cond-mat0601333/text.tex
1: \documentclass[twocolumn,preprintnumbers,amsmath,amssymb,prl]{revtex4}
2: \usepackage{graphicx}
3: 
4: 
5: 
6: 
7: 
8: \begin{document}
9: \title{Small parameter for lattice models with strong interaction}
10: \author{A.N. Rubtsov}
11: \email{alex@shg.ru} \affiliation{Department of Physics, Moscow
12: State University, 119992 Moscow, Russia}
13: \begin{abstract}
14: Diagram series expansion for lattice models with a localized
15: nonlinearity can be renormalized so that diagram vertexes become
16: irreducible vertex parts of certain impurity model. Thus
17: renormalized series converges well in the very opposite cases of
18: tight and weak binding and pretends to describe in a regular way
19: strong-correlated systems with localized interaction. Benchmark
20: results for the classical $O(N)$ models on a cubic lattice are
21: presented.
22: \end{abstract}
23: %
24: 
25: \maketitle
26: 
27: 
28: One of the key problems of the modern condensed-matter theoretical
29: physics is the quest for a regular description of the strong
30: correlations \cite{Fulde}. Roughly speaking, correlations are
31: strong if the nonlinearity and coupling are comparable, and
32: therefore perturbative approaches make no sense.
33: 
34: Consider a lattice model with local coupling between cells with a
35: nonlinearity localized inside each cell. In the two limit cases of
36: the weak and tight binding a series expansion can be performed in
37: powers of the respective small parameters \cite{Balesku}. One of a
38: few regular ways to handle the situation of crossover is to employ
39: the concept of {\it localized} correlations. It basically means
40: that interaction of each cell with the rest of the system can
41: approximately be described as an interaction with certain
42: non-correlated effective bath \cite{Cowley}. The parameters of
43: this bath are to be found self-consistently. If there is, say, one
44: particle per cell, it means that real multi-site problem is
45: approximated by much simpler single-site one ("impurity problem").
46: An example of such an approach is the mean-field description of
47: the classical statistical ensembles of interacting particles. In
48: this scheme, interaction of a particle with its fluctuating
49: surrounding is replaced with interaction of a particle with a
50: constant average field. There is a self-consistent equation for
51: this field. An accuracy can be dramatically increased for change
52: the mean-field by Gaussian-bath approximation. In this case, the
53: surrounding is approximated by certain Gaussian ensemble. It is
54: important that almost the same argumentation can be presented for
55: very wide class of systems. Particularly, so-called dynamical
56: mean-field theory (DMFT) for strongly interacting fermions is
57: essentially a Gaussian-bath approximation \cite{DMFT}.
58: 
59: 
60: Physically, Gaussian-bath approximation consists in the assumption
61: that the irreducible part of high momenta is localized inside a
62: single site. The strong point of this assumption is that it is
63: valid for several very different limits for systems with localized
64: nonlinearity. The approximation reproduces leading-order results
65: of the weak- and tight-binding series expansion simultaneously.
66: Indeed, if the nonlinearity is small, the perturbation theory
67: explicitly says that correlations are localized in the first-order
68: of weak-binding approximation. On the other hand, if coupling goes
69: to zero, sites become almost isolated and correlations are
70: obviously localized, no matter how strong the nonlinearity is.
71: Once the two opposite limits are described well, their crossover
72: can be expected to be also somehow depicted. There is some more
73: justifications for the scheme, particularly it becomes exact for a
74: system of an infinite dimensionality.
75: 
76: 
77: 
78: Since an analysis of proper impurity problem gives a good
79: understanding of the properties of strongly-correlated systems, it
80: is desirable to construct a perturbation theory starting from the
81: solution of an impurity problem as a zeroth approximation. A
82: theory of this kind is presented in this paper for lattice models
83: with a localized nonlinearity. We exactly renormalize diagram
84: expansion so that diagram vertexes become irreducible vertex parts
85: of certain impurity problem. An important peculiarity of the
86: proposed approach is that a good convergence is achieved both in
87: tight-binding and weak-binding situations. As an example, we
88: present the results obtained for quantum Ising model in a
89: transverse field.
90: 
91: 
92: %\section{The system}
93: 
94: We consider statistics of the real vector field $\phi$ on a
95: discrete lattice with a localized nonlinearity $\Theta(\phi)$ and
96: the dispersion law $\Omega$. In general, lattice potential can
97: depend on a number of the lattice site.
98: %If the on-site potential was absent,
99: %the average of $\phi$ would have been $\phi^0$.
100: Through the article we number sites by $x$ whereas components of
101: the on-site field $\phi_x$ are numbered by $j$, statistical
102: average is denoted by triangle brackets:
103: \begin{eqnarray}\label{Zphi}
104:     &<...>=Z^{-1} \int ... ~ e^{-\beta E(\phi)} [D\phi]\\    \nonumber
105:     &Z=\int e^{-\beta E(\phi)} [D\phi]\\    \nonumber
106:     &E(\phi)=\sum_x \Theta_x(\phi_x) + \frac{1}{2} (\phi
107:     \Omega \phi).
108: \end{eqnarray}
109:  Scalar
110: product $(\phi \Omega \phi)$ is a sum over both indices: $\sum_{x
111: x' j j'} \phi_{xj} \Omega_{x x' j j'} \phi_{x'j'}$; below we also
112: use external product denoted by the dot: $\phi \cdot \phi$. The
113: notation $(\phi_x \Omega_{xx} \phi_x)$ is used for a scalar
114: product in the on-site subspace: $(\phi_x \Omega_{xx}
115: \phi_x)\equiv \sum_{j j'} \phi_{xj} \Omega_{x x j j'} \phi_{xj'}$.
116: 
117: 
118: 
119: Consider an auxiliary ensemble with energy
120: \begin{equation}\label{auxil}
121: E^{aux}(\phi)=\sum_x \Theta_x(\phi_x)+\frac{1}{2} \left(\phi A
122: \phi\right)+ (\phi^0 (\Omega-A) \phi).
123: \end{equation}
124: We introduced here vector quantity $\phi^{0}$ and a Hermitian
125: tensor $A$ having non-zero terms only at the same $x$-indices:
126: $A_{xx'}=0$ for $x\neq x'$ (through the text we call such
127: quantities '$x$-diagonal'). The values of $A$ and $\phi^0$ are not
128: specified for a while.
129: 
130: Since $A$ is $x$-diagonal, site oscillators of the auxiliary
131: ensemble are uncoupled (one can see that $E^{aux}(\phi)$ splits to
132: a sum of independent single-site energies). Thus, properties of
133: the auxiliary system at each site can be obtained as a solution of
134: a single-site impurity problem.
135: 
136: Denote an average over the auxiliary ensemble by overline and
137: introduce normalized two-point correlator and higher-order
138: irreducible vertex parts at the sites of auxiliary system:
139: \begin{widetext}
140: \begin{eqnarray}\label{gammas}
141:     &g_{12}=\beta \left(\overline{\phi_1\phi_2}-\overline{\phi_1}~\overline{\phi_2}\right)\\
142:     \nonumber
143:     &\gamma^{(3)}_{1'2'3'}=-\beta^3 g_{1'1}^{-1} g_{2'2}^{-1} g_{3'3}^{-1}
144:     \left(\overline{\phi_1 \phi_2 \phi_3} -
145:     \overline{\phi_1 \phi_2} ~ \overline{\phi_3} - \overline{\phi_1 \phi_3}~ \overline{\phi_2}
146:     -\overline{\phi_2 \phi_3} ~ \overline{\phi_1}+2 \overline{\phi_1} ~ \overline{\phi_2} ~
147:     \overline{\phi_3}\right)
148:     \\ \nonumber &\gamma^{(n)}_{1'..n'}=(-\beta)^n g_{1'1}^{-1} ... g_{n'n}^{-1} \cdot ({\rm irreducible ~part ~of ~the}~
149:     n{\rm-th~moment}).
150: \end{eqnarray}
151: \end{widetext}
152: To simplify the notation, we omit site index $x$ and write numbers
153: as subscripts for the on-site coordinates here.
154: 
155: It should be noted that mean-field and Gaussian-bath
156: approximations can be understood as approximations of the system
157: (\ref{Zphi}) by (\ref{auxil}) with a self-consistent requirement
158: $\phi^0=\overline{\phi}$ and certain choice of $A_{xx}$. Indeed,
159: to obtain the mean-field scheme, one should, first, replace all
160: $\phi$ in $E(\phi)$ with its average for all cells except just a
161: single one, and, second, use an average over this particular cell
162: as a guess for $<\phi>$. It is easy to check that the  mean-field
163: corresponds to the simple choice $A_{xx}=\Omega_{xx}$.
164: Gaussian-bath approximation also reduces the system to a
165: single-site problem. The difference is that the remaining degrees
166: of freedom are not frozen in its average positions, but
167: approximated with harmonic oscillators, forming Gaussian bath.
168: These Gaussian degrees of freedom can be integrated out, giving
169: again single-site energy like in (\ref{auxil}) with certain $A$.
170: An approximation for the single-site system by a harmonic
171: oscillator can be found from the requirement that the
172: approximation mimics $\overline{\phi_x}$ and $g_{xx}$. Thus
173: obtained parameters of harmonic oscillators are to be used as in
174: the next iteration of the self-consistent loop as parameters of
175: the oscillators forming the bath. These iterations converge to the
176: point where $x$-diagonal part of the tensor
177: $\left(g+(\Omega-A)^{-1}\right)^{-1}$ vanish. This is in fact a
178: condition for $A$ of the Gaussian-bath approximation.
179: 
180: 
181: 
182: The aim of this paper is to express averages (\ref{Zphi}) via
183: $\overline{\phi}$, $g$ and $\gamma^{(n)}$ by certain regular
184: series. The quantities $A$ and $\phi_0$ will determined by certain
185: self-consistent conditions.
186: 
187: First, we would like to introduce the "dual" variables $f$. Let us
188: consider the basis where $g (\Omega-A) g$ is diagonal and label
189: the states in this basis by $p$. Define the integration $\int [D
190: f] \equiv \int d f_{p_1} ...\int d f_{p_n} ...$ in such a way that
191: the path of integration in each integral depends on a sign of
192: diagonal element $(g (\Omega-A) g)_{pp}$: $f_p$ goes from
193: $-\infty$ to $\infty$ for $(g (\Omega-A) g)_{pp}<0$ and from $-i
194: \infty$ to $i \infty$ overwise. With thus defined $\int [D f]$,
195: the following identity holds:
196: \begin{widetext}
197: \begin{equation}\label{Eq3}
198:     e^{-\frac{1}{2} \beta \left((\phi-\phi^0) (\Omega-A) (\phi-\phi^0)\right)}
199:     =({\rm factor}) \times \int [Df] e^{-\beta \left(((\phi-\phi^0) g^{-1} f) - \frac{1}{2}
200:     (f (g (\Omega-A) g)^{-1} f)\right)}.
201: \end{equation}
202: In fact, the path of integration over $f_p$ is chosen to deliver
203: the convergence of integrals here. After this identity is applied,
204: partition function takes the form
205: \begin{eqnarray}\label{Zfphi}
206:     &Z=\int [D\phi] \int [D f] e^{-\beta E(\phi, f)}\\    \nonumber
207:     &E(\phi,f)=E^{aux}(\phi)
208:     +((\phi-\phi^0) g^{-1} f)  - \frac{1}{2} (f (g (\Omega-A) g )^{-1} f),
209: \end{eqnarray}
210: 
211: 
212: A set of exact relations between the averages over initial and
213: dual ensemble can be established by the integration by parts in
214: (\ref{Zfphi}) with respect to $f$. Particularly,
215: \begin{eqnarray}\label{ExactRelations}
216: &g (\Omega-A) <\phi-\phi^0>= <f>;\\   \nonumber &g (\Omega-A)
217: <(\phi-\phi^0) \cdot (\phi-\phi^0)>(\Omega-A)g = <f \cdot
218: f>+\beta^{-1} g (\Omega-A) g.
219: \end{eqnarray}
220: 
221: 
222: \end{widetext}
223: 
224: The idea of an introduction of new variables is that energy
225: (\ref{Zfphi}) does not contain direct coupling between different
226: sites for the initial variable $\phi$, since $(\phi g^{-1} f)=
227: \sum_x (\phi_x (g^{-1} f)_x)$. Therefore $\phi_x$ can be
228: integrated out at each site, yielding an an expression for the
229: energy in dual variables $E(f)$:
230: \begin{eqnarray}\label{Zf}
231:     &E(f)=\sum_x V_x(f_x)+\frac{1}{2} \left(f   \tilde{\Omega}  f\right)\\
232:     \nonumber
233:     &\tilde{\Omega}=-g^{-1}-g^{-1}(\Omega-A)^{-1}g^{-1}
234: \end{eqnarray}
235: This expression is formally very similar to an initial system
236: (\ref{Zphi}). Physical difference comes, first, from the
237: possibility to choose $A$ and $\phi^0$ in certain optimal way,
238: and, second, from the the non-trivial way of integration over $f$
239: so that $f$ has an imaginary part. In particular, $<f_{xj}^2>$ can
240: be negative or equal zero.
241: 
242: Properties of the dual potential $V(f)$ are determined by the
243: cumulant expansion (\ref{auxil}) for an auxiliary system. The last
244: term in (\ref{Zf}) is chosen in such a way that the second
245: derivative of $V$ vanishes; other derivatives are
246: $V^{(1)}=(\overline{\phi}-\phi^0) g^{-1}$ and $-\beta
247: V^{(n)}=\gamma^{(n)}$ for $n\ge 3$. To describe effects due to
248: nonlinearity, we introduce tensor $\Sigma'$ defined by equation
249: \begin{equation}\label{SigmaG}
250:     <\phi \cdot \phi>=\beta^{-1} (g^{-1}+\Omega-A+\Sigma')^{-1},
251: \end{equation}
252: so that $\Sigma'$ describes  a non-local contribution to the
253: self-energy of the initial system (\ref{Zphi}). It follows from
254: the second line of (\ref{ExactRelations}) that dual two-point
255: correlator can be expressed via $\Sigma$ as follows:
256: \begin{equation}\label{Sigma}
257:      <f \cdot f>= g\frac{-\beta^{-1}}{(g^{-1}-\Sigma')^{-1}+(\Omega-A)^{-1}}g.
258: \end{equation}
259: 
260: 
261: All expressions presented above are exact. Now, we construct a
262: perturbation theory resulting from the series expansion in powers
263: of $V$. Terms of these series can be expressed by the diagrams
264: with $\gamma^{(n)}$ standing at vertexes connected with lines
265: carrying dual two-point correlator $<f \cdot f>$. We consider the
266: series for $\Sigma'$. It can be found from the expansion of the
267: left- and right-hand side of (\ref{Sigma}) in powers of $V$ and
268: $\Sigma'$, respectively. Some of those diagrams are presented in
269: Figure 1. Each diagram is accompanied by an additional factor
270: $-\beta^{-1}$ and a numerical coefficient. It can be found from
271: the above-mentioned expansion of (\ref{Sigma}).
272: 
273: 
274: \begin{figure}
275: \includegraphics[width=\columnwidth]{1_1.eps}
276: \caption{Diagrams contributing $\Sigma'$ and corresponding
277: coefficients. Diagram (a) is small if condition (11) is fulfilled.
278: Diagrams (b) and (c) describe av $\alpha^2$ correction to
279: $\Sigma'$; diagram (c) vanishes for an unordered state of $O(N)$
280: model. }
281: \end{figure}
282: 
283: Formally, single-lag vertexes can occur in the diagrams due to the
284: presence of $V^{(1)}$ in the dual energy (\ref{Zf}). We require
285: such $\phi^0$ that $<f>=0$ to get rid of them.  Then the first
286: line of (\ref{ExactRelations}) requires
287: \begin{equation}\label{f0}
288:     <\phi>=\phi^0.
289: \end{equation}
290: 
291: Since $\gamma^{(n)}$ of any order can in principle appear in the
292: series, a formal small parameter should be introduced to group the
293: terms properly. Consider a special case of a scalar field with
294: small nonlinearity: $U(\phi)=\alpha \phi^4$, $\alpha \to +0$.
295: Estimate how $\alpha$ appears in high derivatives of $V(f)$:
296: $\gamma^{(2 n)}\propto \alpha^n$ and $\gamma^{(2 n-1)}\propto
297: \alpha^n \phi_0$. We propose to group diagrams with respect to
298: their formal smallness in powers of $\alpha$. This is clearly a
299: good choice for a weak-binding limit. Consider the opposite case
300: of tight-binding $\Omega\to 0$ and suppose that our choice of $A$
301: delivers also $(A-\Omega)\to 0$ (note that mean-field and
302: Gaussian-bath schemes satisfy this condition). It is obvious from
303: (\ref{Zf}) that bare two-point correlator appears to be a small
304: quantity in this case. Since the number of lines in a diagram is
305: roughly proportional to its $\alpha$-order, $\alpha$ again appears
306: to be a good small parameter.  We conclude that {\it series
307: expansion converges well both in tight-binding and weak-binding
308: limits} and therefore expect that the theory is suitable for a
309: crossover situation.
310: 
311: Formally, series expansion can be constructed for any $A$, but an
312: appropriate choice should result to better convergence. It is
313: useful to require
314: \begin{equation}\label{ff0}
315:    \beta^{-1} g=<(\phi-\phi_0)\cdot(\phi-\phi_0)>_{xx}.
316: \end{equation}
317: One can formally obtain Gaussian-bath approximation as a result of
318: Gaussian approximation for $<f\cdot f>$ in (\ref{Sigma}), combined
319: with requirements (\ref{f0},\ref{ff0}). Gaussian approximation
320: means $\alpha^0$ order of the theory. Higher terms of an expansion
321: in powers of  $\alpha$ improve the result and, in particular,
322: bring non-local correlations on the scene.
323: 
324: To demonstrate capabilities of the method and to give more
325: technical details, we present here the results obtained in an
326: $\alpha^2$-approximation for classical $O(N)$ models at 3D cubic
327: lattice with the nearest-neighbor coupling. Potential energy is
328: \begin{equation}\label{Ising}
329:     -\sum_{<xx'>} (\phi_x \phi_{x'}),
330: \end{equation}
331: where $\phi_x$ is an $N$-component vector of the constrained
332: length $||\phi_x||^2=(\phi_x \phi_x)=N$. In the previous notation,
333: dispersion law is $\Omega_k=-(\cos k_x+\cos k_y +\cos k_z)$,
334: whereas the constrain makes the system nonlinear.
335:  Sum is over all pairs of the
336: nearest neighbors. Cases of $N=1,2,3$ correspond to Ising, $xy$,
337: and Heisenberg models, respectively. In three dimensions, system
338: (\ref{Ising}) shows a second-order phase transition at finite
339: temperature for $N$. The model is extensively studied. One can
340: note Monte Carlo simulations \cite{HeisenbergMC, Ising},
341: high-temperature expansion \cite{HT} (in our terminology, this is
342: in fact tight-binding approach), renormalisation-group methods
343: \cite{RG} etc. Here we do not pretend to obtain something new for
344: the model itself. It is used to check if the method works for a
345: realistic system.
346: 
347: 
348: For simplicity, let us consider an unordered state $<\phi>=0$, so
349: that only even-lag vertexes appear in diagrams,
350: $\gamma^{(2n+1)=0}$. It can be easily obtained that
351: $g_{12}=\beta^{-1}\delta_{12}$ (irrespective on $N$) and
352: $\gamma^{(4)}_{1234}=\gamma_{N}
353: (\delta_{12}\delta_{34}+\delta_{13}\delta_{24}+\delta_{14}\delta_{23})$
354: with $\gamma_{N=1}=-2/3, \gamma_{N=2}=-1/2, \gamma_{N=3}=-2/5$.
355: Here $\delta$ is a Kronecker delta. Tensor structure of $g$,
356: $\gamma^{(4)}$, as well as $A$, corresponds to the $O(N)$ symmetry
357: of the problem.
358: 
359: 
360: The self-consistent equation for $a$ follows from (\ref{SigmaG})
361: and (\ref{ff0}):
362: \begin{equation}\label{loop}
363:     g=\frac{1}{(2\pi)^3}\int \frac{d^3 k}{g^{-1}+\Omega_k+\Sigma'-A}
364: \end{equation}
365: 
366: Technically, this equation can be solved iteratively, similarly to
367: DMFT-loop technique \cite{DMFT}. Start from certain guess for $A$
368: and $\Sigma'$. For a general case, one should calculate $g(A)$ and
369: $\gamma^{(n)}(A)$ at this point, but for a particular model
370: (\ref{Ising}) it is not necessary as $g$ and $\gamma$ do not
371: depend on $A$, because of the constrain $||\phi_x||^2=N$.
372: Calculate $<f \cdot f>$ from (\ref{Sigma}), given $g, A$, and
373: $\Sigma'$. For thus obtained $<f \cdot f>$ and $\gamma^{(n)}$,
374: calculate new guess for $\Sigma'$ for a given order in $\alpha$
375: (see below expressions for $\Sigma'$ up to $\alpha^2$). Finally
376: calculate the right-hand side of (\ref{loop}), given $g, A$, and
377: new $\Sigma'$. Denote the obtained value by $\tilde{g}$ and take
378: the quantity $A+\xi (g^{-1}-\tilde{g}^{-1})$ as new guess for $A$,
379: where $\xi\lesssim 1$ is a numerical factor. Repeat the loop for
380: new $A$ and $\Sigma'$. For properly chosen initial guess and value
381: of $\xi$ iterations converge to a fixed point satisfying equation
382: (\ref{Ising}).
383: 
384: 
385: Gaussian-bath approximation corresponds to an assumption
386: $\Sigma'=0$. Corrections are given by the diagrams presented in
387: Figure 1. Lines in these diagrams are thick, {i.e.} they
388: correspond to the renormalized correlator $<f \cdot f>$.
389: First-order correction would be given by a simple loop shown as
390: diagram (a) in Figure 1. This diagram corresponds to formula
391: $-\frac{1}{2\beta} \gamma^{(4)} <f_x \cdot f_x>$. It is however
392: absent, because $<f_x \cdot f_x>$ equals zero in Gaussian-bath
393: approximation, as it can be obtained from (\ref{ExactRelations})
394: and (\ref{loop}). So the correction starts actually from
395: $\alpha^2$ terms.
396: 
397: Strictly speaking, once high-order corrections are taken into
398: account, $<f_x \cdot f_x>$ does not vanish anymore. But it remains
399: small. It can be shown that simple loops does not appear in
400: diagrams up to $\alpha^3$ order. Consequently, second-order
401: correction to $\Sigma'$ can be described by just a single diagram
402: (b) drawn in Figure 1, so that $\Sigma'_{xx'}= -\frac{1}{6 \beta}
403: \gamma^2 <f_{x} \cdot f_{x'}>^3$. After the sum over internal
404: indices is taken, it gives $\Sigma'_{xx'}= -\beta^{-1} \gamma^2
405: (1+N/2) f_{x, x'}^3$, where $f_{x,x'}$ is a diagonal element of
406: the dual correlator.
407: 
408: 
409: \begin{figure*}
410: \includegraphics[width=16cm]{1_2.eps}
411: \caption{ Effective soft-mode dispersion for $O(N)$ model at
412: $N=1,2,3$. Dashed line: Gaussian-bath approximation. Solid lines:
413: results for the $\alpha^2$-scheme, from bottom to top: $N=1$
414: (Ising), $N=2$ (xy), and $N=3$ (Heisenberg). In a narrow region
415: near critical point the iteration procedure did not converge, and
416: the lines are not drawn. Symbols are numerical data for $N=1,2,3$:
417: respectively solid circles, open circles and crosses. Dotted line
418: is a fit corresponding to a `purely critical' behavior of $<\phi
419: \cdot \phi>$ in Ising model.}
420: \end{figure*}
421: 
422: 
423: Figure 2 shows data for the effective soft-mode dispersion
424: $\Sigma_0\equiv(\beta <\phi_{k=0}\cdot\phi_{k=0}>)^{-1}$. Dots are
425: numerical data. Data for the lines are obtained by formula
426: $\Sigma_0=g^{-1}+\Omega_{k=0}+\Sigma'-A$. Iteration procedure
427: converges well everywhere except very narrow regions near critical
428: point. Since we were not interested in a description of the
429: critical point itself, we did not put special attention to the
430: source of this divergence.
431: 
432:  Gaussian-bath approximation predicts the
433: same dependence for any $N$, as one can observe from (\ref{loop})
434: with $\Sigma'=0$. This is a serious qualitative drawback, although
435: the quantitative accuracy is not so bad. So, the dependence of
436: $\Sigma_0$ on $N$ in the model is related with a non-local part of
437: correlations, and should be described by higher terms of
438: $\alpha$-corrections. Indeed, $\alpha^2$-correction describes the
439: dependence $\Sigma_0(N)$ well. There is also a dramatic increase
440: of the accuracy in the $\alpha^2$ curves.
441: 
442: For the broken-symmetry state, the theory can be constructed in a
443: very similar manner. Iteration loop now includes also a
444: modification of $\phi^0$ to fulfill the condition (\ref{f0}) at a
445: fixed point. Diagram (c) from Figure 1 appears due to the presence
446: of $\gamma^{(3)}$. We do not show numerical results here, because
447: they are very similar to what we obtain for the unordered state.
448: 
449: 
450: The entire plot range of Figure 2 lies within a critical region.
451: For example, all points for Ising model obey critical scaling law
452: $<\phi_{k=0}^2>\propto (\beta-\beta_c)^{1.24}$ with a good
453: accuracy, as the dashed line shows. Thus the theory performs well
454: deep inside the critical region. However, the very vicinity of the
455: critical point is not correctly described, because we work with
456: perturbation theory of a finite order. It would be worth to
457: construct a renormalization-group theory based on the presented
458: perturbation approach. An attempt to construct a primitive
459: approach of this kind starting from mean-field theory was
460: performed several years ago \cite{dualRG}.
461: 
462: 
463: 
464: There is a comment about an applicability of other approximation
465: schemes. The system is strongly nonlinear, so weak-binding
466: approximation is clearly not valid here. On the other hand,
467: tight-binding description of a low order fails near the phase
468: transition point because of the critical increase of the
469: correlation length. For example, second order of the tight-binding
470: series gives $\Sigma_0=\beta^{-1}(1+4 \beta)^{-1}$. This is very
471: inaccurate; corresponding curve simply lies out of the plot range
472: of Figure 2. More sophisticated schemes, for example, cumulant
473: expansion, also require much higher order to achieve an accuracy
474: comparable with the presented $\alpha^2$ results. It is also
475: important to recall that Gaussian-bath approximation becomes exact
476: in the limit $N\to \infty$, and $1/N$ expansion can be constructed
477: at this point \cite{1/N}. But $1/N$ approach is essentially based
478: on a very particular $O(N)$ symmetry of system (\ref{Ising}).
479: Contrary, our method is developed for a general case and does not
480: require any special symmetry.
481: 
482: 
483: 
484: 
485: In conclusion, we presented an exact renormalization of the
486: diagram-series expansion in terms of the self-consistent impurity
487: model. Renormalized vertexes are irreducible vertex parts of the
488: impurity model. There is an explicit small parameter in the theory
489: for weak-binding and tight-binding limit. A worked example of an
490: $\alpha^2$-approximation for simple 3D models demonstrates that
491: the method performs well for a strong-correlated situation. The
492: example models were chosen because of the physical simplicity; the
493: method itself looks rather general. It would be particularly
494: important to extend it to the systems of identical quantum
495: particles. Another challenge is a construction of
496: renormalization-group theory based on the method.
497: 
498: The work was supported by Dynasty foundation and NWO grant
499: 047.016.005. Author is grateful to M.I. Katsnelson and O.I.Loiko
500: for their interest to this work.
501: 
502: 
503: 
504: \begin{thebibliography}{99}
505: 
506: \bibitem{Fulde} P.Fulde, {\it Electron Correlations in Molecules and Solids}, Springer, Berlin, 2002
507: 
508: \bibitem{Balesku} R. Balesku, {\it Equilibrium and Nonequilibrium Statistical Mechanics}, Wiley, New York, 1975.
509: 
510: \bibitem {Cowley} A.D. Bruce and R.A. Cowley {\it Structural phase transitions}, Taylor and Francis Ltd., London, 1981.
511: 
512: \bibitem{DMFT} A. Georges, G. Kotliar, W. Krauth, and M.J. Rozenberg, Rev. Mod. Phys. {\bf 68}, 13 (1996).
513: 
514: \bibitem {HeisenbergMC} P. Peczak, A.L. Ferrenberg, and D.P. Landau, Phys. Rev. B {\bf 43}, 6087
515: (1991).
516: 
517: \bibitem {Ising} A. M. Ferrenberg and D. P. Landau, Phys. Rev. B {\bf 44}, 5081 (1991).
518: 
519: \bibitem{HT} M. Campostrini, A. Pelissetto, P. Rossi, and E. Vicari, Phys. Rev. E {\bf 65}, 066127
520: (2002).
521: 
522: \bibitem{RG} A. Pelissetto and E. Vicari, Physics Reports {\bf 368}, 549
523: (2002).
524: 
525: \bibitem{cumulant} H.Li, T. Chen, Z. Phys. B {\bf 100}, 283
526: (1996).
527: 
528: \bibitem{dualRG} A.N. Rubtsov, Phys. Rev. B {\bf 66},052107-1 (2002).
529: 
530: \bibitem{1/N} Y. Okabe, M. Oku, R. Abe, Progress of Theoretical Physics {\bf 59}, 1825
531: (1978).
532: 
533: \end{thebibliography}
534: 
535: 
536: 
537: \end{document}
538: