1: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb, prl]{revtex4}
2: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
3: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
4: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
5: \usepackage{graphicx}
6:
7:
8:
9:
10: \begin{document}
11:
12: \title {Is the mean-field approximation so bad?\\
13: A simple generalization yelding realistic critical indices for 3D
14: Ising-class systems}
15: %\title {A simple microscopic approach \\ yelding correct
16: %critical indices for the 3D Ising-class systems}
17:
18: \author{A.N. Rubtsov}
19:
20: \email{alex@shg.ru}
21:
22: \affiliation {Physical Department of Moscow State University,
23: Moscow 119899, Russia}
24:
25: \date{\today}
26:
27: \begin{abstract}
28: Modification of the renormalization-group approach, invoking
29: Stratonovich transformation at each step, is proposed to describe
30: phase transitions in 3D Ising-class systems. The proposed method
31: is closely related to the mean-field approximation. The low-order
32: scheme works well for a wide thermal range, is consistent with a
33: scaling hypothesis and predicts very reasonable values of
34: critical indices.
35: \end{abstract}
36:
37: \pacs{05.70.Fh}
38:
39: \maketitle
40:
41: It is well understood since 1960's that an essential increase of
42: fluctuations accompanies a second-order phase transition. Exact
43: results for 2D Ising model \cite{Ising2D} showed that Landau
44: phenomenology \cite{Landau} was too crude to describe the vicinity
45: of the transition point. Ginzburg demonstrated that it is
46: correlations what breaks Landau approach down \cite{Ginzburg}. He
47: established the parameter which determines the thermal range for
48: the so-called fluctuation region.
49: %The picture of elementary
50: %excitations is broken in this region when the dimensionality of a
51: %system is lower than the upper critical dimension $D_{up}$ [].
52: According to the scaling hypothesis \cite{PP}, the properties of
53: a system in the fluctuation region are determined by the single
54: quantity $r_c$, being the correlation length for fluctuations.
55: This parameter shows a power-law dependence on the external field
56: $h$ and on the thermal interval $t$ from the transition:
57: $r_c\propto h^{-\mu}$, $r_c\propto t^{-\nu}$. Values of $\mu$ and
58: $\nu$ are determined by the universal characteristics, like the
59: dimensionality $D$, the number of components of the order
60: parameter, and the type of the interaction in the system
61: (long-range or short-range). Other critical indices can be
62: expressed via $\mu$ and $\nu$. For instance for the scalar order
63: parameter $\bar{x}$ one obtains $\bar{x}\propto t^{-\beta},
64: \beta=\nu D - \nu/\mu$.
65: %Above $D_{up}$
66: %fluctuations are not crucial: in this case critical indices
67: %coincide the predictions of the phenomenological Landau theory,
68: %which totally neglects fluctuations. Below $D_{up}$ the values of
69: %indices are different.
70:
71: The general framework to obtain critical indices is the
72: renormalization-group (RG) approach \cite{Wilson,RevRG,ERGE}. The
73: RG transformation consists of the two stages. First, one
74: integrates out all $x_k$ with wave vectors $k>k_0/s$, where $k_0$
75: is a cut-off and $s>1$ is a parameter. The main assumption of the
76: method is that the potential energy of thus renormalized system
77: can be truncated to certain simplified form, for example to the
78: form of $\phi^4$ model. The second step is a scale transformation.
79: It is aimed to restore the cut-off for $k$ back to $k_0$ and
80: preserve the original dispersion law for the long-wavelength
81: modes.
82:
83:
84: The RG transformation should be applied recursively; its unstable
85: fixed point corresponds to a phase transition. Lyapunov exponents
86: for the vicinity of the fixed point determine the critical
87: indices. The scaling hypothesis requires exactly two of the
88: Lyapunov exponents to be positive, they correspond to the two
89: independent critical indices.
90:
91: With respect to the value of $s$, two modifications of the RG
92: scheme can be distinguished. Wilson's $\varepsilon$ expansion
93: \cite{Wilson, RevRG} deals with $s \gg 1$. The method is formally
94: valid near the upper critical dimension $D_{up}$ (for the
95: Ising-class classical systems $D_{up}=4$).The ''most divergent``
96: diagram series can be pointed out in this case. This gives the
97: asymptotic expansion in terms of $\varepsilon\equiv D_{up}-D$. In
98: the continuous version of the scheme (so-called exact RG approach)
99: \cite{ERGE}, $s$ is infinitesimally close to unity. Certain
100: decoupling (truncation) should be introduced in this case to solve
101: the integro-differential equation for RG flow. An accuracy is then
102: determined by a form of the decoupling used.
103: %\cite{Wilson,RevRG,ERGE}.
104:
105:
106: All the approaches mentioned above are very well-developed. They
107: allow to calculate critical indices and universal combinations of
108: amplitudes with a high accuracy. Progress is also achieved in the
109: calculation of non-universal quantities; particularly a crossover
110: between Gaussian (Landau-like) and critical behaviour is analyzed
111: \cite{crossover}. However RG is applicable only for
112: long-wavelength excitations since they can be described as a
113: fluctuating field. Therefore, in a practical calculation,
114: short-wavelength excitations should be first integrated out by
115: means of some perturbation expansion, and then RG should be
116: applied. The short- and the long-wavelength excitations should be
117: separated {\it ad hoc}. Such a two-step procedure is not very
118: elegant. At least from the methodological point of view it is
119: desired to develop a scheme which gives all the information in a
120: uniform manner.
121:
122: It is interesting to note that calculations are much simpler
123: outside the critical region. In fact, even the mean-field
124: approximation (MFA) is usually sufficient for the qualitative
125: analysis of this region. In MFA, the non-local interaction in the
126: system is replaced by an effect of a self-consistent average
127: field. The MFA predictions for the transition temperature and
128: thermal behaviour of the order parameter in 3D case are quite
129: reasonable \cite{MFA}. On the other hand MFA gives Landau set of
130: indices and is not very accurate. Note that Landau approach itself
131: has a mean-field nature.
132:
133: In this paper, we present a simple generalization of MFA,
134: performing well for the 3D Ising-class systems both nearby and far
135: from the transition. Particularly the critical behaviour is
136: reproduced correctly: the values of indices are predicted with a
137: few-percent accuracy. Essentially, the present approach consists
138: in the alternating application of the Stratonovich transformation
139: and the most primitive version of the renormalization-group
140: transformation at a finite value of $s$.
141:
142: We consider a classical scalar field with an anharmonic local
143: potential and a short-range harmonic interaction at temperature
144: $T$. The partition function of this system is
145: %\begin{equation}\label{Ini}
146: % V=\int U(x_r+\bar{x}) d {\bf r}+\sum_k \frac{\Omega_k |x_k|^2}{2}.
147: %\end{equation}
148: \begin{eqnarray}\label{Z0}
149: Z&=&\int [Dx] e^{-V/T}; \\ \nonumber
150: V&=&\int U(x_r+\bar{x}) d {\bf r}+\sum_k \frac{\Omega_k
151: |x_k|^2}{2}.
152: \end{eqnarray}
153: Here $U(x+\bar{x})$ is an even on-site potential, $x_k$ is a
154: Fourier transform of the displacement $x_r$ from the average
155: position $\bar{x}$; $\bar{x} \neq 0$ below the transition
156: temperature, and the average of $x_r$ equals zero by definition.
157: To make the definitions unambiguous, we put additional conditions
158: $U(\bar{x})=0;\Omega_{k=0}=0$. The wave vector $k$ does not exceed
159: the unititary cut-off: $k^2 \le 1$. The case of a short-range
160: interaction corresponds to the asymptotic behaviour
161: \begin{equation} \label{om}
162: \Omega_k = \omega k^2, k \to 0.
163: \end{equation}
164:
165:
166: The simplest way to integrate out the modes is to skip $x_k$ with
167: $k>1/s$. This corresponds to the zero-th approximation in the
168: nonlinearity of $U$. The subsequent scaling results in the RG
169: transformation $k \to s k, U \to s^3 U, x \to x/\sqrt{s}$. The
170: defect of thus applied RG method comes from the the increase of
171: correlations (non-linear part of $U$ grows). Therefore, after
172: several iterative transformations, the RG procedure is not valid
173: any more. Although the formal analysis of the fixed point $U(x)=0$
174: can be done, it wrongly gives Landau-like critical indices.
175:
176: To avoid the problem higher-order approximations and the
177: truncation of the potential are used in known schemes described
178: above (a very clear explanation is given in \cite{fermi}). Here we
179: use another procedure: the zero-th approximation is used, but the
180: change of variables is made at each RG step. The change of
181: variables reduces correlations and is aimed to compensate its
182: increase at RG transformation.
183:
184: At the first stage, we integrate out (omit) all the modes with
185: $\Omega_k>\bar{\Omega}$, where $\bar{\Omega}=\frac{3}{4 \pi} \int
186: \Omega_k d^3 k$ is an average of $\Omega_k$. For the rest of the
187: modes, the Stratonovich identity is utilized:
188: \begin{widetext}
189: \begin{equation}\label{Str}
190: \exp\left(-\frac{(\Omega_k-\bar{\Omega})
191: |x_k|^2}{T}\right)=\frac{\bar{\Omega}-\Omega_k}{\pi
192: \bar{\Omega}^2} \int \exp \left( -\frac{1}{T}
193: \left( 2 \bar{\Omega} {\rm Re}(x_k
194: f_k^*)+\frac{\bar{\Omega} \Omega_k |f_k|^2}{\bar{\Omega}-\Omega_k}
195: \right) \right)df_k.
196: \end{equation}
197: Here the integration over the complex variable $f_k=\Re(f_k)+i
198: \Im(f_k) $ denotes the integration over the complex plane: $\int
199: df_k \equiv \int_{-\infty}^{\infty} d \Re(f_k)
200: \int_{-\infty}^{\infty} d \Im(f_k)$.
201:
202: After (\ref{Str}) is substituted into (\ref{Z0}), expressions
203: $\sum_k {\rm Re}(x_k f_k^*) = \int x_r f_r d \bf {r}$ and $\sum_k
204: \bar{\Omega} |x_k|^2 = \int \bar{\Omega} x_r^2$ can be collected
205: in the exponent. Therefore
206: \begin{equation}
207: Z\propto \int [Dx] \int [Df] \exp\left(-\frac{1}{T}
208: \left(\int \left(U(x_r+\bar{x})-\Omega_k f_r x_r + \frac{\bar{\Omega} x_r^2}{2}\right) d {\bf r}
209: +\sum_k \frac{\bar{\Omega} \Omega_k |f_k|^2}{
210: 2 (\bar{\Omega}-\Omega_k)}\right)\right),
211: \end{equation}
212: where $f_r$ is the inverse Fourier transform of $f_k$.
213:
214: Now, we can integrate over $x_r$. Let us introduce the function
215: $F(f)$ accordingly to the equation:
216: \begin{equation}\label{Fdef}
217: \exp \left(-\frac{F(f+\bar{x})-F_0}{T}\right)= \int \exp
218: \left(-\frac{1}{T}\left(U(x+\bar{x})+\frac{\bar{\Omega}(x-f)^2}{2}\right)\right) dx,
219: \end{equation}
220: \end{widetext}
221: where $F_0$ is a constant, defined from the condition
222: $F(\bar{x})=0$.
223:
224: With thus defined $F$, the partition function takes the form
225: \begin{eqnarray}\label{ZF}
226: Z&\propto & \int [Df] e^{-W/T}; \\ \nonumber
227: W&=&\int F(f_r+\bar{x}) d{\bf r}
228: +\sum_k \frac{\bar{\Omega} \Omega_k |f_k|^2}{
229: 2 (\bar{\Omega}-\Omega_k)}.
230: \end{eqnarray}
231:
232: Equation (\ref{ZF}) is formally very similar to Eq.(\ref{Z0}).
233: Function $F$ is even; the dispersion at $k \to 0$ coincides that
234: of Eq.(\ref{Z0}). One can also guess from Eq.(\ref{Str}) that as
235: the average of $x_r$ equals zero, the average of $f_r$ also
236: vanishes. Indeed, if this average would not be zero, $f_{k=0}$
237: would fluctuate around a macroscopically large average value, and
238: (\ref{Str}) would not be fulfilled.
239:
240: We argue here that variables $f_k$ are much less correlated, than
241: $x_k$. Let us consider the parabolic approximation for $F$:
242: $F(\bar{x}+f)\approx \frac{\partial F(\bar{x})}{\partial f} f +
243: \frac{\partial^2 F(\bar{x})}{2 \partial f^2} f^2$. Variables
244: $f_k$ are uncorrelated in this approximation. Obviously $<f>$
245: vanishes only if $\frac{\partial F(\bar{x})}{\partial f}=0$, that
246: gives
247: \begin{equation}\label{MF}
248: \frac{\bar{\Omega}\bar{x}}{T}=
249: \frac{\int x \exp \left(-T^{-1}\left(U(x+\bar{x})+\frac{1}{2}\bar{\Omega}x^2 \right)\right) dx}
250: {\int \exp \left(-T^{-1}\left(U(x+\bar{x})+\frac{1}{2}\bar{\Omega}x^2 \right)\right)
251: dx}.
252: \end{equation}
253:
254: This is exactly the mean-field equation of state for the system
255: (\ref{Z0}). So, even the complete neglect of correlations in $f_k$
256: still gives something reasonable. On the other hand, the parabolic
257: approximation for $U$ directly in Eq.(\ref{Z0}) gives non-sense:
258: since $U(x)$ does not depend on temperature, the system does not
259: show a transition at all in this approximation. Such a comparison
260: convinces us that correlations in $f_k$ are less essential, than
261: in $x_k$. As it was pointed above we expect that this can
262: compensate an increase of correlations in RG transformation.
263:
264: Now, we finish the RG transformation, operating with
265: Eq.(\ref{ZF}). All $f_k$ with $k>1/s$ are to be integrated out
266: (omitted in our approximation). After the scale transformation $k
267: \to s k, f \to x/\sqrt{s}$ is done, the potential energy of the
268: system takes the form
269: \begin{equation}
270: \int U^{(1)}(x_r+\bar{x}^{(1)}) d {\bf r}+\sum_k \frac{\Omega^{(1)}_k
271: |x_k|^2}{2},
272: \end{equation}
273: where
274: \begin{equation} \label{Trans_k}
275: \Omega^{(1)}_k=\frac{s^2 \bar{\Omega}
276: \Omega_k}{\bar{\Omega}-\Omega_{k/s}},
277: \end{equation}
278:
279:
280: \begin{equation} \label{Trans_w}
281: U^{(1)}(x+\bar{x}^{(1)})=s^3 F(x/\sqrt{s}+\bar{x}), ~~\bar{x}^{(1)}=\bar{x} \sqrt{s}.
282: % U^{(1)}(y+\bar{x}^{(1)})=F_0 s^3-\\ T s^3 \ln \left( \int \exp
283: % \left(-\frac{1}{T}\left(U(x+\bar{x})+\frac{\bar{\Omega}(x-y/\sqrt{s})^2}{2}\right)\right)
284: % dx\right).
285: \end{equation}
286: %Here $$, $U(y)=s^3 F(y / \sqrt{s})$, $U^{(1)}(\bar{x}^{(1)})=0$.
287:
288: Such a transformation should be applied recursively, like in the
289: standard RG approach. The analysis can be easily performed
290: numerically.
291: %The proposed modification of the RG method is illustrated in Fig. 1b.
292:
293:
294: The sequence $\Omega_k, \Omega_k^{(1)}, \Omega_k^{(2)}, ...$ is
295: independent of $U$ and $T$. At every step a new
296: $\bar{\Omega}^{(n)}$ are to be calculated. After several
297: transformations $\bar{\Omega}^{(n)}$ and $\Omega_k^{(n)}$ converge
298: to certain limit. Calculation from formula (\ref{Trans_k}) gives
299: \begin{equation}
300: \Omega_k^{(\infty)}=\frac{\omega \bar{\Omega}^{(\infty)} (s^2-1) k^2}
301: {\bar{\Omega}^{(\infty)} (s^2-1)-\omega k^2};
302: \end{equation}
303: $\bar{\Omega}^{(\infty)}$ is implicitly defined from the equation
304: $\bar{\Omega}^{(\infty)}=\frac{3}{4 \pi} \int \Omega_k^{(\infty)}
305: d^3 k$.%, and $\omega$ is defined in (\ref{om}).
306:
307: %The convergence is quite fast; actually $\Omega_k^{(1)}$ is
308: %already very similar to $\Omega_k^{(\infty)}$.
309:
310: Now, let us consider the properties of the sequence $U, U^{(1)},
311: U^{(2)},...$ at $\bar{x}=0$. Transformation $U^{(n)} \to
312: U^{(n+1)}$ has a non-trivial unstable fixed point $U_f(x)$. This
313: function is shown in the upper panel of Fig.1. Consider small
314: deviations from this fixed point: $U^{(n)}(x)=U_f(x)+u_n(x)$. The
315: linearization of the transformation (\ref{Trans_w}) gives
316: %\begin{widetext}
317: \begin{equation}\label{lin}
318: u_{n+1}(y)=u^0+s^3 \int u_n(x) A(x,y) dx;
319: \end{equation}
320: $$A=\exp\left(- \frac{U_f(x)- s^{-3} U_f(y) + F + \frac{\bar{\Omega}}{2} (x-y/\sqrt{s})^2}{T}\right).$$
321: %\end{widetext}
322: where $u^0$ stands to fulfill the condition $u_{n+1}(0)=0$. The
323: analysis of this operator shows that among its eigenvalues only
324: two exceed unity. Denote them $s^{\lambda_e}$ and $s^{\lambda_o}$;
325: they correspond to an even and odd eigenvector, respectively. Thus
326: the theory is consistent with the scaling hypothesis, as it was
327: mentioned in an introductory part.
328:
329: \begin{figure}
330: \includegraphics{1}
331: \caption{\label{fig:epsart} Upper panel: the fixed point of the
332: transformation (\ref{Trans_w}) at $s=2.5$. Lower panel: critical
333: index $\nu$ {\it vs.} scaling factor $s$. Note a very weak
334: dependence of $\nu(s)$.}
335: \end{figure}
336:
337: Exponents $\lambda_e$ and $\lambda_o$ are the inverse of the
338: critical indices $\nu$ and $\mu$, respectively \cite{RevRG,
339: fermi}. In the ideal case, critical indices should be independent
340: of the particular $U(x)$ and $\Omega_k$, as well as of the value
341: of the factor $s$. Indeed, after a proper re-scaling of the units,
342: the values of $T$ and $\bar{\Omega}$ drop out from
343: Eq.(\ref{Trans_w}). Therefore in our model $\nu$ and $\mu$ can
344: depend only on $s$. Further, it turns out that $\mu=2/5$ at any
345: $s$, as it should be for the local-potential approximation
346: \cite{ERGE}. What is about $\nu$, it depends on $s$ very weakly.
347: The numerical result for $\nu(s)$ is shown in the lower panel of
348: Fig.1.
349:
350:
351:
352: On the other hand it is important to point out that the crude
353: universal properties of the model enter the transformation
354: (\ref{Trans_w}) essentially. The behaviour $\Omega_{k\to 0}
355: \propto k^2$ corresponds to a short-range range interaction. Such
356: behaviour and the dimensionality of the model determine the scale
357: transformation, {\it i.e.} powers of $s$ in Eq.(\ref{Trans_w}). It
358: is also directly reflected in the formula that the order parameter
359: is a scalar.
360:
361: For the more detailed analysis it is reasonable to consider an
362: extreme value of $\nu(s)$, because in this case small variations
363: of $s$ do not affect $\nu$. This gives $s \approx 2.5, ~~\nu
364: \approx 0.652$. The obtained values of indices are collected in
365: Table 1. An agreement with the numerical results for the 3D Ising
366: model is even better than one could expect. Since the scaling
367: hypothesis holds in the model, the values of all other critical
368: indices can be determined and should have a small error bar. As an
369: example, the calculated value of $\beta$ is presented in Table~1.
370:
371: \begin{table}
372: \caption{%\label{tab:table1}
373: Values of the critical indices obtained from the Lyapunov
374: exponents for the transformation (\ref{lin}) at $s=2.5$. Results
375: of the Landau theory and known values \cite{RevRG} for 3D Ising
376: model are given for the comparison.}
377: \begin{ruledtabular}
378: \begin{tabular}{lccc}
379: %\hline
380: Index& Estimation from Eq.(\ref{lin}) & 3D Ising & Landau theory\\
381: $\nu$& 0.652 & 0.63 & 1/2\\
382: $\mu$ & 2/5 & 0.403 & 1/3\\
383: $\beta$& 0.326 & 0.327 & 1/2\\
384: \end{tabular}
385: \end{ruledtabular}
386: \end{table}
387:
388:
389: The advantage of the present approach is that it allows to
390: estimate not only critical indices, but also the (non-universal)
391: thermodynamic quantities in a wide thermal range. Here we
392: calculate the dependence of the order parameter $\bar{x}$ on
393: temperature for the 3D Ising model. The model is defined as a
394: discrete cubic lattice with $U(x)=\delta(x-1)+\delta(x+1)$ and the
395: nearest-neighbors interaction $\Omega_k=3-\cos \pi k_x-\cos \pi
396: k_y-\cos \pi k_z$. The initial Brillouin zone is a cube
397: $-1<k_{x,y,z}<1$; at the first step the value $\bar{\Omega}=3$ is
398: used and all the modes with $|k|>1/s$ are integrated. The
399: procedure for further steps is exactly as described above. To
400: simplify the calculation, we put
401: $\bar{\Omega}^{(n)}=\bar{\Omega}^{(\infty)}$ at these steps; this
402: simplification almost does not change the result obtained. To
403: determine the value of $\bar{x}$ we use the condition
404: \begin{equation}\label{c_eta}
405: \frac{\partial}{\partial x} U_n(\bar{x}_n)=0, n \to \infty
406: \end{equation}
407: (compare with the derivation of formula (\ref{MF})). Numerically,
408: it is enough to calculate $U_n$ up to $n \approx 8$. The result
409: for $\bar{x}^2(T)$ is presented in Fig.2. The mean-field and
410: experimental (Monte-Carlo) data are sketched in the same figure
411: for comparison. Far from the transition $U^{(1)}$ is almost
412: parabolic, therefore the theory passes into MFA. The critical
413: behaviour occurs at the transition point. Note that Ising model is
414: very non-linear, but the scheme still performs good.
415:
416:
417: \begin{figure}
418: \includegraphics{2}
419: \caption{\label{fig:fig2} The dependence of the square of the
420: order parameter on the temperature in the 3D Ising model. Solid
421: curve: the present approach with $s=2.5$; the dot curve:
422: mean-field calculation; the dash curve: numerical results. Thin
423: ticks show the tangents to $\bar{x}^2(T)$ in the transition
424: point.}
425: \end{figure}
426:
427:
428: We conclude that the presented low-order scheme describes
429: correctly all the physics of the second-order phase transitions in
430: 3D systems with a short-range interaction and a scalar order
431: parameter. It gives a unified description both of the fluctuation
432: region and of the system far from the transition point. The method
433: does not pretend to be very accurate; its advantage consists in
434: simplicity. In principle, one can improve the accuracy by a
435: higher-order procedure for the integration in RG transformation.
436: An applicability of the method for another universality classes is
437: the subject of further studies.
438:
439: The work was supported by RFFI foundation (grant 00-02-16253) and
440: by the "Russian Scientific Schools" program (grant 96-1596476).
441:
442:
443: \begin{thebibliography}{33}
444: \bibitem{Ising2D}L. Onsager Phys. Rev. {\bf 65} 117 (1944).
445: \bibitem{Landau}L.D. Landau Phys.Z. SU {\bf 11} 26 (1937), {\it
446: ibid} {\bf 11} 545 (1937).
447: \bibitem{Ginzburg}V.L.Ginzburg Sov.Phys.Solid State {\bf 2} 1824
448: (1960). % FTT 2 2031
449: \bibitem{PP}L.P. Kadanoff Physics {\bf 2} 263 (1966), A.Z. Patashinskii, V.L.
450: Pokrovskii Sov. Phys. JETP {\bf 50} 439 (1966).
451: \bibitem{Wilson} K.G. Wilson and M.E. Fisher Phys. Rev. Lett. {\bf
452: 28} 240 (1972).
453: \bibitem{RevRG} for a recent review and further references, see A. Pelissetto and E. Vicari e-print
454: cond-mat/0012164.
455: \bibitem{ERGE} F.J. Wegner and A. Houghton Phys. Rev. A {\bf 8}.
456: 401 (1973), C. Bagnuls and C. Bervillier, Physics Reports, {\bf
457: 348} 91 (2001).
458: %\bibitem{books} C. Itzykson and J.M. Drouffe Statistical field
459: %theory vol. I Camrgige University Press, Cambrige, England (1989)
460: \bibitem{crossover} C.Bagnus and C. Berviller Phys.Rev. B {\bf 32}.
461: 7209 (1985).
462: \bibitem{MFA} S. Aubry J. Chem. Phys. {\bf 62} 3217 (1975), A.N. Rubtsov, J. Hlinka, and
463: T. Janssen Phys. Rev. E {\bf 61} 126 (2000).
464: \bibitem{fermi} R.Shankar Rev. Mod. Phys. {\bf 66} 129 (1994).
465: \end{thebibliography}
466:
467: \end{document}
468: