cond-mat0312108/v11.tex
1: \documentclass[aps,prl,twocolumn,showpacs, floatfix]{revtex4}
2: %\documentclass[aps,prl,preprint,showpacs,floatfix]{revtex4}
3: \usepackage{amsmath,amssymb}
4: \usepackage[dvips]{graphicx}
5: 
6: %%%%%%%%%%%%%%%%%%%%%%%%%
7: %%% Additional filez required:
8: %%% - Plots: det_10.eps, 100.eps, ene.eps
9: %%% - Bibliography: detMC_1.bib
10: %%%%%%%%%%%%%%%%%%%%%%%%%
11: 
12: \DeclareMathOperator{\Sp}{Tr}
13: 
14: \begin{document}
15: 
16: \title{Truncated-Determinant Diagrammatic Monte Carlo
17: for Fermions with Contact Interaction}
18: \author{Evgueni Bourovski}
19: \author{Nikolay Prokof'ev}
20: \author{Boris Svistunov}
21: \affiliation{Physics Department, University of Massachusetts,
22: Amherst, USA, 01003} \affiliation{Russian Research Center
23: ``Kurchatov Institute'', 123182 Moscow }
24: \begin{abstract}
25: For some models of interacting fermions the known solution to the
26: notorious sign-problem in Monte Carlo (MC) simulations is to work
27: with macroscopic fermionic determinants; the price, however, is
28: a macroscopic scaling of the numerical effort spent on elementary
29: local updates. We find that the {\it ratio} of two macroscopic determinants
30: can be found with any desired accuracy by considering truncated (local in
31: space and time) matices. In this respect, MC for interacting fermionic
32: systems becomes similar to that for the sign-problem-free bosonic
33: systems with system-size independent update cost. We demonstrate
34: the utility of the truncated-determinant method by simulating
35: the attractive Hubbard model within the MC scheme based on partially summed
36: Feynman diagrams. We conjecture that similar approach may be
37: useful in other implementations of the sign-free determinant
38: schemes.
39: \end{abstract}
40: \pacs{02.70.Ss, 03.75.Ss, 71.10.Fd}
41: %\date{\today}
42: \maketitle
43: 
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: 
46: Monte Carlo (MC) methods are a unique tool for studying large
47: interacting systems. The most severe limitation on their
48: applicability is imposed by the so-called sign-problem (SP) when
49: relevant contributions to statistics alternate in sign and almost
50: exactly compensate each other in the final answer \cite{book}.
51: Frustrating interactions and anticommutation relations for fermion
52: operators are typically at the origin of the sign-problem. In this
53: Letter, we address the case when quantum statistics is
54: non-positive only because of the fermion exchange cycles.
55: 
56: One solution to the fermion SP is offered by the determinant Monte
57: Carlo (DetMC) (see, e.g. \cite{Loh92}). The idea is that all
58: contributions to the many-body statistics obtained by exchanging
59: fermion places (in a certain representation) can be written as a
60: product of two determinants---for spin-up and spin-down
61: species---and in cases when the two real determinants coincide
62: the result is positive definite. In Metropolis-type algorithms
63: \cite{Metr53}, MC updates are accepted with
64: probabilities proportional to the ratio of final and initial
65: configuration weights; in DetMC the corresponding acceptance
66: ratio, $R$, is based on the ratio of large determinants.
67: Unfortunately, calculating determinants ratio for macroscopically
68: large matrices is very expensive numerically: even with
69: tricks involving the Hubbard-Stratonovich tranformation the
70: algorithm proposed by Blankenbecler, Scalapino
71: and Sugar \cite{ScaSu@PRL81,BSS81} still requires $L^{2d}$
72: operations per update for a $d$-dimensional system with $L$
73: lattice points per dimension. The same scaling is true for the
74: continuous-time scheme \cite{Romb99}. In contrast, for bosonic
75: systems with local interactions the number of operations per
76: update is small and system-size independent, i.e. they can be
77: simulated $L^{2d}$ times faster!
78: 
79: Since the bottleneck of DetMC is the calculation of $R$, one may
80: question the paradigm of calculating it ``exactly'': In any case,
81: computer operations always involve systematic round-off errors.
82: The other example is provided by (pseudo)random number
83: generators---they are {\it always} imperfect and result in
84: systematic errors equivalent to small errors in $R$ which,
85: however, remain practically undetectable (for good generators) in
86: final results. The heuristic explanation of why small errors in
87: $R$ do not ruin the simulation is as follows. The Metropolis
88: algorithm is a scheme with strong relaxation towards equilibrium
89: distribution, and local configuration updates may be viewed as the
90: result of dissipative coupling to the thermal bath (this picture
91: is often used to model dissipative kinetics \cite{book}).
92: Uncontrolled errors in $R$ may then be regarded as a small
93: stochastic noise in the relaxational dynamics. As such, it only
94: slightly modifies the equilibrium state and its properties. This
95: is a standard argument in the linear response theory.
96: 
97: It seems natural then to suggest that if the goal is to simulate
98: the result with $n$-digit accuracy, there is no need to calculate
99: the acceptance ratio with accuracy much higher than $n$ digits.
100: Often, we simply ignore this issue because getting $R$ with
101: machine precision does not cost any extra CPU time. In determinant
102: methods, however, there is a potential of huge efficiency gains if
103: approximate values of $R$ can be calculated much faster. We
104: demonstrate the feasibility of this approach by showing that the
105: ratio of two macroscopic determinants can be found with high
106: accuracy by considering truncated matrices dealing only with the
107: local (in space-imaginary time) structure of the configuration
108: space. The computational cost of updates in the corresponding
109: ``truncated-determinant'' scheme is system-size independent---an
110: efficiency increase  $\propto L^{2d}$ for large $L$.
111: 
112: In what follows, we discuss the solution of the Hubbard model for
113: fermions within a simple diagrammatic MC scheme based on Feynman
114: diagrams partially summed over fermion propagator permutations
115: \cite{Romb99,Rub}. The resulting diagram weight is the square of
116: the determinant composed of finite-temperature fermion
117: propagators. We explain how the determinant ratio for local
118: updates may be calculated using truncated matrices, and
119: demonstrate the feasibility of the proposed approach. We also show
120: that going to larger system sizes has little effect on the scheme
121: performance. Finally, we conjecture that large efficiency gains
122: are expected in other sign-problem free DetMC schemes, e.g. in
123: lattice QCD simulations with quark fields \cite{QCD}. It is also
124: worth noting that in the diagrammatic DetMC scheme the update cost
125: does not depend directly on the lattice period, which is a big
126: advantage for simulations of dilute systems. By ``dilute" we mean
127: dilute with respect to the lattice, but not necessarily with
128: respect to the interaction; the latter can be effectively large
129: due to a resonance on a (quasi)-bound state. Correspondingly, the
130: new scheme is very promising for the study of ultra cold fermionic
131: systems in the regime of strong Feshbach resonant interaction,
132: including the crossover from the Bardeen-Cooper-Schrieffer pairing
133: to the Bose-Einstein condensation of molecules \footnote{Barbara
134: Goss Levi, Physics Today \textbf{56}, 18 (2003), and references
135: therein; for recent results see D. S. Petrov, C. Salomon and G. V.
136: Shlyapnikov, cond-mat/0309010; M. Greiner, C. A. Regal, and
137: Deborah S. Jin, Phys. Rev. Lett. \textbf{92}, 040403 (2004); S.
138: Jochim \textit{et al.}, Science, \textbf{32}, 2101 (2003); M. W.
139: Zwierlein \textit{et al.}, Phys. Rev. Lett. \textbf{91}, 250101
140: (2003).}.
141: 
142: %%%%%%%%%%%%%%%%%%%%%%%%%% MODEL-METHOD %%%%%%%%%%%%%%%%%%%%%%%%%%%
143: 
144: \textit{Model and method.} We consider interacting
145: lattice fermions with the Hubbard Hamiltonian $H = H_0 + H_1$:
146: %\begin{gather}
147: \begin{equation}
148: H_0 = \sum_{\mathbf{k}\sigma} \, (\epsilon_\mathbf{k} -
149: \mu)c^{\dagger}_{\mathbf{k}\sigma} c_{\mathbf{k}\sigma}, \; \;\;\;
150: % \label{ham_0}\\
151: H_1 = U\sum_{\bf x}n_{{\bf x} \uparrow}n_{{\bf x} \downarrow}, %
152: \label{ham_1}
153: \end{equation}
154: %\end{gather}
155: where $c^{\dagger}_{{\bf x} \sigma}$ is the fermion creation
156: operator, $n_{{\bf x}\sigma} = c^{\dagger}_{{\bf x} \sigma}c_{{\bf
157: x}\sigma}$, $\sigma =\, \uparrow, \downarrow$ is the spin index,
158: ${\bf x}$ runs over the $L^d$ points of the simple cubic lattice,
159: $\mathbf{k}$ runs over the corresponding Brillouin zone,
160: $\epsilon_\mathbf{k} =-2t \sum_{\alpha = 1}^d \cos k_{\alpha} a $
161: is the tight-binding dispersion law, and $\mu$ is the chemical
162: potential. For definiteness and numerical tests, we confine
163: ourselves to the $d=2$ spacial lattice with periodic boundary
164: conditions. We use the hopping amplitude, $t$, and lattice
165: constant $a$ as units of energy and distance, respectively.
166: 
167: Following Refs.~\cite{Romb99,Rub} we start with writing
168: the statistical operator in the interaction representation,
169: %$e^{-\beta H } = e^{-\beta H_0 } \mathcal{T}_{\tau }\exp
170: %\{-\int_0^\beta H_1 (\tau ) d\tau  \}$,
171: \begin{equation}
172: e^{-\beta H } = e^{-\beta H_0 } \mathcal{T}_{\tau }\exp
173: \{-\int_0^\beta H_1 (\tau ) d\tau  \}
174: \end{equation}
175: where $H_1 (\tau ) = e^{H_0 \tau } H_1 e^{-H_0 \tau }$ and
176: $\mathcal{T}_{\tau }$ is the time ordering operator, and expanding
177: it in powers of $H_1$:
178: \begin{eqnarray}
179: Z &=& \sum_{p=0}^{\infty} (-U)^p  \sum_{{\bf x}_1 \dots  {\bf x}_p
180:  } \int_{\tau_1 < \tau_2 <\dots <\tau_p< \beta}
181: \left( \prod_{i=1}^p d\tau_i \!\right) ~~~~~~~~~~ \nonumber \\
182: &\times & \Sp \left[ e^{-\beta H_0 }
183:  \prod_{i=1}^p c^{\dagger}_{\uparrow}(\mathbf{x}_i\tau_i)
184: c_{\uparrow}(\mathbf{x}_i\tau_i)
185: c^{\dagger}_{\downarrow}(\mathbf{x}_i\tau_i)
186: c_{\downarrow}(\mathbf{x}_i\tau_i) \right] .
187: \label{Z}
188: \end{eqnarray}
189: This expansion for the partition function generates
190: standard Feynman diagrams \cite{Fetter-Walecka}.
191: Graphically, each term is a set of
192: four-point vertices with two incoming (spin-up and spin-down), and
193: two outgoing (spin-up and spin-down) lines which connect vertices.
194: Each line is associated with the imaginary time fermion
195: propagator, $G_{\sigma}(\mathbf{x}_i-\mathbf{x}_j,\tau_i-\tau_j;
196: \mu,\beta) = - \Sp \left[ \mathcal{T}_\tau e^{-\beta H_0}
197: c_{\sigma}(\mathbf{x}_i\tau_i)
198:                c^{\dagger}_{\sigma}(\mathbf{x}_j\tau_j) \right]$.
199: 
200: A straightforward MC sampling of diagrammatic series would be
201: impossible because of the sign-problem. However, if for a given
202: configuration of $p$ vertices,
203: $\mathcal{S}_p = \{ ( \mathbf{x}_j,\tau_j ) ,\; j=1,\ldots,p \} $,
204: one sums over all $(p!)^2$ ways of connecting them by propagators,
205: then the result can be written as a {\it product of two
206: determinants}, one for spin up, and another for spin down (see
207: e.g. \cite{Loh92}). The differential weight of the vertex
208: configuration (or vertex diagram) is then
209: \begin{equation}
210: d\mathcal{P}(\mathcal{S}_p)= (-U)^p   \, \det\mathbf{A}^{\uparrow}
211: (\mathcal{S}_p) \det\mathbf{A}^{\downarrow}(\mathcal{S}_p)\,
212: \prod_{i=1}^p d\tau_i \;, \label{P}
213: \end{equation}
214: where
215: $\mathbf{A}^{\sigma}(\mathcal{S}_p)$ are $p\times p$ matrices:
216: $A^{\sigma}_{ij}= G_{\sigma}(\mathbf{x}_i-\mathbf{x}_j,\tau_i-\tau_j)$.
217: For equal number of up- and down-particles, $ \det\mathbf{A}^{\uparrow}
218: \det\mathbf{A}^{\downarrow} = [\det\mathbf{A} ]^2$, and negative
219: $U$ the vertex diagram weight is always positive. [At half
220: filling, $n_{\uparrow}+n_{\downarrow}=1$, the sign of $U$ changes
221: when hole representation is used for one of the spin components,
222: so this scheme may be also used for the repulsive Hubbard model.]
223: 
224: MC sampling in the vertex configuration space $(p,\mathcal{S}_p)$
225: can be performed by standard MC rules (see, e.g.
226: Refs.~\cite{DMC,Rub}) using just one pair of complementary updates
227: ${\cal D}$ and ${\cal C}$: in ${\cal D}$ one selects at random one
228: of the vertices and suggests to delete it from the configuration;
229: in ${\cal C}$  an additional vertex is suggested to be inserted at
230: some point randomly selected in the space-time box $\beta \times
231: L^d$. These updates decrease/increase the rank of $\mathbf{A}$ by
232: one. The acceptance ratio for the ${\cal D}$/${\cal C}$ pair of
233: updates is then based on the ratio of two determinants
234: \begin{equation}
235: R_{p} =  { \det \mathbf{A}(\mathcal{S}_{p+1}) /
236:            \det \mathbf{A}(\mathcal{S}_{p}) } \;,
237: \label{upd_ratio}
238: \end{equation}
239: where $ \mathcal{S}_{p+1} = \{ \mathcal{S}_p ,(\mathbf{x}_{p+1},
240: \tau_{p+1}) \} $ (we omit the spin index for brevity).
241: 
242: %%%%%%%%%%%%%%%%%%%%% FIGURE OF SPHERES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
243: %\begin{figure}[htb]
244: %\includegraphics[scale=0.45]{circles.eps}
245: %\includegraphics[width = \columnwidth,keepaspectratio=true]{circles.eps}
246: %
247: %\includegraphics[bb =110 200 700 720 ,width = 0.8\columnwidth]{circles.eps}
248: %\vspace{+1.cm}
249: %\caption{Truncated vertex configurations include all vertices
250: %(denoted by crosses) inside circles of various radii $\ell$ from
251: %the center.} \label{fig:circles}
252: %\end{figure}
253: %%%%%%%%%%%%%%%%%%%%%  FIGURE OF SPHERES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
254: 
255: The bottleneck of this simple scheme is in evaluating $R_{p}$ when
256: $p$ is macroscopically large.  The typical number of vertices is
257: determined by the number of particles, interaction strength, and
258: inverse temperature as $p \propto N \beta U $. The
259: truncated-determinant idea is to calculate $R_p$ much faster at
260: the expense of accuracy using the following conjecture originating
261: from physical, rather than mathematical, arguments. The vertex
262: configuration represents a sequence of virtual particle collisions
263: in the many-body system, and it is likely that {\it local} changes
264: in its structure depend only on the immediate neighborhood of the
265: updated region. [We note that the idea of employing the local
266: nature of the fermion-boson coupling has been used in
267: \cite{ScaSu@PRL81}, but it has not been extended to the fermionic
268: determinant.] Quantitatively, we define a norm, $\| \dots \|$, or
269: a distance, between vertices in space-time (several choices are
270: discussed below), and construct a truncated vertex configuration,
271: $\mathcal{S}_p^{(\ell)}$, such that all points in
272: $\mathcal{S}_p^{(\ell)}$ satisfy
273: \begin{equation}
274: \| (\mathbf{x}_j,\tau_j) - (\mathbf{x}_{p+1},\tau_{p+1}) \|
275: \leqslant \ell  \;. \label{local}
276: \end{equation}
277: Correspondingly, $ \mathcal{S}_{p+1}^{(\ell)} = \{
278: \mathcal{S}_p^{(\ell)} ,(\mathbf{x}_{p+1} \tau_{p+1}) \} $.
279: %The procedure is illustrated in Fig.~\ref{fig:circles}.
280: We may now use truncated configurations to calculate
281: the ratio (\ref{upd_ratio}) approximately as
282: \begin{equation}
283: R_p^{(\ell)} = \det \mathbf{A}( \mathcal{S}_{p+1}^{(\ell)} )
284: / \det \mathbf{A}(\mathcal{S}_{p}^{(\ell)}).
285: \label{ratio_cut}
286: \end{equation}
287: Clearly,  when $\ell \to L$ we recover the exact ratio. Our
288: conjecture is then that $R_p^{(\ell)}$ quickly converges to $R_p$
289: and there exists a healing length in the $(\mathbf{x},\tau)$-space
290: characterizing this convergence. If this is the case, then $\ell $
291: may be considered as a microscopic (system-size independent)
292: parameter controlling the accuracy and efficiency of simulation.
293: 
294: The proper choice of the norm $\| \dots \|$ depends on system
295: parameters. In the strongly correlated case
296: the natural units of distance and time are provided
297: by the Fermi momentum, $k_F$, and Fermi energy $\epsilon_F$.
298: One possibility is then
299: \begin{equation}
300: \|(\mathbf{x},\tau) - (\mathbf{x}',\tau ' ) \| =
301: \sqrt{
302:  k_F^2 ( \mathbf{x} - \mathbf{x}') ^2 +
303:  \epsilon_F^2 (\tau -\tau') ^2
304:  } \;.
305: \label{norm_sph}
306: \end{equation}
307: Geometrically, this measure results  in a set of vertices
308: $\mathcal{S}_{p}^{(\ell)}$ inside the  space-time ellipsoid
309: centered at $(\mathbf{x}_\mathrm{p+1},\tau_\mathrm{p+1})$. For
310: dense systems $\|(\mathbf{x},\tau) - (\mathbf{x}',\tau') \| = \max
311: \{ |\mathbf{x}-\mathbf{x}'| , |\tau - \tau'| \}$ is equally
312: appropriate and our data for the largest system were obtained
313: using this measure. At temperatures comparable to $\epsilon_F$ one
314: may account for all vertices in the $\hat{\tau }$-direction and
315: simply write $ \|(\mathbf{x},\tau) - (\mathbf{x}',\tau')
316: \|_\mathrm{cyl} = k_F | \mathbf{x} - \mathbf{x}' | $. The
317: corresponding geometrical figure is a $\beta$-cylinder. Similarly,
318: in small systems one may consider truncating configurations only
319: in time direction.
320: 
321: %%%%%%%%%%%%%%%%%%%%%%%% NUMERICAL RESULTS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
322: 
323: \textit{Numerical results and discussion.}
324: Our tests of the truncated-determinant scheme were done for
325: the attractive Hubbard model with $U=-4$, $\mu=-2$, $\beta=10$,
326: and periodic boundary conditions.
327: First, we simulated a small $L^2= 4^2$ cluster
328: for which the ground state energy (of 10 particles) is known from
329: exact diagonalization studies \cite{Husslein97}. Since the spatial
330: dimension is so small we truncate vertex configurations only in the
331: imaginary time direction.
332: In Fig.~\ref{fig:Hubb} we show how the result for energy converges
333: to the exact value. We stress, that at all stages of the MC simulation
334: we never even write the full configuration determinant, which rank
335: is about $2.5$ times larger than typical values of $p$ for the
336: cutoff radius $2$.
337: 
338: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FIGURE $Hussein-Energy$ %%%%%%%%%%%%%
339: \begin{figure}[htb]
340: %
341: \includegraphics[width = 0.99\columnwidth,keepaspectratio=true]{ene.eps}
342: %
343: \caption{Energy and density dependence on the imaginary
344: time cutoff for the Hubbard model for $L=4$.
345: Dots represent the MC data, and solid lines are the exact
346: diagonalization results for 10 particles \cite{Husslein97}.}
347: \label{fig:Hubb}
348: %
349: \end{figure}
350: 
351: In Fig.~\ref{fig:100} we present our data for the $L^2=100^2$
352: system---now, using determinant truncation both in space and in
353: time directions, Eq.~(\ref{norm_sph}). Remarkably, the convergence
354: is achieved around the same value of the truncation radius, which
355: proves that the computational cost per update is not subject to
356: macroscopic scaling.  It is instructive to see how data
357: convergence for energy correlates with the typical errors
358: introduced by the approximate calculation of $R_p$. In
359: Fig.~\ref{fig:ratios} we show examples of $R_p$ dependence on the
360: truncation radius for a number of randomly selected MC
361: configurations. Clearly, quite large fluctuations in $R_p$ are
362: statistically ``averaged out'' in the final result for energy. We
363: are not aware of any other method capable of simulating fermionic
364: systems of comparable size.
365: 
366: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FIGURE $Hussein-Energy$ %%%%%%%%%%%%%
367: \begin{figure}[htb]
368: %
369: \includegraphics[width = 0.99\columnwidth,keepaspectratio=true]{100.eps}
370: %
371: \caption{Potential energy dependence on the truncation
372: radius for the Hubbard model for $L=100$.}
373: \label{fig:100}
374: %
375: \end{figure}
376: 
377: 
378: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FIGURE $T=\epsilon_F/10$ %%%%%%%%%%%%%
379: \begin{figure}[htb]
380: %
381: \includegraphics[width = 0.99\columnwidth,keepaspectratio=true]{det_10.eps}
382: %\includegraphics[width = 8.5cm]{det_10.eps}
383: %
384: %\vspace{1.cm}
385: \caption{Determinant ratios $R_{p}^{(\ell)}$ as functions of
386: $\ell$  for randomly selected MC configurations for the $L=100$
387: system.} \label{fig:ratios}
388: \end{figure}
389: 
390: Apart from the Hubbard model tests, we have also verified that the
391: use of truncated-determinants for randomly seeded vertex
392: configurations works as nicely to speed up the calculation of
393: $R_p$.
394: 
395: 
396: The benchmark DetMC method by Blankenbecler, Scalapino and Sugar
397: (BSS) \cite{ScaSu@PRL81,BSS81} is based on the Trotter-Suzuki
398: imaginary time slicing, and the Hubbard-Stratonovich
399: transformation \cite{Hirsch83}. The rank of the matrix used in the
400: calculation of the acceptance ratio equals the number of lattice
401: sites $L^{d}$, and $\sim L^{2d}$ operations are required to find
402: its determinant. The necessity of handling large matrices,
403: although made possible by this method, requires both elaborate
404: finite-size scaling analysis of MC data \cite{Sewer02}, and
405: special efforts for the calculation stabilization at low
406: temperatures \cite{Loh92}. The contour-distortion stabilization
407: techniques (see, e.g. \cite{Rom97,Baeurle02}) help to alleviate
408: the sign problem, but still suffer from the severe scaling of the
409: computational cost per update. More recently, Rombouts, Heide, and
410: Jachowicz \cite{Romb99} improved the BSS scheme by formulating it
411: in the $\tau$-continuum. It is easy to directly compare this
412: scheme with ours because the starting point is exactly the
413: same---the expansion of the statistical operator in powers of $U$.
414: Rombouts {\it et al.} used the auxiliary Ising variables to
415: decompose four-point vertices into the sums of single-particle
416: exponentials, and arrived at the number of operations for
417: performing one update scaling as $L^{2d}$. At this point we notice
418: that while we work with the same vertex configuration structure,
419: in our scheme there is no extra summation over the auxiliary
420: variables, and the calculation of the acceptance ratio is
421: system-size independent.
422: 
423: Recently, a substantial improvement in lattice QCD simulations has
424: been achieved by including quark-loop effects (see, e.g.
425: \cite{QCD}). After the quarks are integrated out, their effects
426: are described by the macroscopic positive definite determinant,
427: $\det \mathbf{A}(U)$, where $U$ is the configuration of gluon
428: matrices. The conjecture is that the ratio $\det
429: \mathbf{A}(U)/\det \mathbf{A}(U')$ for local updates of gluon
430: matrices, $U \to U'$, can be calculated by accounting only for the
431: immediate neighborhood of the updated lattice bond.
432: 
433: We doubt that the truncated-determinant schemes will help to speed
434: up simulations with the sign-problem. If the average configuration
435: sign is small, the answer is determined by small differences
436: between the sign-positive and sign-negative contributions, i.e.
437: each contribution has to be calculated to much higher accuracy
438: than would be sufficient in the positive definite case. As a
439: result, higher and higher precision is required for $R_p$, and
440: advantages of our approach quickly vanish.
441: 
442: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
443: \bibliography{detMC_2}
444: 
445: \end{document}
446: