1: % PORUKA: privremeno je format onaj koji trosi manje papira
2: %
3: % ispell v12.tex <- gruba provjera engl.
4: % dvips -o v12.ps v12.dvi & <- napravim *.ps fajl
5: % dvips -p 9 -l 11 -o v12.ps v12.dvi & <- napravim *.ps fajl
6: % ghostview v12.ps & <- pogledam ga
7: %
8: % zaglavlje uzeto iz clanka
9: %---------------------------
10: % \documentstyle{elsart}
11: \documentclass{elsart} % za latex2e: \cal -> \mathcal
12: \usepackage{graphicx}
13: % \input /home/glumac/latex/croat.tex
14:
15: \begin{document}
16:
17: \newcommand{\JPA}{J. Phys. A}
18: \newcommand{\JPC}{J. Phys. C}
19: \newcommand{\JSP}{J. Stat. Phys.}
20: \newcommand{\PA}{Physica A}
21: \newcommand{\PRB}{Phys. Rev. B}
22: \newcommand{\PRE}{Phys. Rev. E}
23: \newcommand{\PRL}{Phys. Rev. Lett.}
24:
25: \newcommand{\be}{\begin{equation}}
26: \newcommand{\bea}{\begin{eqnarray}}
27: \newcommand{\cF}{{\mathcal F}}
28: \newcommand{\cG}{{\mathcal G}}
29: \newcommand{\cH}{{\mathcal H}}
30: \newcommand{\cO}{{\mathcal O}}
31: \newcommand{\cq}{complex-$q$ }
32: \newcommand{\cS}{complex-$S$ }
33: \newcommand{\ee}{\end{equation}}
34: \newcommand{\eea}{\end{eqnarray}}
35: \newcommand{\etl}{{\it et al.}\,\,} % ? etal vec postoji u latex2e ??
36: \newcommand{\req}{real-$q$ }
37: \newcommand{\vfi}{\varphi}
38:
39:
40: % ========== PROMIJENITI KASNIJE: ======================
41: % za pravi format: razkomentirati 3 linije i izbaciti 4 sljedece ( do =======)
42: \renewcommand
43: %% \baselinestretch 2 % s proredom
44: %% \baselineskip 24pt %
45: \baselinestretch 1 % bez proreda
46: \baselineskip 15pt %
47: \setlength\textwidth{19cm} % ve"ta "sirina teksta
48: \setlength\textheight{23cm} % ve"ta visina teksta
49: % \vsize=9.0truein
50: % \hsize=6.3truein
51: % =======================================================
52:
53: \begin{frontmatter}
54:
55: \title{Complex-$q$ zeros of the partition function of the
56: Potts model with long-range interactions}
57:
58: \author{Zvonko Glumac\thanksref{corresp}},
59: \thanks[corresp]{Corresponding author. tel.: (385)-031-208-900;
60: fax: (385)-031-208-905;
61: e-mail: zvonko@vrabac.ifs.hr.}
62:
63: \address{Faculty of Electrical Engineering,
64: Kneza Trpimira 2B, 31 000 Osijek, Croatia}
65: \author{Katarina Uzelac}
66:
67: \address{ Institute of Physics, P.O.B. 304,
68: Bijeni\v{c}ka 46, HR-10000 Zagreb, Croatia}
69:
70:
71: \begin{abstract}
72: The zeros of the partition function of the ferromagnetic
73: $q$-state Potts model with long-range interactions in the \cq plane
74: are studied in the mean-field case, while preliminary numerical results are
75: reported for the finite $1d$ chains with power-law decaying interactions.
76: In both cases, at any fixed temperature, the zeros lie on the
77: arc-shaped contours, which cross the positive real axis at the
78: value for which the given temperature is transition temperature.
79: For finite number of spins the positive real axis is free of zeros, which
80: approach to it in the thermodynamic limit.
81: The convergence exponent of the zero closest to the positive \req axis is
82: found to have the same value as the temperature critical exponent $1/\nu$.
83:
84:
85: \leftline{PACS: 05.50.+q, 64.60.Cn}
86: \end{abstract}
87:
88: \begin{keyword}
89: Phase transitions, $q$-state Potts model, Complex-$q$ zeros,
90: Long-range interactions
91: \end{keyword}
92:
93: \end{frontmatter}
94:
95:
96: \section{Introduction} \label{sec-int}
97:
98: Several decades ago, it was shown that the study of distribution
99: of the partition-function zeros in the complex plane of an appropriate
100: variable, can provide relevant informations on phase transitions.
101: In their pioneering work, Lee and Yang \cite{YL52,LY52} have shown
102: the connection between the existence of phase transition and the
103: distribution of partition function zeros of the ferromagnetic Ising model
104: in the complex plane of a symmetry breaking field.
105: Later on, \cite{F65,OKSK68}, similar connection was established for
106: the partition function zeros in the complex-temperature plane.
107: In both cases, the singularities at the transition point
108: are closely related to the way that the zeros approach (either with
109: temperature or with the size of the system) to the positive real axis.
110:
111: The partition function zeros were investigated for number of models,
112: including the Potts model both in complex-field \cite{KC98,KC99} and
113: in complex-temperature \cite{B90,M91,CHW96,KCCH99} plane.
114: In the Potts model a relevant quantity is also the number of states
115: $q$, often considered as a continuous parameter, which is formally
116: possible on the basis of its graphical representation \cite{FK69} as it
117: will be discussed later in text.
118: This gives the motivation to investigate the zeros for this model also
119: in the plane of complex $q$.
120: Such an idea was already formulated \cite{Bx86,Bx87} but mostly in
121: context of chromatic polynomials in connection with studies of the
122: ground state of the anti-ferromagnetic Potts model
123: \cite{ST97a,ST97b,RST98,ST99a,ST99b,S99,SS00}.
124: Only a few works on complex-$q$ plane address the ferromagnetic
125: Potts model at finite temperatures.
126: The papers by Chang and Shrock, \cite{CS00}, study the loci of
127: partition function zeros in $2d$ Potts model with different boundary
128: conditions.
129: In their very recent paper Kim and Creswick \cite{KC01} give a systematic
130: study of both Fisher and complex-$q$ zeros in the $2d$ Potts model with
131: short-range interactions and find the similarity between the scaling property
132: of complex-$q$ zeros and the den Nijs expression for the thermal critical
133: exponent.
134:
135: In the present paper we investigate the zeros in the complex-$q$ plane
136: in connection with the first- and second-order phase
137: transitions in the ferromagnetic Potts model.
138: To this purpose we consider here the Potts model with long-range
139: power-law decaying interactions, which
140: already in one dimension displays diversity of critical regimes
141: and embraces most of the nontrivial aspects of phase
142: transitions, characteristic of the Potts model in general.
143: The limiting case of this model, where all the interactions are equal
144: (and which represents the mean-field case of the model) has the
145: advantage to be exactly solvable for arbitrary $q$.
146:
147: In the next section we describe briefly the starting
148: Hamiltonian and its graphical representation. In subsequent sections
149: we present the results for the mean-field (Section 3) and the
150: power-law decaying
151: interactions case (Section 4).
152:
153:
154:
155: \section{Hamiltonian}
156:
157: The ferromagnetic Potts model with interactions of arbitrary range is
158: described by the Hamiltonian
159: \be \label{eq:lrham}
160: H = - \sum_{i,j}\; J_{i,j}\cdot
161: \delta\,(s_i, s_{j}),
162: \ee
163: where $s_i$ denotes the Potts spin at site $i$ of a
164: $d$-dimensional lattice,
165: which can take $q$ values, and $J_{i,j}\ge 0$ denotes the ferromagnetic
166: interaction between spins at sites $i$ and $j$.
167:
168:
169: An impressive amount of studies of this model in the case of
170: nearest neighbour interactions (for a review see Wu \cite{W82})
171: reveal its very complex critical behaviour which varies with $q$ and
172: comprises transitions of different orders and belonging to various
173: universality classes.
174:
175: The model with long-range interactions was studied less,
176: and mostly in two special cases.
177: The first is the case when all the interactions are equal and
178: it represents the mean-field limit of the model.
179: It can be approached by the saddle point approximation
180: \cite{KMS54,SF73,MS74,PL76} and in the thermodynamic limit
181: the exact expressions, in particular those for critical
182: temperature and exponents may be derived for all values of $q$.
183: Second is the case with power-law interactions
184: decaying with distance as $1/r^{d+\sigma}$.
185: The type of phase transition there depends in addition
186: on the parameter of range $\sigma$, which in certain aspects has
187: similar effects as the change of dimensionality \cite{LB96}.
188: The phase transition is there present in one dimension
189: \cite{D69,ACCN88,GU93} and by
190: varying $\sigma$, series of different regimes are met, which
191: makes this $1d$ model a useful paradigm for most of important
192: features of the model in general (including also the first-order
193: phase transition).
194: It was a subject of number of recent studies
195: \cite{GU93,LB95,C95,UG97,GU98,M99,BDD99,KL00,LM01}.
196:
197: The graphical representation of the Potts model \cite{FK69}
198: holds for arbitrary range of interactions.
199: The partition function for the Hamiltonian (\ref{eq:lrham})
200: may be expressed as
201: \be \label{eq:zlropci}
202: Z_N = \sum_{{\rm all}\;\cG} \,
203: \prod_{\rm {active\atop links}} v_{i,j} \;\,q^{n(\cG)},
204: \ee
205: where the summation is taken over all possible graphs $\cG$.
206: Each graph represents one possible configuration of
207: connections between the $N$ spins, called active links,
208: each of which contributes with a factor $v_{i,j} = \exp(J_{i,j}/{k_B\,T}) - 1$.
209: $n(\cG)$ denotes the number of
210: of disconnected parts of the graph called clusters.
211: The product is taken over all the active links of a considered graph.
212:
213: Whatever the interactions are, the number of clusters that a single graph
214: contains ranges from one cluster (low temperature limit, where
215: all spins are interconnected), up to at most $N$ clusters
216: (high temperature limit where no active links are present and each
217: spin is a cluster of size one).
218: Thus, by collecting all the graphs with the same number of clusters,
219: the partition function becomes simply an $N$-th order polynomial in $q$
220: \be \label{eq:Npol}
221: Z_N = \sum_{n = 1}^N\,a_n(T,J_{i,j},N)\;q^n.
222: \ee
223: This made possible to treat the number of states $q$ as a continuous parameter.
224: It also gives the possibility to examine the zeros of the partition function
225: as the zeros of the $N$-th order polynomial in $q$.
226:
227: The coefficients $a_{n}$ are functions of temperature, the finite size
228: and interactions, and are quite complicated in the general case
229: (\ref{eq:lrham}), but can be written in more compact form for
230: the special cases considered here.
231:
232:
233: \section {The mean-field case}
234:
235:
236: The simplest case of long-range interactions is the one where all the
237: interactions are set to be equal, $J_{i,j}=J$, which reduces the
238: Hamiltonian (\ref{eq:lrham}) to the form
239: \be \label{eq:mfham}
240: \cH_{MF} = H_{MF}/{k_B\,T} = -
241: \frac{K}{N} \sum_{i=1}^{N-1}\;\sum_{j=1}^{N-i}\;
242: \delta\,(s_i, s_{i+j}),
243: \ee
244: where $K=J/{k_B\,T}$, and division by $N$ is necessary to keep
245: the summation finite.
246: This model corresponds to the mean-field limit. It was solved by the
247: saddle point approximation \cite{KMS54}.
248: The phase transition is there of the first order for $q > 2$
249: and of the second order for $q \le 2$.
250: The exact expressions for the inverse of the transition temperature
251: $K_t$ is given by
252: \be \label{eq:Kexact1}
253: K_t = 2\,\frac{q-1}{q-2}\,\ln (q-1),
254: \hspace{1cm}
255: \ee
256: for $q > 2$,
257: and by $K_c = q$ for $q \le 2$.
258: Exact analytical expressions may be derived also for the
259: critical exponents matching the values obtained in the
260: mean-field regime of the short-range interaction model \cite{PL76}
261: (except those related to the correlation function, as it
262: will be discussed later in text).
263:
264: The graph representation (\ref{eq:zlropci}) reduces in the mean-field limit to
265: \be \label{eq:zmf}
266: Z_N = \sum_{{\rm all}\;\cG} \, v^{b(\cG)}\,q^{n(\cG)},
267: \ee
268: where the active links have all the same strength $v = \exp(K/N) - 1$,
269: independently of distance between the spins they connect and
270: $b(\cG)$ denotes the number of active links in a given graph $\cG$.
271: Written as a polynomial in $q$ the above equation is equal to
272: \be \label{eq:Npolmf}
273: Z_N = \sum_{n = 1}^N\,a_n(v, N)\;q^n,
274: \ee
275: where the coefficients $a_n$ are real and positive and depend on
276: temperature and size $N$ only.
277: (Notice, that, since the number of active bonds may reach up to $N\,(N-1)/2$,
278: the above partition function could equally be written as a polynomial
279: in temperature variable $v$, which would be of order $N\,(N-1)/2$.)
280:
281:
282: \subsection{Numerical results}
283:
284: For relatively small sizes the evaluation of coefficients and
285: finding the zeros of the polynomial (\ref{eq:Npolmf}) in the
286: complex plane may easily be performed numerically.
287: Numerical results for sizes up to $N = 35$ spins are presented in
288: Figs. 1 (a)-(c) for the three different temperatures,
289: $K = 0.5, 2, 4.540457$.
290: These temperatures are the transition temperatures of the mean-field model
291: with $q = 0.5, 2$, and $8$ respectively.
292: Each figure displays the data obtained for several sizes.
293:
294: One may observe that in all the three cases the zeros lie on contours
295: similar but different from circle
296: which cross the real axis near the value of $q$, for which the given
297: temperature is the transition temperature.
298: The zeros are absent from the positive real axis, but approach to
299: it with increasing $N$.
300:
301: We were thus interested to perform a more precise extrapolation of this
302: crossing point in the thermodynamic limit and studied the convergence
303: of the zero closest to the positive real axis
304: with $N$.
305: Further, we have examined the possible connection of the convergence
306: of the closest zero with $N$ to the scaling properties of the model
307: near criticality.
308:
309: To this purpose we have used Burlisch and Stoer (BST) extrapolation
310: procedure \cite{BS64,HS88}.
311: This algorithm was shown \cite{HS88} to converge rapidly, needs a modest
312: amount of data and, in advantage over the similar algorithm by van den
313: Broeck and Schwartz (VBS) \cite{VBS79}, it is less sensitive to the rounding
314: errors.
315:
316: In short, BST algorithm extrapolates a sequence of the form
317: \be \label{eq:seq}
318: A(N) = A_{\infty} + a_1\,\cdot\,N^{-\omega} +
319: a_2\,\cdot\,N^{-2\,\omega}
320: + \cdots
321: \ee
322: in the $N\,\to\,\infty$ limit by a set of transformations recursively
323: given by
324: \bea
325: A_m^{(N)} & = & A_{m - 1}^{(N + 1)} \\
326: & + & \left(A_{m - 1}^{(N + 1)} - A_{m - 1}^{(N)}\right) \, \cdot \,
327: \left[ \left(\frac{N + m}{N}\right)^{\omega}\,
328: \left( 1 -
329: \frac{A_{m - 1}^{(N + 1)} - A_{m - 1}^{(N)}}{A_{m - 1}^{(N + 1)}
330: - A_{m - 2}^{(N + 1)}}
331: \right) - 1 \right]^{-1}, \nonumber
332: \eea
333: where $A_{-1}^{(N)} = 0$ and $A_0^{(N)} = q_{1, 2}^{(0)}(N)$
334: are the input data and $\omega$ is a free parameter.
335:
336: If the original series is of the form:
337: \be \label{eq:seq12}
338: A(N) = A_{\infty} + c_1\,\cdot\,N^{-\omega_1} +
339: c_2\,\cdot\,N^{-\omega_2}
340: + \cdots
341: \ee
342: it incorporates the leading convergence exponent as $\omega$,
343: while the higher order approximants converge as $N^{-\omega^{'}}$,
344: where $\omega^{'}=min(2\omega_1,\omega_2)$.
345: Although it was shown, that the extrapolations are insensitive to the choice
346: of $\omega$ in a wide range of values, it influences its convergence
347: speed. Thus, by maximising the convergence of higher approximants one can in
348: the same time obtain both the extrapolated value and the leading convergence
349: exponent. Also, when the extrapolated value is known in advance,
350: $\omega$ yields the leading convergence exponent.
351:
352: We have examined the convergence of the real and imaginary parts
353: of the zero closest to the real axis, denoted as $q_{1, 2}^{(0)}(N)$
354: for several temperatures using the numerical results ranging from
355: $N = 21$ to $N = 35$.
356:
357: In the extrapolation of both real and imaginary part
358: $q_{1, 2}^{(0)}(N)$,
359: the same parameter $\omega$ was used and was fixed by the constraint
360: $q_2^{(0)}(N\,\to\,\infty)\,\to\,0$.
361:
362: The results of extrapolations are presented on Table 1.
363: By $Q_0$ is denoted the value of $q$ for which the considered temperature
364: is known to be the transition temperature.
365: All the data were obtained by requiring that condition $q_2^{(0)}(N)\,\to\,0$
366: should be fulfilled with precision of $10^{-6}$.
367:
368: \begin{table}[hbt] \label{tb:rez12}
369: \caption{ The BST extrapolation results of the positions of
370: closest zeros.}
371: \begin{center} \begin{tabular}{ccccccc}
372: & & & & & & \\
373: \hline
374: $K$ & 0.1 & 0.5 & 1.0 & 2.0 & 2.772589 & 4.540457 \\
375: $Q_0$ & 0.1 & 0.5 & 1.0 & 2.0 & 3.0 & 8.0 \\
376: $q_1$ & 0.09998 & 0.504 & 0.95 & 2.05 & 3.02 & 8.00009 \\
377: $\omega$ & 0.34 & 0.33 & 0.33 & 0.55 & 0.91 & 1.19 \\
378: \hline
379: & & & & & & \\
380: \end{tabular} \end{center} \end{table}
381:
382: The presented results indicate that: \
383:
384: (i) loci of complex-$q$ zeros intersect the positive \req axis at
385: value $q_1 = Q_0$. The values agree up to a few percents.
386:
387: (ii) Power-law convergence of the closest zero
388: is observed in all the cases considered.
389:
390: The convergence exponent $\omega$ represents the scaling with
391: size of the distance from the real axis to the closest zero.
392: Before proceeding with the finite-size scaling analysis it is important to
393: notice that here the scaling is not done with the linear size, but
394: rather with volume, more precisely, with number of spins,
395: since in the infinitely coordinated model the distance as well as the
396: dimensionality lose their meaning.
397: In this respect the considered mean-field limit differs from the standard
398: mean-field approach to the short-range interaction model.
399: As pointed out by Botet \etl \cite{BJP82,BJ83} the basic scaling
400: quantity in the infinitely coordinated system is not the correlation length,
401: but the "correlation number"
402: $N_c$, which scales with temperature with the critical exponent
403: $\nu^{*}=\nu_{MF} \cdot d_c$. The exponent $\nu_{MF}$ and $d_c$
404: are, respectively, the mean-field value of the critical exponent $\nu$
405: and
406: the upper critical dimension (the dimension above which the mean-field regime
407: sets on) of the short-range model.
408:
409: Thus, the temperature critical exponent for the present model,
410: where only the scaling with the number of spins can be applied
411: will be $1/\nu^{*}$ and not $1/\nu_{MF}$.
412: In the Potts model $\nu_{MF}=1/2$ both for $q < 2$ and $q=2$, while
413: $d_c=4$ for $q=2$, but $d_c=6$ for $q<2$ for symmetry reasons.
414: This gives $\nu^{*}=2$ for $q=2$ and $\nu^{*}=3$ for $q<2$.
415:
416: The results presented in Table 1 clearly show that
417: in the regime of second-order phase transition ($q \leq 2$) the
418: convergence exponent coincides up to a few percents
419: with the temperature critical exponent $1/\nu^{*}$.
420: This result should be compared to the one in complex temperature plane,
421: where it was shown \cite{IPZ83} that the distance of the zero
422: closest to the real axis scales as the temperature critical exponent.
423: Similar behaviour in a complex-$q$ plane was recently observed in
424: the $2d$ ferromagnetic Potts model by Kim and Creswick \cite{KC01},
425: but the precision of their results did not permit them to be conclusive
426: whether the obtained convergence exponent was indeed equal to the
427: exponent $1/\nu$.
428:
429: In cases $q=3$ and $q=8$ we obtain approximately linear convergence,
430: proportional to the system size, which is the form of scaling that is
431: expected for the first-order transition.
432:
433:
434: \subsection{Large $N$ expansion}
435:
436: In the limit $N \gg 1$
437: some analytical results may easily be derived
438: by extending the mean-field solution by Kihara \etl
439: For the large number of spins $N$,
440: the free energy density, without the terms which are constant
441: in the order parameter $S$, is given by
442: \bea \label{eq:slen}
443: & \cF & (K, q, S) = \\ & - & \,S^2\,\frac{K}{2}\,\frac{q-1}{q}
444: +
445: \frac{1+(q-1)S}{q}\,\ln\,\left[1+(q-1)S\right] +
446: (q-1)\,\frac{1-S}{q}\,\ln(1-S). \nonumber
447: \eea
448: The partition function is given by
449: \be
450: Z_N (K, q)
451: \sim
452: % = \, \exp\left[ NK/2q + N \ln q\right]
453: \,\int_0^1 \exp\,(- N\,\cF)\,dS,
454: \ee
455: where the proportionality sign stands for all the omitted constant terms.
456:
457: We limit our considerations here to the first-order regime of
458: temperatures, where the free energy density at the transition
459: shows two equal minima, and the partition function integral can be
460: approximately solved by a saddle point method
461: \be
462: Z_N \sim \, 1 + \exp\,[-\,N \cF(S_{min})\,]. \label{eq:pf}
463: \ee
464: The first term on the right hand side denotes the minimum at $S=0$,
465: and the second term is the
466: contribution of the second minimum at $S=S_{min} > 0$,
467: which is obtained from
468: the equation for the extremum
469: \be
470: K S = \ln \frac{1+(q-1)S}{1-S}. \label{eq:ext}
471: \ee
472:
473: The straightforward way to find the complex-$q$ zeros would consist
474: of two steps:
475: first, by using (\ref{eq:ext}) one should represent $\cF$ as a function
476: of $K$ and $q$ only;
477: second, one should solve the equation $Z_N = 0$
478: by analytic continuation of $q$ to the complex plane.
479: Since $\cF = \cF_R + i \cF_I$ then becomes complex,
480: the vanishing of $Z_N$, as given by (\ref{eq:pf}), requires
481: \be \label{eq:uvjeti}
482: \cF_R = \, 0, \hspace{1cm} N \cF_I = \, \pm \, (2 m + 1) \pi,
483: \;\;\;\;m = 0, 1, \ldots
484: \ee
485: These two equations determine the position of complex conjugate
486: pairs of zeros $(q_1^{(m)}, \pm\,q_2^{(m)})$
487: in the complex-$q$ plane.
488:
489: The first of the above mentioned steps cannot be performed in the
490: straightforward way.
491: Instead, we will solve (\ref{eq:ext}) in variable $q$
492: \be
493: q = \frac{1-S}{S} [ \exp(K S) - 1 ], \label{eq:qS}
494: \ee
495: and represent $\cF$ as function of $S$ and $K$.
496: \be \label{eq:FS}
497: \cF (K, S) = \frac{K}{2}\frac{S^2}{1-S}
498: \frac{(1-S)\exp(K S) + 1}{\exp(K S) - 1} + \ln (1-S).
499: \ee
500: The equations (\ref{eq:uvjeti})
501: are then solved
502: in the \cS plane and, by equation (\ref{eq:qS}),
503: mapped back into the complex-$q$ plane.
504:
505: The solutions $q^{(m)}_{1,2}$ should comply with both equations
506: (\ref{eq:uvjeti}).
507: Since the first of equations (\ref{eq:uvjeti}) is $N$-independent,
508: it contains all the possible positions of zeros for arbitrary $N$
509: including the limit $N\to\infty$. The zeros of the system of
510: particular size $N$ are then determined as intersections of
511: this curve with solutions of the second equation in (\ref{eq:uvjeti}).
512: For large $N$ these intersections appear to be equally distributed along the
513: the curve $\cF_R = 0$, so that it can be drawn as the contour
514: containing zeros. (see Fig. 2 as an illustration).
515: The curve crosses the positive \req axis
516: at a point which is transition point for a given temperature.
517:
518: Since the locus of zeros has a shape similar to a circle,
519: we draw the circular contour (long dashed line) in order to stress
520: the difference.
521: The parameters of this circle are determined by the facts, that its
522: center lies on the \req axis (since zeros appear in complex conjugate
523: pairs only) and that it intersects the positive \req axis at $q$ solution
524: of (\ref{eq:Kexact1}) and the negative \req axis at $q = - |q|$,
525: solution of
526: \be
527: K = 2\,\frac{|q|+1}{|q|+2}\,\ln (|q|+1). \label{eq:Kexact3}
528: \ee
529:
530: We were not able to give analytical expressions for all the zeros
531: of the model, but the location of zeros closest to the positive real
532: $S$ or $q$ axis is a relatively easy task.
533:
534: Since we limit here to the first-order transitions,
535: the temperature will be fixed through the parameter $Q_0 > 2$
536: \be \label{eq:KQ}
537: K = 2\,\frac{Q_0-1}{Q_0-2}\,\ln (Q_0-1),
538: \ee
539: In the thermodynamic limit the second minimum of $\cF (S)$
540: appears at $S_0 = (Q_0-2)/(Q_0-1)$.
541: For large but finite $N$, we will expand free energy (\ref{eq:FS})
542: around the minimum, $S = S_0 + s$.
543: The small $s \ll S_0$ vanishes as $N\,\to\,\infty$, so we
544: expect to have, to leading order in $1/N$,
545: \be \label{eq:s1s2}
546: s = s_1 + i\,s_2, \hspace{1cm} s_1 = \frac{a}{N^{x_1}},
547: \hspace{0.5cm} s_2 = \frac{b}{N^{x_2}},
548: \ee
549: for real constants $a, b, x_1, x_2$.
550: The expansion of (\ref{eq:FS}) and the constraint (\ref{eq:uvjeti})
551: give the equation
552: \bea
553: \cF & = & (Q_0 - 1)\,\left[
554: 1 - \ln\,(Q_0 - 1)\,(A_1 - B_1)
555: \right] \, s + \cO(s^2)
556: = 0 \pm\,i\,\frac{2\,m + 1}{N}\,\pi, \nonumber \\
557: A_1 & = & \left(\frac{K - Q_0 + 1}{Q_0} +
558: \frac{2}{Q_0 - 2} \right), \;
559: B_1 = \frac{Q_0 - 1}{Q_0\,(Q_0 - 2)}\,K - 1.
560: \eea
561: The closest zero has the smallest imaginary
562: part, so its index is $m = 0$.
563: By inserting (\ref{eq:s1s2}) in the above equation,
564: one reads
565: \be \label{eq:konv1}
566: a = 0, \hspace{1cm}
567: b = \frac{\pi}{(Q_0 - 1) - \ln\,(Q_0 - 1)\,(A_1 - B_1)},
568: \hspace{1cm} x_2 = 1,
569: \ee
570: while $x_1$ is not defined.
571: Mapping back to the complex-$q$ plane by using the equation (\ref{eq:qS})
572: gives
573: \bea \label{eq:fp1qN}
574: q_1^{(0)}(N) & = & Q_0 + \frac{(Q_0 - 1)^2}{Q_0 -2}\,
575: (K - Q_0)\,\frac{a}{N^{x_1}} \nonumber \\
576: & + & \frac{(Q_0 - 1)^2}{Q_0 -2}\,\left\{
577: \frac{K}{2}\,[K - 2\,(Q_0 - 1)] + \frac{Q_0 - 1}{Q_0 - 2}\,(Q_0 - K)
578: \right\}\,\left(
579: \frac{a^2}{N^{2\,x_1}} - \frac{b^2}{N^{2\,x_2}}
580: \right) \nonumber \\
581: & + & \cO\left(N^{-3\,x}\right), \\
582: q_2^{(0)}(N) & = & \frac{(Q_0 - 1)^2}{Q_0 -2}\,
583: (K - Q_0)\,\frac{b}{N^{x_2}} \nonumber \\
584: & + & \frac{(Q_0 - 1)^2}{Q_0 -2}\,\left\{
585: \frac{K}{2}\,[K - 2\,(Q_0 - 1)] + \frac{Q_0 - 1}{Q_0 - 2}\,(Q_0 - K)
586: \right\}\,\frac{2\,a\,b}{N^{x_1 + x_2}} \nonumber \\
587: & + & \cO\left(N^{-3\,x}\right). \nonumber
588: \eea
589: Finally, by inserting values for $a, b, x_2$ from (\ref{eq:konv1}),
590: one obtains the large-$N$ dependence of the
591: distance of the closest zero from the real axis in the case of the
592: first-order transitions.
593: \bea \label{eq:q12N}
594: q_1^{(0)}(N) & = & Q_0 - \frac{(Q_0 - 1)^2}{Q_0 -2}\,\left\{
595: \frac{K}{2}\,[K - 2\,(Q_0 - 1)] + \frac{Q_0 - 1}{Q_0 - 2}\,(Q_0 - K)
596: \right\}\,\frac{b^2}{N^2} \nonumber \\
597: & + & \cO\left(N^{-3\,x}\right), \nonumber \\
598: q_2^{(0)}(N) & = & \frac{(Q_0 - 1)^2}{Q_0 -2}\,
599: (K - Q_0)\,\frac{b}{N} + \cO\left(N^{-3\,x}\right),
600: \eea
601: The above expressions can be checked numerically.
602: For illustration, the numerical data of the closest zeros,
603: for the temperature
604: critical for $q=8$
605: are presented in Table 2
606: for different sizes $N$.
607:
608: \begin{table}[htb] \label{tb:Nq8}
609: \caption{ The location of the closest zeros for the temperature
610: given by the relation (\ref{eq:KQ}) for $Q_0 = 8$. $N$ denotes the
611: number of the system considered.}
612: \begin{center} \begin{tabular}{ccc}
613: & & \\
614: \hline
615: $N$ & $q_1^{(m = 0)}(N)$ & $q_2^{(m = 0)}(N)$ \\
616: 5000&7.99999543669504&$ -8.679091351481774\,\cdot\,10^{-3}$ \\
617: 6000&7.99999683103614&$ -7.232577480179549\,\cdot\,10^{-3}$ \\
618: 7000&7.99999767178073&$ -6.199352825621011\,\cdot\,10^{-3}$ \\
619: 8000&7.99999821745666&$ -5.424434119811201\,\cdot\,10^{-3}$ \\
620: 9000&7.99999859157044&$ -4.821719459791061\,\cdot\,10^{-3}$ \\
621: 10000&7.99999885917194&$ -4.339547669719317\,\cdot\,10^{-3}$ \\
622: 11000&7.99999905716679&$ -3.945043440975050\,\cdot\,10^{-3}$ \\
623: 12000&7.99999920775815&$ -3.616289894012384\,\cdot\,10^{-3}$ \\
624: 13000&7.99999932495365&$ -3.338113800831173\,\cdot\,10^{-3}$ \\
625: 14000&7.99999941794470&$ -3.099677139486969\,\cdot\,10^{-3}$ \\
626: 15000&7.99999949296515&$ -2.893032025972416\,\cdot\,10^{-3}$ \\
627: \hline \end{tabular} \end{center}
628: \end{table}
629:
630: The data $q_{1, 2}^{(0)}(N)$ from Table 2 and analogous series of data for
631: different temperatures corresponding to $q=3$ and $q=2.1$ were extrapolated
632: to the thermodynamic limit by using the BST algorithm.
633:
634: The extrapolation parameter $\omega$ was fixed by the constraint
635: $q_2^{(0)}(N)\,\to\,0$.
636: After we have made the extrapolations of $q_{1, 2}^{(0)}(N)$
637: and found the values of $q_1(\infty)$ and $\omega$, we have performed
638: alternative calculation of convergence exponents of $q_{1, 2}^{(0)}(N)$.
639: By assuming the convergence of the $q_{1, 2}^{(0)}(N)$ to be of the form
640: \be \label{eq:q12alfa}
641: q_1^{(0)}(N) -
642: q_1(\infty) = C_1\,\cdot\,N^{-\alpha_1 } + \cdots,
643: \hspace{0.9cm}
644: q_2^{(0)}(N) = C_2\,\cdot\,N^{-\alpha_2 } + \cdots,
645: \ee
646: the convergence exponents $\alpha_{1, 2}(N)$ for real and imaginary part
647: of the closest zeros, can be expressed as
648: \bea \label{eq:alfa12}
649: \alpha_1(N) & = & - \frac{\ln\left\{\left[q_1^{(0)}(N) - q_1(\infty)\right]/
650: \left[q_1^{(0)}(N + 1) - q_1(\infty)\right]
651: \right\}}{\ln[N/(N + 1)]}, \nonumber \\
652: \alpha_2(N) & = & - \frac{\ln\left[
653: q_2^{(0)}(N) / q_2^{(0)}(N + 1) \right] }{\ln[N/(N + 1)]} .
654: \eea
655: The above data for $\alpha_{1, 2}(N)$ are extrapolated
656: again by the BST algorithm.
657: The complete set of extrapolated results is summarised
658: in the Table 3.
659:
660: \begin{table}[hbt] \label{tb:bst1ord}
661: \caption{ The results of the BST extrapolations performed on the
662: closest zeros, $q_{1, 2}(\infty)$, followed by parameter
663: $\omega$ and convergence exponents $\alpha_{1, 2}$,
664: for set of temperatures determined by $Q_0$ and (\ref{eq:KQ}).
665: The predicted convergence exponents $x_{1, 2}$ as given
666: by (\ref{eq:q12N}) are in the
667: last two columns. }
668: \begin{center}
669: \begin{tabular}{ cccccccc }
670: & & & & & & & \\
671: \hline
672: $Q_0$ & $q_1(\infty)$ & $q_2(\infty)$ & $\omega$ &
673: $\alpha_1$ & $\alpha_2$ & $x_1$ & $x_2$ \\
674: 8 & 8.0000000000001 & $3\,\cdot\,10^{-13}$ & 1.0000 &
675: 1.9999988 & 1.000000002 & 2 & 1 \\
676: 3 & 3.000000000005 & $4\,\cdot\,10^{-12}$ & 1.00000 &
677: 2.00003 & 0.99999997 & 2 & 1 \\
678: 2.1 & 2.1002 & $9\,\cdot\,10^{-6}$ & 0.98 & 1.85 & 0.95 & 2 & 1 \\
679: \hline
680: & & & & & & & \\
681: \end{tabular}
682: \end{center}
683: \end{table}
684:
685: We observe a full agreement between the analytically given values for
686: $q$ and convergence exponents on one side, and the numerically
687: calculated and extrapolated values for $q$ and exponents on the other side.
688: Although the scaling in the case of the first-order phase transition
689: is trivial and linearly proportional to system size, the results can be used
690: to check the convergence of numerical results.
691: For $q=3$ and $q=8$ the convergence exponent has improved compared to
692: the earlier numerical results given in Table 1, confirming that larger
693: error in this case comes from the small size of considered systems and
694: that scaling is indeed linear with size as expected for the first-order
695: phase transition.
696: The loss of precision in extrapolations persists only for the
697: case corresponding to $q=2.1$ and should be attributed to the
698: crossover effects due to vicinity of the border between the first-
699: and second-order transitions.
700:
701: \vspace{1cm}
702:
703: For the partition function zeros lying on the curve in the appropriate
704: complex plane, the normalised density of zeros $g$, is defined as
705: \cite{LY52}
706: \be
707: \label{eq:defg}
708: g = \frac{1}{N}\,\frac{d n}{d l}, \hspace{1cm} \int\,g\,dl = 1,
709: \ee
710: where $d n$ is the number of zeros inside the arc of the curve of
711: the length $d l$. In the case of the field- or temperature-driven
712: second-order transitions, the density of zeros vanishes
713: at the transition with an exponent connected to
714: critical exponents \cite{CK97}. On the contrary, at the first-order
715: transition point, the density of zeros remains constant.
716: To the end of this paragraph, we will examine the behaviour of
717: density of complex-$q$ zeros in the vicinity of first-order transition point.
718: To this purpose, we notice that the large-$N$ expansion
719: of $q_{1, 2}(N)$ for the closest zeros (those with $m =0$)
720: presented earlier in text,
721: may be directly generalised to other values of $m$
722: as long as the condition $(2 m + 1) \ll N$ is fulfilled.
723: This leads to
724: \bea
725: q_1^{(m)}(N) - q_1(\infty) & = &
726: C_1\,\cdot\,\left(
727: \frac{2 m + 1}{N}\right)^{\alpha_1} + \cdots,
728: \nonumber \\
729: q_2^{(m)}(N) & = & C_2\,\cdot\,\left(
730: \frac{2 m + 1}{N}\right)^{\alpha_2} + \cdots,
731: \eea
732: with the constants $q_1(\infty), C_1$ and $C_2$ given by
733: the relations (\ref{eq:q12N}), where $1/N$ is changed by
734: $(2 m + 1)/N$.
735:
736: To calculate numerically density of zeros (\ref{eq:defg}),
737: we choose two neighbouring zeros, with $d l$ denoting
738: geometrical distance between them, \\
739: $ d l = \left\{ \left[q_1^{(m + 1)} - q_1^{(m)}\right]^2 +
740: \left[q_2^{(m + 1)} - q_2^{(m)}\right]^2\right\}^{1/2}$.
741: In that case density of zeros near positive real-$q$ axis becomes
742: \bea
743: g(m) = & \frac{1}{N} & \,\left\{
744: C_1^2\,\left(\frac{2 m + 1}{N}\right)^{2\,\alpha_1}\,\left[\left(1 +
745: \frac{2}{2 m + 1}\right)^{\alpha_1} - 1\right]^2
746: \right. \nonumber \\
747: & + & \left.
748: C_2^2\,\left(\frac{2 m + 1}{N}\right)^{2\,\alpha_2}\,\left[\left(1 +
749: \frac{2}{2 m + 1}\right)^{\alpha_2} - 1\right]^2
750: \right\}^{-1/2}.
751: \eea
752: The edge $(m = 0)$ density of zeros $g_0$, tends to constant value
753: in the thermodynamic limit.
754: \be \label{eq:gN1}
755: g_0 = \frac{1}{2\,C_2}\,\left[
756: 1 - 8\,\frac{C_1^2}{C_2^2}\, \frac{1}{N^2} + \cdots
757: \right] \;\to\; \frac{1}{2\,C_2},
758: \ee
759:
760: For example, for $Q_0 = 8$ and $K(Q_0 = 8) = 4.540457014462$
761: this constant is
762: \be \label{eq:constg}
763: \frac{1}{2\,C_2} = \left[ 2\,\frac{(Q_0 - 1)^2}{Q_0 -2}\,
764: (K - Q_0)\,b \right]^{-1}
765: = 0.011521939 \cdots ,
766: \ee
767: For illustration, we check this result numerically by calculating
768: for the values of $Q_0$ and $K$
769: from above, and $N = 10\,000$, the
770: twenty zeros closest to the positive \req axis and corresponding densities.
771: They are presented in Table 4.
772:
773: \renewcommand
774: %% \baselinestretch 2 % s proredom
775: %% \baselineskip 24pt %
776: \baselinestretch 1 % bez proreda
777: \baselineskip 15pt %
778: \begin{table} \label{tb:gq8}
779: \caption{ The twenty zeros closest to the positive \req axis
780: and the corresponding densities are
781: shown at temperature determined from the relation
782: (\ref{eq:KQ}) and $Q_0 = 8$ and for the $ N = 10\,000$ spins.}
783: \begin{center} \begin{tabular}{cccc}
784: & & & \\
785: \hline
786: $m$ & $q_1^{(m)}$ & $q_2^{(m)}$ & $g(m)$ \\
787: 0&7.99999885917192&-0.004339547669719&0.0115219529910602 \\
788: 1&7.99998973259138&-0.013018627057442&0.0115219974088187 \\
789: 2&7.99997147966529&-0.021697658591626&0.0115220714335465 \\
790: 3&7.99994410086352&-0.030376610375500&0.0115221750581617 \\
791: 4&7.99990759689083&-0.039055450520507&0.0115223082726076 \\
792: 5&7.99986196868649&-0.047734147149695&0.0115224710640261 \\
793: 6&7.99980721742419&-0.056412668400979&0.0115226634165474 \\
794: 7&7.99974334451167&-0.065090982430549&0.0115228853117394 \\
795: 8&7.99967035159032&-0.073769057415945&0.0115231367280630 \\
796: 9&7.99958824053479&-0.082446861559540&0.0115234176412815 \\
797: 10&7.99949701345249&-0.091124363091702&0.0115237280243068 \\
798: 11&7.99939667268302&-0.099801530274076&0.0115240678473335 \\
799: 12&7.99928722079755&-0.108478331402754&0.0115244370775645 \\
800: 13&7.99916866059815&-0.117154734811648&0.0115248356796385 \\
801: 14&7.99904099511710&-0.125830708875537&0.0115252636152771 \\
802: 15&7.99890422761603&-0.134506222013369&0.0115257208435568 \\
803: 16&7.99875836158505&-0.143181242691353&0.0115262073207999 \\
804: 17&7.99860340074195&-0.151855739426127&0.0115267230005962 \\
805: 18&7.99843934903115&-0.160529680787899&0.0115272678338162 \\
806: 19&7.99826621062264&-0.169203035403569&0.0115278417687404 \\
807: \hline
808: & & & \\
809: \end{tabular} \end{center} \end{table}
810: \renewcommand
811: %% \baselinestretch 2 % s proredom
812: %% \baselineskip 24pt %
813: \baselinestretch 1 % bez proreda
814: \baselineskip 15pt %
815:
816: The limiting value of the density at the transition point, is extracted
817: by the BST extrapolation procedure (\ref{eq:seq}) simply
818: by changing variable $1/N$ into $(2\,m + 1)/N$.
819: \be \label{eq:g0}
820: g_m = g_0 +{\tilde g}\,\left(\frac{2\,m + 1}{N}\right)^{\omega} + \cdots.
821: \ee
822: The extrapolations give $g_0 = 1.152194\,\cdot\,10^{-2}$ for a
823: wide range of $0.3 < \omega < 3.0$, which reproduces the analytically
824: obtained value (\ref{eq:constg}) in six significant digits.
825:
826: We can conclude that the density of the partition function zeros in the
827: complex-$q$
828: plane behave in similar way as the density of zeros in the
829: complex-temperature \cite{F65} or complex-field \cite{CK97} planes:
830: in thermodynamic limit
831: it has constant value at the first-order transition point.
832:
833: \section{The power-law decaying interactions} \label{sec-lr}
834:
835: The second case that we have considered is the ferromagnetic Potts model
836: in one dimension with power-law decaying interactions. When taking the
837: periodic boundary conditions, it is described by the reduced Hamiltonian
838: \be
839: \cH_{PL} = - \frac{H_{PL}}{k_B\,T} =
840: K \sum_{i=1}^{N-1}\;\sum_{j=1}^{N-i}\; \delta\,(s_i, s_{i+j})
841: \left[\frac{1}{j^{1+\sigma}}+\frac{1}{(N-j)^{1+\sigma}}\right].
842: \ee
843: Although in one dimension, the model is nontrivial and has a phase
844: transition at nonzero temperature for all $q$ when
845: $0 < \sigma \leq 1$ \cite{D69,ACCN88,GU93}.
846: This transition is of the mean-field type for low enough values of
847: $\sigma$, $\sigma< \sigma_c(q)$, and is a
848: non-trivial second-order phase transition for $\sigma > \sigma_c(q)$.
849: In the classical regime $\sigma < \sigma_c(q)$, the transition is of the
850: first order when $q > 2$ \cite{UG97,GU98,BDD99}.
851: The exact position of the line separating the two regimes, $\sigma_c(q)$,
852: is a difficult and still open question \cite{GU98,BDD99,UG00}.
853:
854: In the graph representation, the difference from the mean-field case is that
855: the contribution of active links depends on distance between the spins
856: they connect.
857: The contribution of active bonds may then have $N/2$ (for $N$ even)
858: or $(N - 1)/2$ (for $N$ odd) distinct values
859: given by $v_j = \exp[K/j^{1+\sigma}+K/(N-j)^{1+\sigma}] - 1$.
860: If we denote by $b_j(\cG)$ the number of active links of the length
861: $j$ or $N - j$ in a graph $\cG$, the partition function becomes
862: \be \label{eq:zlr}
863: Z_N = \sum_{{\rm all}\;\cG} \, v_1^{b_1(\cG)}\,v_2^{b_2(\cG)}\,\cdots
864: \,q^{n(\cG)}
865: \ee
866: or, in form of polynomial in $q$,
867: \be \label{eq:Npollr}
868: Z_N = \sum_{n = 1}^N\,a_n(v_j, N)\;q^n.
869: \ee
870: The real and positive coefficients $a_n$ depend on temperature and $N$,
871: but also on the parameter of range $\sigma$.
872: This makes the calculations more demanding for the computer memory.
873:
874: We have performed exact numerical calculations on chains with
875: $N \leq 9$ spins for two sets of parameters,
876: ($\sigma = 0.3$, $K = 0.576$) and ($\sigma = 0.8$ $K = 0.8230$).
877: As in the previous case, the chosen temperatures are the critical
878: temperatures known for certain values of $q$. They correspond to
879: $q = 5$ and $q = 2$ respectively.
880: Remark, however, that in the present case the critical temperatures
881: are not known exactly. The values used here were obtained by the finite-range
882: scaling approach \cite{GU93}.
883: The first and second set correspond respectively to the to the first-
884: and second-order phase transition.
885:
886: The numerical results for the loci of complex-$q$ zeros for sizes
887: $N = 6, \cdots, 9$
888: are presented in Figures 3 and 4.
889: In both examples are obtained the zeros lying on the arc-shaped curves,
890: while the positive real axis is free of zeros. The zeros closest to it
891: approach to it with increasing $N$.
892:
893: As in the mean-field case, we have examined the convergence of the
894: closest zeros
895: to the thermodynamic limit by the BST algorithm, assuming the power-law form.
896: The analysis included the data obtained for chains with $N = 5$ to $9$.
897: The parameter $\omega$ was determined, as earlier, by requiring that
898: $q_2 \,\to\,0$ in the thermodynamic limit.
899: The results are presented in Table 5.
900:
901: \renewcommand
902: %% \baselinestretch 2 % s proredom
903: %% \baselineskip 24pt %
904: \baselinestretch 1 % bez proreda
905: \baselineskip 15pt %
906: \begin{table}[hbt] \label{tb:bstLR}
907: \caption{ The loci of closest zeros and its BST extrapolation at
908: $(K = 0.576, \sigma = 0.3)$ and $(K = 0.823, \sigma = 0.8)$}
909: \begin{center} \begin{tabular}{cccc} \hline
910: & $ K = 0.576$ & $\sigma = 0.3$ & \\ \hline
911: $N$ & $q_1^{(0)}(N)$ & $q_2^{(0)}(N)$ & $\omega$ \\
912: 5 & -0.447413829093298& 2.78806303559461 & \\
913: 6 & 0.0734103695976816& 2.88316232503579 & \\
914: 7 & 0.486108948338677& 2.88942370075814 & \\
915: 8 & 0.818954244961081& 2.85269647435491 & \\
916: 9 & 1.09269817849075& 2.79489503015230 & \\
917: $\infty$ & 4.2 & $3\,\cdot\,10^{-3}$ & 0.89 \\ \hline
918: & $ K = 0.823$ & $\sigma = 0.8$ \\ \hline
919: $N$ & $q_1^{(0)}(N)$ & $q_2^{(0)}(N)$ & $\omega$ \\
920: 5 &-0.418572384698087 & 3.69577294340190 & \\
921: 6 & 0.179124021423254 & 3.55212427658784 & \\
922: 7 & 0.581821465484680 & 3.37723067427777 & \\
923: 8 & 0.866098195024489 & 3.20669717997192 & \\
924: 9 & 1.074705295884560 & 3.05042285906340 & \\
925: $\infty$ & 2.01 & $3\,\cdot\,10^{-3}$ & 0.41 \\
926: \hline \end{tabular} \end{center} \end{table}
927: \renewcommand
928: %% \baselinestretch 2 % s proredom
929: %% \baselineskip 24pt %
930: \baselinestretch 1 % bez proreda
931: \baselineskip 15pt %
932:
933: In the case of second order phase transition the expected
934: crossing point with the real axis is obtained with precision better than 1\%.
935: The leading convergence exponent given by $\omega$,
936: is equal to 0.41,
937: which is close to the value for the temperature critical exponent
938: $1/\nu =0.48$ obtained earlier by finite-range scaling
939: for $q=2$ and $\sigma=0.8$ \cite{GU93}.
940:
941: In the case of the first-order transition the fit is more difficult.
942: The discrepancy between the obtained crossing point and the
943: value expected for the considered temperature
944: is larger, about 20\%, while the convergence exponent, which
945: in context of finite-size scaling should be linear for a first-order
946: phase transition, differs by about 11\%.
947: Larger error in determination
948: of the value of $q$, may be attributed to both lower precision of
949: the data for critical temperature obtained by finite-range scaling
950: \cite{GU93,GU98}
951: when the values of $\sigma$ are low and to the proximity of
952: the second-order transition regime in the $(q,\sigma)$ plane,
953: which might induce crossover effects for small sizes.
954:
955: Since in these preliminary calculations we dispose with small number
956: of data, we have not calculated the density of zeros in the case of
957: power-law model.
958:
959:
960: \section{Conclusion} \label{sec-zklj}
961:
962: We have studied the \cq zeros of the partition function of the ferromagnetic
963: Potts model with long-range interactions in context of the first- and
964: second-order phase transitions in this model.
965: Two cases are presented: the mean-field limit and a few examples
966: of the case with power-law decaying interactions.
967:
968: We conclude that the positions of the zeros may be analysed in a
969: similar way as the zeros in the plane of complex field or
970: temperature and that they can give useful informations on the
971: phase transition in ferromagnetic Potts models.
972:
973: Numerical results for relatively small systems show the similar picture
974: as in the case of zeros in the complex field or temperature plane.
975: In all the cases considered the zeros lie on smooth curves
976: that intersect the real positive axis at value for which the given
977: temperature is the transition temperature.
978: For finite systems the positive real axis is free of zeros, but they
979: approach to it with increasing size.
980:
981: The distance of the closest zero from the real axis is found to scale
982: with size as the temperature critical exponent $1/\nu$.
983: In the mean-field case, this result could be obtained with a good precision
984: for series of different temperatures corresponding to different transition
985: regimes. In the case of power-law interactions, the preliminary results
986: obtained with less precision suggest the same behaviour.
987: This result is similar to behaviour in the complex temperature plane,
988: and it supports recent findings by Kim and Creswick \cite{KC01}
989: in the short-range Potts model in $2d$.
990:
991: In the end, let us mention some technical advantages of the analysis
992: by complex-$q$ zeros compared to the complex-temperature zeros
993: in the model with long-range interactions.
994: The partition function of the mean-field model consisting of $N$
995: spins can
996: be written as a $N$-th order polynomial in $q$, or roughly
997: $N^2$-th order polynomial in temperature variable. Consequently,
998: this allows better numerical precision of calculations in the complex-$q$
999: plane then those in the complex-temperature plane.
1000: For a model with power-law decaying interactions the possibilities
1001: in complex temperature plane are even more restricted.
1002: Model with $N$ spins, has $N/2$ (for $N$ even)
1003: different parameters involving
1004: temperature, so that the partition function cannot be defined as a
1005: polynomial in some temperature variable at all.
1006:
1007: Further investigations of the complex-$q$ zeros should be done in
1008: future in at least two directions.
1009: We plan to perform a more extended numerical calculations of the
1010: complex-$q$ zeros in the case of power-law decaying interactions.
1011: This should allow us to obtain the convergence exponent for the closest
1012: zero with more precision.
1013: In addition, it would be interesting to examine a still
1014: open question of the borderline $q_{c}(\sigma)$ dividing
1015: the first- from the second-order transition regimes within the
1016: formalism of complex-$q$ zeros.
1017: Also, on the basis of the results obtained in this paper, a possible
1018: connection between the zeros in \cq and complex-temperature
1019: plane deserves further investigation.
1020:
1021: \newpage
1022:
1023: \begin{thebibliography}{9}
1024: %% \baselineskip=12pt plus 0.2pt minus 0.2pt
1025: \renewcommand
1026: %% \baselinestretch 2 % s proredom
1027: %% \baselineskip 24pt %
1028: \baselinestretch 1 % bez proreda
1029: \baselineskip 15pt %
1030: \bibitem{YL52} C. N. Yang and T. D. Lee, Phys. Rev. {\bf 87} (1952) 404.
1031: \bibitem{LY52} T. D. Lee and C. N. Yang, Phys. Rev. {\bf 87} (1952) 410.
1032: \bibitem{F65} M. Fisher, in Lectures in theoretical physics vol. 12C p. 1
1033: (University of Colorado Press, Boulder, 1965).
1034: \bibitem{OKSK68} S. Ono, Y. Karaki, M. Suzuki and C. Kawabata,
1035: J. Phys. Soc. Japan {\bf 25} (1968) 54.
1036: \bibitem{KC98} S-Y. Kim and R. J. Creswick, \PRL\ {\bf 81} (1998) 2000.
1037: \bibitem{KC99} S-Y. Kim and R. J. Creswick, Physica A {\bf 281} (2000) 252.
1038: \bibitem{B90} G. Bhanot, \JSP\ {\bf 60} (1990) 55.
1039: \bibitem{M91} P. Martin, Potts models and related problems in
1040: statistical mechanics, World Scientific Publishing
1041: Co. Pte. Ltd., 1991, Ch. XI, and references therein.
1042: \bibitem{CHW96} C-N. Chen, C-K. Hu and F. Y. Wu,
1043: \PRL\ {\bf 76} (1996) 169.
1044: \bibitem{KCCH99} S-Y. Kim, R. J. Creswick, C-N. Chen and C-K. Hu,
1045: Physica A {\bf 281} (2000) 262.
1046: \bibitem{FK69} P. W. Kasteleyn and C. M. Fortuin, J. Phys. Soc. Jpn.
1047: Suppl. {\bf 26} (1969) 11;
1048: C. M. Fortuin and P. W. Kasteleyn, Physica {\bf 57} (1972) 536.
1049: \bibitem{Bx86} R. J. Baxter, \JPA\ {\bf 19} (1986) 2821.
1050: \bibitem{Bx87} R. J. Baxter, \JPA\ {\bf 20} (1987) 5241.
1051: \bibitem{ST97a} R. Shrock and S-H. Tsai, \PRE\ {\bf 55} (1997) 5165.
1052: \bibitem{ST97b} R. Shrock and S-H. Tsai, \PRE\ {\bf 56} (1997) 1342.
1053: \bibitem{RST98} M. Ro\v cek, R. Shrock and S-H. Tsai, \PA\ {\bf 259}
1054: (1998) 367.
1055: \bibitem{ST99a} R. Shrock and S-H. Tsai, \PA\ {\bf 275} (1999) 429.
1056: \bibitem{ST99b} R. Shrock and S-H. Tsai, \JPA\ {\bf 32} (1999) L195.
1057: \bibitem{S99} A. D. Sokal, Physica A {\bf 279} (2000) 324.
1058: \bibitem{SS00} J. Salas and A. D. Sokal, cond-mat/0004330,
1059: to be published in J. Stat. Phys.
1060: \bibitem{CS00} S-C. Chang and R. Shrock,
1061: Physica A {\bf 286} (2000) 189.
1062: \bibitem{KC01} S-Y. Kim and R. J. Creswick, \PRE\ {\bf 63} (2001) 0066107.
1063: \bibitem{W82} F. Y. Wu, Rev. Mod. Phys. {\bf 54} (1982) 235.
1064: \bibitem{KMS54} T. Kihara, Y. Midzuno and T. Shizume,
1065: J. Phys. Soc. Jpn. {\bf 9} (1954) 681.
1066: \bibitem{SF73} J. P. Straley and M. E. Fisher, \JPA\ {\bf 6} (1973) 1310.
1067: \bibitem{MS74} L. Mittag and J. Stephen, \JPA\ {\bf 7} (1974) L109.
1068: \bibitem{PL76} R. G. Priest and T. C. Lubensky, \PRB\ {\bf 13} (1976) 4159.
1069: \bibitem{LB96} E. Luijten and H. W. J. Bl\"ote, \PRL\ {\bf 76} (1996) 1557.
1070: \bibitem{D69} F. J. Dyson, Commun. Math. Phys. {\bf 12} (1969) 91.
1071: \bibitem{ACCN88} M. Aizenman, J. T. Chayes, L. Chayes and C. M. Newman,
1072: \JSP\ {\bf 50} (1988) 1.
1073: \bibitem{GU93} Z. Glumac and K. Uzelac, \JPA\ {\bf 26} (1993) 5267.
1074: \bibitem{LB95} E. Luijten and H. W. J. Bl\"ote,
1075: Int. J. Mod. Phys. C {\bf 6} (1995) 359.
1076: \bibitem{C95} S. A. Cannas, \PRB\ {\bf 52} (1995) 3034;
1077: S. A. Cannas, A. C. N. de Magalh\~aes and
1078: F. A. Tamarit, \PRB\ {\bf 61} (2000) 11521.
1079: \bibitem{UG97} K. Uzelac and Z. Glumac, Fizika B {\bf 6} (1997) 133.
1080: \bibitem{GU98} Z. Glumac and K. Uzelac, \PRE\ {\bf 58} (1998) 4372.
1081: \bibitem{M99} J. L. Monroe, \JPA\ {\bf 32} (1999) 7803.
1082: \bibitem{BDD99} E. Bayong, H. T. Diep and V. Dotsenko,
1083: \PRL\ {\bf 83} (1999) 14.
1084: \bibitem{KL00} M. Krech and E. Luijten, \PRE\ {\bf 61} (2000) 2058.
1085: \bibitem{LM01} E. Luijten and H. Me{\ss}ingfeld, \PRL\ {\bf 86} (2001) 5305.
1086: \bibitem{BS64} R. Bulirsch and J. Stoer, Numer. Math. {\bf 6} (1964) 413.
1087: \bibitem{HS88} M. Henkel and G. Sch{\" u}tz, \JPA\ {\bf 21} (1988) 2617.
1088: \bibitem{VBS79} J. M. Vanden Broeck and L. W. Schwartz,
1089: SIAM J. Math. Anal. {\bf 10} (1979) 658.
1090: \bibitem{BJP82} R. Botet, R. Jullien and P. Pfeuty, \PRL\ {\bf 49} (1982) 478.
1091: \bibitem{BJ83} R. Botet and R. Jullien, \PRB\ {\bf 28} (1983) 3955.
1092: \bibitem{IPZ83} C. Itzykson, R.B. Pearson, and J. B. Zuber,
1093: Nucl. Phys. B {\bf 220}, 415 (1983)
1094: \bibitem{CK97} R. J. Creswick and S-Y. Kim, \PRE\ {\bf 56} (1997) 2418.
1095: \bibitem{UG00} K. Uzelac and Z. Glumac, \PRL\ {\bf 85} (2000) 5255.
1096:
1097: \end{thebibliography}
1098:
1099: \newpage
1100:
1101: \centerline{\Large Figure caption}
1102:
1103: \vspace{1cm}
1104:
1105: {\bf Fig. 1}: The loci of complex-$q$ zeros of the partition function of
1106: the mean-field Potts
1107: model calculated at inverse temperatures
1108: (a) $K = K_c(q = 0.5) = 0.5$, (b) $K = K_c(q = 2) = 2$ and
1109: (c) $K = K_t(q = 8.0) = 4.540457$ for
1110: $N = 25$ (open diamonds), $N = 30$ (filled triangles)
1111: and $N = 35$ (filled circles) spins.
1112:
1113: {\bf Fig. 2}: The locus of complex-$q$ zeros of the partition function of
1114: the mean-field Potts
1115: model calculated at inverse temperature
1116: $K = K_{MF}(q = 8) = 4.540457$.
1117: The full line denotes loci of zeros in the limit $N\to\infty$,
1118: obtained by the approach by the saddle point approach
1119: as a solution of the equation $\cF_R = 0$.
1120: Long dashed line is a circle (described in text) drawn as a guide to the eye.
1121:
1122: {\bf Fig. 3}: The complex-$q$ zeros of the partition function of
1123: the $1 d$
1124: Potts model with power-law interactions at the $FRS$ estimate of first-order
1125: transition temperature
1126: $K = K_t(q = 5, \sigma = 0.3) = 0.576$,
1127: for
1128: $N = 6$ (open diamonds), $N = 7$ (filled triangles),
1129: $N = 8$ (open squares) and $N = 9$ (filled circles) spins.
1130:
1131: {\bf Fig. 4}: The complex-$q$ zeros of the partition function of the $1 d$
1132: Potts model with power-law interactions at the $FRS$
1133: estimate of second-order transition temperature
1134: $K = K_c(q = 2, \sigma = 0.8) = 0.8230$,
1135: for
1136: $N = 6$ (open diamonds), $N = 7$ (filled triangles),
1137: $N = 8$ (open squares) and $N = 9$ (filled circles)
1138: spins.
1139:
1140:
1141:
1142: \vfill
1143: \eject
1144:
1145: \centerline{\includegraphics{Fig1a.eps}}
1146:
1147: \centerline{\includegraphics{Fig1b.eps}}
1148:
1149: \centerline{\includegraphics{Fig1c.eps}}
1150:
1151: \centerline{\includegraphics{Fig2.eps}}
1152:
1153: \centerline{\includegraphics{Fig3.eps}}
1154:
1155: \centerline{\includegraphics{Fig4.eps}}
1156:
1157: \end{document}
1158:
1159:
1160: