the constant function $f \equiv 1$, these Euler Maclaurin
208: formulas compute the number
209: of lattice points in $\Delta$ in terms of the volumes
210: of ``dilations" of $\Delta.$
211: A small sample of the literature on the problem of counting lattice
212: points in convex polytopes is given in
213: \cite{pick,Md,V,KK,Mo,Po,DR,BDR,Ha};
214: see the survey \cite{BP} and references therein.
215:
216: These formulas are closely related to the Riemann Roch formula
217: from algebraic geometry via the correspondence between
218: polytopes and toric varieties.
219: Under this correspondence, regular polytopes correspond
220: to smooth toric varieties, and the Khovanskii-Pukhlikov
221: formula was motivated in this way, although it was proved combinatorially.
222: Cappell and Shaneson derived their formula
223: from their theory of characteristic classes of singular algebraic
224: varieties and had the key idea of using the operator $\L$ as we do here.
225: Guillemin obtained his formula from the equivariant
226: Kawasaki-Riemann-Roch formula and methods coming from
227: symplectic geometry and the theory of geometric quantization.
228: Brion and Vergne employed a method
229: that is closer to that in the original proof of Khovanskii and
230: Pukhlikov, using Fourier analysis.
231:
232: To illustrate the relation to toric varieties, let us sketch
233: the symplectic-geometric proof for the case of a regular polytope,
234: following Guillemin \cite{Gu:book}. This approach will not be
235: used elsewhere in this paper.
236: A regular integral polytope $\Delta \subset \R^n$
237: determines a smooth K\"ahler toric variety $(M,\omega)$,
238: and geometric quantization gives rise to a virtual representation $Q(M)$
239: of the torus $T^n$.
240: The dimension $\dim Q(M)$ of this quantization
241: is equal to the number of lattice points in $\Delta$.
242: (This result (see \cite[Corollary 2.23]{Oda})
243: is an expression of the ``quantization commutes with reduction"
244: principle in symplectic geometry \cite{GS}.
245: According to this principle, $\dim Q(M)^c = \dim Q(M_c)$
246: for each lattice point $c \in \Z^n \subset {\rm Lie}(T^n)^*$,
247: where $Q(M)^c$ is the subspace of $Q(M)$ on which
248: $T^n$ acts through the character given by $c$,
249: and where $M_c$ is the reduced space of $M$ at $c$. Because $M$
250: is a toric variety, $M_c$ is a point if $c \in \Delta$
251: and is empty otherwise.)
252: On the other hand, by the Hirzebruch-Atiyah-Singer generalization
253: of the classical Riemann-Roch formula, we have
254: $\dim Q(M) = \int_M \exp(c_1(L)) \Td(TM)$,
255: where $c_1(L) = [\omega]$ is the Chern class of the
256: pre-quantization line bundle
257: and $\Td(TM)$ is the Todd class of the tangent bundle.
258: Expressing $M$ as a reduction of a linear torus action
259: on $\C^d$ (where $d$ is the number of facets of $\Delta$),
260: the tangent bundle stably splits into line bundles
261: $L_1,\ldots,L_d$,
262: and the above integral is obtained by applying
263: the Khovanskii-Pukhlikov differential operator $\prod \Td(\deldel{h_i})$
264: to the integral $\int_M \exp(\omega + \sum h_i c_1(L_i))$.
265: The Duistermaat-Heckman theorem on the variation
266: of reduced symplectic structures implies that this integral
267: is equal to the volume of the polytope $\Delta(h)$
268: that is obtained from $\Delta$ by shifting the $i$th facet
269: by a distance $h_i$, for $i=1,\ldots,d$.
270: Hence, the number of lattice points in $\Delta$
271: is obtained by applying the Khovanskii-Pukhlikov operator
272: to the volume of $\Delta(h)$.
273:
274: The Euler Maclaurin formulas due to Khovanskii-Pukhlikov,
275: Cappell-Shaneson, Guillemin, and Brion-Vergne are all
276: exact formulas, valid for sums of exponential or polynomial functions.
277: Cappell and Shaneson \cite{CS:private} have also investigated the
278: problem of deriving an Euler Maclaurin formula with remainder.
279: In a previous paper \cite{PNAS}, we stated and proved an Euler Maclaurin
280: formula with remainder for the sum of the values of an arbitrary smooth
281: function on the lattice points in a {\em regular}
282: polytope, and adumbrated a generalization to the case of
283: simple integral polytopes. The purpose of this paper is to state
284: and prove an Euler Maclaurin formula with remainder
285: for simple lattice polytopes (Theorem \ref{main}). The key
286: ingredients in the proof of this theorem, as in \cite{PNAS},
287: are a variant of the Euler Maclaurin formula in one dimension,
288: given in Proposition \ref{twistedemray}, which
289: by iteration gives a formula for \emph{orthants},
290: along with a combinatorial result, given
291: in Proposition \ref{weightedlawrence}, which shows how the sum
292: of the values of
293: a function over the lattice points in a polytope can be decomposed into sums
294: over such orthants.
295:
296: Our Euler Maclaurin formula with remainder is stated in Theorem \ref{main}
297: for functions of compact support.
298: In Section \ref{sec:estimates} we show how to extend it to symbols
299: (in the sense of Hormander, see e.g.\ \cite{horm}).
300: This is the content of Theorem \ref{syms}.
301: As a corollary, we deduce an exact Euler Maclaurin formula for polynomials.
302:
303: The early references to the Euler Maclaurin formula are Euler
304: \cite{Eu} and Maclaurin \cite{Ma}.
305: \label{Poisson-page}
306: Apparently, Poisson
307: \cite{Poi} was the first to give a remainder formula.
308: See also \cite{H}.
309:
310: % -------------------------------------------------------------------------
311: \section{Weighted sums in one dimension}
312: % -------------------------------------------------------------------------
313: \labell{sec:EM dim=1}
314:
315: \subsection*{Exact Euler Maclaurin}
316:
317: Here is a brief proof of the exact Euler Maclaurin formula
318: \eqref{EM exact weighted}; cf.~\cite{BV}.
319: First, we prove this formula
320: when $f(x)$ is an exponential function: $f(x) = e^{\lambda x}$
321: with $|\lambda| < 2 \pi.$ The formula then becomes
322: \begin{equation} \labell{EM for exponential}
323: \sum_{[a,b]}{'} e^{\lambda x}
324: = \left. \L(\deldel{h_1}) \L(\deldel{h_2}) \int_{a-h_1}^{b+h_2}
325: e^{\lambda x} dx \right|_{h_1 = h_2 = 0} .
326: \end{equation}
327: An explicit computation,
328: which uses the facts that the constant term
329: in the formal power series $\L(S)$ is one and that $\L(-S) = \L(S)$,
330: shows that
331: \begin{equation} \labell{equation for LN}
332: \left. \Ltk(\deldel{h_1}) \Ltk(\deldel{h_2}) \int_{a-h_1}^{b+h_2}
333: e^{\lambda x} dx \right|_{h_1 = h_2 = 0}
334: = \Ltk(\lambda) \int_a^b e^{\lambda x} dx
335: \end{equation}
336: where $\Ltk(S)$ is the truncation of the power series $\L(S)$
337: at the even integer $2k$.
338: The radius of convergence of the power series $\L(\lambda)$ is $2\pi$
339: because the zeros of $\tanh(\lambda/2)$ that are nearest
340: to the origin are at $\pm 2 \pi i$. Hence, for $|\lambda| < 2\pi$,
341: the expression in \eqref{equation for LN} converges as $k\to \infty$ to
342: \begin{equation} \labell{LL1}
343: \L(\lambda) \int_a^b e^{\lambda x} dx
344: = \frac{ \lambda/2 }{ \tanh(\lambda/2) }
345: \left( \frac{ e^{\lambda b} }{\lambda}
346: - \frac{ e^{\lambda a} }{\lambda} \right)
347: = \half \left( e^{\lambda/2} + e^{-\lambda/2} \right)
348: \frac{ e^{\lambda b} - e^{ \lambda a } }{ e^{\lambda/2} - e^{-\lambda/2} },
349: \end{equation}
350: which is equal to the left hand side of \eqref{EM for exponential}
351: by the formula for a geometric sum.
352:
353: Moreover, this convergence is uniform on any closed sub-disk
354: of $|\lambda| < 2\pi$, because $\L(\lambda)$ is a power series
355: and $\int_a^b e^{\lambda x} dx$ is bounded away from $0$ and from $\infty$.
356: Recall that differentiation commutes with uniform limits of holomorphic
357: functions (as a consequence of the Cauchy formula).
358: It follows that the derivative $\deldel{\lambda}$
359: commutes with the infinite order differential operators $\L(\deldel{h_i})$
360: on the right hand side of \eqref{EM for exponential}.
361: Comparing the Taylor coefficients of $\lambda^n$ on the left and right
362: hand sides of \eqref{EM for exponential}, we get a similar formula
363: for $f(x) = x^n$, and hence for all polynomials.
364:
365: \begin{Remark} \labell{exp to pol}
366: In higher dimensions, the problem of obtaining a formula for polynomials
367: from a formula for exponentials is addressed in \cite{KP2}.
368: \end{Remark}
369:
370: \subsection*{Weighted sums and ordinary sums}
371:
372: The Todd function is defined by
373: $$ \Td(S) = \frac{S}{1 - e^{-S}}.$$
374: The Bernoulli numbers are the coefficients $b_k$ in
375: $$ \Td(-S) = \frac{S}{e^S - 1}
376: = 1 + \sum_{k \geq 1} \frac{1}{k!} b_k S^k . $$
377:
378: (We are following the conventions in Bourbaki \cite{Bo}.)
379: Since
380: \begin{equation*}
381: \begin{aligned}
382: \L(S) &= \left( S/2 \right) \frac{e^{S/2} + e^{-S/2}}{e^{S/2} - e^{-S/2}}
383: = \half \left( \frac{S}{1 - e^{-S}} + \frac{S}{e^S - 1} \right)\\
384: &= \half \left( \Td(S) + \Td(-S) \right) ,
385: \end{aligned}
386: \end{equation*}
387: the coefficients $b_{2k}$ in the power series expansion \eqref{LS}
388: are the even Bernoulli numbers.
389: Since
390: $$ \Td(S) - \Td(-S) = \frac{S}{1 - e^{-S}} - \frac{S}{e^S - 1} = S,$$
391: we have $b_{2k+1} = 0$ for all $k \geq 1$, so the only difference
392: between $\Td(S)$ and $\L(S)$
393: is the absence of the linear term in $\L(S)$.
394: Replacing $\L(\cdot)$ by $\Td(\cdot)$ on the right hand side
395: of the Euler Maclaurin formula \eqref{EM exact weighted}
396: results in a formula for the ordinary sum of the values of $f$
397: over the integers in $[a,b]$.
398: However, we work with the formula for the weighted sum \eqref{wted sum},
399: because this formula avoids ``boundary effects" which occur
400: in the formulas for the ordinary sum.
401:
402: \subsection*{Euler Maclaurin with remainder}
403:
404: As before,
405: we will denote the truncation of the power series $\L(S)$
406: at the even integer $2k$ by $\Ltk(S)$. Then
407: \begin{equation} \labell{Ltk}
408: \Ltk(\deldel{h})
409: = 1 + \frac{1}{2!} b_2 \frac{\del^2}{\del h^2}
410: + \ldots
411: + \frac{1}{(2k)!} b_{2k} \frac{\del^{2k}}{\del h^{2k}}
412: \end{equation}
413: is a differential operator with constant coefficients involving only even
414: order derivatives. In particular,
415: if $g(h)$ is a function with $2k$ continuous derivatives, then
416: \begin{equation} \labell{plusminus}
417: \left.\Ltk(\deldel{h})(g(h))\right|_{h=0}
418: = \left.\Ltk(\deldel{h})(g(-h))\right|_{h=0}.
419: \end{equation}
420:
421: The classical Euler Maclaurin summation formula with remainder
422: can be formulated in the following way.
423: \begin{Proposition}[Euler Maclaurin with remainder for intervals]
424: \labell{EMintervalgrm}
425: Let $f(x)$ be a function with $m \geq 1$ continuous derivatives
426: and let $k = \lfloor m/2 \rfloor$.
427: Then
428: \begin{multline} \labell{emr1d}
429: {\sum_{[a,b]} }' f = \left. \Ltk(\deldel{h_1}) \Ltk(\deldel{h_2})
430: \int_{a-h_1}^{b+h_2} f(x) dx
431: \right|_{h_1 = h_2 = 0}
432: + (-1)^{m-1} \int_a^b P_m(x) f^{(m)}(x) dx
433: \end{multline}
434: with
435: \begin{equation} \labell{P and B}
436: P_m(x) = \frac{B_m(\{ x \})}{m!},
437: \end{equation}
438: where $B_m(x)$ is the $m$th Bernoulli polynomial (see below)
439: and where $\{ x \} = x - \lfloor x \rfloor$
440: is the fractional part of $x$.
441: Moreover, the function $P_m(x)$ is given by
442: \begin{equation} \labell{P2k}
443: P_{2k}(x) = (-1)^{k-1} \sum_{n=1}^\infty \frac{2\cos2\pi nx}{(2\pi n)^{2k}}
444: \end{equation}
445: if $m=2k$ is even and by
446: \begin{equation} \labell{P2kp1}
447: P_{2k+1}(x) =
448: (-1)^{k-1} \sum_{n=1}^\infty \frac{2\sin2\pi n x}{(2\pi n)^{2k+1}}
449: \end{equation}
450: if $m=2k+1$ is odd.
451: \end{Proposition}
452:
453: Up to minor changes in notation, this result is formula \textbf{298}
454: in \cite{Kn}.
455:
456: Note that if $f$ is a polynomial
457: then \eqref{emr1d} becomes an exact formula when $m$ is greater
458: than the degree of $f$.
459:
460: \medskip
461:
462: Let us recall the proof of Proposition \ref{EMintervalgrm}.
463: Consider the difference between integration and summation
464: as the difference between two distributions:
465: $$ P_0 (x) := 1 - \sum_{k \in \Z} \delta(x-k). $$
466: Integrating this, with the choice of constant of integration
467: such that the integral from $0$ to $1$
468: of the result vanishes, gives a distribution $P_1$ which
469: is given by the function
470: \begin{equation} \labell{P1x}
471: P_1(x) = x - \lfloor x \rfloor - \half
472: \end{equation}
473:
474: \noindent at non-integral points.
475: This is the famous zigzag function studied by Gibbs \cite{Gibbs}.
476: \label{Gibbs-page}
477: Notice that $P_1$ is 1-periodic and odd.
478:
479: If $f$ is a continuously differentiable function on $[0,1]$, we get
480: $$
481: \int_0^1 P_1(x) f'(x) dx
482: = \left. \phantom{\int} P_1(x) f(x) \right|_0^1 - \int_0^1 f(x) dx
483: = \half f(0) + \half f(1) - \int_0^1 f(x) dx,
484: $$
485: so
486: $$\half f(0) + \half f(1) = \int_0^1 f(x) dx + \int_0^1 P_1(x) f'(x) dx.$$
487: Summing the corresponding expressions over the intervals
488: $[a,a+1]$, $[a+1,a+2]$, $\ldots$, $[b-1,b]$,
489: where $a <b$ are integers, gives
490: \begin{equation} \labell{sum over intervals}
491: \sum_{[a,b]}{'} f = \int_a^b f(x) dx + \int_a^b P_1(x) f'(x) dx.
492: \end{equation}
493: This is the first of a series of expressions relating the sum to an integral.
494:
495: We now take successive anti-derivatives:
496: for $m \geq 2$,
497: define $P_m(x)$ inductively by the conditions that
498: $\frac{d}{dx} P_m(x) = P_{m-1}(x)$ and $\int_0^1 P_m(x) dx = 0$.
499: Then $P_m(x)$ is 1-periodic and satisfies $P_m(-x) = (-1)^m P_m(x)$.
500: Also, $P_m(x)$ is continuously differentiable $m-2$ times;
501: in particular, it is continuous if $m \geq 2$.
502: Starting from \eqref{sum over intervals}, we integrate by parts:
503: \begin{eqnarray*}
504: \sum_{[a,b]}{'} f
505: & = & \int_a^b f(x) dx + \left. P_2 f' \right|_a^b - \int_a^b P_2(x) f''(x) dx \\
506: & = & \int_a^b f(x) dx + \left. P_2 f' \right|_a^b
507: - \left. P_3 f'' \right|_a^b + \int_a^b P_3(x) f^{(3)}(x) dx\\
508: & = & \int_a^b f(x) dx + \left. P_2 f' \right|_a^b - \left. P_3 f'' \right|_a^b
509: + \left. P_4 f^{(3)} \right|_a^b - \int_a^b P_4(x) f^{(4)}(x) dx \\
510: & \vdots &
511: \end{eqnarray*}
512: Noting that $P_n(a) = P_n(b) = P_n(0)$ and $P_{2n+1}(0) = 0$
513: (because $P_{2n+1}$ is odd), we get, setting $k = \lfloor m/2 \rfloor$,
514: \begin{multline} \labell{dog}
515: \sum_{[a,b]}{'} f =
516: \int_a^b f(x) dx + P_2(0) f' |_a^b + P_4(0) f^{(3)}|_a^b + \ldots
517: + P_{2k}(0) f^{(2k-1)}|_a^b \\
518: + (-1)^{m-1} \int_a^b P_m(x) f^{(m)}(x) dx.$$
519: \end{multline}
520:
521: Consider the polynomial
522: \begin{equation} \labell{L2k[]}
523: L^{[2k]}(S) := 1 + P_2(0) S^2 + P_4(0) S^4 + \ldots + P_{2k}(0) S^{2k}.
524: \end{equation}
525: From \eqref{dog} we get the remainder formula
526: \begin{multline} \labell{emr1d[]}
527: \sum_{[a,b]}{'} f =
528: L^{[2k]} (\deldel{h_1}) L^{[2k]} (\deldel{h_2})
529: \left. \int_{a \pm h_1}^{b \pm h_2} f(x) dx \right|_{h_1 = h_2 = 0}
530: + (-1)^{m-1} \int_a^b P_{m}(x) f^{(m)}(x) dx
531: \end{multline}
532: for a function $f$ of type $C^m$, where $k = \lfloor m/2 \rfloor$.
533:
534: This formula becomes exact when $f$ is a polynomial
535: and when $m$ is sufficiently large.
536: We therefore have, by comparison with Equation \eqref{EM exact weighted},
537: \begin{equation} \labell{L[]=L}
538: L^{[2k]}= \Ltk.
539: \end{equation}
540: This and \eqref{emr1d[]} give \eqref{emr1d}.
541:
542: It remains to derive the expressions \eqref{P and B},
543: \eqref{P2k}, and \eqref{P2kp1} for the functions $P_{m}(x)$.
544:
545: The Bernoulli polynomials $B_m(x)$ are characterized by the properties
546: $$ B_0(x) =1 \ , \ \ \frac{d}{dx} B_m(x) = m B_{m-1}(x) \ , \ \
547: \text{and} \ \int_0^1 B_m(x) dx = 0 \ \ \text{ for } \ m \geq 1.$$
548: In particular,
549: $$ B_1(x) = x - \half = P_1(x) \quad \text{ for all } \ 0 < x < 1. $$
550: (See \eqref{P1x}.)
551: Integrating, we get that $P_m(x) = B_m(x) / m!$ for all $0 < x < 1$,
552: and since $P_m$ is 1-periodic, we get \eqref{P and B}.
553:
554: The Poisson summation formula says that, as distributions,
555: $$ \sum_{k \in \Z} \delta(x-k) = \sum_{n \in \Z} e^{2\pi i n x},$$
556: so
557: $$ P_0(x) = 1 - \sum_{k \in \Z} \delta(x-k)
558: = -2 \sum_{n=1}^\infty \cos 2\pi n x,$$
559: as distributions.
560: Integrating this, with the choice of constant of integration
561: such that the integral from $0$ to $1$ of the result vanishes,
562: gives the expressions \eqref{P2k} and \eqref{P2kp1}.
563:
564: The idea of using these Fourier expansions goes back to Wirtinger \cite{Wi}.
565: See Knopp \cite{Kn}, pp.~521--524.
566: \label{Wi-page}
567:
568: \begin{Remark}
569: From \eqref{L2k[]} and \eqref{P2k} we see that
570: the coefficients of the polynomials $L^{[2k]}(S)$ are
571: $$\displaystyle { P_{2k}(0) =
572: (-1)^{k-1} \frac{2}{(2\pi)^{2k}} \sum_{n \geq 1} \frac{1}{n^{2k}} }
573: = (-1)^{k-1} \frac{2}{(2\pi)^{2k}} \zeta(2k),$$
574: where $\zeta(\cdot)$ is the Riemann zeta function.
575: Comparing this with the coefficients of the polynomials $L^{2k}(S)$,
576: we get, from \eqref{L[]=L} and \eqref{Ltk}, that
577: $$\zeta(2k) = (-1)^{k-1} \frac{(2\pi)^{2k}}{2} \frac{1}{(2k)!} b_{2k},$$
578: reproducing Euler's famous evaluation
579: of the Bernoulli numbers in terms of the Riemann zeta function.
580: \label{zeta-page}
581: \end{Remark}
582:
583: Proposition \ref{EMintervalgrm}, when applied to a $C^{m}$ function $f(x)$
584: of compact support, implies a similar formula for an infinite ray:
585: \begin{multline} \labell{emr1dray}
586: \half f(a) + f(a+1) + f(a+2) + \ldots \\
587: = \left. \Ltk(\deldel{h}) \int_{a\pm h}^{\infty} f(x) dx
588: \right|_{h=0}
589: + (-1)^{m-1} \int_a^\infty P_{m}(x) f^{(m)}(x) dx$$
590: \end{multline}
591: for either choice of $\pm$, where $k = \lfloor m/2 \rfloor$.
592: Indeed, we need only choose $b$ so large that the support of $f(x)$
593: is contained in the set $\{ x < b \}$,
594: and then apply Proposition \ref{EMintervalgrm}.
595: Conversely, if we know (\ref{emr1dray}) for
596: functions of compact support, then we can conclude (\ref{emr1d}).
597: Indeed, multiply $f$ by a smooth function of compact support which is
598: equal to one in a neighborhood of $[a,b]$ and observe that
599: \begin{equation}
600: \labell{1dweighedLawrence}
601: {\sum_{[a,b]} }'={\sum_{[a,\infty)} }'-{\sum_{[b,\infty)} }'
602: \end{equation}
603: when applied to a function of compact support.
604:
605: \subsection*{Twisted Euler Maclaurin with remainder for a ray.}
606:
607: In extending the Euler Maclaurin formula to higher dimensions,
608: we will need an expression for the ``twisted weighted sum"
609: $$ \half f(0) + \sum_{n=1}^\infty \lambda^n f(n) $$
610: when $\lambda \neq 1$ is a root of unity, say, of order $N$,
611: in terms of the integrals of $f$.
612:
613: Let
614: \begin{equation} \labell{def:Q0}
615: Q_{0,\lambda}(x) = - \sum_{n \in \Z} \lambda^n \delta(x-n).
616: \end{equation}
617: This is an $N$-periodic distribution since $\lambda^N = 1$.
618: We will take successive anti-derivatives of this distribution,
619: with the constants chosen so that the integrals from $0$ to $N$ vanish:
620: \begin{equation} \labell{def:Qm}
621: \frac{d}{dx} Q_{m,\lambda} (x) = Q_{m-1,\lambda} (x)
622: \quad \text{and} \quad \int_0^N Q_{m,\lambda}(x) dx = 0 .
623: \end{equation}
624: With this choice of constants, $Q_{m,\lambda}(x)$ is $N$-periodic
625: for each $m$.
626:
627: Let us now look more closely at $Q_{1,\lambda}(x)$. We have
628: $$ \frac{d}{dx} {\bf 1}_{[n,n+1]}(x) = \delta(x-n)-\delta(x-(n+1)),$$
629: so
630: $$ \frac{d}{dx} \sum_{n \in \Z} \lambda^n{\bf 1}_{[n,n+1]} (x) =
631: \sum_{n \in \Z} (\lambda^n - \lambda^{n-1}) \delta(x-n)
632: = \frac{1-\lambda}{\lambda} Q_{0,\lambda} (x).$$
633: Thus,
634: $$ Q_{1,\lambda}(x) = \frac{\lambda}{1-\lambda}
635: \sum_{n \in \Z} \lambda^n {\bf 1}_{[n,n+1]} (x),
636: $$
637: because the right hand side is periodic of period $N$
638: and its integral over $[0,N]$ vanishes since
639: $1 + \lambda + \lambda^2 + \cdots + \lambda^{N-1} = 0$.
640: If $f$ is a continuously differentiable function of compact support,
641: we have
642: \begin{multline*}
643: \int_0^\infty Q_{1,\lambda}(x) f'(x) dx =
644: \frac{\lambda}{1-\lambda} \left. \sum_{n=0}^\infty \lambda^n f
645: \right|_n^{n+1}
646: = - \frac{\lambda}{1-\lambda} f(0) + \lambda f(1) + \lambda^2 f(2)
647: + \cdots
648: \end{multline*}
649: so
650: \begin{equation} \labell{firsttwistedemray}
651: \half f(0) + \sum_{n\geq 1} \lambda^n f(n) =
652: \left( \half + \frac{\lambda}{1-\lambda} \right) f(0)
653: + \int_0^\infty Q_{1,\lambda}(x) f'(x) dx.
654: \end{equation}
655:
656: Successively applying integration by parts to \eqref{firsttwistedemray},
657: we get
658: \begin{multline*}
659: \half f(0) + \lambda f(1) + \lambda^2 f(2) + \ldots \\
660: = \left( \half + \frac{\lambda}{1 - \lambda} \right) f(0)
661: - Q_{2,\lambda}(0) f'(0) - \int_0^\infty Q_{2,\lambda}(x) f''(x) dx \\
662: = \left( \half + \frac{\lambda}{1 - \lambda} \right) f(0)
663: - Q_{2,\lambda}(0) f'(0) + Q_{3,\lambda}(0) f''(0)
664: + \int_0^\infty Q_{3,\lambda}(x) f^{(3)}(x) dx \\
665: \vdots \\
666: \begin{aligned}
667: = \left( \half + \frac{\lambda}{1 - \lambda} \right) f(0)
668: - Q_{2,\lambda}(0) f'(0) + \ldots
669: + (-1)^{k-1} Q_{k,\lambda}(0) f^{(k-1)}(0) \\
670: \quad + (-1)^{k-1} \int_0^\infty Q_{k,\lambda}(x) f^{(k)}(x) dx.
671: \end{aligned}
672: \end{multline*}
673: Since
674: $$ (-1)^{m-1} f^{(m-1)}(0) = \left.
675: \left( \deldel{h} \right)^{m} \int_{-h}^\infty f(x) dx \right|_{h=0},$$
676: we have thus proved this ``twisted Euler Maclaurin formula for a ray":
677:
678: \begin{Proposition}\labell{twistedemray}
679: Let
680: $${\bfM}^{k,\lambda}(S) =
681: \left( \frac12 + \frac{\lambda}{1-\lambda} \right) S
682: + Q_{2,\lambda}(0) S^2
683: + Q_{3,\lambda}(0) S^3
684: + \cdots
685: + Q_{k,\lambda}(0) S^k,$$
686: for a root of unity $\lambda \neq 1$. Then
687: \begin{multline} \labell{twistedemrayeq}
688: \half f(0) + \lambda f(1) + \lambda^2 f(2) + \cdots =
689: \left.
690: {\bfM}^{k,\lambda}(\deldel{h}) \int_{-h}^\infty f(x) dx \right|_{h=0} \\
691: + (-1)^{k-1}\int_0^\infty Q_{k,\lambda}(x) f^{(k)}(x) dx
692: \end{multline}
693: if $f \in C_c^{k}(\R)$.
694: \end{Proposition}
695:
696: \begin{Remark}
697: Another twisted Euler Maclaurin formula for a ray appeared in \cite{T}.
698: \end{Remark}
699:
700: We will now show that the polynomials that appear in Proposition
701: \ref{twistedemray} have the following symmetry property:
702: \begin{equation}
703: \labell{Dandlambdainverse}
704: \bfM^{m,\lambda^{-1}}(S)= \bfM^{m,\lambda}(-S).
705: \end{equation}
706:
707: The linear term transforms according to \eqref{Dandlambdainverse}
708: because
709: $$ \half + \frac{\lambda\inv}{1 - \lambda\inv}
710: = \half - \left( 1 + \frac{\lambda}{1 - \lambda} \right)
711: = - \left( \half + \frac{\lambda}{1 - \lambda} \right).$$
712: %
713: %$$\half + \frac{\lambda}{1-\lambda}
714: %+ \half + \frac{\lambda^{-1}}{1-\lambda^{-1}}
715: %= 1 + \frac{\lambda}{1-\lambda} + \frac1{\lambda-1} = 0.$$
716: For the other terms to transform correctly, we need to check that
717: \begin{equation} \labell{lambdaolambdainv}
718: Q_{m,\lambda^{-1}}(0) = (-1)^m Q_{m,\lambda}(0)
719: \quad \text{for all} \quad m \geq 2.
720: \end{equation}
721:
722: As in the non-twisted formula,
723: it is convenient to work with the Fourier expansions of the
724: $Q_{m,\lambda}$'s. Let
725: $$ \lambda = e^{2\pi i \frac{j}{N}}.$$
726: We are assuming that $\lambda \neq 1$, and so $j \not \equiv 0 \mod N$.
727: Then we can write the distribution $Q_{0,\lambda}(x)$ as
728: $$ Q_{0,\lambda}(x) = - \sum_{n \in \Z} \lambda^n \delta(x-n)
729: = - e^{ (2 \pi i \frac{j}{N}) x } \sum_{n \in \Z} \delta(x-n). $$
730: Writing
731: $$ \sum_{n \in \Z} \delta(x-n) = \sum_{r \in \Z} e^{2\pi i rx },$$
732: we see that the Fourier series of $Q_{0,\lambda}(x)$ is
733: $$ Q_{0,\lambda}(x) = - \sum_{r \in \Z} e^{ 2 \pi i (\frac{j}{N} +r) x}. $$
734: The indefinite integral, chosen so that the integral over $[0,N]$
735: vanishes, is obtained by dividing each Fourier summand by the coefficient
736: of the exponent, so we have the following Fourier series:
737: \begin{equation}\labell{fourierq}
738: Q_{m,\lambda}(x) =
739: - \sum_{r \in \Z} \frac {e^{2\pi i(\frac{j}{N}+r)x}}
740: {\left(2\pi i(\frac{j}{N}+r)\right)^m}.
741: \end{equation}
742: Setting $x=0$, replacing $j$ by $-j$, and replacing $r$ by $-r$
743: in the sum, pulls out a factor of $(-1)^m$,
744: which gives the required equation \eqref{lambdaolambdainv},
745: and which implies \eqref{Dandlambdainverse}.
746:
747: \begin{Remark} \labell{M vs L}
748: For $\lambda = 1$, if we define
749: $$ \bfM^{k,1}(S) = \L^{2 \lfloor k/2 \rfloor}(S) \quad \text{and} \quad
750: Q_{k,1} = P_k,$$
751: then \eqref{twistedemrayeq} boils down to \eqref{emr1dray}.
752: So, with this notation, \eqref{twistedemrayeq} also holds for
753: $\lambda = 1$.
754: Notice that if $\lambda \neq 1$ then $M^{k,\lambda}(S)$ is a multiple
755: of $S$, and that if $\lambda = 1$ then $M^{k,\lambda}(S) = 1 + $
756: a multiple of $S$.
757: Finally, the symmetry property \eqref{Dandlambdainverse} continues to hold
758: for $\lambda = 1$ because the polynomials $\L^{2k}$ are symmetric.
759: \end{Remark}
760:
761: % ----------------------------------------------------------------
762: \section{The polar decomposition of a simple polytope}
763: % ----------------------------------------------------------------
764: \labell{sec:decompose}
765:
766: In this section we decompose a polytope into an ``alternating sum
767: of polyhedral cones". See \cite{V} or \cite{lawrence}.
768: We will give a ``weighted version" of this decomposition.
769:
770: Our polytopes are always compact and convex.
771: A compact convex polytope $\Delta$ in $\R^n$ is a compact set
772: which can be obtained as the intersection of finitely many
773: half-spaces, say,
774: \begin{equation}\labell{halfspaces}
775: \Delta = H_1 \cap \ldots \cap H_d.
776: \end{equation}
777: We assume that \eqref{halfspaces} is an intersection
778: with the smallest possible $d$, so that the $H_i$'s are uniquely
779: determined up to permutation. We order them arbitrarily.
780: The \emph{facets} (codimension one faces) of $\Delta$ are
781: $$ \sigma_i = \Delta \cap \del H_i \quad , \quad i = 1, \ldots, d .$$
782:
783: Alternatively, a compact convex polytope is the convex hull
784: of a finite set of points in $\R^n$. If we take this set to be minimal,
785: it is uniquely determined, and its elements are the \emph{vertices}
786: of $\Delta$.
787:
788: For each vertex $v$ of $\Delta$, let $I_v \subset \{ 1, \ldots, d \}$
789: encode the set of facets that contain $v$, so that
790: $$ i \in I_v \quad \text{if and only if} \quad v \in \sigma_i .$$
791:
792: Assume that $\Delta$ is \emph{simple}, so that each vertex is
793: the intersection of exactly $n$ facets.
794: For each $i \in I_v$, there exists a unique edge at $v$ which does not
795: belong to the facet $\sigma_i$; choose any vector $\alpha_{i,v}$
796: in the direction of this edge.
797: (At the moment, these ``edge vectors" are only determined
798: up to a positive scalar. Later, when discussing integral polytopes,
799: we will make a specific choice of the edge vectors.)
800:
801: The \emph{tangent cone} to $\Delta$ at $v$ is
802: $$
803: \bfC_v =
804: \{ v + r (x-v) \ \mid \ r \geq 0 \, , \, x \in \Delta \}
805: = v + \sum_{j \in I_v} \R_{\geq 0} \alpha_{j,v}.
806: $$
807: We will ``polarize" these tangent cones by flipping some of their edges
808: so that they all ``point in the same direction".
809: This direction is specified by the choice of a ``polarizing vector":
810: a vector
811: $\xi \in {\R^n}^*$, such that $\langle \xi , \alpha_{j,v} \rangle$
812: is non-zero for all $v$ and $j$. With this choice, we define
813: the \emph{polarized edge vectors} to be
814: \begin{equation} \labell{flipalpha}
815: \valpha^{\sharp}_{i,v} = \begin{cases}
816: \valpha_{i,v} & \text{ if } \l< \xi , \valpha_{i,v} \r> < 0, \\
817: - \valpha_{i,v} & \text{ if } \l< \xi , \valpha_{i,v} \r> > 0 ,
818: \end{cases}
819: \end{equation}
820: and the \emph{polarized tangent cone} to be
821: \begin{equation} \labell{pol tan cone}
822: \bfC_v^\sharp = v + \sum_{j \in I_v} \R_{\geq 0} \alpha_{j,v}^\sharp.
823: \end{equation}
824:
825: We define the ``weighted characteristic function",
826: \begin{equation} \labell{def:wted char}
827: {\bf 1}^w_\Delta(x),
828: \end{equation}
829: to be the function on $\R^n$ that takes the value $0$ on the exterior
830: of $\Delta$, the value $1$ on the interior of $\Delta$, and
831: the value $1/2^k$ on the relative interior of a codimension $k$ face
832: of $\Delta$.
833: So, for example, for an interval $[a,b]$ on the line, the function
834: ${\bf 1}^w_{[a,b]}(x)$ assigns the value
835: one to points $a<x<b$, zero to points outside the interval,
836: and $\frac12$ to the points $a$ and $b$.
837: We use a similar definition for a polyhedral cone.
838:
839: \begin{Proposition}[\protect{Weighted polar decomposition
840: of a simple polytope}]
841: \labell{weightedlawrence}
842: Let $\Delta$ be a simple polytope.
843: For any choice of polarizing vector, we have
844: \begin{equation} \labell{wted L}
845: {\bf 1}^w_\Delta (x) = \sum_v(-1)^{\#v} {\bf 1}^w_{\bfC_v^\sharp}(x),
846: \end{equation}
847: where the sum is over the vertices $v$ of $\Delta$,
848: where $\bfC_v^\sharp$ is the polarized tangent cone,
849: where ${\bf 1}^w_\Delta (x)$
850: and ${\bf 1}^w_{\bfC_v^\sharp}(x)$ are the weighted characteristic
851: functions, and where $\#v$ denotes the number of
852: edge vectors at $v$ whose signs were changed by
853: the polarizing process \eqref{flipalpha}.
854: \end{Proposition}
855:
856: This theorem is illustrated for the case of a triangle in Figure
857: \ref{fig:Lawrence}.
858:
859: \begin{figure}
860: \setlength{\unitlength}{0.00004in}
861: \begin{center}
862: \psfig{figure=wtLawrence.ps,width=\textwidth}
863: \end{center}
864: \caption{Polar decomposition of a triangle}
865: \label{fig:Lawrence}
866: \end{figure}
867:
868:
869: \vskip.2in
870: \noindent
871: \begin{proof}
872:
873: We will prove this equality in two steps.
874: First, for each $x$ we will find a polarization
875: such that the equality \eqref{wted L} holds. Second,
876: we will show that the right hand side of the equality
877: \eqref{wted L} is independent of the choice of polarization.
878:
879: Suppose that $x \not \in \Delta$.
880: Let $\xi$ be a polarizing vector such that
881: $\left< \xi , x \right> > \left< \xi , y \right> $
882: for all $y \in \Delta$. Then all the polarized cones
883: ``point away from $x$" and do not contain $x$.
884: Formula (\ref{wted L}) for the polarizing vector $\xi$,
885: when evaluated at $x$, states that $0=0$.
886:
887: For every polarization there exists exactly one vertex $v$
888: for which $\bfC_v^\sharp = \bfC_v$, namely, the vertex $v$
889: such that $\langle \xi , v \rangle$ is maximal. Conversely, for every
890: vertex $v$ there exists a polarization such that $\bfC_v^\sharp = \bfC_v$.
891:
892: Suppose that $x \in \Delta$ is in the relative interior
893: of a face $F$ of codimension $k$. Let $v$ be any vertex of $F$.
894: Let $\xi$ be a polarizing vector such that $\bfC_v^\sharp = \bfC_v$.
895: Let $F_v$ be the codimension $k$ face of $\bfC_v$ that contains $F$.
896: Then $ {\mathbf 1}_\Delta^w(x)
897: = {\mathbf 1}_{C_v^\sharp}^w (x) = \left( \frac{1}{2} \right)^k$.
898: For each other vertex, $v'$, the cone $C_{v'}^\sharp$ is disjoint
899: from the relative interior of $F$, so ${\textbf 1}_{C_v^\sharp}^w(x) = 0$.
900: Formula (\ref{wted L}) for this polarization, when evaluated at $x$,
901: states that $\left(\half\right)^k = \left(\half\right)^k$.
902:
903: To keep track of the different possible choices of polarization, we let
904: $$ E_1,\ldots,E_N $$
905: denote all the different codimension one subspaces of ${\R^n}^*$
906: that are equal to
907: $$ \alpha_{j,v}^\perp =
908: \{ \eta \in {\R^n}^* \mid \langle \eta , \alpha_{j,v} \rangle =0 \} $$
909: for some $j$ and $v$.
910: (For instance, if no two edges of $\Delta$ are parallel,
911: then the number $N$ of such hyperplanes
912: is equal to the number of edges of $\Delta$.)
913: A vector $\xi$ can be taken to be a ``polarizing vector"
914: if and only if it does not belong to any $E_j$.
915: The ``polarized cones" $\bfC_v^\sharp$ only depend on the
916: connected component of the complement
917: \begin{equation} \labell{tDelta}
918: {\R^n}^* \ssminus (E_1 \cup \ldots \cup E_N)
919: \end{equation}
920: in which $\xi$ lies.
921: Any two polarizing vectors can be connected by a path $\xi_t$
922: in ${\R^n}^*$ which crosses the ``walls" $E_j$ one at a time.
923: We finish by showing that the right hand side of formula (\ref{wted L})
924: does not change when the polarizing vector $\xi_t$ crosses
925: a single wall, $E_k$.
926:
927: As $\xi_t$ crosses the wall $E_k$,
928: the sign of the pairing $\langle \xi_t , \alpha_{j,v} \rangle$ flips
929: exactly if $E_k = \alpha_{j,v}^\perp$.
930: For each vertex $v$, denote by $\bfS_v(x)$ and $\bfS_v'(x)$
931: its contributions to the right hand side of formula
932: \eqref{wted L} before and after $\xi_t$ crossed the wall.
933: The vertices for which these contributions differ are exactly those
934: vertices that lie on edges $e$ of $\Delta$ which are perpendicular
935: to $E_k$. They come in pairs because each edge has two endpoints.
936:
937: Let us concentrate on one such an edge, $e$, with endpoints, say,
938: $u$ and $v$. Let $\alpha_e$ denote an edge vector at $v$ that points
939: from $v$ to $u$ along $e$.
940: Suppose that the pairing $\langle \xi_t , \alpha_e \rangle $
941: flips its sign from negative to positive as $\xi_t$ crosses the wall;
942: (otherwise we switch the roles of $v$ and $u$).
943: The ``polarized tangent cones" to $\Delta$ at $v$, before and after
944: $\xi_t$ crosses the wall, are
945: \begin{equation*} \begin{aligned}
946: \bfC_v^\sharp &= v + \sum_{j \in I_e} \R_{\geq 0} \alpha_{j,v}^\sharp
947: + \R_{\geq 0} \alpha_e
948: \quad \text{and} \\
949: (\bfC_v^\sharp)' & = v + \sum_{j \in I_e} \R_{\geq 0} \alpha_{j,v}^\sharp
950: - \R_{\geq 0} \alpha_e
951: \end{aligned} \end{equation*}
952: where $I_e \subset \{ 1, \ldots, d\}$ encodes the facets that contain $e$.
953: (The $\alpha_{j,v}^\sharp$ are the same for the different $\xi_t$'s
954: because the pairings
955: $ \langle \xi_t , \alpha_{j,v}^\sharp \rangle $ do not flip sign
956: when $\xi_t$ crosses the wall for $j \in I_e$.)
957: The cones $\bfC_v^\sharp$ and $(\bfC_v^\sharp)'$ have a common facet
958: and their union is
959: $$ \bfC_e^\sharp := v + \sum_{j \in I_e} \R_{\geq 0} \alpha_{j,v}^\sharp
960: + \R \alpha_e.$$
961: This union only depends on the edge $e$ and not on the endpoint $v$.
962: (This uses the assumption that the polytope $\Delta$ is simple
963: and follows from the fact that
964: $\alpha_{j,u} \in \R_{\geq 0} \alpha_{j,v} + \R \alpha_e$.)
965:
966: The contributions of $v$ to the right hand side of (\ref{wted L})
967: before and after
968: $\xi_t$ crosses the wall are
969: \begin{equation} \labell{bfSv}
970: \bfS_v(x) = \varepsilon {\bf 1}^w_{\bfC_v^\sharp} \quad \text{and} \quad
971: \bfS_v'(x) = -\varepsilon {\bf 1}^w_{(\bfC_v^\sharp)'}
972: \end{equation}
973: where $\varepsilon \in \{ -1 , 1 \}$.
974: Their difference is plus/minus the weighted characteristic function
975: of $\bfC_e^\sharp$:
976: \begin{equation} \labell{difference}
977: \bfS_v(x) - \bfS_v'(x) = \varepsilon {\bf 1}^w_{\bfC_e^\sharp}
978: \end{equation}
979: The contributions of the other endpoint, $u$, have opposite signs
980: than the respective contributions \eqref{bfSv} of $v$,
981: and their difference is minus/plus the characteristic function
982: of $\bfC_e^\sharp$. Hence, the differences $\bfS_v(x)-\bfS_v'(x)$
983: and $\bfS_u(x)-\bfS_u'(x)$, for the two endpoints $u$ and $v$ of $e$,
984: sum to zero.
985: \end{proof}
986:
987: % ---------------------------------------------------------------------
988: \section{Euler Maclaurin formula with remainder for regular polytopes.}
989: % ---------------------------------------------------------------------
990: \labell{sec:EM regular}
991: We extend our notation \eqref{wted sum}
992: to a weighted sum over a simple integral polytope~$\Delta$:
993: $$ \sum_{\Delta \cap \Z^n}{'} f := \sum_{x \in \Delta \cap \Z^n}
994: {\bf 1}_\Delta^w(x) f(x), $$
995: where ${\bf 1}_\Delta^w(x)$ is the weighted characteristic function,
996: introduced in \eqref{def:wted char}, which is equal to $1/2^k$
997: when $x$ lies in the relative interior of a face of $\Delta$
998: of codimension $k$.
999: We use similar notation for the weighted sum of a compactly supported
1000: function $f$ over a simple polyhedral cone.
1001:
1002: With this notation, Proposition \ref{weightedlawrence}
1003: gives the following decomposition. Let $\xi \in {\R^n}^*$
1004: be a ``polarizing vector". Then
1005: \begin{equation} \labell{wheee}
1006: \sum_{\Delta \cap \Z^n}{'} f = \sum_v (-1)^{\# v}
1007: \sum_{\bfC_v^\sharp \cap \Z^n}{'} f
1008: \end{equation}
1009: where we sum over the vertices $v$ of $\Delta$,
1010: where $\bfC_v^\sharp$ is the polarized tangent cone to $\Delta$ at $v$
1011: (see \eqref{pol tan cone}), and where $\# v$ is the number of edge vectors
1012: at $v$ that are flipped by the polarization process \eqref{flipalpha}.
1013:
1014: \bigskip
1015:
1016: For the standard closed orthant
1017: $\bfO = \prod_{i=1}^n \R_{\nonneg}$ in $\R^n$, we have
1018: $${\sum_{\bfO \cap \Z^n}}' g
1019: = {\sum_{m_1 \in \Z_{\nonneg}}}' \cdots
1020: {\sum_{m_n \in \Z_{\nonneg}}}' g(m_1,\dots,m_n) , $$
1021: for any function $g$ of compact support.
1022:
1023: By performing $n$ iterations of \eqref{emr1dray}, we obtain
1024: an Euler Maclaurin formula with remainder for the standard orthant:
1025: Let $m$ be an integer and let $k = \lfloor m/2 \rfloor$.
1026: If $g$ is a $C^{mn}$ function of compact support and $m\geq 1$, then
1027: for any choice of the $\pm$'s we have
1028: \begin{equation} \labell{EMstandardorthant}
1029: {\sum_{\bfO\cap \Z^n}}' g = \left. \prod_{i=1}^n
1030: \Ltk(\deldel{h_i}) \int \limits_{\bfO(\pm h_1,\dots,\pm h_n)} g(x) dx
1031: \right|_{h_1=\cdots=h_n=0} + R^{\st}_{m}(g),
1032: \end{equation}
1033: where
1034: $$\bfO(h_1,\dots,h_n)=\{t \mid t_i + h_i \geq 0 \ \forall i \}$$
1035: denotes the shifted orthant, and where the remainder term
1036: is given by
1037: \begin{multline} \labell{standardremainder1}
1038: R^{\st}_{m}(g) = \\
1039: \sum_{I\subsetneq\{1,\dots,n\}} (-1)^{(m-1)(n-|I|)} \prod_{i\in I}
1040: \Ltk(\deldel{h_i})
1041: \left.
1042: \int\limits_{\bfO(\pm h_1,\dots,\pm h_n)}
1043: \prod_{i\not\in I}P_{m}(x_i)\prod_{i\not\in I}
1044: \left(\frac{\partial}{\partial x_i}\right)^{m} g(x) dx_1 \cdots dx_n
1045: \right|_{h=0} .
1046: \end{multline}
1047: This remainder can also be expressed as a sum of integrals
1048: over the orthant of bounded periodic functions times various
1049: partial derivatives of $f$ of order no less than $m$ and no more than $mn$.
1050: This fact follows from the formula
1051: \begin{equation} \labell{formula with varphi}
1052: \deldel{h_i} \int\limits_{\bfO(h_1, \ldots, h_n)}
1053: \varphi dx_1 \cdots dx_n =
1054: - \int\limits_{\bfO(h_1, \ldots, h_n)}
1055: \frac{\del \varphi}{\del x_i} dx_1 \cdots dx_n.
1056: \end{equation}
1057: We can apply iterations of this formula
1058: with $i \in I$ to the $I$'th summand of \eqref{standardremainder1}
1059: because the non-smooth functions $P_m$ are only applied to the
1060: variables $x_j$ for $j \not \in I$. We get
1061: \begin{equation} \labell{standardremainder}
1062: R^{\st}_{m}(g) = \sum_{I \subsetneq \{ 1, \ldots, n \} }
1063: (-1)^{(m-1)(n-|I|)} \int\limits_\bfO
1064: \prod_{i \in I} \Ltk (-\deldel{x_i})
1065: \prod_{i \not \in I} P_m(x_i)
1066: \prod_{i \not \in I} (\deldel{x_i})^m
1067: g(x) dx_1 \ldots dx_n.
1068: \end{equation}
1069:
1070: \bigskip
1071:
1072: A regular integral orthant $\bfC$ is the image of the standard orthant $\bfO$
1073: via an affine transformation
1074: of the form
1075: \begin{equation} \labell{affine}
1076: (t_1,\ldots,t_n) \mapsto x = v + t_1 \alpha_1 + \ldots + t_n \alpha_n
1077: \end{equation}
1078: where $\alpha_1,\ldots,\alpha_n$ generate $\Z^n$
1079: and where $v \in \Z^n$.
1080: If $u_1,\ldots,u_n \in {\R^n}^*$
1081: is the dual basis to $\alpha_1,\ldots,\alpha_n$
1082: then the image of $\bfO(h_1,\ldots,h_n)$ under this transformation
1083: is given by the inequalities
1084: \begin{equation}\labell{uiss}
1085: \langle u_i , x \rangle-\langle u_i , v \rangle+h_i\geq 0.
1086: \end{equation}
1087: We denote this expanded orthant by ${\bf C}(h)$. If $f$ is a $C^{mn}$
1088: function of compact support and
1089: $$ g(t_1,\ldots,t_n) = f(t_1 \alpha_1 + \ldots + t_n \alpha_n) $$
1090: is its pullback under the transformation \eqref{affine}, then
1091: $${\sum_{{\bfC}\cap \Z^n}}'f={\sum_{\bfO \cap \Z^n}}'g \quad \text{and} \quad
1092: \int\limits_{{\bfC}(h)}f (x) dx
1093: =\int\limits_{\bfO(h)}g(t) dt,$$
1094: and
1095: so we have an Euler Maclaurin formula for regular orthants:
1096: \begin{equation}
1097: \labell{EMregularorthant}
1098: {\sum_{{\bf C}\cap \Z^n}}'f=
1099: \left. \prod_{i=1}^n\Ltk(\deldel{h_i})
1100: \int\limits_{{\bf C}(\pm h_1,\dots,\pm h_n)}f(x)dx
1101: \right|_{h_1=\cdots=h_n=0} +R^{\bf C}_{m}(f),
1102: \end{equation}
1103: where
1104: $$R^{\bf C}_{m}(f)=R^{\st}_{m}(g).$$
1105:
1106: \bigskip
1107:
1108: Let $\Delta$ be a compact convex polytope.
1109: As in Section \ref{sec:decompose},
1110: $\Delta$ can be written as an intersection of half-spaces
1111: \begin{equation} \labell{Delta}
1112: \Delta = H_1 \cap \ldots \cap H_d, \quad \text{where} \quad
1113: H_i = \{ x \mid \langle u_i , x \rangle + \mu_i \geq 0 \}
1114: \quad \text{ for } i=1,\ldots,d
1115: \end{equation}
1116: and $d$ is the number of facets of $\Delta$.
1117: The vector $u_i \in {\R^n}^*$ can be thought of as the inward normal
1118: to the $i$th facet of $\Delta$;
1119: a-priori it is determined up to multiplication by a positive number.
1120: If all the vertices of $\Delta$ are integral, then the $u_i$'s can
1121: be chosen to belong to the dual lattice ${\Z^n}^*$, and we can
1122: fix our choice of the $u_i$'s by imposing
1123: the normalization condition that the $u_i$'s be primitive lattice elements,
1124: that is, that no $u_i$ can be expressed as a multiple of a lattice
1125: element by an integer greater than one.
1126: (The fact that a normal vector $u$ to a facet $\sigma$
1127: can be chosen to be integral is a consequence of Cramer's rule.
1128: Indeed, we can choose integral edge vectors
1129: $\beta_1,\ldots,\beta_n$ that emanate from a vertex on $\sigma$
1130: such that $\beta_1,\ldots\beta_{n-1}$ span the tangent plane to $\sigma$
1131: and $\beta_n$ is transverse to $\sigma$.
1132: Solving the linear equations $\left< u , \beta_1 \right> =$
1133: $\ldots$ $=\left< u , \beta_{n-1} \right> = 0$ and
1134: $\left< u , \beta_{n} \right> = 1$,
1135: we get an inward normal vector $u$ with rational entries;
1136: clearing denominators, we may assume that $u$ is actually integral.)
1137:
1138: We can then consider the ``dilated polytope" $\Delta(h_1,\ldots,h_d)$,
1139: which is obtained by shifting the $i$th facet outward
1140: by a ``distance" $h_i$. More precisely,
1141: $$ \Delta(h) = \bigcap_{i=1}^d \{ x \mid \langle u_i , x \rangle
1142: + \mu_i + h_i \geq 0 \}
1143: \quad \text{where} \quad h = (h_1,\ldots,h_d). $$
1144:
1145: \bigskip
1146:
1147: Now assume that $\Delta$ is simple.
1148: Then $\Delta(h)$ is simple if $h$ is sufficiently small.
1149: The polar decomposition of $\Delta(h)$ involves ``dilated orthants".
1150: However, dilating the facets of $\Delta$ outward results in dilating
1151: some facets of $\bfC_v^\sharp$ inward and some outward.
1152: Explicitly, for $i \in I_v = \{ i_1, \ldots, i_n \} $,
1153: the inward normal vector to the $i$th facet of $\bfC_v^\sharp$ is
1154: \begin{equation} \labell{u i v sharp}
1155: u_{i,v}^\sharp = \begin{cases}
1156: u_i & \text{ if } \alpha_{i,v}^\sharp = \alpha_{i,v} \\
1157: -u_i & \text{ if } \alpha_{i,v}^\sharp = -\alpha_{i,v} .
1158: \end{cases}
1159: \end{equation}
1160: Hence, the dilated orthants that occur on the right hand side
1161: of the polar decomposition of $\Delta(h)$
1162: are $\bfC_v^\sharp(h_{i_1,v}^\sharp, \ldots, h_{i_n,v}^\sharp)$,
1163: where
1164: \begin{equation} \labell{h i v sharp}
1165: h_{i,v}^\sharp = \begin{cases}
1166: h_i & \text{ if } \alpha_{i,v}^\sharp = \alpha_{i,v} \\
1167: -h_i & \text{ if } \alpha_{i,v}^\sharp = - \alpha_{i,v}. \end{cases}
1168: \end{equation}
1169: This subtlety in the signs does not effect the formula for regular polytopes,
1170: because of the symmetry of $\Ltk$; however, it is needed to derive
1171: the formula for simple polytopes.
1172:
1173: If $\Delta$ is a regular integral polytope,
1174: then the $C_v^\sharp$'s are regular integral orthants,
1175: and the ``dilated orthants" are exactly as in \eqref{uiss}.
1176: We then have, by \eqref{wheee}, \eqref{EMregularorthant},
1177: and the symmetry of $\Ltk$,
1178: \begin{multline}
1179: \labell{cow}
1180: {\sum_{\Delta\cap \Z^n}}'f
1181: = \sum_v (-1)^{\#v}{\sum_{\bfC_v^\sharp \cap \Z^n}}'f \\
1182: = \sum_v(-1)^{\#v}\left(\left. \prod_{i\in I_v=\{i_1,\dots,i_n\}}
1183: \Ltk(\deldel{h_i})
1184: \int\limits_{\bfC_v^\sharp (\pm h_{i_1},\dots,\pm h_{i_n})}f(x)dx
1185: \right|_{h=0} +R^{\bfC_v^\sharp}_{m}(f)\right).
1186: \end{multline}
1187: We may multiply the differential operator $\prod \Ltk(\deldel{h_i})$
1188: in the above expression by any number of operators of the form
1189: $\Ltk(\deldel{h_j})$ where $j \not\in I_v$,
1190: since all that will remain of these operators is
1191: the constant term $1$, all actual differentiations yielding zero.
1192: The right hand side of \eqref{cow} is then equal to
1193: \begin{multline*}
1194: \left. \prod_{i=1}^d \Ltk(\deldel{h_i}) \sum_v (-1)^{\# v}
1195: \int\limits_{\bfC_v^\sharp (h_{i_1,v}^\sharp,\dots,h_{i_n,v}^\sharp)} f(x)dx
1196: \right|_{h=0} + R^{\bfC_v^\sharp}_{m}(f) \\
1197: = \left. \prod_{i=1}^d\Ltk(\deldel{h_i})\int\limits_{{\Delta}(h_1,\dots,h_d)}
1198: f(x)dx\right|_{h=0}+
1199: S_\Delta^{m}(f)
1200: \end{multline*}
1201: where
1202: \begin{equation}
1203: \labell{polytoperemainder}
1204: S_\Delta^{m}(f):=\sum_v(-1)^{\#v}R^{\bfC_v^\sharp}_{m}(f)
1205: \end{equation}
1206: and where $\{ i_1, \ldots, i_n \} = I_v$ in the $v$'th summand.
1207:
1208: Notice that both
1209: $${{\sum_{\Delta\cap \Z^n}}'f}$$
1210: and
1211: $$\left. \prod_{i=1}^d\Ltk(\deldel{h_i})\int\limits_{{\Delta}(h_1,\dots,h_d)}
1212: f(x)dx\right|_{h=0}$$
1213: do not
1214: depend on the choice of polarization, and both vanish on any function $f$
1215: whose support
1216: is disjoint from the polytope. So the same must be true of the remainder.
1217: We have proved the following result:
1218:
1219: \begin{Theorem}[\cite{PNAS}] \labell{EMregular}
1220: Let $m \geq 1$ be an integer.
1221: Let $\Delta \subset \R^n$ be a regular integral
1222: polytope and $f$ a $C^{mn}$ function
1223: of compact support on $\R^n$. Choose a ``polarizing vector" for $\Delta$.
1224: Then
1225: $${\sum_{\Delta\cap \Z^n}}'f=
1226: \left. \prod_{i=1}^d\Ltk(\deldel{h_i})\int\limits_{{\Delta}(h_1,\dots,h_d)}
1227: f(x)dx\right|_{h=0}+ S_\Delta^{m}(f)$$
1228: where $k = \lfloor m/2 \rfloor$ and
1229: where $S_\Delta^{m}(f)$ is given by \eqref{polytoperemainder}.
1230: This remainder can be expressed as a sum of integrals over orthants
1231: of bounded periodic functions times various partial derivatives
1232: of $f$ of order no less than $m$ and no more than $mn$. Finally,
1233: this remainder is independent of the choice of
1234: polarization and is a distribution supported on the polytope $\Delta$.
1235: \end{Theorem}
1236:
1237: % -----------------------------------------------------------------------------
1238: \section{Finite groups associated to a simple integral polytope and its faces.}
1239: % -----------------------------------------------------------------------------
1240: \labell{sec:groups}
1241:
1242: In order to extend Theorem \ref{EMregular} to simple integral polytopes
1243: that may not be regular, we must extend the Euler Maclaurin
1244: formula for regular orthants (\ref{EMregularorthant}) to a
1245: formula that is valid for simple orthants that may not be regular.
1246: In this section we analyze
1247: certain finite groups that arise in this generalization.
1248:
1249: Let $\bfC$ be a simple integral orthant.
1250: This means that we can write $\bfC$ as the intersection of $n$ half-planes
1251: in general position,
1252: \begin{equation} \labell{general position}
1253: \bfC = H_1 \cap \ldots \cap H_n \quad \text{where} \quad
1254: H_i = \{ x \mid \langle u_i , x \rangle + \mu_i \geq 0 \}
1255: \quad \text{ for } i=1,\ldots,n,
1256: \end{equation}
1257: and that the vertex of $\bfC$ is in $\Z^n$.
1258: The $u_i$'s are inward normals to the facets of $\bfC$,
1259: and we choose them to be primitive elements
1260: of the dual lattice ${\Z^n}^*$. (See Section \ref{sec:EM regular}.)
1261: We choose $\alpha_1,\ldots,\alpha_n$ to be the dual basis,
1262: that is,
1263: $$ \l< u_j , \alpha_i \r> = \begin{cases}
1264: 1 & j=i \\ 0 & j \neq i.
1265: \end{cases}$$
1266: The $\alpha_i$'s are edge vectors of $\bfC$, but they might not be integral:
1267: they generate a lattice in $\R^n$ which is a finite extension of $\Z^n$.
1268: This extension is trivial exactly if $\Delta$ is regular at $v$,
1269: that is, if the $u_i$'s, generate the dual lattice ${\Z^n}^*$.
1270: To the cone $\bfC$ we associate the finite group
1271: \begin{equation} \labell{Gamma}
1272: \Gamma := {\Z^n}^* / \sum \Z u_i .
1273: \end{equation}
1274: So this group is trivial exactly if the cone $\bfC$ is regular.
1275:
1276: \begin{Lemma} \labell{character}
1277: In the above setting,
1278: \begin{equation} \labell{the character}
1279: \gamma \mapsto e^{2 \pi i \left< \gamma , x \right> }
1280: \end{equation}
1281: is a well defined character on $\Gamma$ whenever
1282: $x = m_1 \alpha_1 + \ldots + m_n \alpha_n$ where $m_j$ are integers.
1283: This character is trivial if and only if $x \in \Z^n$.
1284: \end{Lemma}
1285:
1286: \begin{proof}
1287: Let $ \tilde{\gamma} \in {\Z^n}^*$ be an element that represents $\gamma$.
1288: If we expand it as a combination of the basis elements $u_j$,
1289: so that
1290: $$ \tilde{\gamma} = b_1 u_1 + \ldots + b_n u_n,$$
1291: then $\left< \tilde{\gamma} , \alpha_j \right> = b_j$ for all $j$.
1292: If $\tilde{\gamma}'$ is another element of ${\Z^n}^*$
1293: that represents $\gamma$, then it differs from $\tilde{\gamma}$
1294: by an integral combination of the $u_i$'s. So
1295: $\left< \tilde{\gamma}' , \alpha_j \right> $
1296: differs from $ \left< \tilde{\gamma} , \alpha_j \right> $
1297: by an integer.
1298: Hence, when $x$ is an integral combination of the $\alpha_j$'s,
1299: the pairing
1300: $\left< \tilde{\gamma} , x \right> $ is well defined modulo $\Z$,
1301: and so $e^{2 \pi i \left< \gamma,x \right>}$ is well defined.
1302:
1303: Finally, the character \eqref{the character} is trivial
1304: if and only if $\left< \tilde{\gamma} , x \right> $ is an integer
1305: for all $\gamma \in {\Z^n}^*$, and this holds if and only if $x \in \Z^n$.
1306: \end{proof}
1307:
1308: \bigskip
1309:
1310: Let $\Delta$ be a simple integral polytope in $\R^n$,
1311: given by \eqref{Delta}.
1312: For each vertex $v$ of $\Delta,$ let $I_v \subset \{ 1 , \ldots , d \}$
1313: denote the set of facets of $\Delta$ that meet at $v$.
1314: The normal vectors $u_i$,
1315: $i \in I_v$, form a basis of ${\R^n}^*$.
1316: We choose
1317: \begin{equation} \labell{alpha i v}
1318: \valpha_{i,v} \in \R^n \quad , \quad i \in I_v
1319: \end{equation}
1320: to be the dual basis.
1321:
1322: Given any face $F$ of the polytope $\Delta$,
1323: let $I_F$ denote the set of facets of $\Delta$ which meet at $F$.
1324: Because $\Delta$ is simple, the vectors $u_i$, for $i \in I_F$,
1325: are linearly independent.
1326: Let $N_F \subseteq {\R^n}^*$ be the subspace
1327: $$ N_F = \span \{ u_i \mid i \in I_F \}.$$
1328:
1329: \begin{Remark}
1330: It is natural to define the tangent space to the face $F$ to be
1331: $TF = \span \{x-y \mid x,y \in F\}$,
1332: the normal space to $F$ to be the quotient $\R^n / TF$,
1333: and the co-normal space to be the dual of the normal space.
1334: With these definitions, $N_F$ is the co-normal space to $F$.
1335: \end{Remark}
1336:
1337: To each face $F$ of $\Delta$ we associate a finite abelian
1338: group $\Gamma_F$. Explicitly, the lattice
1339: $$\vV_F = \sum_{i \in I_F} \Z u_i \, \subset \, N_F $$
1340: is a sublattice of $N_F \cap {\Z^n}^*$ of finite index, and
1341: the finite abelian group associated to the face $F$ is
1342: the quotient
1343: \begin{equation} \labell{def GammaF}
1344: \Gamma_F := (N_F \cap {\Z^n}^*) / \vV_F.
1345: \end{equation}
1346: If $F=v$ is a vertex, this is the same as the finite abelian group
1347: associated to the tangent cone $\bfC_v$ as in \eqref{Gamma}.
1348:
1349: Let $E$ and $F$ be two faces of $\Delta$ with $F \subseteq E$.
1350: This inclusion implies that $I_E \subseteq I_F$, and hence
1351: \begin{equation} \labell{uis}
1352: \{ u_i \}_{i \in I_E} \subseteq \{ u_i \}_{i \in I_F}.
1353: \end{equation}
1354: Because these sets are bases of the vector spaces $N_E$ and $N_F$,
1355: we have an inclusion
1356: $$N_E \subseteq N_F.$$
1357: Because the sets occurring in \eqref{uis}
1358: are $\Z$-bases of the lattices $\vV_E$ and $\vV_F$,
1359: we have $N_E \cap \vV_F = \vV_E$.
1360: Hence, the natural map from $\Gamma_E = ({\Z^n}^* \cap N_E) / \vV_E$
1361: to $\Gamma_F = ({\Z^n}^* \cap N_F) / \vV_F$ is one to one,
1362: and it provides us with a natural inclusion map:
1363: $$ \text{if \ $F \subseteq E$ \ then \ $\Gamma_E \subseteq \Gamma_F$.} $$
1364:
1365: We define a subset $\Gamma_F^\flat$ of $\Gamma_F$ by
1366: \begin{equation} \labell{def:GammaFsharp}
1367: \Gamma_F^\flat
1368: := \Gamma_F \ssminus
1369: \bigcup_{\text{faces } E \text{ such that } E \supsetneq F} \Gamma_E.
1370: \end{equation}
1371: Then
1372: \begin{equation} \labell{Gammas}
1373: \Gamma_v = \bigsqcup_{\{F:v \in F\}} \Gamma_F^\flat.
1374: \end{equation}
1375:
1376: Recall that
1377: \begin{equation} \labell{lambda j v}
1378: \lambda_{\gamma,j,v} :=
1379: e^{ 2 \pi i \left< \gamma ,\alpha_{j,v} \right> } ,
1380: \quad \text{ for } \gamma \in \Gamma_v
1381: \quad \text{ and } j \in I_v,
1382: \end{equation}
1383: is well defined, by Lemma \ref{character}. It is a root of unity.
1384: We will need the following results.
1385:
1386: \begin{Claim} \labell{claim1}
1387: If $\gamma \in \Gamma_F$ and $j \in I_F$,
1388: then $\lambda_{\gamma,j,v}$ is the same for all $v \in F$.
1389: \end{Claim}
1390:
1391: This allows us to define $\lambda_{\gamma,j,F}$
1392: for $\gamma \in \Gamma_F$ and $j \in I_F$ such that
1393: $$ \lambda_{\gamma,j,F} = \lambda_{\gamma,j,v}
1394: \quad \text{ for } \gamma \in \Gamma_F
1395: \text{ and } j \in I_F, \text{ if } v \in F .$$
1396:
1397: \begin{Claim} \labell{claim2}
1398: If $\gamma \in \Gamma_F$ and $j \in I_v \ssminus I_F$
1399: then $\lambda_{\gamma,j,v}$ is equal to one.
1400: \end{Claim}
1401:
1402: This allows us to define
1403: $\lambda_{\gamma,j,F} = 1$ when $\gamma \in \Gamma_F$
1404: and $j \in \{ 1 , \ldots, d \} \ssminus I_F $.
1405: Then
1406: \begin{equation} \labell{horse}
1407: \lambda_{\gamma,j,F} = \lambda_{\gamma,j,v}
1408: \quad \text{ for } \gamma \in \Gamma_F
1409: \text{ and } 1 \leq j \leq d , \text{ if } v \in F
1410: \end{equation}
1411: and
1412: \begin{equation} \labell{lambda is one}
1413: \lambda_{\gamma,j,F} = 1 \quad \text{ for } \gamma \in \Gamma_F
1414: \text{ if } j \not\in I_F.
1415: \end{equation}
1416:
1417: \begin{Claim} \labell{claim3}
1418: If $\gamma \in \Gamma_F^\flat$ and $j \in I_F$,
1419: then $ \lambda_{\gamma,j,F} \neq 1$.
1420: \end{Claim}
1421:
1422: \begin{proof}[Proof of Claims \ref{claim1}, \ref{claim2}, and \ref{claim3}]
1423: Let $\gamma \in \Gamma_F$ be represented by
1424: $$ \tilde{\gamma} =
1425: \sum_{i \in I_F} b_i u_i \in N_F \cap {\Z^n}^* $$
1426: for some $b_i \in \R$. (See \eqref{def GammaF}.)
1427:
1428: Let $v \in F$. Because $\alpha_{j,v}$, $j \in I_v$,
1429: is a dual basis to $u_j$, $j \in I_v$, we have
1430: $$ \left< \tilde{\gamma} , \alpha_{j,v} \right>
1431: = \begin{cases} b_j & j \in I_F \\ 0 & j \in I_v \ssminus I_F.
1432: \end{cases}$$
1433: Hence,
1434: $$ \lambda_{\gamma,j,v}
1435: = e^{2 \pi i \left< \tilde{\gamma} , \alpha_{j,v} \right> }
1436: = \begin{cases} e^{ 2 \pi i b_j} & j \in I_F \\
1437: 1 & j \in I_v \ssminus I_F \end{cases}
1438: $$
1439: is independent of $v$ and is equal to $1$
1440: if $j \in I_v \ssminus I_F$.
1441: This prove Claims \ref{claim1} and \ref{claim2}.
1442:
1443: Let $j \in I_F$. If
1444: $ \lambda_{\gamma,j,F} := e^{ 2 \pi i b_j } $
1445: is equal to one, then $b_j$ is an integer, so
1446: \begin{equation} \labell{in GammaE}
1447: \tilde{\gamma}' =
1448: \sum_{i \in I_F \ssminus \{ j \} } b_i u_i
1449: \end{equation}
1450: also represents $\gamma$. Let $E \supset F$
1451: be the face such that $I_E = I_F \ssminus \{ j \}$.
1452: Then, by \eqref{in GammaE}, $\gamma \in \Gamma_E$.
1453: This proves Claim \ref{claim3}.
1454: \end{proof}
1455:
1456: % -----------------------------------------------------------------------------
1457: \section{Euler Maclaurin formula with remainder for simple integral polytopes.}
1458: % -----------------------------------------------------------------------------
1459: \labell{sec:EM simple}
1460:
1461: Let us begin by deriving an Euler Maclaurin formula with remainder
1462: for a simple integral orthant.
1463:
1464: We recall the set-up of Section \ref{sec:groups}.
1465: Let $\bfC$ be a simple integral orthant. Let $u_1, \ldots, u_n$
1466: be the inward normals to its facets, chosen to be primitive
1467: elements of the dual lattice ${\Z^n}^*$,
1468: and let $\alpha_1 , \ldots, \alpha_n$
1469: be the dual basis to the $u_i$'s, so that
1470: $$ \bfC = v + \sum_{j=1}^n \R_{\nonneg} \alpha_j.$$
1471: Let
1472: $$ \Gamma = {\Z^n}^* / \sum \Z u_j $$
1473: be the finite group associated to $\bfC$.
1474: By Lemma \ref{character},
1475: $\gamma \mapsto e^{ 2 \pi i \left< \gamma, x \right> }$
1476: defines a character on $\Gamma$
1477: whenever $x \in \sum \Z \alpha_j$,
1478: and this character is trivial if and only if $x \in \Z^n$.
1479: By a theorem of Frobenius, the average value of a character
1480: on a finite group is zero if the character is non-trivial
1481: and one if the character is trivial. So
1482: $$ \frac{1}{| \Gamma |} \sum_{\gamma \in \Gamma}
1483: e^{ 2 \pi i \left< \gamma, x \right> } = \begin{cases}
1484: 1 & \text{if } x \in \Z^n \\
1485: 0 & \text{if } x \not \in \Z^n \\
1486: \end{cases} $$
1487: for all $x \in \sum \Z \alpha_j$.
1488: Then, for a compactly supported function $f(x)$ on $\R^n$,
1489: \begin{eqnarray}
1490: \nonumber
1491: {\sum_{\bfC \cap \Z^n}}{'} f
1492: & = & {\sum_x}' \left( \frac{1}{|\Gamma|}
1493: \sum_{\gamma \in \Gamma} e^{2\pi i \left< \gamma , x \right>} \right) f(x) \\
1494: \labell{eqAA}
1495: & = & \frac{1}{|\Gamma|} \sum_{\gamma \in \Gamma} {\sum_x}{'}
1496: e^{2\pi i \left< \gamma , x \right>} f(x)
1497: \end{eqnarray}
1498: where we sum over all
1499: \begin{equation} \labell{x}
1500: x = v + m_1 \alpha_1 + \ldots + m_n \alpha_n ,
1501: \end{equation}
1502: where the $m_i$'s are non-negative integers.
1503:
1504: The cone $\bfC$ is the image of the standard orthant $\bfO$
1505: under the affine map
1506: \begin{equation} \labell{affine map}
1507: (t_1,\ldots,t_n) \mapsto x = v + t_1 \alpha_1 + \ldots + t_n \alpha_n .
1508: \end{equation}
1509: This map sends the lattice $\Z^n$ onto the lattice
1510: $\sum \Z \alpha_j$. The inverse transformation is given by
1511: \begin{equation} \labell{inverse transformation}
1512: t_i = \left< u_i , x-v \right> .
1513: \end{equation}
1514:
1515: Let us concentrate on one element $\gamma \in \Gamma$.
1516: Because $v \in \Z^n$, from \eqref{x} we get
1517: $$ e^{2\pi i \left< \gamma , x \right>}
1518: = \prod_{j=1}^n \lambda_{j} ^{m_j}
1519: \quad \text{where} \quad
1520: \lambda_{j} = e^{2\pi i\left< \gamma , \alpha_j \right> } ,$$
1521: so that the inner sum in (\ref{eqAA}) becomes
1522: \begin{equation} \labell{eqBB}
1523: {\sum_{m_1 \geq 0}}' \lambda_{1}^{m_1} \cdots
1524: {\sum_{m_n \geq 0}}' \lambda_{n}^{m_n} g(m_1,\ldots,m_n),
1525: \end{equation}
1526: where
1527: $$ g(t_1,\ldots,t_n) =
1528: f(v + t_1 \alpha_1 + \ldots + t_n \alpha_n). $$
1529:
1530: Recall that we had the twisted remainder formula
1531: $$ {\sum_{m \geq 0}}' \lambda^m g(m)
1532: = \left. \bfM^{k,\lambda}(\deldel{h}) \int_{-h}^\infty g(t) dt \right|_{h=0}
1533: + (-1)^{k-1} \int_0^\infty Q_{k,\lambda}(t) g^{(k)}(t) dt$$
1534: for all compactly supported functions $g(x)$ of type $C^k$,
1535: where $k \geq 1$, where $\lambda$ is a root of unity,
1536: and where $M^{k,\lambda}$ is a polynomial of degree $\leq k$.
1537: (See \eqref{twistedemrayeq} and Remark \ref{M vs L}.)
1538:
1539: Iterating this formula, the sum in \eqref{eqBB} can be written as
1540: $$
1541: \bfM^{k,\lambda_{1}}(\deldel{h_1}) \int_{-h_1}^\infty \cdots
1542: \bfM^{k,\lambda_{n}}(\deldel{h_n}) \int_{-h_n}^\infty
1543: g(t_1,\ldots,t_n) dt_1 \cdots dt_n
1544: + R^{st}_k(\lambda_{1},\dots,\lambda_{n};g)
1545: $$
1546: \begin{equation} \labell{integral over Oh}
1547: = \prod_{i=1}^n \bfM^{k,\lambda_i}(\deldel{h_i})
1548: \int\limits_{\bfO(h)} g(t_1,\ldots,t_n) dt_1 \cdots dt_n
1549: + R^{st}_k(\lambda_{1},\ldots,\lambda_{n};g)
1550: \end{equation}
1551: with
1552: $$ \bfO(h) = \{ (t_1,\ldots,t_n) \ | \ t_i \geq -h_i \text{\ for all $i$} \}$$
1553: and where the remainder is given by
1554: \begin{multline*}
1555: R^{st}_k(\lambda_1,\dots,\lambda_n ; g) = \\
1556: \sum_{I \subsetneq \{ 1, \ldots, n \} } (-1)^{(k-1)(n-|I|)}
1557: \prod_{i \in I} \bfM^{k,\lambda_i}(\deldel{h_i})
1558: \int\limits_{O(h)}
1559: \left.
1560: \prod_{i \notin I} Q_{k,\lambda_i} (t_i)
1561: \prod_{i \notin I} \frac{\del^{k}}{\del t_i ^{k}}
1562: g(t_1,\ldots,t_n) dt_1 \cdots dt_n
1563: \right|_{h=0} .
1564: \end{multline*}
1565: Using \eqref{formula with varphi} to
1566: express this as a sum of integrals over the (non-shifted) orthant,
1567: we get
1568: \begin{multline} \labell{remainder}
1569: R^{st}_k(\lambda_1,\dots,\lambda_n ; g) = \\
1570: \sum_{I \subsetneq \{ 1, \ldots, n \} } (-1)^{(k-1)(n-|I|)}
1571: \int\limits_{\bfO}
1572: \prod_{i \in I} \bfM^{k,\lambda_i}( - \deldel{t_i})
1573: \prod_{i \notin I} Q_{k,\lambda_i} (t_i)
1574: \prod_{i \notin I} \frac{\del^{k}}{\del t_i ^{k}}
1575: g(t_1,\ldots,t_n) dt_1 \cdots dt_n .
1576: \end{multline}
1577:
1578: We now perform in \eqref{integral over Oh} and \eqref{remainder}
1579: the change of variable given by the transformation
1580: \eqref{inverse transformation}.
1581: The integrals over $\bfO(h)$ and $\bfO$
1582: get replaced by integrals over $\bfC(h)$ and $\bfC$
1583: times the Jacobian factor $|\Gamma|$,
1584: the function $g(t)$ gets replaced by the function $f(x)$,
1585: and the partial derivative $\deldel{t_i}$ gets replaced by
1586: the directional derivative $D_{\alpha_i}$.
1587: Summing the expressions \eqref{integral over Oh} over $\gamma \in \Gamma$
1588: and dividing by $|\Gamma|$, we get, by \eqref{eqAA},
1589: \begin{equation} \labell{cat}
1590: \sum_{\bfC \cap \Z^n}{'} f = \sum_{\gamma \in \Gamma}
1591: \prod_{i=1}^n \bfM^{k,\lambda_{\gamma,i}}(\deldel{h_i})
1592: \left. \int_{\bfC(h)} f(x) dx \right|_{h=0} + R_k^{\bfC} (f)
1593: \end{equation}
1594: where
1595: $$ \lambda_{\gamma,j} := e^{2 \pi i \left< \gamma, \alpha_j \right> }$$
1596: and where
1597: \begin{multline} \labell{remainder C}
1598: R^{\bfC}_k(f) := \\
1599: \sum_{\gamma \in \Gamma}
1600: \sum_{I \subsetneq \{ 1, \ldots, n \} } (-1)^{(k-1)(n-|I|)}
1601: \int\limits_{\bfC}
1602: \prod_{i \in I} \bfM^{k,\lambda_{\gamma,i}} (-D_{\alpha_i})
1603: \prod_{i \not \in I}
1604: Q_{k,\lambda_{\gamma,i}} ( \left< u_i , x-v \right> )
1605: \prod_{i \not \in I} (D_{\alpha_i})^k f(x) dx .
1606: \end{multline}
1607:
1608: \bigskip
1609:
1610: Let $\Delta$ be a simple polytope, given by \eqref{Delta}.
1611: Choose a polarizing vector for $\Delta$ and let $\bfC_v^\sharp$
1612: denote the polarized tangent cones.
1613: The inward normals to the facets of $\bfC_v^\sharp$ are
1614: given by \eqref{u i v sharp}, the dual basis to these vectors
1615: is $\alpha_{i,v}^\sharp$, $i \in I_v$,
1616: and the roots of unity that appear in the Euler Maclaurin formula
1617: for $\bfC_v^\sharp$ are then
1618: \begin{equation} \labell{lambda sharp}
1619: \lambda_{\gamma,i,v}^\sharp
1620: = e^{2 \pi i \langle \gamma, \alpha_{i,v}^\sharp \rangle } = \begin{cases}
1621: \lambda_{\gamma,i,v} & \text{ if } \alpha_{i,v}^\sharp = \alpha_{i,v} \\
1622: \lambda_{\gamma,i,v}\inv & \text{ if } \alpha_{i,v}^\sharp = -\alpha_{i,v} .
1623: \end{cases}
1624: \end{equation}
1625: Also recall that the polar decomposition of $\Delta(h)$
1626: involves the dilated orthants
1627: $C_v^\sharp(h_{i_1,v}^\sharp, \ldots, h_{i_n,v}^\sharp )$
1628: where $h_{i,v}^\sharp$ are as in \eqref{h i v sharp}.
1629:
1630: Let $k \geq 1$ be an integer.
1631: For any compactly supported function $f$ on $\R^n$ of type $C^{nk}$,
1632: the polar decomposition of $\Delta(h)$ and the formula \eqref{cat}
1633: give
1634: \begin{multline} \labell{sum for simple}
1635: {\sum_{\Delta \cap \Z^n}}' f =
1636: \sum_v (-1)^{\# v} {\sum_{\bfC_v^\sharp \cap \Z^n}}' f \\
1637: = \sum_v (-1)^{\# v} \sum_{\gamma \in \Gamma_v}
1638: \prod_{i \in I_v = \{ i_1, \ldots, i_n \} }
1639: \bfM^{k,\lambda_{\gamma,i,v}^\sharp}(\deldel{h_{i,v}^\sharp}) \left.
1640: \int_{\bfC_v^\sharp(h_{i_1,v}^\sharp,\ldots,h_{i_n,v}^\sharp)}
1641: f(x) dx \right|_{h=0}
1642: + R_k^{\Delta}(f),
1643: \end{multline}
1644: where the remainder is given by
1645: \begin{equation} \labell{remdef}
1646: R^\Delta_k(f) := \sum_v (-1)^{\#v} R^{C_v^\sharp}_k(f) .
1647: \end{equation}
1648: Note that either $h_{i,v}^\sharp = h_{i}$ and
1649: $\lambda_{\gamma,i,v}^\sharp = \lambda_{\gamma,i,v}$,
1650: or $h_{i,v}^\sharp = -h_{i}$ and
1651: $\lambda_{\gamma,i,v}^\sharp = \lambda_{\gamma,i,v}\inv$.
1652: By the symmetry property \eqref{Dandlambdainverse},
1653: this gives
1654: $$ \bfM^{k,\lambda_{\gamma,i,v}^\sharp} (\deldel{h_{i,v}^\sharp})
1655: = \bfM^{k,\lambda_{\gamma,i,v}} (\deldel{h_i}) .$$
1656:
1657: For $j \not\in I_v$, because $\lambda_{\gamma,j,v} = 1$
1658: (see \eqref{lambda is one}),
1659: we have
1660: $\bfM^{k,\lambda_{\gamma,j,v} }(\deldel{h_j}) = 1 + $powers
1661: of $\deldel{h_j}$.
1662: Also, the cone
1663: $C_v^\sharp( h_{i_1,v}^\sharp, \ldots, h_{i_n,v}^\sharp)$
1664: is independent of $h_j$ for $j \not \in I_v$.
1665: Therefore, \eqref{sum for simple} is further equal to
1666: \begin{equation} \labell{eqA}
1667: \left. \sum_v (-1)^{\# v} \sum_{\gamma \in \Gamma_v}
1668: \prod_{j=1}^d \bfM^{k,\lambda_{\gamma,j,v}} (\deldel{h_j})
1669: \int\limits_{\bfC_v^\sharp(h_{i_1,v}^\sharp, \ldots, h_{i_n,v}^\sharp)}
1670: f(x) dx \right|_{h=0} + R^\Delta_k(f)
1671: \end{equation}
1672: where in the $v$'th summand $\{ i_1, \ldots, i_n \} = I_v$.
1673:
1674: Define
1675: \begin{equation}
1676: \labell{def of Mk gamma F}
1677: \bfM^k_{\gamma,F} = \prod_{j=1}^d \bfM^{k,\lambda_{\gamma,j,F}}(\deldel{h_j})
1678: \quad \text{ for } \gamma \in \Gamma_F .
1679: \end{equation}
1680: Then we have, by \eqref{horse},
1681: \begin{equation} \labell{Fv}
1682: \bfM^k_{\gamma,F} = \bfM^k_{\gamma,v} \quad \text{ whenever }
1683: \gamma \in \Gamma_F \text{ and } v \in F,
1684: \end{equation}
1685: where we identify $\gamma \in \Gamma_F$ with its image
1686: under the inclusion map $\Gamma_F \hookrightarrow \Gamma_v$.
1687:
1688: Then \eqref{eqA} is equal to
1689: \begin{eqnarray}
1690: \nonumber
1691: \left.\sum_v (-1)^{\# v} \sum_{\gamma \in \Gamma_v}
1692: \bfM_{\gamma,v}^k \int_{C_v^\sharp(h_{i_1,v}^\sharp,\ldots,h_{i_n,v}^\sharp)}
1693: f(x) dx \right|_{h=0} + R^\Delta_k(f) \\
1694: \labell{eqB}
1695: = \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1696: \bfM_{\gamma,F}^k \sum_{v \in F} (-1)^{\# v}
1697: \int_{C_v^\sharp(h_{i_1,v}^\sharp,\ldots,h_{i_n,v}^\sharp)}
1698: f(x) dx \right|_{h=0} + R^\Delta_k(f) ,
1699: \end{eqnarray}
1700: by \eqref{Gammas} and \eqref{Fv}.
1701: In the interior summation we may now add similar summands that correspond
1702: to $v \not \in F$.
1703: These summands make a zero contribution to (\ref{eqB})
1704: for the following
1705: reason. If $v \not \in F$ then there exists $i \in I_F \ssminus I_v$.
1706: Because $i \not \in I_v = \{ i_1, \ldots, i_n \}$, the cone
1707: $C_v^\sharp(h_{i_1,v}^\sharp , \ldots h_{i_n,v}^\sharp)$
1708: is independent of $h_i$.
1709: So it is enough to show that $\bfM_{\gamma,F}^k$ is a multiple
1710: of $\deldel{h_i}$.
1711: But because $\gamma \in \Gamma_F^\flat$ and $i \in I_F$, we have
1712: $\lambda_{\gamma,i,F} \neq 1$. (See Claim \ref{claim3}.)
1713: By Remark \ref{M vs L}, this
1714: implies that $\bfM^{k,\lambda_{\gamma,i,F}}(\deldel{h_i})$,
1715: which is one of the factors in $\bfM_{\gamma,F}^k$, is a multiple
1716: of $\deldel{h_i}$. Hence, \eqref{eqB} is equal to
1717: \begin{multline}
1718: \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1719: \bfM_{\gamma,F}^k \sum_{\text{all } v} (-1)^{\# v}
1720: \int_{C_v^\sharp( h_{i_1,v}^\sharp ,\ldots, h_{i_n,v}^\sharp )}
1721: f(x) dx \right|_{h=0} + R^\Delta_k(f) \\
1722: = \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1723: \bfM_{\gamma,F}^k \int_{\Delta(h)} f(x) dx \right|_{h=0}
1724: + R^\Delta_k(f) .\\
1725: \end{multline}
1726:
1727: We have therefore proved our main result:
1728:
1729: \begin{Theorem}\labell{main}
1730: Let $\Delta$ be a simple integral polytope in $\R^n$ and let
1731: $f\in C^{nk}_c(\R^n)$ be a compactly supported function on $\R^n$
1732: for $k \geq 1$. Choose a polarizing vector for $\Delta$. Then
1733: $${\sum_{\Delta \cap \Z^n}}' f = \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1734: \bfM_{\gamma,F}^k \int_{\Delta(h)} f(x) dx \right|_{h=0}
1735: + R^\Delta_k(f) $$
1736: where $\bfM_{\gamma,F}^k$ are differential operators
1737: defined in \eqref{def of Mk gamma F} and where the
1738: remainder $R^\Delta_k(f)$ is given by equation (\ref{remdef}).
1739: Moreover, the differential operators $\bfM_{\gamma,F}^k$
1740: are of order $\leq k$ in each of the variables $h_1,\ldots,h_d$.
1741: Also, the remainder is a sum of integrals
1742: over orthants of bounded periodic functions times various partial derivatives
1743: of $f$ of order no less than $k$ and no more than $kn$.
1744: Finally, this remainder is independent of the choice of polarization
1745: and is a distribution supported on the polytope $\Delta$.
1746: \end{Theorem}
1747:
1748: % =========================================================
1749: \section{Estimates on the remainder and
1750: Euler Maclaurin formulas for symbols and
1751: for polynomials}
1752: \labell{sec:estimates}
1753: % =========================================================
1754:
1755: We first recall a definition from the theory of partial
1756: differential equations. A smooth function $f \in C^\infty(\R^n)$
1757: is called a \textbf{symbol of order} $\mathbf{N}$
1758: if for every $n$-tuple of non-negative integers
1759: $a :=(a_1,\dots,a_n)$,
1760: there exists a constant $C_a$ such that
1761: $$|\partial_1^{a_1}\dots \partial_n^{a_n} f(x) | \leq C_a (1 + |x|)^{N - |a|}$$
1762: where $|a| = \sum_i a_i$. In particular, a polynomial of degree
1763: $N$ is a symbol of order $N$.
1764: Note that if $f$ is a symbol of order $N$ on $\R^n$ then its
1765: derivatives of order $a$ are in $L^1$ if $N < |a| - n$.
1766: In this section we will
1767: show that the Euler Maclaurin formula of Theorem \ref{main}
1768: can be extended to symbols, and gives rise in this way
1769: to an {\em exact} Euler Maclaurin formula for polynomials.
1770:
1771: To make further progress,
1772: we first require an estimate on the remainder term $R^\Delta_k(f)$.
1773: We recall (see Theorem \ref{main}) that this remainder
1774: can be expressed as a sum of integrals over orthants
1775: of bounded periodic functions time various partial derivatives of $f$
1776: of order no less than $k$ and no more than $kn$.
1777: Explicitly, from \eqref{remdef} and \eqref{remainder C} we get
1778: \begin{multline} \labell{def of RkDelta}
1779: R_k^\Delta (f) = \sum_v (-1)^{\# v} \sum_{\gamma \in \Gamma_v}
1780: \sum_{I \subsetneq I_v} (-1)^{(k-1)(n-|I|)} \\
1781: \int\limits_{\bfC_v^\#}
1782: \prod_{j \in I}
1783: \bfM^{k,\lambda_{\gamma,j,v}^\sharp} (-D_{\alpha_{j,v}^\sharp})
1784: \prod_{j \not \in I}
1785: Q_{k,\lambda_{\gamma,j,v}^\sharp}
1786: (\langle u_{j,v}^\sharp, x-v \rangle )
1787: \prod_{j \not \in I} (D_{\alpha_{j,v}^\sharp})^k
1788: f(x) dx .
1789: \end{multline}
1790:
1791: The functions $Q_{k,\lambda}$ are bounded and periodic,
1792: and $M^{k,\lambda}(-D_{\alpha_j})$ are differential operators
1793: of order $k$. In particular, each integrand in the formula
1794: for $R_k^\Delta(f)$ is dominated by a constant times
1795: $$ \sup_{ \{ j_1,\ldots,j_n \} } |\del_1^{j_1} \cdots \del_n^{j_n} f| $$
1796: where the supremum is taken over all $n$-tuples $\{ j_1,\ldots,j_n \}$
1797: with $k \leq j_1 + \ldots + j_n \leq nk$.
1798: Consequently, $R_k^\Delta(f)$ is well defined (by the same formula
1799: \eqref{def of RkDelta}) for any smooth function $f$ whose derivatives
1800: of order between $k$ and $nk$ are integrable on $\R^n$.
1801: In particular, it is defined when $f$ is a symbol of order less than $k-n$.
1802: Moreover, we get an estimate for the remainder:
1803:
1804: \begin{Proposition} \labell{polyest}
1805: The distribution $R_k^\Delta(\cdot)$ extends to symbols of order
1806: less than $k-n$, by the explicit expression \eqref{def of RkDelta} above,
1807: and satisfies an estimate of the form
1808: $$|R^\Delta_k(f)|
1809: \leq K(k,\Delta) \cdot {\rm sup}_{\{j_1,\dots,j_n\}}
1810: |\partial_1^{j_1}\dots\partial_n^{j_n} f|_{L_1(\R^n)},$$
1811: where the supremum is taken over all $n$-tuples $\{j_1,\cdots,j_n\}$
1812: with $k \leq j_1 + \cdots + j_n \leq nk.$
1813: \end{Proposition}
1814:
1815: By applying this estimate we will obtain the following Euler Maclaurin formula
1816: for symbols.
1817:
1818: \begin{Theorem}\labell{syms}
1819: Let $\Delta$ be a simple integral polytope in $\R^n$,
1820: let $f$ be a symbol of order $N$ on $\R^n$, and choose $k \geq N + n + 1$.
1821: Then
1822: %
1823: \begin{equation} \labell{EM formula for symbols}
1824: {\sum_{\Delta \cap \Z^n}}' f =
1825: \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1826: \bfM_{\gamma,F}^k \int_{\Delta(h)} f(x) dx \right|_{h=0}
1827: + R^\Delta_k(f)
1828: \end{equation}
1829: where $\bfM_{\gamma,F}^k$ are differential operators defined in
1830: \eqref{def of Mk gamma F}
1831: and where the remainder term $R_k^\Delta(f)$
1832: is defined in \eqref{def of RkDelta}.
1833: Moreover, the differential operators $\bfM_{\gamma,F}^k$
1834: are of order $\leq k$ in each of the variables $h_1,\ldots,h_d$.
1835: Also, the remainder $R_k^\Delta(f)$ satisfies
1836: the estimate of Proposition \ref{polyest}.
1837: \end{Theorem}
1838:
1839: In the case where $f$ is a polynomial, this formula gives rise to
1840: an {\em exact} Euler Maclaurin formula.
1841:
1842: \begin{Corollary}\labell{polys}
1843: Let $p$ be a polynomial on $\R^n,$ and choose $k \geq {\rm deg~} p+ n + 1.$
1844: Then
1845: %
1846: $${\sum_{\Delta \cap \Z^n}}' p =
1847: \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1848: \bfM_{\gamma,F}^k \int_{\Delta(h)} p(x) dx \right|_{h=0}
1849: .$$
1850: \end{Corollary}
1851:
1852: \begin{proof}[Proof of Theorem \ref{syms}:]
1853: Let $\chi$ be a smooth function on $\R^n$ which is equal to one
1854: on some open ball about the origin that contains $\Delta$
1855: and which is supported in some larger ball, say, of radius $R$.
1856: Define $\chi_\lambda(x) := \chi(x/\lambda)$ for all $\lambda \geq 1.$
1857: Then the function
1858: $$ f_\lambda:= f \chi_\lambda$$
1859: is a smooth compactly supported function on $\R^n$.
1860: We apply Theorem \ref{main} to obtain
1861: \begin{equation} \labell{EM f lambda}
1862: {\sum_{\Delta \cap \Z^n}}' f_\lambda =
1863: \left.\sum_F \sum_{\gamma \in \Gamma_F^\flat}
1864: \bfM_{\gamma,F}^k \int_{\Delta(h)} f_\lambda(x) dx \right|_{h=0}
1865: + R^\Delta_k(f_\lambda),
1866: \end{equation}
1867: which is valid for any $k \geq 1$.
1868:
1869: Since $f_\lambda$ equals $f$ on a neighborhood of $\Delta$
1870: if $\lambda \geq 1$, the left hand side and the first of the
1871: two summands on the right hand side in \eqref{EM f lambda}
1872: are equal to the corresponding terms in \eqref{EM formula for symbols}.
1873: Thus, to deduce \eqref{EM formula for symbols} from \eqref{EM f lambda},
1874: it suffices to prove the following claim:
1875: \begin{equation} \labell{limit claim}
1876: \lim_{\lambda\to \infty}R^\Delta_k(f_\lambda) = R^\Delta_k(f)
1877: \quad \text{ if } \quad k \geq N + n + 1 .
1878: \end{equation}
1879:
1880: To prove this claim, we apply the estimate in Proposition \ref{polyest}
1881: to the difference $R_k^\Delta(f) - R_k^\Delta(f_\lambda)
1882: = R_k^\Delta(f(1-\chi_\lambda))$. We expand each of the derivatives
1883: appearing in the estimate into a finite sum of products of derivatives
1884: of $f$ and derivatives of $1-\chi_\lambda$.
1885: The leading term has the form $g (1 - \chi_\lambda)$
1886: where $g$ is a derivative of $f$ of order $q$ with $q \geq k$.
1887: Because $f$ is a symbol of order $N$, the function $g$ is dominated
1888: by a constant times the function $x \mapsto (1+|x|)^{N-q}$.
1889: Because $q \geq k \geq N+n+1$, the function $g$ is in $L_1$.
1890: It follows that the $L_1$ norm of $g (1-\chi_\lambda)$
1891: converges to zero as $\lambda \to \infty$.
1892: Each of the remaining terms in the expansion has the form
1893: \begin{equation} \labell{the term}
1894: \lambda^{-s} g(\cdot) \tilde{\chi} (\cdot/\lambda)
1895: \end{equation}
1896: where $g$ is a derivative of $f$ of order $q$
1897: and $\tilde{\chi}$ is a derivative of $1-\chi$ of order $s$,
1898: and where $s \geq 1$ and $s+q \geq k$.
1899: The function $g$ is dominated by a constant times the
1900: function $x \mapsto (1+|x|)^{N-q}$; the function $\tilde{\chi}$ is bounded;
1901: the function $\tilde{\chi} (\cdot/\lambda)$ is supported on the ball
1902: of radius $\lambda R$ about the origin.
1903: Hence, the $L_1$ norm of the product $g(\cdot) \tilde{\chi} (\cdot/\lambda)$
1904: is bounded by a constant times $\lambda^{N-q+n}$ if $N-q+n$ is non-negative
1905: and by a constant otherwise.
1906: Since $s \geq 1$ and $s+q \geq k \geq N+n+1$, the $L_1$ norm
1907: of the term \eqref{the term}
1908: is bounded by some constant multiple of $\lambda^{-1}$.
1909: Letting $\lambda \to \infty$, we see that each of the terms in the estimate
1910: approaches zero as $\lambda \to \infty$.
1911: This implies \eqref{limit claim}.
1912: Theorem \ref{syms} follows.
1913: \end{proof}
1914:
1915: \begin{thebibliography}{GDR}
1916:
1917: \bibitem[BP]{BP}
1918: A.\ Barvinok and J.\ E.\ Pommersheim,
1919: \emph{An algorithmic theory of lattice points in polyhedra},
1920: New perspectives in algebraic combinatorics (Berkeley, CA, 1996--97),
1921: 91--147, Math.\ Sci.\ Res.\ Inst.\ Publ., \textbf{38},
1922: Cambridge Univ.\ Press, Cambridge, 1999.
1923:
1924: \bibitem[BDR]{BDR}
1925: M.~Beck, R.~Diaz, and S.~Robins,
1926: \emph{The Frobenius problem, rational polytopes, and Fourier-Dedekind sums},
1927: J.~Number Theory \textbf{96} (2002), no.~1, 1--21.
1928:
1929: \bibitem[Bo]{Bo}
1930: N.\ Bourbaki, \emph{Functions d'Une Variable R\'{e}ele},
1931: Chapitre VI: D\'{e}veloppments Tayloriens G\'{e}neralis\'{e}s,
1932: Formule Summatoire d'Euler-Maclaurin (1951).
1933:
1934: \bibitem[BV]{BV}
1935: M.~Brion and M.~Vergne, \emph{Lattice points in simple polytopes},
1936: Jour.\ Amer.\ Math.\ Soc.\ \textbf{10} (1997), 371--392.
1937:
1938: \bibitem[CS1]{CS:bulletin}
1939: S.~E.~Cappell and J.~L.~Shaneson, \emph{Genera of algebraic varieties
1940: and counting lattice points}, Bull.\ A.~M.~S.\ \textbf{30} (1994), 62--69.
1941:
1942: \bibitem[CS2]{CS:EM}
1943: S.~E.~Cappell and J.~L.~Shaneson, \emph{Euler-Maclaurin expansions
1944: for lattices above dimension one},
1945: C.\ R.\ Acad.\ Sci.\ Paris Sér.\ I Math.\ \textbf{321} (1995), 885--890.
1946:
1947: \bibitem[CS3]{CS:private}
1948: S.~E.~Cappell and J.~L.~Shaneson, unpublished.
1949:
1950: \bibitem[Da]{Da}
1951: V.~I.~Danilov, \emph{The geometry of toric varieties},
1952: Russ.\ Math.\ Surv.\ \textbf{33} (1978) no.~2, 97--154.
1953:
1954: \bibitem[DR]{DR} R.\ Diaz and S.\ Robins, \emph{The Ehrhart Polynomial
1955: of a Lattice Polytope}, Ann.\ Math.\ (2) \textbf{145} (1997),
1956: no.~3, 503--518,
1957: and \emph{Erratum: ``The Ehrhart polynomial of a lattice polytope"},
1958: Ann.\ Math.\ (2) \textbf{146} (1997), no.~1, 237.
1959:
1960: \bibitem[Eu]{Eu}
1961: L.~Euler, Commentarii Acad.\ Petrop.\ \textbf{6} (1732--3)
1962: and \textbf{8} (1736).
1963:
1964: % \bibitem[GV]{GV}
1965: % A.~N.~Varchenko and I.~M.~Gelfand, \emph{Heaviside functions of
1966: % a configuration of hyperplanes.} (Russian), Funktsional.\ Anal.\
1967: % i Prilozhen.\ \textbf{21} (1987), no.~4, 1--18, 96.
1968: % English translation: Functional Anal.\ Appl.\ \textbf{21} (1987),
1969: % no.~4, 255--270.
1970:
1971: \bibitem[Gi]{Gibbs} J. W. Gibbs. ``Fourier Series.'' Nature 59, 200 and 606,
1972: 1899.
1973:
1974: \bibitem[Gu1]{Gu:book}
1975: V.~Guillemin, \emph{Moment maps and combinatorial invariants
1976: of Hamiltonian $T^n$-spaces}, Progress in Math.\ \textbf{122},
1977: Birkhauser, 1994.
1978:
1979: \bibitem[Gu2]{Gu}
1980: V.~Guillemin, \emph{Riemann-Roch for toric orbifolds}, J.\ Diff.\ Geom.\
1981: \textbf{45} (1997), 53--73.
1982:
1983: \bibitem[GS]{GS}
1984: V.~Guillemin and S.~Sternberg, \emph{Geometric quantization
1985: and multiplicities of group representations}, Invent.\ Math.\ \textbf{67}
1986: (1982), 515--538.
1987:
1988: \bibitem[Har]{H}
1989: G.\ H.\ Hardy, \emph{Divergent Series},
1990: Second Edition, 1991 (unaltered), Chelsea Pub.\ Col, New-York, N.Y.
1991:
1992: \bibitem[Hat]{Ha}
1993: A.~Hattori and M.~Masuda, \emph{Theory of multi-fans},
1994: Osaka J.~Math.\ \textbf{40} (2003), 1--68.
1995:
1996: \bibitem[Ho]{horm} L. Hormander. Lectures on nonlinear hyperbolic
1997: differential equations. New York: Springer, 1997.
1998:
1999: \bibitem[KK]{KK}
2000: J.~M.~Kantor and A.~G.~Khovanskii,
2001: \emph{Une application du th\'eor\`eme de Riemann-Roch combinatoire
2002: au polyn$\hat{o}$me d'Ehrhart des polytopes entiers de $R\sp d$},
2003: C.\ R.\ Acad.\ Sci\. Paris S\'er.\ I Math.\ \textbf{317} (1993),
2004: no.\ 5, 501--507.
2005:
2006: \bibitem[KSW]{PNAS}
2007: Y. Karshon, S. Sternberg, and J. Weitsman.
2008: \emph{The Euler-Maclaurin formula for simple integral polytopes},
2009: Proc.\ Nat.\ Acad.\ Sci.\ \textbf{100} no.~2 (2003), 426--433.
2010:
2011: \bibitem[Kh1]{Kh1} A.~G.~Khovanski,
2012: \emph{Newton polyhedra and toroidal varieties},
2013: Funktsional.\ Anal.\ i Prilozhen.\ \textbf{11} (1977), no.~4, 56--64.
2014: English tranlation:
2015: Func.\ Anal.\ Appl.\ \textbf{11} (1977), no.~4, 289--296 (1978).
2016:
2017: \bibitem[Kh2]{Kh2} A.~G.~Khovanski,
2018: \emph{Newton polyhedra and the genus of complete intersections},
2019: Funktsional.\ Anal.\ i Prilozhen.\ \textbf{12} (1978), no.~1, 51--61.
2020: English tranlation:
2021: Func.\ Anal.\ Appl.\ \textbf{12} (1978), no.~1, 38--46 (1978).
2022:
2023: \bibitem[KP1]{KP1} A.~G.~Khovanskii and A.~V.~Pukhlikov,
2024: \emph{Finitely additive measures of virtual polytopes},
2025: Algebra and Analysis \textbf{4} (1992), 161--185;
2026: translation in St.\ Petersburg Math.\ J.\ \textbf{4} (1993),
2027: no.~2, 337--356.
2028:
2029: \bibitem[KP2]{KP2} A.~G.~Khovanskii and A.~V.~Pukhlikov,
2030: \emph{The Riemann-Roch theorem for integrals and sums of quasipolynomials
2031: on virtual polytopes}, Algebra and Analysis \textbf{4} (1992), 188--216,
2032: translation in St.\ Petersburg Math.\ J.\ (1993), no.~4, 789--812.
2033:
2034: \bibitem[Kn]{Kn} K.\ Knopp, \emph{Theory and application of infinite series},
2035: Translated from the
2036: Second German Edition by Miss R.~C.~Young L.\`{e}s Sc., Blackie and Son
2037: Limited, London and Glasgow (1928), Chapter XIV.
2038:
2039: \bibitem[L]{lawrence}
2040: J.\ Lawrence, \emph{Polytope volume computation},
2041: Math.\ Comp.\ \textbf{57} (1991), no.\ 195, 259--271.
2042:
2043: \bibitem[Ma]{Ma}
2044: C.~Maclaurin, \emph{A treatise of Fluxions}, Edinburgh (1742).
2045:
2046: \bibitem[Md]{Md} L.~J.~Mordell, \emph{Lattice points in a tetrahedron
2047: and generalized Dedekind sums}, J.~Indian Math.\ Soc.\ (N.S.)
2048: \textbf{15} (1951), 41--46.
2049:
2050: \bibitem[Mo]{Mo}
2051: R.\ Morelli, \emph{Pick's theorem and the Todd class of a toric variety},
2052: Adv.\ Math.\ \textbf{100} (1993), no.\ 2, 183--231.
2053:
2054: \bibitem[Od]{Oda}
2055: T.\ Oda, \emph{Convex bodies and algebraic geometry}, Ergebnisse der
2056: Mathematik, Springer, 1988.
2057:
2058: \bibitem[Pi]{pick} G. Pick, \emph{Geometrisches zur Zahlentheorie},
2059: Sitzenber.~Lotos (Prague) \textbf{19} (1899), 311--319.
2060:
2061: \bibitem[Poi]{Poi}
2062: S.~D.~Poisson, M\'{e}moires Acad.\ sciences Inst.\ France \textbf{6} (1823).
2063:
2064: \bibitem[Pom]{Po}
2065: J.\ E.\ Pommersheim,
2066: \emph{Toric varieties, lattice points and Dedekind sums},
2067: Math.\ Ann.\ \textbf{295} (1993), 1--24.
2068:
2069: \bibitem[S]{S} J. Shaneson, \emph{Characteristic Classes, Lattice
2070: Points, and Euler-MacLaurin Formulae}, Proceedings of the International
2071: Congress of Mathematicians, Zurich, 1994. Basel: Birkhauser Verlag,
2072: 1995.
2073:
2074: \bibitem[T]{T} R.\ M.\ Trigub,
2075: \emph{A Generalization of the Euler-Maclaurin Formula},
2076: Mathematical Notes \textbf{61}, no.~2, 1997, p.~253--257,
2077: translated from Matematicheskie Zametki \textbf{61}, no.~2,
2078: p.~312--316.
2079:
2080: \bibitem[V]{V} A.~N.~Varchenko,
2081: \emph{Combinatorics and topology of the arrangement of affine hyperplanes
2082: in the real space.} (Russian)
2083: Funktsional.\ Anal.\ i Prilozhen.\ \textbf{21} (1987), no.~1, 11--22.
2084: English translation: Functional Anal.\ Appl.\ \textbf{21} (1987),
2085: no.~1, 9--19.
2086:
2087: \bibitem[W]{Wi} W. Wirtinger \emph{Einige Anwendungen der
2088: Euler-Maclaurin'schen Summenformel, insbesondere auf eine Aufgabe
2089: von Abel} Acta. Math. \textbf{26}
2090: (1902) 255- 271.
2091:
2092:
2093: \end{thebibliography}
2094:
2095: \end{document}
2096:
2097: Victor's paper: http://www.intlpress.com/JDG/archive/vol.45/Index.html
2098:
2099: Brion-Vergne's: http://www.ams.org/jams/1997-10-02/
2100:
2101: