hep-lat0501023/v2.tex
1: %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %%   This file is part of the APS files in the REVTeX 4 distribution.
5: %%   Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %%   Copyright (c) 2001 The American Physical Society.
9: %%
10: %%   See the REVTeX 4 README file for restrictions and more information.
11: %%
12: %
13: % This is a template for producing manuscripts for use with REVTEX 4.0
14: % Copy this file to another name and then work on that file.
15: % That way, you always have this original template file to use.
16: %
17: % Group addresses by affiliation; use superscriptaddress for long
18: % author lists, or if there are many overlapping affiliations.
19: % For Phys. Rev. appearance, change preprint to twocolumn.
20: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21: %  Add 'draft' option to mark overfull boxes with black boxes
22: %  Add 'showpacs' option to make PACS codes appear
23: %  Add 'showkeys' option to make keywords appear
24: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
25: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
26: %\documentclass[pre,twocolumn,groupedaddress]{revtex4}
27: \documentclass[prd,twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}  
28: %\documentclass[aps,pre,preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
29: %\usepackage{graphicx}%
30: \usepackage{epsfig}%
31: \usepackage{longtable}
32: %\usepackage{revsymb}
33: % You should use BibTeX and apsrev.bst for references
34: % Choosing a journal automatically selects the correct APS
35: % BibTeX style file (bst file), so only uncomment the line
36: % below if necessary.
37: \bibliographystyle{apsrev}
38: 
39: \begin{document}
40: \setlength{\topmargin}{-0.25in}
41: % Use the \preprint command to place your local institutional report
42: % number in the upper righthand corner of the title page in preprint mode.
43: % Multiple \preprint commands are allowed.
44: % Use the 'preprintnumbers' class option to override journal defaults
45: % to display numbers if necessary
46: \preprint{U. of IOWA Preprint}
47: 
48: %Title of paper
49: \title{An example of optimal field cut in lattice gauge perturbation theory}
50: 
51: % repeat the \author .. \affiliation  etc. as needed
52: % \email, \thanks, \homepage, \altaffiliation all apply to the current
53: % author. Explanatory text should go in the []'s, actual e-mail
54: % address or url should go in the {}'s for \email and \homepage.
55: % Please use the appropriate macro foreach each type of information
56: 
57: % \affiliation command applies to all authors since the last
58: % \affiliation command. The \affiliation command should follow the
59: % other information
60: % \affiliation can be followed by \email, \homepage, \thanks as well.
61: \author{L.Li}
62: \author{Y. Meurice}
63:  \altaffiliation[Also at ]{Obermann Center for Advanced study}%Lines break
64: \email[]{ yannick-meurice@uiowa.edu}
65: %\homepage[]{Your web page}
66: %\thanks{This work was supported in part by the DOE}
67: %\altaffiliation{}
68: \affiliation{Department of Physics and Astronomy\\ The University of Iowa\\
69: Iowa City, Iowa 52242 \\ USA
70: }
71: 
72: %Collaboration name if desired (requires use of superscriptaddress
73: %option in \documentclass). \noaffiliation is required (may also be
74: %used with the \author command).
75: %\collaboration can be followed by \email, \homepage, \thanks as well.
76: %\collaboration{}
77: %\noaffiliation
78: 
79: \date{\today}
80: 
81: \begin{abstract}
82: We discuss the weak coupling expansion of a one plaquette $SU(2)$ lattice gauge theory. We show that the conventional perturbative series for the partition function has a zero radius of convergence and is asymptotic. 
83: The average plaquette is discontinuous at $g^2=0$. 
84: However, the fact that $SU(2)$ is compact provides a perturbative 
85: sum that converges toward the correct answer for positive $g^2$. This alternate method amounts to introducing a specific coupling dependent field cut, that turns the coefficients into  
86: $g$-dependent quantities. Generalizing to an arbitrary field
87: cut, we obtain a regular power series with a finite radius of convergence. At any order in the modified perturbative procedure, and for a given coupling, it is possible 
88: to find at least one (and sometimes two) values of the field cut that provide 
89: the exact answer.  This optimal field cut can be determined 
90: approximately using the strong coupling expansion. This allows us to interpolate 
91: accurately 
92: between the weak and strong coupling regions. We discuss the extension of the method to 
93: lattice gauge theory on a $D$-dimensional cubic lattice.
94: 
95: % insert abstract here
96: \end{abstract}
97: 
98: % insert suggested PACS numbers in braces on next line
99: %\pacs{05.50.+q, 11.10.Hi, 64.60.Ak, 75.40.Cx}
100: \pacs{11.15.-q, 11.15.Ha, 11.15.Me, 12.38.Cy}
101: % insert suggested keywords - APS authors don't need to do this
102: %\keywords{}
103: 
104: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
105: \maketitle
106: 
107: % body of paper here - Use proper section commands
108: % References should be done using the \cite, \ref, and \label commands
109: \section{Introduction}
110: Lattice gauge theory incorporates 
111: essential features of the strong interactions at short distance (asymptotic freedom) 
112: and large distance (confinement). 
113: Expansions in $1/\beta=g^2/2N$ and $\beta$ usually provide good approximations for the average value of gauge invariant quantities in the limit of small or 
114: large $\beta$. However, calculations in the intermediate region often 
115: require a numerical approach. 
116: 
117: There exists a  general method for calculating Wilson' s or Polyakov's 
118: loops in powers of $1/\beta$ 
119: \cite{heller84} in pure $SU(N)$ gauge theories (see Ref. \cite{capitani02} for
120: a more complete set of references on lattice perturbation theory). 
121: 
122: Much effort has been devoted calculating 
123: \begin{equation}
124: P\equiv \left\langle(1/N_p)\sum_p (1-(1/N)ReTrU_p)\right\rangle
125: \end{equation}
126: where $U_p$ denotes the usual product of links along 
127: a $1\times1$ plaquette and $N_p$ the number of plaquettes. $P$ can be obtained by  taking the 
128: derivative with respect to $\beta$ of the free energy density. 
129: Exact calculations of the coefficients of $P$ up to order 3 in $1/\beta$
130: \cite{alles93} and numerical calculations at order 8 \cite{direnzo95} and 10 \cite{direnzo2000} are available. 
131: The accuracy of the weak and strong coupling expansions at successive orders is shown in Fig. \ref{fig:su3} for $SU(3)$ in 4 dimensions. The figure makes clear that 
132: in the region $5<\beta<6$, none of the expansions (in powers of $\beta$ or $1/\beta$) is accurate.  
133: Unfortunately, this region is precisely the ``scaling window'' where one can extract 
134: information relevant to the continuum limit. In addition, the convergence of the 
135: weak coupling expansion is not completely understood. The analysis of the numerical series 
136: \cite{rakow2002} may suggest the possibility of a finite radius of convergence (the center of the circle being $g^2=0$).
137: This possibility is not expected on general grounds and is  
138: in contradiction with the discontinuity of $P$ when $g^2$ changes sign 
139: \cite{gluodyn04,anlqcd}. 
140: We are not aware of any independent argument in favor of a finite 
141: radius of convergence, and the most likely outcome is that the factorial 
142: growth of the series takes over at higher order. In Ref. 6, this order is 
143: estimated to be approximately 25, which is out of reach of numerical 
144: calculations.
145: \begin{figure}[ht]
146: \includegraphics[width=3.4in,angle=0]{su3-pl.eps}
147: \caption{$P$ versus $\beta$ for $SU(3)$ in 4 dimensions. The solid line represents 
148: the numerical values; the dashed lines on the left, successive orders in the strong 
149: coupling expansion; the dot-dash lines on the right, successive orders in  the 
150: weak coupling expansion.
151: \label{fig:su3}}
152: \end{figure}
153: 
154: In this article, we show that for a $SU(2)$ lattice gauge model on 
155: a single plaquette, the weak 
156: coupling expansion is asymptotic (has zero radius of convergence), 
157: but that it is possible 
158: to modify the perturbative procedure in order to get a convergent 
159: sum, which is an ``expansion'' in powers of $1/\beta$ but 
160: with $\beta$-dependent coefficients, that 
161: is accurate even in the strong coupling region. This work is motivated by 
162: recent results obtained in the case of scalar field theory \cite{convpert,optim03}
163: where the answer for similar questions in the case of nontrivial models 
164: can be guessed correctly by 
165: considering a single site integral. This is briefly reviewed in Sec. 
166: \ref{sec:motivations}. The main point is that the large order behavior of 
167: perturbation theory is related to large field configurations and that by 
168: cutting-off these configurations appropriately, we can obtain a series 
169: that converges to a value exponentially close to the exact one 
170: (for instance, errors of order ${\rm e}^{-\lambda\phi^4_{max}}$ for the simple 
171: integral discussed in Ref. \cite{optim03}).
172: Hereafter, we follow the same path for lattice gauge theories.
173: 
174: The $SU(2)$ model is introduced in Sec. \ref{sec:model} where we also discuss its 
175: connection to Bessel functions. It is worth noting that for 
176: compact groups, there are no large field contributions and consequently, the 
177: factorial growth of the perturbative series comes from adding the tails of integration, 
178: as done in asymptotic analysis of integrals \cite{bleistein} and in the conventional procedure used in lattice perturbation theory\cite{heller84}. This is explained in Section 
179: \ref{sec:weak}. In many quantum problems, the lack of convergence can be traced 
180: to the behavior of the model at negative coupling (see, however, Ref. \cite{bender98} 
181: for a proper definition). This question has been discussed for lattice gauge models in 4 dimensions \cite{gluodyn04}. In Sec. \ref{sec:neg}, we argue that there should be an essential 
182: singularity at zero coupling for the one plaquette model. In Sec. \ref{sec:instanton}, 
183: we show that the regular perturbation series (with the integration tails added) misses 
184: ``instanton effects'' of the form 
185: $\beta^{-1}{\rm e}^{-2\beta}$. 
186: 
187: In Sec. \ref{sec:cut}, we propose to modify the conventional perturbative method by 
188: introducing a field cut. With this modification, the series converges toward a 
189: value which in general is different than the exact one. However, at a given order in 
190: the weak coupling expansion, it is possible to pick an 
191: optimal field cut such that for a given coupling the answer is exact.
192: For the integral studied in this article, we found at least one solution 
193: at every order. This is not necessarily the case in general. For the 
194: integral studied in Ref. \cite{optim03}, we were able to prove that no such 
195: solution exists at odd order and that we could only minimize the error in that case.
196: In Sec. \ref{sec:optim}, we use the strong coupling expansion to determine approximately
197: this optimal field cut. In this approach, the field cut is given as a power series 
198: in $\beta$. A numerical study indicates that this series has a finite radius of 
199: convergence which increases with the order in $1/\beta$ considered. 
200: The method that we propose allows us to interpolate between the weak and strong 
201: coupling region. This is depicted in Fig. \ref{fig:pico}, which is the prototype 
202: of what we expect to accomplish in general. In the conclusions, we consider the implementation of the method for 
203: $D$-dimensional models and discuss three practical ways to calculate the 
204: modified coefficients.
205: \begin{figure}[ht]
206: \includegraphics[width=3.4in,angle=0]{pic0.eps}
207: \caption{Number of significant digits for the $SU(2)$ one plaquette integral, at 
208: order 7 in the weak coupling (dotted line W7), at order 2 in the strong coupling (empty circles SC2) and at order 7 of the modified perturbative method proposed here with 
209: an optimal field cut determined pointwise using the strong coupling expansion at order 2 (solid line W7S2).   
210: \label{fig:pico}}
211: \end{figure}
212: 
213: \section{Motivations}
214: \label{sec:motivations}
215: 
216: A common challenge for quantum field theorists consists in 
217: finding accurate methods in regimes where existing expansions break down. 
218: In the renormalization group language, this amounts to finding acceptable interpolations for the flows in intermediate regions between fixed points.
219: A discussion of this question for lattice gauge theories 
220: can be found in Refs. \cite{kogut79,kogut80}. 
221: In the case of scalar field theory, the weak coupling expansion is unable to reproduce the exponential suppression of the large field configurations operating 
222: at strong coupling. This problem can be cured \cite{convpert} by 
223: introducing a large field cutoff $\phi _{max}$ which eliminates Dyson's instability.
224: One is then considering a slightly different problem, however a judicious choice of $\phi _{max}$ can be used to reduce or eliminate \cite{optim03} the discrepancy with the original problem (i. e., the problem with no field cutoff). 
225: This optimization procedure can be approximately 
226: performed using the strong coupling expansion and naturally bridges the gap between 
227: the weak and strong coupling expansions. 
228: 
229: The study of the simple integral 
230: \begin{equation}
231: \int_{-\infty}^{+\infty}d\phi e^{-\frac{1}{2}\phi^2-\lambda \phi^4}\neq \sum_0^{\infty}
232: \frac{(-\lambda)^l}{l!} \int_{-\infty}^{+\infty}d\phi e^{-\frac{1}{2}\phi^2}\phi^{4l}
233: \label{eq:int}
234: \end{equation}
235: provides a good understanding about the role of large field configurations in 
236: the perturbative series. It helps identifying general features of the the effect 
237: of a field cut.
238: In particular, the dependence of the accuracy of the modified perturbative series
239: on the coupling and the field cut, is qualitatively very similar for the 
240: integral, the anharmonic oscillator and the hierarchical model in 3 dimensions
241: (see the similarity among the three parts of Fig. 2 in \cite{convpert}).
242: 
243: In order to generalize this procedure to gauge theory, we will first consider the simplest 
244: possible gauge model, namely the one plaquette integral
245: \begin{equation}
246: Z(\beta,N)=\int \prod_{l\in p} dU_l {\rm e} ^{-\beta(1-\frac{1}{N}Re TrU_p)}\ ,
247: \end{equation}
248: %with the usual convention
249: %\begin{equation}
250: %\beta=2N/g^2
251: %\end{equation}
252: After fixing the gauge so that $U=1$ on three sides of the plaquette, $Z$ becomes an integral over a single 
253: link
254: \begin{equation}
255: Z(\beta,N)=\int dU {\rm e} ^{-\beta(1-\frac{1}{N}Re TrU)}
256: \end{equation}
257: For an arbitrary gauge fixing prescription, $TrU$ becomes $Tr(U(g)U)$ with $g$ arbitrary 
258: and $Z$ is $g$-independent by virtue of the invariance of the Haar measure $dU$.
259: This integral and its moments appear in the strong coupling expansion 
260: \cite{kogut79,munster80,falcioni80,smit82}
261: and in the mean field treatment \cite{drouffe83} of $SU(N)$ gauge theories. In the one plaquette model,
262: \begin{equation}
263: P =-\frac{d}{d\beta}{\rm ln}Z\ .
264: \end{equation} 
265: The accuracy of successive orders in the $\beta$ and $1/\beta$ described in the 
266: following sections is shown in Fig. \ref{fig:su2one} that can be compared with Fig. \ref{fig:su3}.
267: \begin{figure}[ht]
268: \includegraphics[width=3.4in,angle=0]{su2-pl.eps}
269: \caption{ $P$ versus $\beta$ for $SU(2)$ on one plaquette. The solid line represents 
270: the numerical values; the dashed lines on the left, successive orders in the strong 
271: coupling expansion; the dot-dash lines on the right, successive order in  the 
272: weak coupling expansion.
273: \label{fig:su2one}}
274: \end{figure}
275: \section{The model considered here}
276: \label{sec:model}
277: 
278: In the following, we specialize the discussion to the case $N=2$ for which the Haar measure is very simple. From now on, the reference to $N$ will be dropped and we will use the notation $Z(\beta)$ for $Z(\beta,2)$. The explicit form is:
279: \begin{equation}
280: 	Z(\beta)=\frac{1}{\pi}\int_0^{2\pi}d\omega \sin^2(\omega/2){\rm e} ^{-\beta(1-\cos(\omega/2))}\ .
281: 	\label{eq:zsu2}
282: \end{equation}
283: Setting $u={\rm cos}(\omega/2)$, 
284: \begin{equation}
285: 	Z(\beta)=\frac{2}{\pi}\int_{-1}^1du \sqrt{1-u^2}{\rm e}^{-\beta(1-u)}
286: 	\label{eq:bform}
287: \end{equation}
288: and one recognizes from Eq. (8.431) of Ref. \cite{integrals} that the integral can be expressed in terms of the modified Bessel function $I_1$:
289: \begin{equation}
290: Z(\beta)=2 {\rm e}^{-\beta}I_1(\beta)/\beta
291: \end{equation}
292: Using the Taylor expansion Eq. (8.445) in Ref. \cite{integrals}, we can write
293: \begin{equation}
294: 2 I_1(\beta)/\beta=\sum_{l=0}^{\infty} \frac{1}{l!\Gamma(l+2)}(\beta/2)^{2l} \ .
295: \end{equation}
296: As in the case of the integral of Eq. (\ref{eq:int}), the presence of the factorial at the denominator implies that the strong coupling expansion 
297: (in powers of $\beta =4/g^2$) converges over the entire complex plane.
298: 
299: \section{The weak coupling expansion}
300: \label{sec:weak}
301: 
302: Assuming $\beta>0$, we set $t=\beta(1-u)$ in Eq. (\ref{eq:bform}) yields 
303: \begin{equation}
304: 	Z(\beta)=(2/\beta)^{3/2}\frac{1}{\pi}\int_0^{2\beta}dt t^{1/2}
305: 	{\rm e}^{-t}\sqrt{1-(t/2\beta)}
306: 	\label{eq:tint}
307: \end{equation}
308: If we expand the square root in the integral and exchange the sum and the integral (the validity of this procedure will be discusssed in Sec. \ref{sec:cut}), 
309: we obtain a converging sum:
310: \begin{equation}
311: \label{eq:notail}
312: 	Z(\beta)=(\beta\pi)^{-3/2} 2^{1/2} 
313: 	\sum_{l=0}^{\infty} A_l(2\beta)\beta^{-l}\ ,
314: 	\end{equation}
315: 	with 
316: 	\begin{equation}
317: 	A_l(x)\equiv 2^{-l}
318: 	\frac{\Gamma(l+1/2)}{l!(1/2-l)}\int_0^{x}dt {\rm e}^{-t}t^{l+1/2}\ ,
319: 	\label{eq:al}
320: \end{equation}
321:  
322: The convergence of the 
323: sum in Eq. (\ref{eq:notail}) can be established from the bounds 
324: \begin{equation}
325: \frac{{\rm e}^{-2\beta}}{l+3/2}(2\beta)^{l+3/2}	<\int_0^{2\beta}dt {\rm e}^{-t}t^{l+1/2} < \frac{1}{l+3/2}(2\beta)^{l+3/2}\ ,
326: \end{equation}
327: and the fact that $\Gamma(l+1/2)/l! <1$ for $l\geq1$. (This is true for $l=1$ and can be proved by induction multiplying the inequality by $(l+1/2)/(l+1) <1$).
328: Consequently, the sum converges at the same rate as $\sum l^{-2}$. 
329: Note that as in the case of the 
330: ground state  of the quantum mechanical double well, the first term is positive but all the remaining terms are negative.
331: 
332: Obviously, Eq. (\ref{eq:notail}) is not a power series in $\beta^{-1}$ since the ``coefficients'' $A_l(2\beta)$ are $\beta$-dependent. 
333: We can now obtain the conventional asymptotic expansion by two different methods. The first one consists in adding the tails to the integrals in Eq. (\ref{eq:al}), or in other words by replacing the incomplete gamma function by the gamma function. 
334: This is a standard procedure in asymptotic expansions of integrals \cite{bleistein}.
335: 
336: On then obtain the asymptotic expansion 
337: \begin{eqnarray}
338: &\ &	Z(\beta)\sim(\beta\pi)^{-3/2} 2^{1/2} \times \nonumber \\ 
339: &\ &	\sum_{l=0}^{\infty} (2\beta)^{-l}
340: 	\frac{(\Gamma(l+1/2))^2(l+1/2)}{l!(1/2-l)},
341: 	\label{eq:tail}
342: \end{eqnarray}
343: The terms of this sum now grow like $l!/2^l$ and the series is asymptotic. 
344: As all the signs are negative for $l\geq 1$, the Borel transform has singularities 
345: on the positive real axis.
346: 
347: It is instructive to rederive the expansion of Eq. (\ref{eq:tail}) by following the steps of lattice 
348: perturbation theory \cite{heller84}. We first set $\omega =gA$ in Eq. (\ref{eq:zsu2}) and expand the action and the Haar measure in powers of $g$. This leaves us with 
349: the integral of a power series in $g$ over the range 0 to $2\pi/g$ for $A$. 
350: The asymptotic
351: series (\ref{eq:tail}) is then recovered by letting the range of integration 
352: go to infinity. As the two methods amount to calculate the coefficients 
353: with different variables of 
354: integration, we obtain the same series, as can be checked explicitly up to 
355: high order. We emphasize that in lattice gauge theory with compact groups, 
356: there are no large field contributions. It is only for practical reasons that 
357: the tails of integration are added. In the one plaquette example, calculating 
358: $A_l(2 \beta)$ instead of $A_l(\infty)$ is a very minor problem, however, this is 
359: a technical challenge in the case on a $D$-dimensional lattice
360: 
361: \section{Behavior at negative $\beta$}
362: \label{sec:neg}
363: 
364: From the integral representation Eq. (\ref{eq:bform}), the change $\beta \rightarrow -\beta$ can be made by changing $u\rightarrow-u$ and multiplying by ${\rm e}^{2\beta}$. This implies, 
365: \begin{equation}
366: Z(-\beta)={\rm e}^{2\beta} Z(\beta)\ ,
367: \label{eq:minbet}
368: \end{equation}
369: and 
370: \begin{equation}
371: P(\beta)+P(-\beta)=2	
372: \label{eq:sumrule}
373: \end{equation}
374: A similar equation \cite{gluodyn04} can be found for a $SU(2)$ pure gauge model on a cubic lattice. Since $lim_{\beta\rightarrow+\infty}P(\beta)=0$,
375: the limits $g^2\rightarrow 0^{\pm}$ differ by 2 and a converging 
376: series in $g$ about 0 is impossible.
377: 
378: The discontinuity in the values of $P$ near $g^2=0$ appears in a simpler model 
379: where the integration over $SU(2)$ is replaced by a sum over the two elements of its 
380: center:
381: \begin{equation}
382: 	Z_{center}=\sum_{U=\pm \openone}{\rm e} ^{-\beta(1-\frac{1}{2}Re TrU)}=1+{\rm e}^{-2\beta}\ .
383: 	\label{eq:zcenter}
384: \end{equation}
385: This implies 
386: \begin{equation}
387: P_{center}=\frac{2}{1+{\rm e}^{2\beta}}\ .
388: \end{equation}
389: The center model satisfies Eqs. (\ref{eq:minbet}) and (\ref{eq:sumrule}).
390: Note that Eq. (\ref{eq:zcenter}) makes clear that $Z_{center}$ has an essential singularity
391: at $g=0$. The asymptotic behavior of $P_{center}$ at large $|\beta|$ in the complex plane is $2(1-{\rm e}^{2\beta}+\dots)$ if $Re\beta <0$, and $2{\rm e}^{-2\beta}+\dots$, if $Re\beta >0$, with Stokes lines running along the imaginary axis.
392: 
393: This simplified example makes clear that the usual perturbation series is obtained by making modifications of order ${\rm e}^{-2\beta}$ (the effect of the tails of integration). 
394: We now proceed to estimate the order ${\rm e}^{-2\beta}$ corrections to 
395: the integral over the whole $SU(2)$.
396: 
397: \section{Accuracy of regular perturbation theory}
398: \label{sec:instanton}
399: 
400: In the study of scalar models \cite{convpert}, we have shown that if we plot the 
401: accuracy of perturbative series at successive orders, an envelope 
402: setting the
403: boundary of the accuracy that can be reached at any order using the usual perturbation theory 
404: appears. In the case of the quantum mechanical double-well, this envelope coincides 
405: very precisely with the instanton effect.
406: We expect the limitation in accuracy to be of the general form 
407: $g^A{\rm e}^{-B/g^2}$. For the model considered here, 
408: the limitation of accuracy has this generic form and we will see that 
409: the effect is of order $\beta^{-1}{\rm e}^{-2\beta}$. 
410: 
411: For $\beta$ not too small, the low orders the asymptotic series Eq. (\ref{eq:tail}) overestimate $Z$. As we let the order increase, the series 
412: crosses the true value and then start to grossly underestimate the true value.
413: At each order, there is a special value of $\beta$ for which the truncated 
414: series coincides with the exact 
415: answer. This explains the ``spikes'' seen  in Fig. \ref{fig:envelope}.
416: \begin{figure}[ht]
417: \includegraphics[width=3.4in,angle=0]{fig5.eps}
418: \caption{Number of correct significant digits as a function of $\beta$ at successive orders
419: of the regular perturbative series Eq. (\ref{eq:tail}) for $Z(\beta)$. As the order increases from 1 to 15, the 
420: curves ($W1,\ W2,\dots)$ get lighter. The thick solid line is ${\rm log}_{10}(\beta^{-1}{\rm e}^{-2\beta}/Z)$.
421: \label{fig:envelope}}
422: \end{figure}
423: 
424: If we assume that for a particular value of $\beta$, the converging sum, Eq. (\ref{eq:notail}) with the integrals running from 0 to $2\beta$,  truncated at order $K$ is a good approximation of $Z(\beta)$, then the error $\delta Z(\beta, K)$ made by using the regular perturbative series, 
425: Eq. (\ref{eq:tail}) with the integrals running from 0 to $\infty$, truncated at the same order, is in good approximation
426: \begin{eqnarray}
427: 	\delta Z(\beta, K)\simeq (\beta \pi)^{-3/2} 2^{1/2} &\times & \nonumber \\ 
428: 	\sum_{l=0}^{K} (2\beta)^{-l}
429: 	\frac{\Gamma(l+1/2)}{l!(1/2-l)}\int_{2\beta}^{\infty}&dt& {\rm e}^{-t}t^{l+1/2}\ ,
430: 	\label{eq:delz}
431: \end{eqnarray}
432: 
433: Integrating by parts, dropping terms of order $\beta^{-1}$ and summing the 
434: resulting series, we obtain 
435:  
436: \begin{equation}
437: 	\delta Z(\beta, K)\approx A_K{\rm e}^{-2\beta}\beta^{-1}2\pi^{-3/2}\ , 
438: \end{equation}
439: with 
440: \begin{equation}
441: A_K=-\sum_{l=0}^{K}\frac{\Gamma(l-1/2)}{l!}\ .
442: \end{equation}
443: The coefficient $A_K$ slowly decreases when $K$ increases. For instance, $A_5=0.872\dots$, 
444: $A_{10}=0.624\dots$. In the limit of large $K$, 
445: $A_K$ becomes zero. This comes from  the 
446: resummation
447: \begin{equation}
448: \sum_{l=1}^{\infty}\frac{\Gamma(l-1/2)}{l!}=\int_0^{\infty}dt t^{-3/2}(1-{\rm e}^{-t})=2\pi^{1/2}\ ,
449: \label{eq:gamma}
450: \end{equation}
451: which is exactly the $l=0$ term. In practice, when the order is not too large, 
452: $\beta^{-1}{\rm e}^{-2\beta}$ is a good order of magnitude estimate for the 
453: envelope discussed above as can be seen in Fig. \ref{fig:envelope}.
454: 
455: 
456: \section{Modified model with an arbitrary field cut}
457: \label{sec:cut}
458: In this section,we consider a modified partition function $Z(\beta, t_{max})$ where the integration range 
459: of Eq. (\ref{eq:tint}) takes a fixed, $\beta$-independent, value 
460: $t_{max}$. 
461: When $t_{max}<2\beta$, the Taylor series for the square root converges absolutely and 
462: uniformly over the whole range of integration. It is thus justified to interchange 
463: the sum and the integral and we have 
464: \begin{equation}
465: \label{eq:mod}
466: 	Z(\beta, t_{max})=(\beta\pi)^{-3/2} 2^{1/2} 
467: 	\sum_{l=0}^{\infty} A_l(t_{max})\beta^{-l}\ .
468: \end{equation}
469: The original partition function as expressed in Eq. (\ref{eq:notail}) is obtained by 
470: taking the limit $t_{max} \rightarrow 2\beta$. 
471: Since the integral with upper boundary $t_{max}$ is obviously continuous in that limit and since we can use the $l^{-2}$ suppression to prove the continuity of the sum, 
472: the validity of Eq. (\ref{eq:mod}) extends which justifies 
473: Eq. (\ref{eq:notail}). If $t_{max}>2\beta$, the sum diverges and the integral is ill-defined.
474: The regular perturbation series is obtained by taking the 
475: limit $t_{max}\rightarrow\infty$. In the graphs, we use the notation ``WK'' for the truncation of the regular perturbative series at order $K$. One should however keep in mind that,  
476: for instance in $W7$, the last term is of order $(1/\beta)^{7+3/2}$ due to the 
477: prefactor $\beta^{-3/2}$ in Eq. (\ref{eq:mod}). 
478: 
479: Eq. (\ref{eq:mod}) is now a regular series in $(1/\beta)$. It has a finite radius 
480: of convergence. In order to calculate this radius, we notice that 
481: for large $l$, $\int _0^{t_{max}}dt t^l $ gets most of its contribution from the 
482: region between $t_{max}(1-1/l)$ and $t_{max}$. Consequently, one can replace 
483: ${\rm e}^{-t}t^{1/2}$ by ${\rm e}^{-t_{max}}t_{max}^{1/2}$ without affecting the 
484: asymptotic behavior of the coefficients of the series. 
485: If we perform this change directly in the integral Eq. (\ref{eq:tint}), 
486: the integral can be calculated explicitly. One can then conclude that 
487: $Z(\beta, t_{max})$ has a non-analytical part 
488: proportional to $(1-(t_{max}/2\beta))^{3/2}$. The series defined by Eq. (\ref{eq:mod}) converges if 
489: $(1/\beta)\leq (2/t_{max})$. Numerical studies of the series with conventional estimators confirm 
490: this argument.
491: Note that the finite radius of convergence of the series 
492: Eq. (ref{eq:mod}) is not in 
493: disagreement with the discontinuity of the original model at $1/\beta =0$, 
494: because this series coincides with the original model only when $(1/\beta)=(2/t_{max})$.
495: 
496: Can a truncation of the series of Eq. (\ref{eq:mod}) at order $K$ be a good approximation of the original integral Eq. (\ref{eq:zsu2})? The answer depends on $K$, $t_{max}$ and 
497: $\beta$. It is clear that if $K$ is large enough and $\beta$ slightly above  $2/t_{max}$, 
498: then one should get a reasonable approximation. This statement is confirmed by Fig. \ref{fig:intg2} where the accuracy of 
499: Eq. (\ref{eq:mod}) with $t_{max}=8$ truncated at orders 1 to 15 is displayed as a function of $\beta$. In this particular case, the values 
500: of $\beta\geq 4$ are within the radius of convergence. As the order increases, the spikes in this region (the right half of the Fig. \ref{fig:intg2}) move toward 4. In addition, there is 
501: another set of spikes, outside the radius of convergence 
502: (on the left half of the Fig. \ref{fig:intg2}) and moving in the opposite 
503: direction when the order increases.
504: \begin{figure}[ht]
505: \includegraphics[width=3.4in,angle=0]{fig6more.eps}
506: \caption{Number of correct significant digits as a function of $\beta$  for a fixed cut $t_{max}=8$. As the 
507: order increases from 1 to 15 ($W1,\ W2,\dots$), the curves become lighter.
508: \label{fig:intg2}}
509: \end{figure}
510: 
511: A more global information regarding the location of the spikes is displayed in Fig. \ref{fig:2sol}.  It shows that the ``second solution'',  outside the radius of 
512: convergence, disappears beyond some critical value of $\beta$. As the order in the weak 
513: coupling increases, both solutions get closer to the $t_{max}=2\beta$ line.
514: 
515: \begin{figure}[ht]
516: \includegraphics[width=3.4in,angle=0]{pic1.eps}
517: \caption{ Location of the exact matching between the series Eq. (\ref{eq:mod}) at 
518: order 6, 7 and 8 and $Z(\beta)$ in the $\beta$-$t_{max}$ plane. The dashed lines 
519: represent the solution within the radius of convergence and the empty circles the other 
520: solution.
521: \label{fig:2sol}}
522: \end{figure}
523: 
524: \section{Optimization}
525: \label{sec:optim}
526: In this section, we discuss an approximate method to find the optimal value of $t_{max}$ corresponding to a given order $K$ and a given value of $\beta$. 
527: In a general situation, we do not know accurately the value 
528: of the quantity that we are calculating (the equivalent of $Z$ here).
529: Consequently, we need to find an approximation that allows us to 
530: consistently estimate this quantity and the way its order $K$ approximation 
531: changes with the field cutoff in order to impose an approximate 
532: matching condition.
533: For this purpose, we will use the strong coupling expansion (power series in 
534: $\beta$) which provides information complementary to the weak coupling. 
535: Now, the crucial point is that the field cut allows us to control the $(1/\beta)$
536: in the integral Eq. (\ref{eq:tint}), because (except in the exponential) all the factor $(1/\beta)$ 
537: appear together with a factor $t$. 
538: In other words, except for the exponential,it is a function of $t/\beta$.
539: 
540: We would like to match the the strong coupling 
541: expansion 
542: \begin{equation}
543: Z(\beta)=1-\beta+(5/8)\beta^2 +\dots
544: \end{equation}
545: discussed in Sec. \ref{sec:model}, with
546: the truncated expansion of Eq. (\ref{eq:mod}) which can be rewritten as
547: 	\[ \pi^{-3/2}2\sum_{l=0}^K\frac{\Gamma(l+1/2)}{l!(1/2-l)}\int_0^{t_{max}/\beta}ds 
548: 	{\rm e}^{-s\beta}(s/2)^{l+1/2}\ .\]
549: The control of $s=t/\beta$ can be achieved by imposing that $t_{max}/\beta$ is approximately constant. We can then improve order by order in $\beta$ by setting
550: \begin{equation}
551: (t_{max}/\beta)=c_0 (K)+c_1(K)\beta +\dots
552: \label{eq:apbet}
553: \end{equation} 
554: The only non-trivial  part is to solve the zeroth order (in $\beta$) equation 
555: \begin{equation}
556: F_K(c_0 (K))=1\ ,
557: \label{eq:fk}
558: \end{equation}
559:  with 
560: \begin{equation}
561: F_K(x)=-4(\pi)^{-3/2}\sum_{l=0}^K\frac{\Gamma(l-1/2)	(x/2)^{l+3/2}}{l!(l+3/2)}
562: \ .
563: \end{equation}
564: We have checked that for $K$ going from 1 to 40,  Eq. (\ref{eq:fk}) has exactly two solutions on the positive real axis with one solution on each side of 2. As $K$ increases, the 2 roots get closer.  
565: They should coalesce at 2 in the large $K$ limit. This follows from the fact that  $F_{\infty}(2)=1$ 
566: and $F'_{\infty}(2)=0$ (as can be shown by using the the same method as for Eq. 
567: (\ref{eq:gamma})).
568: The higher order coefficients $c_l(K)$ corresponding to each 
569: of the two solutions at order 0 can then be found by solving linear equations. 
570: This procedure provides an approximate value of the optimal $t_{max}$ which apparently 
571: converges toward the correct numerical value. This is illustrated in Fig. \ref{fig:appropt}. 
572: \begin{figure}
573: \includegraphics[width=3.4in,angle=0]{pic2.eps}
574: \caption{Approximate locations in the $(\beta,t_{max})$ plane of the matching between the order 6 weak coupling 
575: expansion and $Z(\beta)$. The two solid lines are the two numerical solutions 
576: at that order (as in Fig. \ref{fig:2sol}). The dash line (empty circles) represent the first (second) approximate solutions at order $0,\dots,4$ in $\beta$.
577: \label{fig:appropt}}
578: \end{figure}
579: If we plug the two approximate values of $t_{max}$ of Eq. (\ref{eq:apbet}) in Eq. (\ref{eq:mod}) truncated at order $K$, we obtain 
580: approximate values of $Z(\beta)$. The accuracy of this procedure is displayed in 
581: Fig. \ref{fig:sd2} in the case $K=6$. It appears clearly that the first solution 
582: (the one within the radius of convergence with $t_{max}<2\beta$) is significantly more accurate than the 
583: other solution (with $t_{max}>2\beta$). Similar features were observed for $K$ up to 20. 
584: \begin{figure}[ht]
585: \includegraphics[width=3.4in,angle=0]{pic5.eps}
586: \caption{Significant digits obtained from Eq. (\ref{eq:mod}) truncated at order 6 using 
587: $t_{max}/\beta$ at order 0 (squares), 1 (circles), 2 (triangles) and 3 (stars). 
588: The first solution with $t_{max}<2\beta$ is showed with filled symbols, while the 
589: second solution is showed with empty symbols.
590: \label{fig:sd2}}
591: \end{figure}
592: 
593: We can now compare the accuracy of the method proposed here with the weak and strong 
594: coupling expansions. The case $K=6$ is displayed in Fig. (\ref{fig:sd}). In the weak 
595: coupling region ($\beta>10$) the accuracy of our procedure merges with the regular 
596: perturbation series. In the strong coupling region ($\beta<0.1$), our procedure 
597: is more accurate than the regular expansion in powers of $\beta$ by several 
598: significant digits. As $\beta\rightarrow 0$, the accuracy of our procedure 
599: with $t_{max}/\beta$ determined at order $m$ in $\beta$ increases at the 
600: same rate as the regular strong coupling expansion at order $m$ in $\beta$ maintaining 
601: the difference in accuracy approximately constant. In the intermediate region where 
602: none of the conventional expansions work well (except at the perturbative spike), our procedure 
603: maintains a very good accuracy interpolating smoothly between the two regimes.
604: 
605: \begin{figure}[ht]
606: \includegraphics[width=3.4in,angle=0]{pic3.eps}
607: \caption{ Significant digits obtained from Eq. (\ref{eq:mod}) truncated at order 6 using the first solution for 
608: $t_{max}/\beta$ at order 0 to 3 compared to the weak coupling expansion at order 6 
609: (dotted line W6) and the strong coupling expansion at order 0 to 2 (empty circles SC)
610: \label{fig:sd}}
611: \end{figure}
612: 
613: \section{Asymptotic behavior of $c_l(K)$}
614: \label{sec:asy}
615: 
616: In this section, we study empirically the asymptotic behavior of the coefficients $c_l(K)$ appearing in the expansion of $t_{max}/\beta$ Eq. (\ref{eq:apbet}). At fixed $K$ large $l$ , 
617: Fig. \ref{fig:ck1} suggests that 
618: \begin{equation}
619: c_l(K)\propto(G(K))^l	\ .
620: \end{equation}
621: In addition, it appears that $G(K)$ decreases with $K$ approximately like $1/K$.
622: This behavior implies a finite radius convergence $G(K)^{-1}$ for the $\beta $ expansion in Eq. (\ref{eq:apbet}), increasing linearly with $K$. This is good news 
623: for the interpolation between the weak and strong coupling region since
624: as we increase the weak order $K$, we increase the range of validity in $\beta$.
625: \begin{figure}[ht]
626: \includegraphics[width=3.4in,angle=0]{clk_1.eps}
627: \caption{ $\ln |c_l(K)|$ versus $l$ for $K=2, \ 6\ ,10$.
628: \label{fig:ck1}}
629: \end{figure}
630: 
631: The large-$K$ behavior of $c_l(K)$ has also been studied numerically. 
632: The results for $l$ up to 5 are shown in Fig. \ref{fig:ck2}.
633: \begin{figure}[ht]
634: \includegraphics[width=3.4in,angle=0]{clk_2.eps}
635: \caption{$\ln |c_0(K)-2|$ versus $\ln K$; $\ln |c_l(K)|$ versus $\ln K$ for $l=1,\dots 5$.
636: \label{fig:ck2}}
637: \end{figure}
638: At fixed $l$ large $K$ , the linear fits used in Fig. \ref{fig:ck2} suggest that
639: \begin{equation}
640: c_0 (K)\simeq 2+O(1/K)	\ , 
641: \end{equation}
642: and 
643: \begin{equation}
644: c_l(K)\propto K^{-l-1+\alpha(l)}	\ , 
645: \end{equation}
646: for $l>0$, with $\alpha(l)$ small.
647: This behavior is expected, since as the order 
648: increases, we are getting close to the
649: exact expansion Eq. (\ref{eq:notail}) with $t_{max}=2\beta$ ($c_0=2,\ c_l=0$ for $l>0$).
650: The values of $\alpha(l)$ decrease when we reduce the set of points fitted to
651: larger values of $K$. If we use $K=35$ to 45 for the fit, we have approximately
652: $\alpha(l)\simeq l/10$.
653: 
654: 
655: \section{Conclusions}
656: We have shown that for the one-plaquette model, the introduction of a properly chosen
657: field cut can provide a high accuracy
658: in regions where the usual perturbative method is not accurate.
659: The strong coupling expansion provides an efficient way to determine the optimal cut
660: and interpolate between the small and the large $\beta$ regions. 
661: Apparently, the 
662: accuracy of the calculations improves whenever we increase the order in either 
663: the weak or the strong coupling expansion. Given these positive results, we are compelled 
664: to implement the method in the case of lattice gauge theory 
665: on a $D$ dimensional cubic lattice. Two steps are necessary.
666: First, we need to define the theory with a field cut (the analog of Eq. (\ref{eq:tint}) with $2\beta$ replaced by $t_{max}$). Second, we need to 
667: expand relevant quantities such as $P$ for the modified theory in powers 
668: of $1/\beta$ (the analog of Eq. (\ref{eq:mod})). 
669: Note that in the calculation of P using a perturbative series, 
670: the complex zeroes of Z will play an important role. This question remains to 
671: be examined in detail.
672: 
673: The implementation of the first step is straightforward. One can insert 
674: 1 in the partition function in the following way:
675: \begin{equation}
676: 1=\prod_p\int_0^{t_{max}}dt_p\delta(1-(1/N)ReTr(U_p)-t_p)\ .
677: \end{equation}
678: If we could perform the integration over the $U_{link}$, we would get an effective theory 
679: for the new variables $t_P$. Note that the procedure is gauge invariant since 
680: $TrU_p$ is. The ``size'' of a configuration can be defined in several ways.
681: For instance, we could use the value of $ Max_p \left\{ t_p \right\} $ or $(1/N_p)\sum_p t_p$ to decide
682: if we have a large or a small field configuration. We can then order the configurations according to the chosen indicator. Given a (sufficiently large) set 
683: of Monte Carlo configurations, one can define the expectation values with a cut 
684: by averaging only over configurations for which the chosen indicator is below 
685: a certain value. The correlations between the two size indicators mentioned 
686: above are now being studied for $SU(3)$ in 4 dimensions.
687: 
688: The implementation of the second step requires technologies that are now being 
689: developed in the scalar case. As it seems only possible to make analytical 
690: calculations for small or large field cuts, numerical methods seem unavoidable.
691: For the purpose of independent verification, it is important to consider different 
692: methods. We are presently working on three different approaches:
693: \begin{enumerate}
694: \item
695: The conventional 
696: approach \cite{heller84} but with the  $A_\mu^a$ having a finite range of integration.
697: This type of approach works well in the scalar case \cite{lilipro}
698: \item
699: The stochastic approach \cite{direnzo95} where $A_\mu^a$ is expanded as 
700: power series in $1/\beta$. For the lowest order field, the implementation 
701: of a cut is obvious but not for higher order fields. This problem is being considered 
702: with simple examples.
703: \item
704: Fits from numerical data at large $\beta$. This method \cite{cookpro} 
705: allowed to extract at least 
706: 2 coefficients of conventional perturbation. As we mentioned above, it is easy to 
707: implement the field cut with Monte Carlo methods. The advantage of this method is
708: that it does not require the use of the Campbell-Baker-Hausdorff (for a short review see Ref. \cite{cbh}) formula.
709: \end{enumerate}
710: We expect that the use of theses three methods will allow us to 
711: contruct perturbative series with a finite radius of convergence as above. 
712: We hope that this radius of convergence will be sufficiently large to 
713: reach the scaling window. Ultimately, we expect to be able to replace the numerical calculation of the 
714: coefficients by approximate analytical formulas, as it seems possible to do 
715: in the case of the anharmonic oscillator \cite{interp}.
716: 
717: 
718: 
719: 
720: \begin{acknowledgments}
721: This research was supported in part by the Department of Energy
722: under Contract No. FG02-91ER40664.  
723: We thank Antonio Gonzalez Arroyo for discussions on Bessel functions.
724: % put your acknowledgments here.
725: \end{acknowledgments}
726: 
727: % Create the reference section using BibTeX:
728: %\bibliography{c:/papers/mainbib}
729: \begin{thebibliography}{23}
730: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
731: \expandafter\ifx\csname bibnamefont\endcsname\relax
732:   \def\bibnamefont#1{#1}\fi
733: \expandafter\ifx\csname bibfnamefont\endcsname\relax
734:   \def\bibfnamefont#1{#1}\fi
735: \expandafter\ifx\csname citenamefont\endcsname\relax
736:   \def\citenamefont#1{#1}\fi
737: \expandafter\ifx\csname url\endcsname\relax
738:   \def\url#1{\texttt{#1}}\fi
739: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
740: \providecommand{\bibinfo}[2]{#2}
741: \providecommand{\eprint}[2][]{\url{#2}}
742: 
743: \bibitem[{\citenamefont{Heller and Karsch}(1985)}]{heller84}
744: \bibinfo{author}{\bibfnamefont{U.~M.} \bibnamefont{Heller}} \bibnamefont{and}
745:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Karsch}},
746:   \bibinfo{journal}{Nucl. Phys.} \textbf{\bibinfo{volume}{B251}},
747:   \bibinfo{pages}{254} (\bibinfo{year}{1985}).
748: 
749: \bibitem[{\citenamefont{Capitani}(2003)}]{capitani02}
750: %For more references on lattice perturbation theory see:
751: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Capitani}},
752:   \bibinfo{journal}{Phys. Rept.} \textbf{\bibinfo{volume}{382}},
753:   \bibinfo{pages}{113} (\bibinfo{year}{2003}), \eprint{hep-lat/0211036}.
754: 
755: \bibitem[{\citenamefont{All\'es et~al.}(1994)\citenamefont{Alles, Campostrini,
756:   Feo, and Panagopoulos}}]{alles93}
757: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{All\'es}},
758:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Campostrini}},
759:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Feo}}, \bibnamefont{and}
760:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Panagopoulos}},
761:   \bibinfo{journal}{Phys. Lett.} \textbf{\bibinfo{volume}{B324}},
762:   \bibinfo{pages}{433} (\bibinfo{year}{1994}), \eprint{hep-lat/9306001}.
763: 
764: \bibitem[{\citenamefont{Di~Renzo et~al.}(1995)\citenamefont{Di~Renzo, Onofri,
765:   and Marchesini}}]{direnzo95}
766: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Di~Renzo}},
767:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Onofri}}, \bibnamefont{and}
768:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Marchesini}},
769:   \bibinfo{journal}{Nucl. Phys.} \textbf{\bibinfo{volume}{B457}},
770:   \bibinfo{pages}{202} (\bibinfo{year}{1995}), \eprint{hep-th/9502095}.
771: 
772: \bibitem[{\citenamefont{Di~Renzo and Scorzato}(2001)}]{direnzo2000}
773: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Di~Renzo}} \bibnamefont{and}
774:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Scorzato}},
775:   \bibinfo{journal}{JHEP} \textbf{\bibinfo{volume}{10}}, \bibinfo{pages}{038}
776:   (\bibinfo{year}{2001}), \eprint{hep-lat/0011067}.
777: 
778: \bibitem[{\citenamefont{Horsley et~al.}(2002)\citenamefont{Horsley, Rakow, and
779:   Schierholz}}]{rakow2002}
780: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Horsley}},
781:   \bibinfo{author}{\bibfnamefont{P.~E.~L.} \bibnamefont{Rakow}},
782:   \bibnamefont{and}
783:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Schierholz}},
784:   \bibinfo{journal}{Nucl. Phys. Proc. Suppl.} \textbf{\bibinfo{volume}{106}},
785:   \bibinfo{pages}{870} (\bibinfo{year}{2002}), \eprint{hep-lat/0110210}.
786: 
787: \bibitem[{\citenamefont{Li and Meurice}(2005)}]{gluodyn04}
788: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Li}} \bibnamefont{and}
789:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}},
790:   \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{71}},
791:   \bibinfo{pages}{016008} (\bibinfo{year}{2005}), \eprint{hep-lat/0410029}.
792: 
793: \bibitem[{\citenamefont{Li and Meurice}(2004)}]{anlqcd}
794: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Li}} \bibnamefont{and}
795:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}}
796:   (\bibinfo{year}{2004}), \eprint{hep-lat/0411020}.
797: 
798: \bibitem[{\citenamefont{Meurice}(2002)}]{convpert}
799: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}},
800:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{88}},
801:   \bibinfo{pages}{141601} (\bibinfo{year}{2002}), \eprint{hep-th/0103134}.
802: 
803: \bibitem[{\citenamefont{Kessler et~al.}(2004)\citenamefont{Kessler, Li, and
804:   Meurice}}]{optim03}
805: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Kessler}},
806:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Li}}, \bibnamefont{and}
807:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}},
808:   \bibinfo{journal}{Phys. Rev.} \textbf{\bibinfo{volume}{D69}},
809:   \bibinfo{pages}{045014} (\bibinfo{year}{2004}), \eprint{hep-th/0309022}.
810: 
811: \bibitem[{\citenamefont{Bleistein and Handelsman}(1986)}]{bleistein}
812: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Bleistein}} \bibnamefont{and}
813:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Handelsman}},
814:   \emph{\bibinfo{title}{Asymptotic Expansions of Integrals}}
815:   (\bibinfo{publisher}{Dover}, \bibinfo{address}{New York},
816:   \bibinfo{year}{1986}).
817: 
818: \bibitem[{\citenamefont{Bender and Boettcher}(1998)}]{bender98}
819: \bibinfo{author}{\bibfnamefont{C.~M.} \bibnamefont{Bender}} \bibnamefont{and}
820:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Boettcher}},
821:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{80}},
822:   \bibinfo{pages}{5243} (\bibinfo{year}{1998}).
823: 
824: \bibitem[{\citenamefont{Kogut et~al.}(1979)\citenamefont{Kogut, Pearson, and
825:   Shigemitsu}}]{kogut79}
826: \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{Kogut}},
827:   \bibinfo{author}{\bibfnamefont{R.~B.} \bibnamefont{Pearson}},
828:   \bibnamefont{and}
829:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Shigemitsu}},
830:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{43}},
831:   \bibinfo{pages}{484} (\bibinfo{year}{1979}).
832: 
833: \bibitem[{\citenamefont{Kogut and Shigemitsu}(1980)}]{kogut80}
834: \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{Kogut}} \bibnamefont{and}
835:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Shigemitsu}},
836:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{45}},
837:   \bibinfo{pages}{410} (\bibinfo{year}{1980}).
838: 
839: \bibitem[{\citenamefont{Munster and Weisz}(1981)}]{munster80}
840: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{M\"unster}} \bibnamefont{and}
841:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Weisz}},
842:   \bibinfo{journal}{Nucl. Phys.} \textbf{\bibinfo{volume}{B180}},
843:   \bibinfo{pages}{13} (\bibinfo{year}{1981}).
844: 
845: \bibitem[{\citenamefont{Falcioni et~al.}(1981)\citenamefont{Falcioni, Marinari,
846:   Paciello, Parisi, and Taglienti}}]{falcioni80}
847: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Falcioni}},
848:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Marinari}},
849:   \bibinfo{author}{\bibfnamefont{M.~L.} \bibnamefont{Paciello}},
850:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Parisi}}, \bibnamefont{and}
851:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Taglienti}},
852:   \bibinfo{journal}{Phys. Lett.} \textbf{\bibinfo{volume}{B102}},
853:   \bibinfo{pages}{270} (\bibinfo{year}{1981}).
854: 
855: \bibitem[{\citenamefont{Smit}(1982)}]{smit82}
856: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Smit}}, \bibinfo{journal}{Nucl.
857:   Phys.} \textbf{\bibinfo{volume}{B206}}, \bibinfo{pages}{309}
858:   (\bibinfo{year}{1982}).
859: 
860: \bibitem[{\citenamefont{Drouffe and Zuber}(1983)}]{drouffe83}
861: \bibinfo{author}{\bibfnamefont{J.-M.} \bibnamefont{Drouffe}} \bibnamefont{and}
862:   \bibinfo{author}{\bibfnamefont{J.-B.} \bibnamefont{Zuber}},
863:   \bibinfo{journal}{Phys. Rept.} \textbf{\bibinfo{volume}{102}},
864:   \bibinfo{pages}{1} (\bibinfo{year}{1983}).
865: 
866: \bibitem[{\citenamefont{Gradshteyn and Ryzhik}(1980)}]{integrals}
867: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Gradshteyn}} \bibnamefont{and}
868:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Ryzhik}},
869:   \emph{\bibinfo{title}{Table of Integrals, Series and Products}}
870:   (\bibinfo{publisher}{Academic Press}, \bibinfo{address}{London},
871:   \bibinfo{year}{1980}).
872: 
873: \bibitem[{\citenamefont{Li and Meurice}()}]{lilipro}
874: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Li}} \bibnamefont{and}
875:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}}, \bibinfo{note}{to
876:   be published}.
877: 
878: \bibitem[{\citenamefont{J.~Cook and Meurice}()}]{cookpro}
879: \bibinfo{author}{\bibfnamefont{J. ~Cook}, \bibnamefont{L.~Li}} \bibnamefont{and}
880:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}}, \bibinfo{note}{to
881:   be published}.
882: 
883: \bibitem[{\citenamefont{Zachos}()}]{cbh}
884: \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Zachos}},
885:   \bibinfo{note}{http://www.hep.anl.gov/czachos/CBH.pdf}.
886:   
887:   \bibitem[{\citenamefont{Li and Meurice}()}]{interp}
888: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Li}} \bibnamefont{and}
889:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Meurice}}, \bibinfo{note}{
890:   U. of Iowa preprint, hep-th/0503047}.
891: 
892: \end{thebibliography}
893: 
894: \end{document}
895: