hep-th0208181/ll.tex
1: %\documentstyle[preprint,revtex]{aps}
2: %\documentstyle[multicol,aps,psfig]{revtex}
3: %\documentclass[twocolumn,showpacs,preprintnumbers,superscriptaddress,
4: %nofootinbib]{revtex4}
5: %\usepackage{graphicx}
6: \documentstyle[multicol,aps,psfig]{revtex}
7: 
8: %\setlength{\topmargin}{0.15in}
9: \begin{document}     
10: 
11: %\draft
12: 
13: \preprint{aaaa}
14: 
15: \title{Complex singularities of the critical potential in the
16: large-$N$ limit}
17: \author{Y. Meurice{\footnote {e-mail:yannick-meurice@uiowa.edu}}  \\
18: {\it Department of Physics and Astronomy, The University of Iowa, 
19: Iowa City, Iowa 52242, USA{\footnote{permanent address}}}\\
20: and\\
21: {\it Fermilab, PO Box 500, Batavia,
22: Illinois 60510-0500, USA}}
23: 
24: \maketitle
25: \begin{abstract}
26: We show with two numerical examples that the conventional expansion 
27: in powers of the field for  
28: the critical potential
29: of 3-dimensional $O(N)$ models in the 
30: large-$N$ limit, does not converge for values of $\phi ^2$
31: larger than some  critical value. This can be explained by the existence of
32: conjugated branch points in the complex $\phi ^2$ plane.
33: Pad\'e approximants $[L+3/L]$ for the critical potential 
34: apparently converge at large
35: $\phi^2$. This allows high-precision
36: calculation of the fixed point in a more 
37: suitable set of coordinates. 
38: We argue that the singularities are
39: generic and not an artifact of the large-$N$ limit. We show that
40: ignoring these singularities 
41: may lead to inaccurate approximations.
42: 
43: \end{abstract}
44: 
45: %\pacs{PACS: 11.10.-z, 11.15.Bt, 12.38.Cy, 31.15.Md}
46: \begin{multicols}{2}\global\columnwidth20.5pc
47: %\multicolsep=8pt plus 4pt minus 3pt
48:  
49: \section{Introduction}
50: Since the early days of the 
51: renormalization group (RG) method\cite{wilson}, 
52: 3-dimensional scalar models 
53: have been identified 
54: as an important laboratory to discuss the
55: existence of non-trivial fixed points and 
56: the large cut-off (or small lattice spacing)
57: limit of field theory models.
58: In the case of $N$-components models with 
59: an $O(N)$ invariant Lagrangian, the RG transformation
60: becomes particularly simple in the large-$N$ limit \cite{ma73}.
61: The construction of the effective potential for these models 
62: is discussed in Refs. \cite{col,town,app}.
63: Later, motivated by perturbative results indicating
64: the existence of an UV stable tricritical 
65: fixed point for $N$ large enough\cite{pisarski}, a new mechanism 
66: allowing spontaneous breakdown of scale invariance and dynamical 
67: mass generation was found in the large-$N$ limit\cite{bmb}.
68: In the following, we call this mechanism the ``BMB mechanism''.
69: It was argued\cite{amit84} that the BMB mechanism is compatible with
70: a zero vacuum energy and a better understanding 
71: of this question might suggest
72: a solution to the cosmological constant problem.
73: Spontaneous breaking of scale invariance 
74: is also discussed \cite{bardeen} with related methods in four-dimensional
75: models of clear interest in the context of particle physics.
76: However, doubts were cast\cite{david84} about the fact that the 
77: BMB mechanism is generic
78: and it is commonly believed that it disappears at finite $N$.
79: 
80: In this article we report results which force us to reconsider 
81: the way we think about non-trivial fixed points.
82: We usually think of the RG flows as taking place in a
83: space of bare couplings or more generally in a space of functions.
84: The necessity for this 
85: more general point of view appears quite clearly in exact renormalization
86: group equations \cite{exact}. Unfortunately, it seems impossible 
87: to decide {\it a-priori} 
88: which space of functions should be considered 
89: to study the RG flows.
90: It is clear from perturbation theory that near the Gaussian
91: fixed point, low dimensional polynomial approximations 
92: of the local potential should be adequate.
93: However, it is not clear that this kind of 
94: approximation should be valid
95: far away from the Gaussian fixed point and in particular 
96: near non-trivial fixed points. 
97: 
98: In the following, we concentrate on the non-trivial fixed point 
99: found numerically in the case $N=1$ by K. Wilson \cite{wilson}. 
100: This fixed point is located on a hypersurface 
101: of second order phase transition which separates the symmetric phase from 
102: the broken symmetry phase. 
103: In the following we call this fixed point the Heisenberg fixed 
104: point (HFP for short) as in Ref. \cite{david84}. 
105: It should not be confused with the fixed point relevant for the 
106: BMB mechanism and which is not studied in detail here.
107: The main result of the article is 
108: that the bare potential corresponding to the HFP
109: has singularities in the 
110: complex plane and that
111: ignoring these singularities may 
112: lead to inaccurate approximations. 
113: These claims are based on explicit calculations 
114: performed in the large-$N$ limit for two $O(N)$ invariant models
115: reviewed in section
116: \ref{sec:models}. These two models are: 
117: 1) a model with a $k^2$ kinetic term
118: together with a sharp cut-off, the sharp cut-off model (SCM) for
119: short; 2) Dyson's hierarchical model (HM)\cite{dyson,baker}. 
120: %In both cases,
121: %it is possible to write 
122: %the inverse of the first derivative of the bare
123: %potential corresponding to the non-trivial fixed point in close form. 
124: 
125: Before entering into
126: technical details, three points should be clear. First, all the results 
127: presented here are based on the analysis of long numerical series
128: and no attempt is made to give rigorous proofs. Second, 
129: in order to understand some of the statements made below, 
130: the reader should be
131: aware that even though, at leading order in the large-$N$ approximation, 
132: the critical exponents take $N$-independent 
133: values, the same
134: approximation provides finite $N$ approximate HFP which are
135: $N$-dependent.
136: A more precise formulation of this statement can be found in
137: Sections \ref{sec:models} and  \ref{sec:fp5}.
138: Third, we only work in 3 dimensions. The precise meaning of this
139: statement for the hierarchical model is explained at the end of
140: section \ref{sec:models}.
141: 
142: 
143: In section \ref{sec:hfp}, we review the basic equations 
144: \cite{ma73,david84} defining the 
145: HFP for the SCM. We then show that the definition can be extended 
146: naturally for the HM. The correctness of this definition is
147: verified later in the paper.
148: In section \ref{sec:series}, we present the methods used to calculate
149: the critical potential expanded as a Taylor series in $\phi^2$.
150: The coefficients of this expansion are called the critical couplings.
151: The main conclusion that we can infer 
152: from our numerical results 
153: is that the Taylor series
154: is inadequate for large values of $\phi^2$. First of all, one
155: half of the critical couplings are negative. If we truncate the Taylor
156: series at an order such that the coefficient of the highest order is 
157: negative, we obtain an ill-defined functional integral.
158: In addition, the absolute value of the critical couplings grows
159: exponentially with the order and the expansion has a finite radius of
160: convergence. Consequently, the idea that the
161: critical potential associated with the HFP 
162: can be approximated by polynomials should be
163: reconsidered.
164: 
165: It is nevertheless possible to define the critical theory by using 
166: Pad\'e approximants for the critical potential. In section
167: \ref{sec:pade}, we show that sequences of
168: approximants converge toward the expected function 
169: in a way very reminiscent of the
170: case of the anharmonic oscillator were the convergence 
171: can be proven rigorously \cite{loeffel69}.
172: In addition, the zeroes and the poles of these
173: approximants are located far away from the real positive axis and 
174: follow patterns that strongly suggest the existence of two complex
175: conjugated branch points. 
176: 
177: The complex singularities of the critical potential should not be
178: interpreted as a failure of the RG approach but rather as an 
179: artifact of the system of coordinates used. In section \ref{sec:fp5},
180: we present consistent arguments showing 
181: that in a different system of coordinates \cite{kw,guide}, the 
182: function associated with the HFP is free of singularities.
183: In this system of coordinates, finite dimensional truncation 
184: is a meaningful procedure which, in the case of the HM,  
185: allows comparison 
186: with independent numerical calculations at finite $N$.  
187: An example of such a calculation is presented in the case $N=5$.
188: 
189: In section \ref{sec:disc}, 
190: we discuss the errors associated with two 
191: approximate procedures that can be used to deal with the singularities.
192: The first procedure which is justified in the context of perturbation 
193: theory and does not require an understanding of the 
194: singularities, consists in 
195: truncating the potential at order $(\phi^2)^3$. The second 
196: procedure consists 
197: in restricting the range of integration of $\phi^2$ to the radius 
198: of convergence of the critical potential. 
199: If the range of integration is large enough, this second procedure
200: generates small errors \cite{convpert}.
201: As far as the calculation of the HFP in the system of coordinates 
202: of Section \ref{sec:fp5} is concerned, 
203: both procedures have a low accuracy for both models 
204: considered.
205: In the conclusions, we explain why we believe that the singularities
206: persist at finite $N$ and we discuss implications 
207: of the existence of these singularities for other problems.
208: 
209: \section{Models}
210: \label{sec:models}
211: 
212: We consider lattice models defined by the partition function
213: \begin{equation}
214: Z(\vec{J})=\prod _x\int d^N\phi_x {\rm e}^{-S
215: +\sum_x\vec{J}_x\vec{\phi}_x}\ ,
216: \end{equation}
217: with
218: \begin{equation}
219: S=-{1\over 2}\sum_{xy}
220: \vec{\phi}_x\Delta_{xy}\vec{\phi}_y+\sum_xV_o(\phi^2_x)\ .
221: \end{equation}
222: We use the notation $\phi^2_x \equiv \vec{\phi}_x.\vec{\phi}_x$ and 
223: $\Delta_{xy}$ is a symmetric
224: matrix with negative eigenvalues, such as discrete versions of the 
225: Laplacian. For the simplicity of the
226: presentation, we will assume that $\sum_x\Delta_{xy}=0$. If it is not the
227: case, one can always subtract the zero mode from $\Delta$ and compensate it
228: with a new term in $V$.
229: 
230: Defining the rescaled potential
231: \begin{equation}
232: V_0(X)=NU_0({X\over N})\ ,
233: \label{eq:u0}
234: \end{equation}
235: and performing a Legendre transform from the source $\vec{J}$ to the 
236: external classical field $\vec{\phi}_c$, one can show that 
237: in the large $N$ limit\cite{david84} that
238: $M^2\equiv 2\partial V_{eff}/\partial \phi^2_c$ obeys the self-consistent
239: equation 
240: \begin{equation}
241: 2U_0'(\phi_c^2+f_{\Delta}(M^2))=M^2\ ,
242: \end{equation}
243: where $f_{\Delta}(M^2)$ is the one-loop integral corresponding to the
244: quadratic form $\Delta$ and a mass term $M^2$. 
245: The prime denotes the derivative with respect to the $O(N)$ invariant 
246: argument.
247: The explicit form of
248: $f$ for the two models discussed in the following are given in Eqs. 
249: (\ref{eq:fscm}-\ref{eq:fhm}). 
250: Precise definitions of $\phi ^2_c$ and the effective potential 
251: $V_{eff}$ are given in \cite{david84}.
252:  
253: Up to now, all the quantities introduced are dimensionless.
254: They can be interpreted as dimensionful quantities expressed in 
255: cut-off units.
256: Let us consider two models, the first one with a rescaled potential
257: $U_0$, a UV cutoff $\Lambda$  and a quadratic form $\Delta$ 
258: and a second  model with a rescaled potential
259: $U_{0,S}$, a UV cutoff $\Lambda/S$ and a quadratic form $\Delta_S$. 
260: For $D=3$ and in the large-$N$ limit, the two models have the same 
261: dimensionful zero-momentum Green's functions provided that:
262: \begin{eqnarray}
263: \label{eq:rg}
264: &U&'_{0,S}(\phi^2)= \\ 
265: &S&^2U_0' \biggl(\bigl(\phi^2-f_{\Delta_S}(2U'_{0,S}(\phi^2)\bigr)/S+
266: f_{\Delta}\bigl((2/S^2) U'_{0,S}(\phi^2)\bigr)\biggr)\nonumber
267: \end{eqnarray}
268: In two special cases, the dimensionless expression for the one-loop
269: diagram is independent of the cut-off. In other words,
270: $f_{\Delta}=f_{\Delta_S}\equiv f$ and the fixed point equation becomes
271: very simple \cite{ma73,david84}.
272: 
273: We now discuss the two models where this simplification occurs.
274: In the SCM, $\Delta$ becomes $k^2$ in the momentum
275: representation (Fourier modes). The momentum cutoff is sharp: $k^2\leq 1$ (in 
276: cut-off units). This is why we call this model 
277: the sharp cutoff model.
278: The non-renormalization of the kinetic term is justified in the 
279: large-$N$ limit \cite{david84}. For this model,
280: \begin{equation}
281: f_{SCM}(z)=\int _{|k|\leq 1}{{d^3 k}\over {(2\pi)^3}}{1\over {k^2+z}}\ .
282: \label{eq:fscm}
283: \end{equation}
284: 
285: By construction \cite{baker}, 
286: the kinetic term of Dyson's hierarchical model (HM)
287: is not renormalized and we have  
288: \begin{equation}
289: f_{HM}(z)=\sum_{n=0}^{\infty}{{2^{-n-1}}\over{b(c/2)^n+z}}\ ,
290: \label{eq:fhm}
291: \end{equation}
292: with $c=2^{1-2/D}$ and $b$=${\beta c\over{2-c}}$. 
293: The inverse temperature $\beta$ and the parameter 
294: $c$ appear in the hamiltonian of the HM in a way that 
295: is explained in section II of Ref. \cite{gam3}. The parameter $c$ 
296: is related to the dimension $D$ by considering the scaling of a 
297: massless Gaussian field. In the following we will consider the 
298: case $c=2^{1/3}$ ($D=3$) exclusively. 
299: In addition, $\beta$ will be set to 1 as in
300: other fixed point calculations \cite{gam3}.
301: Different values of $\beta$ can be introduced by a trivial rescaling.
302: Note also that the cutoff cannot be changed continuously for the
303: HM, because the invariance of $f$ is only valid when we integrate 
304: the degrees of freedom of a the largest momentum shell (corresponding 
305: to the hierarchically nested blocks in configuration space)
306: all at the same time.
307: For the HM, the density of sites is reduced by a factor 2 at each RG 
308: transformation. The linear dimension (``lattice spacing'') 
309: is thus increased by by a factor $2^{1/D}$ and the cutoff decreased by
310: the same factor.
311: Consequently, for the HM, 
312: Eq. (\ref{eq:rg}) should be understood only with $S=2^{q/D}=2^{q/3}$ 
313: and $q$ integer. 
314: 
315: \section{The HFP}
316: \label{sec:hfp}
317: 
318: In this section, we review the construction of the HFP for the SCM,
319: and we show that the construction can be extended in a natural (but
320: non-trivial) way for the HM. The first step consists in finding all
321: the fixed points of the RG Eq. (\ref{eq:rg}).
322: Following Refs. \cite{ma73,david84}, we introduce the inverse function: 
323: \begin{equation}
324: F(2U'_0(\phi^2))=\phi^2\ ,
325: \end{equation}
326: and the function $H(z)\equiv F(z)-f(z)$ ,
327: %\begin{equation}
328: %\end{equation}
329: where the one-loop function $f$ has been defined in the previous section for
330: the two models considered. With these notations, 
331: the fixed point equation corresponding to Eq. (\ref{eq:rg}) is simply 
332: \begin{equation}
333: H(z)=SH(z/S^2)\ .
334: \label{eq:fpeqh}
335: \end{equation}
336: 
337: For the SCM, $S$ is allowed to vary continuously in Eq. (\ref{eq:fpeqh})
338: and the general solution is 
339: \begin{equation}
340: F(z)=f_{SCM}(z)+Kz^{1/2}\ .
341: \end{equation}
342: For the HM, $S$ can only be an integer power of $2^{1/3}$ and the
343: general solution has an infinite number of free parameters:
344: \begin{equation}
345: F(z)=f_{HM}(z)+\sum_qK_q z^{1/2+iq\omega}\ ,
346: \label{eq:hmfp}
347: \end{equation}
348: with
349: \begin{equation}
350: \omega\equiv {3\pi\over{{\rm ln}2}}\simeq 13.6\ ,
351: \label{eq:omega}
352: \end{equation}
353: and $q$ runs over positive and negative integers.
354: The only restriction on the constants $K$ and $K_q$ is that $F$ should
355: have a well defined inverse which is real when $F (=\phi^2)$ is real and
356: positive.
357: 
358: It is clear from Eqs. (\ref{eq:fscm}) and (\ref{eq:fhm}) that for both
359: models $f(z)$ has singularities along the negative real axis and that, in
360: general, $F(z)$ cannot be defined for $z$ real and negative. 
361: This imposes restrictions on the choice of the constants $K$ and $K_q$.
362: For instance, in the case of 
363: the SCM, when $K$ takes a large positive value, it 
364: is impossible to reach small values of $F=\phi^2$ when $z\geq 0$ and
365: the fixed point has no obvious physical interpretation. However, there
366: is a special positive value of $K$ for which the singularity of
367: $f_{SCM}$ is exactly canceled and an analytic continuation for $z<0$
368: is possible. Its exact value can be calculated \cite{david84} 
369: by decomposing $f_{SCM}$ into 
370: a regular part $f_{SCM,reg.}$ and a singular part $f_{SCM,sing.}$. 
371: Using elementary trigonometric identities, one finds 
372: \begin{equation}
373: f_{SCM,reg.}(z)={1\over{2\pi^2}}(1+z^{1/2}{\rm Arctan}(z^{1/2}))\ ,
374: \end{equation}
375: and 
376: \begin{equation}
377: f_{SCM,sing.}(z)=-{1\over{4\pi} }z^{1/2}\ .
378: \end{equation}
379: Consequently, if we choose $K={1\over{4\pi}}$, $F$ reduces to $f_{SCM,reg.}$.
380: 
381: A more detailed analysis \cite{david84}, shows that this value of $K$
382: is the only positive value of $K$ for which $U'$ can be defined for 
383: any real positive $\phi^2$. On the other hand, for negative $K$, one 
384: obtains a line of fixed points ending (for $K=0$) at the fixed point
385: relevant for the BMB mechanism.
386: Given the isolation of the fixed point with
387: $K={1\over{4\pi}}$, it is easy to identify it with the HFP. We denote the 
388: corresponding inverse function by $F^{\star}_{SCM}(z)=f_{SCM,reg.}(z)$.
389: As promised this function is analytical in a neighborhood of the
390: origin and has a Taylor  expansion:
391: \begin{equation}
392: F^{\star}_{SCM}(z)={1\over{2\pi^2}}(1+z-{z^2\over 3}+{z^3\over
393:   5}+\dots)
394: \end{equation}
395: This expansion has a radius of convergence equal to 1 due to
396: a logarithmic singularity at $z=-1$. However, as we will see in
397: section \ref{sec:series}, this expansion allows us to construct an 
398: inverse power
399: series and $U_0$.
400: 
401: 
402: In the case of the HM, the decomposition into a regular and singular 
403: part is more tedious. Fortunately, this problem is a particular case 
404: of a problem solved in section 5 of Ref. \cite{osc} where
405: Eq. (5.6) with $A=c^2$, $B=c^{-1}$ and $f(z)=G(z/b)$ yields
406: \begin{equation}
407: f_{HM,sing.}(z)=-{\omega \over{4b}}\sum_q{\bigl( {z/b}\bigr)
408:   ^{1/2+iq\omega}
409: \over
410:   {\rm sin}(\pi(1/2+iq\omega))}\ ,
411: \label{eq:fhms}
412: \end{equation}
413: with $b$ and $c$ defined in section \ref{sec:models}.
414: 
415: If we compare this expression with the general solution of the fixed
416: point equation (\ref{eq:hmfp}),
417: we see that in both expressions, the power  
418: $z^{1/2+iq\omega}$ appears for all positive and negative 
419: integer values of $q$. There exists a unique choice of the $K_q$ in Eq. 
420: (\ref{eq:hmfp}) which cancels exactly the singular part of $f_{HM}$.
421: We call the corresponding fixed point the HFP of the HM.
422: The numerical closeness with the finite $N$ HFP discussed in section
423: \ref{sec:fp5} 
424: confirms the validity of this analogic definition.
425: We call the corresponding inverse function $F_{HM}^{\star}$. Using
426: Eq. (5.5) of Ref. \cite{osc}, we find 
427: \begin{equation}
428: F_{HM}^{\star}(z)=f_{HM,reg.}={1\over{2b}}\sum_{l=0}^{\infty}\biggl({-z\over
429:   b}\biggr)^l{1\over{1-c^{2l-1}}}\ .
430: \label{eq:fexp}
431: \end{equation}
432: This expansion has a radius of convergence $bc^2=2.7024\dots$ for the
433: choice
434: of parameters used here.
435: 
436: It is possible to check the accuracy of the expansion given in Eq. 
437: (\ref{eq:fexp}) by using the identity 
438: $F_{HM}^{\star}(z)=f_{HM}(z)-f_{HM,sing.}$.
439: Note that the two terms of the r.h.s. cannot be defined separately on
440: the negative real axis. On the real positive axis, $f_{HM,sing.}$ is
441: dominated by the $q=0$ term. Numerically, 
442: \begin{equation}
443: K_0={3\pi\over{4b^{3/2}{\rm ln}2}}=1.530339\dots .
444: \label{eq:ko}
445: \end{equation}
446: The terms with $q=\pm 1$ produce log-periodic oscillations of 
447: amplitude $1.7 \times 10^{-18}$. The terms with larger $|q|$ have a
448: much smaller amplitude. These findings are consistent with the log-periodic
449: oscillations found numerically in HT expansions \cite{osc1,osc}.
450: The oscillatory terms are very small along the
451: positive real axis. However, in the complex plane, if we write $z=r{\rm
452:   e}^{i\theta}$, the amplitude is multiplied by ${\rm
453:   e}^{-q\omega\theta}$
454: which compensates the suppression of the denominators in Eq. (\ref{eq:fhms}), 
455: if $\theta \rightarrow +\pi (-\pi)$
456: when $q<0$ ($q>0$). In conclusion, along the real positive axis, we can 
457: use the approximation
458: \begin{equation}
459: F_{HM}^{\star}(z)\simeq f_{HM}(z)+K_0 z^{1/2}\ ,
460: \label{eq:fstarapp}
461: \end{equation}
462: %with $K_0$ given in Eq. (\ref{eq:ko}).
463: with an accuracy of 18 significant digits, but this approximation is 
464: certainly not valid near the negative real axis.
465: 
466: \section{Calculation of the critical potential $U_0^{\star}$}
467: \label{sec:series}
468: 
469: In the previous section, we have provided power series for 
470: the inverse
471: function $F(z)$ corresponding to the HFP of the SCM and the HM. 
472: We can use these series to define $F(z)$ on the negative real axis. 
473: In both cases, as we move toward more negative values of $z$, $F$ becomes 
474: zero within the radius of convergence of the expansion. The situation
475: is illustrated in Fig. \ref{fig:foz} for the HM. 
476: \begin{figure}
477: \centerline{\psfig{figure=foz.EPS,width=3.3in}}
478: \vskip10pt
479: \caption{$F_{HM}^{\star}(z)$ versus $z$ }
480: \label{fig:foz}
481: \end{figure} 
482: Numerically, we have $F_{HM}^{\star}(-1.5107\dots)=0$ and 
483: $F_{SCM}^{\star}(-0.6948\dots)=0$. We then reexpand the series
484: about that value of $z$ (which corresponds to $F=\phi^2=0$) and invert
485: it. The resulting series is an expansion of $2U_0^{\star}$' in $\phi^2$.
486: After integration, and up to an arbitrary constant $u_0$, we obtain a 
487: Taylor series for the critical potential $U_0^{\star}$ corresponding 
488: to the HFP. We denote the expansion as 
489: \begin{equation}
490: U_0^{\star}(\phi^2)=\sum_{n=0}^{\infty} u_n (\phi^2)^n\ .
491: \label{eq:uudef}
492: \end{equation}
493: The precise determination of the zero of $F$ is obtained by Newton's
494: method with a large order polynomial expansion. This expansion is then
495: reexpanded about the zero and the large order coefficients in the original
496: expansion have an effect on the low order coefficients 
497: of the reexpanded series. 
498: We have checked that the order was sufficiently large 
499: to stabilize the results presented hereafter. 
500: \begin{figure}
501: \centerline{\psfig{figure=lnco.eps,width=3.3in}}
502: \vskip10pt
503: \caption{Natural Logarithm of the absolute value of the coefficients
504: $u_n$ of the critical potential $U_0^{\star}$ defined in Eq. (\ref{eq:uudef}) 
505: for the SCM 
506: (filled squares) and the HM (empty circles).}
507: \label{fig:lnco}
508: \end{figure} 
509: 
510: 
511: The absolute values of the first 50 coefficients of both models 
512: are shown in Fig. \ref{fig:lnco}.
513: In both cases, it appears clearly that the absolute value grows at an
514: exponential rate. Linear fits of the right part of Fig. \ref{fig:lnco}
515: suggest a radius of convergence of order 0.11 for the SCM and 2.5 for
516: the HM. The signs of both series follow the periodic pattern: + - - + +
517: - + + - -
518: for the SCM and + + - - for the HM. This suggests singularities in the
519: complex plane at an angle ${k\pi\over 5}$ with respect to the positive
520: real axis ($k=$ 1, 3, 7, 9) for the SCM and along the imaginary axis
521: for the HM. This analysis is confirmed by an analysis of the poles 
522: of Pad\'e approximants presented in the next section.
523: 
524: \section{Pad\'e approximants of $U_0^{\star}$}
525: \label{sec:pade}
526: 
527: At this point, our series expansion of the critical potential does 
528: not allows us to define the critical theory as a functional integral.
529: As $\phi^2$ exceeds the critical values estimated in the previous
530: section, the power series is unable to reproduce the expected function
531: $U_0^{\star}$. The situation is illustrated in Fig. \ref{fig:uuu} for the HM.
532: \begin{figure}
533: \centerline{\psfig{figure=uuhm.EPS,width=3.3in}}
534: \vskip10pt
535: \caption{$U_0^{\star}(\phi ^2)$ for the HM 
536: with a parametric plot (filled squares), the
537: series truncated at order 50 (thick solid line) and Pad\'e
538: approximants [4/1] (thin line slightly above the squares) and 
539:  [5/2] (thin line closer to the squares). The constant has been fixed
540:  in such a way that the value at the minimum is zero.}
541: \label{fig:uuu}
542: \end{figure}
543: 
544: The numerical values of $U_0^{\star}$ in Fig. \ref{fig:uuu} 
545: have been calculated
546: using a parametric representation (with $z$ as the parameter). 
547: We have calculated pairs of 
548: values
549: \begin{equation}
550: \left(F^{\star}(z),{1\over 2}\bigl(zF^{\star}(z)-\int_0^z
551: dz'F^{\star}(z')\bigr)\right)\ , 
552: \label{eq:para}
553: \end{equation}
554: for various real positive values of $z$.  
555: A simple graphical analysis performed by representing $U_0^{\star}$ as a surface
556: on Fig. \ref{fig:foz} shows that each pair of values in 
557: in Eq. (\ref{eq:para}) corresponds to 
558: a pair $(\phi^2, U_0^{\star}(\phi^2))$ with the arbitrary constant
559: in $U_0^{\star}$ fixed in such a way that $U_0^{\star}$ vanishes at its minimum. 
560: We have calculated 
561: $F^{\star}$ by using the independent but approximate Eq. (\ref{eq:fstarapp}).
562: As explained in section \ref{sec:hfp},
563: the approximate expression is only valid 
564: for $z$ real and positive and should give 18 correct significant digits. 
565: In Fig. \ref{fig:uuu}, we have used the values $z=2U'_0=0,\ 0.25, 0.5,
566: \dots$.
567: This is why the filled 
568: squares only appear when the derivative of $U_0^{\star}$ is positive.
569: Unlike Eq. (\ref{eq:fexp}) which has a radius of convergence 1, the
570: approximate expression  Eq. (\ref{eq:fstarapp}) remains valid 
571: for large positive values of $z$. It is thus
572: possible to check if Pad\'e approximants can be used to represent 
573: the critical potential beyond the radius of convergence of its 
574: Taylor expansion. Fig. \ref{fig:uuu} shows that low order approximants
575: are close to the parametric curve. As the order increases, the curves
576: coalesce with the parametric curve and a more refined description is
577: necessary. 
578: 
579: In Fig. \ref{fig:padeconv}, we give the accuracy reached 
580: by various approximants for the HM with a broad range of $\phi^2$ 
581: (more than 4 times the radius of convergence). As the order of
582: the approximants increase the accuracy increases but at a rate which 
583: is slower for larger values of $\phi^2$. The figure is very similar 
584: to sequences of Pad\'e approximants obtained for the 
585: ground state of the anharmonic oscillator
586: (see Fig. 1 of Ref. \cite{convpert}),
587: where the convergence can be proven rigorously \cite{loeffel69}.
588: Note that the slow convergence at large $\phi^2$ 
589: is not a serious problem, since the 
590: contributions for 
591: large $\phi^2$ are exponentially suppressed in the functional integral.
592: The choice of $[L+3/L]$ approximants is discussed in more detail
593: below.
594: Up to now, we only discussed the HM. Following the same procedure for 
595: the SCM, we obtain very similar figures (with a different $\phi^2$
596: scale)
597: which we have not displayed.
598: \begin{figure}
599: \centerline{\psfig{figure=padeconvhm.EPS,width=3.3in}}
600: \vskip10pt
601: \caption{Number of correct significant digits obtained with Pad\'e 
602: approximants [9/6], [13/10], [17/14] and [21/18] 
603: for various values of $\phi ^2$ for the HM.}
604: \label{fig:padeconv}
605: \end{figure} 
606: 
607: The singularities of $U_0^{\star}$ in the complex $\phi^2$ plane can be
608: inferred from the location of the zeroes and poles of the Pad\'e 
609: approximants. As $L$ becomes large, regular patterns appear.
610: Examples are shown in Fig. \ref{fig:rootscm} for the SCM and 
611: Fig. \ref{fig:roothm} for the HM.
612: In both cases, the zeroes and 
613: poles approximately alternate along two lines 
614: ending where singularities were expected from the analysis of
615: coefficients
616: in Section \ref{sec:series}. This pattern suggests \cite{baker96} 
617: the existence of two
618: complex conjugated branch points at the end of these lines.
619: 
620: 
621: \begin{figure}
622: \centerline{\psfig{figure=rootsczz.EPS,width=3.3in}}
623: \vskip10pt
624: \caption{Real and imaginary parts of the roots of the denominator 
625: (filled squares) and numerator (crosses) of 
626: a [26/23] Pad\'e approximant for the SCM  
627: The solid circle has a radius 0.11 and the
628: two solid lines make angles $\pm {3\pi\over 5}$ with respect to the
629: positive real axis.}
630: \label{fig:rootscm}
631: \end{figure}
632: \begin{figure}
633: \centerline{\psfig{figure=roothmzz.EPS,width=3.3in}}
634: \vskip10pt
635: \caption{Real and imaginary parts of the roots of the denominator 
636: (filled circles) and numerator (crosses) of 
637: a [26/23] Pad\'e approximant for the HM. 
638: The solid circle has a radius 2.5. Two roots farther away on the
639: imaginary axis and one root farther away near the negative real axis 
640: are not displayed.}
641: \label{fig:roothm}
642: \end{figure} 
643: 
644: The choice of $[L+3/L]$ approximants is easily justified for the 
645: SCM. At large $|z|$, $f_{SCM}(z)\propto 1/z$ and and
646: $F^{\star}_{SCM}(z)
647: \simeq {1\over{4\pi} }z^{1/2}$. For large $|\phi^2|$, 
648: $U_0^{\star}$'$\simeq 8\pi^2 (\phi^2)^2$ and 
649: $U_0^{\star}\simeq{8\pi^2\over{3}} (\phi^2)^3$. 
650: Consequently a $[L+3/L]$ approximant should have the correct asymptotic
651: behavior. More precisely, 
652: if $a_{L+3}$ and $b_L$ are the leading coefficients of the numerator
653: and
654: denominator of a Pad\'e $[L+3/L]$ respectively, we expect that when $L$ is
655: large
656: \begin{equation}
657: {a_{L+3}\over b_L}\rightarrow{8\pi^2\over 3}
658: \label{eq:limas}
659: \end{equation}
660: Defining a quantity 
661: \begin{equation}
662: E_L\equiv 1-{{3a_{L+3}}\over{8\pi^2 b_L}}\ ,
663: \end{equation}
664: that measures the departure from the expected asymptotic behavior, we
665: see from Fig. \ref{fig:sdco} that as $L$ increases, the
666: discrepancy
667: diminishes exponentially.
668: \begin{figure}
669: \centerline{\psfig{figure=sdco.EPS,width=3.3in}}
670: \vskip10pt
671: \caption{Log$|E_L|$ versus $L$.}
672: \label{fig:sdco}
673: \end{figure} 
674: 
675: In the case of the HM,  the situation is more 
676: intricate. From Eq. (\ref{eq:fstarapp}), we may be tempted to
677: conclude that the two cases are similar. Unfortunately, Eq. (\ref{eq:fstarapp})
678: is a real equation not a complex one. In the complex plane, the terms 
679: with $q\neq 0$ become important near the negative real axis and no
680: simple simple limit as in Eq. (\ref{eq:limas}) applies. However,
681: if  we need $U_0^{\star}$ only along the real positive axis, Fig. \ref{fig:padeconv}
682: justifies the use of the $[L+3/L]$ sequence of approximants.
683: 
684: \section{The HFP in a convenient set of coordinates}
685: \label{sec:fp5}
686: 
687: As explained in the introduction
688: we can think that the RG flows move in a space of functions.
689: The system of coordinates for this space can be chosen in a way which
690: is convenient to make approximations. A particularly convenient system
691: of coordinates for the HM consists in considering the 
692: Fourier transform of the local measure of integration \cite{kw,guide}.
693: In this system of coordinates and at leading order in
694: the $1/N$ expansion, the HFP for a given $N$ reads: 
695: \begin{equation}
696: R^{\star}(\vec{k})\propto \int d^N\phi {\rm e}^{-{b\over 2}\phi^2-
697: NU_0^{\star}(\phi^2/N)+i\vec{k}.\vec{\phi}}\ .
698: \label{eq:fpint}
699: \end{equation}
700: The quadratic term proportional to $b$ is due to the fact that 
701: the quadratic form $\Delta$ for the HM has a zero mode.
702: We then Taylor expand
703: \begin{equation}
704: R^{\star}(\vec{k})=1+\sum_{n=1}^{\infty}a_n (k^2)^n\ ,
705: \label{eq:fp}
706: \end{equation}
707: and consider the $a_n$ as the our new set of coordinates.
708: The advantage of this representation is that it is possible 
709: to make very accurate calculations by using polynomial 
710: approximations \cite{kw,guide,gam3} 
711: of the infinite sum in Eq. (\ref{eq:fp}). In this section and the 
712: next section, 
713: we discuss the details of the calculations for the HM.
714: The case of the SCM shares many similarities with the HM 
715: and is discussed briefly at the 
716: end of each section.
717: 
718: We have performed a numerical calculation of the $a_n$ of the HM
719: using Eq. (\ref{eq:fpint}) in the 
720: particular case $N=5$. The study of the ratios of successive
721: coefficients displayed in Fig. \ref{fig:decayrat} indicates that 
722: the $|a_n|$ decay faster than $1/n!$ and that $R^{\star}(\vec{k})$ is 
723: analytical over the entire complex $k^2$ plane in contrast to
724: $U_0^{\star}(\phi^2)$ which has a finite radius of convergence in the complex
725: $\phi^2$ plane.
726: 
727: The good convergence of $R^{\star}(\vec{k})$ can be explained 
728: by an approximate calculation. 
729: The $\phi$ integral that is performed in the calculation
730: of the $a_n$ has a positive integrand with a peak moving to larger 
731: values of $|\phi|$ when $n$ increases. For sufficiently large values
732: of $n$, we can replace $U_0^{\star}$ by its asymptotic behavior on the 
733: positive real $\phi^2$ axis 
734: which can be derived from the approximate Eq. (\ref{eq:fstarapp}) for the HM:
735: \begin{equation}
736: R^{\star}(\vec{k})\sim \int d^N\phi {\rm e}^{-(1/(6N^2K_0^2))
737: (\phi^2)^3+i\vec{k}.\vec{\phi}}\ .
738: \end{equation}
739: With this approximation, the $a_n$ can be expressed  exactly in terms of 
740: gamma functions and a simple calculation yields
741: \begin{equation}
742: -{a_n\over{a_{n-1}}}\simeq {{(6N^2K_0^2)^{1/3}\Gamma((N+2n)/6)}\over{4n(n-1+N/2)
743: \Gamma((N+2(n-1))/6)}}\ .
744: \label{eq:asympratios}
745: \end{equation}
746: Note that there are no free
747: parameters
748: in this formula.
749: Fig. \ref{fig:decayrat} shows that Eq. (\ref{eq:asympratios}) is a
750: very good approximation of the ratios obtained numerically from
751: Eq. (\ref{eq:fpint}). 
752: 
753: We have also calculated the $a_n$ corresponding to the HFP for $N=5$ 
754: using the numerical method developed in the case $N=1$ 
755: in Ref. \cite{gam3} and which
756: can be extended easily for arbitrary $N$. In brief, 
757: it consists of finding the stable manifold by fine tuning the
758: temperature and then iterating the RG transformation 
759: in order to get rid of the irrelevant directions. This procedure is 
760: very accurate and completely 
761: independent of the approximations made in this article. Remarkably, 
762: we found that even though $N=5$ is not a large number, the first
763: coefficients obtained in the leading order in the $1/N$ approximation
764: coincide with about two significant digits with the accurate values 
765: found numerically with $N=5$. 
766: As the order increases, the accuracy
767: degrades slowly. This is explained in more detail below. However, 
768: the ratios of successive coefficients still follows 
769: closely the asymptotic
770: prediction obtained from Eq. (\ref{eq:asympratios}). 
771: This strongly suggests that the $(\phi^2)^3$ asymptotic behavior of the
772: critical potential persists at finite $N$.
773: \begin{figure}
774: \centerline{\psfig{figure=decayrat2.EPS,width=3.3in}}
775: \vskip10pt
776: \caption{Ratios of successive coefficients for the HM, using 
777: the leading order Eq. (\ref{eq:fpint}) (stars), the asymptotic 
778: formula Eq. (\ref{eq:asympratios}) (continuous line) and the 
779: numerical fixed point (empty circles). Same results for the SCM: leading
780: order (filled squares) and asymptotic (dashed line). In all cases, $N=5$.}
781: \label{fig:decayrat}
782: \end{figure} 
783: 
784: Except for the comparison with independent numerical calculations at finite 
785: $N$, the same calculations can be performed for the SCM with minor changes 
786: ($b\rightarrow 0$ and $K_0\rightarrow K$). The results are also shown in 
787: Fig. \ref{fig:decayrat} where one can see that the agreement with the
788: asymptotic formula is very good even at low order.
789: 
790: \section{Discussion of alternate procedures}
791: \label{sec:disc}
792: In section \ref{sec:pade}, we have shown that the Pad\'e approximants
793: provide accurate values of $U_0^{\star}$ far beyond its radius of
794: convergence.
795: In order to estimate the error 
796: on the new coordinates $a_n$ 
797: due to the 
798: use of approximants for 
799: $U_0^{\star}$, we can vary the range
800: of integration and change the approximants. For instance the 
801: values of $a_n$ of the HM 
802: used in Fig. \ref{fig:decayrat} have been calculated
803: using a range of integration $|\phi |<20$ and a [26/23] Pad\'e
804: approximant. For the values of $n$ considered here, 
805: changing the range of integration has effects smaller
806: than the errors due to numerical integration (which has an accuracy of 
807: about 11
808: significant digits in our calculation) 
809: provided that we include values up to $|\phi|\simeq 4.9$.
810: Restricting the range of integration to smaller values produces 
811: sizable effects. As an example, the small effects due a 
812: restriction to $|\phi |<4.4$ are shown in Fig. \ref{fig:various}.
813: Similarly, the values of $a_n$ are not very sensitive to small changes in 
814: the Pad\'e approximants. Sizable effects are obtained by changing 
815: the order of the numerator and denominator by approximately 10.
816: For instance, the effects of using a [14/11] approximant are shown in
817: Fig. \ref{fig:various}. 
818: \begin{figure}
819: \centerline{\psfig{figure=various3.EPS,width=3.3in}}
820: \vskip10pt
821: \caption{Number of significant digits common with our best estimate
822: for the $a_n$ obtained for the HM 
823: from Eq. (\ref{eq:fpint}) with $n$ (the order) 
824: going from 1 to 20.  The alternate procedure are the truncation at 
825: order $(\phi^2)^3$ (filled circles), the $N=5$ accurate 
826: numerical result (filled
827: square), no Pad\'e approximants but a truncation of the range of
828: integration close to the radius of convergence (empty circles), 
829: a restriction of the range of integration for $\phi<4.4$ (stars), and
830: a [14/11] Pad\'e (diamonds).}
831: \label{fig:various}
832: \end{figure} 
833: 
834: 
835: Having demonstrated that we can calculate the first 20 coefficients
836: $a_n$, at leading order 
837: in the $1/N$ expansion, 
838: with at least 10 significant digits, we can now discuss the 
839: errors associated with other procedures mentioned in the introduction. 
840: The first procedure consists in truncating $U_0^{\star}$ keeping only the 
841: terms up to order $(\phi^2)^3$. 
842: This is procedure inspired by 
843: perturbation theory amounts to keep only the relevant and 
844: marginal directions near the Gaussian fixed point. 
845: From Fig. \ref{fig:various}, we 
846: see that this procedure generates errors which are of the same order
847: as the errors due to the use of the leading $1/N$ approximation.
848: Consequently, this procedure is quite unsuitable to study the 
849: correction to this approximation. Slightly better results are 
850: obtained by keeping as many terms as possible in the expansion 
851: (up to 50 in our calculation) but restricting the range of integration
852: in such way that we stay within the radius of convergence. Given 
853: the rescaling of Eq. (\ref{eq:u0}) this means that for $N=5$, we need 
854: to restrict the integration to $|\phi|< \sqrt{5\times 2.5}\simeq 3.54$
855: which is substantially smaller than the 
856: acceptable field cutoff 4.9 mentioned above.
857: As one can see from Fig. \ref{fig:various}, this creates errors which 
858: are between one and two orders of magnitude smaller than the $1/N$ 
859: corrections. This is better but it compares poorly with what can reached 
860: with Pad\'e approximants.
861: 
862: Again, except for the comparison with independent numerical 
863: calculations at finite 
864: $N$, the same calculations can be performed for the SCM with minor changes.
865: Results very similar to those shown in Fig. \ref{fig:various} for 
866: the HM can be produced. Since it contains essentially the same 
867: information, it has not been displayed. It should 
868: however be noted that 
869: the number of significant digits obtained with the two alternate procedures 
870: are lower than in the case of the HM. In the case of the truncation of the 
871: range of integration, we need to restrict to 
872: $|\phi|< \sqrt{5\times 0.11}\simeq 0.74$ while a range of about 2 is 
873: required in order to obtain an accuracy consistent with the method of 
874: numerical integration.
875: 
876: \section{Conclusions}
877: We have shown in two differents models where the critical potential
878: can be calculated at leading order in the $1/N$ expansion that these
879: potentials have finite radii of convergence due to singularities in 
880: the complex plane. Do such a results persist at finite $N$?
881: In the case of the HM, the behavior of the ratios at finite $N$ shown
882: in Fig. \ref{fig:decayrat} strongly suggests that at large 
883: real positive $\phi^2$, 
884: the critical potential still grows like $(\phi^2)^3$. Can an infinite 
885: sum converging over the entire complex plane have this kind
886: of behavior? This is certainly not impossible 
887: (e.g., $(\phi^2)^3+{\rm e}^{-\phi^2}$), however it requires
888: cancellations that we judge unlikely to happen. Consequently,
889: we conjecture that the singularities observed are generic 
890: rather than being an artifact of the large-$N$ limit.
891: 
892: We have observed that in a system of coordinates where 
893: the HFP can be approximated by polynomials, the procedure 
894: which consists in considering bare potential truncated at order 
895: $(\phi^2)^3$ describes the HFP with a 
896: low accuracy. We are planning to investigate if similar problems 
897: appear near tricritical fixed points. In  particular, 
898: reconsidering the RG flows in 
899: a larger space of bare parameters may affect the generic 
900: dimension of
901: the intersections of hypersurface of various codimensions and help us 
902: finding a more general realization of spontaneous breaking of 
903: scale invariance with a
904: dynamical generation of mass.
905: 
906: Our results have qualitative similarities common with those 
907: of Refs. \cite{pathos}:
908: we found some ``pathologies'' which force us to look at the 
909: RG transformations in a more open-minded way.
910: We are planning \cite{prog} 
911: to compare in more detail, the leading order results 
912: presented here with finite $N$ results, as suggested in Ref. 
913: \cite{comellas} for the local potential approximation.
914: Another 
915: issue regarding the $O(N)$ models and which would deserve a more 
916: detailed investigation 
917: is the question of first order
918: phase transitions\cite{hazenfratz,van}.
919: 
920: \begin{acknowledgments}
921: We thank the Theory group of Fermilab for its hospitality while 
922: this work was completed and especially B. Bardeen for conversations 
923: about dynamical mass generation.
924: This research was supported in part by the Department of Energy
925: under Contract No. FG02-91ER40664.
926: \end{acknowledgments}
927: 
928: \begin{thebibliography}{10}
929: 
930: \bibitem{wilson}
931: K. Wilson, Phys.\ Rev.\ D  {\bf 6}, 419 (1972).
932: \bibitem{ma73}
933: E. Ma, Rev. Mod. Phys. {\bf 45}, 589 (1973).
934: \bibitem{col}
935: S. Coleman, R. Jackiw and H. Politzer, Phys. Rev. D {\bf 10} 2491 (1974).
936: \bibitem{town}
937: P. Townsend, Phys. Rev. D {\bf 14} 1715 (1976).
938: \bibitem{app}
939: T. Appelquist and U. Heinz, 
940: Phys. Rev. D {\bf 24} 2169 (1981).
941: \bibitem{pisarski}
942: R. Pisarski, Phys. Rev. Lett. {\bf 48}, 574 (1982).
943: \bibitem{bmb}
944: W. Bardeen, M. Moshe and M. Bander,  Phys. Rev. Lett. {\bf 52}, 1188 (1984).
945: \bibitem{amit84}
946: D. Amit and E. Rabinovici, Nucl. Phys. {\bf B 257}, 371 (1985).
947: \bibitem{bardeen}
948: W. Bardeen, C. Leung and S. Love,  Phys. Rev. Lett. {\bf 56}, 1230 (1986);
949: W. Bardeen and M. Moshe, Phys.\ Rev.\ D  {\bf 34}, 1229 (1986);
950: W. Bardeen, Fermilab preprint CONF-88/149-T (1988).
951: \bibitem{david84}
952: F. David, D. Kessler and H. Neuberger, Phys. Rev. Lett. {\bf 53}, 2071 (1984);
953: F. David, D. Kessler and H. Neuberger, Nucl. Phys. {\bf B 257}, 695 (1985).
954: \bibitem{exact}
955: There is a large literature on this question. Extensive lists of references 
956: can be found in two recent reviews: 
957: C. Bagnuls and C. Bervillier, Phys. Reports {\bf 348},
958: 91 (2001); J. Berges, N. Tetradis and C. Wetterich,  Phys. Reports {\bf 362},
959: 223 (2002).
960: \bibitem{dyson}
961: F. Dyson, Comm.\ Math.\ Phys.\ {\bf 12}, 91 (1969).
962: \bibitem{baker}
963: G. Baker, Phys.\ Rev.\ B{\bf 5}, 2622 (1972). 
964: 
965: \bibitem{loeffel69}
966: J. Loeffel, A. Martin, B. Simon, and A. Wightman, Phys. Lett. B {\bf 30},  656
967:   (1969).
968: 
969: \bibitem{kw}
970: H. Koch and P. Wittwer, Comm. Math. Phys. {\bf 164}, 627 (1994).
971: \bibitem{guide}
972: J. Godina, Y. Meurice, and M. Oktay, Phys. Rev. D {\bf 57}, 6326 (1998).
973: \bibitem{convpert}
974: Y. Meurice, Phys. Rev. Lett. {\bf 88}, 141601 (2002).
975: 
976: \bibitem{gam3}
977: J. Godina, Y. Meurice, and M. Oktay, Phys. Rev. D {\bf 57}, R6581 (1998);
978: J. Godina, Y. Meurice, and M. Oktay, Phys. Rev. D {\bf 59},  096002  (1999).
979: \bibitem{osc}
980: Y. Meurice, G. Ordaz and S. Niermann, J. Stat. Phys. {\bf 87}, 363 (1997).
981: \bibitem{osc1}
982: Y. Meurice, G. Ordaz and V. G. J. Rodgers, Phys. Rev. Lett. {\bf 75}, 
983: 4555 (1995).
984: %
985: \bibitem{baker96}
986: G. Baker and P. Graves-Morris, {\em Pad\'e Approximants} (Cambridge University
987:   Press, Cambridge, 1996).
988: \bibitem{pathos}
989: A. van Enter, R. Fernandez and A. Sokal, J. Stat. Phys. {\bf 72}, 879 (1993);
990: A. van Enter and  R. Fernandez, Phys. Rev. E {\bf 59}, 5165 (1999). 
991: \bibitem{prog}
992: J. J. Godina, L. Li, Y. Meurice and B. Oktay, work in progress.
993: \bibitem{comellas}
994: J. Comellas and A. Travesset, Nucl. Phys. {\bf B 498}, 539 (1997).
995: \bibitem{hazenfratz}
996: P. Hasenfratz and A. Hasenfratz, Nucl. Phys. {\bf B 295}, 1 (1988).
997: \bibitem{van}
998: A. van Enter and S. Shlosman, cond-mat/0205455.
999: \end{thebibliography}
1000: \end{multicols}
1001: \end{document}
1002: