1: \documentclass[11pt]{amsart}
2: \usepackage{vmargin}
3: \usepackage{harvard}
4: \usepackage{psfig}
5: \usepackage{graphics}
6:
7: %%%% theorem-like environments %%%%
8:
9: \newtheorem{theorem}{Theorem}[section]
10: \newtheorem*{untheorem}{Theorem}
11: \newtheorem{lemma}[theorem]{Lemma}
12: \newtheorem*{unlemma}{Lemma}
13: \newtheorem{proposition}[theorem]{Proposition}
14: \newtheorem*{unproposition}{Proposition}
15: \newtheorem{corollary}[theorem]{Corollary}
16: \newtheorem*{uncorollary}{Corollary}
17:
18: \theoremstyle{definition}
19: \newtheorem{definition}[theorem]{Definition}
20: \newtheorem{example}[theorem]{Example}
21: \newtheorem{notation}[theorem]{Notation}
22:
23: \theoremstyle{remark}
24: \newtheorem{remark}[theorem]{Remark}
25: \newtheorem*{unremark}{Remark}
26: \newtheorem*{claim}{Claim}
27:
28: %%%%roman letters in enumerate%%%%
29:
30: \makeatletter
31: \def\romenumi{%
32: \def\theenumi{\roman{enumi}}%
33: \def\p@enumi{\theenumi}%
34: \def\labelenumi{(\@roman\c@enumi)}}
35: \makeatother
36:
37: %%%%"no proof" symbol%%%%
38:
39: \newcommand{\noproof}{\hfill\qedsymbol}
40:
41: %%%%new operators%%%%
42:
43: \DeclareMathOperator{\supp}{supp}
44: \DeclareMathOperator{\re}{Re}
45: \DeclareMathOperator{\im}{Im}
46: \DeclareMathOperator{\Arg}{Arg}
47: \DeclareMathOperator{\Res}{Res}
48: \DeclareMathOperator{\hess}{\mathcal{H}}
49: \DeclareMathOperator{\proj}{proj}
50: \DeclareMathOperator{\rk}{rank}
51: \DeclareMathOperator{\cone}{\bb{\kappa}}
52: \DeclareMathOperator{\dir}{\b{dir}}
53: \DeclareMathOperator{\grad}{\nabla}
54: \DeclareMathOperator{\sign}{sgn}
55:
56: %%%%formatting%%%%%
57:
58: \numberwithin{equation}{section}
59:
60: \setpapersize{USletter}
61: \setmarginsrb{20mm}{20mm}{20mm}{20mm}{12pt}{10mm}{12pt}{10mm}
62:
63: %%%% concepts whose notation may be changed %%%%%%%
64:
65: \def\sing{\mathcal{V}}
66: \def\torus{T}
67: \def\disk{D}
68: \def\numer{{G}}
69: \def\denom{{H}}
70: \def\mult{{\psi}}
71: \def\simp{{\Delta}}
72: \def\meas{\mu}
73: \def\direc{{\bf dir}}
74: %\def\direc{{\bb{\delta}}}
75: \def\dom{{\mathcal D}}
76: \def\logdom{{\rm log} \dom}
77: \def\rate{\gamma}
78: \def\hyper{\mathcal{H}}
79: \def\nbd{{\mathcal N}}
80: \def\one{{\bf 1}}
81: \def\zero{{\bf 0}}
82: \def\GF{{\mbox{GF-sequence}}}
83: \def\cio{C^\infty_0}
84: \def\finv{\eta}
85: \def\X{\Xi}
86:
87: %%%%% abbreviations %%%%%
88:
89: \def\wlog{\mbox{without loss of generality}}
90: \def\b#1{\mathbf{#1}}
91: \def\bb#1{\boldsymbol{#1}}
92: \def\w#1{\widehat{#1}}
93: \def\wh{\w{\b{w}}}
94: \def\zh{\w{\b{z}}}
95: \def\thh{\w{\bf \theta}}
96: \def\zt{{\tilde{z}}}
97: \def\qt{{\tilde{q}}}
98: \def\bt{{\tilde{b}}}
99: \def\gt{{\tilde{g}}}
100: \def\ft{{\tilde{f}}}
101: \def\psit{{\tilde{\psi}}}
102: \def\R{{\mathbb{R}}}
103: \def\CC{{\mathbb{C}}}
104: \def\RP{{\mathbb{RP}}}
105: \def\N{{\mathbb{N}}}
106: \def\Z{{\mathbb{Z}}}
107: \def\A{{\mathcal A}}
108: \def\S{{\mathcal S}}
109: \def\C{{\mathcal C}}
110: \def\B{{\mathcal B}}
111: \def\ee{\varepsilon}
112: \def\dd{\delta}
113:
114: %%%%%%%%%%%%%%END OF TOPMATTER%%%%%%%%%%%%%%%%%%%%%%
115:
116: \begin{document}
117:
118: \title[Asymptotics of multivariate sequences I]{Asymptotics of multivariate sequences, part I: smooth points of the singular variety}
119: \author{Robin Pemantle}
120: \address{Department of Mathematics, Ohio State University, Columbus OH 43210}
121: \email{pemantle@math.ohio-state.edu}
122: \thanks{Research supported in part by NSF grant DMS-9803249}
123: \author{Mark C. Wilson}
124: \address{Department of Mathematical Sciences, University of Montana,
125: Missoula, MT 59812}
126: \email{wilsonm@member.ams.org}
127:
128: \subjclass{Primary 05A16. Secondary 32A20, 41A60.}
129:
130: \keywords{Generating function, recurrence, linear difference equation,
131: contour methods, central limit, oscillating integral, Cauchy integral formula.}
132:
133: \begin{abstract}Given a multivariate generating function
134: $F(z_1 , \ldots , z_d) = \sum a_{r_1 , \ldots , r_d} z_1^{r_1}
135: \cdots z_d^{r_d}$, we determine asymptotics for the coefficients.
136: Our approach is to use Cauchy's integral formula near singular
137: points of $F$, resulting in a tractable oscillating integral. This
138: paper treats the case where the singular point of $F$ is a smooth
139: point of a surface of poles. Companion papers will treat singular
140: points of $F$ where the local geometry is more complicated, and
141: for which other methods of analysis are not known.
142: \end{abstract}
143:
144: \maketitle
145:
146: \section{Introduction}
147:
148: The generating function $F(z) := \sum_{r=0}^\infty a_r z^r$ for
149: the sequence $a_0 , a_1 , a_2 , \ldots$ is one of the most useful
150: constructions in combinatorics. If the function $F$ has a simple
151: description, it is usually not too hard to obtain $F$ as a formal
152: power series once one understands a recursive or combinatorial
153: description of the numbers $\{ a_r \}$. One may then analyze the
154: analytic properties of $F$ in order to obtain asymptotic
155: information about the sequence $\{ a_r \}$. While still part art
156: and part science, this latter analytic step has become quite
157: systematized. \citeasnoun{stanley;enumerative-combinatorics1} in
158: his introduction to enumerative combinatorics gives the example of
159: the function $F(z) = \exp (z + \frac{z^2}{2})$, from which he says
160: ``it is routine (for someone sufficiently versed in complex
161: variable theory) to obtain the asymptotic formula $a_r = 2^{-1/2}
162: r^{r/2} e^{-r/2 + \sqrt{r} - 1/4}$.'' Routine, in this case, means
163: a single application of the saddle point method. When $F$ has
164: singularities in the complex plane, the analysis is often more
165: direct: the location of the singularities and the behavior of $F$
166: near these determine almost algorithmically the asymptotic
167: behavior of the sequence $\{ a_r
168: \}$. For those not sufficiently versed in complex variable
169: theory, two useful sources are \citeasnoun{henrici;complex-analysis2} and
170: \citeasnoun{odlyzko;asymptotic-methods}. The transfer theorems of
171: \citeasnoun{flajolet-odlyzko;singularity-analysis}
172: encapsulate much of this knowledge in a very useful way; see also
173: \citeasnoun{wilf;GFology} for an elementary introduction.
174:
175: When the sequence $a_0 , a_1 , a_2 , \ldots$ is replaced by a
176: multidimensional array $\{ a_{r_1 , \ldots , r_d} \}$, things
177: become much more hit and miss. Let us use boldface to denote
178: vectors in $\CC^d$ or $\N^d$, and use multi-index notation, so
179: that $a_\b{r}$ denotes the multi-index $a_{r_1 , \ldots , r_d}$ and
180: $\b{z}^\b{r}$ denotes the product $z_1^{r_1} \cdots z_d^{r_d}$ which
181: we will sometimes write in expanded form for clarity. The
182: generating function $F : \CC^d \to \CC$ is defined analogously to
183: the one-dimensional generating function by
184: $$F(\b{z}) = \sum_{\b{r} \in \N^d} a_\b{r} \b{z}^\b{r} .$$
185: Surprisingly, techniques for extracting asymptotics of $\{ a_\b{r}
186: \}$ from the analytic properties of $F$ were, until recently,
187: almost entirely missing. In a survey of asymptotic methods,
188: \citeasnoun{bender;asymptotic-methods} says:
189: \begin{quote}
190: Practically nothing is known about asymptotics for recursions in
191: two variables even when a generating function is available.
192: Techniques for obtaining asymptotics from bivariate generating
193: functions would be quite useful.
194: \end{quote}
195: In the intervening 25 years, some results have appeared,
196: addressing chiefly the case where the array $\{ a_\b{r} \}$ obeys a
197: central limit theorem. Common to all of these is the following
198: method. Treat $\{ a_\b{r} \}$ as a sequence of
199: $(d-1)$-dimensional arrays indexed by $r_d$; show that the $n^{th}$
200: $(d-1)$-dimensional generating function is roughly the $n^{th}$
201: power of a given function; use this approximation to invert the
202: characteristic function and obtain a Central Limit Theorem. We
203: refer to these methods as $\GF$ methods. The other body of work
204: on multivariate sequences, which we will call the diagonal method,
205: is based on algebraic extraction of the diagonal, as found in
206: \citeasnoun{hautus-klarner;diagonal} (
207: see also \citeasnoun{furstenberg;diagonal} and later \citeasnoun{lipshitz;diagonal} for an algebraic description of the scope of this method; variants are described in
208: \citeasnoun{stanley;enumerative-combinatorics2} and \citeasnoun{pippenger;equicolourable}).
209:
210: The most fundamental $\GF$ result is probably
211: \citeasnoun{bender-richmond;limit-multivariate2},
212: with extensions appearing in later work of the same
213: authors. \citeasnoun{flajolet-sedgewick;multivariate} present a version of the
214: same idea which holds in much greater generality.
215: \citeasnoun{gao-richmond;limit-multivariate4} go beyond the central limit case,
216: using the transfer theorems of
217: \citeasnoun{flajolet-odlyzko;singularity-analysis} to handle
218: functions that are products of powers with powers of logs. Recent
219: work of Bender and Richmond
220: \cite{bender-richmond;admissible,bender-richmond;products}
221: extends the applicability of the central limit results to many
222: problems of combinatorial interest; see also
223: \cite{hwang;stirling,hwang;convergence-rates}, where
224: more precise asymptotics are given, and
225: \citeasnoun{hwang;local-limit-deviations}, which extends some results
226: to the combinatorial schemes of
227: \citeasnoun{flajolet-soria;gaussian-exponential}. This does not
228: exhaust the recent work on the problem of multivariable
229: coefficient extraction, but does circumscribe it.
230:
231: The present paper, together with forthcoming companion papers,
232: takes aim at a large class of multivariable coefficient extraction
233: problems, for which a fair amount of information can be read off
234: in a systematic way. An ultimate goal (not our only goal) is to
235: systematize the extraction of multivariate asymptotics
236: sufficiently that it may be automated, say in Maple. Everything
237: we do, we do with complex contour integration. In this regard,
238: our methods are most similar to those of
239: \citeasnoun{bertozzi-mckenna;queueing}, who, as we do, provide a general
240: framework for harnessing the multivariable theory of residues for
241: exact and series computation of coefficients. A more detailed
242: description of our method will be given in
243: Section~\ref{ss:results}, but here is an outline.
244:
245: \begin{quote}
246: ~~
247:
248: (1) Use the multidimensional Cauchy integral formula to represent
249: $a_{\b{r}}$ as an integral over a $d$-dimensional torus inside
250: $\CC^d$.
251:
252: (2) Expand the surface of integration across a point $\b{z}$ where
253: $F$ is singular, and use the residue theorem to represent
254: $a_\b{r}$ as a $(d-1)$-dimensional integral of one-variable residues.
255: The choice of $\b{z}$ determines the directions in which asymptotics
256: may be computed.
257:
258: (3) Put this in the form of an integral $\int \exp (\lambda
259: f(\b{z})) \psi (\b{z}) \, d\b{z}$ for which the large-$\lambda$
260: asymptotics can be read off from the theory of oscillating
261: integrals.
262: \end{quote}
263:
264: \begin{figure}
265: \centerline{\psfig{figure=taxonomy.eps,width=6in}}
266: \vspace{10pt}
267: \caption{Classification of generating functions}
268: \label{fig:taxonomy}
269: \end{figure}
270:
271: In the rest of this introductory section, we describe the scope of
272: our methods. Figure~\ref{fig:taxonomy} depicts a classification of generating functions and illustrates the remainder of this paragraph. If a
273: formal power series is nowhere convergent, analytic methods are
274: useless. Among those power series converging in some neighborhood
275: of the origin, there are three possibilities: a function may be
276: entire, may have singularities around which analytic continuations
277: exist, or it may be defined only on some bounded subset of
278: $\CC^d$. Our methods are tailored to the second class. The third
279: class, although in some sense generic, seldom arises in any
280: problem for which the generating function may be effectively
281: described. Incomplete asymptotic information is available via
282: Darboux' method; details of this method in the univariate case are
283: given in \citeasnoun{henrici;complex-analysis2} and
284: \citeasnoun{odlyzko;asymptotic-methods}. The first class can
285: and does arise frequently. Our methods are simply not equipped to
286: handle entire functions, and systematizing the asymptotic analysis
287: of coefficients of entire generating functions remains an
288: important open problem.
289:
290: For the remainder of this paper, we will assume that the formal
291: power series $F$ converges in a neighborhood of the origin and may
292: be analytically continued everywhere except a set $\sing$ of complex
293: dimension $d-1$ which we call the {\em singular variety}. The
294: point $\b{z}$ in step~2 is an element of $\sing$, and the behavior of
295: $\sing$ near $\b{z}$ greatly affects the subsequent analysis in step~3.
296: This paper addresses the case where $\b{z}$ is a smooth point of
297: $\sing$ at which $F$ has a pole. The forthcoming companion papers
298: will address cases where $\b{z}$ is a multiple point
299: or a cone point. We do not know whether cases
300: where $\b{z}$ is a cusp of $\sing$ arise, but if so, the subsequent
301: analysis has mostly been carried out in the work of
302: \citeasnoun{arnold-varchenko;asymptotics-integrals}.
303:
304: The chief purpose of this study is to give a solution to the
305: problem of asymptotic evaluation of coefficients that is as
306: general as possible. An important part of this is re-derivation
307: in a general setting of results obtainable via $\GF$ or {\em ad
308: hoc} methods. We show in Section~\ref{ss:classify} how unifying
309: these results allows us to show that our method successfully finds
310: asymptotics for every function in a certain large class. Familiar
311: examples from this class include: lattice path counting, various
312: known generating functions for polyominos and stacked balls,
313: enumeration of Catalan trees by number of components or
314: surjections by image cardinality (see
315: \citeasnoun{flajolet-sedgewick;multivariate},
316: stopping times for certain random walks
317: (see \citeasnoun{larsen-lyons;coalescing}), as well as the examples
318: given in the $\GF$ papers of
319: \citeasnoun{bender;central-local} and
320: \citeasnoun{bender-richmond;limit-multivariate2}: ordered set partitions
321: enumerated by number of blocks, permutations enumerated by rises,
322: and Tutte polynomials of recursive sequences of graphs.
323:
324: Nevertheless, our pursuit of this problem was also motivated by
325: some specific applications which we mention briefly now and
326: discuss more thoroughly later. These are cases where known
327: methods do not suffice to obtain complete asymptotic information.
328: There is a class of tiling enumeration problems for which an
329: explicit three variable rational generating function may be
330: obtained. This class includes the Aztec Diamond domino tilings of
331: \citeasnoun{cohn-elkies-propp;aztec}. Asymptotics in the so-called {\em
332: region of fixation} are obtained from analysis of the smooth
333: points of $\sing$ (Theorem~\ref{th:higher d} below), while
334: asymptotics in the region of positive entropy are derived from
335: analysis of the cone point. \citeasnoun{cohn-pemantle;fixation} applies a
336: cone point analysis to a tiling enumeration problem for which the
337: only previous results are some pictures via simulation
338: (http://www.math.harvard.edu/\~{}cohn/picture.gif) . Another
339: motivation has been to solve the general multivariable linear
340: recursion. Depending on whether one allows forward recursion in
341: some of the variables, one obtains either rational or algebraic
342: generating functions. The general rational function may have any
343: of the types of singularities mentioned above: smooth points,
344: nodes, cones, cusps, branchpoints, etc. Even the simple rational
345: generating function $1 / (3 - 3z - w + z^2)$ of
346: Example~\ref{ex:cubic} requires two separate analyses in order to
347: get asymptotics in all directions. We will see that
348: Theorem~\ref{th:simple} gives asymptotics in one region, while
349: Theorem~\ref{th:2d} is required for other directions.
350:
351: Asymptotics derived near smooth pole points nearly always exhibit
352: central limit behavior. Smooth pole points are the topic of this
353: first paper, and are exactly the case to which existing methods
354: may apply. While one function of this paper is to lay foundations
355: for the cases in which the singularity is more complicated, there
356: are several ways in which it improves upon available analyses of
357: the smooth case.
358:
359: First, most of the existing results assume that the singular point
360: $\b{z} \in \sing$ has positive real coordinates, and that it is
361: strictly minimal in a sense defined in the next section. This
362: assumption often holds when the coefficients $\{ a_\b{r} \}$ are
363: nonnegative reals, though it will fail if, for example, there is
364: any periodicity. The assumption always fails when the
365: coefficients $\{ a_\b{r} \}$ have mixed signs, as is the case for
366: example with the generating functions $(1 - zw)/(1 - 2zw +w^2)$
367: and $1 / (1 - 2zw + w^2)$ for the Chebyshev polynomials of the
368: first and second kinds \cite[page 50]{comtet;advanced}. $\GF$ methods may
369: be adapted to some of these situations. Indeed, the presentation
370: of these methods by \citeasnoun[Theorem~9.7]{flajolet-sedgewick;multivariate}
371: accomplishes this adaptation in great generality. But certainly
372: there are cases such as the rational generating function
373: $1 / (1 - z - w + \beta zw)$, where the points $\b{z}$ with given
374: moduli form a continuum and standard $\GF$ methods are not
375: sufficient.
376:
377: Second, our methods obtain automatically a full asymptotic
378: expansion of $a_{r_1 , \ldots , r_d}$ in decreasing powers of the
379: indices $r_j$. This is certainly not inherent in the existing
380: results, whose relatively short proofs involve inversion of the
381: characteristic function (see however
382: \citeasnoun{hwang;stirling} and \citeasnoun {hwang;central-limit-deviations} for
383: something in this direction). The expansion to $n$ terms is
384: completely effective in terms of the first $n$ partial derivatives
385: of $1/F$ at $\b{z}$, as is the error bound.
386:
387: Third, these results explicitly cover the case where the pole at
388: $\b{z}$ has order greater than 1. The behavior in this case is not
389: according to the central limit theorem. The only existing work
390: addressing this case is \citeasnoun{gao-richmond;limit-multivariate4},
391: and they require
392: nonnegativity assumptions, as mentioned above. In the case where
393: $F = G^k$ is an exact power, one could attempt first to solve the
394: problem for $G$ and then to take the $k$-fold convolution. This
395: is much harder than the present approach, as may be seen by the
396: rather involved computation in \citeasnoun{cohn-elkies-propp;aztec}.
397:
398: Fourth, the potential for increasing the scope to new applications
399: seems greater for contour methods than for $\GF$ methods. The
400: contour method reduces the asymptotic problem to the problem of an
401: oscillating integral near a singularity, which can almost
402: certainly be done. By contrast, the $\GF$ method requires first
403: an understanding of the sequence of $(d-1)$-dimensional generating
404: functions arising from the given $d$-dimensional generating
405: function, and then another result in order to transfer this
406: information to asymptotics of the coefficients $a_\b{r}$.
407:
408: Fifth, although our results in the case of smooth pole points are
409: often similar to those obtained by $\GF$ methods, our hypotheses
410: are quite different. In Section~\ref{ss:classify} we show how our
411: hypotheses may be universally established for functions that
412: generate nonnegative values and are meromorphic through their
413: domain of convergence.
414:
415: Finally, we compare our method to recent results from the diagonal
416: method. It is known \cite{lipshitz;diagonal} that the diagonal
417: sequence
418: $a_{n,n,\ldots ,n}$ of a multivariate sequence with rational
419: generating function has a generating function satisfying a linear
420: differential equation over rational functions. Much is known
421: about how to compute this equation (see for example
422: \citeasnoun{chyzak-salvy;symbolic-integration}).
423: If one wants asymptotics on the diagonal, or in
424: any direction where the coordinate ratios are rational numbers
425: with small denominators, then these methods give results that are
426: in theory at least as good as ours. The method, however, is
427: inherently non-uniform in the direction, so there is no hope of
428: extending it to larger sets of directions, which is what we
429: accomplish in the present work.
430:
431: The remainder of the paper is organized as follows. In the next
432: section we set forth notation and define the terms necessary to
433: state the main results of the paper. The main results are stated
434: in Section~\ref{ss:results}, and examples are given. The next
435: section contains a proof of these results, modulo the computation
436: of some oscillating integrals. This computation is carried out in
437: Section~\ref{ss:oscillate}. Section~\ref{ss:classify} outlines
438: some details of taxonomy and discusses universality of the method
439: of complex contour integration. The final section states some
440: open problems.
441:
442: \section{Notation and Preliminaries}
443:
444: The main results of this paper give asymptotics valid under
445: certain geometric assumptions on $\sing$ and computable from some
446: quantities that are in turn effectively computable from the
447: generating function $F$. Thus in addition to setting out basic
448: notation, we need to define some terms related to the geometry of
449: $\sing$ and some quantities associated with $F$.
450:
451: \subsection{Notation}
452:
453: Throughout the paper, $F = \sum a_\b{r} \b{z}^\b{r}$ will denote a
454: function on $\CC^d$ analytic in a neighborhood of the origin. The
455: (open) domain of convergence of the power series will be denoted
456: $\dom$. For $\b{z} \in \CC^d$, let $\torus (\b{z})$ denote the torus
457: consisting of points $\b{w}$ with $|w_j| = |z_j|$ for $1 \leq j \leq
458: d$ and let $\disk (\b{z})$ denote the closed polydisk of points
459: $\b{w}$ with $|w_j| \leq |z_j|$ for $1 \leq j \leq d$. Recall (see
460: \citeasnoun{hormander;intro}) that the domain $\dom$ is a union of tori
461: $\torus (\b{z})$ and is logarithmically convex, that is, the set
462: $$\logdom := \{ \b{x} \in \R^d : (e^{x_1} , \ldots , e^{x_d})
463: \in \dom \}$$
464: is a convex subset of $\R^d$ and is an order ideal, that is, it is
465: closed under $\leq$ in the coordinatewise partial order.
466:
467: We assume throughout that $F = \numer / \denom$, where both
468: $\numer$ and $\denom$ are analytic in a neighborhood of
469: $\disk (\b{z})$ for some point $\b{z}$. In particular, every
470: meromorphic function satisfies this condition
471: \footnote{The greater
472: generality allows us to cover examples such as the generating
473: function for self-avoiding random walks
474: \cite{chayes-chayes;self-avoiding} or percolation paths in the
475: subcritical regime \cite{campanino-chayes-chayes;percolation}. In
476: these cases, all the work is in showing the function is
477: meromorphic in a neighborhood of
478: $\disk (\b{z})$. Without further knowledge, the authors then
479: conclude central limit behavior.}. The set where
480: $\denom$ vanishes will be denoted $\sing$. Many of our examples will
481: be in dimension 2, in which case we will often use $z$ and
482: $w$ in place of $z_1$ and $z_2$, use $(z,w)$ in place of
483: $\b{z}$, and use $(r,s)$ in place of $(r_1 , r_2)$. We sometimes
484: need to treat $\CC^d$ as $\CC^{d-1} \times \CC$ (although symmetry of
485: the coordinates is preserved most of the time). Accordingly, when
486: the dimension is greater than~2, we use $\widehat{\b{z}}$ to denote
487: $(z_1 , \ldots , z_{d-1})$. Partial derivatives will be denoted
488: $H_1$ for $\frac{\partial H}{\partial z_1}$ and so forth;
489: in dimension~2 we will also use $H_z$ and $H_w$.
490:
491: As is usual for asymptotic analyses, we let $f \sim g$ denote
492: $f / g \to 1$, with the limit taken at infinity unless otherwise
493: specified. The function $f$ is said to be {\em rapidly
494: decreasing} if $f(x) = O(x^{-N})$ for every $N$, and is said to be
495: {\em exponentially decreasing} if $f(x) = O(e^{-cx})$ for some $c
496: > 0$. We also use the symbol ``$\sim$'' to denote asymptotic
497: expansion. Thus
498: $$f \sim \sum b_n g_n$$
499: is normally taken to mean that $f - \sum_{n=0}^N b_n g_n = o(b_N
500: g_N)$, where $b_n\in \mathbb{C}$ and $\{ g_n \}$ is a fixed sequence of functions such that
501: $g_{n+1} = o(g_n)$ for each $n$. We broaden this to allow $b_n = 0$
502: when $n \neq 0$, so that the remainder term need only be $o(g_n)$
503: and not $o(b_n g_n)$. In particular, if
504: $$f(x) \sim g(x) \cdot \sum_{n = 0}^\infty c_n x^{-n}$$
505: with $c_0 = 1$, then we say we have obtained a full asymptotic
506: expansion for $f$ in decreasing powers of $x$ with leading term
507: $g$.
508:
509: \subsection{Geometry of $\sing$} \label{ss:geom}
510:
511: As in the one-dimensional case, the points of $\sing$ nearest the
512: origin are the most important. Accordingly we define a point $\b{z}
513: \in \sing$ to be {\em minimal} if $\sing \cap \disk (\b{z}) \subseteq
514: \torus (\b{z})$; we say that $\b{z}$ is {\em locally minimal}
515: if the analogous relation holds with $\sing$ replaced by a neighborhood of $\b{z}$ in $\sing$. Divide the minimal
516: points of $\sing$ into three types. Say that
517: $\b{z}$ is {\em strictly minimal}, {\em finitely minimal} or {\em
518: toral}, according to whether the cardinality of
519: $\sing \cap \disk (\b{z})$ is 1, finite, or infinite. When infinite, the
520: intersection must be uncountable. If $\b{z}$ is a minimal point of
521: $\sing$ then the interior of $\disk (\b{z})$ is contained in $\dom$,
522: so the assumption that $\numer$ and $\denom$ are analytic on a
523: neighborhood of $\disk (\b{z})$ is just a little stronger than what
524: is true automatically.
525:
526: A {\em simple pole} of $F$ is a point $\b{z} \in \sing$ where $\denom$
527: vanishes to order 1. Equivalently, the gradient $\grad \denom$
528: does not vanish. Let $\b{z}$ be a simple pole of $F$ and assume for
529: specificity that $\denom_d$ is nonzero at $\b{z}$. By the implicit
530: function theorem, there is a neighborhood of $\b{z}$ where
531: $\sing$ may be parametrized by $z_d = g(z_1 , \ldots , z_{d-1})$ for
532: some analytic function $g$. We will always use $g$ to denote this
533: parametrization.
534:
535: We will see later (in the proof of Theorem~\ref{th:classify}) that
536: under some hypotheses on $F$, minimal points of $\sing$ are always
537: found in the positive real orthant. A relation true in complete
538: generality is the following.
539:
540: \begin{lemma} \label{lem:dir}
541: Let $\b{z}$ be a simple pole of $F$ and suppose that $z_d \denom_d$
542: does not vanish there. If $\b{z}$ is locally minimal then for all
543: $j < d$, the quantity $z_j \denom_j / (z_d \denom_d)$ is real and
544: nonnegative.
545: \end{lemma}
546:
547: \begin{proof} Given $\theta$ and $j$, let $\b{z}^{(\theta)}$
548: be the result of varying $\b{z}$ by multiplying the $j^{th}$
549: coordinate by $e^{i \theta}$ and adjusting the last coordinate so
550: as to remain on $\sing$ (that is, \hfill \\ $z^{(\theta)}_d = g(z_1 ,
551: \ldots , z_{j-1} , z_j e^{i \theta} , z_{j+1} , \ldots ,
552: z_{d-1})$). Differentiating the relation $\denom (\b{z}^{(\theta)})
553: = 0$ implicitly with respect to
554: $\theta$ at 0 yields
555: \begin{equation} \label{eq:minimal}
556: i z_j \denom_j + \denom_d {d z^{(\theta)}_d \over d \theta} = 0 .
557: \end{equation}
558: By minimality of $\b{z}$, we know that the modulus of
559: $z^{(\theta)}_d$ has a minimum at $\theta = 0$, hence
560: $(d z^{(\theta)}_d / d \theta) / z_d$ is purely imaginary.
561: Plugging this into~(\ref{eq:minimal}) proves that $z_j \denom_j /
562: (z_d \denom_d)$ is real. If $z_j \denom_j / (z_d \denom_d) =
563: -\beta < 0$ then $\sing$ has a tangent vector at $\b{z}$ in the
564: direction $- z_j e_j - \beta z_d e_d$, where $e_j$ is the
565: $j^{th}$ coordinate vector. This contradicts minimality.
566: Hence $z_j \denom_j / (z_d \denom_d) \geq 0$. \end{proof}
567:
568: \begin{definition}
569: Define $\direc (\b{z})$ to be the equivalence class of (complex)
570: scalar multiples of the vector $(z_1 \denom_1 , \ldots , z_d
571: \denom_d)$, defined whenever $z_j \denom_j$ does not vanish for
572: all $j$. By the previous lemma, when $\b{z}$ is a minimal pole of
573: $F$ with nonzero coordinates, $\direc (\b{z})$ is a well defined
574: element of
575: $\RP^{d-1}$.
576: \end{definition}
577: %
578: The importance of $\direc$ is that analysis of $F$ near $\b{z}$
579: yields asymptotic information about $a_\b{r}$ with $\b{r}
580: \in \direc (\b{z})$. The function $\direc$ appears in $\GF$
581: method literature as ${\bf m}$. When $\b{z} \in \partial \dom$ is
582: on the boundary of the domain of convergence, $\direc (\b{z})$ is
583: the normal to the support hyperplane of the convex set $\logdom$
584: at the point $(\log |z_1| , \ldots , \log |z_d|)$.
585:
586: We now define a few more quantities associated with $F$ and $g$.
587: Again, we will reserve the names of these functions, so as not to
588: burden the notation with subscripts and arguments. If $\b{z}$ is a
589: simple pole of $F$ with $z_d \denom_d$ not vanishing there, define
590: a function $\psi$ on a neighborhood of $\widehat{\b{z}}$ by
591: \begin{equation} \label{eq:psi}
592: \psi (\widehat{\b{w}}) = - \lim_{w \rightarrow g(\widehat{\b{w}})}
593: (w - g(\widehat{\b{w}})) {F (\widehat{\b{w}} , w) \over w} .
594: \end{equation}
595: Suppose now that $\widehat{\b{w}} \in \torus (\widehat{\b{z}})$ and
596: write
597: $w_j = z_j e^{i \theta_j}$. For fixed $\b{r}$ with $r_d \neq 0$,
598: define a function $f$ on a neighborhood of $\widehat{\b{z}}$ in
599: $\torus (\widehat{\b{z}})$ by
600: \begin{equation} \label{eq:f}
601: f (\widehat{\b{w}}) = \log \left ( {g (\widehat{\b{w}}) \over
602: g(\widehat{\b{z}})} \right ) + i \sum_{j=1}^{d-1} {r_j \over r_d}
603: \theta_j .
604: \end{equation}
605:
606: We will be parametrizing integrals over $\torus (\widehat{\b{z}})$
607: by ${\bf \theta}$, so we will want the above function expressed in
608: terms of $\widehat{\bf \theta}$. We therefore compose with the
609: map $M$ taking $\widehat{\bf \theta}$ to $\widehat{\b{w}}$ defined
610: by
611: $M(\theta_1 , \ldots , \theta_{d-1}) = (z_1 e^{i \theta_1} ,
612: \ldots , z_{d-1} e^{i \theta_{d-1}})$, and define the functions
613: $\gt: = g \circ M , \ft := f \circ M , \psit := \psi \circ M$.
614:
615: Although it is not obvious yet, $\ft$ will always vanish at
616: $\zero$ to at least two orders (Lemma~\ref{lem:stat} below), and
617: the hypothesis $Q \neq 0$ in Theorem~\ref{th:simple} is equivalent
618: to $\ft$ having nonvanishing quadratic term. For ease of
619: reference, Table~\ref{table1} summarizes the foregoing definitions,
620: stratified by how many times the given data $\numer$ and $\denom$
621: have been manipulated.
622:
623: %\pagebreak
624:
625: \begin{table}
626: \label{table1}
627: \caption{Reserved notation in remainder of this article}
628: \begin{tabbing}
629: Given \= information: \\
630: \> the function $F$ in the form $\numer / \denom$ \\[3ex]
631: First level: \\
632: \> $g$ parametrizes the zero set, $\sing$ of $\denom$ \\
633: \> $\direc(\b{z})$ is the coordinatewise product $(\grad \denom)\cdot(\b{z})$
634: in projective space \\[3ex]
635: Second level: \\
636: \> $\psi$ is the residue in $z_d$ of $F / z_d$ at points $(\widehat{\b{z}} ,
637: g(\widehat{\b{z}}))$ \\
638: \> $f$ is $\log g$, plus a term linear in $\log z_j$ and depending
639: on $\b{r}$. \\[3ex]
640: Third level: \\
641: \> $\psit , \gt$ and $\ft$ are $\psi, g$ and $f$ expressed in
642: terms of ${\bf \theta}$
643: \end{tabbing}
644: ~~\\[2ex]
645: \end{table}
646:
647: \section{Statement of results, with examples} \label{ss:results}
648:
649: Before going on, we pause to state a prototype of our results in
650: the simplest possible setting, namely where the number of
651: variables is~2, the functions $\numer$ and $g$ are as
652: nondegenerate as possible, and only the leading term asymptotic is
653: given. The proof is in Section~\ref{ss:proofs}.
654:
655: \begin{theorem} \label{th:simple}
656: Let $F = \numer / \denom$ be a meromorphic function of two
657: variables, not singular at the origin. Define
658: $$Q (z,w) := - w^2 \denom_w^2 z \denom_z - w \denom_w z^2 \denom_z^2 - w^2 z^2
659: \left ( \denom_w^2 \denom_{zz} + \denom_z^2 \denom_{ww}
660: - 2 \denom_z \denom_w \denom_{zw} \right ) .$$
661: Then
662: $$a_{r,s} \sim {\numer (z,w) \over \sqrt{2 \pi}} z^{-r} w^{-s}
663: \sqrt{- w \denom_w \over s Q}$$
664: uniformly as $(z,w)$ varies over a compact set of strictly
665: minimal, simple poles of $F$ on which $Q$ and $\numer$ are
666: nonvanishing, and $(r,s) \in \direc (z,w)$.
667: \end{theorem}
668:
669: \noindent{\em Remarks:} Usually the expression in the radical will
670: be positive real, as will the coefficients $a_{rs}$. The result
671: is true in general, though, as long as the square root is taken to
672: be $-w \denom_w$ times the principal root of $Q / (- w
673: \denom_w^3)$. Also note that when $(r,s) \in \direc (z,w)$ then
674: the expression
675: $w \denom_w / s$ is coordinate-invariant, that is, equal to $z \denom_z / r$.
676: Thus the given expression for $a_{r,s}$ has the expected symmetry.
677:
678: \begin{example}[Lattice paths] \label{ex:lattice} ~~ \end{example}
679:
680: Let $a_{r,s}$ be the number of nearest-neighbor paths from the
681: origin to $(r,s)$ moving only north, east and northeast; these are
682: sometimes called {\em Delannoy numbers}~\citeasnoun[page
683: 185]{stanley;enumerative-combinatorics2}. The generating function
684: is $F(z,w) = 1 / (1 - z - w - zw)$. The zero set $\sing$ of
685: $\denom = 1 - z - w - zw$ is given by
686: $w = (1-z)/(1+z)$, and the minimal points of $\sing$ are those where
687: $w \in [0,1]$. With the help of relations that hold when $\b{z} \in \sing$
688: we may compute as follows.
689: \begin{eqnarray*}
690: \denom_z & = & - 1 - w \\
691: -z \denom_z & = & 1 - w \\
692: Q & = & (1-z) (1-w) (1-zw) \\
693: {z \denom_z \over w \denom_w} & = & {1-w \over 1-z} =
694: {1 - w^2 \over 2w}
695: \end{eqnarray*}
696: with $\denom_w$ and $-w \denom_w$ given by reversing $z$ and $w$.
697: As $z$ varies over $[\ee , 1-\ee]$, the functions $Q$ and $\numer
698: := 1$ do not vanish. The minimal pair $(z,w)$ that solves $(r,s)
699: \in \direc (z,w)$ is given by $z = (\sqrt{r^2 + s^2} - s)/r$ and
700: $w = (\sqrt{r^2 + s^2} - r)/s$. Theorem~\ref{th:simple} then
701: gives
702: \begin{eqnarray*}
703: a_{rs} & \sim & \left ({\sqrt{r^2 + s^2} - s \over r} \right)^{-r}
704: \left ( {\sqrt{r^2 + s^2} - r \over s} \right )^{-s}
705: \sqrt{1 \over 2 \pi} \sqrt{{1-z \over s} {1 \over 1- zw}}
706: \\[1ex]
707: & = & \left ({\sqrt{r^2 + s^2} - s \over r} \right)^{-r}
708: \left ( {\sqrt{r^2 + s^2} - r \over s} \right )^{-s}
709: \sqrt{1 \over 2 \pi} \sqrt{r s \over (r+s-\sqrt{r^2 + s^2})^2
710: \sqrt{r^2 + s^2}} \, ,
711: \end{eqnarray*}
712: uniformly when $r/s$ and $s/r$ remain bounded. In particular,
713: when $r=s=n$, this gives the following formula for the $n^{th}$
714: diagonal coefficient (which may alternatively be obtained by
715: computing the diagonal generating function $(1 - 6s+ s^2)^{-1/2}$
716: according to the method given in
717: \citeasnoun[Section~6.3]{stanley;enumerative-combinatorics2}:
718: $$(\sqrt{2} - 1)^{-2n} \sqrt{1\over 2 \pi} {2^{-1/4} \over
719: 2 - \sqrt{2}} \, .$$
720:
721: The computations in Theorem~\ref{th:simple} in terms of the values
722: and derivatives of $\numer$ and $\denom$ are explicit. As we
723: state more general theorems, it becomes cumbersome and in fact
724: obfuscating to give formulae for the expansion coefficients
725: directly in terms of derivatives of $\numer$ and $\denom$. This
726: is one reason we have already introduced the functions in
727: Table~\ref{table1}. It should be emphasized, however, that while we use
728: higher level quantities in the statements of subsequent theorems,
729: each expansion coefficient can be computed from finitely many
730: derivatives of $\numer$ and $\denom$. We begin with a relatively
731: explicit computation for the general two-variable case.
732:
733: For $k$ at least 2, we define constants
734: \begin{eqnarray}
735: A_+ (k,l) & := & {1 \over k} \Gamma \left ({l+1 \over k} \right )
736: \label{eq:plus} \\[2ex]
737: A (k,l) & := & {1 \over k} \Gamma \left ( {l+1 \over k} \right )
738: \left ( 1 + e^{\sign\Arg (c_k) i \pi (l - {l+1 \over k} )} \right ) \,
739: \mbox{ if } k \mbox{ is odd}, \label{eq:k odd} \\[2ex]
740: A (k,l) & := & {2 \over k} \Gamma \left ( {l+1 \over k} \right )
741: \mbox{ if } k,l \mbox{ are even},\label{eq:k even} \\[2ex]
742: A (k,l) & := & 0 \mbox{ if } k \mbox{ is even and } l \mbox{ is
743: odd.} \nonumber
744: \end{eqnarray}
745: Let
746: $$y (x) = f(x)^{1/k} = c_k^{1/k} x \left ( 1 + {f(x) - c_k x^k
747: \over c_k x^k} \right )^{1/k} \, ,$$
748: where $c_k$ is the first nonvanishing Taylor coefficient of $f (x)
749: = \sum_{j=k}^\infty c_j x^j$ and the argument of
750: $c_k^{1/k}$ is taken between $-\pi/(2k)$ and
751: $\pi / (2k)$. Let $\finv$ denote the inverse function to $y$
752: and let $\{ b_j \}$ be the Taylor coefficients of
753: $(\psit \circ \finv) \cdot \finv'$. Clearly each $\{ b_j \}$
754: is determined by finitely many partial derivatives of $\numer$ and
755: $\denom$, and the index $l_0$ of the first nonvanishing $b_l$ is
756: the same as the order of vanishing of $\psit$ at 0. The
757: coefficients $b_l$ are easily computed from the coefficients
758: $\bt_j := \psit^{(j)} (0) / j!$ and $c_j := \ft^{(j)} (0) / j!$;
759: in particular, if $\ft \sim c_k x^k$ near 0 then
760: \begin{equation} \label{eq:b}
761: b_{l_0} = \bt_{l_0} c_k^{-1/k} \, .
762: \end{equation}
763:
764: \begin{theorem} \label{th:2d}
765: Let $F = \numer / \denom = \sum a_{rs} z^r w^s$ have a strictly
766: minimal, simple pole at $(z,w)$. Let $k$ be the order of vanishing
767: of $\ft$ at 0. Let $l_0$ be the order to which $\numer$ vanishes
768: near $(z,w)$ on $\sing$, that is, the largest $l$ such that $G(z',w')
769: = O(|z-z'|^l + |w-w'|^l)$ as $(z' , w') \to (z,w)$ in $\sing$. Then
770: there is a full asymptotic expansion
771: \begin{equation}\label{eq:full 2d}
772: a_{r,s} \sim {1 \over 2 \pi} z^{-r} w^{-s} \sum_{l \geq l_0}
773: \A (k,l) b_l s^{-(l+1) / k} \, ,
774: \end{equation}
775: where $\A (k,l)$ denotes $A (k,l)$ if $\im \{ c_k \} \geq 0$ and
776: $\overline{A(k,l)}$ otherwise. The expansion is uniform as $(z,w)$
777: varies over a compact set of strictly minimal poles with $(r,s)
778: \in \direc (z,w)$ and $k$ and $l_0$ not changing.
779: \end{theorem}
780:
781: \begin{figure}
782: \scalebox{0.6}{\input{p1-fig2.tex}}
783: \vspace{10pt}
784: \caption{$\sing$ for Example~\ref{ex:cubic}}
785: \label{fig:cuberoot}
786: \end{figure}
787:
788: \begin{example}[Cube root asymptotics] \label{ex:cubic}
789:
790: Let $F(z,w) = 1 / (3 - 3z - w + z^2)$. The set $\sing$ is the set
791: $\{ w = z^2 - 3z + 3 \}$ and $g(z) = z^2 - 3z + 3$. The point
792: $(1,1)$ is in $\sing$, indicating that the maximal exponential growth
793: rate will be zero. Indeed, for directions above the diagonal,
794: Theorem~\ref{th:simple} or~\ref{th:2d} may be used at the minimal
795: points $\{ (z, g(z)) : 0 < z < 1 \}$, while each direction below
796: the diagonal corresponds to a pair of complex minimal points
797: fitting the hypotheses of Corollary~\ref{cor:finitely}; the result
798: is that the coefficients decay exponentially at a rate that is
799: uniform over compact subsets of directions not containing the
800: diagonal.
801:
802: The interesting behavior is near the diagonal. The relevant
803: minimal point is $(1,1)$, where $z^r w^s \equiv 1$ and the decay
804: is sub-exponential. Computing $\ft''(0)$ via
805: equation~(\ref{eq:ft''}) below gives
806: $$\ft'' (z) = -3 {z (z^2 - 4z + 3) \over (z^2 - 3z + 3)^2} \, .$$
807: This vanishes when $z = 1$, and computing further, we find that
808: $\ft$ vanishes to order exactly 3 here, with $c_3 := \ft'''(0) /
809: 3! = i$. Along with $\psit (0) = 1$, this then results in an
810: asymptotic expansion whose leading term is given by
811: $$a_{r,r} \sim {1 \over 2 \pi} A (3,0) i^{-1/3} (1 + e^{- i \pi /
812: 3}) r^{-1/3} = {\Gamma(2/3) \over 6 \sqrt{3} \pi} r^{-1/3}\, .$$
813: In Section~\ref{ss:open} we discuss the question of computing
814: asymptotics ``in the gaps'' so as to be able to conclude that
815: $\limsup \log a_\b{r} / \log |\b{r}| = -1/3$ or even $\limsup |\b{r}|^{1/3}
816: a_\b{r} = {\Gamma(2/3) \over 6 \sqrt{3} \pi}$.
817: \end{example}
818:
819: For more than two variables a result holds similar to the
820: two-variable result.
821: %
822: \begin{theorem} \label{th:higher d}
823: Let $F = \numer / \denom = \sum a_\b{r} \b{z}^\b{r}$ have a strictly
824: minimal, simple pole at $\b{z}$. Suppose $z_d \denom_d$ does not
825: vanish. If the Hessian of $\ft$ at $\b{z}$ is nonsingular, then
826: there is an expansion
827: $$a_\b{r} \sim \b{z}^{-\b{r}} \sum_{l \geq l_0} C_l r_d^{(1-d-l)/2}$$
828: where $l_0$ is the degree to which $\numer$ vanishes on $\sing$ near
829: the point $\b{z}$ . When
830: $\numer$ does not vanish at $\b{z}$ then $l_0 = 0$ and
831: $$C_0 = (2 \pi)^{(1-d)/2} \hess^{-1/2} { \numer (\b{z}) \over z_d H_d}$$
832: where $\hess$ is the determinant of the Hessian at $\b{z}$.
833: \end{theorem}
834:
835: \begin{example}[Domino tilings] \label{ex:arctic}
836:
837: Random perfect tilings of planar regions by dominos have been a
838: subject of some interest, since the analysis by \citeasnoun{fisher;dimer}
839: of this
840: model for dimer packing uncovered an exact expression for the
841: partition function of the ensemble. A generating function is
842: given in \citeasnoun{cohn-elkies-propp;aztec} which allowed the authors to
843: determine, after some cumbersome analysis, which parts of a
844: diamond-shaped region (a union of lattice squares approximating
845: the region $|x| + |y| \leq k$) were asymptotically deterministic
846: and which contained randomness in the limit as the edge size of
847: the diamond grew.
848:
849: An easier analysis in the region of non-randomness is available
850: via Theorem~\ref{th:higher d} together with a slightly more
851: informative generating function than was used by
852: \citeasnoun{cohn-elkies-propp;aztec}.
853: In particular, let
854: $$F(x,y,z) = \sum_{t=0}^{\infty} \sum_{|r| + |s| \leq t}
855: a_{r,s,t} x^r y^s z^t$$
856: be the generating function for the probability $a_{r,s,t}$ that
857: the tile covering position $(r,s)$ of a random diamond of size $t$
858: will be horizontal. For brevity, we omit formal descriptions of
859: the diamond and its indexing. We remark that the use of negative
860: indices (for each fixed $t$, the sum $\sum_{|r| + |s| \leq t}$ is
861: a polynomial in $x, x^{-1}, y$ and $y^{-1}$) does not require any
862: alterations in the theory (see \citeasnoun{cohn-pemantle;fixation} for
863: justification), and that the natural way to parametrize directions
864: is by the pair $(r/t , s/t)$ which varies over the diamond $|r/t|
865: + |s/t| = 1$. From \citeasnoun{cohn-elkies-propp;aztec} or from the
866: generation algorithm in \citeasnoun{gessel-ionescu-propp}, one
867: finds
868: $$F(x,y,z) = {z/2 \over 1 - (x + x^{-1} + y + y^{-1}) z/2
869: + z^2} \, .$$
870: \citeasnoun{cohn-pemantle;fixation} show that whenever $(r,s,t)$ satisfy
871: $$t = \sqrt{r^2 + s^2 + 2 \sqrt{r^2 + 1} \sqrt{s^2 + 1} - s} \, ,$$
872: then there is a smooth minimal point $(x,y,z)$ on the pole
873: manifold of $F$ for which $(r,s,t) \in \direc (x,y,z)$, yielding
874: exponential decay in the direction $(r,s,t)$. The set of
875: directions so parametrized turns out to be the region between the
876: diamond $|r/t| + |s/t| = 1$ and the inscribed circle $(r/t)^2 +
877: (s/t)^2 = 1/2$. Thus they recover the description of the region
878: of non-randomness as the complement of the inscribed circle. They
879: also obtain descriptions of the region of fixation for related
880: tiling problems in which no other analysis has been carried out.
881: \end{example}
882:
883: The extension of all of the above results to finitely minimal
884: points is routine.
885:
886: \begin{corollary} \label{cor:finitely}
887: Suppose $\b{z}$ is a finitely minimal point of $\sing$ with
888: $\sing \cap \torus (\b{z}) = \{ \b{z}_1 , \ldots , \b{z}_n \}$. Then
889: $$a_\b{r} \sim \sum_{j=1}^n E_j (\b{r})$$
890: where $E_j (\b{r})$ is the asymptotic expression given by the
891: previous theorems with $\b{z} = \b{z}_j$. In other words, if there
892: are finitely many points on $\sing \cap \torus (\b{z})$, then sum the
893: contributions as if each were strictly minimal.
894: \noproof
895: \end{corollary}
896:
897: \begin{example}[Chebyshev polynomials] \label{ex:Tsch}
898:
899: Let $F(z,w) = 1 / (1 - 2zw + w^2)$ be the generating function for
900: Chebyshev polynomials of the second kind \cite{comtet;advanced}; of
901: course asymptotics for these are well known and easy to derive by
902: other means. To use Corollary~\ref{cor:finitely}, first find the
903: minimal points for the direction $(r,s)$, which are $(i (\beta -
904: \beta^{-1})/2 , i \beta)$ for $\beta = \pm \sqrt{s-r \over s+r}$.
905: Computing $Q = 4 a^2 (1 - a^2)$ and summing the two contributions
906: then gives
907: $$a_{rs} \sim \sqrt{2 \over \pi} (-1)^{(s-r)/2} \left ( {2r \over
908: \sqrt{s^2 - r^2}} \right )^{-r} \left ( \sqrt{s-r \over s+r}
909: \right )^{-s} \sqrt{s+r \over r(s-r)}$$
910: when $r+s$ is even and zero otherwise, uniformly as $r/s$ varies
911: over compact subsets of $(0,1)$.
912:
913: \end{example}
914:
915: \section{Proofs of main results} \label{ss:proofs}
916:
917: Half of each theorem is easy and follows directly from Cauchy's
918: formula
919: \begin{equation} \label{eq:cauchy}
920: a_\b{r} = \left ( {1 \over 2 \pi i} \right )^d
921: \int_T \b{w}^{-\b{r} - \one} F(\b{w}) \, d\b{w}
922: \end{equation}
923: where the multi-exponent $\b{r} - \one$ means $(r_1 - 1 , \ldots ,
924: r_d - 1)$. Indeed, if $\b{z}$ is a minimal point of $\sing$ then
925: letting $T$ approach $\torus (\b{z})$ from the inside, we see that
926: $|\b{z}^\b{r}| a_\b{r}$ does not increase exponentially. If, furthermore,
927: the hyperplane through $(\log |z_1| , \ldots , \log |z_d|)$ normal
928: to $\b{r}$ is not a support hyperplane for $\logdom$, then some $\b{x}
929: \in \logdom$ has $\b{x} \cdot \b{r} > (\log |z_1| , \ldots , \log |z_d|)
930: \cdot \b{r}$, and integrating on the torus $T(e^\b{x})$ shows that
931: $|\b{z}^\b{r}| a_\b{r}$ decreases exponentially. All the work,
932: therefore, is in showing the converse, namely that when the
933: hyperplane normal to $\b{r}$ is a support hyperplane, then
934: $\b{z}^{-\b{r}}$ does give the right exponential order for $a_\b{r}$.
935: This is done by evaluating $a_\b{r}$.
936:
937: Theorems~\ref{th:simple}--\ref{th:higher d} all begin with the
938: reduction of an iterated Cauchy integral to an oscillating
939: integral in one fewer dimension.
940:
941: \begin{lemma} \label{lem:X}
942: Let $\b{z}$ be a strictly minimal simple pole of $F=\numer/\denom$.
943: Assume that $z_d \denom_d \neq 0$. For a neighborhood
944: $\widetilde{\nbd}$ of $\zero$ in $\R^{d-1}$ define a quantity
945: \begin{equation}\label{eq:X}
946: \X := (2 \pi)^{1-d} \b{z}^{-\b{r}} \int_{\widetilde{\nbd}} \exp (- r_d \ft
947: (\widehat{\bf \theta})) \psit (\widehat{\bf \theta}) \,d\widehat{\bf \theta}.
948: \end{equation}
949: Then the quantity
950: $$|\b{z}^\b{r}| \left | a_\b{r} - \X \right |$$
951: decreases exponentially as $\widetilde{\nbd}$ remains fixed and
952: $\b{r} \to \infty$.
953: \end{lemma}
954:
955: \begin{proof}
956: For $\ee \in (0 , |z_d|)$, let $T$ be the torus $T(\b{z})$ shrunk in
957: the last coordinate by $\ee$, that is, the set of $\b{w}$ for which
958: $|w_j| = |z_j|$, $j < d$ and $|w_d| = |z_d| - \ee$. Write
959: Cauchy's formula as an iterated integral
960: \begin{equation} \label{eq:iter}
961: a_\b{r} = \left ( {1 \over 2 \pi i} \right )^d
962: \int_{T(\widehat{\b{z}})} \widehat{\b{w}}^{-\widehat{\b{r}} - \one}
963: \left [ \int_{\C_1} w_d^{-r_d} F(\b{w}) \,
964: {d w_d \over w_d} \right ] \, d \widehat{\b{w}} \, .
965: \end{equation}
966: Here $\C_1$ is the circle of radius $|z_d| - \ee$. Let $K
967: \subseteq T(\widehat{\b{z}})$ be a compact set not containing
968: $\widehat{\b{z}}$. For each fixed $\widehat{\b{w}} \in K$, the function
969: $F(\widehat{\b{w}} , \cdot)$ has radius of convergence greater than
970: $|z_d|$. Hence the inner integral in equation~(\ref{eq:iter})
971: is $O(|z_d|+ \delta)^{-r_d}$ for some $\delta > 0$. By continuity
972: of the radius of convergence,we may integrate over $K$ to see that
973: $$|\b{z}^\b{r}| \int_{K \times \C_1} \b{w}^{-\b{r} - \one} F(\b{w}) \, d\b{w}$$
974: decreases exponentially. Thus if $\nbd$ is any neighborhood of
975: $\widehat{\b{z}}$ in $T(\widehat {\b{z}})$, the quantity
976: $$|\b{z}^\b{r}| \left | a_\b{r} - \left ( {1 \over 2 \pi i} \right )^d
977: \int_\nbd \widehat{\b{w}}^{-\widehat{\b{r}} - \one}
978: \left [ \int_{\C_1} {F(\b{w}) \over w_d^{r_d + 1}} \, dw_d
979: \right ] d \widehat{\b{w}} \right |$$
980: decreases exponentially. Thus we have reduced the problem to an
981: integral over a neighborhood of $\widehat{\b{z}}$.
982:
983: Near $\b{z}$ there is a parametrization $w_d = g(\widehat{\b{w}})$ of
984: $\sing$. Let $\C_2$ be the circle of radius $|z_d| + \ee$. Then
985: when $\nbd$ is sufficiently small compared to $\ee$, the image of
986: $\nbd$ under $g$ is disjoint from $\C_2$. Fix such a neighborhood.
987: For any $\widehat{\b{w}} \in \nbd$, the function $F(\widehat{\b{w}} ,
988: \cdot)$ has a single simple pole in the annulus bounded
989: by $\C_1$ and $\C_2$, occurring at $g(\widehat{\b{w}})$. The
990: residue in the last variable of $F$ at $g(\widehat{\b{w}})$ is equal
991: to
992: \begin{equation} \label{eq:defres}
993: R(\widehat{\b{w}}) := - \psi (\widehat{\b{w}}) g(\widehat{\b{w}})^{-r_d}
994: \end{equation}
995: where $\psi$ is defined in~(\ref{eq:psi}). Therefore, for each
996: fixed $\widehat{\b{w}} \in \nbd$,
997: $$\int_{\C_1} {F(\b{w}) \over w_d^{r_d + 1}} \, d w_d
998: = \int_{\C_2} {F(\b{w}) \over w_d^{r_d + 1}} \, d w_d
999: - 2 \pi i R(\widehat{\b{w}}) .$$
1000: But $|\b{z}^\b{r} \int_{\C_2} F(\b{w}) dw_d / \b{w}^{\b{r} + \one}|$ is
1001: bounded by a constant multiple of $(1 + \ee / |z_d|)^{-r_d}$ (the
1002: constant depending on the maximum of $F$ on $\C_2$) and hence
1003: $|\b{z}^\b{r}| |a_\b{r} - X|$ is exponentially decreasing, where
1004: \begin{eqnarray} \label{eq:res}
1005: X & = & (2 \pi i)^{1-d} \int_\nbd (\widehat{\b{w}})^{-\widehat{\b{r}}
1006: - 1} g(\widehat{\b{w}})^{-r_d} \psi (\widehat{\b{w}}) \,
1007: d \widehat{\b{w}} \\[1ex]
1008: & = & (2 \pi i)^{1-d} \b{z}^{-\b{r}} \int_\nbd {\widehat{\b{w}}^{-\widehat{\b{r}}}
1009: \over \widehat{\b{z}}^{-\widehat{\b{r}}}} {d \widehat{\b{w}} \over
1010: \prod_{j=1}^{d-1} w_j} \left ( {g(\widehat{\b{w}}) \over g(z_d)}
1011: \right )^{-r_d} \psi (\widehat{\b{w}}) \nonumber
1012: \end{eqnarray}
1013: Changing variables to
1014: $w_j = z_j e^{i \theta_j}$ and $dw_j = i w_j d\theta_j$ turns the
1015: quantity $X$ into
1016: $$(2 \pi)^{1-d} \b{z}^{-\b{r}} \int_{\widetilde{\nbd}} \prod_{j=1}^{d-1}
1017: e^{-i r_j \theta_j} \psit (\widehat{\bf \theta}) \left ( {
1018: g(\widehat{\b{w}}) \over g(\widehat{\b{z}})} \right )^{-r_d}
1019: \, d\widehat{\bf \theta} $$
1020: and plugging in the definitions of $f$ and $\ft$ at~(\ref{eq:f})
1021: above yields
1022: $$(2 \pi)^{1-d} \b{z}^{-\b{r}} \int_{\widetilde{\nbd}} \exp (- r_d
1023: \ft (\widehat{\bf \theta})) \psit (\widehat{\bf \theta}) \,
1024: d\widehat{\bf \theta}$$
1025: which is none other than $\Xi$. \end{proof}
1026:
1027: \begin{unremark} It is possible to compute from Cauchy's
1028: integral formula in a more coordinate-free way as follows. There
1029: is a unique holomorphic $(d-1)$-form $\omega_F$ on $\sing$ for which
1030: $\omega \wedge d\denom = \numer \, dz_1 \wedge \cdots \wedge dz_d$.
1031: Let $\Omega$ be a ($d+1$)-manifold that is a homotopy from a small
1032: torus to a torus at infinity. Then $M := \Omega \cap \sing$ is a
1033: $(d-1)$-manifold and $a_\b{r} = (2 \pi i)^{-d} \int_M \b{w}^{-\b{r} - \one} dF$
1034: in the sense of currents, which is none other than $\int_M
1035: \b{w}^{\b{r} - \one} \omega_F$. See \citeasnoun{kenyon-pemantle} for a
1036: more thorough discussion of the foregoing. The manifold $M$ is
1037: any member of a certain homology class in $\sing$ with the coordinate
1038: axes removed, and choosing $M$ to pass through the stationary
1039: phase point for the integrand replicates the selection of $\b{z}$
1040: with $\b{r} \in \direc(\b{z})$. Although more canonical, the
1041: coordinate-free method is less suitable for explicit computation,
1042: so we do not pursue it further here. Suffice it to point out that
1043: the conclusion of Theorem~\ref{th:higher d} may of course be
1044: written in terms more evidently symmetric, as was done in
1045: Theorem~\ref{th:simple}.
1046: \end{unremark}
1047:
1048: Equation~(\ref{eq:X}) is easily recognized as the standard form
1049: for an oscillating integral. The only unusual feature is that the
1050: phase is neither real nor purely imaginary. This presents no
1051: difficulties, but it does necessitate the statement of a result in
1052: Section~\ref{ss:oscillate} that is a little different from the
1053: usual results on purely oscillating integrals, found in, for example,
1054: \citeasnoun{stein;oscillatory-integrals} or
1055: \citeasnoun{bleistein-handelsman;asymptotic-expansions-integrals}. We first
1056: establish that $\widehat{\bf \theta} = \zero$ is a stationary
1057: phase point for the function $\ft$ when $\b{r} \in \direc (\b{z})$.
1058: %
1059: \begin{lemma} \label{lem:stat}
1060: The quantity $\ft (\zero)$ always vanishes. If $\b{r} \in \direc
1061: (\b{z})$ then $\grad \ft (\zero) = \zero$ and the real part of $\ft$
1062: has a strict minimum at $\zero$.
1063: \end{lemma}
1064:
1065: \begin{proof} The first statement is immediate. To prove
1066: the second, let $j \leq d-1$ and see from the definition of $f$
1067: that
1068: $$r_d f_j (\widehat{\b{z}}) = {r_d g_j (\widehat{\b{z}}) \over
1069: g(\widehat{\b{z}})} + {r_j \over z_j}.$$
1070: By definition of $\direc$, the ratio $r_j / (z_j \denom_j)$ is
1071: some constant $c$ independent of $j$, hence
1072: $$c^{-1} r_d f(\b{z}) = g_j (\b{z}) \denom_d (\b{z}) + \denom_j (\b{z}) .$$
1073: The right hand side of this is the derivative of $\denom (w_1 ,
1074: \ldots , w_{d-1} , g(\widehat{\b{w}}))$ with respect to $w_j$ at
1075: $\widehat{\b{z}}$. By definition of $g$ this vanishes, and hence
1076: $f_j (\widehat{\b{z}}) = 0$. But $\ft_j (\zero) = i z_j f_j (\b{z})$, so
1077: the gradient of $\ft$ must vanish at $\zero$. Finally, observe
1078: that $\re \{ \ft (\widehat{\bf \theta}) \} = -\log |\gt
1079: (\widehat{\bf \theta}) / z_d|$. By strict minimality of $\b{z}$,
1080: the modulus of $g(\widehat{\b{w}}) = \gt (
1081: \widehat{\bf \theta})$ is greater than $|z_d|$ for any
1082: $\widehat{\b{w}} \in \torus (\widehat{\b{z}})$. \end{proof}
1083:
1084: We now prove Theorems~\ref{th:simple},~\ref{th:2d}
1085: and~\ref{th:higher d} in reverse order. We see from
1086: Lemma~\ref{lem:X} that proving any of these theorems amounts to
1087: evaluating the quantity $\X$ in equation~(\ref{eq:X}). From
1088: Lemma~\ref{lem:stat} we see that $\zero$ is a stationary point for
1089: the function $\ft$ as long as $\b{r} \in \direc (\b{z})$. The
1090: function $\ft$ is in general complex valued, but we will see in
1091: Theorem~\ref{th:nondeg} that it may be treated as if it were real
1092: valued, given the strict minimality of the zero guaranteed by
1093: Lemma~\ref{lem:stat} and the nonsingularity hypothesis. In
1094: particular the leading term of the integral in~(\ref{eq:X}) is
1095: $(2 \pi)^{(d-1)/2} \psit (\zero) r_d^{(1-d)/2}$ divided by the
1096: product of the square roots of the eigenvalues of the Hessian.
1097: Once we have identified $\psit (\zero) = \psi (\zero)$ as $\numer
1098: (\zero) / (z_d \denom_d)$, the theorem follows directly from
1099: Theorem~\ref{th:nondeg}.
1100:
1101: Theorem~\ref{th:2d} follows from the more explicit asymptotic
1102: development given in Corollary~\ref{cor:full 2d}. Finally, to
1103: prove Theorem~\ref{th:simple}, it remains to compute the quantity
1104: $\ft'' (0)$ in terms of the partial derivatives of $\denom$.
1105: First we compute the derivatives of $g$.
1106:
1107: \begin{lemma} \label{lem:der g}
1108: In a neighborhood of $(z,w)$, $\psi$ and the derivatives of $g$
1109: are as follows.
1110: \begin{eqnarray}
1111: g'(z) & = & -{\denom_z \over \denom_w} \label{eq:g'} \\[2ex]
1112: g''(z) & = & - {1 \over \denom_w} \left [ \denom_{zz} - 2
1113: {\denom_z \over \denom_w} \denom_{zw} + {\denom_z^2 \over
1114: \denom_w^2} \denom_{ww} \right ] . \label{eq:g''} \\[2ex]
1115: \psi (z) & = & {\numer (z,w) \over - w \denom_w(z,w)} . \nonumber
1116: \end{eqnarray}
1117: \end{lemma}
1118:
1119: \begin{proof} Differentiate the equation $\denom(z,g(z)) = 0$
1120: to get $\denom_z + g'(z) \denom_w = 0$ which is the same
1121: as~(\ref{eq:g'}). Differentiate again to get
1122: $$\denom_{zz} + 2 g' \denom_{zw} + g'' \denom_w + (g')^2 \denom_{ww} = 0$$
1123: and use~(\ref{eq:g'}) to eliminate $g'$, giving~(\ref{eq:g''}).
1124: The formula for $\psi$ follows from the definitions of $\psi$ and
1125: of the partial derivative.
1126: \end{proof}
1127:
1128: \noindent{\sc Proof of Theorem}~\ref{th:simple}~{\sc via direct
1129: computation:} We know from Lemma~\ref{lem:stat} that $\ft$
1130: vanishes to order at least two at 0. To compute $\ft''(0)$,
1131: observe first that $\ft'' - \log \gt$ is linear in $\theta$, so
1132: $\ft'' = (\log \gt)''$. When $Z = z e^{i \theta}$, we
1133: have $(d/d\theta) = iZ (d/dZ)$, so
1134: $$\ft'' = i Z {d \over dZ} \left ( iZ {d \log g \over dZ} \right )
1135: = - Z {d \over dZ} \left ( {Z g' \over g} \right ) \, .$$
1136: Expanding this yields
1137: \begin{equation} \label{eq:ft''}
1138: \ft'' = - Z {g' + Z g'' \over g} + {Z^2 (g')^2 \over g^2} \, .
1139: \end{equation}
1140:
1141: By our assumption, $G$ does not vanish at $(z,w)$, so as long as
1142: $\ft'' (0) \neq 0$, we may use Theorem~\ref{th:2d} to
1143: conclude that the leading term asymptotic for $a_{r,s}$ is the
1144: $k=2, l=0$ term of~(\ref{eq:full 2d}). The term $b_0$ there is
1145: equal to
1146: $$\psit(0) \eta' (0) = \psi (z) \sqrt{2 / \ft''(0)} = {\numer (z,w)
1147: \over - w \denom_w (z,w)} \sqrt{2 \over \ft'' (0)} .$$
1148: Thus from Theorem~\ref{th:2d},
1149: $$a_{r,s} \sim {A (2,0) \over 2 \pi} z^{-r} w^{-s} {\numer (z,w)
1150: \over w \denom_w (z,w)} \sqrt{2 \over s \ft'' (0)} .$$
1151: Now evaluate this using the value $A (2,0) = \sqrt{\pi}$ and
1152: equation~(\ref{eq:ft''}) along with~(\ref{eq:g'})
1153: and~(\ref{eq:g''}) to obtain
1154: $$a_{r,s} \sim {1 \over \sqrt{2 \pi}} z^{-r} w^{-s} {\numer (z,w)
1155: \over w \denom_w (z,w)} \sqrt{(-w \denom_w (z,w))^3 \over s Q}$$
1156: where
1157: $$Q = (- w \denom_w (z,w))^3 \ft''(0) = (- w \denom_w (z,w))^3
1158: z {-g'(z) - z g''(z) \over g(z)} + {z^2 (g'(z))^2 \over (g(z))^2} \,
1159: .$$
1160: With the help of Lemma~\ref{lem:der g} we see (using $g(z) = w$)
1161: that
1162: $$Q = (- w \denom_w)^3 \left [ -z {\denom_z \over -w
1163: \denom_w} - z^2 {1 \over -w \denom_w}
1164: \left (\denom_{zz} - 2 {\denom_z \over \denom_w} \denom_{zw}
1165: + {\denom_z^2 \over \denom_w^2} \denom_{ww} \right )
1166: + {z^2 \denom_z^2 \over w^2 \denom_w^2} \right ] \, ,$$
1167: evaluated at $(z,w)$, which simplifies to the expression in
1168: Theorem~\ref{th:simple}. We see also that the nonvanishing
1169: hypotheses on $Q$ is enough to guarantee $\ft''(0) \neq 0$, which
1170: finishes the proof of Theorem~\ref{th:simple}.
1171: \noproof
1172:
1173: \section{Some oscillating integrals} \label{ss:oscillate}
1174:
1175: The oscillating integrals we require are integrals over a
1176: neighborhood of zero in $\R^d$ of the complex-valued integrand:
1177: $$\int_\nbd \exp (- \lambda f (\b{x})) \psi (\b{x}) \, d\b{x}$$
1178: where $f(\zero) = 0, \grad f (\zero) = \zero$ and $\re \{ f
1179: \} \geq 0$. They are not difficult to compute, but since the
1180: standard references assume $f$ is either real or purely imaginary,
1181: we sketch the development of these results. We mostly follow the
1182: exposition of \citeasnoun{stein;oscillatory-integrals}, adapting
1183: it to complex-valued phase functions and simplifying it to take
1184: advantage of the decay of the magnitude of the integrand in this
1185: case.
1186:
1187: We begin with one-dimensional results. Let $\cio$ denote the
1188: class of smooth functions with compact support. The following
1189: proposition is a well known consequence of Watson's Lemma (see,
1190: for example, \citeasnoun[Ch.~2, Theorem~1]{wong;asymptotic-integrals}).
1191: %
1192: \begin{proposition} \label{pr:laplace}
1193: Let $\psi \in \cio (\R)$ and denote $b_j = \psi^{(j)} (0) / j!$.
1194: Then as $\lambda \to \infty$, there is an asymptotic development
1195: $$\int_0^\infty \exp (- \lambda x^k) \psi (x) \, dx \sim
1196: \sum_{l=0}^\infty A_+ (k,l) b_l \lambda^{-(l+1)/k},$$
1197: where, as in~(\ref{eq:plus}),
1198: $$A_+ (k,l) := k^{-1} \Gamma \left ( {l+1 \over k} \right ) .$$
1199: \noproof
1200: \end{proposition}
1201:
1202: We extend this to more general one-sided integrals by a complex
1203: change of variables. Given any analytic, complex-valued function
1204: $f$ on an interval $[0,B]$, suppose that $f(0) = 0$, that
1205: $f' \neq 0$ on $(0,B]$, and let $k \geq 1$ be the minimal so that
1206: $f^{(k)} (0) \neq 0$. Let $\psi \in \cio$ vanishing to order $l
1207: \geq 0$ at 0. Denote $c_j = f^{(j)} (0) / j!$ and $b_j =
1208: \psi^{(j)} (0) / j!$. The real part of $c_k$ is necessarily
1209: nonnegative. Define a function $y$ on $[0,B]$ by
1210: $$y(x) = f(x)^{1/k} = c_k^{1/k} x \left ( 1 + {f(x) - c_k x^k
1211: \over c_k x^k} \right )^{1/k} \, ,$$
1212: where the argument of $c_k^{1/k}$ is between $-\pi/(2k)$ and $\pi
1213: / (2k)$. The quantity $f(x) - c_k x^k$ is $O(x^{k+1})$ near zero,
1214: so $y$ is analytic near 0, and, in particular, is a diffeomorphism
1215: between $[0,B]$ and a contour $\gamma$ from 0 to some $B^*$. Let
1216: $F$ invert $y$. The derivatives of
1217: $F$ at 0 are easy to compute formally and the first $j+1$ starting
1218: from the $k^{th}$ depend only on the first $j$ coefficients of $f$
1219: starting at $c_k$. Define
1220: \begin{eqnarray}
1221: \psi^* & = & (\psi \circ F) \cdot F' \; ; \nonumber \\
1222: b^*_j & = & \psi^*\,^{(j)} (0) / j! \; . \label{eq:bs}
1223: \end{eqnarray}
1224: %
1225: \begin{theorem} \label{th:one-sided}
1226: Let $f$ be analytic (complex-valued) on an interval
1227: $[0,B]$. Assume that $f(0) = 0$, that $f' \neq 0$ on $(0,B]$,
1228: and $\re \{ f \}$ has a strict minimum at 0. Let $k \geq 2$ be
1229: minimal such that $f^{(k)} (0) \neq 0$ and $m$ be minimal so that
1230: the real part of $f^{(m)} (0)$ does not vanish. Let $\psi \in
1231: \cio$, let $l$ be minimal such that $\psi^{(l)} (0) \neq 0$,
1232: and denote $c_j := f^{(j)} (0) / j!$, $b_j := \psi^{(j)} (0) /
1233: j!$. Define $b^*_j$ as in~(\ref{eq:bs}). Then there is an
1234: asymptotic development
1235: \begin{equation} \label{eq:series}
1236: \int_0^B \exp (- \lambda f(x)) \psi (x) \, dx \sim
1237: \sum_{j=l}^\infty A_+ (k,j) b^*_j \lambda^{-(j+1)/k} .
1238: \end{equation}
1239: The constant in the $O(\lambda^{-(N+1)/k})$ term depends continuously (only) on the derivatives of $f$ and $\psi$ up to $(N+1) m / k - 1$.
1240: \end{theorem}
1241:
1242: \begin{proof} Changing variables to $y = f(x)^{1/k}$,
1243: the integral becomes
1244: $$\int_\gamma \exp (- \lambda y^k) \psi^* (y) \, dy ;$$
1245: the curve $\gamma$ is the image of $[0,B]$ under $y$, so
1246: $\gamma' (0) = c_k^{1/k}$ and $\gamma$ remains in the
1247: right half plane, strictly except at 0. For $0 < N < M$ write
1248: $\psi^*$ as $P_M + y^{M+1} R_M$, where $P_M$ is a polynomial of
1249: degree $M$ and $R_M$ is bounded; this can be done since $\psi^*$
1250: may be approximated by a degree $M$ polynomial to within
1251: $O(y^{M+1})$ at 0.
1252:
1253: First, evaluate
1254: $$\int_\gamma \exp (- \lambda y^k) P_M (y) \, dy$$
1255: by moving the contour. Replace $\gamma$ by two line segments, the
1256: first of which goes along the positive real axis to some distance
1257: $\ee$ and the second of which is strictly in the right half plane
1258: (we assumed $\re \{ f \} > 0$ except at 0). The integral along the
1259: second segment is exponentially small since the integrand is.
1260: Hence the combined contribution is the series~(\ref{eq:series})
1261: out to the $j=M$ term.
1262:
1263: Next, bound
1264: $$\left | \int_\gamma \exp (- \lambda y^k) y^{M+1} R_M (y) \, dy
1265: \right | \, .$$
1266: With $C$ representing different constants in different lines, we
1267: now observe that on $\gamma$ we have $\re \{ - y^k \} < - C
1268: |y|^m$. Thus, parametrizing $\gamma$ by arc-length, an upper bound
1269: is given by
1270: $$\int_0^\infty \exp (- \lambda C t^m) t^{M+1} C |R_M (\gamma
1271: (t))| \, dt .$$
1272: This is easily seen to be bounded above by
1273: $C \lambda^{-(M+2)/m}$ where $C$ depends on the first $M$
1274: derivatives of $f$ and $\psi$. Choosing $M \geq m(N+1)k - 1$ we
1275: have a remainder term that is $O(\lambda^{-(N+1)/k})$, proving the
1276: theorem.
1277: \end{proof}
1278:
1279: The value of a two-sided integral follows as a corollary.
1280: %
1281: \begin{corollary} \label{cor:full 2d}
1282: Assume the hypotheses of Theorem~\ref{th:one-sided}, with $f$ now
1283: defined on an interval $[-B,B]$. Then there is an asymptotic
1284: development
1285: \begin{equation}\label{eq:two-sided}
1286: \int_{-B}^B \exp (- \lambda f(x)) \psi (x) \, dx \sim
1287: \sum_{j=l}^\infty A (k,j) b^*_j \lambda^{-(j+1)/k} .
1288: \end{equation}
1289: with $A (k,j)$ given by~(\ref{eq:k odd}) and~(\ref{eq:k even}).
1290: The bounds on the remainder terms each depend continuously on finitely many
1291: derivatives of $f$ and $\psi$ on $[-B,B]$.
1292: \end{corollary}
1293:
1294: \begin{proof} The two-sided integral is the sum of two
1295: one-sided integrals on intervals $[0,B]$ and $[-B,0]$. The
1296: integral over $[-B,0]$ may be written as an integral over $[0,B]$
1297: of the function $\exp (- \lambda f(-x)) \psi (-x) \, dx$. With
1298: $b_l^*$ still denoting the coefficients resulting form the application
1299: of Theorem~\ref{th:one-sided} to the first integral, let
1300: $\check{b_l^*}$ denote the coefficients when
1301: Theorem~\ref{th:one-sided} is applied to the second integral. In
1302: order to add the two integrals, we write $\check{b_l^*}$ in terms
1303: of $b_l^*$ by means of the following routine computation.
1304:
1305: Let $c_k := c^k e^{i \alpha}$ with $c > 0$ and $|\alpha| \leq \pi
1306: / 2$ and define the analytic quantity $R$ so that
1307: $$y(x) = \left [ c_k x^k (1 + R(x))^k \right ]^{1/k} =
1308: c e^{i \alpha / k} x (1 + R(x)) .$$
1309: If $k$ is odd, then then the hypothesis $\re \{ f
1310: \} \geq 0$ implies that $c_k$ is purely imaginary. We have
1311: $$\check{y} (x) = \left [ - c_k x^k (1 + R(-x))^k \right ]^{1/k} =
1312: c e^{- i \alpha / k} x (1 + R(-x)) = - y(-x) e^{- 2 i \alpha / k}
1313: .$$
1314: Writing $\finv$ for the inverse function to $y$ and $\check{\eta}$
1315: for the inverse function to $\check{y}$ we then have
1316: $$\check{\finv} (x) = - \finv (-e^{2 i \alpha / k} x) \, .$$
1317: Hence, letting $\C_l [ \cdot ]$ denote the coefficient of
1318: $y^l$,
1319: \begin{eqnarray*}
1320: \check{b_l^*} & = & \C_l \left [ \psi ( - \check{\finv} (x))
1321: \cdot \check{\finv}' (x)\right ] \\[1ex]
1322: & = & \C_l \left [ \psi ( \finv (-e^{2 i \alpha / k} x))
1323: \cdot e^{2 i \alpha / k} \cdot \finv' (-e^{2 i \alpha / k} x) \right ]
1324: \end{eqnarray*}
1325: and thus
1326: $$\check{b_l^*} = (-1)^l e^{2 i \alpha (l+1) / k} b_l^* \, .$$
1327: When $k$ is even, the computation is similar but easier, resulting
1328: in
1329: $$\check{b_l^*} = (-1)^l b_l^* \, .$$
1330: Now observe that if $k$ is odd, hence $c_k$ is purely imaginary,
1331: then $e^{2 i \alpha (l+1) / k} = e^{\pm i \pi (l+1) / k}$
1332: according to the sign of the argument of $c_k$. Setting $A (k,l)
1333: = (1 + (-1)^l) A_+ (k,l)$ if $k$ is even and
1334: $(1 + e^{\sign\Arg (c_k) i \pi (l+1) / k} A_+ (k,l)$
1335: if $k$ is odd, we recover the definition in~(\ref{eq:k odd})
1336: and~(\ref{eq:k even}) and prove the Corollary.
1337: \end{proof}
1338: %
1339: \begin{theorem} \label{th:nondeg}
1340: Let $f$ be a smooth complex-valued function on a neighborhood of
1341: $\zero$ in $\R^d$ such that $\re \{ f \} \geq 0$ with equality only at $\zero$. Suppose further that $\grad f(\b{0})=0$, and that the Hessian (matrix of second partials) of $f$ has eigenvalues with positive real parts. Let $\hess$ denote the Hessian determinant at 0. Then for $\psi \in \cio$, there is an asymptotic expansion
1342: $$\int \exp (- \lambda f (\b{x})) \psi (\b{x}) \, d\b{x} \sim
1343: \sum_{j \geq l} C_j \lambda^{-(l+d)/2}$$
1344: where $l$ is the degree of vanishing of $\psi$ at $\zero$. If $l =
1345: 0$ then $C_0 = \psi (\zero) (2 \pi)^{d/2} \hess^{-1/2}$. The
1346: choice of square root is determined by $\hess^{-1/2} =
1347: \prod_{j=1}^d \mu_j^{-1/2}$ where $\mu_j$ are the eigenvalues of
1348: the Hessian and the principal square root is taken in each case.
1349: \end{theorem}
1350:
1351: \begin{proof} Let $Q = \sum_{i,j=1}^d q_{i,j} z_i z_j$
1352: be the quadratic form determined by the Hessian at the origin.
1353: Denote the eigenvalues of $Q$ by $\{ \mu_j : 1 \leq j \leq d \}$
1354: and note that each $\mu_j$ has nonnegative real part.
1355:
1356: Step 1: change coordinates to make $f$ exactly equal to the
1357: quadratic form $Q$. Indeed since $f(\b{x}) = Q(\b{x})/2 + O(|\b{x}|^3)$,
1358: and the Hessian is nondegenerate, there is a locally smooth change
1359: of variables $\{ x_j (\b{z}) : 1 \leq j \leq d \}$ such that $f(\b{z})
1360: = Q(\b{x} (\b{z}))/2$ and the Jacobian at the origin is 1.
1361:
1362: Step 2: normalize by $\hess^{1/2}$. For any quadratic form $Q$
1363: there is a linear change of variables $\b{y} (\b{x})$ such that
1364: $Q(\b{x}) =
1365: \sum_{j=1}^d y_j^2$. The change of variables matrix $P$
1366: satisfies $P P^T = M(Q)$, the symmetric matrix representing $Q$.
1367: Changing variables to $\b{y}$ introduces an integrating factor of
1368: $\det P$ which is a square root of $\hess$ since $M(Q)$ is just the
1369: Hessian. Let $\nbd'$ be the region of integration over which
1370: $\b{y}$ varies when $\b{z}$ varies over an appropriately small
1371: neighborhood of $\zero$.
1372:
1373: Step 3: Expand $\psit$ into monomials. The function $\psi$ has
1374: now become $\psit$, where $\psit(\zero) = \hess^{-1/2} \psi
1375: (\zero)$ and the sign of the square root will be chosen later. We
1376: may expand $\psit$ into monomials, using the same argument as in
1377: the proof of Theorem~\ref{th:one-sided} to show the remainder term
1378: can be made
1379: $O(|\b{y}|^N)$ for any $N$. It remains to evaluate the integral over the
1380: region of integration, $\nbd'$ of
1381: $$\int_{\nbd'} \exp (- \lambda \sum_{j=1}^d y_j^2) \psit (\b{y}) \, d\b{y}$$
1382: when $\psit$ is a monomial.
1383:
1384: Step 4: move the region of integration to the real $d$-space. Let
1385: $\nbd''$ be the projection of $\nbd'$ onto $\R^d$ by setting the
1386: imaginary part to zero. We claim that changing the region of
1387: integration from $\nbd'$ to $\nbd''$ alters the integral by an
1388: amount rapidly decreasing in $\lambda$. To show this, let
1389: $\Omega$ be the region $\{ \re \{ \b{x} \} + i t \im \{ \b{x} \} :
1390: \b{x} \in \nbd' , t \in [0,1] \}$. The boundary of $\Omega$
1391: (as a manifold) is composed of $\nbd', \nbd''$
1392: (with opposite signs) together with $S := \{ \re \{ \b{x} \} + i t
1393: \im \{ \b{x} \} : \b{x} \in \partial \nbd' , t \in [0,1] \}$. For any
1394: $d$-form $\omega$, $\int_\Omega d\omega = \int_{\partial \Omega}
1395: \omega$. When $\omega = \exp (- \lambda \sum_{j=1}^d \mu_j y_j^2)
1396: \b{y}^\b{r} dy_1 \wedge \cdots \wedge dy_d$ is a holomorphic $d$-form,
1397: we see that $d\omega$ vanishes (being the sum of $\partial /
1398: \partial \overline{z_j}$ terms) so that
1399: $$\int_{\nbd'} \omega = \int_{\nbd''} \omega + \int_S \omega .$$
1400: We know that $\re \{ \sum_j \mu_j y_j^2 \}$ is bounded away from 0
1401: on $\partial \nbd'$, and its minimal value on $S$ lies on
1402: $\partial \nbd'$, hence the integral over $S$ decays
1403: exponentially.
1404:
1405: Step 5: evaluate the integral. Factoring $\int_{\nbd''} \b{y}^\b{r}
1406: \exp (- \lambda \sum_{j=1}^d y_j^2)$ into one-dimensional
1407: integrals and plugging into Proposition~\ref{pr:laplace} yields an
1408: asymptotic expansion whose leading term (when $l=0$) is equal to
1409: $(2 \pi)^{d/2} \psi (\zero) \hess^{-1/2}$. When $f (\b{z})$ is the function $\sum_{j=1}^d z_j^2$, then the positive square root is taken. The
1410: choice of square root must be continuous in the analytic topology
1411: on functions having nondegenerate Hessians and having eigenvalues
1412: with positive real parts, and the only such choice is the product
1413: of the principal square roots of the eigenvalues of the Hessian.
1414: \end{proof}
1415:
1416: \section{Classification of cases} \label{ss:classify}
1417:
1418: For purposes of classification some natural questions are:
1419: \begin{enumerate}
1420: \romenumi
1421: \item what are all possible local geometries of minimal points of $\sing$?
1422: \item which of these can be handled by variants of the methods in this paper?
1423: \item are these sufficient to yield a good approximation to $a_\b{r}$ no
1424: matter what the direction, $\b{r} / |\b{r}|$, and no matter which
1425: generating function in the class, say, of functions meromorphic in
1426: a neighborhood of their domain of convergence?
1427: \end{enumerate}
1428: To make the last question more concrete, consider the simplest
1429: possible example, namely binomial coefficients, where $F =
1430: 1/(1-z-w)$ and $\sing$ is a complex line. There are no singular
1431: points here, but how do we know that as $(z,w)$ varies over
1432: minimal points of $\sing$, the direction $\direc (z,w)$ will cover
1433: all of $\RP^1$?
1434:
1435: This question will be answered by Theorem~\ref{th:classify}, but
1436: first we need to add some detail to the geometric discussion begun
1437: in Section~\ref{ss:geom}. It will be evident that quite a few
1438: cases need to be considered, some of which require new tools and
1439: some of which require only minor modifications. Accordingly, the
1440: results will appear in several papers, currently under
1441: preparation. In other words, a discussion of taxonomy will
1442: necessarily refer to results not yet published, and we will
1443: indicate to the best of our knowledge which ones are expected to
1444: be routine.
1445:
1446: Given a point $\b{z} \in \sing$, we extend the definition of
1447: $\direc (\b{z})$ to mean the set of limits of $\direc (\b{y})$ as $\b{y}
1448: \to \b{z}$ along smooth points. When $\b{z}$ is minimal, this is just
1449: the set of normals to support hyperplanes of $\logdom$ at the
1450: point $(\log |z_1| , \ldots , \log |z_d|)$, so this is consistent
1451: with the old definition. As we will see shortly, $\direc (\b{z})$
1452: may be a $(d-1)$-dimensional subset of $\RP^{d-1}$ when $\b{z}$ is a
1453: critical point of $\sing$.
1454:
1455: When $\denom$ has a repeated factor, the residue computation in
1456: equation~(\ref{eq:defres}) must be replaced by one involving the
1457: derivative. The remainder of the computation proceeds without a
1458: hitch as before. Details are given in \citeasnoun{pemantle-wilson;toral}.
1459: For the remainder of the taxonomy, we assume $\denom$ to
1460: be square-free. Toral smooth points may be handled by methods
1461: exactly the same as strictly minimal points. The inner integrand
1462: in~(\ref{eq:iter}) will in this case have its maximal modulus on a
1463: set of dimension larger than zero. A modification of the
1464: necessary oscillating integral computation that works in this case
1465: is also given in \citeasnoun{pemantle-wilson;toral}.
1466:
1467: If $\b{z} \in \sing$ is not smooth, all the first partials vanish. The
1468: expansion of $\denom (\b{x})$ near $\b{z}$ is then a sum of terms of
1469: degrees 2 and higher. We call $\b{z}$ a {\em homogeneous point} of
1470: degree $k$ if this expansion contains terms
1471: $(x_j - z_j)^k$ for each $j = 1 ,\ldots , d$, and contains no
1472: terms of total degree less than $k$.
1473:
1474: \begin{lemma} \label{lem:homogeneous}
1475: If $\b{z}$ is a locally minimal point of $\sing$ with nonzero
1476: coordinates, and $F$ is meromorphic in a neighborhood of $\b{z}$
1477: then $\b{z}$ is homogeneous.
1478: \end{lemma}
1479:
1480: \begin{proof} Passing to $F(z_1 x_1 , \ldots , z_d x_d)$ if
1481: necessary, we may assume $\b{z} = \one$. Setting $x_j = 1$ for all
1482: but one index $j$, we cannot obtain the zero function (by
1483: minimality), and so some term in the expansion around $\one$ is a
1484: pure power of $(x_j - 1)$, and we denote the minimal degree such
1485: term by $c_j (x_j - 1)^{k_j}$. If $\b{z}$ is not a homogeneous
1486: point, then there is some $j$ for which some monomial has total
1487: degree lower than $k_j$. Assume without loss of generality that
1488: $j=d$. The function $F(x , x , \ldots , x , y)$ then has a
1489: minimal degree pure $y-1$ term $c_0 (y-1)^k$,
1490: $k := k_d$, and some term $c' (x-1)^a (y-1)^b$ with $a+b < k$.
1491: In other words, the Newton Polygon of $F(x, \ldots , x , y)$
1492: around $(1,1)$ has a support line passing through $(0,k)$ with
1493: slope $-p/q$ in lowest terms, and $p > q$. It is well known that
1494: we may describe the solutions $y(x)$ of the equation
1495: $$F(1+x , \ldots , 1+x , 1 + y) = 0$$
1496: as follows. Write
1497: $$H := (y-1)^k (c_0 + c_1 (y-1)^{-p} (x-1)^q + c_2(y-1)^{-2p} (x-1)^{2q}
1498: + \cdots + c_s (y-1)^{-sp} (x-1)^{sq})$$
1499: for the polynomial collecting all the terms on this support line.
1500: Then for each $q^{th}$ root of unity, $\omega$, and each root
1501: $\lambda$ of $\sum c_{s-j} \lambda^j = 0$, there is a solution
1502: $y = \lambda^{1/p} x^{q/p} (\omega + o(1))$ as $x \to 0$.
1503: A proof may be found in \citeasnoun{brieskorn-knorrer}.
1504:
1505: Varying $x$ over the set $|\pi - \arg (x)| \leq \pi / 4$, we see
1506: that the solutions $y(x)$ must sometimes be in this set as well.
1507: For those $x$, the points $(1+x , \ldots , 1+x , 1+y)$ will be in
1508: $\sing \cap \disk (\one) \setminus \torus (\one)$, violating
1509: minimality of $\one$. By contradiction, we have shown that no
1510: monomial in the expansion around $\one$ has lower total degree
1511: than any pure power term, hence $\one$ is minimal. \end{proof}
1512:
1513: Continuing the taxonomy, suppose that $\b{z}$ is a homogeneous point
1514: of $\sing$ of degree $k \geq 2$. We say that $\b{z}$ is a {\em
1515: multiple point} if $\sing$ is locally the union of $k$ analytic
1516: surfaces. Algebraically, this means that the leading (order $k$)
1517: terms in the expansion of $\denom$ near $\b{z}$ factors into linear
1518: pieces. If the homogeneous point $\b{z}$ is not a multiple point,
1519: we say it is a {\em cone point}. When $d=2$ there are no cone
1520: points, since any homogeneous polynomial in 2 variables factors
1521: completely over $\CC$.
1522:
1523: Our understanding of cone points is not yet complete, but an
1524: analysis involving cone points is underway in \citeasnoun{cohn-pemantle;fixation}.
1525: For multiple points, most of the story is given in
1526: \citeasnoun{pemantle-wilson;multiple}. In particular, the following theorem
1527: is proved there.
1528: %
1529: \begin{theorem}[\citeasnoun{pemantle-wilson;multiple}]
1530: \label{th:PW2000a}
1531: Let $\b{z}$ be an isolated, minimal, multiple point of $\sing$ with
1532: multiplicity $k$. Let $S \subseteq \RP^{d-1}$ be the set of
1533: outward normals to support hyperplanes to $\logdom$ at the point
1534: $(\log |z_1| , \ldots , \log |z_d|)$. Then there is an integer $p \geq 0$
1535: and a polynomial function $\phi : S \to \R$ such that the
1536: asymptotic expansion
1537: \begin{equation}\label{eq:PW2000a}
1538: a_\b{r} \sim \b{z}^{-\b{r}} \phi (\b{r}) \sum_j C_j (r_d)^{k-p/2-j/2}
1539: \end{equation}
1540: holds uniformly as $\b{r}$ varies over compact subsets of the
1541: interior of $S$.
1542: \noproof
1543: \end{theorem}
1544:
1545:
1546: The extension to toral multiple points is given in
1547: \citeasnoun{pemantle-wilson;toral}. If a multiple point is not
1548: isolated or toral, then the degree of multiplicity, $k$, must be
1549: less than the dimension,
1550: $d$. This cannot happen of course when $d=2$, but does happen
1551: when $d \geq 3$. The method for handling this case, toral or
1552: otherwise, is given in \citeasnoun{pemantle-wilson;toral}. That paper
1553: will also contain some subcases of the isolated multiple point
1554: case, namely when the sheets of $\sing$ intersect non-transversely.
1555:
1556: Having more or less completed the taxonomy, we now discuss when we
1557: can guarantee that our methods yield asymptotics in all
1558: directions.
1559: %
1560: \begin{theorem} \label{th:classify}
1561: Let $F = \numer / \denom = \sum a_{r,s} z^r w^s$ be the quotient
1562: of analytic functions $\numer , \denom : \CC^2 \to \CC$. Suppose
1563: that the coefficients $a_{r,s}$ are all nonnegative, and that
1564: $F(z,0)$ and $F(0,w)$ are not entire. Then for every direction
1565: $\alpha \in \RP^1$ there is a minimal $\b{z} \in \sing$ with
1566: $\alpha \in \direc (\b{z})$.
1567: \end{theorem}
1568:
1569: \begin{proof} Let $(x,y)$ be any point on the boundary of
1570: $\logdom$. For $u < e^x$ and $v < e^y$ the power series for $F$
1571: is convergent at $(u,v)$. As $u \uparrow e^x$ and $v \uparrow
1572: e^y$ therefore, $F(u,v)$ is finite and increasing. On the other
1573: hand, the power series for $F$ is not absolutely convergent on
1574: $\torus (e^x , e^y)$, since we know $F$ to have some singularity
1575: on this torus. Hence $F(u,v) \uparrow \infty$ as $(u,v) \uparrow
1576: (e^x,e^y)$. Since $F$ is meromorphic, it must have a pole at
1577: $(e^x , e^y)$, hence $(e^x , e^y) \in \sing$ and is a minimal point of $\sing$.
1578: As $(x,y)$ varies over the boundary of $\logdom$, we let $\gamma
1579: \subseteq \sing$ denote the curve traced out by this minimal point.
1580:
1581: Pick any $\alpha \in \RP^1$. The convex set $\logdom$ has
1582: horizontal and vertical support hyperplanes (by non-entirety of
1583: $F(z,0)$ and $F(0,w)$), and therefore has a support hyperplane
1584: normal to $\alpha$; let $(x,y)$ be a point of intersection of this
1585: support plane with $\logdom$. We have just seen that $\b{z} (\alpha)
1586: := (e^x , e^y)$ is a minimal point of $\sing$. If $\b{z}$ is a smooth
1587: point of $\sing$ then $\alpha \in \direc (\b{z})$: either $\b{z}$ is
1588: finitely minimal, in which case Theorem~\ref{th:2d} applies, or it
1589: is toral, in which case the toral version of this theorem from
1590: \citeasnoun{pemantle-wilson;toral} applies.
1591:
1592: Assume now that $\b{z}$ is not a smooth point. By
1593: Lemma~\ref{lem:homogeneous}, $\b{z}$ is a homogeneous point, and
1594: since $d=2$, $\b{z}$ is a multiple point. Theorem~\ref{th:PW2000a}
1595: then shows that $\alpha \in \direc (\b{z})$ in this case as well.
1596: This finishes the proof.
1597: \end{proof}
1598:
1599: \section{Further details and open questions} \label{ss:open}
1600:
1601: The theorems in this and subsequent papers give estimates that are
1602: uniform away from the boundary of the domain in which they are
1603: valid. In order for all of these to be patched together so as to
1604: give estimates valid now matter how $\b{r} \to \infty$, one must
1605: determine the bandwidth around the boundary for which the boundary
1606: estimates on either side hold. For instance, suppose $(z,w)$ is a
1607: multiple point of degree 2 and that $\direc (z,w)$ is the set of
1608: slopes between $1/2$ and 2. It appears that the asymptotic
1609: estimate in \citeasnoun{pemantle-wilson;multiple} holding near the line $\{
1610: s = 2r \}$ can be written so it is valid out to $s = 2r + c
1611: \sqrt{r}$. If the estimate for the region $s/r > 2 + \ee$ can be
1612: widened so it holds to $s = 2r + c \sqrt{r}$ and a description
1613: given that is valid in the regime $(s-2r)/\sqrt{r} \to c$, then
1614: the estimates will patch together completely.
1615:
1616: Another natural question is the universality of the method when
1617: the coefficients have mixed signs. We conjecture that
1618: Theorem~\ref{th:classify} still holds, in the sense that for every
1619: direction there is point $\b{z} \in \sing$ for which integration near
1620: $\b{z}$ yields correct asymptotics. What we know is that $\b{z}$ may
1621: no longer be minimal. For example, if $\numer = 1$ and
1622: $$\denom = (1 - (2/3) w - (1/3) z) (1 + (1/3) w - (2/3) z )$$
1623: then the point $(3/2 , 3/4)$ is not minimal but yields asymptotics
1624: in the diagonal direction; one sees this by integrating along a
1625: deformed torus rather than along $\torus (3/2 , 3/4)$. In fact we
1626: conjecture that such a deformation always exists, but the topology
1627: seems not transparent enough to yield an easy proof.
1628:
1629: The class of algebraic functions is in some ways almost as nice as
1630: the set of rational functions, and nicer than the meromorphic
1631: functions. For one thing, an algebraic function is determined by
1632: a finite amount of data, and may thus easily be input into a
1633: symbolic math package.
1634: \citeasnoun{gao-richmond;limit-multivariate4} give an analysis of
1635: algebraic and logarithmic singularities, but sometimes the
1636: relevant singularities for algebraic functions are poles. For
1637: example, in Larsen and Lyons' analysis
1638: \cite{larsen-lyons;coalescing} of merge times for coalescing
1639: particles, they find an algebraic function of the form
1640: $$F(z,w) = {\chi (z,w) \over w - 1 - \sqrt{1-z}}$$
1641: with $\chi$ analytic. The branch of the square root is chosen so
1642: that at the origin the denominator is 2, not 0. There is a
1643: branchline at $z=1$, but for all directions in $\RP^1$, there is a
1644: smooth pole on the curve $w = 1 + \sqrt{1-z}$ yielding asymptotics
1645: in the desired direction. It is natural to ask when this will
1646: happen, and how one can tell effectively. Some questions of
1647: effectiveness are addressed in \citeasnoun{pemantle-wilson;multiple} and
1648: \citeasnoun{pemantle-wilson;toral},
1649: but there is probably substantial room for improvements on an
1650: algorithmic level.
1651:
1652: \noindent{\bf Acknowledgements:} We thank Steve
1653: Wainger for his generous help, including pointing out
1654: equation~(\ref{eq:X}). Thanks also to Andreas Seeger for his help
1655: with the material in Section~\ref{ss:oscillate}. We are indebted
1656: to Manuel Lladser for many helpful comments on an earlier draft.
1657:
1658: \bibliographystyle{agsm}
1659: \bibliography{asymp}
1660: \end{document}
1661: