math0503198/art.tex
1: \documentclass[12pt]{article}
2: 
3: \usepackage{amsmath,a4wide,amssymb,amsthm}
4: \usepackage{graphicx}
5: 
6: \newtheorem{theorem}{Theorem}
7: \newtheorem{lemma}[theorem]{Lemma}
8: \newtheorem{prop}[theorem]{Proposition}
9: \newtheorem{cor}[theorem]{Corollary}
10: \newtheorem{defi}[theorem]{Definition}
11: \newtheorem{ass}[theorem]{Assumption}
12: 
13: \numberwithin{equation}{section}
14: \numberwithin{theorem}{section}
15: 
16: \newcommand{\bs}{\boldsymbol}
17: 
18: \title{\bf On $q$-functional equations\\
19: and excursion moments
20: }
21: 
22: \author{\sc Christoph~Richard\\
23: \\
24: {\it Fakult\"at f\"ur Mathematik, Universit\"at Bielefeld,}\\
25: {\it Postfach 10 01 31, 33501 Bielefeld, Germany}}        
26: 
27: \begin{document}
28: 
29: \maketitle
30: 
31: \begin{abstract}
32: We analyse $q$-functional equations arising from tree-like combinatorial 
33: structures, which are counted by size, internal path length, and certain 
34: generalisations thereof. The corresponding counting parameters are labelled
35: by a positive integer $k$. We show the existence of a joint limit distribution 
36: for these parameters in the limit of infinite size, if the size generating 
37: function has a square root as dominant singularity. The limit distribution 
38: coincides with that of integrals of $k$-th powers of the standard Brownian 
39: excursion. Our approach yields a recursion for the moments of the limit
40: distribution. It can be used to analyse asymptotic expansions of the 
41: moments, and it admits an extension to other types of singularity.
42: \end{abstract}
43: 
44: Keywords: simply generated trees, $q$-difference equation, Brownian excursion, 
45: limit distribution
46: 
47: \section{Introduction and main results}
48: 
49: \subsection{A combinatorial motivation}
50: 
51: When studying combinatorial classes, a functional equation of the form
52: %
53: \begin{equation}\label{form:geneq}
54: G(x)=P(x,G(x))
55: \end{equation}
56: %
57: frequently arises, where $G(x)$ is the generating function of the
58: class, and $P(x,y)$ is a formal power series in two variables with real
59: coefficients.
60: Prominent examples are classes of simply generated trees, counted by 
61: number of vertices \cite{MM78}, classes of directed square lattice paths 
62: counted by length \cite{BF02}, and classes of square
63: lattice polygons, counted by perimeter \cite{Bou96}. Indeed, there exist 
64: combinatorial bijections between corresponding models of trees, paths, 
65: and polygons, see e.g.~\cite{S99} for Catalan trees, Dyck paths, 
66: and staircase polygons. Eqn.~\eqref{form:geneq} reflects a combinatorial 
67: decomposition of the given class. If $P(0,y)\equiv0$ and $\frac{\partial P}{\partial y}
68: (0,y)\not\equiv1$, then eqn.~\eqref{form:geneq} admits a unique solution 
69: $G(x)\in\mathbb R[[x]]$ such that $G(0)=0$. Here, $\mathbb R[[x]]$ denotes the 
70: ring of formal power series in $x$ with coefficients in $\mathbb R$. Often, the 
71: coefficients of $P(x,y)$ are non-negative real numbers. Then, the series $G(x)$ is a power 
72: series with non-negative coefficients, typically analytic at $x=0$ with a
73: square root as dominant singularity, see e.g.~\cite[Thm.~10.6]{O95}, \cite{D97},
74: or \cite[Ch.~VII.4]{FS07}, and references therein.
75: 
76: \smallskip
77: 
78: We are interested in certain deformations of the above equation. This is done by 
79: introducing a new variable $q$, such that the limit $q\to1$ reduces to
80: the original equation. For example, the functional equation 
81: %
82: \begin{equation}\label{form:geneq2}
83: G(x,q)=P(x,G(qx,q))
84: \end{equation}
85: %
86: defines a formal power series in $x$ with polynomial coefficients, i.e., 
87: $G(x,q)\in\mathbb R[q][[x]]$, if eqn.~\eqref{form:geneq} defines a power series 
88: $G(x)\in\mathbb R[[x]]$. The above equation is a \emph{$q$-difference equation},
89: see e.g.~\cite{VRSZ03} and references therein. It appears 
90: in classes of simply generated trees, counted by number of vertices and internal 
91: path length (i.e., the sum of the vertex distances to the root), 
92: in classes of directed square lattice paths, counted 
93: by length and area under the path, and in classes of square
94: lattice polygons, counted by perimeter and area. For some models, 
95: an explicit expression for its generating function $G(x,q)$ is known, see 
96: e.g.~\cite{Bou96,PB95a,PO95} and references therein. Such an expression 
97: typically contains $q$-products and has a natural boundary $|q|=1$. 
98: An interesting question concerns the statistics of the additional 
99: counting parameter, e.g., in a uniform ensemble in the limit of large system size. 
100: It is known (\cite{T91, D99}, see also \cite{R02}) that eqn.~\eqref{form:geneq2} leads, for certain 
101: $q$-difference equations and after appropriate normalisation, to the \emph{Airy 
102: distribution} as the limit distribution for the 
103: additional parameter. This distribution is known to also describe the area under a 
104: Brownian excursion, see the following subsection. Note that generalisations 
105: of eqn.~\eqref{form:geneq} other than eqn.~\eqref{form:geneq2} have also been 
106: studied previously. A class of equations, 
107: which leads to Gaussian limit laws, is discussed in \cite{D97}. 
108: 
109: 
110: \smallskip
111: 
112: The idea of iterating the above deformations has been considered by Duchon 
113: \cite{D98,D99}. The deformation variables may be denoted by $q_k$, where 
114: $k\in\{1,\ldots,M\}$, and the associated counting parameters 
115: are called \emph{parameters of rank $k+1$}. For eqn.~\eqref{form:geneq2}, 
116: an example is given by
117: %
118: \begin{equation}\label{form:geneq3}
119: G(x,q_1,\ldots,q_M)=P(x,G(xq_1\cdot\ldots\cdot q_M,q_1q_2\cdot\ldots\cdot q_M,
120: q_2q_3\cdot\ldots\cdot q_M,\ldots,q_M)).
121: \end{equation}
122: %
123: Here, $G(x,q_1,\ldots, q_M)=\sum_n p_n(q_1,\ldots, q_M) x^n$ is a formal power
124: series in $x$, with polynomial coefficients $p_n(q_1,\ldots, q_M)\in\mathbb 
125: R[q_1,\ldots, q_M]$. The name ``parameter of rank $k+1$'' reflects that 
126: $k$ is the smallest integer $r$, such that the degree of the polynomial 
127: $p_n(q_1,\ldots, q_M)$ in $q_k$ is bounded by 
128: $c\,n^{r+1}$ for some constant $c$ (see \cite[Lemma~1]{D99} and \cite{D98}). 
129: We will call an equation like eqn.~\eqref{form:geneq3} a \emph{$q$-functional equation} 
130: (Definition \ref{defi:qfunc}). 
131: Again the question arises, under which conditions a limit distribution for the additional
132: counting parameters exists. 
133: In this paper, we shall show this for a class of deformations 
134: which we call \emph{$q$-shift} (Definition \ref{def:qshift}), the main assumption 
135: on the $q$-functional equation being that the solution of the undeformed equation, 
136: eqn.~\eqref{form:geneq}, is analytic at the origin, with a square root as dominant singularity 
137: (Assumption \ref{ass}). See Theorem~\ref{theo:probdist} for a precise statement. The 
138: resulting limit distributions appear to be related to distributions of integrals of $k$-th 
139: powers of the standard Brownian excursion. We will obtain a recursion for the moments 
140: of the joint distribution.  Our approach is based on the multivariate \emph{moment method}
141: (see e.g.~\cite{JLR00, Bill95}). The univariate case 
142: $M=1$ has been studied previously \cite{T91, FPV98, D99, R02}, and recursions for $M=2$ have been 
143: studied in \cite{NT03b}.
144: 
145: \smallskip
146: 
147: Before we consider general $q$-functional equations in Section~\ref{sec:qfunc}, let us 
148: first discuss the above questions in more detail for the simple example of Dyck paths.
149: 
150: \subsection{Dyck paths and Brownian excursions}\label{sec:Dyck}
151: 
152: We review the connection between Dyck paths and Brownian excursions.
153: This relates, in particular, the moments of height
154: of a random Dyck path to Brownian excursion moments.
155: Since the generating function of Dyck paths, counted by length and moments 
156: of their height, provides a simple example of a solution of a  $q$-functional 
157: equation, this also motivates the appearance of excursion moments in 
158: $q$-functional equations. We will state a central result of the paper in 
159: Theorem~\ref{theo:excmom}, which characterises the Brownian excursion random 
160: variables by a recursion for their moments.
161: 
162: \medskip
163: 
164: Let $\mathbb N_0=\mathbb N \cup \{0\}$ and $\mathbb R_{\ge0}=\{x\in\mathbb 
165: R: x\ge0\}$. A {\it Dyck path of length $2n$}, where $n\in \mathbb N_0$, is a map 
166: $y:[0,2n]\to\mathbb R_{\ge0}$, where $y(0)=y(2n)=0$ and $|y(i)-y(i+1)|=1$ 
167: for $i\in\mathbb N_0$ and $i<2n$. For non-integral argument, $y(s)$ is 
168: defined by linear interpolation. The values $y(s)$ are called the \emph{heights} 
169: of the path. An {\it arch of length $2n$} is a Dyck path of length $2n$, where $n>0$ and 
170: $y(s)>0$ for $s\in (0,2n)$. An example is given in Figure \ref{fig:dyck}.
171: %
172: \begin{figure}[htb]
173: \begin{center}
174: \begin{minipage}[b]{0.65\textwidth}
175: \center{\includegraphics[width=8cm]{dyck}}
176: \end{minipage}
177: \end{center}
178: \caption{\label{fig:dyck}
179: \small A Dyck path of length $2n=10$. It is a sequence of two arches of lengths 8 and 2.}
180: \end{figure}
181: %
182: 
183: \medskip
184: 
185: We are interested in a probabilistic interpretation of Dyck paths in a uniform
186: ensemble where, for fixed length, each of the finitely many Dyck paths occurs with 
187: the same probability. For $0\le s\le 2n$, let $\widetilde Y_{n}(s)$ denote the height 
188: of a random Dyck path of length $2n$. It is well known, see e.g.~\cite{BKR72}, that 
189: the average maximal height $\widetilde h(n)$ of a random Dyck path of length $2n$ has 
190: the asymptotic form $\widetilde h(n)\sim \sqrt{\pi n}$ as $n\to\infty$. 
191: In order to obtain a finite positive limit as $n$ approaches infinity, 
192: and in order to normalise the domain, we introduce the normalised height 
193: $Y_{n}(t)=(2n)^{-1/2}\widetilde Y_{n}(2nt)$, where $0\le t\le 1$. 
194: 
195: The sequence $\{Y_{n}(t)\}_{n\in\mathbb N}$ is a sequence of stochastic 
196: processes defined on $(C[0,1],{||\cdot ||_{\infty}})$ with the Borel 
197: $\sigma$-algebra, which converges in distribution to the standard Brownian 
198: excursion $e(t)$ of duration $1$, see \cite{Bel72,A93, MM03}. 
199: This implies convergence in distribution of sequences of continuous bounded 
200: functionals of $Y_n(t)$ towards the corresponding excursion functionals. 
201: (We refer to \cite{B99} for background about convergence of probability measures.)
202: There 
203: is a more recent result \cite[Thm.~9]{D03} that also
204: %
205: \begin{equation*}
206: \lim_{n\to\infty}\mathbb E_n[F(Y_{n}(t))] = \mathbb E[F(e(t))]
207: \end{equation*}
208: %
209: for continuous functionals $F:C[0,1]\to\mathbb R$ of \emph{polynomial} growth, 
210: i.e., for functionals such that there exists an $r\ge0$ with $|F(y)|\le 
211: ||y||_\infty^r$ for all $y\in C[0,1]$. The above property is called \emph{polynomial 
212: convergence}. In particular, polynomial convergence implies
213: \emph{moment convergence} for such a functional, i.e., convergence of the
214: sequence of moments of order $l$, $\{\mathbb E_n[F(Y_{n}(t))^l]\}_{n\in
215: \mathbb N}$, for every $l\in\mathbb N_0$.  
216: For the functional of polynomial growth $F(y)=\int_0^1 y^k(t) {\rm d}t$, 
217: $k\in\mathbb N$, we call the random variable $\int_0^1e^k(t){\rm d}t$ 
218: the \emph{$k$-th excursion moment}.
219: 
220: \medskip
221: 
222: As counting parameters for Dyck paths of length $2n$, we consider the parameters 
223: %
224: \begin{equation}\label{form:dyckpara}
225: x_{k,n}=\sum_{i=0}^{2n} y^k(i) \qquad (k=1,\ldots,M).
226: \end{equation}
227: %
228: These are sums of $k$-th powers of heights, which we call \emph{$k$-th moments of 
229: height}. The parameter $x_{1,n}$ is the area under the Dyck path, the parameter 
230: $x_{2,n}$ is called the \emph{moment of inertia} of the Dyck path \cite{NT03b}.
231: Let $\widetilde X_{k,n}$ denote the $k$-th moment of height of a random Dyck
232: path of length $2n$. In terms of $\widetilde Y_{n}(s)$, 
233: the random variables  $\widetilde X_{k,n}$ are expressed as
234: %
235: \begin{equation*}
236: \widetilde X_{k,n}=\sum_{i=0}^{2n} \widetilde Y_{n}^k(i).
237: \end{equation*}
238: %
239: The scaling of the average height of a random Dyck path with its length 
240: suggests considering normalised random variables $X_{k,n}$, defined by
241: %
242: \begin{equation}\label{form:DyckRV}
243: (X_{1,n},X_{2,n},\ldots, X_{M,n})= \left(\frac{\widetilde X_{1,n}}{n^{(1+2)/2}},
244: \frac{\widetilde X_{2,n}}{n^{(2+2)/2}},\ldots, 
245: \frac{\widetilde X_{M,n}}{n^{(M+2)/2}}\right).
246: \end{equation}
247: %
248: Then, in terms of $Y_{n}(t)$, the normalised random variables 
249: $X_{k,n}$ are given by
250: %
251: \begin{equation}\label{form:Dyckran}
252: X_{k,n}=2^{(k+2)/2}\sum_{i=0}^{2n} Y^k_{n}\left(\frac{i}{2n}\right)\cdot 
253: \frac{1}{2n}.
254: \end{equation}
255: %
256: Since the above sum is a Riemann sum, one is led to expect convergence
257: in distribution and moment convergence of the sequence $\{X_{k,n}\}_{n\in\mathbb N}$ to 
258: $2^{(k+2)/2}\int_0^1 e^k(t){\rm d}t$, due to the convergence properties of 
259: $\{Y_{n}(t)\}_{n\in\mathbb N}$. This is the statement of the following proposition.
260: 
261: \begin{prop}\label{prop:reldyckexc}
262: For $k\in \{1,\ldots,M\}$, let $X_k=\int_{0}^1 e^k(t)\,{\rm d}t$ denote the $k$-th 
263: excursion moment. The sequence of Dyck path random variables eqn.~\eqref{form:DyckRV} 
264: converges to the normalised excursion moments $(2^{(1+2)/2}X_1,\ldots,2^{(M+2)/2}X_M)$ 
265: in distribution,
266: %
267: \begin{equation}\label{form:DyEx}
268: (X_{1,n},X_{2,n},\ldots, X_{M,n}) \stackrel{d}{\longrightarrow} 
269: (2^{(1+2)/2}X_1,2^{(2+2)/2}X_2,\ldots,2^{(M+2)/2}X_M) \qquad (n\to\infty).
270: \end{equation}
271: %
272: We also have moment convergence.
273: \end{prop}
274: 
275: \begin{proof}
276: 
277: For $k\in\{1,\ldots,M\}$, define $Y_{k,n}=\int_0^1 Y^k_n(t){\rm d}t$. As stated
278: above, convergence in distribution and moment convergence holds for the 
279: sequence $\left\{(Y_{1,n},\ldots,Y_{M,n})\right\}_{n\in\mathbb N}$, 
280: see \cite{Bel72,A93} and \cite[Thm.~9]{D03}. We argue that the sequences 
281: $\{(X_{1,n},\ldots,X_{M,n})\}_{n\in\mathbb N}$ and $\{(2^{(1+2)/2}Y_{1,n},
282: \ldots,2^{(M+2)/2}Y_{M,n})\}_{n\in\mathbb N}$ converge in distribution and for 
283: moments to the same limit.
284: 
285: Fix $k\in\{1,\ldots,M\}$ and consider, for a Dyck path $y$ of length $2n$ 
286: and for $m\in\{0,1,\ldots,2n-1\}$, the elementary estimate $(\max\{0,y(m)-1\})^k
287: \le \int_m^{m+1}y^k(s){\rm d}s\le(y(m)+1)^k$. Summing over $m$ yields, together 
288: with the binomial theorem, the estimate
289: %
290: \begin{equation}\label{form:estmom}
291: \sum_{l=0}^k \binom{k}{l}(-1)^l x_{l,n}\le\int_0^{2n} y^k(s){\rm d}s
292: \le \sum_{l=0}^k \binom{k}{l} x_{l,n}.
293: \end{equation}
294: %
295: This estimate translates to the distribution functions of the corresponding 
296: random variables. In terms of the random variables $X_{k,n}$ and 
297: $Y_{n}(t)$, we get
298: %
299: \begin{equation}\label{form:cdis}
300: \sum_{l=0}^k \binom{k}{l}(-1)^l \frac{X_{l,n}}{n^{(k-l)/2}}\stackrel{d}{\le}
301: 2^{(k+2)/2}\int_0^{1} Y_{n}^k(t){\rm d}t \stackrel{d}{\le} \sum_{l=0}^k 
302: \binom{k}{l}\frac{X_{l,n}}{n^{(k-l)/2}},
303: \end{equation}
304: %
305: where the superscript $\stackrel{d}{}$ indicates that the relation is to be 
306: understood via distribution functions. Since the sequence of random variables 
307: $\{X_{l,n}/n^{(l-k)/2}\}_{n\in\mathbb N}$ converges in probability to zero, 
308: for $l=0,\ldots,k-1$, the random variables on the l.h.s.~and on the r.h.s.~of 
309: eqn.~\eqref{form:cdis} converge in distribution to the same limit. We conclude 
310: that the sequences $\{X_{k,n}\}_{n\in\mathbb N}$ and $\{2^{(k+2)/2}\int_0^1 Y_{n}^k (t)
311: {\rm d}t\}_{n\in\mathbb N}$ converge in distribution to the same limit. 
312: Moment convergence follows similarly from eqn.~\eqref{form:estmom}. 
313: 
314: Since the above estimates also hold jointly in $k\in\{1,\ldots,M\}$, the 
315: statement of the proposition follows.
316: \end{proof}
317: 
318: \noindent \textbf{Remark.} 
319: The above statement concerns Dyck paths in a uniform ensemble. The same 
320: argument as above leads to analogous results for more general ensembles of 
321: Dyck paths, which arise from the depth first process of simply generated trees 
322: \cite{A93, MM03}. Compare Theorem~\ref{theo:probdist} for a further generalisation.
323: 
324: \medskip
325: 
326: Due to the above result, Brownian excursion functionals may be
327: studied via their discrete Dyck path counterparts. Below, we will analyse the 
328: asymptotic behaviour of the moments of the random variables 
329: $(X_{1,n},\ldots,X_{M,n})$ eqn.~\eqref{form:DyckRV}, using the $q$-functional 
330: equation which the generating function of Dyck paths obeys. This implies a 
331: certain recursion for the moments of the joint distribution of the 
332: excursion moments $(X_1,\ldots,X_M)$.
333: We have the following result, which is obtained by combining the statements of 
334: Proposition~\ref{prop:reldyckexc} and Proposition~\ref{prop:dyckmom} below.
335: For $M$-dimensional vectors, we use the abbreviation $\bs k=(k_1,\ldots,k_M)$ 
336: and write $\bs0=(0,\ldots,0)$. For $i\in\{1,\ldots,M\}$, the unit vector $\bs e_i$ is 
337: defined by $(\bs e_i)_j=\delta_{i,j}$ for $j=1,\ldots,M$, and $\bs l\le\bs k$ for 
338: vectors $\bs l=(l_1,\ldots,l_M)$ and $\bs k=(k_1,\ldots,k_M)$ means that $l_i\le k_i$ 
339: for $i=1,\ldots,M$. A multivariate power series with complex coefficients is called 
340: \emph{entire}, if it converges for arbitrary complex arguments.
341: 
342: \begin{theorem}[Excursion moments]\label{theo:excmom}
343: 
344: For $k\in\{1,\ldots, M\}$, let $X_k=\int_{0}^1 e^k(t)\,{\rm d}t$ denote the $k$-th 
345: excursion moment. The moments of the joint distribution of $(X_1,\ldots,X_M)$ 
346: are given by 
347: %
348: \begin{equation}\label{eqn:exmomexpl}
349: \frac{\mathbb E[X_1^{k_1}\cdot\ldots\cdot X_M^{k_M}]}{k_1!\cdot\ldots\cdot k_M!}
350: =\frac{\sqrt{2\pi}}{\Gamma(\gamma_{k_1,\ldots,k_M})}2^{\gamma_{k_1,\ldots,k_M}}
351: \frac{f_{k_1,\ldots,k_M}}{2},
352: \end{equation}
353: %
354: where $\gamma_{k_1,\ldots,k_M}=\gamma_{\bs k}=-1/2+\sum_{i=1}^M(1+i/2)k_i$,
355: and where $\Gamma(z)$ is the Gamma function. 
356: For ${\bs k}\ne {\bf 0}$, the numbers $f_{k_1,\ldots,k_M}=f_{\bs k}$ are characterised 
357: by the recursion
358: %
359: \begin{equation}\label{form:exc2}
360: \gamma_{{\bs k}-{\bs e}_1}
361: f_{{\bs k}-{\bs e}_1}+\sum_{i=1}^{M-1}2(i+1)(k_i+1)
362: f_{{\bs k}-{\bs e}_{i+1}+{\bs e}_i}+
363: \sum_{{\bf 0}\le{\bs l}\le{\bs k}}f_{\bs l}
364: f_{{\bs k}-{\bs l}}=0,
365: \end{equation}
366: %
367: with boundary conditions $f_{\bf 0}=-4$ and $f_{\bs k}=0$, if $k_j<0$ for 
368: some $j\in\{1,\ldots, M\}$. 
369: The moment generating function $\mathbb E[\mathrm{e}^{t_1 X_1+\ldots+t_M X_M}]$ 
370: is entire. Hence, the joint distribution is uniquely defined by its moments. \qed
371: \end{theorem}
372: 
373: \noindent \textbf{Remarks.} \textit{i)} The theorem asserts that the numbers 
374: $f_{\bs k}$ can be recursively computed from eqn.~\eqref{form:exc2}. To 
375: see this, we remark that the above equation has $-f_{\bs k}=2f_{\bs k}
376: f_{\bs 0}/8$ as a term, and that all other numbers $f_{\bs l}$ in the equation 
377: satisfy $\bs l\prec \bs k$ for a suitable total order $\prec$, which is 
378: specified below in Definition~\ref{def:totord}. \\
379: \textit{ii)} In probability theory, Louchard's theorem \cite{L84} leads
380: to a characterisation of a certain Laplace transform of the moment generating 
381: function, for some excursion functionals including excursion moments. It 
382: is however difficult to extract moment values or moment recursions from 
383: these expressions.  For $M=1$, this has been done in \cite{L85}.
384: The moment generating function of the marginal distribution for $M=2$ has been
385: obtained in \cite[Thm.~2.4]{NT03b}. Moment values or recursions for $M>2$ 
386: have apparently not been derived.\\
387: \textit{iii)} 
388: We obtain Theorem~\ref{theo:excmom} by studying the corresponding Dyck
389: path functionals. Our discrete approach has been used previously \cite{NT03b}.
390: It led to a combinatorial derivation \cite[Thm.~3.1]{NT03b} of Louchard's formula 
391: for integrals over Brownian excursion polynomials. It also led to the values 
392: $\mathbb E[(X_1)^{k_1} (X_2)^{k_2}]$, see \cite[Table~2]{NT03b}.
393: For arbitrary $M$, our result is announced in \cite{R04}. 
394: [A factor 1/2 is missing on the r.h.s.~of eqn.~(6) in \cite{R04}].\\
395: \textit{iv)} Assuming only convergence in distribution of $(X_{1,n},\ldots,X_{M,n})$, 
396: the above result can be used to provide an alternative proof of moment
397: convergence.
398: 
399: \medskip
400: 
401: Let us briefly discuss the moments of the marginal distributions for $M\le4$.
402: For $M=1$, the first few coefficients $\mathbb E[(X_1)^k]$ are $1$, 
403: $\frac{1}{4}\sqrt{2\pi}$, $\frac{5}{12}$, $\frac{15}{128}\sqrt{2\pi}$, 
404: $\frac{221}{1008}$, $\frac{565}{8192}\sqrt{2\pi}$. These are related to the 
405: \emph{Airy distribution} \cite{FL01,J07,KMM07}, i.e., the distribution of area under the 
406: Brownian excursion $\sqrt{8} \, e(t)$. Explicit expressions are 
407: known for the moment generating function, for the density of the distribution,
408: and for the moments.
409: 
410: For $M=2$, the first few moments $\mathbb E[(X_2)^k]$
411: are $1$, $\frac{1}{2}$, $\frac{19}{60}$, $\frac{631}{2520}$, 
412: $\frac{1219}{5040}$, $\frac{92723}{332640}$, $\frac{1513891}{4036032}$. 
413: An explicit expression for the corresponding moment generating function 
414: \cite{NT03b} is
415: %
416: \begin{equation*}
417: \mathbb E[\mathrm{e}^{tX_2}]=\sum_{k=0}^\infty\frac{\mathbb E[(X_2)^k]}{k!} t^k=
418: \left( \frac{\sqrt{2t}}{\sin(\sqrt{2t})}\right)^{3/2}.
419: \end{equation*}
420: %
421: The first few moments $\mathbb E[(X_3)^k]$ are $1$, 
422: $\frac{3\sqrt{2\pi}}{16}$, $\frac{207}{560}$, $\frac{11907\sqrt{2\pi}}{65536}$, 
423: $\frac{88655283}{108908800}$, $\frac{1165359069\sqrt{2\pi}}{1476395008}$, and 
424: for the first few moments $\mathbb E[(X_4)^k]$ we have  $1$, $\frac{1}{2}$, 
425: $\frac{251}{840}$, $\frac{288751}{1201200}$, $\frac{19093793}{76236160}$, 
426: $\frac{105169404203}{3259095840000}$. It remains an open problem to find
427: explicit expressions for the moment generating functions.
428: 
429: \subsection{Dyck paths and $q$-functional equations}
430: 
431: We discuss the functional equation which the generating function of Dyck 
432: paths obeys, when counted by length and $k$-th moments of height,
433: $k=1,\ldots,M$. We then state in Theorem~\ref{theo:probdist} a convergence 
434: result for the limit distribution of counting parameters related to general 
435: $q$-functional equations. This is the main result of our paper.
436: 
437: \medskip
438: 
439: Let $\bs u=(u_0,u_1,\ldots,u_M)$ denote formal (commutative) variables. For a Dyck 
440: path $p$ of length $2n$, its \emph{weight} $w(p)$ is given by $w(p)=
441: u_0^{n_{}} u_1^{x_{1,n}}\cdot\ldots\cdot u_M^{x_{M,n}}$, where $x_{k,n}$
442: is the $k$-th moment of height eqn.~\eqref{form:dyckpara}. For the set $\cal D$ 
443: of Dyck paths, its generating function is the formal power 
444: series $D({\bs u})=\sum_{p\in\cal D} w(p)$. For the set $\cal A$ of 
445: arches, its generating function is the formal power series $A({\bs 
446: u})=\sum_{p\in\cal A} w(p)$. The variables $u_1,\ldots, u_M$ are interpreted 
447: as deformation variables. They may all be set equal to unity, in which case 
448: the generating function reduces to that of Dyck paths (arches), counted by length only.
449: The generating functions satisfy the following functional
450: equation.
451: 
452: \begin{theorem}[Dyck path generating function, cf.~\cite{NT03b}]\label{theo:dyck}
453: The generating functions of Dyck paths $D({\bs u})$ and of arches 
454: $A({\bs u})$ satisfy
455: %
456: \begin{equation}\label{eqn:dyckfeq}
457: \begin{split}
458: D(u_0,\ldots,u_M)&=1+D(u_0,\ldots,u_M)A(u_0,\ldots,u_M),\\
459: A(u_0,u_1,\ldots,u_M)&=u_0 u_1\cdot\ldots\cdot u_M D(v_0,\ldots,v_M),
460: \end{split}
461: \end{equation}
462: %
463: where the monomials $v_k=v_k({\bs u})$ are given by
464: %
465: \begin{equation*}
466: v_0({\bs u})=u_0u_1^2u_2^2\cdot\ldots\cdot u_M^2, \qquad v_k({\bs u})
467: =\prod_{l=k}^M u_l^{\binom{l}{k}} \qquad (k=1,\ldots,M).
468: \end{equation*}
469: %
470: \qed
471: \end{theorem}
472: 
473: \noindent \textbf{Remarks.} \textit{i)} For $M=2$, the above theorem appears in 
474: \cite[Sec.~3.1]{NT03b}. For arbitrary $M$, it is (incorrectly) stated in \cite[p.~713]{NT03b}.
475: The proof for $M=2$ given in \cite{NT03b} generalises to arbitrary $M$. It uses
476: a last passage decomposition argument and the additivity of the counting parameters with
477: respect to the sequence construction, yielding the first equation in the theorem. 
478: Note that the arch decomposition of Dyck paths, as depicted in Figure \ref{fig:dyck},
479: implies the equivalent equation $D(\bs u)=1/(1-A(\bs u))$. 
480: For $n\in\mathbb N$, there is a bijection between the set of arches of length $2n$ and 
481: the set of Dyck paths of length $2n-2$, by identifying an arch, where the bottom layer 
482: has been removed, with the corresponding Dyck path. This implies, after a short calculation 
483: (see also \cite[Thm.~1]{R06}), the second equation in Theorem \ref{theo:dyck}.\\
484: \textit{ii)} The induced functional equation for the generating function $E(\bs u)=
485: D(\bs u)-1$ is an example of a $q$-functional equation as in Definition \ref{defi:qfunc},
486: with a square-root singularity in the generating function in the ``undeformed'' case 
487: $(u_1,\ldots,u_M)=(1,\ldots,1)$. See the following section and Section~\ref{sec:Dyckrev}.
488: 
489: \medskip
490: 
491: For a probabilistic interpretation of the counting parameters, as in the previous subsection,
492: let $p_{n_0,n_1,\ldots,n_M}$ denote the number of Dyck 
493: paths of length $2n_0$, where the $k$-th moment of height has the value $n_k$, 
494: for $k=1,\ldots,M$. We clearly have $0<\sum_{n_1,\ldots, n_M}p_{n_0,n_1,\ldots,n_M}<\infty$
495: for $n_0>0$. In the uniform ensemble, the random variables $\widetilde X_{k,n_0}$, which assign 
496: the $k$-th moment of height to a random Dyck path of length $2n_0$, have the joint distribution
497: 
498: \begin{equation}\label{eq:pxt}
499: \mathbb P(\widetilde X_{1,n_0}=n_1, \ldots, \widetilde X_{M,n_0}=n_M)=
500: \frac{p_{n_0,n_1,\ldots,n_M}}{\sum_{n_1,\ldots, n_M}p_{n_0,n_1,\ldots,n_M}}.
501: \end{equation}
502: %
503: The moments $\widetilde m_{k_1,\ldots,k_M}(n_0)$ of the joint distribution are given by
504: 
505: \begin{equation*}
506: \widetilde m_{k_1,\ldots,k_M}(n_0)=\frac{\sum_{n_1,\ldots, n_M} n_1^{k_1}
507: \cdot\ldots\cdot n_M^{k_M}p_{n_0,n_1,\ldots,n_M}}{\sum_{n_1,\ldots, n_M}
508: p_{n_0,n_1,\ldots,n_M}}.
509: \end{equation*}
510: %
511: In the previous subsection, we introduced the normalised
512: random variable eqn.~\eqref{form:DyckRV},
513: %
514: \begin{equation}\label{form:DyckRV2}
515: (X_{1,n_0},X_{2,n_0},\ldots, X_{M,n_0})= \left(\frac{\widetilde X_{1,n_0}}{n_0^{(1+2)/2}},
516: \frac{\widetilde X_{2,n_0}}{n_0^{(2+2)/2}},\ldots, 
517: \frac{\widetilde X_{M,n_0}}{n_0^{(M+2)/2}}\right).
518: \end{equation}
519: %
520: The normalised random variable $(X_{1,n_0},\ldots,X_{M,n_0})$ has moments 
521: $m_{k_1,\ldots,k_M}(n_0)$, given by
522: %
523: \begin{equation*}
524: m_{k_1,\ldots,k_M}(n_0)=\frac{\widetilde m_{k_1,\ldots,k_M}(n_0)}{
525: n_0^{(1+2)k_1/2+(2+2)k_2/2+\ldots+(M+2)k_M/2}}.
526: \end{equation*}
527: We argued above that these numbers should tend to a finite limit, as 
528: $n_0$ approaches infinity. As we will see below, this is indeed the 
529: case. Hence we may define
530: %
531: \begin{equation}\label{eqn:momlim}
532: m_{k_1,\ldots,k_M}=\lim_{n_0\to\infty}m_{k_1,\ldots,k_M}(n_0).
533: \end{equation}
534: 
535: A careful analysis of the functional equation eqn.~\eqref{eqn:dyckfeq}, which 
536: will be performed in the general case from Section \ref{sec:ana} onwards, 
537: yields the following result for the numbers $m_{k_1,\ldots,k_M}$. Its proof 
538: is deferred until Section~\ref{sec:Dyckrev}.
539: 
540: \begin{prop}\label{prop:dyckmom}
541: The normalised moments $m_{k_1,\ldots,k_M}$ of Dyck paths eqn.~\eqref{eqn:momlim}
542: are given by
543: %
544: \begin{equation}\label{eqn:dyckres}
545: \frac{m_{k_1,\ldots,k_M}}{k_1!\cdot\ldots\cdot k_M!}=
546: \frac{1}{f_{\bs0} u_c^{\gamma_{\bs k}-\gamma_{\bs 0}}}
547: \frac{\Gamma(\gamma_{\bs 0})}{\Gamma(\gamma_{\bs k})}
548: f_{\bs k},
549: \end{equation}
550: %
551: where $u_c=1/4$, where $\Gamma(z)$ denotes the Gamma function, and where the numbers 
552: $\gamma_{\bs k}$ and $f_{\bs k}$ are defined in eqn.~\eqref{form:exc2}. 
553: The moments have an entire exponential generating function. Hence, they define 
554: a unique random variable with moments $m_{\bs k}$.
555: \end{prop}
556: 
557: \noindent \textbf{Remarks.}
558: \textit{i)} Proposition~\ref{prop:dyckmom} implies convergence in distribution
559: and moment convergence of the normalised random variables eqn.~\eqref{form:DyckRV2}. 
560: This yields an alternative
561: proof of the convergence statement of Proposition~\ref{prop:reldyckexc}. However,
562: it does not establish a connection between the limit random variable and
563: the excursion moments, as in Proposition~\ref{prop:reldyckexc}.\\
564: \textit{ii)} Since the moments $m_{\bs k}$ define a unique random variable, the 
565: result for the excursion moments eqn.~\eqref{eqn:exmomexpl} follows from 
566: Proposition~\ref{prop:dyckmom} by Proposition \ref{prop:reldyckexc}. \\
567: 
568: 
569: \medskip
570: 
571: For the solution of a general $q$-functional equation (Definition \ref{defi:qfunc}), 
572: we obtain a similar result. Write $G(\bs u)=\sum_{n_1,\ldots,n_M}p_{n_0,n_1,\ldots,n_M}
573: u_0^{n_0}u_1^{n_1}\cdot\ldots\cdot u_M^{n_M}$ for such a solution. Under mild assumptions, 
574: the corresponding random variables in eqn.~\eqref{eq:pxt} are well-defined, and the same 
575: asymptotic analysis as above can be performed. We have the following theorem which is, 
576: in conjunction with Theorem \ref{theo:excmom}, the main result of our paper. Its proof is 
577: deferred until Section~\ref{sec:mom} and Section~\ref{sec:Dyckrev}.
578: %
579: \begin{theorem}[Limit distribution for $q$-functional equations]\label{theo:probdist}
580: Let a $q$-functional equation Definition \ref{defi:qfunc} with solution 
581: %
582: \begin{equation*}
583: G(u_0,u_1,\ldots,u_M)=\sum_{n_1,\ldots,n_M}p_{n_0,n_1,\ldots,n_M}u_0^{n_0}u_1^{n_1}
584: \cdot\ldots\cdot u_M^{n_M}
585: \end{equation*}
586: as in Proposition~\ref{prop:uni} be given, and let Assumption~\ref{ass} be satisfied.
587: Assume that the numbers $A_i$ in Proposition \ref{prop:rek} are
588: positive, i.e., $A_i>0$ for $i=0,\ldots,M-1$. Then, the random variables 
589: $(\widetilde X_{1,n_0},\ldots,\widetilde X_{M,n_0})$ eqn.~\eqref{eq:pxt} are
590: well-defined for almost all $n_0$. The following conclusions hold.
591: 
592: \begin{itemize}
593: \item[i)]
594: The numbers $m_{k_1,\ldots,k_M}$ eqn.~\eqref{eqn:momlim}, which are derived from the
595: moments of the normalised random variables $(X_{1,n_0},\ldots,X_{M,n_0})$ 
596: eqn.~\eqref{form:DyckRV2}, define a unique random variable
597: $(Y_1,\ldots,Y_M)$ with moments $m_{k_1,\ldots,k_M}$. We have convergence in distribution,
598: \begin{equation*}
599: (X_{1,n_0},\ldots, X_{M,n_0}) \stackrel{d}{\to} (Y_1,\ldots,Y_M)\qquad (n_0\to\infty),
600: \end{equation*}
601: and we have moment convergence.
602: 
603: \item[ii)]
604: The limiting random variable $(Y_1,\ldots,Y_M)$ is explicitly given by
605: %
606: \begin{equation*}
607: (Y_1,\ldots,Y_k)=(c_1X_1,\ldots,c_MX_M),
608: \end{equation*}
609: % 
610: where the constants $c_k>0$ and the random variables $X_k$ are
611: %
612: \begin{equation*}
613: c_k =2^{\frac{k+2}{2}}\frac{2\mu_0\cdot\ldots\cdot \mu_{k-1}}{4^k k!}, \qquad
614: X_k=\int_{0}^1 e^k(t)\,{\rm d}t \qquad (k=1,\ldots,M).
615: \end{equation*} 
616: %
617: Here, the numbers $\mu_i>0$ are defined in Proposition \ref{prop:rek}, and $e(t)$ 
618: denotes a standard Brownian excursion of duration 1.
619: 
620: \end{itemize}
621: \end{theorem}
622: 
623: \noindent \textbf{Remarks.} \textit{i)} A recursion for the moments of the limit distribution
624: appears in Theorem~\ref{theo:excmom} above, see also Proposition \ref{prop:rek} below.\\
625: \textit{ii)} For Dyck paths counted by length and $k$-th moments of height, the 
626: above statements follow already from Proposition~\ref{prop:reldyckexc}. More 
627: generally, as was argued in the remark following 
628: Proposition~\ref{prop:reldyckexc}, the above result can be shown to hold for models 
629: of simply generated trees \cite{MM78}, counted by number of vertices and $k$-th moments 
630: of internal path length. This follows from the polynomial convergence of the 
631: depth first process derived from simply generated trees towards the Brownian excursion 
632: \cite{Bel72, A93, MM03, D03}. For a general $q$-functional equation, such a connection 
633: is not known to exist.\\
634: \textit{iii)} Our method of proof also allows to study corrections to the asymptotic behaviour, 
635: compare the discussion in Section~\ref{sec:dom}, and \cite{R02} for the case $M=1$. These 
636: cannot be obtained by the methods of \cite{MM03}. 
637: 
638: \subsection{Structure of the paper}
639: 
640: The remainder of the paper is organised as follows. In Section~\ref{sec:qfunc}, we 
641: introduce $q$-shifts (Definition~\ref{def:qshift}) and $q$-functional equations 
642: (Definition~\ref{defi:qfunc}). Our results rely on an application of the 
643: multivariate moment method (see e.g.~\cite[Ch.~6.1.]{JLR00} or \cite[Sec.~30]{Bill95}). 
644: Hence in Section~\ref{sec:gk}, eqn.~\eqref{form:facarmom}, we introduce 
645: \emph{factorial moment generating functions}, which are derivatives of the solution 
646: of the functional equation, evaluated at $u_1=\ldots=u_M=1$. We study their properties 
647: by a combinatorial analysis of derivatives of the functional equation, using a 
648: multivariate generalisation of Faa di Bruno's formula \cite{CS96}. In Section 
649: \ref{sec:ana}, we study the singular behaviour of the factorial moment generating 
650: functions, in the case of a square root singularity as the dominant singularity 
651: of the size generating function. Proposition~\ref{prop:rek} gives a recursion for 
652: the amplitudes, which describe the leading singular behaviour of the factorial moment 
653: generating functions.  Our method is also called \emph{moment pumping} \cite{FPV98}. 
654: We employ in Section~\ref{sec:dom} an alternative 
655: (rigorous) method to obtain the recursion, which originates from the 
656: \emph{method of dominant balance} of statistical mechanics \cite{PB95a, R02}. 
657: This method is generally easier to apply than an analysis of the functional equation 
658: via Faa di Bruno's formula, and yields an algorithm for obtaining the amplitude recursion.
659: In addition, the method allows to analyse corrections to the limiting behaviour. 
660: The behaviour of the moments follows then by standard methods 
661: from singularity analysis of generating functions \cite{FO90,FS07}.
662: In Proposition~\ref{theo:pde}, we give a quasi-linear partial differential equation 
663: for the generating function of the amplitudes. Growth estimates for the amplitudes 
664: (and hence for the moments) are obtained in Section~\ref{sec:growth}, by an analysis 
665: of the (singular) partial differential equation. Existence and uniqueness of a limit distribution 
666: is then guaranted by L\'evy's continuity theorem, see Section~\ref{sec:mom}. The connection 
667: to Brownian excursions follows in Section~\ref{sec:Dyckrev}, by a comparison with 
668: Dyck paths. Possible applications of our method are discussed in a concluding section.
669: 
670: \section{$q$-functional equations}\label{sec:qfunc}
671: 
672: Let $\mathbb C [[\bs u]]$ denote the ring of formal power series with complex coefficients 
673: in the (commutative) variables ${\bs u}=(u_0,u_1,\ldots, u_M)$. Let $\mathbb C [\bs u]$ 
674: denote the ring of polynomials with complex coefficients in the variables ${\bs u}$. We 
675: set ${\bs u}_0=(u_0,1,\ldots,1)$. For ${\bs u}=(u_0,u_1,\ldots,u_M)$ and ${\bs n}\in
676: \mathbb N_0^{1+M}$, we define $\bs u^{\bs n}$ to be the monomial ${{\bs u}^{\bs n}}=
677: u_0^{n_0}\cdot\ldots\cdot u_M^{n_M}$. We employ the notation ${\bs u}_+=(u_1,\ldots,u_M)$, 
678: the plus sign indicating that the first component of $\bs u$ is omitted. For $n\in\mathbb 
679: N$ and $r>0$, let $D^n_r$ denote the open polydisc 
680: %
681: \begin{equation*}
682: D^n_r=\{ (x_1,\ldots, x_n)\in\mathbb C^n: |x_k|<r \;\mbox{for all}\; k=1,\ldots, n\}.
683: \end{equation*}
684: % 
685: Let $\mathcal{H}_r(\bs x)$ denote the ring of power series in $\bs x=(x_1,\ldots,x_n)$ 
686: with complex coefficients, which are convergent in $D^n_r$. 
687: 
688: \begin{defi}[$q$-shift]\label{def:qshift}
689: Fix $M\in\mathbb N$. Let for $k=0,\ldots,M$ formal power series 
690: $v_k(\bs u)\in\mathbb C[[\bs u]]$ be given, and write
691: $\bs v=\bs v(\bs u)=(v_0(\bs u),
692: v_1(\bs u),\ldots,v_M(\bs u))$. Assume that there is a
693: number $d=d(\bs v)$ satisfying $1<d\le\infty$, such that
694: %
695: \begin{equation*}
696: v_0(\bs u)\in \mathcal{H}_d(u_1,\ldots,u_M)[[u_0]],\qquad
697: v_k(\bs u)\in \mathcal{H}_d(u_k,\ldots,u_M) \qquad (k=1,\ldots, M).
698: \end{equation*}
699: %
700: Assume that $\bs v(\bs u)$ satisfies
701: %
702: \begin{equation*}
703: {\bs v({\bs u}_0)={\bs u}_0}, \qquad
704: v_0(0,{\bs u}_+)\equiv0, \qquad \frac{\partial v_k}
705: {\partial u_k}({\bs u}_0)\equiv1 \qquad (k=0,\ldots, M).
706: \end{equation*} 
707: %
708: Let $r=r(\bs v)$ be a number $0<r\le\infty$ such that
709: %
710: \begin{equation*}
711: \left\{(v_1(\bs u_+),\ldots, v_M(\bs u_+) ): \bs u_+\in D_d^M\right\}\subseteq D_r^M.
712: \end{equation*}
713: %
714: Then, $\bs v$ is called a \emph{$q$-shift}, $d(\bs v)$ is called
715: the \emph{domain of $\bs v$}, and $r(\bs v)$ is called the 
716: \emph{range of $\bs v$}.
717: 
718: \end{defi}
719: 
720: \noindent \textbf{Remarks.}
721: \textit{i)} For $M=1$, a simple example of $q$-shift appears in 
722: the $q$-difference equation eqn.~\eqref{form:geneq2}, with
723: $v_0(u_0,u_1)=u_0u_1$ and $v_1(u_1)=u_1$. This motivates 
724: the name for the generalisation. \\
725: \textit{ii)} For a $q$-shift ${\bs v}$, its $k$-th component 
726: $v_k(\bs u)$ does not depend on $u_l$, where $l=0,\ldots,k-1$,
727: and it does depend on $u_k$. Since ${\bs v({\bs u}_0)={\bs u}_0}$,
728: one may interpret $\bs v(\bs u)$ for $\bs u\ne\bs u_0$
729: as a ``deformation'' of $\bs u_0$. The condition $v_0(0,{\bs u}_+)\equiv0$
730: is imposed to ensure that composition of formal power series is well-defined, 
731: see below.\\
732: \textit{iii)}
733: The identity $id: {\bs u}\mapsto {\bs u}$ is a $q$-shift with infinite domain and 
734: range. For two $q$-shifts ${\bs v}$ and ${\bs w}$, their composition 
735: ${\bs v}\circ {\bs w}$ is well-defined if $r(\bs w)\le d(\bs v)$. As is readily checked, 
736: ${\bs v}\circ {\bs w}$ is a $q$-shift in that case, with domain $d({\bs v}\circ {\bs w})=d(\bs w)$ 
737: and range $r({\bs v}\circ {\bs w})=r(\bs v)$. If for a $q$-shift ${\bs v}$  we have 
738: $r(\bs v)\le d(\bs v)$, we can thus consider \emph{iterated} $q$-shifts ${\bs v}^{[n]}$, where
739: %
740: \begin{equation*}
741: {\bs v}^{[0]}=id, \qquad {\bs v}^{[n]}={\bs v}\circ{\bs v}^{[n-1]} \qquad (n\in\mathbb N).
742: \end{equation*}
743: %
744: The power series ${\bs v}^{[n]}$ is a $q$-shift for $n\in \mathbb N_0$.\\
745: \textit{iv)} An important subclass (with infinite domain and range) are \emph{polynomial} 
746: $q$-shifts, i.e., $q$-shifts $\bs v$ satisfying $v_k(\bs u) \in 
747: \mathbb C[\bs u]$ for $k=0,\ldots, M$. Examples are given by 
748: the monomial $q$-shifts
749: %
750: \begin{equation*}
751: v_k({\bs u}) = {\bs u}^{{\bs n}_k},
752: \end{equation*}
753: %
754: where ${\bs n}_k\in\mathbb N_0^{1+M}$, $({\bs n}_k)_k=1$ 
755: and $({\bs n}_k)_l=0$ for $l=0,\ldots,k-1$ and $k=0,\ldots,M$. These
756: appear in eqns.~\eqref{form:geneq2} and \eqref{form:geneq3}, and for 
757: Dyck paths in eqn.~\eqref{eqn:dyckfeq}. In these examples, 
758: we have $({\bs n}_l)_{l+1}\ne 0$ for $l=0,\ldots, M-1$.
759: 
760: \medskip
761: 
762: For a formal power series $G({\bs u})\in \mathcal{H}_d(\bs u_+)[[u_0]]
763: \subseteq\mathbb C [[\bs u]]$ and a $q$-shift ${\bs v}$ satisfying
764: $r(\bs v)\le d$, we define $H({\bs u}) \in \mathcal{H}_{d(\bs v)}(\bs u_+)[[u_0]]$ by
765: %
766: \begin{equation*}
767: H({\bs u})= G({\bs v}({\bs u})).
768: \end{equation*}
769: %
770: This is well-defined since $\mathcal{H}_d({\bs v}_+({\bs u}))\subseteq 
771: \mathcal{H}_{d(\bs v)}(\bs u_+)$ and $\mathbb C[[v_0({\bs u})]]\subseteq
772: \mathcal{H}_{d(\bs v)}(\bs u_+)[[u_0]]$, due to $v_0(0,{\bs u}_+)\equiv0$. 
773: We are interested in derivatives of $H({\bs u)}$. For 
774: clarity of presentation, we will use the multi-index notation, and 
775: Greek indices $\bs\mu, \bs\nu,\bs\rho$ will denote vectors with 
776: non-negative integer entries. For $F({\bs u}) \in\mathbb C [[\bs u]]$ and 
777: $\bs\nu=(\nu_0,\ldots,\nu_M)\in
778: \mathbb N_0^{1+M}$, we write the derivative $F_{\bs \nu}({\bs u}) 
779: \in\mathbb C [[\bs u]]$ of $F(\bs u)$ of order $\bs\nu$ as
780: %
781: \begin{equation*}
782: F_{\bs\nu}({\bs u}) = \partial_0^{\nu_0}\cdots 
783: \partial_M^{\nu_M}F({\bs u}),\qquad \partial_k^{\nu_k}=
784: \frac{\partial^{\nu_k}}{\partial u_k^{\nu_k}}\qquad (k=0,\ldots,M),
785: \end{equation*}
786: %
787: where we use the convention that $F_{\bs 0}(\bs u)=F(\bs u)$.
788: If $G({\bs u}) \in \mathcal{H}_d(\bs u_+)[[u_0]]\subset\mathbb 
789: C[[\bs u]]$, we also have $G_{\bs\nu}({\bs u})\in\mathcal{H}_d(\bs u_+)
790: [[u_0]]$. Set $|\bs\nu|=\nu_0+\ldots+\nu_M$ and $\bs\nu!=\nu_0!
791: \cdot\ldots\cdot \nu_M!$. We define a total order in $\mathbb N_0^{1+M}$
792: as follows.
793:  
794: \begin{defi}[total order $\prec$]\label{def:totord}
795: For $\bs\mu =(\mu_0,\ldots,\mu_M)\in\mathbb N_0^{1+M}$ and 
796: $\bs\nu=(\nu_0,\ldots,\nu_M)\in\mathbb N_0^{1+M}$, we write 
797: $\bs\mu \prec\bs\nu$ if either $|\bs\mu |<|\bs\nu|$, or if $|\bs\mu |=|\bs\nu|$, 
798: there exists an index  $k\in\{0,\ldots,M\}$, such that
799: $\mu_k>\nu_k$ and $\mu_i=\nu_i$ for $i=0,\ldots,k-1$.
800: \end{defi}
801: 
802: The total order introduced above will be used to label terms appearing
803: in derivatives of $H({\bs u)}$. We have the following lemma.  For 
804: $i\in\{0,\ldots,M\}$, let ${\bs e}_i\in\mathbb C^{1+M}$ denote the 
805: unit vector in direction $i$, given by $({\bs e}_i)_k=\delta_{i,k}$ for $k=0,\ldots,M$.
806: 
807: \begin{lemma}\label{lemma:H}
808: Let $G({\bs u}) \in \mathcal{H}_d(\bs u_+)[[u_0]]$, and let a 
809: $q$-shift $\bs v$ with domain $d(\bs v)$ and range $r(\bs v)\le d$ be given. 
810: Set $H({\bs u)}=G({\bs v(\bs u)})\in \mathcal{H}_{d(\bs v)}(\bs u_+)[[u_0]]$ 
811: and fix $\bs\nu\ne {\bf 0}$.
812: %
813: \begin{itemize}
814: 
815: \item[i)] 
816: We have $H_{\bs\nu}({\bs u})\in\mathcal{H}_{d(\bs v)}(\bs u_+)[[u_0]]$.
817: For every $\bs \mu$ satisfying $\mathbf{0}\ne\bs\mu \preceq\bs\nu$, there
818: exists a coefficient $A(\bs\mu,\bs u)\in\mathcal{H}_{d(\bs v)}(\bs u_+)[[u_0]]$, 
819: independent of the choice of $G({\bs u})$, such that
820: %
821: \begin{equation}\label{form:chain}
822: H_{\bs\nu}({\bs u}) = 
823: \sum_{{\bf 0}\ne\bs\mu \preceq\bs\nu}
824: G_{\bs\mu }({\bs v}({\bs u})) \cdot A(\bs\mu ,{\bs u}).
825: \end{equation}
826: 
827: \item[ii)] The coefficient $A(\bs\mu,{\bs u})$ in eqn.~\eqref{form:chain}
828: is, for $\bs\mu=\bs\nu$, given by
829: %
830: \begin{equation}\label{form:anuu}
831: A(\bs\nu,{\bs u})=\prod_{k=0}^M \left( \frac{\partial v_k}{\partial u_k}
832: ({\bs u}) \right)^{\nu_k}.
833: \end{equation}
834: %
835: In particular, we have $A(\bs\nu,{\bs u_0})=1$. 
836: 
837: \item[iii)]
838: Fix $i\in\{0,\ldots,M-1\}$. If $\nu_{i+1}>0$, 
839: we have $\bs\nu-{\bs e}_{i+1}+{\bs e}_i \prec \bs\nu$.
840: The coefficient $A(\bs\mu,{\bs u})$ in eqn.~\eqref{form:chain}
841: is, for $\bs\mu=\bs\nu-{\bs e}_{i+1}+{\bs e}_i$ and 
842: $\bs u=\bs u_0$, given by
843: %
844: \begin{equation*}
845: A(\bs\nu-{\bs e}_{i+1}+{\bs e}_i,{\bs u_0})= 
846: \nu_{i+1}\frac{\partial v_i}{\partial u_{i+1}}({\bs u_0}).
847: \end{equation*}
848: 
849: \item[iv)] 
850: For real numbers $r,r_0,\ldots,r_M$, where $r_{k+1}>r_k$ for 
851: $k=0,\ldots,M-1$, define $\alpha_{\bs\mu }=r+\sum_{k=0}^M r_k\mu_k$. 
852: Then, for indices $\bs\mu \prec\bs\nu$ in eqn.~\eqref{form:chain} 
853: satisfying $A(\bs\mu ,{\bs u})\not\equiv0$, we have
854: $\alpha_{\bs\mu }<\alpha_{\bs\nu}$.
855: \end{itemize}
856: \end{lemma}
857: 
858: \noindent {\bf Remarks.} \textit{i)} The coefficient $A(\bs\mu ,{\bs u})$ 
859: in eqn.~\eqref{form:chain} might be chosen to vanish. E.g., fix $\bs\nu=(1,1)$ 
860: and consider $\bs\mu =(0,1)$. We have $\bs\mu\prec\bs
861: \nu$ but might choose $A(\bs\mu ,{\bs u})\equiv0$, as is readily verified by an 
862: explicit calculation, using the $q$-shift property $v_k({\bs u}) \in 
863: \mathbb C[[u_k,\ldots,u_M]]$. \\
864: \textit{ii)} The property \textit{iv)} will be used for exponent estimates
865: in Proposition~\ref{prop:est} below.
866: 
867: \begin{proof}
868: {\it i)} This is an application of the chain rule. Successive differentiation leads to
869: %
870: \begin{equation}\label{form:succ}
871: H_{\bs\nu}({\bs u}) = 
872: \sum_{0\le|\bs\mu |\le|\bs\nu|}
873: G_{\bs\mu }({\bs v}({\bs u})) \cdot A(\bs\mu ,{\bs u}),
874: \end{equation}
875: %
876: where $A(\bs\mu ,{\bs u})$ is independent of $G({\bs u})$. To analyse the 
877: effect of the $q$-shift property $v_k({\bs u})\in\mathbb {\cal H}_{d(\bs v)}(u_k,\ldots, u_M)$, 
878: consider for $k\in\{0,\ldots, M\}$ the equation
879: %
880: \begin{equation}\label{form:d1}
881: \frac{\partial H}{\partial u_k}({\bs u}) =
882: \sum_{l=0}^M \frac{\partial G}{\partial v_l}({\bs v}({\bs u}))\frac{\partial v_l}
883: {\partial u_k}({\bs u})=\sum_{l=0}^k \frac{\partial G}{\partial v_l}({\bs v}
884: ({\bs u}))\frac{\partial v_l}{\partial u_k}({\bs u}).
885: \end{equation}
886: %
887: It shows that derivatives of $H({\bs u})$ w.r.t.~$u_k$ do not contribute to derivatives of 
888: $G({\bs v}({\bs u}))$ w.r.t.~$v_{k+1},\ldots,v_M$. A similar statement holds for 
889: higher derivatives. This implies that the numbers $\nu_k$ in eqn.~\eqref{form:succ}
890: can contribute to the numbers $\mu_0,\ldots,\mu_k$ only. If $|\bs\mu |=|\bs\nu|$, we thus 
891: have  $\nu_k\le \mu_0+\ldots+\mu_k$ for $k\in\{0,\ldots,M\}$. We show that $\bs\mu 
892: \preceq\bs\nu$. Assume w.l.o.g. that $|\bs\mu |=|\bs\nu|$. If $\mu_0>\nu_0$, we have  
893: $\bs\mu \prec \bs\nu$, and the claim follows. Otherwise, we have $\mu_0=\nu_0$. 
894: Thus, $\nu_k$ does not contribute to $\mu_0$ for $k\in\{1,\ldots,M\}$. The previous 
895: argument yields that $\nu_k\le \mu_1+\ldots+\mu_k$ for $k\in\{1,\ldots,M\}$. Now repeat 
896: the above argument until $\mu_0=\ldots=\mu_{M-1}=\nu_0=\ldots=\nu_{M-1}$. Then 
897: $\mu_M=\nu_M$ since $|\bs\mu |=|\bs\nu|$. Thus $\bs\mu =\bs\nu$, and the statement 
898: is shown. 
899: 
900: \noindent {\it ii)}
901: This is seen by an explicit calculation using eqn.~\eqref{form:d1}. Successively applying 
902: derivatives w.r.t. $u_0$, $u_1$, $\ldots,u_M$ yields eqn.~\eqref{form:anuu}. Together 
903: with the $q$-shift property ${\partial v_k}/{\partial u_k}({\bs u_0})=1$, it follows that 
904: $A(\bs\nu,{\bs u_0})=1$. Statement \textit{iii)} is shown by an analogous calculation.
905: 
906: \noindent {\it iv)} Consider first the case $|\bs\mu |=|\bs\nu|$. Note that, by differentiation, 
907: the numbers $\nu_k$ do not contribute to $\mu_{k+1},\ldots,\mu_M$, for $k\in\{0,\ldots,M\}$. 
908: Thus, the contribution of $\nu_k$ to $\alpha_{\bs\mu }$ is maximal if $\mu_k=\nu_k$ for 
909: $k=0,\ldots,M$.  We get $\bs\mu =\bs\nu$ and $\alpha_{\bs\mu }=\alpha_{\bs\nu}$. Since 
910: this maximum is unique, $\bs\mu \prec \bs\nu$ implies $\alpha_{\bs\mu }<\alpha_{\bs\nu}$.
911: Let now $|\bs\mu |<|\bs\nu|$. For $k\in\{0,\ldots,M\}$, denote by ${\widetilde \nu}_k$ the 
912: number of derivatives w.r.t.~$u_k$, which contribute to $\bs\mu $. Set $\widetilde{\bs\nu}=
913: ({\widetilde \nu}_0, \ldots, {\widetilde \nu}_M)$. Then clearly $|\bs\mu |=|\widetilde{\bs\nu}|$ 
914: and $\alpha_{\widetilde{\bs\nu}}<\alpha_{\bs\nu}$. The reasoning for the case $|\bs\mu |=
915: |\bs\nu|$ can now be applied to the present case, with $\bs\nu$ replaced by $\widetilde{\bs\nu}$.
916: This yields $\alpha_{\bs\mu }\le\alpha_{\widetilde{\bs\nu}}$. Thus $\alpha_{\bs\mu }<
917: \alpha_{\bs\nu}$, and the statement is shown.
918: \end{proof}
919: 
920: \begin{defi}[$q$-functional equation]\label{defi:qfunc}
921: Let $P(\bs u, y_1,\ldots, y_N)$ be a formal power series $P(\bs u,
922: \bs y)\in\mathcal{H}_{d}(\bs u_+)[[u_0,\bs y]]$, for
923: a number $d$ satisfying $1<d\le\infty$. Assume that 
924: $P(0, {\bs u}_+, \bs 0)\equiv 0$, and that $\frac{\partial P}{\partial y_j}(0, 
925: {\bs u}_+,\bs 0) \equiv0$ for $j=1,\ldots,N$. Let ${\bs v^{(\it j)}}$ be a $q$-shift
926: with domain $d(\bs v^{(j)})$ and range $r(\bs v^{(j)})\le d$, for $j=1,\ldots,N$.
927: 
928: \begin{itemize}
929: 
930: \item[i)]
931: 
932: If the above assumptions are satisfied, the equation
933: %
934: \begin{equation}\label{form:funceq}
935: G({\bs u})=P(\bs u, H^{(1)}({\bs u)},\ldots, H^{(N)}
936: ({\bs u)}),
937: \end{equation}
938: %
939: where $H^{(j)}({\bs u)}=G({\bs v^{(\it j)}(\bs u)})$ for $j=1,\ldots, N$,
940: is called a {\em $q$-functional equation}.
941: \item[ii)]
942: Let the above assumptions be satisfied. If there is a number $d(G)$ such that
943: %
944: \begin{equation*}
945: 1< d(G)\le \min_{1\le j\le N}\{d(\bs v^{(j)})\},
946: \end{equation*}
947: %
948: and a formal power series $G(\bs u)\in\mathcal{H}_{d(G)}(\bs 
949: u_+)[[u_0]]$ satisfying eqn.~\eqref{form:funceq}, then $G(\bs u)$ 
950: is called a \emph{solution} of the $q$-functional equation.
951: \end{itemize}
952: \end{defi}
953: 
954: \noindent \textbf{Remarks.} 
955: \textit{i)} An example of a $q$-functional equation for $M=1$ is 
956: given by the $q$-difference equation eqn.~\eqref{form:geneq2}. 
957: This motivates the name for the generalisation. Examples for $M>1$ 
958: appear in eqn.~\eqref{form:geneq3}. Examples of $q$-functional equations 
959: frequently satisfy $P(\bs u, \bs y)\in\mathbb C[\bs u, \bs y]$ 
960: with $d=\infty$, and with monomial $q$-shifts. We infer from the 
961: functional equation eqn.~\eqref{eqn:dyckfeq} for Dyck paths that 
962: the power series $E({\bs u})=D({\bs u})-1$ satisfies the 
963: $q$-functional equation
964: %
965: \begin{equation}\label{form:exfunc}
966: E({\bs u}) = u_0u_1\cdot\ldots\cdot u_M\left( E({\bs u})+1\right)
967: \left( E({\bs v(\bs u)})+1\right).
968: \end{equation}
969: %
970: Specialising to ${\bs u}={\bs u_0}$ yields a quadratic equation, which
971: can be explicitly solved for $E({\bs u_0})$. We get the well-known
972: result
973: %
974: \begin{equation*}
975: E({\bs u_0})=\frac{1-2u_0-\sqrt{1-4u_0}}{2u_0}.
976: \end{equation*}
977: 
978: \noindent \textit{ii)} If $Q(u,y)\in\mathbb 
979: C[[u,y]]$ satisfies $Q(0,0)=0$ and $\frac{\partial Q}{\partial y}(0,y)
980: \not\equiv1$, a formal power series $G(u)\in\mathbb C[[u]]$ such that 
981: $G(0)=0$ is uniquely defined as the solution of the equation 
982: $G(u)=Q(u,G(u))$. When studying such 
983: equations, we may assume without loss of generality that $\frac{\partial Q}
984: {\partial y}(0,y)\equiv0$. The above definition reduces to this setup, when 
985: restricting to $\bs u=\bs u_0$. \\
986: 
987: 
988: \medskip
989: 
990: A result about solutions of $q$-functional equations is given by the following proposition.
991: We use the vector notation ${\bs H(\bs u)}=(H^{(1)}({\bs 
992: u)},\ldots, H^{(N)}({\bs u)})$. 
993: 
994: \begin{prop}\label{prop:uni}
995: The $q$-functional equation of Definition~\ref{defi:qfunc},
996: %
997: \begin{equation}\label{form:func}
998: G({\bs u})=P(\bs u, \bs H(\bs u)),
999: \end{equation}
1000: %
1001: has a unique solution $G({\bs u}) \in \mathcal{H}_{d(P)}(\bs u_+)[[u_0]]$
1002: satisfying $G(0,\bs u_+)\equiv0$, where $d(P)=\min_{1\le j\le N}\{d(\bs v^{(j)})\}$.
1003: \end{prop}
1004: 
1005: \begin{proof}
1006: If $G({\bs u})=\sum_{n_0\ge1} p_{n_0}({\bs u}_+)u_0^{n_0}$, then
1007: $G({\bs v}^{(\it j)}({\bs u}))=\sum_{n_0\ge1} p_{n_0}({\bs v}_+^{(\it j)}
1008: ({\bs u}))\left(v_0^{(\it j)}({\bs u})\right)^{n_0}$. For $n_0\ge1$ fixed, we take the coefficient 
1009: of $u_0^{n_0}$ in eqn.~\eqref{form:func}. Due to $v_0^{(j)}(0,{\bs u}_+)
1010: \equiv0$ for $j=1,\ldots,N$ and the assumptions on the derivatives of $P({\bs y},{\bs u})$, we 
1011: get the expression
1012: %
1013: \begin{equation}\label{form:Wr}
1014: p_{n_0}({\bs u}_+)=W_{n_0}\left({\bs u}_+, \left\{ p_1({\bs v}_+^{(\it j)}
1015: ({\bs u})), \ldots,p_{n_0-1}({\bs v}_+^{(\it j)}({\bs u}) )\right\}_{j=1}^N
1016: \right),
1017: \end{equation}
1018: %
1019: for a power series $W_{n_0}\left({\bs u}_+, \bs p\right)\in\mathcal{H}_{d(P)}
1020: (\bs u_+)\left[\bs p\right]$ in $M+N(n_0-1)$ variables. Thus $p_{n_0}
1021: ({\bs u}_+)$ is determined recursively in terms of $p_l({\bs u}_+)$, 
1022: where $l=1,\ldots, n_0-1$. The recursion also shows that $p_{n_0}({\bs 
1023: u}_+)\in \mathcal{H}_{d(P)}(\bs u_+)$. Thus $G({\bs u}) \in \mathcal{H}_{d(P)}
1024: (\bs u_+)[[u_0]]$, and the proposition is proved.
1025: \end{proof}
1026: 
1027: \noindent \textbf{Remarks.} 
1028: \noindent \textit{i)} For a solution $G(\bs u)$ of a $q$-functional 
1029: equation satisfying $G(\bs0)=0$, we will always assume $d(G)=d(P)$ in the following.\\
1030: \textit{ii)} For the solution of a simple $q$-functional 
1031: equation, an explicit expression may be given. This is e.g. the
1032: case for some $q$-difference equations eqn.~\eqref{form:geneq2}, with $P(x,y)$
1033: linear in $y$, see e.g.~\cite{Bou96}, or with $P(x,y)$ quadratic in $y$, see \cite{PB95a}. \\
1034: \textit{iii)} It follows from $G({\bs u}) \in \mathcal{H}_{d(P)}
1035: (\bs u_+)[[u_0]]$ that $G({\bs u_0})\in\mathbb C[[u_0]]$. For derivatives of 
1036: $G({\bs u})$, which are also elements of $\mathcal{H}_{d(P)}(\bs u_+)[[u_0]]$, the 
1037: same conclusion holds. Thus, the derivatives
1038: %
1039: \begin{equation}\label{form:facarmom}
1040: g_{\bs\nu}(u_0) :=\left.\frac{1}{\bs\nu!}
1041: G_{\bs\nu}({\bs u})\right|_{\bs u=\bs u_0}
1042: \end{equation}
1043: %
1044: are formal power series, i.e., $g_{\bs\nu}(u_0)\in\mathbb C[[u_0]]$. They 
1045: are called \emph{factorial moment generating functions}, for reasons to 
1046: be explained in section \ref{sec:mom}.
1047: 
1048: \section{Factorial moment generating functions}\label{sec:gk}
1049: 
1050: Due to the following proposition, the factorial moment generating functions 
1051: can be computed recursively, by successively differentiating the $q$-functional 
1052: equation.
1053: 
1054: \begin{prop}\label{prop:solv}
1055: Let a $q$-functional equation eqn.~\eqref{form:funceq} with solution 
1056: $G(\bs u)$ as in Proposition~\ref{prop:uni} be given. Consider the derivative of order
1057: $\bs\nu\ne {\bf 0}$ of the $q$-functional equation, 
1058: evaluated at $\bs u=\bs u_0$. It is linear in 
1059: $g_{\bs\nu}(u_0)$. Its r.h.s.~is a polynomial in 
1060: $\{g_{\bs\mu }(u_0):\mathbf{0}\neq\bs
1061: \mu \preceq \bs\nu\}$, with coefficients in 
1062: $\mathbb C[[u_0,g_{\bs 0}(u_0)]]$.
1063: \end{prop}
1064: 
1065: In order to prove Proposition \ref{prop:solv}, we analyse partial 
1066: derivatives of eqn.~\eqref{form:funceq}. To this end, we employ a generalization 
1067: of Faa di Bruno's formula \cite{CS96}, adapted to our situation. The following 
1068: lemma will also be used for exponent estimates in the next section. For 
1069: $k\in\{1,\ldots,K\}$, let $f_k(\bs u)\in\mathbb C[[\bs u]]$, and 
1070: define ${\bs f(\bs u)}=(f_1({\bs u}), \ldots,
1071: f_K({\bs u}))$. For $\bs\mu \in\mathbb N_0^{1+M}$, we 
1072: use the notation ${\bs f(\bs u)}_{\bs\mu} =
1073: ((f_1)_{\bs\mu}({\bs u}),\ldots,(f_K)_{\bs\mu}
1074: ({\bs u}))$ for the derivative of $\bs f(\bs u)$ of order $\bs \mu$.
1075: 
1076: \begin{lemma}{\rm (cf.~\cite[Thm.~2.1]{CS96})}\label{lemma:faa}
1077: Let a $q$-functional equation eqn.~\eqref{form:funceq} with solution 
1078: $G(\bs u)$ as in Proposition~\ref{prop:uni} be given.
1079: Its derivative of order $\bs\nu\ne\bs0$ satisfies
1080: %
1081: \begin{equation}\label{form:faa}
1082: G_{\bs\nu}({\bs u})=\sum_{1\le|\bs\lambda|\le |\bs\nu|}
1083: P_{\bs\lambda}(\bs u, \bs H(\bs u))\sum_{s=1}^{|\bs\nu|}
1084: \sum_{p_s(\bs\nu,\bs\lambda)}
1085: (\bs\nu!) \prod_{j=1}^s\frac{[(\bs u, \bs H(\bs u))_{\bs\mu _j}]^{\bs\kappa_j}}
1086: {(\bs\kappa_j!)[\bs\mu _j!]^{|\bs\kappa_j|}},
1087: \end{equation}
1088: %
1089: where the vectors $\bs\lambda$ and $\bs\kappa_j$ are $1+M+N$-dimensional, 
1090: the vectors $\bs\mu _j$ are $1+M$-dimensional, for $j\in\{1,\ldots,s\}$, 
1091: and the summation ranges over
1092: %
1093: \begin{equation*}
1094: \begin{split}
1095: p_s(\bs\nu,\bs\lambda)=\{&(\bs\kappa_1,\ldots, \bs\kappa_s; 
1096: \bs\mu _1,\ldots, \bs\mu _s): |\bs\kappa_i|>0,\\
1097: &{\bf 0}\lhd \bs\mu _1\lhd \ldots \lhd \bs\mu _s, 
1098: \sum_{i=1}^s \bs\kappa_i=\bs\lambda \mbox{ and } \sum_{i=1}^s |\bs\kappa_i| 
1099: \bs\mu _i= \bs\nu \}.
1100: \end{split}
1101: \end{equation*}
1102: %
1103: In the above equation, the total order $\lhd$ is defined as 
1104: $\bs\mu \lhd\bs\nu$ if either $|\bs\mu |<|\bs\nu|$, or
1105: if $|\bs\mu |=|\bs\nu|$, there exists an index $k\in\{0,\ldots,M\}$ such that
1106: $\mu_k<\nu_k$ and $\mu_i=\nu_i$ for $i=0,\ldots,k-1$. 
1107: \qed
1108: \end{lemma}
1109: 
1110: \noindent \textbf{Remark.} If $M=1$, we have $(\mu_0,\mu_1)\lhd(\nu_0,\nu_1)$
1111: if and only if $(\mu_1,\mu_0)\prec(\nu_1,\nu_0)$. The analogous statement for
1112: higher values of $M$ is not true.
1113: 
1114: \begin{proof}[Proof of Proposition \ref{prop:solv}.]
1115: For given $\bs\nu\neq\bs0$, we analyse the values $\bs\mu _j$ appearing in
1116: Lemma \ref{lemma:faa}. We show that $|\bs\mu _j|\ge|\bs\nu|$ for
1117: some value $j$, where $1\le j\le s \le |\bs\nu|$, implies $s=1$ and 
1118: $\bs\mu _1=\bs\nu$. Together with Lemma \ref{lemma:H}, this implies that 
1119: $\bs\mu \preceq \bs\nu$. The linearity will follow from an explicit 
1120: calculation of the term containing $\bs\mu _1=\bs\nu$.
1121: 
1122: \smallskip
1123: 
1124: Assume that $|\bs\mu _j|\ge|\bs\nu|$ for some $j$, where 
1125: $1\le j\le s\le |\bs\nu|$. The explicit form of 
1126: $p_s(\bs\nu,\bs\lambda)$ in Lemma \ref{lemma:faa}
1127: states that $\bs\nu=\sum_{i=1}^s |\bs\kappa_i| \bs\mu _i$. This implies 
1128: that $|\bs\nu|=\sum_{i=1}^s |\bs\kappa_i| |\bs\mu _i|$. According 
1129: to the assumption, this leads to $s=1$,  $|\bs\kappa_1|=1$ and 
1130: $|\bs\mu _1|=|\bs\nu|$. This, in turn, implies that $\bs\mu _1=\bs\nu$.
1131: 
1132: Clearly, the r.h.s.~of eqn.~\eqref{form:faa}, when specialised to $\bs u=\bs u_0$,
1133: is a polynomial in $g_{\bs\mu }(u_0)$ for $\mathbf{0}\neq\bs\mu\preceq\bs\nu$,
1134: with coefficients in $\mathbb C[[u_0,g_{\bs 0}(u_0)]]$. 
1135: To show linearity in $G_{\bs\nu}({\bs u_0})$, we note that 
1136: the possible values of $\bs\kappa_1$ are $\bs\kappa_1={\bs e}_j$, 
1137: where $j\in\{1,\ldots,1+M+N\}$. The sum of the terms with $|\bs\mu _1|=|\bs\nu|$
1138: in the r.h.s.~of eqn.~\eqref{form:faa} is given by
1139: %
1140: \begin{equation*}
1141: \sum_{j=1}^{1+M+N} P_{{\bs e}_j}({\bs u, \bs H(\bs u)})
1142: [({\bs u, \bs H(\bs u)})_{\bs\nu}]^{{\bs e}_j}.
1143: \end{equation*}
1144: %
1145: We now extract terms containing $G_{\bs\nu}({\bs u})$ from this 
1146: expression, using Lemma \ref{lemma:H}, and group them to the l.h.s.~of 
1147: eqn.~\eqref{form:faa}. In the resulting equation, the l.h.s.~$L(\bs u)$, 
1148: when specialised to $\bs u=\bs u_0$, is given by
1149: %
1150: \begin{equation}\label{form:pref}
1151: L(\bs u_0) =\left(1- \sum_{j=1}^{N} \frac{\partial P}{\partial y_j}
1152: ({\bs u_0, \bs G(\bs u_0)})\right) G_{\bs\nu}({\bs u_0}).
1153: \end{equation}
1154: %
1155: Due to the assumptions on $P(\bs u,\bs y)$, the prefactor
1156: of $G_{\bs\nu}({\bs u_0})$ in the above equation is not
1157: identically vanishing. Thus, $G_{\bs\nu}({\bs u_0})$ is contained linearly in 
1158: eqn.~\eqref{form:faa}, specialised to $\bs u=\bs u_0$.
1159: \end{proof}
1160: 
1161: \section{Analytic generating functions}\label{sec:ana}
1162: 
1163: Let $Q(u,y)=P(u,1,\ldots,1,y,\ldots,y)$. The power series $G({\bs u}_0)$
1164: satisfies the equation
1165: %
1166: \begin{equation}\label{pgfeqn}
1167: G({\bs u}_0)=Q(u_0,G({\bs u}_0)).
1168: \end{equation}
1169: %
1170: In the following, we specialise the class of $q$-functional equations. We are 
1171: interested in the case  where $G({\bs u}_0)$ is analytic at $u_0=0$, with 
1172: a square root as dominant singularity. This situation is generic for 
1173: combinatorial constructions, see \cite[Thm.~10.6]{O95}, \cite[Prop.~1]{D97}, or
1174: \cite[Sec.~7.4]{FS07}. Throughout the remainder of the article, we employ 
1175: the following assumption.
1176: 
1177: \begin{ass}\label{ass}
1178: Let a $q$-functional equation \eqref{form:funceq} as in Definition~\ref{defi:qfunc} 
1179: be given. Let numbers $r,s$ such that $0<r,s\le\infty$ be given. Denote by
1180: $\mathcal{H}_{r,d,s}(\bs u,\bs y)$ the ring of power series in 
1181: $(\bs u,\bs y)=(u_0,\bs u_+,\bs y)$, which
1182: are convergent if  $|u_0|<r$, if $|u_k|<d$ for $k=1,\ldots, M$, and if $|y_k|<s$ for 
1183: $k=1,\ldots,N$. In addition to the assumptions in Definition~\ref{defi:qfunc},
1184: we assume the following properties.
1185: %
1186: \begin{itemize}
1187: \item[{\it i)}] $P({\bs u},{\bs y})\in\mathcal{H}_{r,d,s}(\bs u, 
1188: \bs y)$, and its Taylor coefficients are non-negative real numbers,
1189: when expanded about $(\bs u,\bs y)=(\bs 0,\bs 0)$.
1190: For $j=1,\ldots,N$, each coordinate series of the $q$-shift 
1191: ${\bs v}^{(j)}({\bs u})$ has non-negative
1192: coefficients only, if expanded about $\bs u=\bs 0$.
1193: 
1194: \item[{\it ii)}] Let $Q(u,y)=P(u,1,\ldots,1,y,\ldots,y)$. There exist numbers $(u_c,y_c)$
1195: satisfying $0<u_c<r$ and $0<y_c<s$, such that 
1196: %
1197: \begin{equation}\label{form:lett}
1198: \begin{split}
1199: Q(u_c,y_c)=&y_c, \qquad \left.\frac{\partial Q}{\partial y}(u_c,y)\right|_{y=y_c}=1, \\
1200: B:=\frac{1}{2}\left.\frac{\partial ^2 Q}{\partial y^2}(u_c,y)\right|_{y=y_c}&>0, \qquad 
1201: C:=\left.\frac{\partial Q}{\partial u}(u,y_c)\right|_{u=u_c}>0.
1202: \end{split}
1203: \end{equation}
1204: %
1205: \item[{\it iii)}] The solution $G(\bs u)$ of the $q$-functional equation as in 
1206: Proposition~\ref{prop:uni} has the property that $G({\bs u_0})=\sum_{n\ge 1} 
1207: p_n u_0^n$ is \emph{aperiodic}, i.e., there exist indices $1\le i<j<k$ such that 
1208: $p_ip_jp_k\ne0$, while $\gcd(j-i,k-i)=1$.
1209: \end{itemize}
1210: \end{ass}
1211: 
1212: \noindent {\bf Remarks.} \textit{i)} When restricting to $\bs u=\bs u_0$,
1213: the above assumptions (together with the assumptions in Definition~\ref{defi:qfunc}) 
1214: reduce to the setup for implicitly defined power series which is usual in enumerative 
1215: combinatorics, see \cite[Thm.~10.6]{O95}, \cite[Prop.~1]{D97}, and \cite[Sec.~7.4]{FS07}. \\
1216: \textit{ii)} Assumption \ref{ass} {\it i)} implies that the coefficients $p_{\bs n}$ in 
1217: $G({\bs u})= \sum_{\bs n\ge 0} p_{\bs n}{\bs u^{\bs n}}$ 
1218: are non-negative. For combinatorial constructions, such positivity assumptions are common. 
1219: However, systems of functional equations, arising from a combinatorial construction 
1220: with positive coefficients, might be reduced to an equation of the above form 
1221: having \emph{negative} coefficients. In that situation, other types of singularity might 
1222: appear. This is, e.g., the case for discrete meanders \cite{NT03}. The methods used in 
1223: this paper can be adapted to treat such cases. $P({\bs u},
1224: {\bs y})\in\mathcal{H}_{r,d,s}(\bs u, \bs y)$ implies that,
1225: for each $(\bs u, \bs y)\in D_r^1\times D_d^M\times D_s^N$, the 
1226: function $P({\bs u}, {\bs y})$ has a convergent series expansion
1227: about $(\bs u, \bs y)$.\\
1228: \textit{iii)} Assumption \ref{ass} {\it ii)} implies that the dominant singularity of 
1229: $G({\bs u}_0)$ is a square root. This type of singularity is generic for
1230: functional equations with positive coefficients. If $P({\bs u},
1231: {\bs y})$ is a \emph{polynomial} in $\bs u$ and $\bs y$, it
1232: can be shown that Assumption \ref{ass} {\it ii)} follows from Assumption \ref{ass} 
1233: {\it i)}, if $Q(u,y)$ is of degree $\ge2$ in $y$ and if $Q(u,0)\not\equiv0$.
1234: By the closure properties of algebraic functions, it then follows, with an adaption
1235: of Proposition \ref{prop:solv}, that all factorial moment generating functions are algebraic.\\
1236: \textit{iv)} Assumption \ref{ass} {\it iii)} ensures that $G({\bs u_0})$ 
1237: has exactly one singularity on its circle of convergence. The case of periodic
1238: $G(\bs u_0)$ may be treated by a slight extension of this setup.
1239: 
1240: \medskip
1241: 
1242: We investigate analytic properties of $g_{\bf 0}(u_0)$. A function 
1243: $f(u)$ is called \emph{$\Delta$-regular} \cite{FFK04} if it is analytic in the 
1244: \emph{indented disc} $\Delta=\Delta(u_c,\eta,\phi)=\{u:|u|\le u_c+\eta, |\mbox{arg}
1245: (u-u_c)|\ge\phi\}$ for some real numbers $u_c>0$, $\eta>0$ and $\phi$, 
1246: where $0<\phi<\pi/2$. Note that $u_c\notin\Delta$, where we employ the 
1247: convention $\mbox{arg}(0)=0$. The set of $\Delta$-regular functions is 
1248: closed under addition, multiplication, differentiation, and integration. Moreover, 
1249: if $f(u)\ne0$ in $\Delta$, then $1/f(u)$ exists in $\Delta$ and is $\Delta$-regular. The following
1250: proposition is a straightforward extension of a well-known result, see 
1251: e.g.~\cite[Thm.~10.6]{O95}, \cite[Prop.~1]{D97}, or \cite[Sec.~7.4]{FS07}.
1252:  
1253: \begin{prop}[cf.~\cite{O95,D97,FS07}]\label{prop:g0}
1254: Given Assumption \ref{ass}, the power series $g_{\bf 0}(u_0)$ is analytic at $u_0=0$, 
1255: with radius of convergence $0<u_c<\infty$. Its analytic continuation is $\Delta$-regular, 
1256: with a square root singularity at $u_0=u_c$, and a local Puiseux expansion
1257: %
1258: \begin{equation}\label{form:genex0}
1259: g_{\bf 0}(u_0) = g_{\bf 0}(u_c) + \sum_{l=0}^\infty f_{{\bf 0},l}
1260: (u_c-u_0)^{-\gamma_{\bs0}+l/2},
1261: \end{equation}
1262: %
1263: where $\gamma_{\bs0}=-1/2$, $f_{\bs 0,0}=-\sqrt{C/B}$, and $g_{\bf 0}(u_c)=
1264: \lim_{u_0\to u_c^-} g_{\bs0}(u_0)<\infty$.
1265: \qed
1266: \end{prop}
1267: 
1268: \noindent \textbf{Remark.} The coefficients $f_{\bs 0,l}$, also called \emph{amplitudes},
1269: can be computed recursively from the functional equation eqn.~\eqref{pgfeqn}, by inserting 
1270: the representation eqn.~\eqref{form:genex0} into
1271: eqn.~\eqref{pgfeqn} and then expanding the resulting equation in $s=\sqrt{u_c-u_0}$.
1272: This technique will be exploited below, when using the method of dominant balance.
1273: 
1274: \medskip
1275: 
1276: The properties of $G({\bs u_0})$ carry over to the factorial moment generating functions 
1277: $g_{\bs\nu}(u_0)$. We will first analyse the general form of the factorial moment generating 
1278: functions, and later provide explicit values for exponents and leading amplitudes.
1279: 
1280: \begin{prop}\label{prop:pui}
1281: Let Assumption \ref{ass} be satisfied. For $\bs\nu\neq\bs0$ arbitrary, the factorial moment 
1282: generating function $g_{\bs\nu}(u_0)$ is analytic at $u_0=0$, 
1283: with radius of convergence $0<u_c<\infty$. Its analytic continuation is $\Delta$-regular.
1284: It has a local Puiseux expansion
1285: %
1286: \begin{equation}\label{form:genex}
1287: g_{\bs\nu}(u_0) = \sum_{l=0}^\infty f_{{\bs\nu},l}(u_c-u_0)^{-\gamma_{\bs\nu}+l/2},
1288: \end{equation}
1289: %
1290: with non-vanishing leading amplitude $f_{{\bs\nu},0}\ne0$ and exponent 
1291: $\gamma_{\bs\nu}\in \frac{1}{2}\mathbb Z$.
1292: \end{prop}
1293: 
1294: \begin{proof}
1295: We prove the proposition by induction on $\bs\nu$ w.r.t.~the total order 
1296: $\prec$ in Definition~\ref{def:totord}. Note that the proof of Proposition 
1297: \ref{prop:solv} yields
1298: %
1299: \begin{equation}\label{form:gl1}
1300: g_{\bs\nu}(u_0) Q_1(u_0,g_{\bf 0}(u_0))=Q_{\bs\nu}( 
1301: u_0, g_{\bf 0}(u_0),\{ g_{\bs\mu}(u_0)\}_{\bs 0 \ne\bs\mu \prec\bs\nu}),
1302: \end{equation}
1303: %
1304: where $Q_1(u,y)=1-\frac{\partial Q}{\partial y}(u,y)$ is analytic for 
1305: $|u|<r$, $|y|<s$. The function $Q_{\bs\nu}( u,y,\widetilde 
1306: {\bs y})$ is a polynomial in $\widetilde {\bs y}$ and 
1307: analytic for $|u|<r$, $|y|<s$. Due to the closure properties of $\Delta$-regular and 
1308: analytic functions, both $Q_1(u_0,g_{\bf 0}(u_0))$ and $Q_{\bs\nu}( 
1309: u_0,g_{\bf 0}(u_0), \{ g_{\bs\mu}(u_0)\}_{\bs 0 \ne\bs\mu 
1310: \prec\bs\nu})$ are $\Delta$-regular. Due to Proposition \ref{prop:g0}, we 
1311: have $Q_1(u_0, g_{\bf 0}(u_0))\ne0$ in $\Delta$, thus its inverse is $\Delta$-regular. 
1312: This implies that $g_{\bs\nu}(u_0)$ is $\Delta$-regular. 
1313: 
1314: Due to Proposition \ref{prop:g0}, the series $h(s)=y_c+\sum_{l=0}^\infty 
1315: f_{{\bf 0},l}s^{(l+1)}$ is holomorphic at $s=0$ and equals $g_{\bf 0}(u_0)$ 
1316: about $u_0=u_c$, if $s=\sqrt{u_c-u_0}$. We show that eqn.~\eqref{form:gl1} 
1317: leads to local expansions of $g_{\bs\nu}(u_0)$ in terms of 
1318: $s=\sqrt{u_c-u_0}$, which are meromorphic at $s=0$. Consider first the term 
1319: $Q_1(u_0,g_{\bf 0}(u_0))$. The function $Q_1(u,y)$ is holomorphic at $(u,y)=
1320: (u_c,y_c)$. Thus, as composition of holomorphic functions, $\widetilde 
1321: Q_1(s):=Q_1(u_c-s^2,h(s))$ is holomorphic at $s=0$. We have $\widetilde 
1322: Q_1(0)=0$ and $\widetilde Q'_1(0)=-\frac{\partial Q_1}{\partial y}(u_c,y_c) h'(0)
1323: \ne0$. Thus, its inverse  $1/\widetilde Q_1(s)$ is meromorphic at $s=0$, 
1324: with a simple pole. Consider finally the function $Q_{\bs\nu}
1325: (u,y,\widetilde {\bs y})$. It is a polynomial in $\widetilde 
1326: {\bs y}$ and analytic at $(u,y)=(u_c,y_c)$. Thus, after inserting 
1327: the expansions of $g_{\bs\mu}(u_0)$, where we use Proposition 
1328: \ref{prop:g0} and the induction hypothesis, we conclude that 
1329: it is meromorphic in $s$ at $s=0$. 
1330: It follows that $g_{\bs\nu}(u_0)$ is meromorphic in $s$ at $s=0$. 
1331: Since $G(\bs u)$ is non-negative and $g_{\bs0}(u_0)$ 
1332: does not vanish identically, $g_{\bs\nu}(u_0)$ does not vanish identically.  
1333: We thus have local expansions \eqref{form:genex} of $g_{\bs\nu}(u_0)$.
1334: \end{proof}
1335: 
1336: \noindent {\bf Remarks.} \textit{i)} 
1337: It was argued in the preceding proof that $1-\sum_{j=1}^{N} \frac{\partial P}{\partial y_j}
1338: (\bs u_0,{\bs G(\bs u_0)})=\widetilde Q_1(s)$ satisfies $\widetilde Q_1(s)\sim A
1339: \sqrt{u_c-u_0}$ as $u_0\to u_c^{-}$, for some constant $A\ne0$, see also 
1340: eqn.~\eqref{form:pref}. This will be used in the proof of the following proposition. \\
1341: \textit{ii)} The exponents $\gamma_{\bs\nu}$ and, in principle,
1342: all amplitudes $f_{\bs \nu,l}$ in the Puiseux expansion \eqref{form:genex} of 
1343: $g_{\bs\nu}(u_0)$ can be computed recursively from the functional equation 
1344: eqn.~\eqref{pgfeqn}, compare the argument above for the case $\bs\nu=0$. 
1345: Our aim in this section is to determine the exponent $\gamma_{\bs\nu}$ and 
1346: the amplitudes $f_{{\bs\nu},0}$. The method of Section \ref{sec:dom} will allow to 
1347: also obtain information about the amplitudes $f_{{\bs\nu},l}$ for higher values of $l$,
1348: with reasonable effort.
1349: 
1350: \medskip
1351: 
1352: We will first give an estimate of $\gamma_{\bs\nu}$.
1353: 
1354: \begin{prop}\label{prop:est}
1355: Let Assumption \ref{ass} be satisfied. The exponent $\gamma_{\bs\nu}$ in 
1356: the Puiseux expansion of $g_{\bs\nu}(u_0)$ eqns.~\eqref{form:genex0}, \eqref{form:genex},
1357: satisfies the estimate
1358: \begin{equation}
1359: \gamma_{\bs\nu}\le -\frac{1}{2}+\sum_{i=0}^M \left( 1+\frac{i}{2}\right)\nu_i.
1360: \end{equation}
1361: \end{prop}
1362: 
1363: \noindent {\bf Remark.} Under mild additional assumptions, the above estimate
1364: is sharp, i.e., we have $\gamma_{\bs\nu}= -\frac{1}{2}+\sum_{i=0}^M 
1365: \left( 1+\frac{i}{2}\right)\nu_i$. This follows from Proposition \ref{prop:rek}, 
1366: where we show that the amplitudes $f_{{\bs\nu},0}$, for the special choice of 
1367: $\gamma_{\bs\nu}$ as above, have non-zero values. 
1368: Similar estimates can be obtained in situations different from a square root 
1369: singularity. The proof below can be adapted to treat these situations.
1370: 
1371: \begin{proof}
1372: Set ${\widetilde\gamma_{\bs\nu}}
1373: =-r+\sum_{i=0}^M r_i\nu_i$, where $r=1/2$, and $r_i=(1+i/2)$ for $i=0,\ldots,M$.
1374: We have $r_{k+1}>r_k$ for $k=0,\ldots,M-1$. For 
1375: $\bs\nu=\bf0$ and $\bs\nu={\bs e}_0$, we get 
1376: from Proposition \ref{prop:g0} that $\gamma_{\bs\nu}=
1377: {\widetilde\gamma_{\bs\nu}}$.
1378: We prove the proposition by induction on $\bs\nu$ w.r.t.~the total order 
1379: $\prec$ of Definition~\ref{def:totord}, using Proposition \ref{prop:pui}.   
1380: For the remaining induction step, it suffices to show that $\gamma_{\bs\nu}
1381: \le{\widetilde\gamma_{\bs\nu}}$.
1382: 
1383: \smallskip
1384: 
1385: Assume that $\gamma_{\bs\mu }\le{\widetilde\gamma_{\bs\mu }}$ 
1386: holds for $\bs\mu \prec\bs\nu$. We group all terms in 
1387: eqn.~\eqref{form:faa}, evaluated at $\bs u=\bs u_0$, 
1388: which contain $g_{\bs\nu}(u_0)$, to the l.h.s. In the
1389: resulting equation, its l.h.s.~$L(s)$ satisfies asymptotically, 
1390: with $c_L$ a non-zero constant, $L(s)\sim c_L(u_c-u_0)^{-\gamma_{\bs
1391: \nu}+1/2}$ as $u_0\to u_c^-$, see the remark following the proof of 
1392: Proposition \ref{prop:pui}. By the reasoning in the proof of Proposition 
1393: \ref{prop:pui}, the r.h.s.~$R(s)$ of the equation satisfies asymptotically, 
1394: with $c_R$ a nonzero constant, $R(s)\sim c_R(u_c-u_0)^{-\gamma}$ as $u_0\to 
1395: u_c^-$. Using the induction assumption, the exponent $\gamma$ can be estimated by
1396: %
1397: \begin{equation*}
1398: \gamma \le \sum_{i=1}^s \gamma_{\bs\mu _i} |\bs\kappa_i|
1399: \le \sum_{i=1}^s \left( -\frac{1}{2} + \sum_{j=0}^M r_j (\bs\mu _{i})_j\right)|
1400: \bs\kappa_i|=  -\frac{|\bs\lambda|}{2}+\sum_{j=0}^M r_j\nu_j = 
1401: - \frac{(|\bs\lambda|-1)}{2}+\widetilde\gamma_{\bs\nu}.
1402: \end{equation*}
1403: %
1404: In the above expression, we used Lemma \ref{lemma:H} and the properties of 
1405: $p_s(\bs\nu,\bs\lambda)$ in Lemma \ref{lemma:faa}.  Since 
1406: $L(s)$ equals $R(s)$, we have $\gamma_{
1407: \bs\nu}-1/2=\gamma$. Thus, the estimate $\gamma_{\bs\nu}\le
1408: {\widetilde\gamma_{\bs\nu}}$ is satisfied for $|\bs\lambda |\ge2$. 
1409: Let us analyse the terms contributing to $|\bs\lambda|=1$ in Lemma 
1410: \ref{lemma:faa}. We have $s=1$, $|\bs\kappa_1|=1$ and $\bs\mu _1
1411: =\bs\nu$. Now, use Lemma \ref{lemma:H}. If in eqn.~\eqref{form:chain} 
1412: we have $|\bs\mu |=|\bs\nu|-l$ for some $l\in\{1,\ldots, |\bs\nu|\}$, 
1413: it follows that $\widetilde\gamma_{\bs\mu }\le \widetilde\gamma_{\bs\nu}
1414: -l\min_{0\le k\le M}\{r_k\}=\widetilde\gamma_{\bs\nu}-l$. Thus, terms with 
1415: exponent larger than $\widetilde\gamma_{\bs\nu}-1/2$ must have $l=0$. 
1416: If $|\bs\mu |=|\bs\nu|$ but $\bs\mu \ne \bs\nu$, 
1417: we have $\widetilde\gamma_{\bs\mu }\le \widetilde\gamma_{\bs\nu}-
1418: |r_j-r_k|$ for some indices $j, k\in\{0,\ldots, M\}$ satisfying $j\ne k$. 
1419: This means that the exponent estimate $\gamma_{\bs\nu}-1/2\le-1/2+
1420: \widetilde\gamma_{\bs\nu}$ is 
1421: satisfied as long as $|r_j-r_k|\ge 1/2$ for $j, k\in\{0,\ldots,M\}$ and $j\ne k$. This 
1422: condition is satisfied for the particular choice of exponents $r_i=(1+i/2)$ of 
1423: the proposition. Thus, the proposition has been proved.
1424: \end{proof}
1425: 
1426: The reasoning in the proof of Proposition \ref{prop:est} can be refined in 
1427: order to obtain a recursion for the amplitudes $f_{{\bs\nu},0}$. 
1428: For vectors $\bs x=(x_1,\ldots, x_K)\in\mathbb R^K$ and
1429: $\bs y=(y_1,\ldots,y_K)\in\mathbb R^K$, we write 
1430: ${\bs x}\le {\bs y}$ if $x_i\le y_i$
1431: for $i=1,\ldots,K$. In the proof, we will use the ``large oh'' symbol:
1432: For functions $f(x)$ and $h(x)$, we write $f(x)={\cal O}(h(x))$ as 
1433: $x\to x_0$ in a domain $D$, if there exists a positive constant $C$ and 
1434: a neighbourhood $N(x_0)$ of $x_0$ such that $|f(x)|\le C |h(x)|$ for 
1435: all $x\in N(x_0)\cap D$.  
1436: 
1437: \begin{prop}\label{prop:rek}
1438: Let Assumption \ref{ass} be satisfied and fix $\bs\nu\in\mathbb N_0^{1+M}$.
1439: Then, the amplitudes $f_{\bs\nu,0}=f_{\bs\nu}$ of the
1440: Puiseux expansion eqn.~\eqref{form:genex} are, if ${\bs\nu}\ne {\bf 0}$ and 
1441: ${\bs\nu}\ne {\bs e}_0$, determined by the recursion
1442: %
1443: \begin{equation}\label{form:coefrek}
1444: f_{\bs\nu} = \mu_0 \gamma_{{\bs\nu}-{\bs e}_1}
1445: f_{{\bs\nu}-{\bs e}_1}+\sum_{i=1}^{M-1}\mu_i(\nu_i+1)
1446: f_{{\bs\nu}-{\bs e}_{i+1}+{\bs e}_i}-\frac{1}
1447: {2f_{\bf 0}}\sum_{\substack{{\bs\rho}\ne {\bf 0},{\bs\rho}
1448: \ne {\bs\nu}\\ {\bf 0}\le{\bs\rho}\le{\bs\nu}}}
1449: f_{\bs\rho}f_{{\bs\nu}-{\bs\rho}},
1450: \end{equation}
1451: %
1452: with boundary conditions $f_{\bf 0}=-\sqrt{C/B}<0$, $f_{{\bs e}_0}
1453: =-f_{\bf 0}/2>0$, and $f_{\bs\nu}=0$ if $\nu_j<0$ for some 
1454: $j\in\{1,\ldots,M\}$. We have 
1455: %
1456: \begin{equation*}
1457: \gamma_{\bs\nu}=-\frac{1}{2}+\sum_{i=0}^M \left(1+\frac{i}{2}\right)\nu_i.
1458: \end{equation*}
1459: % 
1460: For $i\in\{0,\ldots,M-1\}$, we have $\mu_i=-A_{i}/
1461: (2Bf_{\bf 0})\ge0$, and the number $A_i\ge0$ is given by
1462: %
1463: \begin{equation}\label{eqn:Ai}
1464: A_i=\sum_{j=1}^{N} \frac{\partial v_i^{(j)}}{\partial u_{i+1}}({\bs u}_c)
1465: \frac{\partial P}{\partial y_j}({\bs u_c, \bs G(\bs u_c)}),
1466: \end{equation}
1467: %
1468: where $\bs u_c=(u_c,1,\ldots,1)$. If $\bs\nu\ne {\bf 0}$, the amplitudes 
1469: satisfy $f_{\bs\nu}\ge0$. This inequality is strict, if $A_i>0$ for $i=0,\ldots,M-1$.
1470: \end{prop}
1471: 
1472: \begin{proof}
1473: The amplitude $f_{\bs0}$ has been determined in Proposition \ref{prop:g0}.
1474: Assume that ${\bs\nu}\ne {\bf 0}$ and ${\bs\nu}\ne {\bs e}_0$.
1475: We group all terms in eqn.~\eqref{form:faa}, evaluated at $\bs u=\bs u_0$,
1476: which contain $g_{\bs\nu}(u_0)$, to the l.h.s.
1477: Using eqns.~\eqref{form:pref} and \eqref{form:facarmom}, the l.h.s.~$L(\bs u_0)$
1478: of the resulting equation is given by
1479: %
1480: \begin{equation*}
1481: L(\bs u_0)=\bs\nu! g_{\bs\nu}(u_0)P_{\bs e_1}({\bs u_0, \bs G(\bs u_0)})/
1482: \left(\frac{{\rm d}G}{{\rm d} u_0}({\bs u}_0)\right).
1483: \end{equation*}
1484: %
1485: Here, we used the identity
1486: %
1487: \begin{equation*}
1488: \frac{{\rm d}G}{{\rm d}u_0}({\bs u}_0)=
1489: \left(\sum_{j=1}^{N} \frac{\partial P}{\partial y_j}({\bs u_0, \bs G(\bs u_0)})
1490: \right)\frac{{\rm d}G}{{\rm d}u_0}({\bs u}_0)+
1491: P_{\bs e_1}({\bs u_0, \bs G(\bs u_0)}),
1492: \end{equation*}
1493: %
1494: which is obtained from differentiating eqn.~\eqref{pgfeqn}.
1495: By the reasoning in the proof of Proposition 
1496: \ref{prop:est}, the r.h.s.~$R(u_0)$ of the resulting equation satisfies asymptotically, 
1497: with $c_R$ a nonzero constant, $R(u_0)\sim c_R(u_c-u_0)^{-(\gamma_{\bs\nu}-1/2)}$ as $u_0\to 
1498: u_c^-$. We collect all terms with exponents $\gamma_{\bs\nu}-1/2$. Due to 
1499: the proof of Proposition \ref{prop:est}, they arise from terms where $|\bs\lambda|=1$ 
1500: or $|\bs\lambda|=2$ in eqn.~\eqref{form:faa}. An explicit analysis using Lemma 
1501: \ref{lemma:H}, whose details we omit, leads to the following expressions. The 
1502: contribution arising from terms where $|\bs\lambda|=2$ is given by
1503: %
1504: \begin{equation*}
1505: \frac{1}{2}\left(\sum_{j,k=1}^{N} \frac{\partial^2P}{\partial y_j\partial y_k}({\bs u_0, \bs 
1506: G(\bs u_0)})\right)\bs\nu!
1507: \sum_{\substack{\bs\rho\ne {\bf 0}, \bs\rho\ne \bs\nu\\{\bf 0}\le 
1508: \bs\rho\le \bs\nu}}g_{\bs\rho}(u_0)g_{\bs\nu-\bs\rho}
1509: (u_0).
1510: \end{equation*}
1511: %
1512: The contribution from terms where $|\bs\lambda|=1$ is given by
1513: %
1514: \begin{equation*}
1515: \sum_{j=1}^{N} \frac{\partial P}{\partial y_j}({\bs u_0, \bs G(\bs u_0)})
1516: \sum_{i=0}^{M-1} (\bs\nu+{\bs e}_i)!
1517: \frac{\partial v_i^{(j)}}{\partial u_{i+1}}({\bs u}_0)
1518: g_{\bs\nu+{\bs e}_i-{\bs e}_{i+1}}(u_0).
1519: \end{equation*}
1520: %
1521: Omitting arguments and normalising the l.h.s., we arrive at the equation
1522: %
1523: \begin{equation*}
1524: \begin{split}
1525: g_{\bs\nu}=&\frac{\sum \frac{\partial^2P}{\partial y_j\partial y_k}}{2P_{\bs e_1}}G'
1526: \left( \sum g_{\bs\rho}g_{\bs\nu-\bs\rho}\right)
1527: +\frac{G'}{P_{\bs e_1}}\sum_{j=1}^N \frac{\partial P}{\partial y_j}
1528: \frac{\partial v_0^{(j)}}{\partial u_1}g_{\bs\nu-{\bs e}_1}'\\
1529: &+\frac{G'}{P_{\bs e_1}}\sum_{j=1}^N \frac{\partial P}{\partial y_j} \sum_{i=1}^M (\nu_i+1)
1530: \frac{\partial v_i^{(j)}}{\partial u_{i+1}} g_{\bs\nu+{\bs e}_i-{\bs e}_{i+1}}
1531: +{\cal O}\left((u_c-u_0)^{-\gamma_{\bs\nu}+1/2}\right)
1532: \end{split}
1533: \end{equation*}
1534: %
1535: as $u_0\to u_c^-$, where $()'$ denotes differentiation w.r.t. $u_0$. We have $G'=g_{{\bs e}_1}$ 
1536: and $f_{{\bs e}_1}=-f_{\bf 0}/2>0$. The implied recursion for the amplitudes $f_{{\bs\nu},0}$
1537: is given by
1538: %
1539: \begin{equation*}
1540: \begin{split}
1541: f_{\bs\nu}=&\frac{\sum  \frac{\partial^2P}{\partial y_j\partial y_k}}{2P_{\bs e_1}}
1542: \left(-\frac{f_{{\bf 0}}}{2}\right)\left( \sum f_{\bs\rho}f_{\bs\nu-\bs\rho}\right)
1543: -\frac{f_{\bf 0}}{2P_{\bs e_1}}\sum_{j=1}^N \frac{\partial P}{\partial y_j}
1544: \frac{\partial v_0^{(j)}}{\partial u_1}\gamma_{\bs\nu-{\bs e}_1} f_{\bs\nu-{\bs e}_1}\\
1545: &-\frac{f_{{\bf 0}}}{2P_{\bs e_1}}\sum_{j=1}^N \frac{\partial P}{\partial y_j}
1546: \sum_{i=1}^M (\nu_i+1)\frac{\partial v_i^{(j)}}{\partial u_{i+1}} f_{\bs\nu+{\bs e}_i-{\bs e}_{i+1}}.
1547: \end{split}
1548: \end{equation*}
1549: %
1550: After rewriting prefactors, using eqn.~\eqref{form:lett} and Proposition 
1551: \ref{prop:pui}, we arrive at eqn.~\eqref{form:coefrek}.
1552: 
1553: We show strict positivity of the amplitudes, if $A_i>0$ for $i=0,\ldots,M-1$. 
1554: For $|{\bs\nu}|\ge2$, all prefactors of $f_{\bs\mu }$ in eqn.~\eqref{form:coefrek} are
1555: non-negative. For the first term, this follows from the estimate
1556: %
1557: \begin{equation*}
1558: \gamma_{{\bs\nu}-{\bs e}_1}
1559: = \left(-\frac{1}{2}-\frac{3}{2}+\sum_{i=0}^M\frac{i+2}{2} \nu_i \right)
1560: \ge -2+|{\bs\nu}|\ge0.
1561: \end{equation*}
1562: %
1563: Moreover, the last sum in eqn.~\eqref{form:coefrek} is over a nonempty set of 
1564: indices $\bs\rho$ and has a strictly positive prefactor. For $|{\bs\nu}|=1$ and 
1565: ${\bs\nu}\ne {\bs e}_1$, the first term in eqn.~\eqref{form:coefrek} is zero, 
1566: whereas the prefactors of $f_{\bs\mu }$ in the second term are strictly positive. 
1567: Note finally that $f_{{\bs e}_1}=\mu_0\gamma_{\bf0}f_{\bf0}>0$, as is readily inferred from 
1568: eqn.~\eqref{form:coefrek}. This leads, by induction, to strictly positive 
1569: amplitudes $f_{\bs\nu}$. The same argument shows that $f_{\bs\nu}\ge{\bf 0}$
1570: for $\bs\nu\ne\bf0$.
1571: \end{proof}
1572: 
1573: \section{Method of dominant balance}\label{sec:dom}
1574: 
1575: We now discuss an altervative method to obtain the recursion for the 
1576: amplitudes $f_{\bs\nu}$ in eqn.~\eqref{form:coefrek}, if 
1577: $\nu_0=0$. It is based on a generating function approach, and generally 
1578: easier to apply than the combinatorial approach in the proof of 
1579: Proposition \ref{prop:rek}, which was based on an application of Faa 
1580: di Bruno's formula. It also allows to analyse corrections to the 
1581: leading singular behaviour of the factorial moment generating functions. 
1582: Introduce parameters $\delta_i=1-u_i$, where $i\in\{1,\ldots,M\}$, and 
1583: set $\bs\delta =(\delta_1, \ldots, \delta_M)$. From now on, we 
1584: consider factorial moment generating functions $g_{\bs\nu}(u_0)$ 
1585: for $\nu_0=0$ only. For ${\bs k}=(k_1,\ldots, k_M)\in\mathbb 
1586: N_0^M$, set $g_{\bs k}(u_0)=g_{(0, \bs k)}(u_0)$, 
1587: $f_{\bs k}(u_0)=f_{(0,\bs k)}(u_0)$, and $\gamma_{
1588: \bs k}(u_0)=\gamma_{(0,\bs k)}(u_0)$.
1589: 
1590: \begin{prop}\label{prop:db1}
1591: Assume that the $q$-functional equation~\eqref{form:funceq} has the solution 
1592: $G(u_0,\bs u_+)\in{\cal H}_{d(P)}(\bs u_+)[[u_0]]$ such that $G(0,\bs u_+)=0$. 
1593: Then ${\widetilde G}(u_0, \bs\delta):=G(u_0,1-\delta_1, \ldots, 1-\delta_M)\in
1594: \mathbb C[[u_0,\bs\delta]]$ is a formal power series, given by
1595: %
1596: \begin{equation*}
1597: {\widetilde G}(u_0,\bs\delta) = 
1598: \sum_{{\bs k} \ge \bs 0} (-1)^{\bs k} 
1599: g_{\bs k}(u_0) {\bs\delta}^{\bs k},
1600: \end{equation*}
1601: %
1602: where $g_{\bs k}(u_0)=g_{(0, \bs k)}(u_0)\in\mathbb C[[u_0]]$ 
1603: are the factorial moment generating functions \eqref{form:facarmom}.
1604: \end{prop}
1605: 
1606: \begin{proof}
1607: Proposition \ref{prop:uni} states that $G({\bs u})\in\mathcal{H}_{d(P)}(\bs u_+)[[u_0]]$.
1608: Thus ${\widetilde G}(u_0,\bs\delta)\in\mathcal{H}_{d(P)-1}(\bs \delta)[[u_0]]\subset\mathbb 
1609: C[[\bs\delta]][[u_0]]=\mathbb C[[u_0,\bs\delta]]$. We thus have
1610: %
1611: \begin{equation*}
1612: {\widetilde G}(u_0,\bs\delta) = 
1613: \sum_{\bs k\ge\bf 0} 
1614: h_{\bs k}(u_0) {\bs\delta}^{\bs k},
1615: \end{equation*}
1616: %
1617: for some $h_{\bs k}(u_0)\in\mathbb C[[u_0]]$. By Taylor's formula, we have
1618: %
1619: \begin{equation*}
1620: h_{\bs k}(u_0)=\left.\frac{1}{\bs k!}
1621: {\widetilde G}_{(0,\bs k)}
1622: (u_0,\bs\delta)\right|_{\bs\delta={\bf  0}}=
1623: \left.\frac{(-1)^{\bs k}}{\bs k!}
1624: G_{(0,\bs k)}({\bs u})\right|_{\bs u= \bs u_0}=
1625: (-1)^{\bs k}
1626: g_{\bs k}(u_0),
1627: \end{equation*}
1628: %
1629: and the proposition is proved.
1630: \end{proof}
1631: 
1632: Let $\mathbb C((s))$ denote the field of formal Laurent series $f(s)=
1633: \sum_{l\ge l_0}f_l s^l$, where $l_0\in\mathbb Z$ and $f_l\in\mathbb C$ 
1634: for $l\ge l_0$. Employing the Puiseux expansions of the factorial moment 
1635: generating functions, we have an alternative representation of the 
1636: generating function $G({\bs u})$.
1637: 
1638: \begin{prop}\label{prop:scalbeh}
1639: Let Assumption \ref{ass} be satisfied.
1640: Replace the coefficients $g_{\bs k}(u_0)$ of ${\widetilde G}
1641: (u_0, \bs\delta)\in\mathbb C[[u_0,\bs\delta]]$ in 
1642: Proposition \ref{prop:db1} by their Puiseux expansions of Propositions 
1643: \ref{prop:g0} and \ref{prop:pui}, and denote the resulting series by 
1644: ${\widetilde G}(s, \bs\delta)\in\mathbb C((s))[[\bs \delta]]$.
1645: It is explicitly given by
1646: %
1647: \begin{equation*}
1648: {\widetilde G}(s, \bs\delta)=G({\bs u_c})+
1649: \sum_{\bs k\ge\bf0}\left(\sum_{l=0}^\infty
1650: (-1)^{\bs k}f_{{\bs k},l}s^{-\gamma_{\bs k}+l/2}
1651: \right) \bs\delta^{\bs k},
1652: \end{equation*}
1653: %
1654: where $\gamma_{\bs k}=-1/2+\sum_{i=1}^M(1+i/2)k_i$,
1655: where the numbers $f_{{\bs k},l}=f_{(0,{\bs k}),l}$ are 
1656: defined in eqn.~\eqref{form:genex0} and in eqn.~\eqref{form:genex},
1657: and where $\bs u_c=(u_c,1,\ldots,1)$.
1658: 
1659: \smallskip
1660: 
1661: Then, the rescaled series $G({\bs u}(s,\bs\epsilon))=
1662: {\widetilde G}(s,\epsilon_1s^3,\epsilon_2s^4,\ldots,\epsilon_Ms^{M+2})
1663: \in\mathbb C[[s,\bs \epsilon]]$ is a \emph{formal power series},
1664: %
1665: \begin{equation}\label{form:scaling}
1666: G({\bs u}(s,\bs\epsilon)) = G({\bs u_c}) + s F\left(s,
1667: \bs\epsilon\right),
1668: \end{equation}
1669: %
1670: where $F(s,\bs\epsilon)\in\mathbb C[[s, \bs\epsilon]]$ is 
1671: given by
1672: %
1673: \begin{equation}\label{form:Fl}
1674: F(s, \bs\epsilon)=\sum_{l=0}^\infty F_l({\bs\epsilon}) s^l,
1675: \qquad F_l(\bs\epsilon)=\sum_{\bs k\ge\bf0} 
1676: (-1)^{\bs k}f_{{\bs k},l}
1677: {\bs\epsilon}^{\bs k}.
1678: \end{equation}
1679: %
1680: \end{prop}
1681: 
1682: \begin{proof}
1683: Due to Propositions \ref{prop:g0} and \ref{prop:pui}, we have 
1684: $g_{\bs k}(u_0)\in\mathbb C((s))$, where $s=\sqrt{u_c-u_0}$. The 
1685: explicit form \eqref{form:scaling} follows immediately from the Puiseux expansions 
1686: in Propositions \ref{prop:g0} and \ref{prop:pui}, together with
1687: the exponent estimate in Proposition \ref{prop:est}. 
1688: \end{proof}
1689: 
1690: \noindent {\bf Remarks.}
1691: $i)$ Equations like eqn.~\eqref{form:scaling} appear in statistical mechanics as
1692: a so-called \emph{scaling Ansatz}, being an assumption on the behaviour of a generating
1693: function near a multicritical point singularity \cite{J00}, see also \cite{BFG03,KS05}. Its validity 
1694: has been proved only in a limited number of examples. Here, we employ the 
1695: different framework of formal power series.\\
1696: $ii)$ Proposition \ref{prop:scalbeh} suggests an alternative strategy to compute the 
1697: singular behaviour of the factorial moment generating functions. We consider the
1698: functional equation for $F(s,\bs\epsilon)\in\mathbb C[[s, \bs\epsilon]]$,
1699: which is induced by the $q$-functional equation for $G(\bs u)$. Writing
1700: $F(s, \bs\epsilon)=\sum_{l\ge0} F_l({\bs\epsilon}) s^l$,
1701: where $F_l({\bs\epsilon}) \in\mathbb C[[\bs\epsilon]]$, then
1702: leads to a partial differential equation for $F_l({\bs\epsilon})$, upon 
1703: expanding the induced functional equation in powers of $s$. Note that this algorithm,
1704: which is computationally involved, can be easily implemented in a computer algebra 
1705: system. The above method is called in statistical mechanics the method of {\em dominant 
1706: balance}, see \cite{PB95a,R02}. It can be applied in our framework, if an exponent 
1707: bound like that of Proposition \ref{prop:est} is known. Such bounds are generally 
1708: easier to obtain than explicit recursions for amplitudes, see the above proofs.
1709: 
1710: \begin{prop}\label{prop:newfeqn}
1711: Let Assumption \ref{ass} be satisfied.
1712: Define the power series ${\bs u}(s,\bs\epsilon)=(u_0(s,\bs\epsilon), 
1713: \ldots, u_M(s,\bs\epsilon))$, where
1714: %
1715: \begin{equation*}
1716: u_0(s,\bs\epsilon)=u_c-s^2, \qquad 
1717: u_i(s,\bs\epsilon)=1-\epsilon_is^{i+2} \qquad (i=1,\ldots,M).
1718: \end{equation*}
1719: %
1720: For a $q$-shift $\bs v$, define the induced $q$-shift $s_{{\bs v}}$ of $s$ and the 
1721: induced $q$-shift $\bs\epsilon_{{\bs v}}=(\epsilon_{1,{\bs v}}, \ldots, 
1722: \epsilon_{M,{\bs v}})$ of $\bs\epsilon$ by
1723: %
1724: \begin{equation*}
1725: \begin{split}
1726: s_{{\bs v}}&=s_{{\bs v}}(s,\bs\epsilon)=
1727: \sqrt{u_c-v_0({\bs u}(s,\bs\epsilon))},\\
1728: \epsilon_{i,{\bs v}}
1729: &=\epsilon_{i,{\bs v}}(s,\bs\epsilon)=
1730: \frac{1-v_i({\bs u}(s,\bs\epsilon))}{s_{{\bs v}}
1731: (s,\bs\epsilon)^{i+2}} \qquad (i=1,\ldots,M).
1732: \end{split}
1733: \end{equation*}
1734: %
1735: We then have $s_{{\bs v}}\in\mathbb C[[s, \bs\epsilon]]$ and 
1736: $\epsilon_{i,{\bs v}}\in\mathbb C[[s, \bs\epsilon]]$ for $i=1,\ldots,M$.
1737: The functional equation for $F(s, \bs\epsilon)\in\mathbb C[[s, 
1738: \bs\epsilon]]$ of Proposition \ref{prop:scalbeh}, induced by 
1739: eqn.~\eqref{form:funceq}, is given by
1740: %
1741: \begin{equation}\label{form:funcind}
1742: G({\bs u}(s,\bs\epsilon))=P( {\bs u}(s,\bs\epsilon),
1743: {\bs H}({\bs u}(s,\bs\epsilon))).
1744: \end{equation}
1745: %
1746: In the above equation, $G({\bs u}(s,\bs\epsilon))$ is given by 
1747: eqn.~\eqref{form:scaling}, and the power series $H^{(j)}({\bs u}
1748: (s,\bs\epsilon))$ are given by
1749: %
1750: \begin{equation*}
1751: H^{(j)}({\bs u}(s,\bs\epsilon)) = G({\bs u_c})+s_{{\bs v}^{(j)}}
1752: F(s_{{\bs v}^{(j)}},\bs\epsilon_{{\bs v}^{(j)}}) \qquad (1\le j\le N).
1753: \end{equation*}
1754: \end{prop}
1755: 
1756: \begin{proof}
1757: This follows from direct computation. Note first that the Taylor expansions 
1758: of the power series $v_k({\bs u})$ about ${\bs u}={\bs u}_0$ 
1759: are of the form
1760: %
1761: \begin{equation}\label{form:taylvk}
1762: v_k({\bs u})= v_k({\bs u}_0) +
1763: \sum_{\substack{\bs l\ne 0\\ \bs l\ge 0}}
1764: \frac{(-1)^{\bs l}}{\bs l !} v_{k,\bs l}
1765: ({\bs u}_0)(1-u_1)^{l_1}\cdot\ldots\cdot(1-u_M)^{l_M} \qquad (k=0,\ldots,M).
1766: \end{equation}
1767: %
1768: Now insert the parametrisations $u_k(s,\bs\epsilon)\in \mathbb 
1769: C[s,\bs\epsilon]$, where $k=0,\ldots,M$. We get $v_k({\bs u}
1770: (s,\bs\epsilon))\in\mathbb C[[{s,\bs\epsilon}]]$. Due to the 
1771: $q$-shift properties, we have
1772: %
1773: \begin{equation*}
1774: \begin{split}
1775: &v_0({\bs u}(s,\bs\epsilon))=u_c-s^2-\frac{\partial v_0}{\partial u_1}
1776: ({\bs u}_c) \epsilon_1s^3+{\cal O}(s^4),\\
1777: &v_k({\bs u}(s,\bs\epsilon))= 1-\epsilon_ks^{k+2}-
1778: \frac{\partial v_k}{\partial u_{k+1}}({\bs u}_c)\epsilon_{k+1}s^{k+3} + 
1779: {\cal O}(s^{k+4}) \qquad (k=1,\ldots,M-1),\\
1780: &v_M({\bs u}(s,\bs\epsilon))= 1-\epsilon_Ms^{M+2}+ {\cal O}(s^{M+3}).
1781: \end{split}
1782: \end{equation*}
1783: %
1784: We thus have $v_0({\bs u}(s,\bs\epsilon))=u_c-s^2-s^3R_0(s,\bs
1785: \epsilon)$ and $v_k({\bs u}(s,\bs\epsilon))=1-s^{k+2}R_k(s,\bs
1786: \epsilon)$ for $k=1,\ldots,M$, where $R_i(s,\bs\epsilon)\in\mathbb 
1787: C[[s, \bs\epsilon]]$ for $i=0,\ldots,M$. For $s_{{\bs v}}$, we get
1788: %
1789: \begin{equation*}
1790: s_{{\bs v}}=\sqrt{u_c-v_0({\bs u}(s,\bs\epsilon))}=
1791: s\sqrt{1-sR_0(s,\bs\epsilon)}.
1792: \end{equation*}
1793: %
1794: We thus have $s_{{\bs v}}\in\mathbb C[[s, \bs\epsilon]]$ 
1795: and $s_{{\bs v}}(0,\bs\epsilon)=0$. We get
1796: %
1797: \begin{equation*}
1798: \epsilon_{i,{\bs v}}=\frac{1-v_i({\bs u}(s,\bs\epsilon))}
1799: {s_{{\bs v}}^{i+2}}=\frac{R_i(s,\bs\epsilon)}{\sqrt{1-sR(s,
1800: \bs\epsilon)}^{i+2}} \qquad (i=1,\ldots,M).
1801: \end{equation*}
1802: %
1803: It follows that $\epsilon_{i,{\bs v}}\in\mathbb C[[s, \bs\epsilon]]$, 
1804: where $i=1,\ldots,M$. Furthermore, we get from eqn.~\eqref{form:taylvk}
1805: %
1806: \begin{equation*}
1807: \epsilon_{i,{\bs v}}(0,\bs\epsilon)=R_i(0,\bs\epsilon)=
1808: \epsilon_i,
1809: \end{equation*}
1810: % 
1811: such that $\bs\epsilon_{{\bs v}}(0,\bs\epsilon)=\bs\epsilon$. 
1812: This ensures that eqn.~\eqref{form:funcind} is well-defined for  
1813: $F(s, \bs\epsilon)\in\mathbb C[[s, \bs\epsilon]]$.
1814: \end{proof}
1815: 
1816: \begin{prop}\label{theo:pde}
1817: Let Assumption \ref{ass} be satisfied. The formal power series
1818: $F_0(\bs\epsilon)\in\mathbb C[[\bs\epsilon]]$
1819: satisfies the singular, quasi-linear partial differential equation of first order
1820: %
1821: \begin{equation}\label{form:res}
1822: \sum_{j=1}^{N} \frac{\partial P}{\partial y_j}(\bs u_c, \bs G(\bs u_c)) \left( \frac{1}{2}
1823: \frac{\partial v_0^{(j)}}{\partial u_1}({\bs u}_c)\epsilon_1 F_0(\bs\epsilon) +
1824: \sum_{i=1}^M h_i^{(j)}(\bs\epsilon) \frac{\partial}{\partial\epsilon_i}
1825: F_0(\bs\epsilon) \right) +B F_0^2(\bs\epsilon)-C=0,
1826: \end{equation}
1827: %
1828: where the numbers $B$, $C$ are defined in eqn.~\eqref{form:lett}, and the 
1829: formal power series $h_i^{(j)}(\bs\epsilon)$ are given by
1830: %
1831: \begin{equation}\label{form:hi}
1832: \begin{split}
1833: &h_i^{(j)}(\bs\epsilon)=\frac{\partial v_i^{(j)}}{\partial u_{i+1}}({\bs u}_c)
1834: \epsilon_{i+1}-\frac{i+2}{2}\frac{\partial v_0^{(j)}}{\partial u_{1}}({\bs u}_c)
1835: \epsilon_1\epsilon_i \qquad (i=1,\ldots, M-1),\\
1836: &h_M^{(j)}(\bs\epsilon)=-\frac{M+2}{2}\frac{\partial v_0^{(j)}}
1837: {\partial u_{1}}({\bs u}_c)\epsilon_1\epsilon_M.
1838: \end{split}
1839: \end{equation}
1840: %
1841: \end{prop}
1842: 
1843: \begin{proof}
1844: Using the expansions in the proof of Proposition \ref{prop:newfeqn}, it 
1845: is readily inferred that the power series $s_{{\bs v}}$ and $\bs
1846: \epsilon_{{\bs v}}$ are, to leading orders in $s$, given by
1847: %
1848: \begin{equation*}
1849: \begin{split}
1850: &s_{{\bs v}^{(j)}}= s+ \frac{1}{2}\frac{\partial v_0^{(j)}}
1851: {\partial u_{1}}({\bs u}_c)\epsilon_1 s^2 +{\cal O}(s^3),\\
1852: &\epsilon_{i,{\bs v}^{(j)}}= \epsilon_i+h_i^{(j)}(\bs\epsilon)s
1853: +{\cal O}(s^2) \qquad (i=1,\ldots, M-1),
1854: \end{split}
1855: \end{equation*}
1856: %
1857: with $h_i^{(j)}(\bs\epsilon)$ as defined in eqn.~\eqref{form:hi}.
1858: Using this result, we compute the expansion of $H^{(j)}({\bs u}(s,\bs\epsilon))$
1859: up to order $s^2$. This yields
1860: %
1861: \begin{equation*}
1862: \begin{split}
1863: H^{(j)}(&{\bs u}(s,\bs\epsilon))  = G({\bs u_c})+
1864: s_{{\bs v}^{(j)}}F(s_{{\bs v}^{(j)}},\bs
1865: \epsilon_{{\bs v}^{(j)}})\\
1866: &= G({\bs u_c}) + F_0(\bs\epsilon)s +
1867: \left( F_1(\bs\epsilon)+\frac{1}{2}\frac{\partial v_0^{(j)}}
1868: {\partial u_{1}}({\bs u}_c)\epsilon_1 F_0(\bs\epsilon)
1869: + \sum_{i=1}^M h_i^{(j)}(\bs\epsilon) \frac{\partial}{\partial\epsilon_i}
1870: F_0(\bs\epsilon)\right)s^2 +{\cal O}(s^3).
1871: \end{split}
1872: \end{equation*}
1873: %
1874: Now expand the functional equation \eqref{form:funcind} to leading orders in 
1875: $s$. Terms of order $s^0$ vanish due to eqn.~\eqref{pgfeqn}, evaluated at 
1876: ${\bs u_0}={\bs u_c}$. Terms of order $s^1$ vanish due the 
1877: condition $\sum_{j=1}^{N} \frac{\partial P}{\partial y_j}(\bs u_c, \bs 
1878: G(\bs u_c))=1$. Terms of order $s^2$ lead to the partial differential 
1879: equation given above.
1880: \end{proof}
1881: 
1882: \noindent {\bf Remarks.}
1883: {\it i)} For $M>2$, it is an open question whether closed form solutions 
1884: for $F_0(\bs\epsilon)$ exist. See \cite{NT03b} for a discussion of the cases
1885: $M=1$ and $M=2$.\\
1886: {\it ii)} The above method can also be used to derive 
1887: partial differential equations characterising the generating functions 
1888: $F_l(s,\bs\epsilon)$ of the amplitudes $f_{{\bs k},l}$ 
1889: for $l>0$. These equations arise in the expansion of the $q$-functional 
1890: equation in $s$ at order $l+2$, see \cite{R02} for examples where $M=1$.
1891: 
1892: \medskip
1893: 
1894: The above theorem leads to an alternative derivation of the recursion 
1895: eqn.~\eqref{form:coefrek} of Proposition \ref{prop:rek} in the case 
1896: $\nu_0=0$.
1897: 
1898: \begin{proof}[Alternative proof of eqn.~\eqref{form:coefrek}]
1899: We set $F_0(\bs\epsilon)=K(-\bs\epsilon)+f_{{\bf 0},0}$ 
1900: and rewrite eqn.~\eqref{form:res} in the form
1901: %
1902: \begin{equation}\label{form:ee1}
1903: K(\bs\epsilon)=\sum_{i=1}^M \frac{i+2}{2}\mu_0\epsilon_1\epsilon_i
1904: \frac{\partial K(\bs\epsilon)}{\partial \epsilon_i}+\sum_{i=1}^{M-1} 
1905: \mu_i\epsilon_{i+1}\frac{\partial K(\bs\epsilon)}{\partial \epsilon_i}
1906: -\frac{\mu_0}{2}\epsilon_1(K(\bs\epsilon)+f_{{\bf 0},0})-\frac{1}
1907: {2f_{{\bf 0},0}}K(\bs\epsilon)^2,
1908: \end{equation}
1909: %
1910: where $K({\bf 0})=0$, and the constants $\mu_i$ are, for $i\in\{0,\ldots,M-1\}$, 
1911: given by $\mu_i=-A_i/(2Bf_{{\bf 0},0})$, with $A_i$ defined in eqn.~\eqref{eqn:Ai}.
1912: This leads to the recursion eqn.~\eqref{form:coefrek} for the coeffients 
1913: $f_{\bs\nu}$ in Proposition \ref{prop:rek}, if we set $\bs
1914: \nu=(0,{\bs k})$.
1915: \end{proof}
1916: 
1917: 
1918: \section{Growth of amplitudes}\label{sec:growth}
1919: 
1920: We are interested in the growth of the coefficents $f_{\bs\nu,0}$,
1921: which appear in Proposition \ref{prop:rek} and in Proposition \ref{theo:pde}, 
1922: in the case $\nu_0=0$. To this end, we study properties of the partial 
1923: differential equation of the associated generating function $F_0(\bs
1924: \epsilon)$ of eqn.~\eqref{form:Fl}, given by
1925: %
1926: \begin{equation*}
1927: F_0(\bs\epsilon)=\sum_{\bs k\ge\bf0} (-1)^{\bs k}
1928: f_{{\bs k},0}{\bs\epsilon}^{\bs k}.
1929: \end{equation*}
1930: 
1931: \begin{prop}\label{theo:gr}
1932: For a $q$-functional equation, let Assumption \ref{ass} be satisfied.
1933: There exist positive real numbers $D, R_1,\ldots,R_M$, such that
1934: %
1935: \begin{equation*}
1936: |f_{{\bs k},0}| \le D (k_1+\ldots+k_M)! \, (R_1)^{k_1}\cdot\ldots
1937: \cdot(R_M)^{k_M}
1938: \end{equation*}
1939: %
1940: for all $\bs k\ge\bf0$.
1941: \end{prop}
1942: 
1943: For the proof of the proposition, we apply the technique of majorising series.
1944: A formal power series $g=\sum_{\bs k} g_{\bs k} 
1945: {\bs x}^{\bs k}$ majorises a formal power series $h=
1946: \sum_{\bs k} h_{\bs k} {\bs x}^{\bs k}$ if 
1947: $|g_{\bs k}|\le |{\bs k}|!h_{\bs k}$ for all 
1948: ${\bs k}\ge{\bf 0}$. We then write $g\ll h$. We have the following relations.
1949: 
1950: \begin{lemma}\label{lem:maj}
1951: Let $g\ll h$. Then 
1952: \begin{equation}
1953: g^2\ll h^2, \qquad x_1g\ll x_1h, \qquad
1954: x_1x_i\frac{\partial g}{\partial x_i}\ll x_1 h, \qquad
1955: x_{j+1}\frac{\partial g}{\partial x_j}\ll x_{j+1}\frac{\partial h}{\partial x_j},
1956: \end{equation}
1957: for $i=1,\ldots,M$ and $j=1,\ldots, M-1$.
1958: \end{lemma}
1959: 
1960: \begin{proof}[Sketch of proof]
1961: These relations are checked by direct computation. If $i=2,\ldots,M$, we have
1962: $x_1x_i\frac{\partial g}{\partial x_i}=\sum k_ig_{{\bs k}-
1963: {\bs e}_1}{\bs x}^{\bs k}$. Thus $|k_ig_{{\bs k}-
1964: {\bs e}_1}|\le (|{\bs k}|-1)! k_ih_{{\bs k}-
1965: {\bs e}_1}\le |{\bs k}|!h_{{\bs k}-{\bs e}_1}$, 
1966: and the statement follows. The remaining assertations are proved similarly.
1967: \end{proof}
1968: 
1969: \begin{proof}[Proof of Proposition \ref{theo:gr}]
1970: Starting with $K(\bs\epsilon)$, defined in eqn.~\eqref{form:ee1}, 
1971: we introduce the majorant equation
1972: %
1973: \begin{equation}\label{form:ee2}
1974: L(\bs\epsilon)=\mu_0\left(\sum_{i=1}^M \frac{i+2}{2}\right) \epsilon_1 
1975: L(\bs\epsilon)
1976: +\frac{\mu_0}{2}\epsilon_1(L(\bs\epsilon)+|f_{{\bf 0},0}|)
1977: +\frac{1}{2|f_{{\bf 0},0}|}
1978: L(\bs\epsilon)^2
1979: + \sum_{i=1}^{M-1} \mu_i\epsilon_{i+1}
1980: \frac{\partial L(\bs\epsilon)}{\partial \epsilon_i}.
1981: \end{equation}
1982: %
1983: The last equation uniquely defines a power series with non-negative 
1984: coefficients and positive radius of convergence, satisfying $L({\bf 0})=0$. 
1985: It belongs to a class of singular partial differential equations with regular 
1986: solutions, which is discussed in 
1987: \cite[Thm.~2.8.2.1 and Sec.~2.9.5]{GT96}. To prove that 
1988: $K(\bs\epsilon)\ll L(\bs\epsilon)$, denote the 
1989: r.h.s. of eqn.~\eqref{form:ee1} by $RK$ and the r.h.s.~of eqn.~\eqref{form:ee2} by 
1990: $\widehat RL$. By construction, $g\ll h$ implies $Rg \ll \widehat R h$. We use 
1991: an iteration argument. Set $K_0=L_0=0$. Clearly $K_0\ll L_0$. Define 
1992: $K_n=RK_{n-1}$ and $L_n=\widehat R L_{n-1}$ for $n\in\mathbb N$. We have 
1993: $K_n\ll L_n$ for $n\in\mathbb N_0$, due to Lemma \ref{lem:maj}. Thus 
1994: $K(\bs\epsilon)\ll L(\bs\epsilon)$ for the 
1995: formal solutions $K(\bs\epsilon)$ of eqn.~\eqref{form:ee1} and 
1996: $L(\bs\epsilon)$ of eqn.~\eqref{form:ee2}. Since 
1997: $L(\bs\epsilon)$ is regular, we have the estimate
1998: %
1999: \begin{equation*}
2000: |f_{{\bs k},0}| \le |{\bs k}|! [\bs\epsilon^{\bs k}]
2001: L(\bs\epsilon)\le D |{\bs k}|! {\bs R}^{\bs k}
2002: \end{equation*}
2003: %
2004: for some real constants $D>0$ and $R_i>0$, where $i=1,\ldots,M$.
2005: \end{proof}
2006: 
2007: 
2008: \section{Moments and limit distributions}\label{sec:mom}
2009: 
2010: In the following, we give a probabilistic interpretation of the obtained results.
2011: This generalises the discussion of Dyck paths in the introduction, before 
2012: Proposition~\ref{prop:dyckmom}, to the case of a general $q$-functional equation,
2013: and will prove part \textit{i)} of Theorem~\ref{theo:probdist}. For technical reasons 
2014: (Lemma~\ref{lem:Levy}), we will not use random variables below, but rather argue
2015: with the associated probability measures.
2016: 
2017: \smallskip
2018: 
2019: For a $q$-functional equation, let Assumption \ref{ass} be satisfied. Then, 
2020: the coefficients $p_{\bs n}$ in a solution $G({\bs u})= \sum_{\bs 
2021: n\ge 0} p_{\bs n}{\bs u^n}$, such that $G(\bs0)=0$, are non-negative. Assume that the 
2022: numbers $A_i$ of Proposition \ref{prop:rek} satisfy $A_i>0$ for $i=0,\ldots, M-1$. 
2023: We then have $f_{{\bs k},0}>0$ for all ${\bs k}\ge {\bf 0}$ and 
2024: ${\bs k}\neq {\bf 0}$. This implies that $\sum_{n_1,\ldots,n_M}
2025: p_{n_0,n_1,\ldots,n_M}>0$ for almost all $n_0$, see eqn.~\eqref{form:coefas} below. 
2026: Fix $N_0\in\mathbb N$ such that strict positivity holds for all $n_0\ge N_0$. 
2027: For $n_0\ge N_0$, we define discrete Borel probability measures $\widetilde\mu_{n_0}$ 
2028: by
2029: %
2030: \begin{equation*}
2031: \widetilde\mu_{n_0}=\sum_{n_1,\ldots,n_M} \frac{p_{n_0,n_1,\ldots,n_M}}{\sum_{m_1,
2032: \ldots,m_M} p_{n_0,m_1,\ldots,m_M}} \delta_{(n_1,\ldots,n_M)} ,
2033: \end{equation*}
2034: %
2035: compare eqn.~\eqref{eq:pxt} in the introduction.
2036: Their corresponding moments are, for ${\bs k}\in\mathbb N_0^M$, given by
2037: %
2038: \begin{equation}\label{form:tilmom}
2039: \widetilde m_{\bs k}(n_0) = \frac{\sum_{n_1,\ldots,n_M} n_1^{k_1}\cdot\ldots
2040: \cdot n_M^{k_M} p_{n_0,n_1,\ldots,n_M}}{\sum_{n_1,\ldots,n_M} p_{n_0,n_1,\ldots,n_M}}.
2041: \end{equation}
2042: 
2043: We are interested in the asymptotic behaviour of the moments $\widetilde 
2044: m_{\bs k}(n_0)$, as $n_0$ tends to infinity. This will be obtained by an 
2045: analysis of the coefficients of the factorial moment generating functions, 
2046: which relies on a transfer lemma \cite[Thm.~1]{FO90}.
2047: 
2048: \begin{lemma}[Transfer lemma \cite{FO90}]
2049: Suppose that $F(z)=\sum_{n\ge0}f_nz^n$ has a singularity at $z=z_c$ and is
2050: $\Delta$-regular, i.e., it is analytic in the domain
2051: %
2052: \begin{equation*}
2053: \Delta=\{z:|z|\le z_c+\eta, |{\rm arg}(z-z_c)|\ge\phi\}
2054: \end{equation*}
2055: %
2056: for some $\eta>0$ and $0<\phi<\pi/2$. Assume that, as $z\to z_c$ in $\Delta$,
2057: %
2058: \begin{equation*}
2059: F(z) = {\cal O} \left( (z_c-z)^\alpha \right)
2060: \end{equation*}
2061: %
2062: for some real $\alpha$. Then, the $n$-th Taylor coefficient $f_n$ of $F(z)$ satisfies
2063: %
2064: \begin{equation*}
2065: f_n={\cal O}(z_c^{-n} n^{-1-\alpha}) \qquad (n\to\infty).
2066: \end{equation*}
2067: \qed
2068: \end{lemma}
2069: 
2070: The following lemma characterises the asymptotic behaviour
2071: of the moments $\widetilde m_{\bs k}(n_0)$, as $n_0$ tends to infinity.
2072: 
2073: \begin{lemma}\label{lem:pos}
2074: Let Assumption \ref{ass} be satisfied. Assume that the 
2075: numbers $A_i$ of Proposition \ref{prop:rek} satisfy $A_i>0$ for 
2076: $i=0,\ldots, M-1$. Then the  moments $\widetilde m_{\bs k}(n_0)$ 
2077: eqn.~\eqref{form:tilmom} are well-defined for almost all $n_0$. They  
2078: are for ${\bs k}\in\mathbb N_0^M$ asymptotically given by
2079: %
2080: \begin{equation*}
2081: \widetilde m_{\bs k}(n_0)= \frac{{\bs k}!}{f_{{\bf 0},0}
2082: u_c^{\gamma_{\bs k}-\gamma_{\bf 0}}}
2083: \frac{\Gamma(\gamma_{\bf 0})}{\Gamma(\gamma_{\bs k})}f_{{\bs k},0}
2084: n_0^{\gamma_{\bs k}-\gamma_{\bf 0}}+{\cal O}(n_0^{\gamma_{\bs k}-
2085: \gamma_{\bf 0}-1/2}) \qquad (n_0\to\infty),
2086: \end{equation*}
2087: %
2088: where $\Gamma(z)$ denotes the Gamma function, and where the numbers 
2089: $f_{{\bs k},0}$ and $\gamma_{\bs k}$ are defined in 
2090: Proposition \ref{prop:rek}.
2091: \end{lemma}
2092: 
2093: \begin{proof}
2094: The functions $g_{\bs k}(u_0)$ are $\Delta$-regular due to 
2095: Proposition \ref{prop:pui}. We infer from eqn.~\eqref{form:genex} that 
2096: $g_{\bs k}(u_0)=f_{{\bs k},0}(u_c-u_0)^{-\gamma_{\bs k}}+
2097: {\cal O}((u_c-u_0)^{-\gamma_{\bs k}+1/2})$ as $u_0\to u_c^-$, where 
2098: $\gamma_{\bf 0}=-1/2$ and $\gamma_{\bs k}>0$ otherwise. The 
2099: coefficient asymptotics of $(u_c-u_0)^{-\gamma_{\bs k}}$ and 
2100: the transfer lemma yield
2101: %
2102: \begin{equation}\label{form:coefas}
2103: [u_0^{n_0}] g_{\bs k}(u_0)=\frac{f_{{\bs k},0}}
2104: {u_c^{\gamma_{\bs k}}\Gamma(\gamma_{\bs k})}u_c^{-n_0}
2105: n_0^{\gamma_{\bs k}-1}\left(1+{\cal O}(n_0^{-1/2})\right) \qquad
2106: (n_0\to\infty),
2107: \end{equation}
2108: %
2109: where $[x^n]f(x)$ denotes the coefficient of $x^n$ in the power 
2110: series $f(x)$. The error term implies that asymptotically 
2111: factorial moments coincide with ordinary moments. The numbers 
2112: $\widetilde m_{\bs k}(n_0)$ are thus well-defined for $n_0$ 
2113: large enough, and asymptotically given by
2114: %
2115: \begin{equation*}
2116: \begin{split}
2117: \widetilde m_{\bs k}(n_0)&=
2118: \frac{\sum_{{\bs n}_+\ge0} {\bs n}_+^{{\bs k}} 
2119: p_{n_0,{\bs n}_+}}{\sum_{{\bs n}_+\ge0} 
2120: p_{n_0,{\bs n}_+}}=\frac{[u_0^{n_0}]g_{\bs k}(u_0)}
2121: {[u_0^{n_0}]g_{\bf 0}(u_0)}{\bs k}!\left( 1+{\cal O}(n_0^{-1/2}) 
2122: \right)\\
2123: &= \frac{{\bs k}!}{f_{{\bf 0},0}u_c^{\gamma_{\bs k}-
2124: \gamma_{\bf 0}}}\frac{\Gamma(\gamma_{\bf 0})}{\Gamma(\gamma_{\bs k})}
2125: f_{{\bs k},0}n_0^{\gamma_{\bs k}-\gamma_{\bf 0}}\left( 1+
2126: {\cal O}(n_0^{-1/2}) \right) \qquad (n_0\to\infty).
2127: \end{split}
2128: \end{equation*}
2129: %
2130: This concludes the proof of the lemma.
2131: \end{proof}
2132: %
2133: The moments $\widetilde m_{\bs k}(n_0)$ diverge as $n_0\to\infty$,
2134: as may be inferred from Lemma \ref{lem:pos}. Introduce normalised Borel
2135: probability measures $\mu_{n_0}$ by
2136: %
2137: \begin{equation}\label{form:normrv}
2138: \mu_{n_0}=\sum_{n_1,\ldots,n_M} \frac{p_{n_0,n_1,\ldots,n_M}}{\sum_{m_1,\ldots,m_M} 
2139: p_{n_0,m_1,\ldots,m_M}} \delta_{(\widetilde n_1,\ldots,\widetilde n_M)},
2140: \end{equation}
2141: %
2142: where $\widetilde n_k=n_kn_0^{-(k+2)/2}$ for $k=1,\ldots,M$. For ${\bs k}
2143: \in\mathbb N_0^M$, denote the corresponding moments by $m_{\bs k}(n_0)$.
2144: We now show that the limits
2145: %
2146: \begin{equation}\label{form:normmom}
2147: m_{\bs k}=\lim_{n_0\to\infty} m_{\bs k}(n_0)=\lim_{n_0\to\infty} 
2148: \frac{\widetilde m_{\bs k}(n_0)}{n_0^{\gamma_{\bs k}-\gamma_{\bf 0}}}
2149: = \frac{{\bs k}!}{f_{{\bf 0},0}u_c^{\gamma_{\bs k}-\gamma_{\bf 0}}}
2150: \frac{\Gamma(\gamma_{\bf 0})}{\Gamma(\gamma_{\bs k})}f_{{\bs k},0}
2151: \end{equation}
2152: %
2153: exist and define a unique Borel probability measure $\mu$, with finite moments 
2154: $m_{\bs k}>0$ at all orders. This will be achieved using L\'evy's 
2155: continuity theorem. We first prove the following lemma.
2156: 
2157: \begin{lemma}\label{lem:momcon}
2158: Let Assumption \ref{ass} be satisfied. Assume that the 
2159: numbers $A_i$ of Proposition \ref{prop:rek} satisfy $A_i>0$ for $i=0,\ldots, M-1$. 
2160: Consider for ${\bs k}\in \mathbb N_0^M$
2161: the numbers $m_{\bs k}\ge0$, defined in eqn.~\eqref{form:normmom} and
2162: eqn.~\eqref{form:tilmom}.
2163: For ${\bs t}\in \mathbb R^M$, we have
2164: %
2165: \begin{equation*}
2166: \lim_{|{\bs k}|\to\infty} \frac{m_{\bs k}
2167: {\bs t}^{\bs k}}{{\bs k}!}=0.
2168: \end{equation*}
2169: %
2170: Equivalently, the exponential generating function of the
2171: numbers $m_{\bs k}$ is entire.
2172: \end{lemma}
2173: 
2174: \begin{proof}
2175: The limit $m_{\bs k}$ eqn.~\eqref{form:normmom} exists for ${\bs k}\in 
2176: \mathbb N_0^M$, due to Lemma~\ref{lem:pos}. Note that
2177: %
2178: \begin{equation*}
2179: \gamma_{\bs k}=-\frac{1}{2}+\sum_{i=1}^M \left(1+\frac{i}{2}\right)k_i\ge
2180: -\frac{1}{2}+\frac{3}{2}|{\bs k}|\ge |{\bs k}|+1+ \lfloor \frac{|{\bs k}|-3}{2}\rfloor.
2181: \end{equation*}
2182: %
2183: Thus $|{\bs k}|\to\infty$ implies $\gamma_{\bs k}\to\infty$. Furthermore, since
2184: $e(n/e)^n\le n!\le e n (n/e)^n$ for $n\in\mathbb N$, we have the estimate
2185: %
2186: \begin{equation*}
2187: \frac{n!}{(n+n_0)!}\le \frac{n (n/e)^n}{((n+n_0)/e)^{n+n_0}}\le \frac{e^{n_0}}{n^{n_0-1}}
2188: \end{equation*}
2189: %
2190: for $n,n_0\in\mathbb N$. Now fix ${\bs t}\in \mathbb R^M$. We estimate
2191: %
2192: \begin{equation*}
2193: \begin{split}
2194: \left|\frac{m_{\bs k}{\bs t}^{\bs k}}{{\bs k}!} \right|&=
2195: \frac{\Gamma(\gamma_0)}{\Gamma(\gamma_{\bs k})}
2196: \frac{|f_{{\bs k},0}|}{f_{{\bf 0},0}u_c^{\gamma_{\bs k}-\gamma_{\bf 0}}} |{\bs t}^{\bs k}|
2197: \le
2198: \frac{\Gamma(\gamma_0)}{(|{\bs k}|+\lfloor \frac{|{\bs k}|-3}{2}\rfloor)!}
2199: \frac{D |{\bs k}|! {\bs R}^{\bs k}}{f_{{\bf 0},0}} |{\bs t}^{\bs k}|
2200: \max (u_c,u_c^{-1})^{\gamma_{\bs k}-\gamma_{\bf 0}}\\
2201: &\le \frac{D\Gamma(\gamma_0)}{f_{{\bf 0},0}} e^{|\bs k|}{\bs R}^{\bs k}
2202: |{\bs t}^{\bs k}|\frac{1}{|{\bs k}|^{\lfloor \frac{|{\bs k}|-5}{2}\rfloor}}
2203: \max (u_c,u_c^{-1})^{\frac{M+2}{2}|{\bs k}|}.
2204: \end{split}
2205: \end{equation*}
2206: %
2207: The r.h.s. vanishes as $|{\bs k}|\to\infty$, which implies the assertion.
2208: The equivalent statement is obvious.
2209: \end{proof}
2210: 
2211: Our proof of claim \textit{i)} in Theorem~\ref{theo:probdist} relies
2212: on an application of L\'evy's continuity theorem \cite[Thm.~23.8]{B96},
2213: which we cite in the following lemma.
2214: 
2215: \begin{lemma}[L\'evy's continuity theorem \cite{B96}]\label{lem:Levy}
2216: For $n\in\mathbb N$, let probability measures $\mu_n$ on the Borel $\sigma$-algebra 
2217: of $\mathbb R^M$ be given. If the sequence $\{\widehat \mu_n\}_{n\in\mathbb N}$ of 
2218: their characteristic functions $\widehat \mu_n$ converges pointwise to a complex 
2219: function $\phi$ which is continuous at the origin, then $\phi$ is the characteristic 
2220: function of a uniquely determined Borel probability measure $\mu$. Moreover, 
2221: $\{\mu_n\}_{n\in\mathbb N}$ converges to $\mu$ weakly. \qed 
2222: \end{lemma}
2223: 
2224: \begin{proof}[Proof of claim \textit{i)} in Theorem~\ref{theo:probdist}.]
2225: Lemma \ref{lem:pos} implies that the Borel probability measures
2226: $\mu_{n_0}$ eqn.~\eqref{form:normrv} are well-defined for almost all 
2227: $n_0\in\mathbb N$. The estimate in Lemma \ref{lem:momcon} implies uniform 
2228: convergence of the sequence of Fourier transforms $\widehat \mu_{n_0}:\mathbb 
2229: R^M\to\mathbb C$ of $\mu_{n_0}$, in every ball of finite radius centred 
2230: at the origin. In particular, we have pointwise convergence of the sequence 
2231: $\{\widehat \mu_{n_0}\}_{n_0\in\mathbb N}$. For $M=1$, 
2232: the corresponding argument is given in the proof of \cite[Thm~6.4.5]{C74}. 
2233: It can be directly extended to arbitrary $M$. Since the functions 
2234: $\widehat \mu_{n_0}$ are continuous, we conclude that the limit function 
2235: $\phi:\mathbb R^M\to\mathbb C$ is continuous at the origin.
2236: Now apply L\'evy's continuity theorem. The limit probability measure $\mu$ 
2237: has moments $m_{\bs k}$ eqn.~\eqref{form:normmom}. The claimed statement 
2238: of the theorem follows, when phrasing the result in terms of the associated 
2239: random variables.
2240: \end{proof}
2241: 
2242: The connection to Brownian motion, which is claimed in part \textit{ii)} of 
2243: Theorem~\ref{theo:probdist}, will be established in the following section.
2244: 
2245: \section{Dyck paths revisited}\label{sec:Dyckrev}
2246: 
2247: We apply our results to the example of Dyck paths of Section \ref{sec:Dyck}. 
2248: The power series solution $E({\bs u_0})$ of eqn.~\eqref{form:exfunc}, specialised to 
2249: $\bs u=\bs u_0$, has radius of convergence $u_c=1/4$, with a square root 
2250: singularity at $u=u_c$, and $E({\bs u_c})=1$. The $q$-functional equation 
2251: eqn.~\eqref{form:exfunc} satisfies Assumption \ref{ass}. Since the random variables 
2252: $(X_{1,n_0},\ldots,X_{M,n_0})$ of eqn.~\eqref{form:DyckRV} have the distribution of 
2253: $\mu_{n_0}$, as defined in eqn.~\eqref{form:normrv}, Proposition 
2254: \ref{prop:dyckmom} follows from the results of the previous section.
2255: 
2256: \begin{proof}[Proof of Proposition \ref{prop:dyckmom}]
2257: The generating function $E({\bs u})$ of Dyck paths satisfies the 
2258: $q$-functional equation eqn.~\eqref{form:exfunc}. Assumption \ref{ass} holds with 
2259: $u_c=1/4$ and $y_c=1$. We have $f_{{\bf 0},0}=-4$, $\gamma_{\bf 0}=-1/2$, 
2260: $\mu_i=(i+1)/4$ for $i=1,\ldots,M-1$ and $\mu_0=1/8$. Thus, Proposition 
2261: \ref{prop:rek} yields eqn.~\eqref{form:exc2}. The distribution of the 
2262: random variables $(X_{1,n_0},\ldots,X_{M,n_0})$ of eqn.~\eqref{form:DyckRV} is that of 
2263: the probability measures $\mu_{n_0}$ in eqn.~\eqref{form:normrv}. By 
2264: part \textit{i)} of Theorem \ref{theo:probdist}, there exists a unique limit probability 
2265: measure $\mu$. We thus get eqn.~\eqref{eqn:dyckres} from 
2266: eqn.~\eqref{form:normmom}.
2267: \end{proof}
2268: 
2269: The connection between Dyck paths and Brownian excursions in 
2270: Proposition~\ref{prop:reldyckexc} leads, for a general 
2271: $q$-functional equation, to an explicit characterisation of the 
2272: limit probability measure $\mu$, resp.~of the associated limit random variable 
2273: $(Y_1,\ldots,Y_M)$.
2274: 
2275: \begin{proof}[Proof of claim \textit{ii)} in Theorem~\ref{theo:probdist}.]
2276: For the given $q$-functional equation, let $F_0(\bs\epsilon)$ be the 
2277: generating function of the leading amplitudes $f_{\bs k,0}$. For $k=1,\ldots,M$ 
2278: and $d_k\in\mathbb R$, define 
2279: $G_0(\bs\epsilon)=F_0(\epsilon_1 d_1,\ldots, \epsilon_M d_M)$.
2280: An explicit calculation using eqn.~\eqref{form:ee1} shows that the power series 
2281: $G_0(\bs\epsilon)$ satisfies the same type of differential equation 
2282: as $F_0(\bs\epsilon)$ does, with $\mu_0$ replaced by $\mu_0 d_1$ and 
2283: $\mu_i$ replaced by $\mu_i d_{i+1}/d_i$, where $i=1,\ldots,M-1$. Now choose 
2284: the values $d_k$, such that the equation for $G_0(\bs\epsilon)$ is 
2285: that of Dyck paths. Noting the relation between Dyck paths and Brownian 
2286: excursions eqn.~\eqref{form:DyEx}, we arrive at the values $c_k=2^{(k+2)/2}
2287: /d_k$, for numbers $c_k$ as in the claim of Theorem~\ref{theo:probdist}.
2288: \end{proof}
2289: 
2290: \section{Concluding remarks}\label{sec:con}
2291: 
2292: We finally stress three central aspects of our approach.
2293: Firstly, the approach yields a \emph{universal} limit distribution -- loosely
2294: spoken, it appears for all models, whose underlying functional equation 
2295: has the same singularity structure. Such a result may be compared to a 
2296: central limit theorem in probability 
2297: theory. For example, models other than Dyck paths, which display a square root 
2298: as dominant singularity of their size generating function, are certain models of 
2299: trees or polygons. For simply generated trees, $q$-functional equations appear when counting 
2300: by number of vertices and by internal path length \cite{T91}. Using the 
2301: above setup, moment recursions for the parameter ``sum of $k$-th 
2302: powers of the vertex distances to the root'' are obtained. For polygon 
2303: models, $q$-functional equations appear when counting by perimeter 
2304: and area \cite{D99,R02}, which is, for column-convex polygons, the 
2305: sum of the column heights. The above setup gives moment recursions
2306: for the parameter ``sums of $k$-th powers of the column heights'', in the 
2307: limit of infinite (horizontal) perimeter.
2308: 
2309: Secondly, our approach is \emph{algorithmic} -- the moment recursion can
2310: finally be deduced from a straightforward calculation, by the method of
2311: dominant balance. Our approach also allows to obtain corrections to the 
2312: asymptotic behaviour, which cannot (easily) be deduced by other methods.
2313: 
2314: Thirdly, the approach is \emph{flexible} -- it may be applied to other classes 
2315: to obtain a universal limit distribution, which only depends on the singularity 
2316: structure of the functional equation. Our generating function approach is particularly
2317: suited for counting parameters, which decompose linearly under the 
2318: cartesian product construction. Examples of such models with a 
2319: rational generating function appear in \cite{R02}. Examples with an inverse 
2320: square root appear in \cite{NT03}. In particular, the discrete counterparts of 
2321: Brownian motion, Brownian bridges, and Brownian meanders can be studied, 
2322: see \cite{NT03,R06}. Also, universality questions for parameters related 
2323: to left and right path lengths in trees \cite{J06,KS06,P06,Bou06} can be studied 
2324: by our methods, compare the discussion in \cite{BJ06}.
2325: 
2326: Within the framework of simply generated trees, an alternative derivation of 
2327: the above moment recursion could be obtained with the techniques of 
2328: \cite{J03}, where the different problem of the (generalised) Wiener index of 
2329: trees was analysed. It would be interesting to 
2330: consider how our methods can be adapted to this problem.
2331: 
2332: \section*{Acknowledgements}
2333: 
2334: The author thanks Philippe Duchon, Philippe Flajolet and Michel 
2335: Nguy$\tilde{\mbox{\rm \^e}}$n Th$\acute{\mbox{\rm \^e}}$ for helpful 
2336: discussions, and Svante Janson for comments on the manuscript. 
2337: He thanks the department LaBRI (Bordeaux) for hospitality 
2338: in autumn 2003, where parts of the problem have been analysed. The
2339: author thanks the referees for suggestions, which improved the presentation 
2340: of the article. Financial support by the German Research Council (DFG) 
2341: is gratefully acknowledged.
2342: 
2343: \begin{thebibliography}{99}
2344: 
2345: \bibitem{A91}
2346: D.J.~Aldous, 
2347: The continuum random tree I,
2348: \textit{Ann.~Prob.~\bf 19} (1991), 1--28.
2349: 
2350: \bibitem{A92}
2351: D.J.~Aldous, 
2352: The continuum random tree II: an overview,
2353: {\it Stochastic Analysis}, eds. M.T.~Barlow and N.H.~Bingham,
2354: Cambridge University Press (1991), 23--70.
2355: 
2356: \bibitem{A93}
2357: D.J.~Aldous, 
2358: The continuum random tree III,
2359: \textit{Ann.~Prob.~\bf 21} (1993), 248--289.
2360: 
2361: \bibitem{BF02}
2362: C.~Banderier and P.~Flajolet,
2363: Basic analytic combinatorics of directed lattice paths,
2364: \textit{Theoret.~Comput.~Sci. \bf 281} (2002), 37--80.
2365: 
2366: \bibitem{B96}
2367: H.~Bauer,
2368: \textit{Probability Theory},
2369: de Gruyter Studies in Mathematics, vol.~23,
2370: de Gruyter, Berlin (1996).
2371: 
2372: \bibitem{Bel72}
2373: B.~Belkin,
2374: An invariance principle for conditioned recurrent random walk
2375: attracted to a stable law,
2376: \textit{Z.~Wahrscheinlichkeitstheorie und Verw.~Gebiete \bf 21} (1972), 45--64.
2377: 
2378: \bibitem{Bill95}
2379: P.~Billingsley, 
2380: \textit{Probability and Measure}, 
2381: 3rd ed., Wiley, New York (1995).
2382: 
2383: \bibitem{B99}
2384: P.~Billingsley, 
2385: \textit{Convergence of Probability Measures},
2386: 2nd ed., Wiley, New York (1999).
2387: 
2388: \bibitem{Bou96}
2389: M.~Bousquet-M\'elou,
2390: A method for the enumeration of various classes of column-convex polygons,
2391: \textit{Discr.~Math. \bf 154} (1996), 1--25.
2392: 
2393: \bibitem{Bou06}
2394: M.~Bousquet-M\'elou, 
2395: Limit laws for embedded trees: applications to the integrated 
2396: superBrownian excursion,
2397: \textit{Random Structures Algorithms \bf 29} (2006), 475--523.
2398: 
2399: \bibitem{BJ06} 
2400: M.~Bousquet-M\'elou and S.~Janson,
2401: The density of the ISE and local limit laws for embedded trees,
2402: \textit{Ann.~Appl.~Probab. \bf 16} (2006), 1597--1632. 
2403: 
2404: \bibitem{BFG03}
2405: J.~Bouttier, P.~Di Francesco and E.~Guitter,
2406: Geodesic distance in planar graphs,
2407: \textit{Nuclear Phys. B \bf 663} (2003), 535--567. 
2408: 
2409: \bibitem{BKR72}
2410: N.G.~de Bruijn, D.E.~Knuth, and S.O.~Rice, 
2411: The average height of planted plane trees, in:
2412: \textit{Graph Theory and Computing},  
2413: pp. 15--22, Academic Press, New York (1972).
2414: 
2415: \bibitem{C74}
2416: K.L.~Chung,
2417: A Course in Probability Theory,
2418: Academic Press, New York (1974).
2419: 
2420: \bibitem{CS96}
2421: G.M.~Constantine and T.H.~Savits,
2422: A multivariate Faa di Bruno formula with applications,
2423: \textit{Trans.~Amer.~Math.~Soc. \bf 348} (1996), 503--520. 
2424: 
2425: \bibitem{D97}
2426: M.~Drmota, 
2427: Systems of functional equations,  
2428: \textit{Random Struct.~Algorithms  \bf 10}  
2429: (1997), 103--124.
2430: 
2431: \bibitem{D03}
2432: M.~Drmota,
2433: Stochastic analysis of tree-like data structures,
2434: \textit{Proc.~Royal~Soc.~Lond.~A \bf 460} (2004), 271--307.
2435: 
2436: \bibitem{D98}
2437: P.~Duchon,
2438: $Q$-grammaires: un outil pour l'\`enumeration,
2439: PhD thesis, Bordeaux University (1998).
2440: 
2441: \bibitem{D99}
2442: P.~Duchon,
2443: $Q$-grammars and wall polyominoes,
2444: {\it Ann.~Comb. \bf 3} (1999), 311--321.
2445: 
2446: \bibitem{FFK04}
2447: J.A.~Fill, P.~Flajolet and N.~Kapur,
2448: Singularity analysis, Hadamard products, and tree recurrences,
2449: \textit{J.~Comput.~Appl.~Math. \bf 174} (2005), 271--313.
2450: 
2451: \bibitem{FL01}
2452: P.~Flajolet and G.~Louchard,
2453: Analytic variations on the Airy distribution,
2454: \textit{Algorithmica \bf 31} (2001), 361--377.
2455: 
2456: \bibitem{FO90}
2457: P.~Flajolet and A.M.~Odlyzko,
2458: Singularity analysis of generating functions,
2459: \textit{SIAM J.~Discr.~Math.~\bf 3} (1990), 216--240.
2460: 
2461: \bibitem{FPV98}
2462: P.~Flajolet, P.~Poblete and A.~Viola,
2463: On the analysis of linear probing hashing,
2464: \textit{Algorithmica \bf 22} (1998), 37--71.
2465: 
2466: \bibitem{FS07}
2467: P.~Flajolet and R.~Sedgewick,
2468: \textit{Analytic Combinatorics},
2469: book in preparation (2007).
2470: 
2471: \bibitem{GT96}
2472: R.~G\'erard and H.~Tahara,
2473: \textit{Singular Nonlinear Partial Differential Equations}, 
2474: Aspects of Mathematics, vol.~E 28, Vieweg, Braunschweig (1996).
2475: 
2476: \bibitem{J00}
2477: E.J.~Janse van Rensburg,
2478: \textit{The Statistical Mechanics of Interacting Walks, Polygons, Animals and 
2479: Vesicles}, Oxford University Press, Oxford (2000).
2480: 
2481: \bibitem{JLR00}
2482: S.~Janson, T.~Luczak and A.~Rucinski,
2483: \textit{Random graphs},
2484: Wiley, New York (2000).
2485: 
2486: \bibitem{J03}
2487: S.~Janson,
2488: The Wiener index of simply generated random trees,
2489: \textit{Random Struct.~Algorithms \bf 22} (2003), 337--358.
2490: 
2491: \bibitem{J06}
2492: S.~Janson,
2493: Left and right pathlengths in random binary trees,
2494: \textit{Algorithmica \bf 46} (2006), 419--429.
2495: 
2496: \bibitem{J07}
2497: S.~Janson,
2498: Brownian excursion area, Wright's constants in graph enumeration, 
2499: and other Brownian areas,
2500: \textit{Probab.~Surv. \bf 4} (2007), 80--145.
2501: 
2502: \bibitem{KMM07}
2503: M.J.~Kearney, S.N.~Majumdar and R.J.~Martin,
2504: The first-paasage area for drifted Brownian motion and the moments of
2505: the Airy distribution,
2506: \textit{preprint} (2007); \texttt{arXiv:0706.2038}.
2507: 
2508: \bibitem{KS05}
2509: C.~Knessl and W.~Szpankowski,
2510: Enumeration of binary trees and universal types,
2511: \textit{Discrete Math. Theor. Comput. Sci. \bf 7} (2005), 313--400.
2512: 
2513: \bibitem{KS06}
2514: C.~Knessl and W.~Szpankowski, 
2515: On the joint path length distribution in random binary trees,
2516: \textit{Stud. Appl. Math. \bf 117} (2006), 109--147.
2517: 
2518: \bibitem{L84}
2519: G.~Louchard,
2520: Kac's formula, L\'evy's local time and Brownian excursion,
2521: \textit{J.~Appl.~Prob. \bf 21} (1984), 479--499.
2522: 
2523: \bibitem{L85}
2524: G.~Louchard,
2525: The {B}rownian excursion area: a numerical analysis,
2526: \textit{Comput.~Math.~Appl. \bf 10} (1985), 413--417;
2527: Erratum, 
2528: \textit{Comput.~Math.~Appl. \bf 12} (1986), 375.
2529: 
2530: \bibitem{MM03}
2531: J.-F.~Marckert and A.~Mokkadem,
2532: The depth first processes of Galton-Watson trees converge 
2533: to the same Brownian excursion,
2534: \textit{Ann. Probab. \bf 31} (2003), 1655--1678.
2535: 
2536: \bibitem{MM78}
2537: A.~Meir and J.W.~Moon,
2538: On the altitude of nodes in random trees,
2539: \textit{Canad.~J.~Math. \bf 30} (1978), 997--1015.
2540: 
2541: \bibitem{NT03c}
2542: M.~Nguy$\tilde{\mbox{\rm \^e}}$n 
2543: Th$\acute{\mbox{\rm \^e}}$,
2544: Distributions de valuations sur les arbres,
2545: PhD thesis, L'\'Ecole Polytechnique, Paris (2003).
2546: 
2547: \bibitem{NT03}
2548: M.~Nguy$\tilde{\mbox{\rm \^e}}$n 
2549: Th$\acute{\mbox{\rm \^e}}$,
2550: Area of Brownian Motion with Generatingfunctionology, in: 
2551: {\em Discrete Random Walks, DRW'03}, 
2552: eds. C.~Banderier and C.~Krattenthaler, 
2553: \textit{Discr.~Math.~and Theoret.~Comput.~Sci. Proceedings AC} (2003), 
2554: 229--242.
2555: 
2556: \bibitem{NT03b}
2557: M.~Nguy$\tilde{\mbox{\rm \^e}}$n 
2558: Th$\acute{\mbox{\rm \^e}}$,
2559: Area and inertial moment of Dyck paths, 
2560: \textit{Combin.~Probab.~Comput. \bf 13} (2004), 697--716.
2561: 
2562: \bibitem{O95}
2563: A.M.~Odlyzko,
2564: Asymptotic enumeration methods,
2565: in: \textit{Handbook of Combinatorics}, vol. 2, 
2566: eds. R.L.~Graham, M.~Gr\"otschel and L.~Lov\'asz,
2567: Elsevier, Amsterdam (1995), 1063--1229.
2568: 
2569: \bibitem{P06}
2570: A.~Panholzer,
2571: Left and right length of paths in binary trees or
2572: on a question of Knuth, 
2573: \textit{Discr.~Math.~and Theoret.~Comput.~Sci. Proceedings AG} (2006), 
2574: 415--418.
2575: 
2576: \bibitem{P95}
2577: T.~Prellberg,
2578: Uniform $q$-series asymptotics for staircase polygons,
2579: \textit{J.~Phys.~A: Math.~Gen. \bf 28} (1995), 1289--1304.
2580: 
2581: \bibitem{PB95a}
2582: T.~Prellberg and R.~Brak,
2583: Critical exponents from non-linear functional equations for
2584: partially directed cluster models,
2585: \textit{J.~Stat.~Phys. \bf 78} (1995), 701--730.
2586: 
2587: \bibitem{PO95}
2588: T.~Prellberg and A.L.~Owczarek,
2589: Stacking models of vesicles and compact clusters,
2590: \textit{J.~Stat.~Phys. \bf 80} (1995), 755--779.
2591: 
2592: \bibitem{R02}
2593: C.~Richard,
2594: Scaling behaviour of two-dimensional polygon models,
2595: \textit{J.~Stat.~Phys. \bf 108} (2002), 459--493.
2596: 
2597: \bibitem{R04}
2598: C.~Richard,
2599: Limit distributions for models of exactly solvable walks,
2600: in: \textit{Oberwolfach reports \bf 1}, Report No. 22/2004, 
2601: Mathematisches Forschungsinstitut Oberwolfach (2004), 1189--1191.
2602: 
2603: \bibitem{R06}
2604: C.\ Richard,
2605: Staircase polygons: moments of diagonal lengths and column heights,
2606: \textit{J.~Phys.: Conf.~Ser. \bf 42} (2006), 239--257.
2607: 
2608: \bibitem{S99}
2609: R.P.~Stanley,
2610: \textit{Enumerative Combinatorics},
2611: vol.~2, Cambridge University Press, Cambridge (1999).
2612: 
2613: \bibitem{T91}
2614: L.~Tak\'acs,
2615: A Bernoulli excursion and its various applications,
2616: \textit{Adv.~Appl.~Prob. \bf 23} (1991), 557--585.
2617: 
2618: \bibitem{VRSZ03}
2619: L.~Di Vizio, J.-P.~Ramis, J.~Sauloy and C.~Zhang,
2620: \'Equations aux $q$-diff\'erences,
2621: \textit{Gaz.~Math. \bf 96} (2003), 20---49.
2622: 
2623: \end{thebibliography}
2624: 
2625: \end{document}
2626: 
2627: 
2628: 
2629: 
2630: 
2631: 
2632: