1: \documentclass[pre,twocolumn,showpacs,amsmath,amssymb,sort]{revtex4}
2: %\documentclass[preprint,pre,amsmath,amssymb,showpacs,sort]{revtex4}
3:
4: \usepackage{graphicx}% Include figure files
5: \usepackage{dcolumn}% Align table columns on decimal point
6: \usepackage{bm}% bold math
7: \usepackage{verbatim}
8: \usepackage{subfigure}
9: \usepackage{dcolumn}
10: \usepackage{xspace}
11: %\usepackage{rotating}
12: %\usepackage{showkeys}
13: \usepackage{ifthen}
14:
15:
16: \newcommand{\pd}{\partial}
17: \newtheorem{conjecture}{Conjecture}
18: \newcommand{\myemph}{}
19: \newcommand{\md}{monomer-dimer\xspace}
20: \newcommand{\dd}{dimer density\xspace}
21: \newcommand{\p}{\rho}
22:
23: %\newcommand{\combine}{false}
24: \newcommand{\combine}{true}
25:
26: \begin{document}
27:
28: %\title{Accurate calculations of an intractable problem -- monomer-dimer model
29: %in two-dimensional rectangular lattices}
30: \title{Monomer-dimer model in two-dimensional rectangular lattices
31: with fixed dimer density}
32:
33: \author{Yong Kong}
34: \email{matky@nus.edu.sg}
35: \affiliation{%
36: Department of Mathematics\\
37: National University of Singapore\\
38: Singapore 117543\\
39: }%
40:
41: \date{\today}
42:
43: \begin{abstract}
44: The classical monomer-dimer model in two-dimensional lattices
45: has been shown to belong to the \emph{``\#P-complete''} class,
46: which indicates the problem is computationally ``intractable''.
47: We use exact computational method to
48: investigate the number of ways to arrange dimers on $m \times n$
49: two-dimensional rectangular lattice strips
50: with fixed \dd $\rho$. For any \dd $0 < \rho < 1$,
51: we find a
52: logarithmic correction term in the finite-size correction of the
53: free energy per lattice site.
54: The coefficient of the logarithmic correction term
55: is exactly $-1/2$.
56: This logarithmic correction term is explained by
57: %This logarithmic correction term fits nicely with
58: the newly developed asymptotic theory of Pemantle and Wilson.
59: %regardless of the parity of the lattice width $n$.
60: %This is in contrast with the situation reported earlier
61: %for fixed number of monomers (or vacancies),
62: %where the logarithmic correction coefficient
63: %depends on the number of monomers present and the
64: %parity of the width of the lattice strip.
65: The sequence of the free energy
66: of lattice strips with cylinder boundary condition
67: converges so fast that very accurate free energy $f_2(\rho)$ for large lattices
68: can be obtained.
69: For example, for a half-filled lattice,
70: $f_2(1/2) = 0.633195588930$, while $f_2(1/4) = 0.4413453753046$
71: and $f_2(3/4) = 0.64039026$.
72: For $\rho < 0.65$, $f_2(\rho)$ is accurate at least to 10 decimal digits.
73: The function $f_2(\rho)$ reaches the maximum value
74: $f_2(\rho^*) = 0.662798972834$ at $\rho^* = 0.6381231$,
75: with $11$ correct digits.
76: This is also the \md constant for two-dimensional rectangular lattices.
77: The asymptotic expressions of free energy near close packing
78: are investigated for finite and infinite lattice widths.
79: For lattices with finite width,
80: dependence on the parity of the lattice width
81: is found. For infinite lattices, the data support the functional
82: form obtained previously through series expansions.
83: %The data presented are the most accurate results for the \md problem
84: %up to date,
85: %and there are few intractable problems that have been calcualted to
86: %such high accuracy.
87:
88:
89: \end{abstract}
90:
91: \pacs{05.50.+q, 02.10.Ox, 02.70.-c}
92:
93: %\keywords{monomer dimer}
94:
95: \maketitle
96:
97:
98: \section{Introduction}
99:
100: The \md problem has received much attention not only from
101: statistical physics but also from theoretical computer science.
102: As one of the classical lattice statistical mechanics models,
103: the \md model was first used to describe
104: the absorption of a binary mixture of molecules of unequal sizes
105: on crystal surface
106: \cite{Fowler1937}.
107: In the model,
108: the regular lattice sites are either covered by monomers or dimers.
109: The diatomic molecules are modeled as rigid dimers
110: which occupy two adjacent sites in a regular lattice and no lattice site
111: is covered by more than one dimer.
112: The lattice sites that are not covered by the dimers
113: are regarded as occupied by monomers.
114: A central problem of the model is to enumerate the
115: dimer configurations on the lattice.
116: In 1961 an elegant analytical solution was found
117: for a special case of the problem,
118: namely when the planar lattice is completely covered by dimers
119: (the close-packed dimer problem, or dimer-covering problem)
120: \ifthenelse{\equal{\combine}{false}}{
121: \cite{Kasteleyn1961,Temperley1961,Fisher1961}}
122: {\cite{Kasteleyn1961,Temperley1961}}.
123: For the general \md problem
124: where there are vacancies (monomers) in the lattice,
125: there is no exact solution.
126: For three-dimensional lattices,
127: there is even no exact solution for the special case of
128: close-packed dimer problem.
129: One recent advance is an analytic solution to the
130: special case of the problem in two-dimensional lattices
131: where there is a single
132: vacancy at certain specific sites on the boundary
133: of the lattice
134: \ifthenelse{\equal{\combine}{false}}{
135: \cite{Tzeng2003,Wu2006,Wu2006b}}
136: {\cite{Tzeng2003,Wu2006}}.
137: The \md problem also serves as a prototypical problem in the field of
138: computational complexity
139: \ifthenelse{\equal{\combine}{false}}{
140: \cite{Garey1979,Welsh1993,Mertens2002}}
141: {\cite{Garey1979}}.
142: It has been shown that
143: two-dimensional \md problem belongs to
144: the \emph{``\#P-complete''} class
145: and hence is computationally intractable
146: \ifthenelse{\equal{\combine}{false}}{
147: \cite{Jerrum1987,Jerrum1990}}
148: {\cite{Jerrum1987}}.
149:
150:
151: Even though there is a lack of progress in the analytical solution
152: to the \md problem,
153: many rigorous results exist, such as
154: series expansions
155: \cite{Nagle1966,Gaunt1969,Samuel1980c},
156: lower and upper bounds on free energy
157: \ifthenelse{\equal{\combine}{false}}{
158: \cite{Bondy1966,Hammersley1968,Hammersley1970,Ciucu1998,Lundow2001,FriedlandP05}}
159: {\cite{Bondy1966,FriedlandP05}},
160: monomer-monomer correlation function of two monomers
161: in a lattice otherwise packed with dimers
162: \ifthenelse{\equal{\combine}{false}}{
163: \cite{Fisher1963,Hartwig1966}}
164: {\cite{Fisher1963}},
165: locations of zeros of partition functions,
166: \cite{Heilmann1972,Gruber1971},
167: and finite-size correction
168: \ifthenelse{\equal{\combine}{false}}{
169: \cite{Ferdinand1967,IzmailianOH03,Izmailian2005}}
170: {\cite{Ferdinand1967}}.
171: Some approximate methods have also been proposed
172: \ifthenelse{\equal{\combine}{false}}{
173: \cite{Chang1939,Baxter1968,Nemirovsky1989,LinL94,KenyonRS96,BeichlOS01}}
174: {\cite{Chang1939,Baxter1968,Nemirovsky1989,LinL94}}.
175: The \md constant $h_d$ (the exponential growth rate) of the
176: number of all configurations with different number of dimers
177: has also been calculated
178: \cite{Baxter1968,FriedlandP05}.
179: By using sequential importance sampling Monte Carlo method,
180: the dimer covering constant for a three-dimensional cubic lattice
181: has been estimated \cite{BeichlS99}.
182: The importance of \md model also comes from the fact that
183: there is one to one mapping between the Ising model and the \md
184: model: the Ising model in the absence of an external field is mapped
185: to the pure dimer model
186: \cite{Kasteleyn1963,Fisher1966,Fan1970,McCoy1973},
187: and the Ising model in the presence of an external field is mapped
188: to the general \md model \cite{Heilmann1972}.
189:
190:
191: The major purposes of this paper are (1) to show it is possible
192: to calculate accurately the free energy of the \md problem
193: in two-dimensional rectangular lattices at a fixed \dd
194: by using the proposed computational methods
195: (Sections \ref{S:method}, \ref{S:cylinder}, \ref{S:max}, and \ref{S:Baxter}),
196: and (2) to use the computational methods
197: to probe the physical properties of the \md model,
198: especially at the high \dd limit (Section \ref{S:hdd}).
199: The high \dd limit is considered to be more difficult
200: and more interesting than the low \dd limit.
201: The major result is the asymptotic expression Eq. \ref{E:f_asympt}.
202: The third purpose of the paper
203: is to introduce the asymptotic theory of Pemantle and Wilson
204: \cite{Pemantle2002}, which not only gives a theoretical explanation of
205: the origin of the logarithmic correction term found by
206: computational methods reported in this paper (Section \ref{S:lc}),
207: but also has the potential to be applicable to other statistical models.
208:
209: %\subsection{Notation and definitions}
210: The following notation and definitions will be used throughout the paper.
211: The configurational grand canonical partition function of the \md system
212: in a $m \times n$ two-dimensional lattice is
213: \begin{equation} \label{E:gpf}
214: Z_{m,n}(x) = a_{N}(m,n) x^N + a_{N-1}(m,n) x^{N-1} + \cdots + a_{0}(m,n)
215: \end{equation}
216: where $a_{s}(m,n)$ is the number of distinct ways to arrange $s$ dimers on the
217: $m \times n$ lattice, $N = \lfloor mn/2 \rfloor$,
218: and $x$ can be taken as the activity of a dimer.
219: The \emph{average}
220: number of sites covered by dimers (twice the average number of dimers)
221: of this grand canonical ensemble
222: is given by
223: \begin{equation} \label{E:theta-general}
224: \theta_{m,n}(x) = \frac{2}{mn} \frac{ \pd \ln Z_{m,n}(x) } {\pd \ln x}
225: = \frac{2}{mn}
226: \frac{ \sum_{s=1}^N a_s(m,n) s x^s } {\sum_{s=0}^N a_s(m,n) x^s} .
227: \end{equation}
228: The limit of this average for large lattices is denoted as $\theta(x)$:
229: %\begin{equation} \label{E:theta-general-large}
230: % \theta(x) = \lim_{m,n \rightarrow \infty} \theta_{m,n}(x) .
231: %\end{equation}
232: $\theta(x) = \lim_{m,n \rightarrow \infty} \theta_{m,n}(x)$.
233: In general we use $\theta_{d} (x)$
234: for the average number of sites covered by dimers
235: in a $d$-dimensional infinite lattice when the dimer activity is $x$.
236:
237: The total number of configurations of dimers is given by
238: $Z_{m,n}(1)$ at $x=1$, and
239: the \emph{monomer-dimer constant} for a two-dimensional infinite lattice
240: is defined as
241: \begin{equation} \label{E:h2_1}
242: h_2 = \lim_{m,n \rightarrow \infty} \frac{\ln Z_{m,n}(1)}{mn} .
243: \end{equation}
244: In general, we denote $h_d$ as the monomer-dimer constant
245: for a $d$-dimensional infinite lattice,
246: and $h_d(x)$ as the grand potential per lattice site
247: at any dimer activity $x$.
248: For a two-dimensional infinite lattice,
249: \begin{equation} \label{E:h2-general}
250: h_2(x) = \lim_{m,n \rightarrow \infty} \frac{\ln Z_{m,n}(x)}{mn}.
251: \end{equation}
252:
253: In this paper we focus on the number of dimer configurations
254: at a given \dd $\rho$.
255: In this sense
256: we are working on the \emph{canonical} ensemble.
257: %where the number of
258: %dimers $s$ (and hence the number of monomers $mn - 2s$) is fixed.
259: The connection between the canonical ensemble and
260: the grand canonical ensemble
261: is discussed in Appendix \ref{S:ee}.
262: We define the \dd for the canonical ensemble as the ratio
263: \begin{equation} \label{E:rho_0}
264: \rho = \frac{2s}{mn}.
265: \end{equation}
266: %so that in closed packed lattices, $\rho=1$.
267: When the lattice is fully covered by dimers, $\rho = 1$.
268: For a $m \times n$ lattice,
269: the number of dimers at a given \dd is
270: $
271: s = \frac{mn \rho}{2}
272: $.
273: In the following we use $a_{m,n}(\rho)$ as
274: the number of distinct dimer and monomer configurations at the given
275: \dd $\rho$.
276: By using this definition, Eq. \ref{E:gpf}
277: can be rewritten as
278: \begin{equation} \label{E:Z_rho}
279: Z_{m,n}(x) = \sum_{0 \le \rho \le 1} a_{m,n}(\rho) x^{mn\rho/2}.
280: \end{equation}
281:
282:
283:
284:
285: The free energy per lattice site at a given \dd $\rho$ is defined as
286: \[
287: f_{m,n} (\rho) = \frac{\ln a_{m,n}(\rho)}{mn}
288: \]
289: and the free energy at a given \dd for a semi-infinite lattice strip
290: $\infty \times n$ is
291: \[
292: f_{\infty, n}(\rho) = \lim_{m \rightarrow \infty} \frac{\ln a_{m,n}(\rho)}{mn}
293: = \lim_{m \rightarrow \infty} f_{m,n} (\rho) .
294: \]
295: For infinite lattices where both $m$ and $n$ go to infinity,
296: %$m, n \rightarrow \infty$:
297: the free energy is
298: \[
299: f_2(\rho) = f_{\infty, \infty}(\rho) =
300: \lim_{m, n \rightarrow \infty} \frac{\ln a_{m,n}(\rho)}{mn}
301: = \lim_{n \rightarrow \infty} f_{\infty, n}(\rho).
302: \]
303: We use the subscript $d$ in $f_d(\rho)$ to indicate
304: the dimension of the infinitely large lattice.
305: From the exact result
306: \ifthenelse{\equal{\combine}{false}}{
307: \cite{Kasteleyn1961,Temperley1961,Fisher1961}}
308: {\cite{Kasteleyn1961,Temperley1961}}
309: we know $f_2(\rho)$ at $\rho = 1$
310: \[
311: f_2(1) = \frac{G}{\pi} = 0.291560904
312: \]
313: where $G$ is the Catalan's constant.
314: For other values of $\rho > 0$, no analytical result is known,
315: although several bounds are developed
316: \ifthenelse{\equal{\combine}{false}}{
317: \cite{Bondy1966,Hammersley1968,Hammersley1970,Ciucu1998,Lundow2001,FriedlandP05}}
318: {\cite{Bondy1966,FriedlandP05}}.
319: We will show below that by using the exact calculation method
320: developed previously
321: \cite{Kong1999,Kong2006,Kong2006b},
322: we can calculate $f_2(\rho)$ at an arbitrary \dd $\rho$ with high accuracy.
323:
324: The article is organized as follows.
325: In Section \ref{S:method}, the computational method is outlined.
326: In Section \ref{S:lc}, we show a logarithmic correction term
327: in the finite-size correction of $f_{m,n} (\rho)$
328: for any fixed \dd $0 < \rho < 1$.
329: The coefficient of this logarithmic correction term
330: is exactly $-1/2$, for both cylinder lattices and lattices
331: with free boundaries.
332: We give a theoretical explanation for this logarithmic correction term
333: and its coefficient using the newly developed asymptotic theory
334: of Pemantle and Wilson \cite{Pemantle2002}.
335: In this section we point out the universality of this
336: logarithmic correction term with coefficient of $-1/2$.
337: This term is not unique to the \md model:
338: a large class of lattice models has this term
339: when the ``density'' of the models is fixed.
340: More discussions of applications of this asymptotic method
341: to the \md model in particular, and statistical models
342: in general, can be found in Section \ref{S:diss}.
343: In Section \ref{S:cylinder} we calculate $f_{\infty, n}(\rho)$
344: on lattice strips $\infty \times n$
345: for $n=1, \dots, 17$ with cylinder boundary condition.
346: The sequence of $f_{\infty, n}(\rho)$
347: on cylinder lattices converges very fast so that we can obtain
348: $f_2 (\rho)$ quite accurately.
349: To the best of our knowledge,
350: the results presented here are the most accurate
351: for \md problem in two-dimensional rectangular lattices
352: at an arbitrary \dd.
353: In Section \ref{S:fb} similar calculations of $f_{\infty, n}(\rho)$
354: are carried out on lattice strips $\infty \times n$ with free boundaries
355: for $n=1, \dots, 16$.
356: Compared with the sequence with cylinder boundary condition,
357: the sequence $f_{\infty, n}(\rho)$ with free boundaries
358: converges slower.
359: In Section \ref{S:max}, the position and values of the
360: maximum of $f_2(\rho)$ are located:
361: $f_2(\rho^*) = 0.662798972834$ at $\rho^* = 0.6381231$.
362: These results give an estimation of the \md constant
363: with $11$ correct digits.
364: The previous best result is with $9$ correct digits \cite{FriedlandP05}.
365: The results are also compared with those obtained
366: by series expansions and field theoretical methods.
367: The maximum value of $f_2(\rho)$ is equal to the two-dimensional \md
368: constant $h_2$.
369: This is one special case of the more general
370: relations between the calculated values in the canonical ensemble
371: and those in the grand canonical ensemble,
372: and these relations are further discussed
373: in Appendix \ref{S:ee}.
374: In Section \ref{S:Baxter},
375: the relations developed in Appendix \ref{S:ee}
376: are used to compare the results of the computational method
377: presented in this paper with
378: those of Baxter \cite{Baxter1968}.
379: For \md model, the more interesting properties
380: are at the more difficult high \dd limit.
381: In Section \ref{S:hdd} asymptotic behavior of the free energy
382: $f_{\infty, n}(\rho)$
383: is examined for high \dd near close packing.
384: For lattices with finite width, a dependence of the free energy
385: $f_{\infty, n}(\rho)$
386: on the parity of the lattice width $n$ is found (Eq. \ref{E:f_m_n}),
387: consistent with the previous results when the number of monomers is
388: fixed \cite{Kong2006b}.
389: The combination of the results in this section and those
390: of Section \ref{S:lc} leads to the asymptotic expression
391: Eq. \ref{E:f_asympt} for near close packing \dd.
392: The asymptotic expression of $f_2(\rho)$,
393: the free energy on an infinite lattice,
394: is also investigated near close packing.
395: The results support the functional forms obtained
396: previously through series expansions \cite{Gaunt1969},
397: but quantitatively the value of the exponent is lower than previously
398: conjectured.
399: In Appendix \ref{S:1d} we put together in one place
400: various explicit formulas for the one-dimensional lattices ($n=1$).
401: These formulas can be used to check the formulas developed
402: for the more general
403: situations where $n > 1$.
404: As an illustration, an explicit application of the
405: Pemantle and Wilson asymptotic method is also given for $n=1$.
406: %
407: %
408: %
409: %In the Appendices, extensive data are listed.
410: %
411: %
412: %
413:
414: \section{Computational methods} \label{S:method}
415:
416: The basic computational strategy is to use exact calculations
417: to obtain a series of partition functions $Z_{m,n}(x)$
418: of lattice strips $m \times n$.
419: Then for a given \dd $\p$, $f_{m, n}(\p)$ can be calculated
420: using arbitrary precision arithmetic.
421: By fitting $f_{m, n}(\p)$ to a given function (Sections \ref{S:cylinder},
422: \ref{S:fb}, and \ref{S:hdd}),
423: $f_{\infty, n}(\p)$ can be estimated with high accuracy.
424: From $f_{\infty, n}(\p)$, $f_2(\p)$ can then be estimated
425: using the special convergent properties of the sequence
426: $f_{\infty, n}(\p)$ on the cylinder lattice strips
427: (Section \ref{S:cylinder}).
428:
429: \subsection{Calculation of the partition functions} \label{SS:pf}
430:
431: The computational methods used here have been described in details
432: previously \cite{Kong1999, Kong2006, Kong2006b}.
433: The full partition functions (Eq. \ref{E:gpf})
434: are calculated recursively for lattice strips on cylinder lattices and lattices
435: with free boundaries.
436: As before,
437: all calculations of the terms $a_{s}(m,n)$
438: in the partition functions use exact integers,
439: and when logarithm is taken to calculate free energy $f_{m,n}(\rho)$,
440: the calculations are done with precisions much higher than the
441: machine floating-point precision.
442: The bignum library used is GNU MP library (GMP)
443: for arbitrary precision arithmetic
444: (version 4.2) \cite{gmp}.
445: The details of the calculations on lattices with free boundaries can be found
446: in Ref. \onlinecite{Kong2006}, so in the following only information
447: on cylinder lattices is given.
448:
449:
450: For a $m \times n$ lattice strip,
451: a square matrix $M_n$ is set up based on two rows of the lattice strip
452: with proper boundary conditions.
453: The vector $\Omega_m$, which consists of partition function of Eq. \ref{E:gpf}
454: as well as other contracted partition functions \cite{Kong1999},
455: is calculated by the following recurrence
456: \begin{equation} \label{E:M}
457: \Omega_m = M_n \Omega_{m-1}.
458: \end{equation}
459: Similar recursive method
460: has also been used for other
461: combinatorial problems, such as calculation of the number of
462: independent sets \cite{Calkin1998}.
463: For a cylinder lattice strip, the matrix $M_n$ is constructed
464: in a similar way as that with free boundaries \cite{Kong2006}.
465: The total valid number ($v_c(n)$) and unique number ($u_c(n)$) of
466: configurations are given respectively by the generating function
467: $\sum_n v_c(n) x^n = x(3+x-x^3)/(1-3x-x^2)/(x-1)/(x+1)$
468: and the formula
469: \[
470: u_c(n) = \sum_{ d \backslash n } \frac{\varphi(d) 2^{n/d}}{2n} +
471: \begin{cases}
472: 2^{(n-1)/2} & \text{if $n$ odd} \\
473: 2^{n/2-1}+2^{n/2-2} & \text{if $n$ even}
474: \end{cases}
475: \]
476: where $\varphi(m)$ is Euler's totient function,
477: which gives the number of integers relatively prime to integer $m$.
478: The size of matrix $M_n$ is $u_c(n) \times u_c(n)$.
479: The first $17$ terms of the sequence $v_c(n)$ are:
480: $3$, $10$, $36$, $118$, $393$, $1297$, $4287$, $14158$,
481: $46764$, $154450$, $510117$, $1684801$,
482: $5564523$, $18378370$, $60699636$, $200477278$, and $662131473$.
483: The first $17$ terms of the sequence $u_c(n)$ are:
484: $2$, $3$, $4$, $6$, $8$, $13$, $18$, $30$, $46$,
485: $78$, $126$, $224$, $380$, $687$, $1224$, $2250$, and $4112$.
486: It is noted that the sequence $u_c(n)$ is exactly the same as that
487: shown in column 2, Table 1 of Ref. \onlinecite{FriedlandP05}.
488: Calculations based on the dominant eigenvalues of the matrices
489: of the cylinder lattice strips for $n=4$, $6$, $8$, and $10$
490: are carried out by Runnels \cite{Runnels1970}.
491: The sizes of $M_n$ for cylinder lattice strips are smaller
492: when compared with the corresponding numbers for lattice strips
493: with free boundaries \cite{Kong2006},
494: which allows for calculations on wider lattice strips.
495: For cylinder lattice strips, full partition functions
496: are calculated for $n=1, \dots, 17$,
497: with length up to $m=1000$ for $n=1, \dots, 13$,
498: $m=880$ for $n=14$,
499: $m=669$ for $n=15$,
500: $m=474$ for $n=16$, and
501: $m=325$ for $n=17$.
502: The corresponding numbers for lattice strips with free boundaries
503: are reported in Ref. \onlinecite{Kong2006}.
504:
505:
506:
507: \subsection{Interpolation for arbitrary \dd $\p$} \label{SS:idd}
508:
509:
510: In this paper the main quantity to be calculated is $f_2(\rho)$.
511: The starting point of the calculations is the full partition function
512: Eq. \ref{E:gpf} for different values of $n$ and $m$.
513: Finite values of $n$ and $m$
514: only lead to discrete values of \dd $\p$, as defined in
515: Eq. \ref{E:rho_0}.
516: For example, when $n=7$ and $m=11$, the number of dimers $s$
517: takes the values of $0, 1, \dots, 38$,
518: and \dd $\p$ of this lattice can only be one of the following values:
519: $0, 2/77, \dots, 76/77$.
520: In general, for fixed finite $m$ and $n$,
521: $\p$ can only be a rational number:
522: \[
523: \rho = \frac{p}{q}
524: \]
525: where $p$ and $q$ are positive integers.
526: When $\rho$ is expressed as a rational number, the number of dimers
527: is given by
528: \begin{equation} \label{E:rho}
529: s = \frac{mn \rho}{2} = \frac{mnp}{2q}.
530: \end{equation}
531: This expression is only meaningful
532: if $mnp$ can be divided by $2q$.
533: When we write the grand canonical partition function in the form
534: of Eq. \ref{E:Z_rho} for finite $m$ and $n$,
535: we implicitly imply that Eq. \ref{E:rho} is satisfied.
536:
537:
538: In the following we use the rational \dd
539: $\rho = p/q$ whenever possible so that
540: the value of $a_{m,n}(\p)$ can be read directly from
541: the partition function of $m \times n$ lattices.
542: Depending on the values of $p$ and $q$,
543: some dimer densities, such as $\p=1/2$, can be realized in many lattices,
544: while others can only be realized in small number of lattices
545: with special combinations of values of $m$ and $n$.
546: %In general, the bigger $p$ and $q$,
547: %the fewer lattices that can realize the $\p$.
548: In many situations
549: it becomes impossible to use rational $\rho$.
550: For example,
551: in Section \ref{S:max}
552: the location of the maximum of $f_2(\rho)$
553: is searched within a very small region of $\rho$,
554: and in Section \ref{S:Baxter}, in order to compare the results from different
555: methods,
556: $\rho$ takes the output values of other
557: computational methods \cite{Baxter1968}.
558: In such situations,
559: if the rational form of $\p$ were used,
560: $p$ and $q$ would become so big that not enough data points
561: which satisfy Eq. \ref{E:rho} could be found for the fitting
562: in the $m \times n$ lattice strip.
563: To calculate $f_{m,n} (\rho)$
564: for an arbitrary real number $\p$ ($0 < \rho < 1$),
565: interpolation of the exact data points is needed.
566: Since full partition functions have been calculated for
567: fairly long lattice strips, proper interpolation procedure
568: can yield highly accurate values of $f_{m,n} (\rho)$
569: for an arbitrary real number $\rho$.
570: For interpolation, we use the standard
571: Bulirsch-Stoer rational function interpolation method
572: \cite{Stoer1980,Press1992}.
573: For any real number $\rho$, Eq. \ref{E:rho} is used to calculate
574: the corresponding number of dimers $s$, which may not be an integer.
575: On each side of this value of $s$, $30$ exact values of
576: $a_{s} (m,n)$ are used (if possible) in the interpolation.
577: If on one side there are not enough exact data points of $a_{s} (m,n)$,
578: extra data points on the other side of $s$ are used
579: to make the total number of
580: exact data points as $60$.
581: For the high \dd case (Section \ref{S:hdd}),
582: the total number of data points used is changed to $30$.
583: We also take care that no extrapolation is used:
584: if $\p$ is greater than the maximum \dd for a given $m \times n$ lattice,
585: the data point from this lattice is not used.
586: Let's look at the above example of $11 \times 7$ lattice again.
587: %As for the above example of the $11 \times 7$ lattice,
588: For this lattice,
589: the highest \dd is $76/77$.
590: If calculation is done for a \dd $\p=0.99$,
591: since $\p=0.99$ is greater than $76/77 \approx 0.987$,
592: the data point from this lattice will not be used in the following steps
593: to avoid inaccuracy introduced by unreliable extrapolations.
594:
595:
596:
597: \subsection{Fitting procedure} \label{SS:fit}
598:
599: The fitting experiments are carried out by using the ``fit'' function
600: of software {\sc gnuplot} (version 4.0) \cite{gnuplot40}
601: on a 64-bit Linux system.
602: %(kernel 2.6.12-1.1381\_FC3smp)
603: %with dual AMD Opteron processors.
604: The fit algorithm implemented is the nonlinear least-squares
605: (NLLS) Levenberg-Marquardt method \cite{Marquart1963}.
606: All fitting experiments use the default value $1$ as the initial value
607: for each parameter, and each fitting experiment is done
608: independently.
609: As done previously \cite{Kong2006,Kong2006b},
610: only those $a_{m,n} (\rho)$ with $m \ge 100$
611: are used in the fitting.
612: Since $a_{m,n} (\rho)$ is calculated for relatively long lattice strips
613: (in the $m$ direction, see Section \ref{SS:pf}),
614: the estimates of $f_{\infty, n}(\p)$
615: are usually quite accurate, up to $12$ or $13$ decimal place.
616: The accuracy for this fitting step
617: is limited by the machine floating-point precision,
618: since {\sc gnuplot} uses machine floating-point representations,
619: instead of arbitrary precision arithmetic.
620: We would have used the GMP library to implement a fitting program
621: with arbitrary precision arithmetic.
622: %as we did in the previous steps.
623: This would increase the accuracy in the estimation of $f_2(\p)$
624: when $\p$ is small.
625: For the major objective of this paper, i.e., to investigate the behavior
626: of $f_2(\p)$ when $\p \rightarrow 1$ (Section \ref{S:hdd}),
627: however, the current accuracy is adequate.
628: At high \dd limit,
629: the convergence of $f_{\infty, n}(\p)$ towards $f_{2}(\p)$
630: is much slower than
631: at low \dd limit.
632: With lattice width $n \le 17$ used for the current calculations,
633: $f_{\infty, n}(\p)$ is far from converging to the machine
634: floating-point precision when $\p \rightarrow 1$.
635:
636:
637: \section{Logarithmic corrections of the free energy
638: at fixed \dd} \label{S:lc}
639:
640: For lattice strips $m \times n$ with a fixed width $n$ and
641: a given \dd $\rho$,
642: the coefficients $a_{m,n}(\rho)$ of the partition functions
643: are extracted to fit the following function:
644: \begin{equation} \label{E:fit}
645: %\begin{multline} \label{E:fit}
646: f_{m,n} (\rho) =
647: \frac{\ln a_{m,n}(\rho)}{m n}
648: = c_0 + \frac{c_1}{m} + \frac{c_2}{m^2} +
649: \frac{c_3}{m^3} + \frac{c_4}{m^4} + \frac{\ell}{n} \frac{\ln(m+1)}{m}
650: \end{equation}
651: %\end{multline}
652: where $c_0 = f_{\infty, n} (\rho)$.
653:
654:
655: For both cylinder lattices and lattices with free boundaries,
656: the fitting experiments clearly show that
657: $\ell = -1/2$, accurate up to at least six decimal place,
658: for any \dd $0 < \rho < 1$.
659: This result holds for both odd $n$ and even $n$.
660: This is in contrast with the results reported earlier
661: for the situation with a fixed number of monomers (or vacancies),
662: where the logarithmic correction coefficient
663: depends on the number of monomers present and the
664: parity of the width of the lattice strip
665: \cite{Kong2006,Kong2006b}.
666: We notice that a coefficient $-1/2$ also appears
667: in the logarithmic correction term
668: of the free energy studied in Ref. \onlinecite{Tzeng2003},
669: which is a special case of the \md problem where there is a single
670: vacancy at certain specific sites on the boundary
671: of the lattice.
672:
673: For the general \md model,
674: to our best knowledge,
675: this logarithmic correction term with coefficient of exactly $-1/2$
676: has not been reported before in the literature.
677: The recently developed multivariate asymptotic theory
678: by Pemantle and Wilson \cite{Pemantle2002},
679: however, gives an explanation of this term and its coefficient.
680: This theory applies to combinatorial problems when the
681: multivariate generating function of the model is known.
682: For univariate generating functions,
683: asymptotic methods are well known and have been used for
684: a long time. The situation is quite different for
685: multivariate generating functions.
686: Until recently, techniques to get asymptotic expressions from
687: multivariate generating functions were ``almost entirely missing''
688: (for review, see Ref. \onlinecite{Pemantle2002}).
689: The newly developed Pemantle and Wilson method applies to a large class of
690: multivariate generating functions in a systematic way.
691: In general the theory applies to generating functions with multiple variables,
692: and for the bivariate case that we are interested in here,
693: the generating function of two variables takes the form
694: \begin{equation} \label{E:F}
695: F(x, y) = \frac{G(x, y)}{H(x, y)} = \sum_{s, m = 0}^{\infty} a_{sm} x^s y^m
696: \end{equation}
697: where $G(x, y)$ and $H(x, y)$ are analytic, and $H(0,0) \ne 0$.
698: In this case, Pemantle and Wilson method gives the asymptotic expression
699: as
700: \begin{equation} \label{E:asympt}
701: a_{sm} \sim \frac{G(x_0, y_0)} {\sqrt{2\pi}} x_0^{-s} y_0^{-m}
702: \sqrt{\frac{-y_0 H_y(x_0, y_0)}{m Q(x_0, y_0)}}
703: \end{equation}
704: where $(x_0, y_0)$ is the positive solution to the two equations
705: %\begin{align} \label{E:asympt_dir}
706: % H(x, y) &= 0 \\
707: % s x \frac{ \pd{H} } {\pd {x}} &= m y \frac{ \pd{H} } {\pd {y}} \notag
708: %\end{align}
709: \begin{equation} \label{E:asympt_dir}
710: H(x, y) = 0, \qquad
711: m x \frac{ \pd{H} } {\pd {x}} = s y \frac{ \pd{H} } {\pd {y}}
712: \end{equation}
713: and $Q(x, y)$ is defined as
714: \begin{widetext}
715: \[
716: -(x H_x) (y H_y)^2 -(y H_y) (x H_x)^2
717: - [(y H_y)^2 (x^2 H_{xx}) + (x H_x)^2 (y^2 H_{yy})
718: - 2 (x H_x) (y H_y)(xy H_{xy})].
719: \]
720: \end{widetext}
721: Here $H_x$, $H_y$, etc. are partial derivatives $\pd{H}/\pd{x}$,
722: $\pd{H}/\pd{y}$, and so on.
723: One of the advantages of the method over previous ones
724: is that the convergence of Eq. \ref{E:asympt} is \emph{uniform}
725: when $s/m$ and $m/s$ are bounded.
726:
727: For the \md model discussed here,
728: with $n$ as the finite width of the lattice strip,
729: $m$ as the length,
730: and $s$ as the number of dimers,
731: we can construct the bivariate generating function $F(x, y)$ as
732: \begin{equation} \label{E:bgf}
733: F(x, y) = \sum_{m=0}^{\infty} \sum_{s=0}^{mn/2} a_s(m,n) x^s y^m
734: = \sum_{m=0}^{\infty} Z_{m,n}(x) y^m.
735: \end{equation}
736: For the \md model, as well as a large class of lattice models
737: in statistical physics,
738: the bivariate generating function $F(x, y)$
739: is always in the form of Eq. (\ref{E:F}),
740: with $G(x, y)$ and $H(x, y)$ as polynomials in $x$ and $y$.
741: In fact, we can get $H(x, y)$ directly from matrix
742: $M_n$ in Eq. (\ref{E:M}).
743: It is closely related to the characteristic function of $M_n$ \cite{Kong1999}:
744: $H(x, y) = \det(M_n - I/y) \times y^w $, where $w$ is the size of the
745: matrix $M_n$.
746: As an illustration, the bivariate generating function
747: $F(x, y)$ for the one-dimensional lattice ($n=1$)
748: is shown in Eq. \ref{E:F_1} of Appendix \ref{S:1d}.
749:
750:
751: When the \dd is fixed, which is the case discussed here,
752: $s = \rho m n /2$.
753: If we substitute this relation into Eq. (\ref{E:asympt_dir}),
754: then we see that the solution $(x_0, y_0)$ of Eq. (\ref{E:asympt_dir})
755: only depends on $\rho$ and $n$, and does not depend on $m$ or $s$.
756: Substituting this solution $(x_0(\rho), y_0(\rho))$ into Eq. (\ref{E:asympt})
757: we obtain
758: \begin{widetext}
759: \begin{equation} \label{E:asympt_md}
760: f_{m,n} (\rho) \sim -\frac{1}{n} \ln(x_0^{\rho n /2} y_0)
761: -\frac{1}{2}\frac{\ln m}{mn}
762: + \frac{1}{mn} \ln \left( G(x_0, y_0)
763: \sqrt{\frac{-y_0 H_y(x_0, y_0)}{2\pi Q(x_0, y_0)}}
764: \right).
765: \end{equation}
766: \end{widetext}
767: From this asymptotic expansion we obtain the
768: logarithmic correction term with coefficient of $-1/2$ exactly,
769: for any value of $n$.
770: In fact, this asymptotic theory predicts that there exists
771: such a logarithmic correction term with coefficient of $-1/2$
772: for a large class of lattice models
773: when the two variables involved are proportional, that is,
774: when the models are at fixed ``density''.
775: For those lattice models which can be described
776: by bivariate generating functions,
777: this logarithmic correction term with coefficient of $-1/2$
778: is universal when those models are at fixed ``density''.
779: For the \md model, this proportional relation is for
780: $s$ and $m$ with $s = \rho m n /2$.
781: An explicit calculation for $n=1$ is shown in Appendix \ref{S:1d}.
782:
783: For a fixed \dd $\rho$ and a fixed lattice width $n$,
784: the first term of Eq. \ref{E:asympt_md} is
785: a constant and does not depend on $m$.
786: we identify it as $f_{\infty,n} (\rho)$
787: \begin{equation} \label{E:asympt_rho}
788: f_{\infty,n} (\rho) = -\frac{1}{n} \ln(x_0^{\rho n /2} y_0) .
789: \end{equation}
790:
791: In all the following fitting experiments, we set $\ell = -1/2$
792: for Eq. \ref{E:fit}.
793:
794:
795:
796:
797:
798: \section{Cylinder lattices} \label{S:cylinder}
799:
800: For the \md problem at a given \dd $\rho$ in cylinder lattice strips,
801: the sequence $f_{\infty, n}(\rho)$ converges very fast to
802: $f_2(\rho)$, as can be seen from a few sample data in Table \ref{T:c}.
803: In the table, values of $f_{\infty, n}(\rho)$ for $\rho = 1/4$, $1/2$,
804: $3/4$, and $1$ are listed.
805: Two obvious features can be observed:
806: (1) The function $f_{\infty, n}(\rho)$ is an increasing function of odd $n$,
807: but a decreasing function of even $n$.
808: Furthermore,
809: for finite integer values of $h$ and $k$,
810: \begin{equation} \label{E:inequ}
811: f_{\infty, 2h}(\rho) > f_2(\rho) > f_{\infty, 2k+1}(\rho).
812: \end{equation}
813: The value $f_{\infty, n}(\rho)$ oscillates around the limit value
814: $f_2(\rho)$
815: from even $n$ to odd $n$.
816: (2) The smaller the value of $\rho$, the faster the rate of convergence
817: of $f_{\infty, n}(\rho)$ towards $f_2(\rho)$.
818: Rational values of $\rho$ are used for the calculations in Table \ref{T:c}
819: and no interpolation of $a_{m,n}(\rho)$ is used.
820: The numbers of data points used in the fitting are listed in the
821: parentheses.
822:
823: As a check of the accuracy of the results, the data at $\rho=1$
824: can be compared with the exact solution.
825: For a cylinder lattice strip $\infty \times n$,
826: the exact expression for $f_{\infty, n}(1)$ reads
827: as \cite{Kasteleyn1961}
828: \begin{equation} \label{E:a_0_e}
829: f_{\infty, n}(1) = \frac{1}{n} \ln
830: %\left[
831: \prod_{i=1}^{n/2}
832: \left[ \sin \frac{(2i - 1)\pi}{n}
833: + \left(1 + \sin^2 \frac{(2i - 1)\pi}{n} \right)^{\frac{1}{2}}
834: \right]
835: %\right]
836: .
837: \end{equation}
838: The results from Eq. \ref{E:a_0_e} are listed as the last column
839: in Table \ref{T:c}.
840: As mentioned in the previous Section,
841: all input data are exact integers and the logarithm of these integers
842: is taken with high precision before the fitting.
843: The only places where accuracy can be lost are in the fitting procedure
844: as well as the approximation introduced by the fitting function
845: Eq. \ref{E:fit}.
846: Comparisons of the data in the last two columns
847: of Table \ref{T:c} show that,
848: as far as the fitting procedure is concerned,
849: the calculation accuracy is up to $12$ or $13$ decimal place.
850:
851: Another check for the accuracy of the fitting procedure
852: is through the exact expression Eq. \ref{E:r1} of one-dimensional
853: strip ($n=1$) at various \dd $\rho$.
854: The data are listed in the first row of Table \ref{T:c}.
855: By using Eq. (\ref{E:asympt_rho})
856: of the Pemantle and Wilson asymptotic method,
857: we can also compare
858: the fitting results with exact asymptotic values for small values
859: of $n$ (data not shown).
860: All these checks confirm consistently
861: that the accuracy of the fitting procedure
862: is up to $12$ or $13$ decimal place.
863: See Section \ref{S:diss} for further discussions on this issue.
864:
865: The fast convergence of $f_{\infty, n}(\rho)$ and
866: the property of Eq. \ref{E:inequ}
867: make it possible to obtain
868: $f_2(\rho)$ quite accurately, especially when $\rho$ is not too close
869: to $1$. Some of the values of $f_2(\rho)$ at
870: rational $\rho = p/q$ for small $p$ and $q$
871: are listed in Table \ref{T:list}.
872: As in Table \ref{T:c}, no interpolation of $a_{m,n}(\rho)$ is used.
873: The numbers in square brackets indicate the next digits
874: for $n=16$ (upper bound) and $n=17$ (lower bound).
875: The data show that when $\rho \le 0.65$,
876: the $f_2(\rho)$ is accurate up to at least $10$ decimal place.
877: It should be pointed out that the data listed are just raw data,
878: showing digits that have already converged
879: for $n=16$ and $n=17$.
880: If the pattern of convergence of these raw data is explored
881: and extrapolation technique is used, as is done in Section \ref{S:max},
882: it is possible to get even more correct digits.
883: As shown in Section \ref{S:max}, the true value of
884: $f_2(\rho)$ is not the average of $f_{\infty, 16}(\rho)$
885: and $f_{\infty, 17}(\rho)$. Instead, it should lie closer
886: to $f_{\infty, 17}(\rho)$.
887: %
888: %
889: %
890: %More data are listed in the tables in the Appendices.
891: %
892: %
893: %
894:
895:
896:
897: %\begin{comment}
898:
899: \begingroup
900: \squeezetable
901: \begin{table*}
902: \caption{
903: The coefficient $c_0$ ($f_{\infty, n}(\rho)$)
904: for different $n$ and $\rho$ on cylinder lattice strips $\infty \times n$.
905: The numbers in parentheses are the number of data points used in the fitting.
906: The first row for $n=1$ is from the exact expression Eq. \ref{E:r1}.
907: The last column is from the exact expression Eq. \ref{E:a_0_e}
908: when $\rho = 1$.
909: Rational $\rho$ is used here and no interpolation of $a_{m,n}(\rho)$ is used.
910: \label{T:c}}
911: \begin{ruledtabular}
912: \begin{tiny}
913: \begin{tabular}{llllll}
914: \input{latex_table_v2.1}
915:
916: \end{tabular}
917: \end{tiny}
918: \end{ruledtabular}
919: \end{table*}
920: \endgroup
921:
922: %\end{comment}
923:
924:
925:
926:
927: \begin{table}
928: \caption{
929: List of $f_2(\rho)$ for different $\rho$.
930: Numbers in square brackets indicate the next digits
931: for $n=16$ (upper bound) and $n=17$ (lower bound).
932: Rational $\rho$ is used here and no interpolation of $a_{m,n}(\rho)$ is used.
933: \label{T:list}}
934: \begin{ruledtabular}
935: \begin{tabular}{ll}
936: $\rho$ & $f_2(\rho)$ \\
937: \hline
938: 0 & 0 \\
939: 1/20 & 0.1334362263587 \\
940: 1/10 & 0.229899144084[8..9] \\
941: 3/20 & 0.310823643168[1..2] \\
942: 1/5 & 0.380638530252[1..2] \\
943: 1/4 & 0.4413453753046 \\
944: 3/10 & 0.4940275921700 \\
945: 1/3 & 0.525010031447[5..6] \\
946: 7/20 & 0.539305666744[5..6] \\
947: 2/5 & 0.5775208675757 \\
948: 9/20 & 0.6088200746799 \\
949: 1/2 & 0.633195588930[4..5] \\
950: 11/20 & 0.650499726669[5..8] \\
951: 3/5 & 0.66044120984[2..5] \\
952: 13/20 & 0.6625636470[2..4] \\
953: 2/3 & 0.661425713[7..8] \\
954: 7/10 & 0.65620036[0..1] \\
955: 3/4 & 0.64039026[3..5] \\
956: 4/5 & 0.6137181[3..4] \\
957: 17/20 & 0.573983[2..3] \\
958: 9/10 & 0.51739[1..2] \\
959: 19/20 & 0.435[8..9] \\
960: 1 & 0.29[0..3] \\
961:
962:
963: \end{tabular}
964: \end{ruledtabular}
965: \end{table}
966:
967:
968:
969:
970:
971:
972: \section{Lattices with free boundaries} \label{S:fb}
973:
974: We also carry out similar calculations for lattice strips
975: on $m \times n$ two-dimensional lattices with free boundaries,
976: for $n = 2, \dots, 16$.
977: A few sample data are shown in Table \ref{T:fb1}.
978: In the table, values of $f_{\infty, n}^{\text{fb}}(\rho)$
979: for $\rho = 1/4$, $1/2$,
980: $3/4$, and $1$ are listed.
981: The data in Table \ref{T:fb1} show that the sequence of
982: $f_{\infty, n}^{\text{fb}}(\rho)$ in lattices with free boundaries
983: converges slower than that in cylinder lattices.
984: Furthermore, in contrast to the situation in cylinder lattices,
985: $f_{\infty, n}^{\text{fb}}(\rho)$ is an increasing function of $n$
986: for $0 < \rho < 1$: $f_{\infty, n}^{\text{fb}}(\rho)$ approaches $f_2(\rho)$
987: (the same value as that for cylinder lattices) monotonically from below.
988: When $\rho = 1$, the functions $f_{\infty, 2k}^{\text{fb}}(1)$
989: and $f_{\infty, 2k+1}^{\text{fb}}(1)$
990: are increasing functions, with
991: $f_{\infty, 2k}^{\text{fb}}(1) > f_{\infty, 2k-1}^{\text{fb}}(1)$
992: and $f_{\infty, 2k}^{\text{fb}}(1) > f_{\infty, 2k+1}^{\text{fb}}(1)$.
993: Due to the slow convergence rate and the lack of property like
994: Eq. \ref{E:inequ},
995: it is difficult to obtain $f_2(\rho)$ reliably from the data
996: on the lattice strips with free boundaries.
997:
998: As we did in the previous Section, we also take advantage of
999: the known exact solution for $\rho=1$ as a check for the
1000: numerical accuracy of the fitting procedure.
1001: The exact result for lattice strips with free boundaries is given by
1002: \cite{Kasteleyn1961}
1003: \begin{equation} \label{E:a_0_e_fb}
1004: f_{\infty, n}^{\text{fb}}(1) = \frac{1}{n} \ln \left[ \prod_{i=1}^{\frac{n}{2}}
1005: \left( \cos \frac{i \pi}{n+1}
1006: + \left(1 + \cos^2 \frac{i \pi}{n+1} \right)^{\frac{1}{2}}
1007: \right) \right].
1008: \end{equation}
1009: The last column in Table \ref{T:fb1} lists the values given by
1010: Eq. \ref{E:a_0_e_fb}, which can be compared with the calculated values
1011: from the fitting experiments in the column next to it.
1012: Again, as shown in the previous Section, the
1013: accuracy at $\rho=1$ is up to $11$ or $12$ decimal place for most of the values
1014: of $n$.
1015:
1016:
1017: %\begin{comment}
1018:
1019:
1020: \begingroup
1021: \squeezetable
1022: \begin{table*}
1023: \caption{
1024: The coefficient $c_0$ ($f_{\infty, n}^{\text{fb}}(\rho)$)
1025: for different $n$ and $\rho$ on lattice strips $\infty \times n$
1026: with free boundaries.
1027: The numbers in parentheses are the number of data points used in the fitting.
1028: The last column is from the exact expression Eq. \ref{E:a_0_e_fb}
1029: when $\rho = 1$.
1030: Rational $\rho$ is used here and no interpolation of $a_{m,n}(\rho)$ is used.
1031: \label{T:fb1}}
1032: \begin{ruledtabular}
1033: \begin{tiny}
1034: \begin{tabular}{llllll}
1035:
1036: \input{latex_table_fb_v2.1}
1037:
1038: \end{tabular}
1039: \end{tiny}
1040: \end{ruledtabular}
1041: \end{table*}
1042: \endgroup
1043:
1044: %\end{comment}
1045:
1046:
1047:
1048:
1049:
1050: \section{Maximum of free energy and the \md constant} \label{S:max}
1051:
1052: It is well known that $f_d(\rho)$ is a continuous concave function of $\rho$
1053: and at certain \dd $\rho^*$, $f_d(\rho)$ reaches its maximum
1054: \cite{Hammersley1966}. However, there is no analytical knowledge of the
1055: location ($\rho^*$) and value ($f_d(\rho^*)$) of the maximum for $d > 1$.
1056: As is shown in Appendix \ref{S:ee},
1057: the maximum of $f_d(\rho)$
1058: is equal to the \md constant: $f_d(\rho^*) = h_d$.
1059: Currently the best value for $h_2$ is given in Ref. \onlinecite{FriedlandP05},
1060: which gives $h_2 = 0.6627989727 \pm
1061: 0.0000000001 $,
1062: with $9$ correct digits.
1063: The location of the maximum, $\rho^*$, is controversial.
1064: Baxter gives the value of $\rho^* = 0.63812311$ \cite{Baxter1968},
1065: while Friedland and Peled state, ``it is reasonable to assume
1066: that the value $p^*$, for which $\lambda_2(p^*) = h_2$, is fairly close to
1067: $p(2) = \frac{9-\sqrt{17}}{8} \approx 0.6096118$''
1068: (Here the original notation is used:
1069: $p$ is our $\rho$ and $\lambda_2(p)$ is our $f_2(\rho)$) \cite{FriedlandP05}.
1070:
1071: In this Section
1072: we use the same computational procedure described in the previous
1073: sections to locate accurately the maximum of $f_2(\rho)$.
1074: Using rational \dd $\rho = p/q$ and choose appropriate $p$ and $q$,
1075: we can locate the maximum to a fairly small region,
1076: as shown in Figure~\ref{F:max_1}.
1077:
1078: \begin{figure}
1079: \centering
1080: \includegraphics[angle=270,width=\columnwidth]{maximum.eps}
1081: \caption{(Color online)
1082: The function of $f_2(\rho)$ in the region of $19/20 \le \rho \le 9/14$.
1083: All data points use rational $\rho$, so no interpolation is used.
1084: \label{F:max_1}}
1085: \end{figure}
1086:
1087:
1088:
1089: With the interpolated data for $a_{m,n} (\rho)$,
1090: we can locate $\rho^*$ and $f_2(\rho^*)$ more accurately.
1091: As shown in Figures \ref{F:max_2} and \ref{F:max_3},
1092: we find that
1093: \[
1094: 0.662798972831 < f_2(\rho^*) < 0.662798972845,
1095: \]
1096: where the value of $f_{\infty, n}(\rho)$ for $n=16$ is used as the
1097: upper bounds, and that for $n=17$ as the lower bounds.
1098: From Figures \ref{F:max_2}, \ref{F:max_3} and \ref{F:max_4}
1099: we can locate $\rho^*$ as
1100: \[
1101: 0.63812310 < \rho^* < 0.63812312.
1102: \]
1103:
1104: The values of $f_{\infty, n}(\rho)$ around $\rho^*$ are listed in
1105: Table \ref{T:max}.
1106: Inspection of the convergent rate of these data for even and odd values
1107: of $n$ suggests that for both sequences, the convergent rate is geometric.
1108: If we assume that
1109: \begin{equation} \label{E:geometry}
1110: f_{\infty, n}(\rho) = f_2(\rho) - \alpha \eta^n,
1111: \end{equation}
1112: then the data points
1113: at $\rho = 0.63812311$ of $n=12$, $14$, and $16$ can be used
1114: to obtain an extrapolated value of $f_2(\rho) = 0.6627989728336$,
1115: while the data points of $n=13$, $15$, and $17$ give another
1116: extrapolation value $f_2(\rho) = 0.6627989728341$.
1117: Together these two extrapolation values converge to
1118: $f_2(\rho^*) = 0.662798972834$,
1119: with $11$ correct digits.
1120:
1121: We can also get the same conclusion graphically from
1122: Figure \ref{F:max_3}.
1123: By inspecting the pattern of the data points of different values of $n$
1124: in the inset of Figure \ref{F:max_3},
1125: we notice that the difference between the data points of $n=14$ and $n=16$
1126: is bigger than the difference between $n=15$ and $n=17$.
1127: This indicates that the true value of $f_2(\rho^*)$
1128: lies closer to the data point of $n=17$ than the data point of $n=16$.
1129: From Figure \ref{F:max_3} we are quite sure that the $11$-th
1130: digit of $f_2(\rho^*)$ is $3$ instead of $4$,
1131: and the $12$-th digit is probably $4$, as indicated by the
1132: two extrapolation values mentioned above.
1133:
1134:
1135: The value of $f_2(\rho^*)$ is in excellent agreement with that
1136: reported in Ref. \onlinecite{FriedlandP05},
1137: which gives $9$ correct digits
1138: (Friedland and Peled also guess correctly the $10$-th digits as $8$).
1139: The value also agrees with that in Ref. \onlinecite{Baxter1968},
1140: which gives $8$ correct digits \cite{FriedlandP05}.
1141: The value of $\rho^*$ is exactly that of Baxter \cite{Baxter1968}.
1142:
1143: By using field theoretical method, Samuel uses the following
1144: relation to transform the activity $x$ into a new variable $\omega$
1145: \cite{Samuel1980c}
1146: \begin{equation} \label{E:sam}
1147: x = \frac{\omega}{(1-4\omega)^2}.
1148: \end{equation}
1149: This relation is very close to the one used by Nagle \cite{Nagle1966}.
1150: By substituting this relation into Gaunt's series expansions \cite{Gaunt1969},
1151: Samuel obtained new series for various lattices, including the
1152: rectangular lattice (Eq. (5.12) of Ref. \onlinecite{Samuel1980c}).
1153: The value of \md constant in two-dimensional rectangular lattice
1154: can be calculated at $x=1$ by using his
1155: series as $0.66279914$. As we can see, this only gives
1156: five correct digits.
1157: Nagle used the following transform \cite{Nagle1966}
1158: \begin{equation} \label{E:nagle}
1159: x = \frac{\omega}{(1-3\omega)^2}.
1160: \end{equation}
1161: By using Gaunt's series, a value of $0.6627988$ is obtained, with
1162: six correct digits.
1163:
1164: %
1165: %
1166: %
1167: %More data for Figure \ref{F:max_3} are listed in Appendix \ref{A:max}.
1168: %
1169: %
1170: %
1171: \begingroup
1172: \squeezetable
1173: \begin{table*}
1174: \caption{
1175: The coefficient $c_0$ ($f_{\infty, n}(\rho)$)
1176: for different $n$ and $\rho$ on cylinder lattice strips $\infty \times n$
1177: around $\rho^*$.
1178: The numbers in parentheses are the number of data points used in the fitting.
1179: \label{T:max}}
1180: \begin{ruledtabular}
1181: \begin{tabular}{lllll}
1182:
1183: \input{latex_table_v6.1}
1184:
1185: \end{tabular}
1186: \end{ruledtabular}
1187: \end{table*}
1188: \endgroup
1189:
1190:
1191: \begin{figure}
1192: \centering
1193: \includegraphics[angle=270,width=\columnwidth]{maximum_6.eps}
1194: \caption{(Color online)
1195: The function of $f_{\infty, n}(\rho)$
1196: in the region of $0.638110 \le \rho \le 0.638130$ for $n=16$ (upper curve)
1197: and $n=17$ (lower curve).
1198: \label{F:max_2}}
1199: \end{figure}
1200:
1201: \begin{figure}
1202: \centering
1203: \includegraphics[angle=270,width=\columnwidth]{maximum_7_14_17.eps}
1204: \caption{(Color online)
1205: The function of $f_{\infty, n}(\rho)$
1206: in the region of $0.6381221 \le \rho \le 0.6381239$ for $n=16$
1207: (upper curve)
1208: and $n=17$ (lower curve).
1209: In the inset the data points from $n=14$ and $n=15$ are also shown.
1210: From top to bottom in the inset:
1211: $n=14$, $n=16$, $n=17$, and $n=15$.
1212: \label{F:max_3}}
1213: \end{figure}
1214:
1215: \begin{figure}
1216: \centering
1217: \includegraphics[angle=270,width=\columnwidth]{maximum_8.eps}
1218: \caption{(Color online)
1219: The function of $f_{\infty, n}(\rho)$
1220: in the region of $0.63812301 \le \rho \le 0.63812319$ for $n=16$.
1221: \label{F:max_4}}
1222: \end{figure}
1223:
1224:
1225: To conclude this section, we compare our results on the maximum of $f_2(\rho)$
1226: with the approximate
1227: formulas of Chang \cite{Chang1939} and Lin and Lai \cite{LinL94}.
1228: The Chang's approximate formula is given below
1229: %\begin{multline*}
1230: \[
1231: f_2^C(\rho) =
1232: -\frac{1}{2} [4 \ln 4 - (4-\rho) \ln(4-\rho)
1233: %\\
1234: + \rho \ln \rho
1235: + 2 (1-\rho) \ln (1-\rho) - 2\rho \ln 4],
1236: \]
1237: %\end{multline*}
1238: which gives $\rho^* = 0.634641$ and $f_2(\rho^*) = 0.661355$.
1239: The approximate formula of Lin and Lai is
1240: \[
1241: f_2^{\text{LL}}(\rho) =
1242: -\frac{\rho}{2} \ln \frac{\rho}{2} - 0.9030 (1-\rho) \ln(1-\rho)
1243: -0.05645 \rho
1244: \]
1245: which gives $\rho^* = 0.638057$ and $f_2(\rho^*) = 0.66282235$.
1246: Although the two formulas are quite simple, they give
1247: effective approximation with respect to $\rho^*$ and $f_2(\rho^*)$.
1248:
1249:
1250:
1251: \begin{comment}
1252:
1253: \begingroup
1254: \squeezetable
1255: \begin{table}
1256: \caption{\label{T:}}
1257: \begin{ruledtabular}
1258: \begin{tabular}{lll}
1259: 19/30 & 0.63333 & 0.662761073 \\
1260: 45/71 & 0.63380 & 0.6627681 \\
1261: 26/41 & 0.63415 & 0.662772835 \\
1262: 59/93 & 0.63441 & 0.66277617 \\
1263: 33/52 & 0.63462 & 0.662778631 \\
1264: 40/63 & 0.63492 & 0.662782015 \\
1265: 47/74 & 0.63514 & 0.66278421 \\
1266: 61/96 & 0.63542 & 0.662786858 \\
1267: 103/162 & 0.63580 & 0.66279006 \\
1268: 7/11 & 0.63636 & 0.662793850 \\
1269: 107/168 & 0.63690 & 0.662796516 \\
1270: 58/91 & 0.63736 & 0.66279802 \\
1271: 51/80 & 0.63750 & 0.66279833 \\
1272: 44/69 & 0.63768 & 0.662798649 \\
1273: 37/58 & 0.63793 & 0.662798912 \\
1274: 67/105 & 0.63810 & 0.66279897 \\
1275: 30/47 & 0.63830 & 0.662798922 \\
1276: 53/83 & 0.63855 & 0.6627987 \\
1277: 23/36 & 0.63889 & 0.662798001 \\
1278: 62/97 & 0.63918 & 0.6627971 \\
1279: 39/61 & 0.63934 & 0.662796501 \\
1280: 55/86 & 0.63953 & 0.66279567 \\
1281: 16/25 & 0.64000 & 0.662793131 \\
1282: 57/89 & 0.64045 & 0.66279 \\
1283: 41/64 & 0.64062 & 0.66278859 \\
1284: 41/64 & 0.64062 & 0.66278859 \\
1285: 25/39 & 0.64103 & 0.662785 \\
1286: 34/53 & 0.64151 & 0.662779939 \\
1287: 43/67 & 0.64179 & 0.6627766 \\
1288: 70/109 & 0.64220 & 0.66277135 \\
1289: 9/14 & 0.64286 & 0.662761744 \\
1290: \end{tabular}
1291: \end{ruledtabular}
1292: \end{table}
1293: \endgroup
1294:
1295: \end{comment}
1296:
1297:
1298:
1299: \section{Comparison with Baxter's results} \label{S:Baxter}
1300: Using variational approach, Baxter calculated $h_2(x)$
1301: (which is $\ln \kappa$ using his notation)
1302: and $\theta(x)$ (which is $2 \rho$ by his notation)
1303: for different values of dimer activity $x$ ($s^2$ by his notation).
1304: By using Eqs. \ref{E:theta} and \ref{E:h2_theta},
1305: we can compare our results
1306: with Baxter's results in his Table II.
1307: For each of his data point at a dimer activity $x$,
1308: we calculate $f_2(\rho)$ with $\rho = \theta(x)$.
1309: Then his $h_2(x)$ is converted to
1310: $f_2^B(\rho) = h_2(x) - \frac{\rho} {2} \ln(x)$.
1311: The comparisons are shown in Table \ref{T:baxter}.
1312: It should be pointed out that in Baxter's data,
1313: extrapolation is used for the sequence to obtain
1314: $\theta(x)$ and $h_2(x)$
1315: when $x^{-1/2}$ is small
1316: ($x^{-1/2} < 0.3$ for $h_2(x)$ and $x^{-1/2} < 0.5$ for $\theta(x)$),
1317: while no extrapolation is used
1318: in our data: we only look at the digits that have been converged
1319: for $n=16$ and $n=17$.
1320: Although the extrapolation used in Baxter's data makes
1321: the comparison less direct, we still see that
1322: the agreement is excellent.
1323: It seems that Baxter's method converges faster for $\rho$ very close to $1$
1324: (again the extrapolation factor has to be considered),
1325: and our method is more accurate when $\rho$ is not too close to $1$.
1326: As in Section \ref{S:cylinder},
1327: we only present the raw data here.
1328: If extrapolation is used,
1329: more correct digits can be obtained.
1330:
1331:
1332:
1333: \begin{table}
1334: \caption{
1335: Comparison with Baxter's results.
1336: Numbers in square brackets indicate the next digits
1337: for $n=16$ (upper bound) and $n=17$ (lower bound).
1338: \label{T:baxter}}
1339: \begin{ruledtabular}
1340: \begin{tabular}{llll}
1341: $x^{-1/2}$ & $\rho$ & $f_2(\rho)$ & $f_2^B(\rho)$ \\
1342: \hline
1343: 0.00 & 1.0 & 0.29[0..3] & 0.291557 \\
1344: 0.02 & 0.994176 & 0.319[2..8] & 0.3194631 \\
1345: 0.05 & 0.9836216 & 0.355[0..2] & 0.35510683 \\
1346: 0.10 & 0.96456376 & 0.4047[5..8] & 0.404771005 \\
1347: 0.20 & 0.924706050 & 0.4810[8..9] & 0.4810887477 \\
1348: 0.30 & 0.8846581140 & 0.536892[1..4] & 0.5368922350 \\
1349: 0.40 & 0.8453815864 & 0.5782845[2..9] & 0.5782845477 \\
1350: 0.50 & 0.8072764728 & 0.608814[3..4] & 0.6088143934 \\
1351: 0.60 & 0.7705280966 & 0.63085609[6..8] & 0.6308560970 \\
1352: 0.80 & 0.7013863228 & 0.655894637[3..5] & 0.6558946374 \\
1353: 1.00 & 0.6381231092 & 0.6627989728[3..4] & 0.6627989726 \\
1354: 1.50 & 0.5042633294 & 0.6349499289380[4..9]
1355: & 0.6349499290 \\
1356: 2.00 & 0.4006451804 & 0.5779686472227[1..4]
1357: & 0.5779686472 \\
1358: 2.50 & 0.3211782498 & 0.5140847735884[4..6]
1359: & 0.5140847737 \\
1360: 3.00 & 0.2603068980 & 0.4528361791290[7..9]
1361: & 0.4528361790 \\
1362: 3.50 & 0.2134739142 & 0.3978378948658[1..3]
1363: & 0.3978378949 \\
1364: 4.00 & 0.17715243204 & 0.3499573614350[0..2]
1365: & 0.3499573615 \\
1366: 4.50 & 0.14869898092 & 0.3088705309099[2..6]
1367: & 0.3088705306\\
1368: 5.00 & 0.126162903820 & 0.273811439807[1..2]
1369: & 0.2738114398 \\
1370:
1371: \end{tabular}
1372: \end{ruledtabular}
1373: \end{table}
1374:
1375:
1376:
1377:
1378:
1379:
1380:
1381:
1382: \section{High \dd near close packing} \label{S:hdd}
1383:
1384: It is well known that phase transition for the \md model
1385: only occurs at $\rho=1$ \cite{Heilmann1972}.
1386: Since the close-packed dimer system is at the critical point,
1387: it is interesting to investigate the behavior of the model
1388: when $\rho \rightarrow 1$.
1389: Using the similar computational procedure outlined before,
1390: the following results are obtained at high \dd limit:
1391: \begin{equation} \label{E:f_m_n}
1392: f_{\infty,n}(\rho) \sim
1393: f_{\infty, n}^{\text{lattice}} (1) +
1394: \begin{cases}
1395: -(1-\rho) \ln(1-\rho) & \text{$n$ is odd}\\
1396: -\frac{1}{2} (1-\rho) \ln(1-\rho) & \text{$n$ is even}\\
1397: \end{cases}
1398: \end{equation}
1399: where $f_{\infty, n}^{\text{lattice}} (1)$
1400: is the free energy of close-packed lattice with width $n$,
1401: and is given, based on the boundary condition,
1402: by Eq. \ref{E:a_0_e} (cylinder lattices)
1403: or Eq. \ref{E:a_0_e_fb} (lattices with
1404: free boundary condition).
1405: Eq. \ref{E:f_m_n} for $n=1$ is verified from the exact result
1406: as shown in Eq. \ref{E:f1}.
1407: The result is also confirmed for other values of $n$
1408: by using the Pemantle and Wilson asymptotic methods for multivariate
1409: generating function, as described in Section \ref{S:lc}.
1410: For space limitation these confirmations are not presented in this paper.
1411:
1412: The dependence of the asymptotic form of $f_{\infty,n}(\rho)$
1413: on the parity of the lattice width $n$
1414: as shown in Eq. \ref{E:f_m_n}
1415: reminds us of the results reported previously
1416: for \md model with fixed number of monomers $v$ \cite{Kong2006b},
1417: in which the coefficient of the logarithmic correction term
1418: of the free energy depends on the parity of the lattice width $n$.
1419: These two results are consistent with each other.
1420: If we substitute the relation $v = (1-\rho) mn$ into Eq. \ref{E:f_m_n},
1421: we will get the logarithmic correction term with coefficient $v$
1422: for odd $n$, and $v/2$ for even $n$.
1423: More discussions about this asymptotic form will be found
1424: in Section \ref{S:diss} (Eq. \ref{E:f_asympt}).
1425:
1426:
1427: We also investigate the behavior of $f_2(\rho)$ (for infinite lattice)
1428: as $\rho \rightarrow 1$.
1429: Since $f_{\infty,n}(\rho)$ does not converge fast enough
1430: as $\rho \rightarrow 1$ (Table \ref{T:c}),
1431: we use weighted average of $f_{\infty,16}(\rho)$
1432: and $f_{\infty,17}(\rho)$ as an approximation of $f_2(\rho)$.
1433: The weights are calculated from the exact results at $\rho=1$.
1434: Fitting these data to the following function
1435: \begin{equation} \label{E:f2_fit}
1436: f_2(\rho) = G/\pi + \frac{\gamma}{2} (1-\rho) \ln{(1-\rho)} + b_1 (1-\rho),
1437: \end{equation}
1438: we obtain $\gamma \approx 1.69775$ and $b_1 \approx 0.427832$.
1439: No other reasonable form of functions other than Eq. (\ref{E:f2_fit})
1440: gives better fit.
1441: Including a term of $(1-\rho)^2$ in Eq. \ref{E:f2_fit}
1442: leads to only slight changes in
1443: the values
1444: of $\gamma$ and $b_1$
1445: The data and the fitting result are shown in Figure \ref{F:f2}.
1446:
1447: \begin{figure}
1448: \centering
1449: \includegraphics[angle=270,width=\columnwidth]{f2_high_dd.eps}
1450: \caption{(Color online)
1451: Fitting result for $f_2(\rho)$ as $\rho \rightarrow 1$.
1452: \label{F:f2}}
1453: \end{figure}
1454:
1455: Using the equivalence between statistical ensembles
1456: discussed in Appendix \ref{S:ee}, we can relate our results
1457: with Gaunt's series expansions \cite{Gaunt1969}.
1458: Plugging $f_2(\rho)$ as in Eq. \ref{E:f2_fit} into
1459: $f(\rho) + \frac{\rho}{2} \ln(x)$ (see Eq. \ref{E:h2}),
1460: differentiating with respect to $\rho$, and solving for $\rho$,
1461: we obtain the average \dd $\theta(x)$ at the activity $x$.
1462: Expressing $x$ as a function of $\theta$, we have
1463: \[
1464: x = \frac{e^{2 b_1 - \gamma}}{(1-\theta)^\gamma}.
1465: \]
1466: This is in the same form of Eq. (3.7) of Gaunt \cite{Gaunt1969}.
1467: If we put in the values of $\gamma$ and $b_1$,
1468: we can estimate the amplitude $A = \exp(2 b_1 - \gamma) = 0.4308$.
1469: Gaunt obtains through series expansions $\gamma = 1.73 \pm 4$
1470: and $A = 0.3030 \pm 4$, and conjectures that $\gamma = 7/4$.
1471: Our results support the same functional form, and
1472: the numerical values are close to these obtained by Gaunt's
1473: series analysis.
1474: As for the conjectured value of $\gamma$,
1475: the current data seem to indicate a value lower than $7/4$.
1476: In fact, the data presented here as well as theoretical arguments
1477: (not shown here) indicate that $\gamma = 5/3$.
1478: More discussion on this constant can be found in the next Section.
1479:
1480:
1481: \section{Discussion}
1482: \label{S:diss}
1483:
1484: In Section \ref{S:lc} we show by computational methods that
1485: there is a logarithmic correction term in the free energy
1486: with a coefficient of $-1/2$.
1487: By introducing the newly developed Pemantle and Wilson asymptotic method,
1488: we give a theoretical explanation of this term.
1489: We also demonstrate that this term is not unique to the \md model.
1490: Many statistical lattice models can be casted in the form of
1491: bivariate (or multivariate) generating functions, and
1492: when the two variables are proportional to each other so that
1493: the system is at a fixed ``density'', we will expect
1494: such a universal logarithmic correction term with coefficient of $-1/2$.
1495: We anticipate more applications of this asymptotic method in
1496: statistical physics in the future.
1497:
1498: The Pemantle and Wilson asymptotic method not only gives
1499: a nice explanation of the logarithmic correction term and its
1500: coefficient found by computational means,
1501: but also gives \emph{exact} numerical values of
1502: $f_{\infty, n} (\rho)$
1503: for small $n$ (the width
1504: of the lattice strips).
1505: These exact values can be used to check the accuracy of the
1506: computational method, as already mentioned in Section \ref{S:cylinder}.
1507: In Section \ref{S:lc} we discuss how this can be done.
1508: The denominator $H(x, y)$ of the bivariate generating functions
1509: is derived from the characteristic function of the matrix $M_n$, and the
1510: size of $M_n$ is given by $u_c(n)$ in Section \ref{S:method} for
1511: cylinder lattices, and in Ref. \onlinecite{Kong2006} for lattices
1512: with free boundaries.
1513: For small $n$, the size of $M_n$ is small enough so that
1514: the characteristic function can be calculated symbolically.
1515: As $n$ increases, however, the size of $M_n$ increases exponentially:
1516: $u_c(17) = 4112$ for cylinder lattice when $n=17$,
1517: and $u_{\text{fb}}(16) = 32896$ when $n=16$.
1518: It is currently impractical to calculate the characteristic functions
1519: symbolically from matrices of such sizes to get $H(x, y)$
1520: of the corresponding bivariate generating functions,
1521: so the Pemantle and Wilson method cannot be applied
1522: when the width of the lattice $n$ becomes larger.
1523: Even when $H(x, y)$ is available,
1524: it is of the order of thousands or higher,
1525: which will lead to instabilities in the numerical calculations.
1526: The computational method utilized here, however,
1527: can still give important and accurate data
1528: in these situations.
1529:
1530: Previously we demonstrated that when the monomer number $v$ or
1531: the dimer number $s$ are fixed, there is also a
1532: logarithmic correction term in the free energy \cite{Kong2006,Kong2006b}.
1533: When the number of dimers is fixed (low \dd limit), the coefficient
1534: of this term equal to the number of dimers.
1535: When the number of monomers is fixed (high \dd limit),
1536: the coefficient, however, depends on the parity of the lattice
1537: width $n$: it equals to $v$ when $n$ is odd, and
1538: $v/2$ when $n$ is even.
1539: In this high \dd limit, as $m \rightarrow \infty$,
1540: \dd $\rho \rightarrow 1$.
1541: In this paper we focus on the situation where the \dd is fixed,
1542: and find that again there is a logarithmic correction term,
1543: but this time
1544: its coefficient equals to $-1/2$ and
1545: does not depend on the parity of the lattice width.
1546: Why does the dependence of the coefficient on the parity
1547: of the lattice width disappear
1548: as \dd $\rho \rightarrow 1$ and the lattice becomes almost completely covered
1549: by the dimers?
1550:
1551: This seemingly paradoxical phenomenon can be explained as follows.
1552: When the number of monomer $v$ is fixed and as $m \rightarrow \infty$,
1553: if we can put $v = (1-\rho)mn$ into Eq. \ref{E:f_m_n},
1554: then the term of $(1-\rho) \ln(1-\rho)$ leads to a term of $v \ln m /(mn)$
1555: when $n$ is odd, and a term of $v \ln m /(2mn)$
1556: when $n$ is even.
1557: At the same time, the logarithmic correction term
1558: with $-1/2$ as coefficient ($-\ln m /(2mn)$,
1559: the second term in Eq. \ref{E:asympt_md}),
1560: gets canceled out by the a term of $- \ln (1-\rho)/(2mn)$
1561: from the third term in Eq. \ref{E:asympt_md} as $m \rightarrow \infty$ and
1562: $\rho \rightarrow 1$.
1563: Putting Eqs. \ref{E:asympt_md} and \ref{E:f_m_n} together,
1564: we have for finite $n$, as $m \rightarrow \infty$ and
1565: $\rho \rightarrow 1$,
1566: \begin{widetext}
1567: \begin{equation} \label{E:f_asympt}
1568: f_{m, n}(\rho) \sim
1569: f_{\infty, n}^{\text{lattice}} (1) +
1570: \left\{
1571: \begin{tabular}{l}
1572: $-(1-\rho) \ln(1-\rho)$ \\
1573: $-\frac{1}{2} (1-\rho) \ln(1-\rho) $
1574: \end{tabular}
1575: - \frac{1}{2mn} \ln m
1576: - \frac{1}{2mn} \ln (1-\rho)
1577: \quad
1578: \begin{tabular}{l}
1579: $n$ is odd\\
1580: $n$ is even\\
1581: \end{tabular}
1582: \right.
1583: \end{equation}
1584: \end{widetext}
1585: This expression can be checked with the explicit formulas
1586: for one-dimensional lattice, Eqs. \ref{E:n1_1} and \ref{E:f1}.
1587: By using the relation $v=(1-\p)mn$, we see from Eq. \ref{E:f_asympt}
1588: that as $m \rightarrow \infty$,
1589: when the monomer number $v$ is fixed,
1590: the dependence of the logarithmic correction term
1591: on the parity of $n$ comes from the second term in the equation;
1592: the third and fourth terms cancel each other out.
1593: On the other hand
1594: when the \dd $\rho$ is fixed, the only logarithmic correction term
1595: comes from the third term of Eq. \ref{E:f_asympt}, with coefficient
1596: $-1/2$.
1597:
1598:
1599: As $n \rightarrow \infty$, $f_2(\rho)$ also has a term of
1600: $(1-\rho) \ln(1-\rho)$ (Eq. \ref{E:f2_fit}),
1601: whose coefficient
1602: is estimated as $-0.85$ (Section \ref{S:hdd}).
1603: This value is closer to $-5/6=-0.83$
1604: than to the conjectured value $-7/8=-0.875$ by Gaunt
1605: \cite{Gaunt1969}.
1606: Runnels' result, however, seems to be closer to Gaunt's result
1607: \cite{Runnels1970}.
1608: It should be recognized that,
1609: as pointed out previously
1610: \cite{Nagle1966,Baxter1968,Gaunt1969,Runnels1970,Samuel1980c} as well as
1611: in the present work, the convergence is poorest
1612: when the \dd is near close packing.
1613: %The discrepancy in the estimated values of $\gamma$ may just
1614: %be due to the errors introduced by the slow convergence.
1615: %To get a definite conclusion about
1616: %the true value of this coefficient,
1617: %further theoretical and computational
1618: %developments are needed.
1619: %investigations need to be
1620: %carried out.
1621: %It should be borne in mind, however, that the convergence
1622: %is slow when the \dd is near close packing,
1623: %and apprixmate weight factors have to be used in these calculations.
1624: On the other hand,
1625: theoretical calculations underway (not shown here due to space limitation)
1626: indeed
1627: indicate that this coefficient of $(1-\rho) \ln(1-\rho)$
1628: for the infinite lattice is $-5/6$,
1629: or equivalently, $\gamma = 5/3$.
1630:
1631:
1632: It is well known
1633: that there is a one to one correspondence
1634: between the Ising model in a rectangular lattice
1635: with zero magnetic field
1636: and a fully packed dimer model
1637: in a decorated lattice \cite{Kasteleyn1963,Fisher1966}.
1638: By using a similar method \cite{Heilmann1972},
1639: it has been shown that
1640: the Ising model in a non-zero magnetic field,
1641: a well-known unsolved problem in statistical mechanics,
1642: can be mapped to a \md model with \dd $\rho<1$.
1643: The investigation of the \md model near close packing
1644: is of interest within this context.
1645:
1646: As mentioned in Section \ref{S:max},
1647: several authors have applied field theoretic methods
1648: to analyze the \md model (for example, Ref. \onlinecite{Samuel1980c}).
1649: In such studies, the \md problem is expressed as a fermionic field theory.
1650: For close-packed dimer model, the expression is a free field theory
1651: with quadratic action, which is exactly solvable as expected.
1652: For general \md model, the expression is an interacting field theory with a quartic interaction term,
1653: and self-consistent Hartree approximation is used to improve the
1654: Feynman rules to derive the series expansions.
1655: The transforms that are obtained using these methods
1656: (such as Eq. \ref{E:sam}, which is similar to Eq. \ref{E:nagle})
1657: make the series expansions converge in the full range of the
1658: dimer activity.
1659: The accuracy of these calculations, however,
1660: is not comparable to the accuracy of the computational method
1661: reported here,
1662: %when $\rho$ is not very close to zero,
1663: possibly due to the limited length of the series expansion.
1664:
1665:
1666:
1667:
1668: \appendix
1669:
1670: \section{Equivalence of statistical ensembles} \label{S:ee}
1671:
1672: Throughout the paper our focus is on the functions $f_{m, n}(\rho)$,
1673: $f_{\infty, n}(\rho)$, or $f_2(\rho)$
1674: at a given \dd $\rho$.
1675: These functions are in essence properties of the canonical ensemble.
1676: In this Appendix we make the connection between $f_2(\rho)$
1677: and the functions of $\theta(x)$ and $h_2(x)$
1678: as defined in Eqs. \ref{E:theta-general} and \ref{E:h2-general},
1679: which are properties of the grand canonical ensemble.
1680: The results are used in
1681: Section \ref{S:hdd} to compare
1682: the results of near close packing \dd with Gaunt's series analysis, and in
1683: Section \ref{S:Baxter}
1684: to compare our results with those of Baxter, whose calculations are
1685: carried out in terms of $\theta(x)$ and $h_2(x)$ \cite{Baxter1968}.
1686:
1687: Suppose at $\rho = \rho^*$, the summand $a_{m,n}(\rho) x^{mn\rho/2}$
1688: in Eq.~\ref{E:Z_rho} reaches its maximum.
1689: By using the standard thermodynamic equivalence between
1690: different statistical ensembles
1691: \cite[Chap. 4]{Hill1987},
1692: we have
1693: \begin{align} \label{E:h2}
1694: h_2(x) &= \lim_{m,n \rightarrow \infty} \frac{\ln Z_{m,n}(x)}{mn} \notag \\
1695: &= \lim_{m,n \rightarrow \infty}
1696: \frac{\ln \sum_{0 \le \rho \le 1} a_{m,n}(\rho) x^{mn\rho/2}}{mn} \notag \\
1697: &= \lim_{m,n \rightarrow \infty}
1698: \frac{\ln a_{m,n}(\rho^*) x^{mn\rho^*/2}}{mn} \notag \\
1699: &= f_2(\rho^*) + \frac{\rho^*}{2} \ln x.
1700: \end{align}
1701: In other words, if we define $F_2(\rho, x) = f_2(\rho) + \frac{\rho}{2} \ln x$,
1702: then
1703: \begin{equation} \label{E:h2_ee}
1704: h_2(x) = \max_{0 \le \rho \le 1} ( f_2(\rho) + \frac{\rho}{2} \ln x)
1705: = \max_{0 \le \rho \le 1} F_2(\rho, x).
1706: \end{equation}
1707: As a special case, the \md constant is the maximum of the function
1708: $f_2(\rho)$ by setting $x=1$
1709: \begin{equation} \label{E:h2_1_ee}
1710: h_2 = \max_{0 \le \rho \le 1} f_2(\rho) = f_2(\rho^*).
1711: \end{equation}
1712: The connection for the average dimer coverage can also be obtained
1713: by using Eqs. \ref{E:theta-general} and \ref{E:h2} as
1714: \begin{equation} \label{E:theta}
1715: \theta(x) = \lim_{m,n \rightarrow \infty} \theta_{m,n}(x)
1716: = \lim_{m,n \rightarrow \infty}
1717: \frac{2}{mn} \frac{ \pd \ln Z_{m,n}(x) } {\pd \ln x}
1718: = \rho^*(x)
1719: \end{equation}
1720: with the understanding that at $\rho^*$,
1721: $F_2(\rho, x) = f_2(\rho) + \frac{\rho}{2} \ln x$,
1722: not $f_2(\rho)$,
1723: reaches its maximum.
1724: Substituting Eq. \ref{E:theta} into Eq. \ref{E:h2}, we obtain
1725: \begin{equation} \label{E:h2_theta}
1726: h_2(x) = f_2( \theta(x) ) + \frac{\theta(x)}{2} \ln x .
1727: \end{equation}
1728:
1729: In Section \ref{S:max},
1730: the excellent agreement of our result of $f_2(\rho^*)$ with the result on
1731: $h_2$ of Friedland and Peled \cite{FriedlandP05}
1732: has already been demonstrated.
1733: In Ref. \onlinecite{FriedlandP05}, what is calculated is actually
1734: $h_2$. Eq. \ref{E:h2_1_ee} makes it possible to compare our result
1735: with that of Friedland and Peled.
1736: Eq. \ref{E:h2_1_ee} is proved as a theorem for the specific \md problem
1737: in Ref. \onlinecite{FriedlandP05} as Theorem 4.1.
1738:
1739:
1740: Since there is an analytical solution to the one-dimensional lattice
1741: problem,
1742: Eqs. \ref{E:h2} and \ref{E:theta} can be confirmed
1743: for the one-dimensional lattice
1744: by explicit calculations, as shown in Appendix \ref{S:1d}.
1745:
1746:
1747:
1748:
1749: \section{Explicit results for one-dimensional lattice} \label{S:1d}
1750:
1751: In this Appendix we summarize some exact results
1752: for the one-dimensional lattice ($n=1$) which are useful to compare and
1753: check the results for lattices with width $n > 1$.
1754: When $n=1$, the problem is a special case of the so-called ``parking problem''
1755: in one-dimensional lattice
1756: in which a $k$-mer covers $k$ consecutive lattice sites in a non-overlapping
1757: way. Various methods exist which lead to closed form solutions to the general
1758: case of interacting $k$-mers (for example, see Refs.
1759: \ifthenelse{\equal{\combine}{false}}{
1760: \onlinecite{Zasedatelev1971,DiCera1996,Kong2001}}
1761: {\onlinecite{Zasedatelev1971,DiCera1996}}).
1762: For the \md model, $k=2$ and there is no interaction between
1763: the dimers.
1764: The number of ways to
1765: put $s$ dimers in the $m \times 1$ lattice is known as
1766: \begin{equation} \label{E:d1}
1767: a_{m, 1}(s) = \binom{m-s}{s} .
1768: \end{equation}
1769: From this expression, in the next subsection we derive the
1770: asymptotic expression of the
1771: free energy by using the traditional method.
1772: As an illustration, later we also give an explicit demonstration of
1773: Pemantle and Wilson's asymptotic method as it is applied
1774: to the bivariate generating function.
1775:
1776: \subsection{Canonical ensemble}
1777:
1778: From the explicit expression of Eq. \ref{E:d1},
1779: we can get the asymptotic expression of the
1780: free energy by using Stirling formula when $0 < \rho < 1$:
1781: \begin{widetext}
1782: \begin{align}
1783: & f_{m,1}(\rho) = \frac{\ln a_{m, 1}(s)}{m}
1784: = f_{\infty,1}(\rho)
1785: - \frac{1}{2m} \ln m
1786: + \frac{1}{2m}
1787: %\left[
1788: \ln \frac{2-\rho}{\rho (1-\rho)}
1789: %\right]
1790: \notag \\
1791: &
1792: + \sum_{j=1}^{\infty} \frac{2^{2j-2} B_{2j}}{ j(2j-1) m^{2j}}
1793: \left[ \frac{1}{(2-\rho)^{2j-1}} - \frac{1}{\rho^{2j-1}}
1794: -\frac{1}{2^{2j-1} (1-\rho)^{2j-1}}
1795: \right] \label{E:n1_1}
1796: \\
1797: &
1798: = f_{\infty,1}(\rho)
1799: - \frac{1}{2m} \ln (m+1) \label{E:n1_2}
1800: + \frac{1}{2m}
1801: %\left[
1802: \ln \frac{2-\rho}{\rho (1-\rho)}
1803: %\right]
1804: \notag \\&
1805: + \sum_{j=1}^{\infty} \frac{1}{m^{2j}}
1806: \left[ \frac{2^{2j-2} B_{2j}}{ j(2j-1) } \left(
1807: \frac{1}{(2-\rho)^{2j-1}} - \frac{1}{\rho^{2j-1}}
1808: -\frac{1}{2^{2j-1} (1-\rho)^{2j-1}}
1809: \right) + \frac{1}{2(2j-1)} \right]
1810: %\notag \\&
1811: -\sum_{j=1}^{\infty} \frac{1}{4j m^{2j+1}}
1812: \end{align}
1813: \end{widetext}
1814: where
1815: \begin{equation} \label{E:r1}
1816: f_{\infty,1}(\rho) = (1 - \frac{\rho}{2}) \ln (1 - \frac{\rho}{2})
1817: -\frac{\rho}{2} \ln \frac{\rho}{2}
1818: - (1-\rho) \ln(1-\rho)
1819: \end{equation}
1820: and $B_{2j}$ are the Bernoulli numbers.
1821:
1822: From Eqs. \ref{E:n1_1} or \ref{E:n1_2}
1823: it is evident that for $n=1$,
1824: the coefficient of the logarithmic correction
1825: term is $\ell = -1/2$ for $0 < \rho < 1$.
1826:
1827: The asymptotic expression of $f_{\infty,1}(\rho)$ (Eq. \ref{E:r1})
1828: at $\rho=1$ is given by
1829: \begin{equation} \label{E:f1}
1830: f_{\infty,1}(\rho) \sim
1831: -(1-\rho) \ln (1-\rho) - (\ln2-1)(1-\rho)
1832: - \sum_{i=1} \frac{(1-\rho)^{2i+1}}{2i(2i+1)}
1833: \end{equation}
1834: From this asymptotic expression we can see that the coefficient of
1835: $(1-\rho) \ln (1-\rho)$ is exactly $-1$,
1836: as in Eq. \ref{E:f_m_n} for odd values of $n$.
1837: By combining Eqs. \ref{E:n1_1} and \ref{E:f1} together
1838: we confirm Eq. \ref{E:f_asympt} for $n=1$
1839: at high \dd limit.
1840:
1841:
1842: \subsection{Grand canonical ensemble}
1843:
1844: In this section we calculate various quantities associated
1845: with the grand canonical ensemble.
1846: The configurational grand canonical partition function (Eq. \ref{E:gpf}) is
1847: \[
1848: Z_{m,1}(x) = \sum_{s=0}^{m/2} \binom{m-s}{s} x^s
1849: \]
1850: To get a closed form of $Z_{m,1}(x)$, we use the WZ method
1851: (Wilf-Zeilberger) \cite{Petkovsek1996}
1852: to get the following recurrence of $Z_{m,1}(x)$
1853: \[
1854: x Z_{m,1}(x) + Z_{m+1,1}(x) - Z_{m+2,1}(x) = 0,
1855: \]
1856: from which we obtain the closed form solution as
1857: \begin{equation} \label{E:Z_1d}
1858: Z_{m,1}(x) =
1859: \frac{1}{\sqrt{1+4x}}
1860: \left[ \beta_1^{m+1} - \beta_2^{m+1}
1861: \right]
1862: \end{equation}
1863: where
1864: \[
1865: \beta_{1,2} = \frac{1 \pm \sqrt{1+4x}}{2}.
1866: \]
1867: To calculate $\theta(x)$ using Eq. \ref{E:theta-general},
1868: we need to evaluate the sum
1869: \[
1870: S(m) = \sum_{s=0}^{m/2} \binom{m-s}{s} s x^s.
1871: \]
1872: Again by using WZ method, we obtain the following recurrence for
1873: $S(m)$
1874: \[
1875: (m+2) x S(m) + (m+1) S(m+1) - m S(m+2) = 0.
1876: \]
1877: To solve this recurrence, we use the generating function of $S(m)$:
1878: $G_S(z) = \sum_{m} S(m) z^m$, and get $G_S(z)$ from the recurrence as
1879: \[
1880: G_S(z) = \frac{x z^2}{ (1-z-xz^2)^2}.
1881: \]
1882: From $G_S(z)$ a closed form expression of $S(m)$ can be found as
1883: %\begin{widetext}
1884: \begin{equation} \label{E:F_1d}
1885: S(m) = \frac{x}{1+4x}
1886: \left[
1887: (m - \frac{1}{\sqrt{1+4x}}) \beta_1^{m}
1888: + (m + \frac{1}{\sqrt{1+4x}}) \beta_2^{m}
1889: \right]
1890: \end{equation}
1891: %\end{widetext}
1892: Substituting Eqs. \ref{E:Z_1d} and \ref{E:F_1d} into Eq. \ref{E:theta-general},
1893: we obtain
1894: \begin{equation} \label{E:theta_1d}
1895: \theta_1(x) = 1 - \frac{1}{\sqrt{1+4x}}
1896: \end{equation}
1897: Using Eq. \ref{E:Z_1d} we can calculate $h_1(x)$ as
1898: \begin{equation} \label{E:h_1}
1899: h_1(x) = \lim_{m \rightarrow \infty} \frac{\ln Z_{m,1}(x)}{m}
1900: = \ln \frac{1+\sqrt{1+4x}}{2}
1901: \end{equation}
1902:
1903: It is known that there are multiple methods to solve
1904: the one-dimensional lattice model. For example, transform matrix method
1905: can also be used \cite{Wu2006c}. In this case, the transform matrix is
1906: \[
1907: T_1 =
1908: \begin{bmatrix}
1909: 0 & x \\
1910: 1 & 1
1911: \end{bmatrix},
1912: \]
1913: whose eigenvalues are $\lambda_{1,2} = (1 \pm \sqrt{1+4x})/2$.
1914: As $m \rightarrow \infty$, $Z_{m,1}(x) \sim \lambda_{1}^m$, so we obtain
1915: $h_1(x)$ as Eq. \ref{E:h_1}.
1916: From $\theta_1(x) = 2 \pd \ln \lambda_{1} / \pd \ln x$ we obtain
1917: Eq. \ref{E:theta_1d}.
1918:
1919:
1920:
1921: \subsection{Equivalence of statistical ensembles}
1922:
1923: The confirmation of the equivalence of ensembles
1924: for the special case of $x=1$ (Eq. \ref{E:h2_1_ee})
1925: has been done in Ref. \onlinecite{FriedlandP05}.
1926: Here we carry out the explicit calculations for the
1927: general case of arbitrary dimer activity $x$.
1928:
1929:
1930: If we take the derivative of function
1931: $F_1(\rho, x) = f_1(\rho) + \frac{\rho}{2} \ln x$,
1932: where $f_1(\rho) = f_{\infty, 1}(\rho)$ is given in Eq. \ref{E:r1},
1933: solve for $\rho$, and retain only the solution in $[0, 1]$, we have
1934: \[
1935: \rho^* = 1 - \frac{1}{\sqrt{1+4x}}
1936: \]
1937: which is the same as Eq. \ref{E:theta_1d}.
1938: If we put the value of $\rho^*$ into $F_1(\rho, x)$,
1939: we obtain the maximum of $F_1(\rho^*, x)$:
1940: \[
1941: \max_{0 \le \rho \le 1} ( f_1(\rho) + \frac{\rho}{2} \ln x)
1942: = \ln \frac{1+\sqrt{1+4x}}{2}
1943: = h_1(x)
1944: \]
1945:
1946: \subsection{Application of Pemantle and Wilson asymptotic method
1947: to the bivariate generating function}
1948:
1949: The bivariate generating function of Eq, \ref{E:bgf}
1950: can also be obtained in multiple ways,
1951: for example by direct summation of Eq. \ref{E:Z_1d},
1952: or by using the characteristic function of $M_1$,
1953: or by using Eq. (23) of Ref. \onlinecite{DiCera1996}
1954: (by setting the interaction parameter $\sigma = 1$ and the size of $k$-mer as 2),
1955: to get
1956: \begin{equation} \label{E:F_1}
1957: F_1(x, y) = \frac{1}{1-y-xy^2}.
1958: \end{equation}
1959: Here $G(x, y) = 1$ and $H(x, y) = 1-y-xy^2$.
1960: Solving the two equations in Eq. \ref{E:asympt_dir} we get
1961: $(x_0, y_0)$ as
1962: $(x_0 = s(m-s)/(m-2s)^2, y_0 = (m-2s)/(m-s) )$.
1963: Substituting $(x_0, y_0)$ into Eq. \ref{E:asympt} we obtain
1964: \[
1965: a_{m,1} \sim \frac{1}{\sqrt{ 2 \pi}}
1966: (m-s)^{m-s + 1/2}
1967: (m-2s)^{-m+2s-1/2} (s)^{-s-1/2}.
1968: \]
1969: By putting $s = \rho m /2$, the first three terms of Eq. \ref{E:n1_1}
1970: are recovered, including the term of logarithmic correction
1971: $-\ln m/(2m)$. Higher order terms can also be obtained if more terms of
1972: the asymptotic expressions are used \cite{Pemantle2002}.
1973:
1974:
1975: \ifthenelse{\equal{\combine}{true}}{
1976:
1977: \begin{thebibliography}{99}
1978: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1979: \expandafter\ifx\csname bibnamefont\endcsname\relax
1980: \def\bibnamefont#1{#1}\fi
1981: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1982: \def\bibfnamefont#1{#1}\fi
1983: \expandafter\ifx\csname citenamefont\endcsname\relax
1984: \def\citenamefont#1{#1}\fi
1985: \expandafter\ifx\csname url\endcsname\relax
1986: \def\url#1{\texttt{#1}}\fi
1987: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1988: \providecommand{\bibinfo}[2]{#2}
1989: \providecommand{\eprint}[2][]{\url{#2}}
1990:
1991: \bibitem[{\citenamefont{Fowler and Rushbrooke}(1937)}]{Fowler1937}
1992: \bibinfo{author}{\bibfnamefont{R.~H.} \bibnamefont{Fowler}} \bibnamefont{and}
1993: \bibinfo{author}{\bibfnamefont{G.~S.} \bibnamefont{Rushbrooke}},
1994: \bibinfo{journal}{Trans. Faraday Soc.} \textbf{\bibinfo{volume}{33}},
1995: \bibinfo{pages}{1272} (\bibinfo{year}{1937}).
1996:
1997: \bibitem[{\citenamefont{Kasteleyn}(1961)}]{Kasteleyn1961}
1998: \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Kasteleyn}},
1999: \bibinfo{journal}{Physica} \textbf{\bibinfo{volume}{27}},
2000: \bibinfo{pages}{1209} (\bibinfo{year}{1961}).
2001:
2002: \bibitem[{\citenamefont{Temperley and Fisher}(1961)}]{Temperley1961}
2003: \bibinfo{author}{\bibfnamefont{H.~N.~V.} \bibnamefont{Temperley}}
2004: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~E.}
2005: \bibnamefont{Fisher}}, \bibinfo{journal}{Philos. Mag.}
2006: \textbf{\bibinfo{volume}{6}}, \bibinfo{pages}{1061} (\bibinfo{year}{1961});
2007: %
2008: %\bibitem[{\citenamefont{Fisher}(1961)}]{Fisher1961}
2009: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Fisher}},
2010: \bibinfo{journal}{Phys. Rev.} \textbf{\bibinfo{volume}{124}},
2011: \bibinfo{pages}{1664} (\bibinfo{year}{1961}).
2012:
2013: \bibitem[{\citenamefont{Tzeng and Wu}(2003)}]{Tzeng2003}
2014: \bibinfo{author}{\bibfnamefont{W.-J.} \bibnamefont{Tzeng}} \bibnamefont{and}
2015: \bibinfo{author}{\bibfnamefont{F.~Y.} \bibnamefont{Wu}}, \bibinfo{journal}{J.
2016: Stat. Phys.} \textbf{\bibinfo{volume}{110}}, \bibinfo{pages}{671}
2017: (\bibinfo{year}{2003}).
2018:
2019: \bibitem[{\citenamefont{Wu}(2006{\natexlab{a}})}]{Wu2006}
2020: \bibinfo{author}{\bibfnamefont{F.~Y.} \bibnamefont{Wu}},
2021: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{74}},
2022: \bibinfo{pages}{020104} (\bibinfo{year}{2006}{\natexlab{a}});
2023: %
2024: %\bibitem[{\citenamefont{Wu}(2006{\natexlab{b}})}]{Wu2006b}
2025: %\bibinfo{author}{\bibfnamefont{F.~Y.} \bibnamefont{Wu}},
2026: % \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{74}},
2027: % \bibinfo{pages}{039907} (\bibinfo{year}{2006}{\natexlab{b}}).
2028: \textbf{\bibinfo{volume}{74}},
2029: \bibinfo{pages}{039907(E)}
2030: (\bibinfo{year}{2006}).
2031:
2032: \bibitem[{\citenamefont{Garey and Johnson}(1979)}]{Garey1979}
2033: \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Garey}} \bibnamefont{and}
2034: \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Johnson}},
2035: \emph{\bibinfo{title}{Computers and Intractability, A Guide to the Theory of
2036: {NP}-Completeness}} (\bibinfo{publisher}{W.H. Freeman and Company},
2037: \bibinfo{address}{New York}, \bibinfo{year}{1979});
2038: %
2039: %\bibitem[{\citenamefont{Welsh}(1993)}]{Welsh1993}
2040: \bibinfo{author}{\bibfnamefont{D.~J.~A.} \bibnamefont{Welsh}},
2041: \emph{\bibinfo{title}{Complexity: Knots, Colourings, and Counting}}, vol.
2042: \bibinfo{volume}{186} of \emph{\bibinfo{series}{London Mathematical Society
2043: Lecture Note Series}} (\bibinfo{publisher}{Cambridge University Press},
2044: \bibinfo{year}{1993});
2045: %
2046: %\bibitem[{\citenamefont{Mertens}(2002)}]{Mertens2002}
2047: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Mertens}},
2048: \bibinfo{journal}{Computing in Science and Engineering}
2049: \textbf{\bibinfo{volume}{4}}, \bibinfo{pages}{31} (\bibinfo{year}{2002}).
2050:
2051: \bibitem[{\citenamefont{Jerrum}(1987)}]{Jerrum1987}
2052: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Jerrum}}, \bibinfo{journal}{J.
2053: Stat. Phys.} \textbf{\bibinfo{volume}{48}}, \bibinfo{pages}{121}
2054: (\bibinfo{year}{1987});
2055: %
2056: %\bibitem[{\citenamefont{Jerrum}(1990)}]{Jerrum1990}
2057: %\bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Jerrum}}, \bibinfo{journal}{J.
2058: % Stat. Phys.} \textbf{\bibinfo{volume}{59}}, \bibinfo{pages}{1087}
2059: % (\bibinfo{year}{1990}).
2060: \textbf{\bibinfo{volume}{59}},
2061: \bibinfo{pages}{1087(E)}
2062: (\bibinfo{year}{1990}).
2063:
2064: \bibitem[{\citenamefont{Nagle}(1966)}]{Nagle1966}
2065: \bibinfo{author}{\bibfnamefont{J.~F.} \bibnamefont{Nagle}},
2066: \bibinfo{journal}{Phys. Rev.} \textbf{\bibinfo{volume}{152}},
2067: \bibinfo{pages}{190} (\bibinfo{year}{1966}).
2068:
2069: \bibitem[{\citenamefont{Gaunt}(1969)}]{Gaunt1969}
2070: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Gaunt}},
2071: \bibinfo{journal}{Phys. Rev.} \textbf{\bibinfo{volume}{179}},
2072: \bibinfo{pages}{174} (\bibinfo{year}{1969}).
2073:
2074: \bibitem[{\citenamefont{Samuel}(1980)}]{Samuel1980c}
2075: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Samuel}}, \bibinfo{journal}{J.
2076: Math. Phys.} \textbf{\bibinfo{volume}{21}}, \bibinfo{pages}{2820}
2077: (\bibinfo{year}{1980}).
2078:
2079: \bibitem[{\citenamefont{Bondy and Welsh}(1966)}]{Bondy1966}
2080: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Bondy}} \bibnamefont{and}
2081: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Welsh}},
2082: \bibinfo{journal}{Proc. Camb. Phil. Soc. Math. Phys. Sci.}
2083: \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{503} (\bibinfo{year}{1966});
2084: %
2085: %\bibitem[{\citenamefont{Hammersley}(1968)}]{Hammersley1968}
2086: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hammersley}},
2087: \bibinfo{journal}{Proc. Camb. Phil. Soc. Math. Phys. Sci.}
2088: \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{455} (\bibinfo{year}{1968});
2089: %
2090: %\bibitem[{\citenamefont{Hammersley and Menon}(1970)}]{Hammersley1970}
2091: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hammersley}} \bibnamefont{and}
2092: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Menon}}, \bibinfo{journal}{J.
2093: Inst. Math. Appl.} \textbf{\bibinfo{volume}{6}}, \bibinfo{pages}{341}
2094: (\bibinfo{year}{1970});
2095: %
2096: %\bibitem[{\citenamefont{Ciucu}(1998)}]{Ciucu1998}
2097: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Ciucu}}, \bibinfo{journal}{Duke
2098: Math. J} \textbf{\bibinfo{volume}{94}}, \bibinfo{pages}{1}
2099: (\bibinfo{year}{1998});
2100: %
2101: %\bibitem[{\citenamefont{Lundow}(2001)}]{Lundow2001}
2102: \bibinfo{author}{\bibfnamefont{P.~H.} \bibnamefont{Lundow}},
2103: \bibinfo{journal}{Disc. Math.} \textbf{\bibinfo{volume}{231}},
2104: \bibinfo{pages}{321} (\bibinfo{year}{2001}).
2105:
2106: \bibitem[{\citenamefont{Friedland and Peled}(2005)}]{FriedlandP05}
2107: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Friedland}} \bibnamefont{and}
2108: \bibinfo{author}{\bibfnamefont{U.~N.} \bibnamefont{Peled}},
2109: \bibinfo{journal}{Adv. Appl. Math.} \textbf{\bibinfo{volume}{34}},
2110: \bibinfo{pages}{486} (\bibinfo{year}{2005}).
2111:
2112: \bibitem[{\citenamefont{Fisher and Stephenson}(1963)}]{Fisher1963}
2113: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Fisher}} \bibnamefont{and}
2114: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Stephenson}},
2115: \bibinfo{journal}{Phys. Rev.} \textbf{\bibinfo{volume}{132}},
2116: \bibinfo{pages}{1411} (\bibinfo{year}{1963});
2117: %
2118: %\bibitem[{\citenamefont{Hartwig}(1966)}]{Hartwig1966}
2119: \bibinfo{author}{\bibfnamefont{R.~E.} \bibnamefont{Hartwig}},
2120: \bibinfo{journal}{J. Math. Phys.} \textbf{\bibinfo{volume}{7}},
2121: \bibinfo{pages}{286} (\bibinfo{year}{1966}).
2122:
2123: \bibitem[{\citenamefont{Heilmann and Lieb}(1972)}]{Heilmann1972}
2124: \bibinfo{author}{\bibfnamefont{O.~J.} \bibnamefont{Heilmann}} \bibnamefont{and}
2125: \bibinfo{author}{\bibfnamefont{E.~H.} \bibnamefont{Lieb}},
2126: \bibinfo{journal}{Commun. Math. Phys.} \textbf{\bibinfo{volume}{25}},
2127: \bibinfo{pages}{190} (\bibinfo{year}{1972}).
2128:
2129: \bibitem[{\citenamefont{Gruber and Kunz}(1971)}]{Gruber1971}
2130: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Gruber}} \bibnamefont{and}
2131: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Kunz}},
2132: \bibinfo{journal}{Commun. Math. Phys.} \textbf{\bibinfo{volume}{22}},
2133: \bibinfo{pages}{133} (\bibinfo{year}{1971}).
2134:
2135: \bibitem[{\citenamefont{Ferdinand}(1967)}]{Ferdinand1967}
2136: \bibinfo{author}{\bibfnamefont{A.~E.} \bibnamefont{Ferdinand}},
2137: \bibinfo{journal}{J. Math. Phys.} \textbf{\bibinfo{volume}{8}},
2138: \bibinfo{pages}{2332} (\bibinfo{year}{1967});
2139: %
2140: %\bibitem[{\citenamefont{Izmailian et~al.}(2003)\citenamefont{Izmailian,
2141: % Oganesyan, and Hu}}]{IzmailianOH03}
2142: \bibinfo{author}{\bibfnamefont{N.~S.} \bibnamefont{Izmailian}},
2143: \bibinfo{author}{\bibfnamefont{K.~B.} \bibnamefont{Oganesyan}},
2144: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{C.~K.} \bibnamefont{Hu}},
2145: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{67}},
2146: \bibinfo{pages}{066114} (\bibinfo{year}{2003});
2147: %
2148: %\bibitem[{\citenamefont{Izmailian et~al.}(2005)\citenamefont{Izmailian,
2149: % Priezzhev, Ruelle, and Hu}}]{Izmailian2005}
2150: \bibinfo{author}{\bibfnamefont{N.~S.} \bibnamefont{Izmailian}},
2151: \bibinfo{author}{\bibfnamefont{V.~B.} \bibnamefont{Priezzhev}},
2152: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Ruelle}}, \bibnamefont{and}
2153: \bibinfo{author}{\bibfnamefont{C.-K.} \bibnamefont{Hu}},
2154: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{95}},
2155: \bibinfo{pages}{260602} (\bibinfo{year}{2005}).
2156:
2157: \bibitem[{\citenamefont{Chang}(1939)}]{Chang1939}
2158: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Chang}},
2159: \bibinfo{journal}{Proc. R. Soc. Series A} \textbf{\bibinfo{volume}{169}},
2160: \bibinfo{pages}{512} (\bibinfo{year}{1939}).
2161:
2162: \bibitem[{\citenamefont{Lin and Lai}(1994)}]{LinL94}
2163: \bibinfo{author}{\bibfnamefont{G.~J.} \bibnamefont{Lin}} \bibnamefont{and}
2164: \bibinfo{author}{\bibfnamefont{P.~Y.} \bibnamefont{Lai}},
2165: \bibinfo{journal}{Physica A} \textbf{\bibinfo{volume}{211}},
2166: \bibinfo{pages}{465} (\bibinfo{year}{1994}).
2167:
2168: \bibitem[{\citenamefont{Baxter}(1968)}]{Baxter1968}
2169: \bibinfo{author}{\bibfnamefont{R.~J.} \bibnamefont{Baxter}},
2170: \bibinfo{journal}{J. Math. Phys.} \textbf{\bibinfo{volume}{9}},
2171: \bibinfo{pages}{650} (\bibinfo{year}{1968}).
2172:
2173: \bibitem[{\citenamefont{Nemirovsky and Coutinho-Filho}(1989)}]{Nemirovsky1989}
2174: \bibinfo{author}{\bibfnamefont{A.~M.} \bibnamefont{Nemirovsky}}
2175: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~D.}
2176: \bibnamefont{Coutinho-Filho}}, \bibinfo{journal}{Phys. Rev. A}
2177: \textbf{\bibinfo{volume}{39}}, \bibinfo{pages}{3120} (\bibinfo{year}{1989});
2178: %
2179: %\bibitem[{\citenamefont{Kenyon et~al.}(1996)\citenamefont{Kenyon, Randall, and
2180: % Sinclair}}]{KenyonRS96}
2181: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Kenyon}},
2182: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Randall}}, \bibnamefont{and}
2183: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Sinclair}},
2184: \bibinfo{journal}{J. Stat. Phys.} \textbf{\bibinfo{volume}{83}},
2185: \bibinfo{pages}{637} (\bibinfo{year}{1996});
2186: %
2187: %\bibitem[{\citenamefont{Beichl et~al.}(2001)\citenamefont{Beichl, O'Leary, and
2188: % Sullivan}}]{BeichlOS01}
2189: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Beichl}},
2190: \bibinfo{author}{\bibfnamefont{D.~P.} \bibnamefont{O'Leary}},
2191: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sullivan}},
2192: \bibinfo{journal}{Phys. Rev. E} \textbf{\bibinfo{volume}{64}},
2193: \bibinfo{pages}{016701} (\bibinfo{year}{2001}).
2194:
2195: \bibitem[{\citenamefont{Beichl and Sullivan}(1999)}]{BeichlS99}
2196: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Beichl}} \bibnamefont{and}
2197: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Sullivan}},
2198: \bibinfo{journal}{J. Comput. Phys.} \textbf{\bibinfo{volume}{149}},
2199: \bibinfo{pages}{128} (\bibinfo{year}{1999}).
2200:
2201: \bibitem[{\citenamefont{Kasteleyn}(1963)}]{Kasteleyn1963}
2202: \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Kasteleyn}},
2203: \bibinfo{journal}{J. Math. Phys.} \textbf{\bibinfo{volume}{4}},
2204: \bibinfo{pages}{287} (\bibinfo{year}{1963}).
2205:
2206: \bibitem[{\citenamefont{Fisher}(1966)}]{Fisher1966}
2207: \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Fisher}},
2208: \bibinfo{journal}{J. Math. Phys.} \textbf{\bibinfo{volume}{7}},
2209: \bibinfo{pages}{1776} (\bibinfo{year}{1966}).
2210:
2211: \bibitem[{\citenamefont{Fan and Wu}(1970)}]{Fan1970}
2212: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Fan}} \bibnamefont{and}
2213: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Wu}}, \bibinfo{journal}{Phys.
2214: Rev. B} \textbf{\bibinfo{volume}{2}}, \bibinfo{pages}{723}
2215: (\bibinfo{year}{1970}).
2216:
2217: \bibitem[{\citenamefont{McCoy and Wu}(1973)}]{McCoy1973}
2218: \bibinfo{author}{\bibfnamefont{B.~M.} \bibnamefont{McCoy}} \bibnamefont{and}
2219: \bibinfo{author}{\bibfnamefont{T.~T.} \bibnamefont{Wu}},
2220: \emph{\bibinfo{title}{The two-dimensional Ising Model}}
2221: (\bibinfo{publisher}{Harvard University Press}, \bibinfo{address}{Cambridge,
2222: Massachusetts}, \bibinfo{year}{1973}).
2223:
2224: \bibitem[{\citenamefont{Pemantle and Wilson}(2002)}]{Pemantle2002}
2225: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Pemantle}} \bibnamefont{and}
2226: \bibinfo{author}{\bibfnamefont{M.~C.} \bibnamefont{Wilson}},
2227: \bibinfo{journal}{J. Comb. Theory Ser. A} \textbf{\bibinfo{volume}{97}},
2228: \bibinfo{pages}{129} (\bibinfo{year}{2002}).
2229:
2230: \bibitem[{\citenamefont{Kong}(1999)}]{Kong1999}
2231: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kong}}, \bibinfo{journal}{J.
2232: Chem. Phys.} \textbf{\bibinfo{volume}{111}}, \bibinfo{pages}{4790}
2233: (\bibinfo{year}{1999}).
2234:
2235: \bibitem[{\citenamefont{Kong}(2006{\natexlab{a}})}]{Kong2006}
2236: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kong}}, \bibinfo{journal}{Phys.
2237: Rev. E} \textbf{\bibinfo{volume}{73}}, \bibinfo{pages}{016106}
2238: (\bibinfo{year}{2006}{\natexlab{a}}).
2239:
2240: \bibitem[{\citenamefont{Kong}(2006{\natexlab{b}})}]{Kong2006b}
2241: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kong}}, \bibinfo{journal}{Phys.
2242: Rev. E} \textbf{\bibinfo{volume}{74}}, \bibinfo{pages}{011102}
2243: (\bibinfo{year}{2006}{\natexlab{b}}).
2244:
2245: \bibitem[{\citenamefont{Granlund}(2006)}]{gmp}
2246: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Granlund}},
2247: \emph{\bibinfo{title}{GNU MP: The GNU Multiple Precision Arithmetic Library}}
2248: (\bibinfo{year}{2006}), \bibinfo{note}{version 4.2},
2249: \urlprefix\url{http://www.swox.com/gmp/}.
2250:
2251: \bibitem[{\citenamefont{Calkin and Wilf}(1998)}]{Calkin1998}
2252: \bibinfo{author}{\bibfnamefont{N.~J.} \bibnamefont{Calkin}} \bibnamefont{and}
2253: \bibinfo{author}{\bibfnamefont{H.~S.} \bibnamefont{Wilf}},
2254: \bibinfo{journal}{SIAM J. Discrete Math.} \textbf{\bibinfo{volume}{11}},
2255: \bibinfo{pages}{54} (\bibinfo{year}{1998}).
2256:
2257: \bibitem[{\citenamefont{Runnels}(1970)}]{Runnels1970}
2258: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Runnels}}, \bibinfo{journal}{J.
2259: Math. Phys.} \textbf{\bibinfo{volume}{11}}, \bibinfo{pages}{842}
2260: (\bibinfo{year}{1970}).
2261:
2262: \bibitem[{\citenamefont{Stoer and Bulirsch}(1992)}]{Stoer1980}
2263: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Stoer}} \bibnamefont{and}
2264: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bulirsch}},
2265: \emph{\bibinfo{title}{Introduction to Numerical Analysis}}
2266: (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{address}{New York, USA},
2267: \bibinfo{year}{1992}), \bibinfo{edition}{2nd} ed.
2268:
2269: \bibitem[{\citenamefont{Press et~al.}(1992)\citenamefont{Press, Teukolsky,
2270: Vetterling, and Flannery}}]{Press1992}
2271: \bibinfo{author}{\bibfnamefont{W.~H.} \bibnamefont{Press}},
2272: \bibinfo{author}{\bibfnamefont{S.~A.} \bibnamefont{Teukolsky}},
2273: \bibinfo{author}{\bibfnamefont{W.~T.} \bibnamefont{Vetterling}},
2274: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~P.}
2275: \bibnamefont{Flannery}}, \emph{\bibinfo{title}{Numerical Recipes in C: The
2276: Art of Scientific Computing}} (\bibinfo{publisher}{Cambridge University
2277: Press}, \bibinfo{address}{New York, NY, USA}, \bibinfo{year}{1992}).
2278:
2279: \bibitem[{\citenamefont{Williams and Kelley}(2004)}]{gnuplot40}
2280: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Williams}} \bibnamefont{and}
2281: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Kelley}},
2282: \emph{\bibinfo{title}{Gnuplot: An Interactive Plotting Program}}
2283: (\bibinfo{year}{2004}), \bibinfo{note}{version 4.0},
2284: \urlprefix\url{http://www.gnuplot.info}.
2285:
2286: \bibitem[{\citenamefont{Marquart}(1963)}]{Marquart1963}
2287: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Marquart}},
2288: \bibinfo{journal}{J. Soc. Indust. Appl. Math.} \textbf{\bibinfo{volume}{11}},
2289: \bibinfo{pages}{431} (\bibinfo{year}{1963}).
2290:
2291: \bibitem[{\citenamefont{Hammersley}(1966)}]{Hammersley1966}
2292: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hammersley}}, in
2293: \emph{\bibinfo{booktitle}{Research Papers in Statistics: Festschrift for J.
2294: Neyman}}, edited by \bibinfo{editor}{\bibfnamefont{F.}~\bibnamefont{David}}
2295: (\bibinfo{publisher}{Wiley}, \bibinfo{address}{London},
2296: \bibinfo{year}{1966}), p. \bibinfo{pages}{125}.
2297:
2298: \bibitem[{\citenamefont{Hill}(1987)}]{Hill1987}
2299: \bibinfo{author}{\bibfnamefont{T.~L.} \bibnamefont{Hill}},
2300: \emph{\bibinfo{title}{Statistical Mechanics : Principles and Applications}}
2301: (\bibinfo{publisher}{Dover Publications}, \bibinfo{address}{New York, USA},
2302: \bibinfo{year}{1987}).
2303:
2304: \bibitem[{\citenamefont{Zasedatelev et~al.}(1971)\citenamefont{Zasedatelev,
2305: Gurskii, and Volkenshtein}}]{Zasedatelev1971}
2306: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Zasedatelev}},
2307: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gurskii}}, \bibnamefont{and}
2308: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Volkenshtein}},
2309: \bibinfo{journal}{Mol. Biol.} \textbf{\bibinfo{volume}{5}},
2310: \bibinfo{pages}{245} (\bibinfo{year}{1971});
2311: %
2312: %\bibitem[{\citenamefont{Kong}(2001)}]{Kong2001}
2313: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kong}}, \bibinfo{journal}{J.
2314: Phys. Chem. B} \textbf{\bibinfo{volume}{105}}, \bibinfo{pages}{10111}
2315: (\bibinfo{year}{2001}).
2316:
2317: \bibitem[{\citenamefont{Cera and Kong}(1996)}]{DiCera1996}
2318: \bibinfo{author}{\bibfnamefont{E.~D.} \bibnamefont{Cera}} \bibnamefont{and}
2319: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Kong}},
2320: \bibinfo{journal}{Biophys. Chem.} \textbf{\bibinfo{volume}{61}},
2321: \bibinfo{pages}{107} (\bibinfo{year}{1996}).
2322:
2323:
2324: \bibitem[{\citenamefont{Petkov\v{s}ek et~al.}(1996)\citenamefont{Petkov\v{s}ek,
2325: Wilf, and Zeilberger}}]{Petkovsek1996}
2326: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Petkov\v{s}ek}},
2327: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Wilf}}, \bibnamefont{and}
2328: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Zeilberger}},
2329: \emph{\bibinfo{title}{A=B}} (\bibinfo{publisher}{AK Peters, Ltd.},
2330: \bibinfo{address}{Wellesley, MA, USA}, \bibinfo{year}{1996}).
2331:
2332: \bibitem[{\citenamefont{Wu}()}]{Wu2006c}
2333: \bibinfo{author}{\bibfnamefont{F.~Y.} \bibnamefont{Wu}}, \bibinfo{note}{private
2334: communication}.
2335:
2336: \end{thebibliography}
2337: }
2338: {
2339: \bibliography{md,md_1}
2340: }
2341:
2342:
2343:
2344: \end{document}
2345: