1: \documentclass[12pt]{article}
2:
3: \usepackage{a4wide}
4: \usepackage{amssymb,amsmath,amsthm}
5:
6: \usepackage{rotating}
7: \usepackage{graphicx}
8: \usepackage{latexsym}
9: \usepackage{bm}
10: \usepackage{mathrsfs}
11:
12:
13: \newcommand{\AGf}{\ensuremath{\mathcal{A}}}
14: \newcommand{\PGf}{\ensuremath{\mathcal{P}}}
15: \newcommand{\QGf}{\ensuremath{\mathcal{Q}}}
16: \newcommand{\PPR}{\ensuremath{\mathcal{P}_{\rm Rot}}}
17: \newcommand{\PP}[1]{\ensuremath{\mathcal{P}^{(#1)}}}
18: \newcommand{\PPM}[2]{\ensuremath{\mathcal{P}^{(#1)}_{#2}}}
19: \newcommand{\StM}[1]{\ensuremath{\mathcal{P}}^{\,\Box\,(#1)}}
20: \newcommand{\StMR}[1]{\ensuremath{\mathcal{P}}_{\,#1}^{\,\Box,(reg)}}
21: \newcommand{\StMM}[2]{\ensuremath{\mathcal{P}}_{#2}^{\,\Box\,(#1)}}
22: \newcommand{\Pol}{\ensuremath{\mathscr{P}}}
23: \newcommand{\Sign}[2]{\ensuremath{{ \rm Sign}\left ( [x^{#2}] #1 \right )}}
24: \newcommand{\La}[2]{\ensuremath{A^{(#1)}_{#2}}}
25: \newcommand{\Lb}[2]{\ensuremath{B^{(#1)}_{#2}}}
26: \newcommand{\Lc}[2]{\ensuremath{C^{(#1)}_{#2}}}
27: \newcommand{\ave}[1]{\langle #1 \rangle}
28:
29: \newcommand{\FM}[1]{\ensuremath{{\cal F}}^{\,\Box\,(#1)}}
30:
31: %\theoremstyle{definition}
32:
33: \newtheorem{theorem}{Theorem}
34: \newtheorem{fact}{Fact}
35: \newtheorem{prop}{Proposition}
36: \newtheorem{lem}{Lemma}
37: \newtheorem{definition}{Definition}
38: \newtheorem{conjecture}{Conjecture}
39:
40:
41: \title{Area distribution and scaling function \\for punctured polygons}
42:
43: \author{\sc Christoph Richard\dag, Iwan Jensen\ddag{} and Anthony J. Guttmann\ddag\\
44: \normalsize
45: \dag Fakult\"at f\"ur Mathematik, Universit\"at Bielefeld,\\
46: \normalsize
47: Postfach 10 01 31, 33501 Bielefeld, Germany\\
48: \normalsize
49: \ddag ARC Centre of Excellence for Mathematics and Statistics of Complex Systems,\\
50: \normalsize
51: Department of Mathematics and Statistics,
52: The University of Melbourne,\\
53: \normalsize
54: Victoria 3010, Australia}
55:
56: \begin{document}
57:
58: \maketitle
59:
60: \centerline{Mathematics Subject Classifications: 05A15, 05A16}
61: \centerline{PACS numbers: 05.50.+q, 05.70.Jk, 02.10.Ox}
62:
63: \begin{abstract}
64: Punctured polygons are polygons with internal holes which are
65: also polygons. The external and internal polygons are of the
66: same type, and they are mutually as well as self-avoiding.
67: Based on an assumption about the limiting area distribution
68: for unpunctured polygons, we rigorously analyse the effect of a
69: finite number of punctures on the limiting area distribution in
70: a uniform ensemble, where punctured polygons with equal perimeter
71: have the same probability of occurrence. Our analysis
72: leads to conjectures about the scaling behaviour
73: of the models.
74:
75: We also analyse exact enumeration data. For staircase polygons
76: with punctures of fixed size, this yields explicit expressions
77: for the generating functions of the first few area moments.
78: For staircase polygons with punctures of arbitrary size,
79: a careful numerical analysis yields very accurate estimates
80: for the area moments.
81: Interestingly, we find that the leading correction term for each area moment
82: is proportional to the corresponding area moment with one less puncture.
83: We finally analyse corresponding quantities for punctured self-avoiding
84: polygons and find agreement with the conjectured formulas to at
85: least 3--4 significant digits.
86: \end{abstract}
87:
88: \maketitle
89:
90: \section{Introduction}
91:
92: The behaviour of planar self-avoiding walks (SAW) and polygons (SAP) is one
93: of the classical unsolved problems, not only of algebraic combinatorics,
94: but also of chemistry and of physics \cite{MSbook,HughesV1,Vanderzande}.
95: In the field of algebraic combinatorics, it is a classical enumeration problem.
96: In chemistry and physics, SAWs and SAPs are used to model a variety of phenomena, including the
97: properties of long-chain polymers in dilute solution \cite{deGennes},
98: the behaviour of ring polymers and vesicles in general \cite{LSF87} and
99: benzenoid systems \cite{GCbook,VGJ02} in particular.
100: Though the qualitative form of the phase diagram \cite{FGW91} is known rigorously,
101: there is otherwise a paucity of rigorous results. However, there are a few conjectures,
102: including the exact values of the critical exponents \cite{N82,N84},
103: and more recently the limit distribution of area and scaling function for SAPs, when
104: enumerated by both area and perimeter \cite{RGJ01,C01,R02, RJG03, RJG04}.
105:
106: Models of planar polygons with punctures arise naturally as cross-sections of
107: three-dimensional vesicle models. In such cross-sections, there may be holes
108: within holes, and the number of punctures may be infinite. In this work, we exclude
109: these possibilities. Whereas our methods can be used to study the former
110: case, the second situation presents new difficulties, which we have not yet
111: overcome\footnote{Since punctured polygons with an unlimited number of punctures have,
112: in contrast to polygons without punctures, an (ordinary) perimeter
113: generating function with zero radius of convergence \cite{GJO01}, both the phase diagram
114: and the detailed asymptotics are clearly going to
115: be very different from those of polygons without punctures. This is discussed
116: further in the conclusion.}.
117:
118: In this work we consider the effect of a finite number of punctures in polygon
119: models, in particular we study staircase polygons and self-avoiding polygons on the
120: square lattice. The perimeter of a punctured polygon \cite{GJWE00, JvRW89} is the
121: perimeter of its boundary (both internal and external) while the area of a
122: punctured polygon is the area of the enclosed by the external perimeter minus the
123: area(s) of any holes\footnote{This has to
124: be distinguished from so-called composite polygons \cite{J99}.
125: The perimeter of a composite polygon is defined as the perimeter of
126: the external polygon only, resulting in asymptotic behaviour different
127: from punctured polygons. Moreover, composite polygons can have more complex
128: internal structure than just other polygons.}.
129: As discussed in section \ref{sec:punc} below, the effect of punctures
130: on the critical point and critical exponents of the area
131: and perimeter generating function has been the subject of previous studies, but
132: the effect of punctures on the critical amplitudes and detailed asymptotics
133: have not, to our knowledge, been previously considered.
134:
135: Apart from the intrinsic interest of the problem, we also believe it to be the appropriate
136: route to study the detailed asymptotics of {\it polyominoes},
137: since punctured polygons are a subclass of polyominoes.
138: While we still have some way to go to understand the polyomino phase diagram,
139: we feel that restricting the problem to this important subclass is the correct route.
140:
141: The make-up of the paper is as follows: In the next section we review the known situation
142: for the perimeter and area generating functions of punctured polygons and polyominoes.
143: In section 3 we review the phase diagram and scaling behaviour of staircase
144: polygons and self-avoiding polygons.
145: In section 4 we rigorously express the asymptotic behaviour of
146: models of punctured polygons in the limit of large perimeter in terms of the
147: asymptotic behaviour of the model without punctures, by refining arguments
148: used in \cite{GJWE00}. This leads, in particular,
149: to a characterisation of the limit distribution of the area of punctured polygons.
150: This result is then used to conjecture scaling functions of punctured polygons.
151: We consider three cases of increasing generality. First, we
152: consider the case of minimal punctures. It is
153: shown that effects of self-avoidance are asymptotically irrelevant, and that elementary area
154: counting arguments yield the leading asymptotic behaviour. We then discuss the
155: case of a finite number of punctures of bounded size, and finally the case of a
156: finite number of punctures of unbounded size. Results for the latter case
157: are given for models with a finite critical perimeter generating function
158: such as staircase polygons and self-avoiding polygons. Whereas the latter two
159: cases are technically more involved, the underlying arguments are similar to the case
160: of minimal punctures. If the critical perimeter generating function of the
161: polygon model without punctures is finite, then all three cases lead, up to
162: normalisation, to the same limit distributions
163: and scaling function conjectures.
164:
165: The next two sections discuss the development and application of
166: extensive numerical data to test the results of the previous section. Moreover,
167: the numerical analysis yields predictions, conjectured to be exact, for the corrections to the
168: asymptotic behaviour. In particular, section 5 describes the very efficient algorithms
169: used to generate the data, while section 6 applies a range of numerical tools to
170: the analysis of the generating functions for
171: punctured staircase polygons and then punctured self-avoiding polygons.
172: Here we wish to emphasise that our work on this problem
173: involved a close interplay between analytical and numerical work.
174: Initially, our intention was to check our predictions for scaling functions by
175: studying amplitude ratios for area moments (given in Table~\ref{tab:ratios}).
176: We subsequently discovered numerically the {\em exact} solutions for minimally
177: punctured staircase polygons. We also obtained very accurate estimates for the
178: amplitudes of staircase polygons with one or two punctures of arbitrary size.
179: From these results we were able to conjecture exact expressions for the amplitudes,
180: which in turn spurred us on to further analytical work in order to prove these results.
181: The final section summarises and discusses our results.
182:
183: \section{Punctured polygons \label{sec:punc}}
184:
185: We consider polygons on the square lattice in this article. In particular,
186: we study self-avoiding polygons and staircase polygons.
187: A self-avoiding polygon on a lattice can be defined as a walk along
188: the edges of the lattice, which starts and ends at the same lattice point, but
189: has no other self-intersections. When counting SAPs, they are generally considered
190: distinct up to translations,
191: change of starting point, and orientation of the walk, so if there are $p_m$
192: SAPs of length or perimeter $m$ there are $2mp_m$ walks (the factor of two
193: arising since the walk can go in two directions). On the square lattice the perimeter
194: of any polygon is always even so it is natural to count polygons by half-perimeter
195: instead of perimeter.
196: The area of a polygon is the number of lattice cells (times
197: the area of the unit cell) enclosed by the perimeter of the polygon.
198: A (square lattice) staircase polygon can be defined as the intersection of two mutually avoiding
199: directed walks starting at the same lattice point, moving only to the right or up and
200: terminating once the walks join at a vertex. Every staircase polygon is a self-avoiding
201: polygon. It is well known that the number $p_m$ of staircase polygons of half-perimeter
202: $m$ is given by the $(m-1)^{th}$ Catalan number, $p_m={2m-2 \choose m-1}/m$, with
203: half-perimeter generating function
204: %
205: \begin{equation}\label{eq:StGf}
206: \hspace{-2cm}
207: \PGf(x) =\sum_{m}p_m x^m= \frac{1-2x-\sqrt{1-4x}}{2} \sim \frac{1}{4}-
208: \frac{1}{2}(1-\mu x)^{2-\alpha}
209: \qquad (\mu x\nearrow 1),
210: \end{equation}
211: %
212: where the connective constant $\mu=4$ and the critical exponent $\alpha=3/2$.
213: Recall that $f(x)\sim g(x)$ as $x\nearrow x_c$ means that $\lim f(x)/g(x)=1$
214: as $x\to x_c$ from below. In addition, as usual, the rhs is understood as the first
215: two leading terms in an asymptotic expansion of the lhs about $x=1/\mu$, see
216: e.g.~\cite[Sec~1]{BH75}.
217:
218: Punctured polygons \cite{GJWE00} are polygons with internal holes which are
219: also polygons (the polygons are mutually- as well as self-avoiding). The perimeter
220: of a punctured polygon is the sum of the external and internal perimeters
221: while the area is the area of the external polygon {\em minus} the areas of the
222: internal polygons. We also consider polygons with {\em minimal} punctures,
223: that is, polygons where the punctures are unit cells (or polygons with perimeter 4
224: and area 1). Punctured staircase polygons are illustrated in figure~\ref{fig:poly}.
225: %
226: \begin{figure}
227: \begin{center}
228: \includegraphics[scale=0.7]{punctured_stair}
229: \end{center}
230: \caption{\label{fig:poly}
231: Examples of the types of staircase polygons we consider in this paper.
232: }
233: \end{figure}
234:
235: We briefly review the situation for SAPs with punctures. Analogous results
236: can be shown to hold for staircase
237: polygons with punctures. Square lattice SAPs with $r$ punctures, counted by area $n$,
238: were first studied by Janse van Rensburg and Whittington
239: \cite{JvRW89}. They proved the existence of an exponential growth constant
240: $\kappa^{(r)}$ satisfying $\kappa^{(r)}=\kappa^{(0)}=\kappa$.
241: Denoting the corresponding number of SAPs by $a_n^{(r)}$ and assuming
242: asymptotic behaviour of the form
243: %
244: \begin{displaymath}
245: a_n^{(r)} \sim A^{(r)}(\kappa^{(r)})^n n^{\beta_r-1}\qquad (n\to\infty),
246: \end{displaymath}
247: %
248: Janse van Rensburg proved \cite{JvR92} that $\beta_r = \beta_0 + r$.
249: These results of course translate to the singular behaviour of the corresponding
250: generating functions, defined by $\AGf^{(r)}(q) = \sum_{n > 0} a_n^{(r)}q^n$.
251:
252: In \cite{GJWE00} Guttmann, Jensen, Wong and Enting studied square lattice
253: SAPs with $r$ punctures counted by half-perimeter $m$. They proved the
254: existence of an exponential growth constant $\mu^{(r)}$ satisfying
255: $\mu^{(r)}=\mu^{(0)}=\mu$. If the corresponding number $p_m^{(r)}$
256: of SAPs is assumed to behave asymptotically as
257: %
258: \begin{displaymath}
259: p_m^{(r)}\sim B^{(r)}(\mu^{(r)})^m m^{\alpha_r-3}\qquad (m\to\infty),
260: \end{displaymath}
261: %
262: they argued, on the basis of a non-rigorous argument, that
263: $\alpha_r = \alpha_0 + \frac32 r$. Their results also translate to the associated
264: half-perimeter generating function $\PP{r}(x) = \sum_{m > 0} p_m^{(r)}x^m$
265: correspondingly.
266:
267: Similar results were obtained for polyominoes enumerated by number of cells
268: (i.e. area) with a finite number $r$ of punctures \cite{GJWE00}. It has been
269: proved that an exponential growth constant $\tau$ exists independently of $r$,
270: which satisfies $4.06258\approx \tau >\kappa \approx 3.97087$, where $\kappa$
271: is the growth constant for SAPs enumerated by area. If the number
272: $a_n^{(r)}$ of polynominoes of area $n$ with $r$ punctures is assumed to
273: satisfy asymptotically
274: %
275: \begin{displaymath}
276: a_n^{(r)} \sim C^{(r)}(\tau^{(r)})^n n^{\gamma_r-1}\qquad (n\to\infty),
277: \end{displaymath}
278: %
279: it has been shown that $\gamma_r=\gamma_0+r$ and hence that, if the exponents
280: $\gamma_r$ exist, they increase by 1 per puncture. It was further conjectured
281: on the basis of extensive numerical studies \cite{GJWE00}, that the number
282: $a_n^{(r)}$ satisfies asymptotically
283: %
284: \begin{displaymath}
285: a_n^{(r)} \sim \tau^n n^{r-1}\sum_{i \ge 0} C_i^{(r)}/n^{i}
286: \qquad (n\to\infty).
287: \end{displaymath}
288: %
289: Notice the conjecture $\gamma_0=0$ and that the correction terms
290: go down by a whole power.
291:
292: For unrestricted polyominoes, that is to say, with no restriction on the number
293: of punctures, it was proved by Guttmann, Jensen and Owczarek \cite{GJO01} that
294: the perimeter generating function has zero radius of convergence.
295: The perimeter is defined to be the perimeter of the boundary plus the
296: total perimeter of any holes.
297: If $p_{m}$ denotes the number of polyominoes, distinct up to a translation,
298: with half-perimeter $m$, they proved that $p_{m} = m^{m/4 + {\rm o}(m)}$,
299: meaning that
300: %
301: $$\lim_{m \to \infty}\frac {\log p_{m}}{m \log {m}} = \frac{1}{4}.$$
302: %
303: An attempt to study the quasi-exponential generating function with coefficients
304: $r_m = p_m/\Gamma(m/4+1)$ was equivocal.
305: For that reason, studying punctured self-avoiding polygons was
306: considered a controlled route to attempt to
307: determine the two-variable area-perimeter generating function of polyominoes.
308:
309: In passing, we note that in \cite{GJ06b} the exact solution of the perimeter generating
310: function for staircase polygons with a staircase hole is conjectured, in the form of
311: an $8^{th}$ order ODE. It is not obvious how to extract particular asymptotic information,
312: notably critical amplitudes from the solution without numerically integrating the ODE.
313: In the following, we will obtain such information by combinatorial arguments,
314: which refine those of \cite{GJWE00}.
315:
316:
317: \section{Polygon models and their scaling behaviour}
318:
319: We review the asymptotic behaviour of self-avoiding polygons
320: and staircase polygons following mainly \cite{FGW91}.
321: For concreteness, consider the fixed perimeter ensemble where, for
322: fixed half-perimeter $m$, each polygon of area $n$ has
323: a weight proportional to $q^n$, for some positive real number $q$.
324: If $0<q<1$, polygons of large area are exponentially suppressed,
325: so that typical polygons should be ramified objects. Since such
326: polygons would closely resemble branched polymers, the phase $0<q<1$
327: is also referred to as the {\it branched polymer phase}. As $q$
328: approaches unity, typical polygons should fill out more, and
329: become less string-like. For $q>1$, polygons of small area are
330: exponentially suppressed, so that typical polygons should become
331: ``fat''. Indeed, they resemble convex polygons \cite{OP99a} and
332: it has been proved \cite{FGW91} that the mean area of polygons of
333: half-perimeter $m$ grows asymptotically proportional to $m^2$.
334: In the extended phase $q=1$, it is numerically very well established that
335: the mean area of polygons of half-perimeter $m$ grows
336: asymptotically proportionally to $m^{3/2}$. In the branched polymer
337: phase $0<q<1$, the mean area of polygons of half-perimeter $m$ is
338: expected to grow asymptotically linearly in $m$, compare also
339: \cite[Thm 7.6]{J00} and \cite[Ch~IX.6, Ex.~12]{FS06}.
340:
341: This change of asymptotic behaviour of typical polygons w.r.t.~$q$
342: is reflected in the singular behaviour of the half-perimeter and area
343: generating function
344: %
345: \begin{displaymath}
346: \PGf(x,q)=\sum_{m,n} p_{m,n} x^m q^n,
347: \end{displaymath}
348: %
349: where $p_{m,n}$ denotes the number of (self-avoiding) polygons of
350: half-perimeter $m$ and area $n$. It has been proved \cite{FGW91}
351: that the free energy
352: %
353: \begin{displaymath}
354: \kappa (q):= \lim_{m \to \infty} \frac{1}{m} \log \left(\sum_n p_{m,n} q^n\right)
355: \end{displaymath}
356: %
357: exists and is finite if $0<q \le 1$. Further, $\kappa(q)$ is log-convex and continuous for these
358: values of $q$. It is infinite for $q>1$.
359: It was proved that for fixed $0<q\le1$, the radius of convergence $x_c(q)$ of $\PGf(x,q)$
360: is given by $x_c(q)= e^{-\kappa(q)}$.
361: For fixed $q>1$, $\PGf(x,q)$ has zero radius of convergence.
362: Fisher et al. \cite{FGW91} obtained rigorous upper and
363: lower bounds on $x_c(q)$.
364: The expected phase diagram, i.e., the radius of convergence of $\PGf(x,q)$
365: in the $x-q$ plane, as estimated numerically from extrapolation of SAP
366: enumeration data by perimeter and area,
367: is sketched qualitatively in figure~\ref{fig:phase}.
368: %
369: \begin{figure}[htbp]
370: \centering
371: \includegraphics[width=5cm]{phase}
372: \caption{\label{fig:phase} A sketch of the phase diagram of self-avoiding polygons.}
373: \end{figure}
374:
375: For $0<q<1$, the line $x_c(q)$ is, for self-avoiding polygons, expected to be a
376: line of logarithmic singularities of the generating function $\PGf(x,q)$.
377: For branched polymers in the continuum limit, the existence of the logarithmic
378: singularity has recently been proved \cite{BI03}.
379: The line $q=1$ is, for $0<x<x_c:=x_c(1)$, a line of finite essential singularities
380: \cite{FGW91}. For staircase polygons, counted by half-perimeter and area, the
381: corresponding phase diagram can be determined exactly, and is qualitatively
382: similar to that of self-avoiding polygons. Along the line $x_c(q)$ the
383: half-perimeter and area generating function diverges with a simple pole, and
384: the line $q=1$ is, for $0<x<x_c$, a line of finite essential
385: singularities \cite{Pr94}.
386:
387: \medskip
388:
389: We will focus on the uniform fixed perimeter ensemble $q=1$
390: in this article.
391: Whereas asymptotic area laws in the fixed perimeter ensemble are
392: expected to be Gaussian for positive $q\ne1$, the behaviour in the uniform
393: fixed perimeter ensemble $q=1$ is more interesting. For
394: staircase polygons, it can be shown that a limit distribution of
395: area exists and is given by the Airy distribution \cite{L84,T91,FL01}.
396: For self-avoiding polygons, it is conjectured that an area limit law exists and
397: is given by the Airy distribution, on the basis of a detailed numerical analysis
398: \cite{RGJ01,RJG03,RJG04}. See subsections \ref{sec:scaling_min} and
399: \ref{sec:scaling_func}.
400:
401: If $p_{m,n}$ denotes the number of polygons of half-perimeter $m$ and area $n$,
402: the existence and the form of a limit distribution can be inferred from the asymptotic
403: behaviour of the factorial moment coefficients $\sum_{n}(n)_k p_{m,n}$,
404: where $(a)_k=a\cdot(a-1)\cdot\ldots\cdot(a-k+1)$. The following
405: result is obtained by standard reasoning \cite{C74}.
406:
407: \begin{prop}\label{theo:general}
408: Let for $m,n\in\mathbb N_0$ real numbers $p_{m,n}$ be given.
409: Assume that the numbers $p_{m,n}$ have the asymptotic form, for $k\in\mathbb N_0$,
410: %
411: \begin{equation}\label{Ecr:ampli}
412: \sum_n (n)_k p_{m,n}\sim A_k x_c^{-m}
413: m^{\gamma_k-1}\qquad (m\to\infty)
414: \end{equation}
415: %
416: for positive real numbers $A_k$ and $x_c$, where $\gamma_k=(k-\theta)/\phi$,
417: with real constants $\theta$ and
418: $\phi>0$. Assume that the numbers $M_k:=A_k/A_0$ satisfy
419: the Carleman condition
420: %
421: \begin{equation}\label{Ecr:growth}
422: \sum_{k=0}^\infty (M_{2k})^{-1/{2k}}=+\infty.
423: \end{equation}
424: %
425: Then, for almost all $m$, the random variables $\widetilde X_m$
426: of area in the uniform fixed perimeter ensemble
427: %
428: \begin{displaymath}
429: \mathbb P(\widetilde X_m=n)=\frac{p_{m,n}}{\sum_np_{m,n}}
430: \end{displaymath}
431: %
432: are well defined. We have
433: %
434: \begin{displaymath}
435: X_m:=\frac{\widetilde X_m}{m^{1/\phi}} \stackrel{d}{\longrightarrow}X \qquad (m\to\infty),
436: \end{displaymath}
437: %
438: for a uniquely defined random variable $X$ with moments $M_k$,
439: where the superscript $\stackrel{d}{}$ denotes convergence in distribution.
440: We also have moment convergence.
441: \end{prop}
442:
443: \begin{proof}[Sketch of proof]
444: A straightforward calculation using Eq.~(\ref{Ecr:ampli}) leads to
445: %
446: \begin{displaymath}
447: \mathbb E[(\widetilde X_m)_k]
448: \sim \frac{A_k}{A_0} m^{k/\phi}
449: \qquad (m\to\infty).
450: \end{displaymath}
451: %
452: It follows that asymptotically
453: the factorial moments are equal to the (ordinary) moments.
454: Thus, the moments of $X_m$ have the same asymptotic form
455: %
456: \begin{displaymath}
457: \mathbb E[(X_m)^k]
458: \sim\frac{A_k}{A_0}=M_k\qquad (m\to\infty).
459: \end{displaymath}
460: %
461: Due to the growth condition Eq.~(\ref{Ecr:growth}), the sequence
462: $(M_k)_{k\in\mathbb N_0}$ defines a unique random variable $X$ with moments $M_k$.
463: Moment convergence of $(X_m)$ implies convergence in distribution,
464: see \cite[Thm 4.5.5]{C74} for the line of arguments.
465: \end{proof}
466:
467: \medskip
468:
469: The assumption Eq.~(\ref{Ecr:ampli})
470: translates, on the level of the half-perimeter and area generating function
471: $\PGf(x,q)$, to a certain asymptotic
472: behaviour of the so-called factorial moment generating functions
473: %
474: \begin{displaymath}
475: g_k(x)=\frac{(-1)^k}{k!}\left.\frac{\partial^k}{\partial q^k}
476: \PGf(x,q)\right|_{q=1}.
477: \end{displaymath}
478: %
479: It can be shown (compare \cite{F99}) that the asymptotic behaviour Eq.~(\ref{Ecr:ampli})
480: implies for $\gamma_k>0$ the asymptotic equivalence
481: %
482: %
483: \begin{equation}\label{eq:relation2}
484: g_k(x)\sim \frac{f_k}{(x_c-x)^{\gamma_k}}\qquad (x\nearrow x_c),
485: \end{equation}
486: %
487: where the amplitudes $f_k$ are related to the amplitudes $A_k$%
488: \footnote{
489: Note that our definition of the amplitudes $A_k$ differs
490: from that in \cite{R02} by a factor of $(-1)^kk!$ and from
491: that in \cite{R05,R06} by a factor of $k!$.}
492: in Proposition~\ref{theo:general} via
493: %
494: \begin{equation}\label{eq:relation}
495: f_k=\frac{(-1)^k}{k!}A_kx_c^{\gamma_k}\Gamma(\gamma_k).
496: \end{equation}
497: %
498: If $-1<\gamma_k<0$, the series $g_k(x)$ is convergent as $x\nearrow x_c$, and
499: the same estimate Eq.~(\ref{eq:relation2}) holds,
500: with $g_k(x)$ replaced by $g_k(x)-g_k(x_c)$, where
501: $g(x_c):=\lim_{x\nearrow x_c} g(x)$.
502: In order to deal with these two different cases, we define for
503: a power series $g(x)$ with radius of convergence $x_c$, the number
504: %
505: \begin{displaymath}
506: g^{(c)}=\left\{
507: \begin{array}{cc}
508: g(x_c)& {\rm if} \,\,|\lim_{x\nearrow x_c} g(x)|<\infty\\
509: 0 & {\rm otherwise}.
510: \end{array}
511: \right.
512: \end{displaymath}
513: %
514: Adopting the generating function
515: point of view, the amplitudes $f_k$ determine the numbers $A_k$ and
516: hence the moments $M_k=A_k/A_0$ of the limit distribution.
517: The formal series $F(s)=\sum_{k\ge0} f_k s^{-\gamma_k}$ will appear
518: frequently in the sequel.
519:
520: \begin{definition}\label{def:fscl}
521: %
522: For the generating function $\PGf(x,q)$ of a class of
523: self-avoiding polygons, denote its factorial moment
524: generating functions by
525: %
526: \begin{displaymath}
527: g_k(x)=\frac{(-1)^k}{k!}\left.\frac{\partial^k}{\partial q^k}
528: \PGf(x,q)\right|_{q=1}.
529: \end{displaymath}
530: %
531: Assume that the factorial moment generating functions satisfy
532: %
533: \begin{equation}\label{form:approx}
534: g_k(x)-g_k^{(c)} \sim \frac{f_k}{(x_c-x)^{\gamma_k}} \qquad (x\nearrow x_c),
535: \end{equation}
536: %
537: with real exponents $\gamma_k$. Then, the formal series
538: %
539: \begin{displaymath}
540: F(s) = \sum_{k\ge0}f_k s^{-\gamma_k}
541: \end{displaymath}
542: %
543: is called the area amplitude series.
544: \end{definition}
545:
546: The area amplitude series is expected to approximate
547: the half-perimeter and area generating function $\PGf(x,q)$ about
548: $(x,q)=(x_c,1)$. This is motivated by the following heuristic
549: argument. Assume that $\gamma_k=(k-\theta)/\phi$ with $\phi>0$ and argue
550: %
551: \begin{eqnarray*}
552: \PGf(x,q)&\approx \sum_{k\ge0}
553: \left(g_k^{(c)}+\frac{f_k}{(x_c-x)^{\gamma_k}}\right)(1-q)^k\\
554: &\approx \left(\sum_{k\ge0}g_k^{(c)}(1-q)^k\right)+(1-q)^{\theta}\left(\sum_{k\ge0}
555: f_k \left( \frac{x_c-x}{(1-q)^{\phi}}\right)^{-\gamma_k}
556: \right).
557: \end{eqnarray*}
558: %
559: In the above calculation, we formally expanded $\PGf(x,q)$
560: about $q=1$ and then replaced the Taylor coefficients by their
561: leading singular behaviour about $x=x_c$.
562: In the rhs of the above expression, the first sum is by assumption finite,
563: and the second term contains the area amplitude
564: series $F(s)$ of combined argument $s=(x_c-x)/(1-q)^\phi$.
565: This motivates the following definition. A class of self-avoiding
566: polygons is a subset of self-avoiding polygons. Prominent examples are,
567: among others \cite{BM96}, self-avoiding polygons and staircase polygons.
568:
569: \begin{definition}\label{def:sf}
570: Let a class of square lattice self-avoiding polygons be given, with
571: half-perimeter and area generating function $\PGf(x,q)$.
572: Let $0<x_c<\infty$ be the radius of convergence
573: of the half-perimeter generating function $\PGf(x,1)$. Assume that there
574: exist a constant $s_0\in[-\infty,0)$, a function
575: ${\cal F}:(s_0,\infty)\to\mathbb R$, a real constant $A$ and real
576: numbers $\theta$ and $\phi>0$, such
577: that the generating function $\PGf(x,q)$ satisfies, for real
578: $x$ and $q$, where $0<q<1$ and $(x_c-x)/(1-q)^\phi\in(s_0,\infty)$,
579: the asymptotic equivalence
580: %
581: \begin{equation}\label{form:scaling}
582: \PGf(x,q)\sim A+(1-q)^\theta {\cal F}\left(\frac{x_c-x}{(1-q)^{\phi}}\right)
583: \qquad (x,q)\longrightarrow (x_c,1).
584: \end{equation}
585: %
586: Then, the function ${\cal F}(s)$ is called a {\it scaling
587: function} of combined argument
588: $s=(x_c-x)/(1-q)^{\phi}$, and $\theta$ and $\phi$
589: are called {\it critical exponents}.
590: \end{definition}
591:
592: \noindent {\bf Remarks.} {\it i)}
593: Due to the restriction on the argument of the scaling function, the
594: limit $(x,q)\to(x_c,1)$ is approached for values $(x,q)$ satisfying
595: $x<x_0(q)$ and $q<1$, where $x_0(q)= x_c-s_0(1-q)^\phi$.
596: %If $s_0$ is a singularity of $F(s)$, the curve $x_0(q)$ describes the
597: %radius of convergence of $\PGf(x,q)$ for $q<1$ near $q=1$.
598: \\
599: {\it ii)} The above scaling form is also suggested by the theory
600: of tricritical scaling, adapted to polygon models \cite{BOP93}. The
601: scaling function describes the leading singular behaviour
602: of $\PGf(x,q)$ about the point $(x_c,1)$ where the two lines of
603: qualitatively different singularities meet. \\
604: {\it iii)} The additional condition $\phi>\theta$ and
605: $\theta\notin \mathbb N_0$ ensure that $\gamma_k\in(-1,\infty)
606: \setminus\{0\}$.
607: Then, by the above argument, it is plausible that there exists
608: an asymptotic expansion of the scaling function ${\cal F}(s)$
609: about infinity coinciding with the area amplitude series
610: $F(s)$, i.e., ${\cal F}(s)\sim F(s)$ as $s\to\infty$. Recall that
611: $s$ is considered to be a real parameter.
612:
613: \medskip
614:
615: For staircase polygons the existence of a scaling form
616: Eq.~(\ref{form:scaling}) has been proved \cite[Thm~5.3]{Pr94},
617: with scaling function ${\cal F}(s):(s_0,\infty)\to\mathbb R$
618: explicitly given by
619: %
620: \begin{equation}\label{sfstair}
621: {\cal F}(s) = \frac{1}{16}\frac{\rm d}{{\rm d}s}\log\mbox{Ai}
622: \left( 2^{8/3} s \right),
623: \end{equation}
624: %
625: with exponents $\theta=1/3$ and $\phi=2/3$ and $x_c=1/4$, where
626: $\mbox{Ai}(x)=\frac{1}{\pi}\int_0^\infty \cos(t^3/3+tx){\rm d}t$ is the
627: Airy function. The constant $s_0$ is such that $2^{8/3}s_0$ is the
628: location of the Airy function zero of smallest modulus. For {\it rooted}
629: SAPs with half-perimeter and area generating
630: function $\PGf_r(x,q)=x\frac{{\rm d}}{{\rm d}x}\PGf(x,q)$, the conjectured form of
631: the scaling function ${\cal F}_r(s):(s_0,\infty)\to\mathbb R$ is \cite{R02}
632: %
633: \begin{displaymath}
634: {\cal F}_r(s) = \frac{x_c}{2\pi}\frac{\rm d}{{\rm d}s}\log\mbox{Ai} \left( \frac{\pi}{x_c}
635: \left( 4A_0\right)^\frac{2}{3} s \right),
636: \end{displaymath}
637: %
638: with the same exponents as for staircase polygons, $\theta=1/3$
639: and $\phi=2/3$. Here, $x_c=0.14368062927(2)$ is the radius of convergence
640: of the half-perimeter generating function $\PGf_r(x,1)$ of (rooted) SAPs, and
641: $A_0=0.09940174(4)$ is the critical amplitude $\sum_n m p_{m,n}\sim A_0
642: x_c^{-m}m^{-3/2}$ of rooted SAPs, which coincides with the critical amplitude
643: $A_0$ of (unrooted) SAPs. Again, the constant $s_0$ is such that the corresponding
644: Airy function argument is the location of the Airy function zero of smallest modulus.
645: This conjecture was based on the conjecture that both models
646: have, up to normalisation constants, the same area amplitude series.
647: The latter conjecture is supported numerically
648: to very high accuracy by an extrapolation of the moment series using exact
649: enumeration data \cite{RGJ01,RJG03}. The conjectured form of the scaling
650: function ${\cal F}(s):(s_0,\infty)\to\mathbb R$ for SAPs is obtained
651: by integration,
652: %
653: \begin{equation}\label{form:unroot}
654: {\cal F}(s) = -\frac{1}{2\pi} \log\mbox{Ai}
655: \left( \frac{\pi}{x_c} \left( 4A_0\right)^\frac{2}{3} s \right)+C(q),
656: \end{equation}
657: %
658: with exponents $\theta=1$ and $\phi=2/3$. In the above formula, $C(q)$
659: is a $q$ dependent constant of integration, $C(q) = \frac{1}{12 \pi}
660: (1 - q)\log(1 - q)$, see \cite{RJG04}.
661: Corresponding results for the triangular and hexagonal lattices can be found in \cite{RGJ01}.
662:
663: For models of punctured polygons with a finite number of punctures, we have
664: qualitatively the same phase diagram as for polygon models without punctures,
665: however with different critical exponents $\theta$ depending on the number of
666: punctures \cite{JvR92,GJWE00}, and hence we expect different scaling functions.
667: We will focus on critical exponents and area limit laws in the uniform
668: ensemble $q=1$ in the following section. This will lead to conjectures for the
669: corresponding scaling functions.
670:
671: \section{Scaling behaviour of punctured polygons \label{sec:scaling}}
672: We briefly preview the main results of this section.
673: In subsection~\ref{sec:scaling_min} we study polygons with a finite number
674: of minimal punctures. Our result assumes a certain asymptotic form for
675: the area moment coefficients for unpunctured polygons. This `assumed' form is
676: known to be true for staircase polygons and many other models and universally
677: accepted as true for self-avoiding polygons. Given this assumption, we prove
678: that the asymptotic behaviour of the area moment coefficients for minimally
679: punctured polygons can be expressed in terms of the asymptotic behaviour of
680: unpunctured polygons.
681: In particular we derive expressions for the leading amplitude of the area moments
682: for punctured polygons in terms of the amplitudes for unpunctured
683: polygons. For staircase polygons this leads to exact formulas for the amplitudes.
684: For self-avoiding polygons the formulas contain certain constants which aren't
685: known exactly but can be estimated numerically to a very high degree of accuracy.
686: In subsection~\ref{sec:scaling_bound} we extend the study and proofs to
687: polygons with a finite number of punctures of {\em bounded} size and then
688: in subsection~\ref{sec:scaling_arbitrary} to models with punctures of arbitrary
689: or unbounded size. Finally in subsection~\ref{sec:scaling_func} we consider
690: the consequences of our results for the area limit laws of punctured polygons
691: and we present conjectures for the scaling functions.
692:
693: \subsection{Polygons with $r$ minimal punctures \label{sec:scaling_min}}
694:
695: For polygon models with rational perimeter generating functions,
696: corresponding models with minimal punctures have been studied in
697: \cite{RG01}. In particular,
698: a method to derive explicit expressions for generating functions
699: of exactly solvable models with a minimal puncture was
700: given \cite[Appendix]{RG01}. It has been applied to Ferrers
701: diagrams, whose perimeter and area generating function satisfies
702: a linear $q$-difference equation, see \cite[Eq.~(54)]{RG01}.
703: The method can also be applied to the model of staircase polygons,
704: whose half-perimeter and area generating function ${\cal P}(x,q)$
705: satisfies the quadratic $q$-difference equation
706: %
707: \begin{equation}\label{sp0}
708: {\cal P}(x,q)=\frac{x^2q}{1-2qx-{\cal P}(qx,q)}.
709: \end{equation}
710: %
711: Let $\StM{r}(x,q)$ denote the half-perimeter and area generating function
712: of staircase polygons with $r$ minimal punctures. We have the following
713: result for the case $r=1$.
714: %
715: \begin{fact}\label{fact}
716: The half-perimeter and area generating function of staircase polygons with a
717: single minimal puncture $\StM{1}(x,q)$ is given by
718: %
719: \begin{equation}\label{sp1}
720: \StM{1}(x,q)=\frac{x^4}{(1-2qx-{\cal P}(qx,q))^2}
721: \left( {\cal P}(qx,q)-qx \frac{\partial{\cal P} }{\partial x}(qx,q)+
722: q \frac{\partial {\cal P}}{\partial q}(qx,q) \right),
723: \end{equation}
724: %
725: where ${\cal P}(x,q)$ satisfies Eq.~(\ref{sp0}).
726: \qed
727: \end{fact}
728: %
729: \noindent {\bf Remarks.}
730: %
731: {\it i)} For a proof of Fact \ref{fact}, proceed along the lines of
732: \cite[Appendix]{RG01}. We do not give the details, since we are mainly
733: interested in asymptotic results, for which we will give an elementary
734: combinatorial derivation, valid for arbitrary $r$. See
735: Proposition \ref{prop:minpunc} and its subsequent extensions.\\
736: %
737: {\it ii)} For polygons with $r$ punctures, their {\bf $k^{th}$} area moment generating
738: functions are defined by $\StM{r}_k(x)= \left(q\frac{\partial}
739: {\partial q}\right)^k \left.\StM{r}(x,q) \right|_{q=1}$. The above equations can be used
740: to obtain explicit expressions for the area moment generating functions $\StM{1}_k(x)$
741: by implicit differentiation. The functions $\StM{1}_k(x)$ also appear in
742: section \ref{sec:stair_min}.\\
743: %
744: {\it iii)} Assuming that $\StM{1}(x,q)$ has scaling behaviour of the form
745: %
746: \begin{displaymath}
747: \StM{1}(x,q) \sim (1-q)^{\theta_1}\FM{1}((x_c-x)(1-q)^{-\phi_1})
748: \end{displaymath}
749: %
750: about $(x,q)=(x_c,1)$, and the necessary analyticity conditions for the
751: validity of the following calculation, we can
752: express the scaling function $\FM{1}(s)$ of staircase polygons with
753: a single minimal puncture in terms of the
754: known scaling function ${\cal F}(s)$ of staircase polygons Eq.~(\ref{sfstair}).
755: From Eq.~(\ref{sp1}) we infer that $\theta_1=-2/3$, $\phi_1=2/3$ and
756: %
757: \begin{equation}\label{stsf}
758: \FM{1}(s)=\frac{1}{24}s{\cal F}'(s)-\frac{1}{48} {\cal F}(s).
759: \end{equation}
760:
761: \medskip
762:
763: In principle, the method of \cite[Appendix]{RG01} can
764: be used to analyse the case of several minimal punctures.
765: However, the analysis becomes quite cumbersome. On the other
766: hand, the previous result suggests simple expressions for
767: the scaling functions of models with several punctures in terms
768: of that without a puncture. Moreover, we expect such
769: a phenomenon also to occur for models where an exact
770: solution does not exist or is not known. This is
771: discussed next. We will asymptotically analyse the area
772: moments of a polygon model with punctures and
773: draw conclusions about their possible scaling behaviour.
774:
775: For a class of punctured self-avoiding
776: polygons, consider their area moment coefficients
777: %
778: \begin{displaymath}
779: p_m^{\Box(r,k)}:= \sum_n n^kp_{m,n}^{\Box(r)},
780: \end{displaymath}
781: %
782: where $p_{m,n}^{\Box(r)}$ denotes the number of polygons
783: in the class with $r$ minimal punctures, $r\in\mathbb N_0$,
784: of half-perimeter $m$ and area $n$. For simplicity of notation,
785: we write $p_{m,n}:=p_{m,n}^{\Box(0)}$ and $p_m^{(k)}:=
786: p_{m}^{\Box(0,k)}$. The area moments in the uniform fixed
787: perimeter ensemble are expressed in terms of
788: the area moment coefficients via
789: %
790: \begin{equation}
791: \mathbb E[(\widetilde X_m^{\Box(r)})^k]=\frac{\sum_n n^k p_{m,n}^{\Box(r)}}
792: {\sum_n p_{m,n}^{\Box(r)}} = \frac{p_m^{\Box(r,k)}}{p_m^{\Box(r,0)}}.
793: \end{equation}
794: %
795:
796: \medskip
797:
798: \begin{prop}\label{prop:minpunc}
799: Assume that, for a class of self-avoiding polygons without punctures,
800: the area moment coefficients $p_m^{(k)}$ have
801: the asymptotic form, for $k\in\mathbb N_0$,
802: %
803: \begin{equation}\label{pmas}
804: p_m^{(k)} \sim A_k x_c^{-m} m^{\gamma_k-1} \qquad (m\to\infty),
805: \end{equation}
806: %
807: for numbers $A_k>0$, $x_c>0$ and exponents $\gamma_k=(k-\theta)/\phi$,
808: where $\theta$ and $\phi$ are real constants and $0<\phi<1$.
809: Then, the area moment coefficient $p_m^{\Box(r,k)}$ of the polygon class
810: with $r\ge1$ minimal punctures is asymptotically given by, for $k\in\mathbb N_0$,
811: %
812: \begin{equation}\label{form:minprk}
813: p_m^{\Box(r,k)} \sim A_k^{(r)} x_c^{-m} m^{\gamma_k^{(r)}-1} \qquad (m\to\infty),
814: \end{equation}
815: where $A_k^{(r)}=A_{k+r}x_c^{2r}/r!$ and $\gamma_k^{(r)}=\gamma_{k+r}$.
816: \end{prop}
817:
818: \begin{proof}
819: We will derive upper and lower bounds on $p_m^{\Box(r,k)}$, which will be
820: shown to coincide asymptotically. Let us call two polygons interacting if
821: their boundary curves have non-empty intersection. An upper bound is obtained
822: by allowing for interaction between all constituents of a punctured polygon.
823: Let a polygon $\Pol$ of half-perimeter $m-2r$ and area $n+r$ be given. The
824: number of ways of placing $r$ squares inside $\Pol$ is clearly less than
825: $(n+r)^r/r!$. We thus have
826: %
827: \begin{displaymath}
828: p_m^{\Box(r,k)}\le {\widetilde p}_m^{(r,k)} := \frac{1}{r!}
829: \sum_{n\ge 1} n^k (n+r)^rp_{m-2r,n+r} =\frac{1}{r!}
830: \sum_{n\ge r+1} (n-r)^k n^rp_{m-2r,n}.
831: \end{displaymath}
832: %
833: By Bernoulli's inequality, we get for ${\widetilde p}_m^{(r,k)}$ the
834: bound
835: %
836: \begin{displaymath}
837: \frac{1}{r!} \sum_{n\ge r+1} \left( n^{k+r}-kr\,n^{k+r-1}\right)
838: p_{m-2r,n}\le {\widetilde p}_m^{(r,k)}\le \frac{1}{r!}
839: \sum_{n\ge r+1} n^{k+r}p_{m-2r,n}.
840: \end{displaymath}
841: %
842: For every polygon of perimeter $2s$ and area $t$ we
843: have $t\ge s-1$. Thus, for $m$ sufficiently large, we can
844: replace the lower bound of summation $r+1$ by zero.
845: In particular, the latter relation is for $m\ge 3r+2$
846: equivalent to
847: %
848: \begin{displaymath}
849: \frac{1}{r!} \left( p_{m-2r}^{(k+r)}-kr\,p_{m-2r}^{
850: (k+r-1)} \right)\le {\widetilde p}_m^{(r,k)}\le
851: \frac{1}{r!} p_{m-2r}^{(k+r)}.
852: \end{displaymath}
853: %
854: The assumption Eq.~(\ref{pmas}) on the asymptotic
855: behaviour of $p_{m}^{(k)}$ then implies that
856: %
857: \begin{displaymath}
858: {\widetilde p}_m^{(r,k)}\sim \frac{x_c^{2r}}{r!}p_m^{
859: (k+r)}\qquad (m\to\infty).
860: \end{displaymath}
861: %
862: We derive a lower bound by subtracting from the upper bound an
863: upper bound on the number of square-square and square-boundary
864: interactions. Clearly, square-square interactions are only
865: present for $r>1$. For a given polygon $\Pol$, the number
866: of square-square interactions of $r$ squares is smaller than the number of
867: interactions between two squares, where the remaining
868: $r-2$ squares may occur at arbitrary positions within the polygon. There
869: are five possible configurations for an interaction between two
870: squares, yielding the upper bound $5(n+r)(n+r)^{r-2}$. Thus, the
871: contribution to ${\widetilde p}_{m}^{(r,k)}$ from square-square
872: interactions is bounded from above by
873: %
874: \begin{displaymath}
875: \sum_{n\ge 1} n^k 5(n+r)(n+r)^{r-2}p_{m-2r,n+r}=
876: 5(r-1)!{\widetilde p}_{m}^{(r-1,k)},
877: \end{displaymath}
878: %
879: which is asymptotically negligible compared to ${\widetilde p}_{m}
880: ^{(r,k)}$. Similarly, the number of configurations arising from
881: square-boundary interactions is bounded from above by $\sum_{j=1}^r
882: 4^j(m-2r)^j(n+r)^{r-j}$. This bound is obtained by estimating the
883: number of configurations of $j$ squares at the boundary by
884: $4^j(m-2r)^j$, the factor 4 arising from edge and vertex interactions,
885: the factor $(n+r)^{r-j}$ accounting for arbitrary positions of the
886: remaining $(r-j)$ squares. We thus get an upper bound
887: %
888: \begin{eqnarray*}
889: \sum_{j=1}^r 4^j (m-2r)^j (r-j)! {\widetilde p}_m^{(r-j,k)}
890: \sim \sum_{j=1}^r 4^j x_c^{2r-2j}(m-2r)^j p_m^{(k+r-j)}\\
891: \sim 4 x_c^{2r-2}m\, p_m^{(k+r-1)}
892: \qquad (m\to\infty).
893: \end{eqnarray*}
894: %
895: By assumption, the latter bound is asymptotically negligible compared to
896: ${\widetilde p}_{m}^{(r,k)}$. Thus, the lower bound is asymptotically
897: equal to the upper bound, which yields the assertion of the proposition.
898: \end{proof}
899:
900: \noindent {\bf Remarks.}
901: %
902: {\it i)} Proposition \ref{prop:minpunc} expresses the asymptotic
903: behaviour of the area moment coefficients of minimally
904: punctured polygons in terms of those of polygons without punctures.
905: The assumption Eq.~(\ref{pmas}) on the growth of the area moment coefficients
906: of the model without punctures is satisfied for the
907: usual polygon models \cite{BM96}. The asymptotic behaviour of some models
908: satisfying $\phi=1$, to which Proposition \ref{prop:minpunc} does not
909: apply, has been studied in \cite{RG01}.\\
910: \noindent {\it ii)}
911: As discussed in the previous subsection, the amplitudes
912: $A_k$ are related to the amplitudes $f_k$ of Eq.~(\ref{form:approx}) by
913: Eq.~(\ref{eq:relation}), if $\gamma_k\in(-1,\infty)\setminus\{0\}$.
914: For staircase polygons, where $\theta=1/3$
915: and $\phi=2/3$, we have explicit expressions for the
916: amplitudes $A_k$. More generally, it has been shown \cite{R02,R05,R06}
917: that, for classes of polygon models whose generating function satisfies a
918: $q$-functional equation with a square root as the dominant singularity
919: of their perimeter generating function, we have $f_k=c_kf_1^kf_0^{1-k}$,
920: where the numbers $c_k$ are, for $k\ge1$, given by
921: %
922: \begin{equation}\label{ck}
923: \gamma_{k-1}c_{k-1}+\frac{1}{4}\sum_{l=0}^{k}c_{k-l}c_l=0, \qquad c_0=1.
924: \end{equation}
925: %
926: The critical point $x_c$ as well as $f_0$ and $f_1$
927: are model dependent constants. For staircase polygons we have
928: $x_c=1/4$, $f_0=-1$ and $f_1=-1/64$.\\
929: \noindent {\it iii)} Rooted self-avoiding polygons are conjectured to also have
930: the exponents $\theta=1/3$ and $\phi=2/3$. In this case the asymptotic form
931: Eq.~(\ref{pmas}) and the form of the amplitudes $A_k$, given in
932: Eqs.~(\ref{eq:relation}) and (\ref{ck}), has been tested for $k\le10$
933: and shown to hold for to a high degree of
934: numerical accuracy \cite{RJG03}. Here $x_c=0.14368062927(2)$ is the
935: radius of convergence of the (rooted) SAP half-perimeter generating function,
936: $f_0=-0.929607(1)$ and $f_1=-x_c/(8\pi)$ are the rooted SAP critical amplitudes
937: as in Eq.~(\ref{form:approx}). We conjecture that the asymptotic form~(\ref{pmas})
938: holds for rooted SAPs for all values of $k$. Accepting this conjecture to be true,
939: Proposition~\ref{prop:minpunc} gives the asymptotic behaviour
940: for rooted self-avoiding polygons with $r$ minimal punctures.
941: By definition, unrooted SAPs have the same amplitudes $A_k$.
942: \\
943: %
944: {\it iv)} The crude combinatorial estimates of interactions in the proof
945: of Proposition~\ref{prop:minpunc} cannot be used to obtain corrections to the asymptotic
946: behaviour. See also the discussion in the conclusion.
947:
948: \subsection{Polygons with $r$ punctures of bounded size \label{sec:scaling_bound}}
949:
950: The arguments in the above proof can be applied to
951: obtain results for polygon models with a finite number
952: of punctures of bounded size.
953: The following theorem generalises Proposition~\ref{prop:minpunc} and serves as
954: preparation for the next section, where the case of a finite number
955: of punctures of arbitrary size is discussed.
956: For a class of punctured self-avoiding polygons, consider
957: their area moment coefficients
958: %
959: \begin{displaymath}
960: p_m^{(r,k,s)}:= \sum_n n^kp_{m,n}^{(r,s)},
961: \end{displaymath}
962: %
963: where $p_{m,n}^{(r,s)}$ denotes the number of polygons in the
964: class of half-perimeter $m$ and area $n$ with $r$ punctures,
965: $r\in\mathbb N_0$, obeying the condition that the sum of the
966: half-perimeter values of the puncturing polygons equals $s$.
967: For simplicity of notation, we write $p_{m,n}:=p_{m,n}^{(0,0)}$,
968: $p_m^{(k)}:=p_{m}^{(0,k,0)}$ and $p_m:=p_m^{(0)}$.
969:
970: \begin{theorem}\label{theo:bounded}
971: Assume that, for a class of self-avoiding polygons without
972: punctures, the area moment coefficients $p_m^{(k)}$ have
973: the asymptotic form, for $k\in\mathbb N_0$,
974: %
975: \begin{displaymath}
976: p_m^{(k)} \sim A_k x_c^{-m} m^{\gamma_k-1} \qquad
977: (m\to\infty),
978: \end{displaymath}
979: %
980: for numbers $A_k>0$, $x_c>0$ and $\gamma_k=(k-\theta)/\phi$,
981: where $\theta$ and $\phi$ are real constants and
982: $0<\phi<1$. Denote its half-perimeter generating function
983: by ${\cal P}(x)=\left(\sum_{m\ge0} x^m\, p_m\right)$.
984: Fix $r\ge1$ und $s\in\mathbb N$ such that $[x^s]({\cal P}(x))^r\ne0$.
985: Then, the area moment coefficient $p_m^{(r,k,s)}$ of the polygon class
986: with $r\ge1$ punctures whose half-perimeter sum equals is
987: asymptotically given by, for $k\in\mathbb N_0$,
988: %
989: \begin{equation}\label{eq:Aks}
990: p_m^{(r,k,s)} \sim A_k^{(r,s)} x_c^{-m}m^{\gamma_k^{(r)}-1}
991: \qquad (m\to\infty),
992: \end{equation}
993: %
994: where $\gamma_k^{(r)}=\gamma_{k+r}$ and $A_{k}^{(r,s)}=\frac{A_{k+r}}{r!}
995: x_c^s[x^s]({\cal P}(x))^r$.
996: \end{theorem}
997:
998:
999: \noindent {\bf Remarks.} {\it i)}
1000: With $s=2$, Theorem~\ref{theo:bounded} reduces
1001: to Proposition~\ref{prop:minpunc}.
1002: By summation, we also obtain the asymptotic behaviour for models with
1003: $r$ punctures of total half-perimeter less or equal to $s$. Note that
1004: we have the formal identity
1005: %
1006: \begin{displaymath}
1007: \sum_{s=0}^\infty x^s[x^s]({\cal P}(x))^r= ({\cal P}(x))^r.
1008: \end{displaymath}
1009: %
1010: The above expressions are convergent for $|x|<x_c$. If $\theta>0$, the sum is also
1011: convergent in the limit $x\nearrow x_c$.\\
1012: {\it ii)} The remarks following the proof
1013: of Proposition~\ref{prop:minpunc} also apply to Theorem~\ref{theo:bounded}.
1014:
1015: \begin{proof}
1016: This proof is a direct extension of the proof of Proposition~\ref{prop:minpunc} to the
1017: case of a finite number of punctures of bounded size. We consider a model of
1018: punctured polygons where, for fixed $s$, the
1019: $r$ punctures of half-perimeter $s_i$ and area $t_i$ satisfy
1020: $s_1+\ldots+s_r= s$.
1021: We give an asymptotic estimate for $p_m^{(r,k,s)}$. Let a polygon
1022: $\Pol$ of half-perimeter $m-|\boldsymbol s|$ and of area $n+
1023: |\boldsymbol t|$, where $|\boldsymbol s|=s_1+\ldots+s_r$ and
1024: $|\boldsymbol t|=t_1+\ldots+t_r$, be given. To obtain an upper
1025: bound for $p_m^{(r,k,s)}$, ignore all interactions between
1026: components of a punctured polygon. Recall that two polygons
1027: interact if their boundary curves have non-empty intersection.
1028: The number of ways of placing $r$ punctures inside $\Pol$ is
1029: clearly smaller than
1030: %
1031: \begin{displaymath}
1032: (n+|\boldsymbol t|)^r/r!.
1033: \end{displaymath}
1034: %
1035: This bound is obtained by considering the number of ways of placing
1036: the lower left corner of each puncture on each square plaquette
1037: inside the polygon. Note that, unlike in the proof of Proposition
1038: \ref{prop:minpunc}, this bound also counts configurations where punctures
1039: protrude from the boundary of $\Pol$. We will compensate for these
1040: over-counted configurations when deriving a lower bound for
1041: $p_m^{(r,k,s)}$. We have
1042: %
1043: \begin{eqnarray}
1044: p_m^{(r,k,s)} &\le {\widetilde p}_m^{(r,k,s)}:= \frac{1}{r!}
1045: \sum_{|\boldsymbol s|= s}\sum_{t_i}\sum_{n\ge 1} n^k (n+|\boldsymbol t|)^rp_{m-
1046: |\boldsymbol s|,n+|\boldsymbol t|}\prod_{i=1}^r p_{s_i,t_i}\nonumber\\
1047: &=\frac{1}{r!} \sum_{|\boldsymbol s|= s}\sum_{t_i}\sum_{n\ge |\boldsymbol t|+1}
1048: (n-|\boldsymbol t|)^k n^r p_{m-|\boldsymbol s|,n}\prod_{i=1}^r \label{form:theo1eq}
1049: p_{s_i,t_i},
1050: \end{eqnarray}
1051: %
1052: where the first sum is over the variables $s_1,\ldots,s_r$
1053: subject to the restriction $|\boldsymbol s|= s$, and the second
1054: sum is over all values of the variables $t_1,\ldots,t_r$.
1055: Note that, for $m$ fixed, all sums are finite.
1056: Invoking Bernoulli's inequality, we obtain the bound
1057: %
1058: \begin{eqnarray*}
1059: \frac{1}{r!} \sum_{|\boldsymbol s|= s}\sum_{t_i}&\sum_{n\ge |\boldsymbol t|+1}
1060: \left( n^{k+r}-k|\boldsymbol t|\, n^{k+r-1}\right)
1061: p_{m-|\boldsymbol s|,n}\prod_{i=1}^r p_{s_i,t_i}\le
1062: {\widetilde p}_m^{(r,k,s)}\\
1063: &\le \frac{1}{r!}\sum_{|\boldsymbol s|= s}\sum_{t_i}\sum_{n\ge |\boldsymbol t|+1}
1064: n^{k+r}p_{m-|\boldsymbol s|,n}\prod_{i=1}^r p_{s_i,t_i}.
1065: \end{eqnarray*}
1066: %
1067: Consider first the asymptotic behaviour of the expression
1068: %
1069: \begin{displaymath}
1070: {\widetilde a}_{m,s}:= \sum_{|\boldsymbol s|= s}\sum_{t_i}
1071: \sum_{n\ge |\boldsymbol t|+1} n^{k+r}p_{m-|\boldsymbol s|,n}
1072: \prod_{i=1}^r p_{s_i,t_i}.
1073: \end{displaymath}
1074: %
1075: If $m\ge |\boldsymbol s|^2+|\boldsymbol s|+2$, then the
1076: lower bound of summation on the index $n$ may be replaced by zero.
1077: This follows from the estimate $t_i\le s_i^2$, being valid
1078: for every self-avoiding polygon of half-perimeter $s_i$ and area $t_i$.
1079: Thus $|\boldsymbol t|\le |\boldsymbol s|^2$, and we argue
1080: that $n\ge m-|\boldsymbol s|-1\ge |\boldsymbol s|^2+1\ge
1081: |\boldsymbol t|+1$. We thus get for $m$ sufficiently large
1082: %
1083: \begin{eqnarray*}
1084: {\widetilde a}_{m,s} = \sum_{|\boldsymbol s|= s}
1085: p_{m-|\boldsymbol s|}^{(k+r)} \prod_{i=1}^r p_{s_i}
1086: \sim p_m^{(k+r)} \left( \sum_{|\boldsymbol s|= s}
1087: x_c^{|\boldsymbol s|}\prod_{i=1}^r p_{s_i}\right) \qquad (m\to\infty),
1088: \end{eqnarray*}
1089: %
1090: where the sum in brackets is finite.
1091: We now analyse the second term in the estimate derived from the
1092: Bernoulli inequality. To this end, define
1093: %
1094: \begin{displaymath}
1095: {\widetilde b}_{m,s}:= \sum_{|\boldsymbol s|= s}\sum_{t_i}
1096: \sum_{n\ge |\boldsymbol t|+1} |\boldsymbol t|n^{k+r-1}
1097: p_{m-|\boldsymbol s|,n}\prod_{i=1}^r p_{s_i,t_i}.
1098: \end{displaymath}
1099: %
1100: Using the estimate $|\boldsymbol t|\le |\boldsymbol s|^2$, we get
1101: %
1102: \begin{eqnarray*}
1103: {\widetilde b}_{m,s} \le s^2 \left(\sum_{|\boldsymbol s|= s}
1104: p_{m-|\boldsymbol s|}^{(k+r-1)} \prod_{i=1}^r p_{s_i}\right)
1105: \sim s^2 p_m^{(k+r-1)} \left( \sum_{|\boldsymbol s|= s}
1106: x_c^{|\boldsymbol s|}\prod_{i=1}^r p_{s_i}\right) \qquad (m\to\infty).
1107: \end{eqnarray*}
1108: %
1109: Now set $b_{m,s}:={\widetilde b}_{m,s}/(x_c^{-m}m^{\gamma_{k}^{(r)}-1})$.
1110: The above estimate yields $\lim_{m\to\infty} b_{m,s}=0$, since $0<\phi<1$.
1111:
1112: We now derive a lower bound for $p_m^{(r,k,s)}$ by subtracting
1113: from ${\widetilde p}_m^{(r,k,s)}$ an upper bound on the contributions
1114: arising from puncture-puncture interactions and from puncture-boundary
1115: interactions. We will show that the lower bound coincides asymptotically
1116: with the upper bound, which then implies the assertion of the theorem
1117: %
1118: \begin{displaymath}
1119: p_m^{(k,r,s)}\sim \frac{A_{k+r}}{r!}
1120: \left( \sum_{|\boldsymbol s|= s} x_c^{|\boldsymbol s|}
1121: \prod_{i=1}^r p_{s_i}\right) x_c^{-m}m^{\gamma_{k+r}-1}
1122: \qquad (m\to\infty).
1123: \end{displaymath}
1124: %
1125: For any polygon $\Pol$, the number of puncture-puncture interactions
1126: between $r>1$ punctures is smaller than the number of
1127: puncture-puncture interactions of two punctures with the remaining
1128: $r-2$ punctures occuring at arbitrary positions in the polygon.
1129: We thus get the upper bound
1130: %
1131: \begin{displaymath}
1132: (t_1+4s_1)t_2(n+|\boldsymbol t|)(n+|\boldsymbol t|)^{r-2}
1133: \le 6 t_1t_2 (n+|\boldsymbol t|)^{r-1},
1134: \end{displaymath}
1135: where we used $t_1\ge s_1-1$. The factor $(t_1+4s_1)t_2$
1136: bounds the number of configurations of two interacting punctures,
1137: and the factor $(n+|\boldsymbol t|)^{r-2}$ arises from allowing
1138: arbitrary positions of the remaining $r-2$ punctures. Define
1139: %
1140: \begin{displaymath}
1141: {\widetilde c}_{m,s}:= \sum_{|\boldsymbol s|= s}\sum_{t_i}
1142: \sum_{n} t_1t_2 n^k (n+|\boldsymbol t|)^{r-1}
1143: p_{m-|\boldsymbol s|,n+|\boldsymbol t|}\prod_{i=1}^r p_{s_i,t_i}
1144: \le s^4 \left(\sum_{|\boldsymbol s|= s}
1145: p_{m-|\boldsymbol s|}^{(k+r-1)}\prod_{i=1}^r p_{s_i}\right),
1146: \end{displaymath}
1147: %
1148: where we used $t_i\le|\boldsymbol t|\le|\boldsymbol s|^2$ for the
1149: last inequality. Setting $c_{m,s}:={\widetilde c}_{m,s}/
1150: (x_c^{-m}m^{\gamma_{k}^{(r)}-1})$, we infer that $\lim_{m\to\infty}
1151: c_{m,s}=0$. We have shown that for fixed $s$ the
1152: puncture-puncture interactions are asymptotically irrelevant.
1153:
1154: We finally estimate the puncture-boundary interactions. This is done
1155: similarly to the above treatment of puncture-puncture interactions.
1156: The number of puncture-boundary interactions is bounded from above by
1157: %
1158: \begin{displaymath}
1159: \sum_{j=1}^r4^j(m-|\boldsymbol s|)^j s_1\cdot\ldots\cdot s_j \,
1160: (n+|\boldsymbol t|)^{r-j},
1161: \end{displaymath}
1162: %
1163: where $j$ punctures interact with the boundary, each contributing
1164: a factor $4(m-|\boldsymbol s|)s_i$, and $r-j$ punctures have arbitrary
1165: positions, each contributing a factor $(n+|\boldsymbol t|)$. Note that
1166: the over-counted configurations in ${\widetilde p}_m^{(r,k,s)}$, which
1167: protrude from the boundary, are compensated for by the above estimate.
1168: Define
1169: %
1170: \begin{eqnarray*}
1171: {\widetilde d}_{m,s}&:= \sum_{|\boldsymbol s|= s}\sum_{t_i}
1172: \sum_{n} (m-|\boldsymbol s|)^j n^k (n+|\boldsymbol t|)^{r-j}s_1\cdot
1173: \ldots\cdot s_j\,p_{m-|\boldsymbol s|,n+|\boldsymbol t|}\prod_{i=1}^r
1174: p_{s_i,t_i}\\
1175: &\le (m-s)^j s^{j} \left(\sum_{|\boldsymbol s|= s}
1176: p_{m-|\boldsymbol s|}^{(k+r-j)}\prod_{i=1}^r p_{s_i}\right).
1177: \end{eqnarray*}
1178: %
1179: Defining $d_{m,s}:={\widetilde d}_{m,s}/
1180: (x_c^{-m}m^{\gamma_{k}^{(r)}-1})$, we infer that $\lim_{m\to\infty}
1181: d_{m,s}=0$. We have shown that for fixed $s$ the puncture-boundary
1182: interactions are asymptotically irrelevant. This completes the proof.
1183: \end{proof}
1184:
1185:
1186: \subsection{Polygons with $r$ punctures of arbitrary size
1187: \label{sec:scaling_arbitrary}}
1188:
1189: For a class of punctured self-avoiding polygons, consider
1190: for $k\in\mathbb N_0$ their area moment coefficients
1191: %
1192: \begin{displaymath}
1193: p_m^{(r,k)}:= \sum_n n^kp_{m,n}^{(r)},
1194: \end{displaymath}
1195: %
1196: where $p_{m,n}^{(r)}:=\sum_{s=0}^\infty p_{m,n}^{(r,s)}<\infty$ denotes
1197: the number of polygons in the class of half-perimeter $m$ and area $n$ with
1198: $r$ punctures of arbitrary size, $r\in\mathbb N_0$. For simplicity
1199: of notation, we write $p_{m,n}=p_{m,n}^{(0)}$ and $p_m^{(k)}=
1200: p_{m}^{(0,k)}$. In the sequel, we will use the area moment generating
1201: functions $\PGf_k(x)=\sum p_m^{(k)}x^m$ of the model without punctures.
1202:
1203: \begin{theorem}\label{theo:arb}
1204: Assume that, for a class of self-avoiding polygons without
1205: punctures, the area moment coefficients $p_m^{(k)}$
1206: have the asymptotic form, for $k\in\mathbb N_0$,
1207: %
1208: \begin{displaymath}
1209: p_m^{(k)} \sim A_k x_c^{-m} m^{\gamma_k-1} \qquad
1210: (m\to\infty)
1211: \end{displaymath}
1212: %
1213: for numbers $A_k>0$, $x_c>0$ and $\gamma_k=(k-\theta)/\phi$,
1214: where $0<\phi<1$. Let $\PGf_k(x)=\sum p_m^{(k)}x^m$ denote the $k^{th}$ area
1215: moment generating function.
1216:
1217: Then, the area moment coefficient $p_m^{(r,k)}$ of the polygon class with $r\ge1$
1218: punctures of arbitrary size is, for $k\in\mathbb N_0$, bounded from above by
1219: %
1220: \begin{displaymath}
1221: p_m^{(r,k)} \le \frac{[x^m]\PGf_{k+r}(x)(\PGf_0(x))^r}{r!}.
1222: \end{displaymath}
1223: %
1224: For finite critical perimeter generating functions, characterised by $\theta>0$,
1225: $p_m^{(r,k)}$ is asymptotically given by, for $k\in\mathbb N_0$,
1226: %
1227: \begin{equation}\label{form:t2}
1228: p_m^{(r,k)} \sim \frac{[x^m]\PGf_{k+r}(x)(\PGf_0(x))^r}{r!} \sim
1229: A_k^{(r)}x_c^{-m} m^{\gamma_{k+r}-1}
1230: \qquad (m\to\infty),
1231: \end{equation}
1232: %
1233: where the amplitudes $A_k^{(r)}$ are given by
1234: %
1235: \begin{equation}\label{eq:Aka}
1236: A_k^{(r)}=\frac{A_{k+r}(\PGf_0(x_c))^r}{r!},
1237: \end{equation}
1238: %
1239: where $\PGf_0(x_c):=\lim_{x\nearrow x_c}\PGf_0(x)<\infty$ is the critical amplitude of the
1240: half-perimeter generating function.
1241: \end{theorem}
1242:
1243:
1244: \noindent {\bf Remarks.} {\it i)}
1245: The asymptotic form Eq.~(\ref{form:t2}) is formally obtained from
1246: Theorem~\ref{theo:bounded} in the limit of infinite puncture size,
1247: see Remark {\it i)} after Theorem~\ref{theo:bounded}. This observation
1248: is also the main ingredient of the following proof, by noting that
1249: the upper bound has the same asymptotic behaviour.
1250: \\
1251: {\it ii)} For staircase polygons, where
1252: $\theta=1/3$ and $\phi=2/3$, the assumptions of Theorem
1253: \ref{theo:arb} are satisfied. For self-avoiding
1254: polygons, we have the numerically very well established
1255: values $\theta=1$ and $\phi=2/3$, which we believe to describe
1256: the asymptotic behaviour of SAPs.
1257: For models satisfying $\theta<0$, the upper bound generally does
1258: not coincide asymptotically with $p_m^{(r,k)}$. An example
1259: of failure is rectangles with a single puncture.
1260:
1261: \begin{proof}
1262: We obtain as in the proof of Theorem~\ref{theo:bounded} an upper bound
1263: ${\widetilde p}_m^{(r,k)}$ for the area moment
1264: coefficients $p_m^{(r,k)}$. It is given by
1265: %
1266: \begin{displaymath}
1267: p_m^{(r,k)}\le{\widetilde p}_m^{(r,k)} :=\frac{1}{r!}\sum_{s=0}^m
1268: \sum_{|{\bf s}|=s}p_{m-|{\bf s}|}^{(k+r)}\prod_{i=1}^r p_{s_i}
1269: =\frac{1}{r!}[x^m] \PGf_{k+r}(x) (\PGf_0(x))^r.
1270: \end{displaymath}
1271: %
1272: Assume in the following that $\theta>0$. The asymptotic behaviour
1273: of the rhs of (\ref{form:t2}) follows by $r$-fold application of
1274: Lemma 1, which is given in the appendix.
1275: Note that, for $M$ arbitrary, we have by definition
1276: %
1277: \begin{displaymath}
1278: p_m^{(r,k)}\ge\sum_{s=0}^M p_m^{(r,k,s)},
1279: \end{displaymath}
1280: %
1281: where $p_m^{(r,k,s)}$ is the number of $r$-punctured polygons,
1282: whose punctures have total perimeter equal to $s$.
1283: Theorem~\ref{theo:bounded} implies that the above sum is,
1284: for $M$ sufficiently large, asymptotically in $m$, arbitrarily close
1285: to the upper bound ${\widetilde p}_m^{(r,k)}$. See also the
1286: remark following Theorem~\ref{theo:bounded}. This
1287: yields the statement of the theorem.
1288: \end{proof}
1289:
1290: \subsection{Limit distribution of area and scaling
1291: function conjectures \label{sec:scaling_func}}
1292:
1293: We first discuss the implications of the previous
1294: results on the asymptotic area law of polygon models
1295: with punctures. By an application of Proposition
1296: \ref{theo:general}, Theorem \ref{theo:bounded} and
1297: Theorem \ref{theo:arb} immediately yield the following
1298: result:
1299:
1300: \begin{theorem}\label{thm:limdist}
1301: Assume that, for a class of self-avoiding polygons without
1302: punctures, the area moment coefficients $p_m^{(k)}$
1303: have the asymptotic form, for $k\in\mathbb N_0$,
1304: %
1305: \begin{displaymath}
1306: p_m^{(k)} \sim A_k x_c^{-m} m^{\gamma_k-1} \qquad
1307: (m\to\infty)
1308: \end{displaymath}
1309: %
1310: for numbers $A_k>0$, $x_c>0$ and $\gamma_k=(k-\theta)/\phi$, where $0<\phi<1$.
1311: Assume further that the numbers $A_k$ satisfy the Carleman condition
1312: %
1313: \begin{displaymath}
1314: \sum_{k\ge0} (A_{2k})^{-1/{2k}}=+\infty.
1315: \end{displaymath}
1316: %
1317: Denote the half-perimeter generating function of the model
1318: by ${\cal P}(x)=\left(\sum_{m\ge0} x^m\, p_m\right)$.
1319: \begin{itemize}
1320: \item[\it i)]
1321: Consider for $r\ge1$ the corresponding model with $r$
1322: punctures of bounded size, whose half-perimeter sum equals $s\in \mathbb N$, such
1323: that $[x^s]({\cal P}(x))^r\ne0$.
1324: Denote the random variables of area in the uniform fixed
1325: perimeter ensemble by $\widetilde X_m^{(r,s)}$. Then,
1326: we have convergence in distribution,
1327: %
1328: \begin{displaymath}
1329: \frac{\widetilde X_m^{(r,s)}}{m^{1/\phi}} \stackrel{d}{\longrightarrow} X^{(r,s)}
1330: \qquad (m\to\infty),
1331: \end{displaymath}
1332: %
1333: for a uniquely defined random variable $X^{(r,s)}$ with moments
1334: %
1335: \begin{displaymath}
1336: \mathbb E[(X^{(r,s)})^k]=\frac{A_k^{(r,s)}}{A_r},
1337: \end{displaymath}
1338: %
1339: where the numbers $A_k^{(r,s)}$ are those of Theorem~\ref{theo:bounded}.
1340: We also have moment convergence.
1341: \item[\it ii)]
1342: Let $\widetilde X_m^{(r)}$ denote the
1343: random variable of area in the fixed perimeter ensemble for
1344: the model with $r\ge1$ punctures of unbounded size. If $\theta>0$, then
1345: we have convergence in distribution,
1346: %
1347: \begin{displaymath}
1348: \frac{\widetilde X_m^{(r)}}{m^{1/\phi}} \stackrel{d}{\longrightarrow} X^{(r)} \qquad (m\to\infty)
1349: \end{displaymath}
1350: %
1351: for a uniquely defined random variable $X^{(r)}$ with moments
1352: %
1353: \begin{displaymath}
1354: \mathbb E[(X^{(r)})^k]=\frac{A_k^{(r)}}{A_r},
1355: \end{displaymath}
1356: %
1357: where the numbers $A_k^{(r)}$ are those of Theorem~\ref{theo:arb}.
1358: We also have moment convergence.
1359: \item[\it iii)]
1360: If $\theta>0$, the random variables $X^{(r)}$ and $X^{(r,s)}$ are related by
1361: %
1362: \begin{displaymath}
1363: x_c^s[x^s]({\cal P}(x))^r\, X^{(r)}=({\cal P}(x_c))^r\, X^{(r,s)},
1364: \end{displaymath}
1365: %
1366: where ${\cal P}(x)$ is the half-perimeter generating function
1367: of the polygon model without punctures, and where $\mathcal{P}(x_c)=
1368: \lim_{x\nearrow x_c}\mathcal{P}(x)<\infty$.
1369: %
1370: \end{itemize}
1371: \qed
1372: \end{theorem}
1373:
1374: \noindent {\bf Remarks.} {\it i)}
1375: For a given polygon model satisfying the assumptions of
1376: Theorem~\ref{thm:limdist}, the area moments satisfy asymptotically
1377: %
1378: \begin{displaymath}
1379: \frac{\mathbb E[(\widetilde X_m^{(r)})^k]}{k!}\sim D_k^{(r)} m^{k/\phi}
1380: \qquad (m\to\infty),
1381: \end{displaymath}
1382: %
1383: for positive numbers $ D_k^{(r)}$.
1384: For classes of polygon models whose generating function satisfies a
1385: $q$-functional equation with a square root as the dominant singularity
1386: of their perimeter generating function, the amplitude ratios $D_k^{(r)}/\left
1387: [ D_1^{(r)} \right ]^k$ are universal, i.e., independent of the constants
1388: $f_0$, $f_1$ and $x_c$, which characterise the underlying model \cite{R02,R05,R06}. This
1389: follows from Eqs.~(\ref{eq:relation}) and (\ref{ck}) by a straightforward calculation. The
1390: numbers are listed in Table~\ref{tab:ratios} for small values of $r$. Note that
1391: the same numbers appear for punctures of bounded size.
1392: %
1393: \begin{table}
1394: \begin{center}
1395: \begin{tabular}{cccc}
1396: \hline \hline
1397: Amplitude & $r=0$ & $r=1$ & $r=2$ \\
1398: \hline
1399: $D_2/D_1^2$ &
1400: 0.530518$\times 10^{-0}$ &
1401: 0.530143$\times 10^{-0}$ &
1402: 0.529356$\times 10^{-0}$
1403: \\
1404: $D_3/D_1^3$ &
1405: 0.198944$\times 10^{-0}$ &
1406: 0.198369$\times 10^{-0}$ &
1407: 0.197361$\times 10^{-0}$
1408: \\
1409: $D_4/D_1^4$ &
1410: 0.592379$\times 10^{-1}$&
1411: 0.588127$\times 10^{-1}$&
1412: 0.581533$\times 10^{-1}$
1413: \\
1414: $D_5/D_1^5$ &
1415: 0.149079$\times 10^{-1}$ &
1416: 0.146994$\times 10^{-1}$ &
1417: 0.144042$\times 10^{-1}$
1418: \\
1419: $D_6/D_1^6$ &
1420: 0.329453$\times 10^{-2}$&
1421: 0.321705$\times 10^{-2}$&
1422: 0.311511$\times 10^{-2}$
1423: \\
1424: $D_7/D_1^7$ &
1425: 0.655743$\times 10^{-3}$ &
1426: 0.632288$\times 10^{-3}$ &
1427: 0.603260$\times 10^{-3}$
1428: \\
1429: $D_8/D_1^8$ &
1430: 0.119654$\times 10^{-3}$&
1431: 0.113600$\times 10^{-3}$&
1432: 0.106501$\times 10^{-3}$
1433: \\
1434: $D_9/D_1^9$ &
1435: 0.202754$\times 10^{-4}$ &
1436: 0.189015$\times 10^{-4}$ &
1437: 0.173673$\times 10^{-4}$
1438: \\
1439: $D_{10}/D_1^{10}$ &
1440: 0.322150$\times 10^{-5}$ &
1441: 0.294132$\times 10^{-5}$ &
1442: 0.264251$\times 10^{-5}$
1443: \\
1444: \hline \hline
1445: \end{tabular}
1446: \end{center}
1447: \caption{\label{tab:ratios}
1448: Universal amplitude ratios for staircase polygons with $r$ punctures.}
1449: \end{table}
1450: %
1451: \\
1452: {\it ii)}
1453: For the above class of models, explicit expressions for the
1454: asymptotic behaviour of their moment generating functions and
1455: their probability distributions can be derived from the area
1456: amplitude series via inverse Laplace transform techniques.
1457: Since the resulting expressions are quite cumbersome, we do not
1458: give them here. For $r=0$ the corresponding limit distribution
1459: of area is the Airy distribution \cite{L84,T91,FL01}. The
1460: extension to $r\ge1$ is straightforward.
1461: As mentioned above, for $r=0$ the amplitude ratios are found
1462: to coincide with those of (rooted) self-avoiding polygons to a
1463: high degree of numerical accuracy \cite{RJG03}. If the conjecture
1464: holds true that they agree {\it exactly}, then the above expressions
1465: for limit distributions also appear for rooted self-avoiding polygons,
1466: for all values of $r$. See Section 6 for a detailed numerical
1467: analysis.
1468:
1469: \medskip
1470:
1471: We now discuss the relations between the area amplitude series
1472: $F(z)$ of the polygon model without punctures and $F^{(r)}(z)$ of
1473: the polygon model with $r$ punctures. Since all of our models have an
1474: entire moment generating function, the Carleman condition is satisfied,
1475: and Theorem \ref{theo:bounded} and Theorem \ref{theo:arb} yield, by a
1476: straightforward calculation, the following result.
1477:
1478: \begin{theorem}\label{thm:amplfct}
1479: Assume that, for a class of self-avoiding polygons, the polygon model without punctures
1480: has an area amplitude series, given by
1481: %
1482: \begin{displaymath}
1483: F(z)=\sum_{k\ge0} f_k z^{-\gamma_k},
1484: \end{displaymath}
1485: %
1486: where $\gamma_k=(k-\theta)/\phi\in(-1,\infty)\setminus\{0\}$ and $0<\phi<1$, and where
1487: the numbers $f_k\ne0$ are related to the amplitudes $A_k$ in Proposition
1488: \ref{theo:general} via Eq.~(\ref{eq:relation}). Denote the half-perimeter
1489: generating function of the model
1490: by ${\cal P}(x)=\sum_{m\ge0} x^m\, p_m$.
1491: %
1492: \begin{itemize}
1493: \item[\it i)]
1494: Assume that $r\ge1$ and $s\in\mathbb N$ are given such that
1495: $[x^s]({\cal P}(x))^r\ne0$.
1496: Then, the corresponding model of punctured polygons with $r$ punctures
1497: of bounded size $s$ has an area amplitude series, given by
1498: %
1499: \begin{displaymath}
1500: F^{(r,s)}(z)=\sum_{k\ge0} f_k^{(r)} z^{-\gamma_k^{(r)}},
1501: \end{displaymath}
1502: %
1503: where $\gamma_k^{(r)}=(k-\theta_r)/\phi_r$. We have
1504: $\theta_r=\theta-r$, $\phi_r=\phi$, and
1505: %
1506: \begin{equation}\label{form:Fr}
1507: F^{(r,s)}(z)=\frac{(-1)^r}{r!}x_c^s[x^s]({\cal P}(x))^r
1508: \sum_{k\ge r} (k)_r f_k z^{-\gamma_k}.
1509: \end{equation}
1510: %
1511: \item[\it ii)]
1512: If $\theta>0$, the corresponding model of punctured polygons with $r\ge1$ punctures
1513: of arbitrary size has an area amplitude series, given by
1514: %
1515: \begin{displaymath}
1516: F^{(r)}(z)=\frac{(-1)^r}{r!}({\cal P}(x_c))^r
1517: \sum_{k\ge r} (k)_r f_k z^{-\gamma_k},
1518: \end{displaymath}
1519: %
1520: where $\mathcal{P}(x_c)=\lim_{x\nearrow x_c}\mathcal{P}(x)<\infty$.
1521: \end{itemize}
1522: %
1523: \qed
1524: \end{theorem}
1525:
1526: \noindent {\bf Remarks.} {\it i)}
1527: Eq.~(\ref{form:Fr}) allows one to derive explicit expressions for the
1528: area amplitude series in terms of $F(z)$. For models where $\theta=1/3$
1529: and $\phi=2/3$ such as staircase polygons,
1530: the area amplitude series $F(z)$ satisfies the Riccati equation
1531: %
1532: \begin{equation}\label{eq:Ricc}
1533: F(z)^2-4f_1F'(z)-f_0^2z=0.
1534: \end{equation}
1535: %
1536: This can be used to show that $F^{(r,s)}(z)$
1537: (and also $F^{(r)}(z)$) is of the form
1538: %
1539: \begin{displaymath}
1540: F^{(r,s)}(z)=\sum_{k=0}^{r+1} p_{k,r}(z)F(z)^{k},
1541: \end{displaymath}
1542: %
1543: where $p_{k,r}(z)$ are polynomials in $z$ of degree not exceeding
1544: $\lceil 3r/2\rceil$, and $p_{r+1,r}(z)$ is not identically vanishing.
1545: Simple expressions for the polynomials
1546: $p_{k,r}(z)$ are not apparent. We note, however, that such
1547: expressions appear as correction-to-scaling functions
1548: of the underlying polygon models without punctures \cite{R02}.\\
1549: {\it ii)}
1550: The model of rooted self-avoiding polygons has been found numerically
1551: to have the same type of area amplitude series as staircase polygons
1552: (with different constants $f_0$ and $f_1$). Similar considerations apply
1553: to the model of unrooted self-avoiding polygons
1554: starting from Eq.~(\ref{form:unroot}).
1555:
1556: \medskip
1557:
1558: We finally discuss the scaling function conjectures implied
1559: by the results of the previous subsections. For staircase polygons,
1560: the area amplitude
1561: function satisfies the differential equation~(\ref{eq:Ricc}).
1562: This differential equation has a unique solution ${\cal F}(z)$
1563: analytic for $\Re(z)\ge0$, having $F(z)$ as an asymptotic
1564: expansion at infinity, ${\cal F}(z)\sim F(z)$ as $z\to\infty$.
1565: The function ${\cal F}(z)$ is explicitly given by
1566: %
1567: \begin{displaymath}
1568: {\cal F}(z)=-4f_1 \frac{{\rm d}}{{\rm d}z}\log\mbox{Ai}
1569: \left(\left(\frac{f_0}{4f_1}\right)^{2/3}z\right),
1570: \end{displaymath}
1571: %
1572: and this function coincides with the scaling function of
1573: staircase polygons Eq.~(\ref{sfstair}).
1574:
1575: In analogy to the above observation, we conjecture that
1576: the area amplitude series for punctured staircase
1577: polygons determine their scaling functions. Likewise, we conjecture that
1578: the area amplitude series for punctured rooted self-avoiding
1579: polygons determine their scaling functions.
1580:
1581: \begin{conjecture}\label{con1}
1582: Let $r\ge1$ and $s\ge2$ be given. For staircase polygons and rooted self-avoiding polygons,
1583: the area amplitude series $F^{(r,s)}(z)$ and $F^{(r)}(z)$ of
1584: Theorem~\ref{thm:amplfct} uniquely define functions ${\cal F}^{(r,s)}(z)$
1585: and ${\cal F}^{(r)}(z)$ analytic for $\Re(z)> z_0$ and non-analytic at $z=z_0$,
1586: for some negative real number $z_0<0$. We conjecture that the functions
1587: ${\cal F}^{(r,s)}(z):(z_0,\infty)\to\mathbb R$ and
1588: ${\cal F}^{(r)}(z):(z_0,\infty)\to\mathbb R$ are scaling functions
1589: as in Definition~\ref{def:sf},
1590: %
1591: \begin{eqnarray*}
1592: {\cal P}^{(r,s)}(x,q) &\sim (1-q)^{1/3-r}{\cal F}^{(r,s)}
1593: \left(\frac{x_c-x}{(1-q)^{2/3}}\right)\\
1594: {\cal P}^{(r)}(x,q) &\sim (1-q)^{1/3-r}{\cal F}^{(r)}
1595: \left(\frac{x_c-x}{(1-q)^{2/3}}\right).
1596: \end{eqnarray*}
1597: %
1598: \end{conjecture}
1599:
1600: \medskip
1601:
1602: \noindent {\bf Remark.}
1603: The above conjecture has the following implications.\\
1604: {\it i)} Staircase polygons with a single minimal
1605: puncture specialise to Eq.~(\ref{stsf}).\\
1606: {\it ii)} Up to constant factors, the scaling
1607: form of the model with $r$ punctures is obtained as the $r^{th}$
1608: derivative w.r.t.~$q$ of the scaling form of the model without
1609: punctures, as can be proved by induction. As derivatives
1610: can be interpreted combinatorially as marking, this reflects the fact
1611: underlying the proofs in this section that punctures may be regarded
1612: as being asymptotically independent, and boundary effects do not
1613: play a role asymptotically.\\
1614: {\it iii)}
1615: Ignoring questions of analyticity, a (formal) calculation yields
1616: that the functions ${\cal F}^{(r)}$ (and ${\cal F}^{(r,s)}$) lead,
1617: for both staircase polygons and (unrooted) self-avoiding polygons,
1618: to the same critical exponents in the branched polymer phase as
1619: those conjectured previously \cite{JvR92,GJWE00}. These are obtained from
1620: the singular behaviour of ${\cal F}^{(r)}$ about the singularity of
1621: smallest modulus on the negative axis, i.e., at the first zero of
1622: the Airy-function on the negative axis, see \cite[Sec~1]{R02}. The
1623: fact that $\PGf^{(r)}(x,q)$ is obtained from $\PGf(x,q)$ by $r$-fold
1624: differentiation yields the result.
1625:
1626:
1627: \section{Computer enumerations \label{sec:enum}}
1628:
1629: Here we briefly outline which algorithms were used to derive the series expansions
1630: for the area moments of punctured polygons. In most cases (SAPs and punctured
1631: staircase polygons) the algorithms are simple generalisations of previous
1632: algorithms already described in detail in other papers, referenced below. In
1633: these cases we give brief details of the length of the series and the amount
1634: of CPU time used. Only in the case of staircase polygons with minimal punctures
1635: did we write a new specific algorithm which we shall describe in some detail.
1636:
1637: The series for punctured self-avoiding polygons were calculated using a simple
1638: generalisation of the parallel version of the algorithm we used previously
1639: to enumerate ordinary SAPs \cite{IJ03}. In each case (SAPs with one or two
1640: minimal punctures and SAPs with one or two arbitrary punctures) we calculated the
1641: area moments up to $k=10$ for SAPs to total perimeter 100. Since the smallest such
1642: SAPs have perimeter 16 and 24 this results in series with 42 and 38 non-zero
1643: terms, respectively. The total CPU time required was about 5000 hours for
1644: each of the once punctured SAP problems and up to 3000 hours for the twice
1645: punctured problems. The bulk of these calculations were performed on the old facility
1646: of the Australian Partnership for Advanced Computing (APAC), which
1647: was a Compaq Server Cluster with ES45 nodes with 1GHz Alpha chips (this facility
1648: has since been replaced by an SGI Altix cluster).
1649:
1650: In \cite{GJ06b} we used a very efficient algorithm to enumerate once punctured
1651: staircase polygons. The algorithm is easily generalised in order to calculate
1652: area moments which we have done to perimeter 718 ($k=1$), 598 ($k=2$) and
1653: 506 ($k=3$ to 10). It is also quite straight-forward to generalise the algorithm
1654: to count twice punctured staircase polygons and in this case we obtained the
1655: series to perimeter 502 for $k=0$, perimeter 450 for $k=1$ and 2 and to perimeter
1656: 302 for $k=3$ to 10. It is also easy to extract data for staircase polygons with
1657: punctures of fixed combined perimeter.
1658: In each case we used around 1000 CPU hours on the APAC Altix cluster which
1659: use 1.6GHz Intel Itanium 2 chips.
1660:
1661:
1662: \begin{figure}
1663: \begin{center}
1664: \includegraphics[scale=0.6]{transfer}
1665: \end{center}
1666: \caption{\label{fig:TM}
1667: Illustration of the transfer matrix boundary line and local updating rules.
1668: }
1669: \end{figure}
1670:
1671: Finally we describe the algorithm used to enumerate minimally punctured staircase
1672: polygons. The algorithm is based on so-called transfer matrix techniques.
1673: The basic idea is to count the number of polygons by bisecting the lattice with
1674: a boundary line. In the left panel of Fig.~\ref{fig:TM} we show how such a boundary
1675: (the first medium thick line) will intersect the polygon in several places.
1676: The first and last occupied edges intersected by the boundary line are the
1677: directed walks constituting the outer staircase polygon. The other occupied
1678: edges (if any) belong to the minimal punctures. In a calculation to maximal
1679: half-perimeter $m$ we need only consider intersections with widths up to $w=m/2$.
1680: Any intersection pattern (or signature) can be specified by a string of occupation
1681: variables, $S = \{\sigma_0,\sigma_1,\ldots \sigma_w \}$, where $\sigma_i=0$, 1 or 2
1682: if edge number $i$ is empty, an occupied outer edge or an occupied edge
1683: part of a minimal puncture, respectively. We could use the same symbol for all
1684: occupied edges but it is convenient to explicitly distinguish between the two cases.
1685: For each signature we keep a generating function which keeps track of the number
1686: of configurations (to the left of the boundary line), that is, it counts the number
1687: of possible partially completed polygons with a given signature. In order
1688: to count the total number of punctured polygons we move the boundary line
1689: to the right column by column with each column built up one vertex at a time.
1690: In the left panel of Fig.~\ref{fig:TM} we have also shown a typical move of the boundary
1691: line, which starts in the position given by the second medium thick line and where
1692: we add two new edges to the lattice by moving the kink in the boundary line to
1693: the position given by the thin lines. As we move the boundary line to a
1694: new position we calculate the associated generating functions (the updating rules will
1695: be given below). Formally we can view this transformation between signatures as a matrix
1696: multiplication (hence our use of the nomenclature {\em transfer matrix algorithm}).
1697: However, as can be readily seen, the transfer matrix is extremely sparse and there
1698: is no reason to list it explicitly (it is given {\em implicitly} by the updating rules).
1699:
1700: We start the calculation with the initial signature $\{1,1,0, \ldots ,0\}$, which
1701: corresponds to inserting the two steps of the outer walks in the lower left corner
1702: (the count of this configuration is 1). As the boundary line is moved it passes over
1703: a vertex and the updating depends on the states of the edges below and to the left of
1704: this vertex. After the move we `insert' the edges to the right and above the vertex.
1705: There are four possible local configurations of the `incoming' edges as illustrated
1706: in the middle panels of Fig.~\ref{fig:TM}: Both edges are empty, one of the edges is
1707: occupied and the other edge is empty or both edges are occupied.
1708:
1709:
1710: \begin{description}
1711: \item{\bf Both edges empty:} If both incoming edges are empty then both outgoing edges
1712: can be empty. Else we may insert two new steps which {\em must} be part of a minimal puncture
1713: (the outgoing edges are in state `2'). This is {\em only} possible if the vertex is in the
1714: interior of the polygon (there is an edge in state `1' both below and above the vertex).
1715: \item{\bf Left edge empty, bottom edge occupied:} The walk occupying the incoming edge must
1716: be continued along an outgoing edge. If the occupied edge is part of the external
1717: polygon (in state `1') there are no restrictions. If the occupied edge is part of
1718: a minimal puncture the walk can only be continued along the edge to the right of the
1719: vertex (otherwise we would not get a minimal puncture).
1720: \item{\bf Left edge occupied, bottom edge empty:} This is similar to the previous
1721: case except that an edge in state `2' must be continued along the edge above the vertex.
1722: \item{\bf Both edges occupied:} If both edges are in state `2' we close the puncture
1723: and the new edges are empty. If the incoming edges are in state `1' we have closed
1724: the outer polygon and then we add the count to the running total for the generating
1725: function.
1726: \end{description}
1727: In the last panel of Fig.~\ref{fig:TM} we show how the updating rules given above
1728: through a sequence of moves of the boundary line gives rise to a minimal puncture.
1729:
1730: The perimeter of the completed polygon is given by the position of the boundary
1731: line when the polygon is closed, e.g., if we have taken $k$ steps in the
1732: $x$-direction (completed $k$ columns) and moved the kink $l$ steps in the $y$-direction
1733: then the outer half-perimeter is $k+l$ and the total half-perimeter in $k+l+2r$.
1734: So we need only keep track of the number of punctures $r$. This is done
1735: by associating a truncated polynomial $P_S(x)=\sum p_r x^r$ with each signature, where
1736: the coefficient $p_r$ is the number of partially completed polygons with $r$
1737: punctures of the signature $S$. As a new signature $S'$ is created from $S$
1738: we set $P_{S'}(x)=P_{S'}(x)+x^{\delta}P_S(x)$, where $\delta =1$ if an additional
1739: puncture is inserted (as in the first case described above) or 0 otherwise.
1740: The extension to the calculation of area moments is described in \cite{IJ00}.
1741:
1742: \section{Numerical analysis \label{sec:numeric}}
1743:
1744: We now turn to our numerical analysis of the series for punctured polygons.
1745: In section~\ref{sec:stair_min} we use our series to determine numerically the
1746: {\em exact} area moment generating functions for minimally punctured
1747: staircase polygons with up to 5 punctures and $k\leq 10$. The resulting
1748: exact expressions for the leading amplitudes are in complete agreement with
1749: the formula derived in Proposition~\ref{prop:minpunc}. In section \ref{sec:stair_fix} we
1750: extend the study to staircase polygons with one puncture of fixed size and two punctures
1751: of fixed combined size. Again we find the exact generating functions and
1752: confirm the formula for the amplitude given in Theorem~\ref{theo:bounded}.
1753: Next, in section~\ref{sec:stair_one}, we analyse area moments for staircase polygons with
1754: a single puncture of arbitrary size. Guided by results obtained from an analysis of the
1755: conjectured exact ODE \cite{GJ06b} satisfied by the perimeter generating function
1756: we carry out a careful numerical analysis of the area moment coefficients.
1757: This allows to obtain accurate estimates for the leading amplitudes
1758: and we confirm the results of Theorem~\ref{theo:arb} to at least 15 significant digits.
1759: Then, in section~\ref{sec:stair_two} we extend our study to staircase polygons
1760: with two staircase punctures of arbitrary size and we again find good agreement
1761: with the exact results. Intriguingly, we find in all of the above cases, that
1762: the amplitude of the leading correction term is a constant times the
1763: corresponding amplitude with one less puncture. Finally, in section~\ref{sec:sapana}
1764: we present the results of our analysis of self-avoiding polygons with one and two punctures
1765: (minimal as well as arbitrary). In this case the numerical evidence is not
1766: quite as convincing, but we do find that the numerical estimates agree with the exact formulas
1767: to at least 3--4 significant digits.
1768:
1769:
1770: \subsection{Staircase polygons with minimal punctures \label{sec:stair_min}}
1771:
1772: In \cite{GJWE00} it was found that the half-perimeter generating function of staircase
1773: polygons with a single minimal puncture is:
1774:
1775: \begin{equation}\label{eq:St1p}
1776: \StM{1}(x) = \frac{1-8x+20x^2-16x^3+2x^4}{2(1-4x)}- \frac{1-6x+10x^2-4x^3}{2\sqrt{1-4x}}.
1777: \end{equation}
1778: This result is also derivable from Eq.~(\ref{sp1}), which gives a functional equation
1779: for the area-perimeter generating function.
1780: It is thus plausible to expect that the generating function of staircase polygons
1781: with $r$ minimal punctures is of a similar form
1782:
1783:
1784: \begin{equation}\label{eq:Stmin}
1785: \StM{r}(x) = \frac{A_r(x)+B_r(x)\sqrt{1-4x}}{(1-4x)^{\gamma_r}},
1786: \end{equation}
1787: where $A_r(x)$ and $B_r(x)$ are polynomials and $\gamma_r = (3r-1)/2$. We find this to be correct
1788: for all the cases we have enumerated that is up to $r=5$:
1789:
1790:
1791: \begin{eqnarray*}
1792: 2A_2(x) &=& {x}^{2}-26{x}^{3}+228{x}^{4}-906{x}^{5}+1709{x}^{6}-1378{x}^{7}+322{x}^{8},\\
1793: 2B_2(x) &=& -{x}^{2}+24\,{x}^{3}-182\,{x}^{4}+586\,{x}^{5}-815\,{x}^{6}+404\,{x}^{7}-32\,{x}^{8}.\\
1794: A_3(x) &=& {x}^{2}-22\,{x}^{3}+197\,{x}^{4}-924\,{x}^{5}+2545\,{x}^{6}-5374\,{x}^{7}+13828\,{x}^{8} \\
1795: &&\!-33634\,{x}^{9}\!+46027\,{x}^{10}\!-24746\,{x}^{11}\!+612\,{x}^{12}\!+256\,{x}^{13}\!+256\,{x}^{14},\\
1796: B_3(x) &=& -{x}^{2}+20\,{x}^{3}-159\,{x}^{4}+642\,{x}^{5}-1509\,{x}^{6}+3176\,{x}^{7}-9040\,{x}^{8}\\
1797: &&+19254\,{x}^{9}-18943\,{x}^{10}+4968\,{x}^{11}+768\,{x}^{12}+256\,{x}^{13}.\\
1798: 2A_4(x) &=& 2\,{x}^{2}-60\,{x}^{3}+809\,{x}^{4}-6564\,{x}^{5}+36321\,{x}^{6}-146436\,{x}^{7}\\
1799: &&+439283\,{x}^{8}-960070\,{x}^{9}+1485167\,{x}^{10}-1823356\,{x}^{11}\\
1800: &&+2728708\,{x}^{12}-4441406\,{x}^{13}+4054296\,{x}^{14}-932228\,{x}^{15}\\
1801: &&-298318\,{x}^{16}-143360\,{x}^{17}+16384\,{x}^{18}-32768\,{x}^{19},\\
1802: 2B_4(x) &=& -2\,{x}^{2}+56\,{x}^{3}-701\,{x}^{4}+5266\,{x}^{5}-26987\,{x}^{6}+100694\,{x}^{7}\\
1803: &&-276415\,{x}^{8}+537888\,{x}^{9}-727683\,{x}^{10}+889018\,{x}^{11}\\
1804: &&-1536634\,{x}^{12}+2199158\,{x}^{13}-1289388\,{x}^{14}-47472\,{x}^{15}\\
1805: &&+26880\,{x}^{16}+50176\,{x}^{17}-6144\,{x}^{18}+8192\,{x}^{19}.\\
1806: 2A_5(x) &=& 2\,{x}^{2}-76\,{x}^{3}+1343\,{x}^{4}-14776\,{x}^{5}+114384\,{x}^{6}-666240\,{x}^{7}\\
1807: &&+3036602\,{x}^{8}-11071408\,{x}^{9}+32642310\,{x}^{10}-77911156\,{x}^{11}\\
1808: &&+148630330\,{x}^{12}-220310536\,{x}^{13}+250700412\,{x}^{14}\\
1809: &&-250317844\,{x}^{15}+290657417\,{x}^{16}-309183568\,{x}^{17}\\
1810: &&+150313538\,{x}^{18}+21743832\,{x}^{19}-15222464\,{x}^{20}+449152\,{x}^{21}\\
1811: &&-3828224\,{x}^{22}-2844672\,{x}^{23}+974848\,{x}^{24}-819200\,{x}^{25},\\
1812: 2B_5(x) &=& -2\,{x}^{2}+72\,{x}^{3}-1203\,{x}^{4}+12506\,{x}^{5}-91510\,{x}^{6}+504084\,{x}^{7}\\
1813: &&-2171612\,{x}^{8}+7467208\,{x}^{9}-20683474\,{x}^{10}+46059704\,{x}^{11}\\
1814: &&-80841764\,{x}^{12}+107986392\,{x}^{13}-111525400\,{x}^{14}\\
1815: &&+114888220\,{x}^{15}-142562573\,{x}^{16}+122527230\,{x}^{17}\\
1816: &&-24478856\,{x}^{18}-17117496\,{x}^{19}-533632\,{x}^{20}-2988544\,{x}^{21}\\
1817: &&-808960\,{x}^{22}+401408\,{x}^{23}-819200\,{x}^{24}.
1818: \end{eqnarray*}
1819:
1820: Likewise the generating functions for
1821: the $k$'th area moment, $\StMM{r}{k}(x)$, is also of the form (\ref{eq:Stmin})
1822: \begin{equation}\label{eq:Stminmom}
1823: \StMM{r}{k}(x) = \frac{A_{r,k}(x)+B_{r,k}(x)\sqrt{1-4x}}{(1-4x)^{\gamma_{r+k}}}.
1824: \end{equation}
1825: We find that the degrees of the polynomials $A_{r,k}(x)$ and $B_{r,k}(x)$ are less than $5r+2k$
1826: for $r\leq5$ and $k\leq 10$. In particular we have:
1827:
1828:
1829: \begin{eqnarray}\label{eq:St1p1m}
1830: \StMM{1}{1}(x) =& \frac{1-14x+72x^2-162x^3+145x^4-34x^5+2x^6}{(1-4x)^{5/2}} \nonumber \\
1831: &-\frac{1-12x+50x^2-82x^3+43x^4-4x^5}{(1-4x)^2}.
1832: \end{eqnarray}
1833:
1834: From these solutions we then calculate the exact leading amplitudes
1835: and indeed we find that
1836: \begin{equation}
1837: \La{r}{k}=A_{k,r}(x_c)/\Gamma (\gamma_{r+k})
1838: =\frac{(-1)^{k+r}(k+r)! x_c^{2r}f_{k+r}}{r! x_c^{\gamma_{k+r}}\Gamma(\gamma_{k+r})}.
1839: \end{equation}
1840: in complete agreement with Eq.~(\ref{eq:relation}).
1841:
1842: We have also looked at the amplitudes $\Lb{r}{k}=B_{k,r}(x_c)/\Gamma (\gamma_{r+k}-\frac12)$
1843: of the correction terms and find, quite remarkably, that they are given simply in terms
1844: of the leading amplitudes with one less puncture:
1845:
1846: \begin{equation}\label{eq:StMB}
1847: \Lb{r}{k}=-\frac18 \La{r-1}{k}.
1848: \end{equation}
1849:
1850:
1851: \subsection{Staircase polygons with staircase punctures of fixed size \label{sec:stair_fix}}
1852:
1853: Next we examine the more general case of staircase polygons with punctures of fixed size. In the
1854: case of one puncture we thus look at staircase polygons with a staircase hole of half-perimeter
1855: $s$ while in the case of two punctures we look at staircase polygons with two staircase holes
1856: whose half-perimeters sum to $s$. As in the previous section we expect the generating
1857: functions to be of the form
1858: \begin{equation}\label{eq:Stfix}
1859: \PPM{r,s}{k}(x) = \frac{A_{r,s,k}(x)+B_{r,s,k}(x)\sqrt{1-4x}}{(1-4x)^{\gamma_{r+k}}}.
1860: \end{equation}
1861: We find this to be true with the degree of the polynomials less than $5r+2(k+s)$.
1862:
1863: For once punctured polygons we calculated the generating functions for $s \leq 25$ and $k \leq 10$.
1864: In Theorem~\ref{theo:bounded} we proved that the leading amplitude
1865: $\La{1,s}{k}= A_{1,s,k}(x_c)/\Gamma(\gamma_{k+1})= A_{k+1}x_c^s p_s$,
1866: where $p_s$ is the number of staircase polygons of half-perimeter $s$. This formula is naturally
1867: confirmed by our numerical results. Of more interest is the sub-leading amplitude $\Lb{1,s}{k}$.
1868: We find that the result for minimally punctured polygons generalises to this case and
1869: $\Lb{1,s}{k} = -b_{1,s} A_k$. We also find that the integer sequence
1870: $d_s=8^{s-1}b_{1,s}=1,5,29,182,\ldots$ is given by the recurrence:
1871: $$8s^2d_s+(s+3)(7s+10)d_{s+1}-(s+3)(s+2)d_{s+2}=0, d_1=0, d_2=1.$$
1872: We note that $d_s$ grows like $8^s$ so that $b_{1,s}$ grows no faster than a polynomial in $s$.
1873:
1874: For twice punctured polygons we calculated the generating functions for $s \leq 10$ and $k \leq 10$.
1875: As a consequence of theorem~\ref{theo:bounded} the leading amplitude is given by
1876: $\La{2,s}{k}= A_{2,s,k}(x_c)/\Gamma(\gamma_{k+2})= A_{k+2}x_c^s \sum_{t=2}^{s-2}p_{s-t}p_t$.
1877: And again we find that $\Lb{2,s}{k} = -b_{2,s} A_{k+1}$, though in this case we haven't
1878: found a recurrence for the integer sequence $2\times 8^{s-2}b_{2,s}=1,9,69,510, \ldots$.
1879:
1880:
1881: \subsection{Staircase polygons with a single staircase puncture \label{sec:stair_one}}
1882:
1883: We now turn to the analysis of the area moments of staircase polygons with a single staircase
1884: puncture of arbitrary size (1-punctured staircase polygons for short).
1885: In a recent paper \cite{GJ06b} we reported on work which led to an exact Fuchsian linear
1886: differential equation of order 8 apparently satisfied by the half-perimeter generating function,
1887: $\PP{1}(x) = \sum_{m\geq 0} p_m^{(1)} x^m$, for 1-punctured staircase polygons
1888: (that is $\PP{1} (x)$ is one of the solutions of the ODE, expanded around the origin).
1889: Our analysis of the ODE showed that the dominant singular behaviour is
1890: \begin{equation}\label{eq:S1}
1891: \PP{1}(x) \sim \frac{A(x)}{(1-4x)} + \frac{B(x) + C(x) \log(1-4x)}{\sqrt{1-4x}}+D(x) (1+4x)^{13/2}.
1892: \end{equation}
1893: The functions $A(x)$--$D(x)$ are regular in the disc $|x| \le 1/4$.
1894: In addition there were a pair of singularities on the imaginary axis at $x=\pm i/2$,
1895: and at the roots of $1+x+7x^2$. Note that the absolute value of these singularities exceeds
1896: $1/4$ and so their contributions to the asymptotic growth of the series coefficients are
1897: exponentially suppressed.
1898:
1899:
1900: We expect that the area moment generating functions, $\PPM{1}{k}(x)$, should have a similar
1901: critical behaviour to that of (\ref{eq:S1}). Indeed our analysis using
1902: differential approximants \cite{AJG89a} revealed that at $x=x_c=1/4$ there
1903: is a triple root with exponents $-\gamma_{k+1}$ and $-\gamma_{k+1}\!+\!1/2$ (twice)
1904: which is indicative that the behaviour is
1905: \begin{equation}\label{eq:S1mom}
1906: \PPM{1}{k}(x) \sim \frac{A(x)+[B(x) + C(x) \log(1-4x)]\sqrt{1-4x}}{(1-4x)^{3k/2+1}}.
1907: \end{equation}
1908: However, the behaviour at the singularity $x=x_-=-1/4$ is a little more complicated.
1909: For the first area moment we find that there is a double root with exponents $5$ and
1910: $13/2$, while for the second area moment we find a triple root at $x_-$ with
1911: exponents $7/2$, $5$ and $13/2$. For higher moments the behaviour is consistent
1912: with a triple root with exponents $(13-3k)/2$, $(10-3k)/2$ and $(7-3k)/2$. That is, the
1913: value of the leading exponent decreases by $3k/2$ and there is in addition a non-analytic
1914: correction to scaling with exponent $3/2$. The upshot of this analysis is that
1915: the asymptotic behaviour of the coefficients of $\PPM{1}{k}(x)$ should be given by
1916: \begin{equation}\label{eq:S1co}
1917: [x^m]\PPM{1}{k}(x)\! \sim\! 4^m\! \left ( \! \sum_{j=0} \!\!
1918: m^{3k/2-j}\! \left (\! a_j\!+\!\frac{1}{\sqrt{m}}\left [ b_j \!+\!c_j \log (m) \right ]\! \right )
1919: \!\!+\!(-1)^m \! \sum_{j=0}\! d_j m^{(-15+3k-j)/2}\! \right )\!,
1920: \end{equation}
1921: where we have ignored the contributions from singularities in the complex plane with absolute values
1922: exceeding $1/4$. Estimates for the amplitudes were obtained by fitting the coefficients
1923: $p_m^{(1,k)}=[x^m]\PPM{1}{k}(x)$ to the form given above using an increasing number of
1924: amplitudes. `Experimentally' we find we need about the same total number of terms at
1925: $x_c$ and $x_-$. So in the fits we used the terms with amplitudes $a_i$, $b_i$ and $c_i$,
1926: $i=0,\ldots,K$ and $d_i$, $i=0,\ldots,3K$. Going only to $K$ with the $d_i$ amplitudes
1927: results in much poorer convergence and going beyond $3K$ leads to no improvement. So we
1928: use the $6K+4$ terms $p_m^{(1,k)}$ with $m$ ranging from $M$ to $M-6K-3$ and solve the
1929: resulting system of $6K+4$ linear equations.
1930:
1931: \begin{figure}[htbp]
1932: \centering
1933: \includegraphics[scale=0.95]{St1_ampl}
1934: \caption{\label{fig:1StA} The relative precision of the estimates for the leading amplitude,
1935: $|a_0 - \La{1}{k}|/\La{1}{k}$, against $1/M$ for the first (top left panel), second
1936: (top right), fifth (bottom left) and tenth (bottom right) area moments.}
1937: \end{figure}
1938:
1939: We compare the amplitude estimates to the predictions in Section~\ref{sec:scaling_arbitrary}
1940: and we find that the estimated leading amplitude $a_0$ is given by the exact formula
1941: \begin{equation}\label{eq:StA}
1942: \La{r}{k}=\frac{(-1)^{k+r}(k+r)! x_c^r f_{k+r}}{r! x_c^{\gamma_{k+r}}\Gamma(\gamma_{k+r})},
1943: \end{equation}
1944: which agrees with Eq.~(\ref{eq:Aka}) since the critical half-perimeter generating,
1945: see Eq.~(\ref{eq:StGf}), for staircase polygons is $\PGf(x_c)=1/4=x_c$.
1946: In \cite{GJ06b} we studied the normalised coefficients
1947: $r_m = p_{m+8}^{(1)}/4^m.$ Using the recurrence relations for $p_m^{(1)}$ (derived from the ODE)
1948: it is easy and fast to generate many more terms $r_m$. We generated the first 100000 terms and saved
1949: them as floats with 500 digit accuracy. We found (to better than 100 digits) that the leading
1950: amplitude of the normalised series was $\tilde{a}_0=1024$. Going back to the normalisation
1951: used in this paper we find $a_0=1024/4^8=1/64$ in agreement with formula~(\ref{eq:StA}).
1952: In figure~\ref{fig:1StA} we plot the relative precision of the estimates for the amplitude,
1953: $|a_0 - \La{r}{k}|/\La{r}{k}$ against $1/M$ for some of the area moments.
1954: Recall that for $k=1$ we have a series of 352 non-zero terms, for $k=2$ we have 292 terms
1955: and for $k=3$--10 we have 246 terms.
1956: In all cases (including for area moments not shown) the relative precision of the estimate
1957: $a_0$ is better than $10^{-15}$ for $K=10$. For the first area moment, where we have a longer
1958: series, the precision is even more impressive, being better than $10^{-20}$. So in all cases we
1959: can confirm the exact prediction (\ref{eq:StA}) for the amplitude to at least 15 significant
1960: digits.
1961:
1962:
1963: \begin{figure}[htbp]
1964: \centering
1965: \includegraphics[scale=0.95]{St1_logampl}
1966: \caption{\label{fig:1StC} The relative precision of the estimates for the amplitude,
1967: $|c_0 - \Lc{1}{k}|/\Lc{1}{k}$, against $1/M$ for the zeroth (leftmost panel)
1968: and tenth (rightmost panel) area moments.}
1969: \end{figure}
1970:
1971: In the previous section we saw that for minimally punctured staircase polygons
1972: the amplitude of the correction term is just a constant times
1973: the leading amplitudes with one less puncture, see Eq.~(\ref{eq:StMB}). It is thus
1974: natural to ask if something similar happens in the more general case.
1975: And indeed we find numerically that the amplitude $\Lc{1}{k}$ of the dominant
1976: correction in (\ref{eq:S1mom}) (the one proportional to the log-term) giving
1977: rise to the $c_j$-terms in (\ref{eq:S1co}) are
1978: \begin{equation}\label{eq:S1C}
1979: \Lc{1}{k} = - \frac{3\sqrt{3}}{8\pi} A_k.
1980: \end{equation}
1981: In figure~\ref{fig:1StC} we have plotted the relative error between the estimate
1982: of $c_0$ and the predicted value $\Lc{1}{k}$ for $k=0$ and 10. The estimates for the
1983: other moments are very similar with the accuracy of the agreement diminishing with
1984: higher moments. So in all cases (\ref{eq:S1C}) has been confirmed to better than 10
1985: digit accuracy.
1986:
1987: For the sub-dominant correction term the amplitude $\Lb{1}{k}$, as approximated
1988: by the term $b_0$ in (\ref{eq:S1co}), is not simply related to $A_k$ and in fact
1989: it even changes sign as $k$ is increased.
1990:
1991:
1992: \subsection{Staircase polygons with two staircase punctures \label{sec:stair_two}}
1993:
1994: In this section we report on our analysis of the series for staircase polygons with
1995: two punctures of arbitrary size. Our first task is to work out the singularity
1996: structure of the perimeter generating function $\PP{2}(x)$ (for which we have
1997: a series with 240 non-zero terms). From the general
1998: considerations of Section~\ref{sec:punc} we expect a singularity at $x=x_c=1/4$ with
1999: exponent $-5/2$, but given the quite complicated singularity structure of $\PP{1}(x)$,
2000: as detailed in Eq.~(\ref{eq:S1}) and below, we would expect similar complications
2001: for $\PP{2}(x)$. We analysed the series for $\PP{2}(x)$ using differential
2002: approximants \cite{AJG89a}. This analysis revealed that there is a triple root at $x=x_c=1/4$
2003: and the exponents had values $-2.499(1)$, $-2.070(5)$ and $-1.78(1)$.
2004: So despite having a series of 240 terms it is still very difficult
2005: to pin down the exponents accurately. Given the values quoted above two possible
2006: scenarios present themselves. Either the exponents have the exact values $-5/2$, $-2$ and
2007: $-2$ or they have the exact values $-5/2$, $-2$ and $-7/4$. The behaviour of
2008: $\PP{1}(x)$ would support the first of these scenarios and below we shall present
2009: evidence from the analysis of the asymptotic form of the coefficients which strongly
2010: supports this behaviour. We also find a double root singularity at $x=x_-=-1/4$
2011: with exponent estimates consistent with the exact values 5 and 11/2.
2012: In addition there are several conjugate pairs of singularities
2013: in the complex plane. The most important of these are at $x=(-1 \pm i \sqrt{3})/8$,
2014: which has magnitude $1/4$ and thus lies equidistant from the origin to $x_c$. So unlike
2015: the situation for once punctured staircase polygons we cannot ignore the complex singularities.
2016: The exponent estimate at this singularity is consistent with the exact value $33/2$.
2017: The singularities at $x=(-1 \pm i \sqrt{3})/8$
2018: are the roots of the polynomial $1+4x+16x^2$ which would indicate that the generating function
2019: contains a term $\sim E(x)(1+4x+16x^2)^{33/2}$. Finally, we find some singularities
2020: with magnitude greater than $1/4$. There are singularities at $x=(1 \pm i \sqrt{3})/6$
2021: (which has magnitude $1/3$) with an exponent $33/2$ (we note that these are the roots
2022: of ($1-3x+9x^2$)) and at $x=\pm i/2$ (magnitude $1/2$) with
2023: an exponent consistent with the value $5$. We also find weak evidence that $\PP{2}(x)$,
2024: just as $\PP{1}(x),$ has a pair of singularities at the roots of $1+x+7x^2$.
2025:
2026: As noted above we need to include the contribution from a conjugate pair of complex
2027: singularities to the asymptotic form of the coefficients. In general this is not as
2028: straight-forward a task as dealing with singularities on the real axis. In \cite{IJ06a}
2029: we examined the generating function of self-avoiding walks on the honeycomb lattice
2030: which has a pair of singularities on the imaginary axis at $x=\pm i/\tau$, arising
2031: from a term of the form $H(x)(1+\tau^2 x^2)^{-\eta}$. This typically produces coefficients
2032: which change sign according to a $++--$ pattern. This can be accommodated by including
2033: terms of the form
2034:
2035: $$
2036: \sim \tau^m m^{\eta-1} \sum_{j\geq 0} (-1)^{\lfloor (m+j)/2 \rfloor} h_j/m^j
2037: $$
2038: in fitting to the asymptotic form of the coefficients.
2039:
2040: Note that the analysis in \cite{IJ06a} clearly demonstrated that, as is done above,
2041: one has to shift the sign-pattern by $j$ when terms proportional to $1/m^j$ are
2042: considered. Terms arising from other complex conjugate pairs of singularities can
2043: give rise to much more complicated sign-patterns. In order to handle such cases we
2044: simply form the Taylor expansion of the simplest possible term arising from the
2045: singularity and take the sign of the appropriate coefficient. Specifically in
2046: order to include terms proportional to $1/m^j$, when looking at the coefficients
2047: $[x^m]\PP{2}(x)$, we take the sign to be the sign of the coefficient of $x^{m+j}$
2048: in the Taylor expansion of the function $(1+4x+16x^2)^{33/2}$. We use the notation
2049: $\Sign{(1\!+\!4x\!+\!16x^2)^{33/2}}{m\!+\!j}$ for this operation.
2050:
2051:
2052: The singularity structure revealed above leads us to fit the coefficients
2053: to a form, which is appropriate if the first scenario (exponents at $x=x_c=1/4$
2054: are $-5/2$, $-2$ and $-2$) is correct
2055: \begin{eqnarray}\label{eq:St2Log}
2056: [x^m]\PP{2}(x)\! =&\! 4^m\! \left ( \sum_{j=0}^K \!
2057: m^{3/2-j}\! \left [ a_j\!+\!\frac{1}{\sqrt{m}}\left [ b_j \!+\!c_j \log (m) \right ]\! \right ]
2058: +(-1)^m \sum_{j=0}^{3K}\! d_j m^{-6-j/2} \right .\nonumber \\
2059: &\left .+ \sum_{j=0}^{3K}\! e_j \Sign{(1\!+\!4x\!+\!16x^2)^{33/2}}{m\!+\!j} m^{-35/2+j}\! \right ),
2060: \end{eqnarray}
2061: where we have ignored the singularities with magnitude exceeding $1/4$.
2062: We also examine the alternative form appropriate if the second scenario
2063: (exponents at $x=x_c=1/4$ are $-5/2$, $-2$ and $-7/4$) is correct
2064: \begin{eqnarray}\label{eq:St2Exp}
2065: [x^m]\PP{2}(x)\! =&\! 4^m\! \left ( \sum_{j=0}^K \!
2066: m^{3/2-j}\left [ a_j\!+\!\frac{b_j}{m^{1/2}}\!+\!\frac{c_j}{m^{3/4}}\! \right ]
2067: +(-1)^m \sum_{j=0}^{3K}\! d_j m^{-6-j/2}\!\right . \nonumber \\
2068: &\left . + \sum_{j=0}^{3K}\! e_j \Sign{(1\!+\!4x\!+\!16x^2)^{33/2}}{m\!+\!j} m^{-35/2+j}\! \right ).
2069: \end{eqnarray}
2070:
2071: \begin{figure}[htbp]
2072: \centering
2073: \includegraphics[scale=0.95]{St2_ampl}
2074: \caption{\label{fig:2StA} Estimates of the leading amplitude $a_0$ and the amplitude $c_0$ of the
2075: sub-dominant term.
2076: The top panels show the results from fitting to the form~(\protect{\ref{eq:St2Log}})
2077: while the bottom panels are results from fitting to the form~(\protect{\ref{eq:St2Exp}}).
2078: The straight line is the exact value of this amplitude $\La{2}{0}$}
2079: \end{figure}
2080:
2081: In figure~\ref{fig:2StA} we have plotted the resulting estimates for the leading amplitude
2082: $a_0$, which we expect are given by the exact value $\La{2}{0} =\frac{5}{3072\sqrt{\pi}}$, and the
2083: amplitude $c_0$ of the sub-dominant terms from the two alternative
2084: asymptotic forms. First we focus on the sub-dominant terms. In the top right
2085: panel we show the estimates when fitting to the form~(\ref{eq:St2Log}) where we have a
2086: sub-dominant term proportional to $m \log (m)$ while the bottom right panel shows the
2087: estimates obtained when fitting to the form~(\ref{eq:St2Exp}) where the sub-dominant
2088: term is $m^{3/4}$ (in both cases the dominant term is proportional to $m^{3/2}$ with
2089: a second sub-dominant term proportional to $m$). In the bottom right panel we note that
2090: the estimates for the amplitude $c_0$ of the term $m^{3/4}$ seems to diverge. As $K$
2091: is increased rather than settle down we find that the slope of the curves of the estimates
2092: plotted vs. $1/M$ {\em increases} which is the opposite of what we would expect if we
2093: are fitting to the correct asymptotic form. We take this as firm numerical evidence that
2094: (\ref{eq:St2Exp}) is incorrect. This could also explain why the corresponding estimates of
2095: $a_0$ (bottom left panel) don't appear to converge to the predicted exact value. In
2096: contrast the estimates for $c_0$ of the term $m\log (m)$ (top right panel) from the
2097: form~(\ref{eq:St2Log}) do seem to converge and the slopes of the estimates plotted vs. $1/M$
2098: decrease with $K$. In this case the estimates for $a_0$ (top left panel) clearly can be
2099: extrapolated to a value consistent with the predicted exact value. Note further
2100: that the top left panel has a resolution along the ordinate which is a factor 5 higher than
2101: in lower right panel so the estimates when fitting to the form~(\ref{eq:St2Log}) are
2102: much more tightly converged. The conclusion is that the numerical evidence clearly
2103: favours the asymptotic form~(\ref{eq:St2Log}) and we believe this to be (if not entirely
2104: correct at least a very good approximation to) the true asymptotic form of the
2105: coefficients of the generating function $\PP{2}(x)$ for twice punctured staircase polygons.
2106:
2107:
2108: \begin{figure}[htbp]
2109: \centering
2110: \includegraphics[scale=0.95]{St2_mom}
2111: \caption{\label{fig:2StM} Estimates of the leading amplitude
2112: $\La{2}{k}$ of various area moments
2113: obtained by fitting to the form~(\protect{\ref{eq:St2Mom}}).
2114: The straight line is the exact value of this amplitude.}
2115: \end{figure}
2116:
2117: Now that we have settled the question of the correct singularity structure of $\PP{2}(x)$
2118: we turn our attention to the analysis of the area moment generating function $\PPM{2}{k}(x)$.
2119: As for once punctured staircase polygons we find that the leading exponent at all singularities
2120: decreases by $3k/2$. So in order to estimate the leading amplitude $a_0$ we fit to the form
2121: \begin{eqnarray}\label{eq:St2Mom}
2122: [x^m]\PPM{2}{k}(x)\! =&\! 4^m\! \left ( \sum_{j=0}^K \!
2123: m^{3(1+k)/2-j}\! \left [ a_j\!+\!\frac{1}{\sqrt{m}}\left [ b_j \!+\!c_j \log (m) \right ]\! \right ]
2124: +(-1)^m \sum_{j=0}^{3K}\! d_j m^{(-12+3k+j)/2} \right .\nonumber \\
2125: &\left .+ \sum_{j=0}^{3K}\! e_j \Sign{(1\!+\!4x\!+\!16x^2)^{33/2}}{m\!+\!j} m^{-(35-3k+j)/2}\! \right ),
2126: \end{eqnarray}
2127: where for simplicity our notation suppresses the $k$-dependence of the amplitudes.
2128: Recall that for $k=1$ and 2 we have series with 214 terms and for $k=3$--10 we
2129: have 140 terms.
2130: We compare the amplitude estimates to the predictions of Section~\ref{sec:scaling}.
2131: The leading amplitude $a_0=\La{2}{k}$ is given by the exact
2132: formula~(\ref{eq:StA}). In figure~\ref{fig:2StM} we show the estimates for the
2133: leading amplitudes for area moments with $k=1$, 2, 5, and 10.
2134: In all cases the amplitudes estimates appears to converge to the predicted exact
2135: value and agreement is found to at least 3 significant digits.
2136:
2137:
2138: \begin{figure}[htbp]
2139: \centering
2140: \includegraphics[scale=0.95]{St2_logampl}
2141: \caption{\label{fig:2StC} The ratio $c_0/\La{1}{k}$ for the zeroth area moment
2142: (leftmost panel) and several different area moments (rightmost panel).}
2143: \end{figure}
2144:
2145: Finally we turn our attention to the amplitude $\Lc{2}{k}$ of the dominant correction term.
2146: In the leftmost panel of figure~\ref{fig:2StC} we have plotted the ratio between
2147: the estimated amplitude $c_0$ and the predicted value of $\La{1}{k}$ for $k=0$ using
2148: several cut-offs. In the rightmost panel we show the same ratio but for
2149: several different moments using the cut-off $K=6$. These plots are consistent
2150: with $\Lc{2}{k} \propto \La{1}{k}$ with the constant of proportionality being
2151: $-0.212(2)$.
2152:
2153: \subsection{Punctured self-avoiding polygons \label{sec:sapana}}
2154:
2155: \begin{table}
2156: \begin{center}
2157: \small
2158: \begin{tabular}{llll}
2159: \hline\hline
2160: \multicolumn{1}{c}{Degrees} & \multicolumn{3}{c}{Once punctured $k=0$} \\
2161: \hline
2162: $[8,9,9,10]$ & $ -0.002022 -0.023965i$ & $ -0.002022+ 0.023965i$ & $ 0.518101$ \\
2163: $[8,9,9,11]$ & $ -0.003847+ 0.034768i$ & $ -0.003847 -0.034768i$ & $ 0.528973$ \\
2164: $[8,9,9,12]$ & $ -0.004553 -0.038104i$ & $ -0.004553+ 0.038104i$ & $ 0.533388$ \\
2165: $[8,9,10,10]$ & $ -0.001830 -0.023143i$ & $ -0.001830+ 0.023143i$ & $ 0.515916$ \\
2166: $[8,9,10,11]$ & $ -0.044131$ & $ 0.060194$ & $ 0.459558$ \\
2167: $[8,9,11,10]$ & $ -0.006365 -0.045090i$ & $ -0.006365+ 0.045090i$ & $ 0.547479$ \\
2168: $[8,10,9,10]$ & $ -0.007054 -0.047578i$ & $ -0.007054+ 0.047578i$ & $ 0.552445$ \\
2169: $[8,10,9,11]$ & $ -0.009398 -0.055135i$ & $ -0.009398+ 0.055135i$ & $ 0.570859$ \\
2170: $[8,10,10,10]$ & $ -0.012495 -0.063935i$ & $ -0.012495+ 0.063935i$ & $ 0.595957$ \\
2171: \hline\hline
2172: \end{tabular}
2173: \end{center}
2174: \caption{\label{tab:SAP1pM0}
2175: Biased estimates of the critical exponents of once punctured SAPs.}
2176: \end{table}
2177:
2178: Before proceeding to the estimation of the amplitudes we briefly have a look
2179: at the critical behaviour of the area moment generating functions for
2180: punctured self-avoiding polygons. In \cite{GJWE00} we analysed the behaviour of
2181: $\PPM{r}{k}(x)$ and found the critical behaviour to be consistent with
2182: $\PPM{r}{k}(x) \sim A(x)+B(x)(x-x_c)^{-\gamma_{k+r}}+ C(x)(x-x_c)^{-\gamma_{k+r}+1/2}$,
2183: where $\gamma_{j} = 3j/2-3/2$.
2184: From previous work \cite{JG99,IJ03} we have very precise estimates for $x_c,$ which
2185: is indistinguishable from the positive root of the polynomial $581x^2+7x-13$, that is,
2186: $x_c=0.1436806292\ldots$. Using this value for $x_c$
2187: we form a $K^{th}$-order {\em biased} differential approximant (DA) to $\PPM{r}{k}(x)$
2188: by matching the coefficients in the polynomials $Q_i(x)$ of degree $N_i$
2189: so that (one) of the formal solutions to the homogeneous differential equation
2190: $$
2191: \sum_{i=0}^K (x-x_c)^iQ_i(x)(x\frac{{\rm d}}{{\rm d}x})^i \tilde{P}(x) = 0
2192: $$
2193: agrees with the first $M=\sum_i (N_i+1)$ series coefficients of $\PPM{r}{k}(x)$.
2194: We are thus `forcing' the DAs to have regular singular points at the origin and $x_c$.
2195: The critical exponents $\gamma_j$ ($j=1\ldots K$) are estimated from the indicial equation
2196: at $x_c$ (note that due to the biasing, $x_c$ is root of order $K$).
2197: If the true singular behaviour at $x_c$ implies a root of degree less than $K$ we
2198: expect that the `true' exponents will be quite well estimated and show
2199: little scatter while the `surplus' exponents will show a lot of random scatter.
2200: In the following we always use $K=3$ and denote the degrees of the polynomials
2201: $Q_i$ as $[N_3,N_2,N_1,N_0]$.
2202:
2203: First we look at the perimeter generating function for once punctured SAPs (for
2204: which we have series with 42 terms).
2205: In this case we have $\gamma_1 =0$ and we thus expect a logarithmic singularity at $x_c$
2206: with a square-root correction term as argued in \cite{GJWE00}.
2207: In Table~\ref{tab:SAP1pM0} we list some exponent estimates obtained from the biased
2208: DAs. The exponent estimates are indeed consistent with the exact values $0,0,1/2$,
2209: which confirms the expected behaviour.
2210:
2211:
2212: \begin{table}
2213: \begin{center}
2214: \small
2215: \begin{tabular}{lllllll}
2216: \hline\hline
2217: \multicolumn{1}{c}{Degrees} & \multicolumn{3}{c}{Once punctured $k=2$}
2218: & \multicolumn{3}{c}{Once punctured $k=5$} \\
2219: \hline
2220: $[8,9,9,10]$ & $ -3.000291$ & $ -2.493921$ & $ -1.154513$
2221: & $ -7.500838$ & $ -6.976201$ & $ -5.107975$ \\
2222: $[8,9,9,11]$ & $ -3.000158$ & $ -2.496645$ & $ -1.258606$
2223: & $ -7.503030$ & $ -6.927432$ & $ -2.422194$ \\
2224: $[8,9,9,12]$ & $ -3.012422$ & $ -2.421400$ & $ -3.323742$
2225: & $ -7.500231$ & $ -6.992184$ & $ -5.461825$ \\
2226: $[8,9,10,10]$ & $ -2.999849$ & $ -2.502858$ & $ -1.447159$
2227: & $ -7.517118$ & $ -6.776728$ & $ -8.010389$ \\
2228: $[8,9,10,11]$ & $ -3.000452$ & $ -2.491160$ & $ -1.022360$
2229: & $ -7.500700$ & $ -6.980331$ & $ -5.251153$ \\
2230: $[8,9,11,10]$ & $ -3.000143$ & $ -2.497148$ & $ -1.307314$
2231: & $ -7.500366$ & $ -6.988804$ & $ -5.406296$ \\
2232: $[8,10,9,10]$ & $ -2.999481$ & $ -2.511762$ & $ -1.668527$
2233: & $ -7.505629$ & $ -6.879658$ & $-14.20958$\\
2234: $[8,10,9,11]$ & $ -2.999906$ & $ -2.501910$ & $ -1.441981$
2235: & $ -7.500535$ & $ -6.984576$ & $ -5.361394$ \\
2236: $[8,10,10,10]$ & $ -2.999647$ & $ -2.507601$ & $ -1.578352$
2237: & $ -7.500946$ & $ -6.974159$ & $ -5.059683$ \\
2238: \hline\hline
2239: \end{tabular}
2240: \end{center}
2241: \caption{\label{tab:SAP1pM25}
2242: Biased estimates of the critical exponents for the 2nd and 5th area moments of once punctured SAPs.}
2243: \end{table}
2244:
2245:
2246: \begin{table}
2247: \begin{center}
2248: \small
2249: \begin{tabular}{lllrlll}
2250: \hline\hline
2251: \multicolumn{1}{c}{Degrees} & \multicolumn{3}{c}{Twice punctured $k=0$}
2252: & \multicolumn{3}{c}{Twice punctured $k=2$} \\
2253: \hline
2254: $[7,8,8,9]$ & $ -1.504250$ & $ -0.968766$ & $ 6.594627$
2255: & $ -4.499111$ & $ -4.006801$ & $ -2.544930$ \\
2256: $[7,8,8,10]$ & $ -1.504232$ & $ -0.968829$ & $ 5.029521$
2257: & $ -4.495871$ & $ -4.039199$ & $ -3.083192$ \\
2258: $[7,8,8,11]$ & $ -1.504209$ & $ -0.968910$ & $-60.90721$
2259: & $ -4.494258$ & $ -4.048789$ & $ -2.701322$ \\
2260: $[7,8,9,9]$ & $ -1.504210$ & $ -0.968909$ & $ 12.03204$
2261: & $ -4.494568$ & $ -4.047486$ & $ -2.836670$ \\
2262: $[7,8,9,10]$ & $ -1.504197$ & $ -0.968951$ & $ 32.28205$
2263: & $ -4.493872$ & $ -4.051402$ & $ -2.672813$ \\
2264: $[7,8,10,9]$ & $ -1.504205$ & $ -0.968923$ & $ 22.27646$
2265: & $ -4.508207$ & $ -3.915977$ & $ -1.950665$ \\
2266: $[7,9,8,9]$ & $ -1.504178$ & $ -0.969017$ & $ 18.18298$
2267: & $ -4.495340$ & $ -4.040712$ & $ -2.812621$ \\
2268: $[7,9,8,10]$ & $ -1.504177$ & $ -0.969020$ & $ 18.81517$
2269: & $ -4.495136$ & $ -4.041237$ & $ -2.679530$ \\
2270: $[7,9,9,9]$ & $ -1.504176$ & $ -0.969022$ & $ 18.58757$
2271: & $ -4.498396$ & $ -4.012876$ & $ -2.698591$ \\
2272: \hline\hline
2273: \end{tabular}
2274: \end{center}
2275: \caption{\label{tab:SAP2pM02}
2276: Biased estimates of the critical exponents for the 0th and 2nd area moments of twice punctured SAPs.}
2277: \end{table}
2278:
2279: In Table~\ref{tab:SAP1pM25} we list some exponent estimates for the 2nd and 5th
2280: area moments of once punctured SAPs. The exponents support the expectations for the
2281: leading and sub-dominant exponent. The third exponent is of no significance--the nature
2282: of the differential approximant forces a third exponent to have some value, but its lack
2283: of convergence suggests it is not, in fact present. In Table~\ref{tab:SAP2pM02} we list exponent
2284: estimates for the 0th and 2nd area moments of twice punctured SAPs (for
2285: which we have series with 38 terms). Similar comments
2286: apply to the three columns of exponent estimates as were made for the 2nd area moments.
2287: In all cases we get a clear confirmation of the critical behaviour observed in \cite{GJWE00}.
2288:
2289: Next we turn our attention to estimates for the critical amplitudes.
2290: Proposition~\ref{prop:minpunc} and Theorem~\ref{theo:arb} tells us that the critical amplitude
2291: of the $k$th area moment of self-avoiding polygons with $r$ (minimal or arbitrary)
2292: punctures is proportional to the critical amplitude of the $(k+r)$th area moment
2293: of unpunctured SAPs (the theorems also give the constants of proportionality).
2294: In order to test these predictions numerically we analyse in this section data
2295: for SAPs with one and two punctures. In all cases we look at the ratio
2296: $r_m=p_m^{(r,k)}/p_m^{(k+r)}$ which should approach the relevant constant of
2297: proportionality. Given the critical behaviour outlined above we expect that these
2298: amplitude ratios can be approximated quite well by the asymptotic form
2299: $$r_m = \sum_{j=0} a_j m^{j/2}$$
2300: So as in the analysis of punctured staircase polygons we obtain estimates for the
2301: leading amplitude $a_0$, by fitting to the above form truncated at some level $K$,
2302: and we then plot these estimates against $1/M$.
2303:
2304:
2305:
2306: \subsubsection{Minimal punctures}
2307:
2308: According to Proposition~\ref{prop:minpunc} the amplitude of the $k$-th area moment of
2309: SAPs with $r$ minimal punctures is
2310:
2311: \begin{equation}
2312: A_k^{(r)}=A_{k+r}x_c^{2r}/r!.
2313: \end{equation}
2314:
2315: We first analyse the area moments of SAPs with a single minimal puncture.
2316: In the left panel of figure~\ref{fig:SapM1A} we have plotted estimates for amplitude ratio
2317: $A_0^{(1)}/A_1$ with $7 \leq K \leq 10$. The prediction for this ratio is $x_c^2$, which
2318: is plotted as a straight line. The estimates obtained with $K=9$ and $10$ are indistinguishable
2319: from the predicted value at the resolution of the plot. Note that the `curvature' of
2320: the plotted values decreases as $K$ increases. We take this to be a very clear indication
2321: that the ratio $r_m$ is very well approximated by the assumed asymptotic form.
2322: In the right panel we plot the estimates for the amplitude ratio $A_k^{(1)}/A_{k+1}$
2323: for $0 \leq k \leq 9$ using the cut-off value $K=10$. For small values of $k$ these plots
2324: give firm numerical evidence for the correctness of Proposition~\ref{prop:minpunc}. For higher
2325: values of $k$ (8 and 9 in particular) the evidence is not quite as firm though nothing
2326: in the plot would suggest a discrepancy with the predicted value. We again emphasise
2327: that in this case we have relatively short series of only 42 terms.
2328:
2329:
2330:
2331: \begin{figure}[htbp]
2332: \centering
2333: \includegraphics[scale=1.0]{SapM1_amplpred}
2334: \caption{\label{fig:SapM1A} Estimates of the amplitude ratio $A_k^{(1)}/A_{k+1}$
2335: for self-avoiding polygons with one minimal puncture.}
2336: \end{figure}
2337:
2338:
2339: Next we analyse the area moments of SAPs with two minimal punctures.
2340: The left panel of figure~\ref{fig:SapM2A} shows estimates for amplitude ratio
2341: $A_0^{(2)}/A_2$ with $7 \leq K \leq 10$. The prediction for the ratio, $x_c^4/2$,
2342: is plotted as a straight line. Again we find that the estimates for $K=10$ are
2343: indistinguishable from the predicted value, though in this case the relative
2344: resolution is coarser than in the previous plot.
2345: In the right panel we plot estimates of the amplitude ratio $A_k^{(2)}/A_{k+2}$
2346: for $0 \leq k \leq 8$ using the cut-off value $K=10$. Again we find that our
2347: numerical analysis confirms the prediction to a high degree of confidence.
2348: Recall that in this case we have series with only 38 terms.
2349:
2350: \begin{figure}[htbp]
2351: \centering
2352: \includegraphics[scale=1.0]{SapM2_amplpred}
2353: \caption{\label{fig:SapM2A} Estimates of the amplitude ratio $A_k^{(2)}/A_{k+2}$
2354: for self-avoiding polygons with two minimal punctures.}
2355: \end{figure}
2356:
2357:
2358: \subsubsection{Arbitrary punctures}
2359:
2360: According to Theorem~\ref{theo:arb} the amplitude of the $k$-th area moment of
2361: SAPs with $r$ arbitrary punctures is
2362:
2363: \begin{equation}
2364: A_k^{(r)}=A_{k+r}({\cal P}(x_c))^r/r!,
2365: \end{equation}
2366: where $ {\cal P}(x_c)$ is the critical amplitude of the half-perimeter generating function.
2367:
2368: The first step in our analysis is to obtain an accurate estimate of ${\cal P}(x_c)$.
2369: In \cite{JG99} we obtained the rather imprecise estimate ${\cal P}(x_c)\approx 0.036$
2370: by evaluating Pad\'e approximants to the generating function. Here we shall
2371: estimate ${\cal P}(x_c)$ directly from the perimeter data. We first tried the form
2372: %
2373: \begin{displaymath}
2374: \sum_{m=0}^M p_{m}x_c^m \sim {\cal P}(x_c)+a_1/M^{1/2}+a_2/M+\cdots
2375: \end{displaymath}
2376: Using the first ten terms in this asymptotic expansion we found (with $M=55$)
2377: that ${\cal P}(x_c)\approx 0.0362642$, but $a_1\approx 4.28\times10^{-9}$ and
2378: $a_2\approx -1.42\time 10^{-7}$, while $a_3 \approx -0.066$. We are therefore
2379: quite confident that $a_1=a_2=0$. Upon further analysis we found convincing evidence
2380: that the correct asymptotic form in fact is
2381:
2382: \begin{displaymath}
2383: \sum_{m=0}^M p_{m}x_c^m \sim {\cal P}(x_c)+b_1/M^{3/2}+b_2/M^{5/2}+\cdots
2384: \end{displaymath}
2385:
2386:
2387: \begin{table}
2388: \begin{center}
2389: \begin{tabular}{cccc}
2390: \hline\hline
2391: $M$ & $K=8$ & $K=10$ & $K=12$ \\
2392: \hline
2393: 45 & 0.036264215181387
2394: & 0.036264215181095
2395: & 0.036264215180475 \\
2396: 46 & 0.036264215181343
2397: & 0.036264215181088
2398: & 0.036264215181354 \\
2399: 47 & 0.036264215181305
2400: & 0.036264215181073
2401: & 0.036264215180915 \\
2402: 48 & 0.036264215181271
2403: & 0.036264215181060
2404: & 0.036264215180994 \\
2405: 49 & 0.036264215181240
2406: & 0.036264215181048
2407: & 0.036264215180970 \\
2408: 50 & 0.036264215181214
2409: & 0.036264215181038
2410: & 0.036264215181001 \\
2411: 51 & 0.036264215181190
2412: & 0.036264215181029
2413: & 0.036264215180972 \\
2414: 52 & 0.036264215181168
2415: & 0.036264215181021
2416: & 0.036264215180972 \\
2417: 53 & 0.036264215181149
2418: & 0.036264215181013
2419: & 0.036264215180969 \\
2420: 54 & 0.036264215181131
2421: & 0.036264215181006
2422: & 0.036264215180967 \\
2423: 55 & 0.036264215181116
2424: & 0.036264215181000
2425: & 0.036264215180962 \\
2426: \hline\hline
2427: \end{tabular}
2428: \end{center}
2429: \caption{\label{tab:SAPampl}
2430: Estimates of the critical SAP amplitude ${\cal P}(x_c)$.}
2431: \end{table}
2432:
2433: \noindent
2434: Indeed, the observation that $a_1$ and $a_2$ vanish follows from the assumed
2435: asymptotic behaviour of $p_m$.
2436: In Table~\ref{tab:SAPampl} we have listed estimates for ${\cal P}(x_c)$ obtained
2437: using various values of $M$ and the cut-off $K$ in the asymptotic form. From this
2438: we confidently estimate that ${\cal P}(x_c)=0.0362642151808(2)$.
2439:
2440: \begin{figure}[htbp]
2441: \centering
2442: \includegraphics[scale=1.0]{Sap1_amplpred}
2443: \caption{\label{fig:Sap1A} Estimates of the amplitude ratio $A_k^{(1)}/A_{k+1}$ for
2444: self-avoiding polygons with one puncture of arbitrary size.}
2445: \end{figure}
2446:
2447: In figure~\ref{fig:Sap1A} we have plotted estimates of the amplitude ratio
2448: $A_k^{(1)}/A_{k+1}$ for self-avoiding polygons with one puncture of arbitrary size.
2449: In the leftmost panel we look at the ratio $A_0^{(1)}/A_{1}$ using different
2450: cut-offs $K$. The rightmost panel shows the ratio $A_k^{(1)}/A_{k+1}$ for
2451: area moments up to $k=9$ using the cut-off $K=10$. The straight line corresponds
2452: to the expected value ${\cal P}(x_c)$, using the estimate for this quantity obtained above.
2453: The estimates in the leftmost panel show
2454: some variation when plotted against $1/M$, but in the limit $M\to \infty$
2455: the estimates appear to converge to the expected value (if the trend holds). As for minimally
2456: punctured SAPs we see an ever closer agreement as $K$ is increased, again indicating
2457: that the assumed asymptotic form is reasonable. Obviously, as seen in the rightmost panel,
2458: the estimates for high area moments are not as close to the expected value. However,
2459: given the trend in these estimates we are confident in stating that our
2460: numerical analysis is consistent with the results of Theorem~\ref{theo:arb}.
2461: The agreement is particularly impressive bearing in mind that we analyse
2462: series with just 42 terms.
2463:
2464: \begin{figure}[htbp]
2465: \centering
2466: \includegraphics[scale=0.95]{Sap2_amplpred}
2467: \caption{\label{fig:Sap2A} Estimates of the amplitude ratio $A_k^{(2)}/A_{k+2}$
2468: for self-avoiding polygons with two punctures of arbitrary size.}
2469: \end{figure}
2470:
2471: Finally in figure~\ref{fig:Sap2A} we have plotted our estimates of the amplitude ratio
2472: $A_k^{(2)}/A_{k+2}$ for self-avoiding polygons with two punctures of arbitrary size.
2473: The straight line corresponds to the expected value ${\cal P}(x_c)^2/2$. Again
2474: all estimates are consistent with the results of Theorem~\ref{theo:arb}, though
2475: the numerical agreement is less convincing, but then again the series have only
2476: 38 terms.
2477:
2478:
2479: \section{Conclusion}
2480:
2481: We have rigorously analysed the effect of punctures on
2482: the area law of polygon models. In particular we obtained
2483: expressions for the leading amplitudes of the area moments
2484: for punctured polygons in terms of the amplitudes for unpunctured
2485: polygons (see Theorems~\ref{theo:bounded} and~\ref{theo:arb}). For staircase
2486: polygons this led to exact formulas for the amplitudes.
2487: For self-avoiding polygons the formulas rely on an assumption about
2488: the asymptotic behaviour of SAPs without punctures. They contain constants not
2489: known exactly but estimated numerically to a very high degree of accuracy.
2490: Our analysis also led to conjectures
2491: about the scaling behaviour of these models (see Conjecture~\ref{con1}).
2492: A proof of these conjectures is an open and difficult problem. Further
2493: numerical support of these conjectures might follow from an analysis
2494: of critical perimeter moments along the lines of \cite{RJG04}.
2495:
2496: The expressions for the amplitudes were thoroughly checked
2497: numerically. For staircase polygons with up to 5 minimal punctures and
2498: staircase polygons with one or two punctures of fixed size
2499: we used our series expansions to find the {\em exact} generating
2500: functions for area moments to order 10. Naturally, in all these cases the
2501: leading amplitude agrees with the proved formulas. Interestingly
2502: we find that the amplitude of the correction term is proportional
2503: to the corresponding leading amplitude with one less puncture.
2504: For staircase polygons with one and two punctures of arbitrary size
2505: a careful asymptotic analysis of the series for the area moments
2506: yielded very accurate estimates for the amplitudes, again confirming
2507: the exact formulas. Finally, we also analysed series for self-avoiding
2508: polygons with one and two punctures (minimal as well as arbitrary).
2509: In this case the numerical evidence is not quite as convincing, but
2510: we did find that the numerical estimates agree with the exact formulas
2511: to at least 3--4 significant digits.
2512:
2513: The numerical analysis also yielded explicit expressions for correction
2514: terms, see Eqs.~(\ref{eq:StMB}), (\ref{eq:S1C}) and subsection \ref{sec:stair_fix}.
2515: These correspond to contributions from puncture-boundary
2516: interactions and from puncture-puncture interactions. It would
2517: be interesting, but difficult to give a combinatorial proof of these results.
2518: The difficulty of any combinatorial proof becomes clearer when
2519: considering the recent closed form solution of a closely related model
2520: of punctured staircase polygons. In \cite{IJ07} one of us (IJ) considered
2521: a model of punctured staircase polygons in which the internal polygon
2522: is rotated by 90 degrees with respect to the outer polygon.
2523: The proofs in this paper never consider such restrictions on the
2524: placement of the internal polygon, that is, the internal polygon can
2525: be placed in any way one pleases. The results therefore carry over
2526: unaltered to the problem of rotated punctured staircase polygons.
2527: Interestingly, this means that the leading asymptotic form of
2528: the coefficients is exactly the same for both models, any differences arising
2529: only from the correction terms. The dominant correction term to the
2530: half-perimeter generating function for punctured staircase polygons
2531: $\PGf(x)$ is $\propto \log(1-4x)/\sqrt{1-4x}$ \cite{GJ06b}.
2532: From the exact closed form solution to the half-perimeter generating function
2533: for rotated punctured staircase polygons $\PPR(x)$ it follows that
2534: the first correction term is $\propto (1-4x)^{-3/4}$ \cite{IJ07}.
2535: These differences indicate that combinatorial arguments for a
2536: proof of sub-dominant behaviour must be quite subtle!
2537:
2538: In this paper, we discussed models of punctured polygons,
2539: where punctures are of the same type as the outer polygon.
2540: More generally, punctures may be built from different
2541: polygon classes. In that situation, Theorem 1 holds, with the obvious
2542: modification, for any collection of polygon models as punctures.
2543: Theorem 2 holds, with the obvious modification, for any collection
2544: of polygon models as punctures, if the corresponding half-perimeter generating
2545: functions are finite at the radius of convergence $x_c$ of the half-perimeter generating
2546: function of the outer polygon. This includes the model considered in
2547: \cite{IJ07} as a special case. We remark that the critical half-perimeter
2548: generating function of the outer polygon may be infinite.
2549:
2550: Our analytical results are based on the observation that
2551: puncture counting can be done using polygon area estimates,
2552: polygon boundary contributions being asymptotically
2553: negligible. In particular, effects of self-avoidance
2554: do not influence the results. This phenomenon is also
2555: expected to hold in higher dimension.
2556: Consider so-called polycubes, the generalisation of
2557: polyominoes to three dimensions. Polycubes have been
2558: enumerated by volume up to 18 cubes, see \cite{AB06}.
2559: Three-dimensional vesicles \cite{W93} are a subclass
2560: of polycubes, having no interior holes. Let a class of three-dimensional
2561: vesicles be given, counted by surface area, with a bounded number
2562: of vesicular holes. Assume that the asymptotic behaviour
2563: of the volume moments is known for the model without
2564: vesicular holes. If $0<\phi<1$ and the critical surface
2565: area generating function is finite, then our method of proof
2566: can be adapted to describe the volume law of three-dimensional
2567: vesicles with vesicular holes. For the full model of closed
2568: self-avoiding orientable surfaces of genus zero on the cubic
2569: lattice, however, there is numerical evidence that $\phi=1$,
2570: see e.g.~the review in \cite{Vanderzande}.
2571:
2572: For models with minimal punctures, the number $p_m^{\Box(1,0)}$
2573: in Eq.~(\ref{form:minprk}) also counts the number of
2574: polygons winding around a fixed point
2575: of the dual lattice. This problem and its generalisation
2576: to $M>1$ mutually avoiding self-avoiding polygons has been
2577: considered previously by Cardy \cite{C00}. It would be
2578: interesting to consider whether this generalisation
2579: can also be treated using the above methods. This
2580: involves the analysis of polygon models satisfying
2581: $\theta=0$. If $\theta<0$, interaction terms are generally
2582: not asymptotically negligible, so that the above
2583: analysis does not yield asymptotically exact estimates.
2584:
2585: A major open question is the problem of self-avoiding polygons
2586: with an unbounded number of punctures. Here, Theorem~\ref{theo:arb}
2587: yields an upper bound.
2588: Let $\QGf_{k}(x)$ denote the $k^{th}$ area moment generating function
2589: for a model of punctured polygons with an arbitrary number of punctures.
2590: $\QGf_{k}(x)$ is a formal power series, usually with zero radius of convergence.
2591: The (asymptotically exact) upper bound for models with $r$
2592: punctures yields an upper estimate for the number of punctured
2593: polygons with an arbitrary number of punctures. It is
2594: %
2595: \begin{displaymath}
2596: \QGf_{k}(x) \ll \sum_{r=0}^\infty \frac{\PGf_{k+r}(x)}{r!}(\PGf_{0}(x))^r,
2597: \end{displaymath}
2598: %
2599: where $\ll$ denotes coefficientwise majorisation. Let $\widetilde \QGf_{k}(x)$
2600: denote the $k^{th}$ factorial area moment generating function
2601: for punctured polygons with an arbitrary number of punctures.
2602: It can be shown that an upper bound is given by
2603: %
2604: \begin{displaymath}
2605: \widetilde \QGf_{k}(x) \ll \sum_{r=0}^\infty \frac{\widetilde \PGf_{k+r}(x)}{r!}
2606: (\widetilde \PGf_{0}(x))^r=\left.\frac{{\rm d}^k}{{\rm d}q^k} {\cal P}(x,q)
2607: \right|_{q={1+\cal P}(x,1)},
2608: \end{displaymath}
2609: %
2610: where $\widetilde \PGf_{k}(x)$ is the $k^{th}$ factorial moment
2611: generating function, and ${\cal P}(x,q)$ is the perimeter and area
2612: generating function of the model without punctures. In particular,
2613: the half-perimeter generating function $\QGf_{0}(x)=\widetilde
2614: \QGf_{0}(x)$ is majorised by
2615: %
2616: \begin{displaymath}
2617: \QGf_{0}(x) \ll {\cal P}(x,1+{\cal P}(x,1)).
2618: \end{displaymath}
2619: %
2620:
2621: Another problem touched upon in this article is punctured
2622: polygons in the fixed area ensemble. We gave a (non-rigorous)
2623: argument for values of the critical
2624: exponent in the branched polymer phase from the crossover
2625: behaviour of the tentative scaling function, thereby confirming
2626: previous results \cite{JvR92}.
2627: In that phase, boundary effects are indeed crucial, such that
2628: our methods of deriving limit distributions cannot be applied in this
2629: situation. On the other hand, area laws are expected to be
2630: of Gaussian type, as is usually the case away from
2631: phase transition points. The same phenomenon is expected
2632: to occur in the fixed perimeter ensembles for $q\ne1$.
2633:
2634:
2635: \section*{E-mail or WWW retrieval of series}
2636:
2637: The series for the generating functions studied in this paper
2638: can be obtained via e-mail by sending a request to
2639: I.Jensen@ms.unimelb.edu.au or via the world wide web on the URL
2640: http://www.ms.unimelb.edu.au/\~{ }iwan/ by following the instructions.
2641:
2642:
2643: \section*{Acknowledgements}
2644:
2645: The calculations presented in this paper would not have been possible
2646: without a generous grant of computer time on the server cluster of the
2647: Australian Partnership for Advanced Computing (APAC). We also used
2648: the computational resources of the Victorian Partnership for Advanced
2649: Computing (VPAC).
2650: CR and AJG would like to acknowledge support by the German Research Council
2651: (Deutsche Forschungsgemeinschaft) within the CRC701.
2652: IJ and AJG gratefully acknowledge financial support from the Australian Research Council.
2653:
2654:
2655: \section*{Appendix}
2656:
2657: We analyse the asymptotic growth of a Cauchy product of sequences
2658: in terms of the asymptotic growth of its constituting sequences.
2659: The following lemma is an extension of \cite[Thm~2]{B74} to the case
2660: of generating functions with equal radii of convergence. Its proof relies
2661: on Lebesgue's dominated convergence theorem \cite[Thm~1.34]{R87}, which
2662: states conditions under which an exchange of limit and sum is allowed:
2663: With $n\in\mathbb N$ and $k\in\mathbb N_0$ let real numbers
2664: $a_{n,k}$ be given. Assume that $\lim_{n\to\infty}a_{n,k}=:a_k\in\mathbb R$
2665: for all $k$ and that for all $k$ there is a bound $|a_{n,k}|\le b_k$
2666: uniformly in $n\in\mathbb N$. Assume that $\sum_{k\ge0} b_k\in
2667: \mathbb R$. Then $\lim_{n\to\infty} \sum_{k\ge0} a_{n,k} =
2668: \sum_{k\ge0} a_k\in\mathbb R$.
2669:
2670: \begin{lem}
2671: Let two sequences $(f_n)_{n\in\mathbb N_0}$ and $(g_n)_{n\in\mathbb N_0}$
2672: of real numbers be given, with generating functions $f(x)=
2673: \sum_{n\ge0}f_n x^n$ and $g(x)=\sum_{n\ge0}g_n x^n$. Assume that both
2674: generating functions have the same positive finite radius of convergence
2675: $x_c$, $0<x_c<\infty$. Assume that the sequences $(f_n)$ and $(g_n)$
2676: satisfy asymptotically
2677: %
2678: \begin{displaymath}
2679: f_n \sim Ax_c^{-n}n^{\gamma-1},\qquad g_n \sim Bx_c^{-n}n^{\delta-1} \qquad
2680: (n\to\infty),
2681: \end{displaymath}
2682: %
2683: for nonzero numbers $A\ne0$, $B\ne0$, and for real constants $\gamma,\delta$
2684: satisfying $\delta<0$ and $\gamma>\delta+1$. Assume that $g(x_c):=
2685: \lim_{x\nearrow x_c} g(x)\ne0$. Then, the Cauchy product of $(f_n)$ and $(g_n)$
2686: satisfies
2687: %
2688: \begin{displaymath}
2689: \sum_{k=0}^n f_{n-k} g_k = [x^n]f(x)g(x)\sim
2690: g(x_c) f_n \qquad (n\to\infty).
2691: \end{displaymath}
2692: %
2693: \end{lem}
2694:
2695: \begin{proof}
2696: Let $f_{n}:=0$ for $n<0$ and define for $n\in\mathbb N$ and for $k\in\mathbb N_0$
2697: %
2698: \begin{displaymath}
2699: a_{n,k}:=\frac{f_{n-k}}{Ax_c^{-n}n^{\gamma-1}}g_k.
2700: \end{displaymath}
2701: %
2702: Below, we derive a bound on $|a_{n,k}|$ uniformly in $n\in\mathbb N$ and
2703: summable in $k\in\mathbb N_0$. Then, Lebesgue's dominated convergence theorem
2704: can be applied to $a_{n,k}$. Since $g(x_c)\ne0$, this
2705: yields the statement of the lemma.
2706:
2707: Note first that the assumption on the asymptotic
2708: behaviour of $f_n$ implies the existence
2709: of a constant $n_0\in\mathbb N$ such that for all $n\ge n_0+k$ we have the
2710: estimate
2711: %
2712: \begin{displaymath}
2713: \left|\frac{f_{n-k}}{Ax_c^{-n}n^{\gamma-1}}\right|
2714: \le 2 x_c^k \left(1-\frac{k}{n}\right)^{\gamma-1}.
2715: \end{displaymath}
2716: %
2717: Fix such $n_0$. We distinguish the three cases $n< k$, $k\le n<n_0+k$, and $n\ge n_0+k$.
2718: If $n< k$, we clearly have $a_{n,k}=0$. For $n\ge n_0+k$,
2719: the above estimate yields
2720: %
2721: \begin{eqnarray*}
2722: \left|a_{n,k}\right|&&=\left|\frac{f_{n-k}}{Ax_c^{-n}n^{\gamma-1}}g_k\right|
2723: \le 2\left(1-\frac{k}{n}\right)^{\gamma-1}x_c^k \left|g_k\right|.
2724: \end{eqnarray*}
2725: %
2726: We will first consider the case $\gamma-1<0$. If $n\ge n_0+k$, we have
2727: $1/(1-k/n)\le1+k/n_0$. This implies
2728: %
2729: \begin{eqnarray*}
2730: \left|a_{n,k}\right| &\le 2^{2-\gamma_1} k^{1-\gamma}x_c^k\left|g_k\right|=:b_k^{(1)}.
2731: \end{eqnarray*}
2732: %
2733: If $k\le n<n_0+k$, we estimate similarly
2734: %
2735: \begin{eqnarray*}
2736: &\hspace{-1.5cm}\left|a_{n,k}\right|=\left|\frac{f_{n-k}}{Ax_c^{-n}n^{\gamma-1}}g_k\right|
2737: \le |A|^{-1}\max\{1,x_c^{n_0}\} \max_{m<n_0}\{|f_m|\} (n_0+k)^{1-\gamma}x_c^k |g_k|\\
2738: &\le |A|^{-1}\max\{1,x_c^{n_0}\}\max_{m<n_0}\{|f_m|\} \cdot
2739: \left\{\begin{array}{cc}n_0^{1-\gamma}|g_0|, & k=0\\
2740: (2n_0)^{1-\gamma}x_c^k k^{1-\gamma}|g_k|, & k\ne0
2741: \end{array}
2742: \right\}=:
2743: b_k^{(2)}.
2744: \end{eqnarray*}
2745: %
2746: Define $b_k:=b_k^{(1)}+b_k^{(2)}$ for $k\in\mathbb N_0$.
2747: Then $|a_{n,k}|\le b_k$ uniformly in $n\in\mathbb N$,
2748: and $\sum_{k\ge0}b_k<\infty$.
2749: Summability follows from
2750: %
2751: \begin{displaymath}
2752: \sum_{k=0}^\infty x_c^k k^{1-\gamma}|g_k|<\infty,
2753: \end{displaymath}
2754: %
2755: since by assumption $x_c^k k^{1-\gamma}|g_k|\sim |B| k^{\delta-\gamma}$ as $k\to\infty$,
2756: where $\gamma-\delta>1$.
2757:
2758: If $\gamma-1\ge0$, uniform estimates are obtained along the same lines.
2759: In that situation, the factors $(1-k/n)^{\gamma-1}$ and $(n_0+k)^{1-\gamma}$
2760: may be replaced by unity, resulting in simpler uniform bounds involving
2761: $\sum_{k\ge0} x_c^k |g_k|<\infty$.
2762: \end{proof}
2763:
2764:
2765: \begin{thebibliography}{10}
2766:
2767: \bibitem{MSbook}
2768: Madras N and Slade G 1993 {\it The Self-Avoiding Walk\/} (Boston: Birkh\"auser)
2769:
2770: \bibitem{HughesV1}
2771: Hughes B~D 1995 {\it Random Walks and Random Environments, Vol {I} Random
2772: Walks\/} (Oxford: Clarendon)
2773:
2774: \bibitem{Vanderzande}
2775: Vanderzande C 1998
2776: {\em Lattice Models of Polymers} Cambridge Lecture Notes in Physics 11
2777: (Cambridge: Cambridge University Press)
2778:
2779: \bibitem{deGennes} de Gennes P G 1979 {\it Scaling Concepts in Polymer Physics}
2780: (Ithaca: Cornell University Press)
2781:
2782: \bibitem{LSF87}
2783: Leibler S, Singh R~R~P and Fisher M~E 1987 Thermodynamic behavior of
2784: two-dimensional vesicles {\em Phys. Rev. Lett.\/} {\bf 59} 1989--1992
2785:
2786: \bibitem{GCbook}
2787: Gutman I and Cyvin S 1989 {\it Introduction to the Theory of Benzenoid Hydrocarbons}
2788: (Berlin: Springer)
2789:
2790: \bibitem{VGJ02}
2791: V\"oge M, Guttmann A~J and Jensen I 2002 On the number of benzenoid hydrocarbons,
2792: {\em J. Chem. Inf. Comput. Sci.\/} {\bf 42} 456--66
2793:
2794: \bibitem{FGW91}
2795: Fisher M~E, Guttmann A~J and Whittington S~G 1991,
2796: Two-dimensional lattice vesicles and polygons,
2797: {\em J.~Phys.~A:~Math.~Gen.} {\bf 24}, 3095--3106
2798:
2799: \bibitem{N82}
2800: Nienhuis B 1982 Exact critical point and critical exponents of O$(n)$ models in two dimensions
2801: {\em Phys. Rev. Letts.\/} {\bf 49} 1062--65
2802:
2803: \bibitem{N84}
2804: Nienhuis B 1984 Critical behavior of two-dimensional spin models and charge
2805: asymmetry in the coulomb gas {\em J. Stat. Phys.\/} {\bf 34} 731--761
2806:
2807: \bibitem{RGJ01}
2808: Richard C, Guttmann A~J and Jensen I 2001 Scaling function and universal amplitude
2809: combinations for self-avoiding polygons {\em J.~Phys.~A:~Math.~Gen. \/} {\bf 34} L495--501
2810:
2811: \bibitem{C01}
2812: Cardy J 2001
2813: Exact scaling functions for self-avoiding loops and branched polymers
2814: {\it J.~Phys.~A:~Math.~Gen.~\bf 34} L665--L672
2815:
2816: \bibitem{R02}
2817: Richard C 2002 Scaling behaviour of two-dimensional polygon models {\em J.
2818: Stat. Phys.\/} {\bf 108} 459--493
2819:
2820: \bibitem{RJG03}
2821: Richard C, Jensen I and Guttmann A~J 2003 Scaling function for self-avoiding
2822: polygons in {\em Proceedings of the International Congress on Theoretical
2823: Physics TH2002 (Paris), Supplement\/} (eds Iagolnitzer D, Rivasseau V and
2824: Zinn-Justin J) (Basel: Birkh\"auser) 267--277.
2825:
2826: \bibitem{RJG04}
2827: Richard C, Jensen I and Guttmann A~J 2004 Scaling function for self-avoiding
2828: polygons revisited {\em J. Stat. Mech.: Th. Exp.\/} P08007
2829:
2830: \bibitem{GJWE00}
2831: Guttmann A~J, Jensen I, Wong L~H and Enting I~G 2000 Punctured polygons and
2832: polyominoes on the square lattice {\em J. Phys. A: Math. Gen.\/} {\bf 33}
2833: 1735--1764
2834:
2835: \bibitem{JvRW89}
2836: Janse~van Rensburg E~J and Whittington S~G 1989 Self-avoiding surfaces
2837: {\em J. Phys. A: Math. Gen.\/} {\bf 22} 4939--4958
2838:
2839: \bibitem{GJO01}
2840: Guttmann A~J, Jensen I and Owczarek A~L 2001 Polygonal polyominoes on the square
2841: lattice {\em J. Phys. A: Math. Gen.\/} {\bf 34} 3721--3733
2842:
2843: \bibitem{J99}
2844: Janse van Rensburg E~J 1999
2845: Models of composite polygons
2846: {\em J. Phys. A: Math. Gen.\/} {\bf 32} 4351--4372
2847:
2848: \bibitem{BH75}
2849: Bleistein N and Handelsman Richard~A 1986
2850: {\em Asymptotic Expansions of Integrals} 2nd ed (New York: Dover)
2851:
2852: \bibitem{JvR92}
2853: Janse~van Rensburg E~J 1992 Surfaces in the hypercubic lattice
2854: {\em J. Phys. A: Math. Gen.\/} {\bf 25} 3529--3547
2855:
2856: \bibitem{GJ06b}
2857: Guttmann A~J and Jensen I 2006 The perimeter generating function of punctured
2858: staircase polygons {\em J. Phys. A: Math. Gen.\/} {\bf 39} 3871--3882
2859:
2860: \bibitem{OP99a}
2861: Prellberg T and Owczarek A 1999 On the asymptotics of the finite-perimeter
2862: partition function of two-dimensional lattice vesicles {\em Commun. Math.
2863: Phys.\/} {\bf 201} 493--505
2864:
2865: \bibitem{J00}
2866: Janse van Rensburg E J 2000
2867: {\it The Statistical Mechanics of Interacting Walks, Polygons,
2868: Animals and Vesicles}
2869: Oxford Lecture Series in Mathematics and its Applications
2870: {\bf 18} (Oxford: Oxford University Press)
2871:
2872: \bibitem{FS06}
2873: Flajolet P and Sedgewick R 2008 \textit{Analytic Combinatorics} in press
2874: (Cambridge: Cambridge University Press)
2875:
2876: \bibitem{BI03}
2877: Brydges D~C and Imbrie J~Z 2003 End-to-end distance from the {G}reen's function
2878: for a hierarchical self-avoiding walk in four dimensions {\em Commun. Math.
2879: Phys.\/} {\bf 239} 523--547
2880:
2881: \bibitem{Pr94}
2882: Prellberg T 1994 Uniform $q$-series asymptotics for staircase polygons
2883: {\em J. Phys. A: Math. Gen.\/} {\bf 28} 1289--1304
2884:
2885: \bibitem{L84}
2886: Louchard G 1984 The Brownian excursion area: a numerical analysis
2887: {\em Comput. Math. Appl. \bf 10} 413--417;
2888: Louchard G 1986
2889: Erratum: ``The Brownian excursion area: a numerical analysis"
2890: {\em Comput. Math. Appl. \bf 12} 375
2891:
2892: \bibitem{T91}
2893: Tak\'acs L 1991 On a probability problem connected with railway
2894: traffic {\em J.~Appl.~Math.~Stoch.~Anal. \bf 4} 1--27
2895:
2896: \bibitem{FL01}
2897: Flajolet P and Louchard G 2001 Analytic variations on the Airy
2898: distribution {\em Algorithmica \bf 31} 361--377
2899:
2900: \bibitem{C74}
2901: Chung K L 1974
2902: {\it A Course in Probability Theory}
2903: (New York: Academic Press) 2nd ed
2904:
2905: \bibitem{F99}
2906: Flajolet P 1999 Singularity analysis and asymptotics of Bernoulli sums
2907: {\em Theoret.~Comput.~Sci.} {\bf 215} 371--381
2908:
2909: \bibitem{R05}
2910: Richard C 2008 On $q$-functional equations and excursion moments
2911: {\em Discrete Math.}, in press; {\tt math.CO/0503198}
2912:
2913: \bibitem{R06}
2914: Richard C 2006 Staircase polygons: moments of diagonal lengths and column
2915: heights {\em J. Phys.: Conf. Ser.} {\bf 42} 239--257
2916:
2917: \bibitem{BM96}
2918: Bousquet-M\'elou M 1996
2919: A method for the enumeration of various classes of column-convex polygons
2920: {\em Discrete~Math.~\bf 154} 1--25
2921:
2922: \bibitem{BOP93}
2923: Brak R, Owczarek A L and Prellberg T 1993
2924: A scaling theory of the collapse transition in geometric cluster
2925: models of polymers and vesicles
2926: {\em J. Phys. A: Math. Gen.\/} {\bf 26} 4565--4579
2927:
2928: \bibitem{RG01}
2929: Richard C and Guttmann A~J 2001 $q$-linear approximants: scaling functions for
2930: polygon models {\em J. Phys. A: Math. Gen.\/} {\bf 34} 4783--4796
2931:
2932: \bibitem{IJ03}
2933: Jensen I 2003 A parallel algorithm for the enumeration of self-avoiding
2934: polygons on the square lattice {\em J. Phys. A: Math. Gen.\/} {\bf 36}
2935: 5731--5745
2936:
2937: \bibitem{IJ00} Jensen I 2000 Size and area of square lattice polygons
2938: {\em J. Phys. A: Math. Gen.\/} {\bf 33} 3533--3543
2939:
2940: \bibitem{AJG89a}
2941: Guttmann A~J 1989 Asymptotic analysis of power-series expansions in {\em Phase
2942: Transitions and Critical Phenomena\/} vol.~13 (eds Domb C and Lebowitz J L) (New
2943: York: Academic) 1--234
2944:
2945: \bibitem{IJ06a}
2946: Jensen I 2006 Honeycomb lattice polygons and walks as a test of series analysis
2947: techniques {\em J. Phys.: Conf. Ser.} {\bf 42} 163--178.
2948:
2949: \bibitem{JG99}
2950: Jensen I and Guttmann A~J 1999 Self-avoiding polygons
2951: on the square lattice {\em J. Phys. A: Math. Gen.\/} {\bf 32} 4867--4876
2952:
2953: \bibitem{IJ07}
2954: Jensen I 2006 Exact perimeter generating function for a model of punctured staircase polygons
2955: Preprint: {\tt cond-mat/0610605}
2956:
2957: \bibitem{AB06}
2958: Aleksandrowicz G and Barequet G 2006
2959: Counting $d$-dimensional polycubes and nonrectangular planar polyominoes
2960: in: {\it Proc. 12th Ann. Int. Computing and Combinatorics Conf. (COCOON), Taipei, Taiwan}
2961: Springer Lecture Notes in Computer Science {\bf 4112} 418--427
2962:
2963: \bibitem{W93}
2964: Whittington S~G 1993
2965: Statistical mechanics of three-dimensional vesicles
2966: {\it J.~Math.~Chem. \bf 14} 103--100
2967:
2968: \bibitem{C00}
2969: Cardy J 2000 Linking numbers for self-avoiding loops and percolation:
2970: application to the spin quantum hall transition {\em Phys.~Rev.~Lett.\/}
2971: {\bf 84} 3507--3510
2972:
2973: \bibitem{B74}
2974: Bender E~A 1974
2975: Asymptotic methods in enumeration
2976: {\em SIAM Rev. \bf 16} 485--515;
2977: Errata {\em SIAM Rev. \bf 18} (1976) 292
2978:
2979: \bibitem{R87}
2980: Rudin W 1987 {\em Real and Complex Analysis} (New York: McGraw-Hill)
2981:
2982: \end{thebibliography}
2983:
2984:
2985: \end{document}
2986: