hep-lat0011092/mg.tex
1: \newif\ifcbtex
2: \cbtexfalse
3: 
4: \ifcbtex
5: 
6: \documentclass[a4paper,12pt]{article}
7: \usepackage{times}
8: \usepackage{cb2}
9: \usepackage{epsf}
10: \usepackage{graphicx}
11: 
12: \pagestyle{cb}
13: \titleline{C.Best: Algebraic Multigrid for Disordered Systems\dots}
14: 
15: \else
16: 
17: \documentclass[preprint,aps]{revtex4}
18: \usepackage{graphicx}
19: 
20: \newcommand{\nnm}{\nonumber}
21: \newcommand{\setZ}{{\bf Z}}
22: \newcommand{\onehalf}{\mbox{$\scriptstyle 1\over 2$}}
23: \newcommand{\const}{\mbox{\rm const}}
24: 
25: \fi
26: 
27: \newcommand{\V}[1]{{\bf #1}}
28: \newcommand{\conj}{+}
29: \renewcommand{\and}{{\&} }
30: 
31: \begin{document}
32: 
33: \preprint{HLRZ2000\_21}
34: \title{Algebraic Multigrid for Disordered Systems\\and Lattice Gauge
35:   Theories}
36: 
37: \ifcbtex
38: 
39: \author{Christoph Best\\
40:   {\small John von Neumann Institute for Computing/DESY,}
41:   {\small 52425 J\"ulich, Germany}\\
42:   {\small c.best@computer.org}}
43: \date{November 30, 2000}
44: \maketitle
45: \vspace*{-1cm}
46: 
47: \else
48: 
49: \author{Christoph Best}
50: \email{c.best@computer.org}
51: \affiliation{%
52:   John von Neumann Institute for Computing/DESY,
53:   52425 J\"ulich, Germany}
54: 
55: \date{November 30, 2000}
56: 
57: \fi
58: 
59: \begin{abstract}
60:   The construction of multigrid operators for disordered linear
61:   lattice operators, in particular the fermion matrix in lattice gauge
62:   theories, by means of algebraic multigrid and block LU decomposition
63:   is discussed. In this formalism, the effective coarse-grid operator
64:   is obtained as the Schur complement of the original matrix.  An
65:   optimal approximation to it is found by a numerical optimization
66:   procedure akin to Monte Carlo renormalization, resulting in a
67:   generalized (gauge-path dependent) stencil that is easily evaluated
68:   for a given disorder field.  Applications to preconditioning and
69:   relaxation methods are investigated.
70: \end{abstract}
71: 
72: \ifcbtex
73: 
74: \vspace*{1cm}
75: {\small
76: \noindent{\bf Table of Contents}
77: 
78: \tableofcontents}
79: 
80: \clearpage
81: 
82: \else
83: 
84: \maketitle
85: 
86: \tableofcontents
87: 
88: \fi
89: 
90: \section{Introduction}
91: 
92: Most of the computer time in lattice gauge theory simulations is spent
93: today on inverting the Dirac matrix, which describes the dynamics of
94: quarks.  While this problem can be tackled efficiently by iterative
95: solvers such as Krylov subspace methods
96: \cite{heplat-9404013,heplat-9608074,gutknecht-wup-1999}, these solvers
97: suffer from critical slowing down in the physically interesting
98: regions of small quark masses. One of the limiting factors here is
99: that the elementary operation in iterative solvers, the matrix-vector
100: multiplication, affects only next neighbors and thus limits the speed
101: of information propagation over the lattice (though it can be improved
102: by ILU preconditioning \cite{heplat-9602019,heplat-0011080}).  Using
103: multigrid methods, one attempts to overcome this limitation by
104: propagating information on a hierarchy of coarser and coarser grids.
105: 
106: Multigrid methods \cite{hackbusch-springer,mccormick-siam} for the
107: Dirac matrix inversion were inspired by Mack's multigrid approach to
108: quantum field theory \cite{mack-1988} and have been proposed and
109: investigated around 1990 by groups in Boston
110: \nocite{juelich-1991}
111: \cite{prd-43-1965,brower-moriarty-rebbi-vicari,brower-edwards-rebbi-1991,
112:   vyas-jue-1991,prd-vyas}, Israel
113: \cite{plb-253-185,solomon-jue-1991,heplat-9204014,lauwers-benav-solomon-1992},
114: Amsterdam \cite{hulsebos-jue-1991,
115:   npb-331-531,plb-272-81,amsterdam-1991}, and Hamburg
116: \cite{kalkreuter-jue-1991,
117:   heplat-9304004,heplat-9310029,heplat-9409008,heplat-9408013,thesis-baeker}.
118: Though it was shown that multigrid methods would, in principle,
119: greatly reduce or eliminate critical slowing down, they have not been
120: able to improve the performance of the Dirac matrix inversion in
121: actual applications.  These classical multigrid methods are based on a
122: geometrical blocking of the lattice that leads to an effective
123: coarse-grid formulation of the original matrix by a Galerkin approach.
124: The choice of the blocking and interpolation kernels is crucial for
125: the performance of the multigrid algorithm; they must be chosen such
126: that the long-range and short-range dynamics decouple as much as
127: possible.  An optimal choice of the blocking kernels thus depends on
128: the gauge field and must in principle be recalculated for each gauge
129: configuration as the solution of a minimization problem. As this is
130: not feasible, various approximations have been introduced in the
131: literature.
132: 
133: % as the
134: % solution of an eigenproblem
135: % \cite{heplat-9304004,heplat-9310029,amsterdam-1991} or a variational
136: % minimization \cite{brower-edwards-rebbi-1991} on each point of the
137: % coarse lattice.
138: 
139: \nocite{wuppertal-1999-proceedings} 
140: 
141: Since then, the algebraic multigrid method \cite{siam-ruge-stueben,
142:   reusken-wup-1999,notay-wup-1999,brandt-2000,zhang-2000} has emerged
143: as a proven new method in numerical mathematics which does not rely on
144: a geometrical decomposition of the lattice, but solely on the matrix
145: itself.  It starts out from a thinned lattice that contains a certain
146: subset of the original lattice sites retaining their original values
147: and constructs an effective operator on this lattice by means of an
148: block $LU$ decomposition.  Thus, instead of geometric proximity, the
149: actual matrix elements determine in the coarsening procedure.  For
150: lattice gauge theory, this would automatically take the dynamics of
151: the gauge field, which appears as a phase factor in the matrix
152: elements, into account. (A similar reasoning was already cited by
153: Edwards, Goodman, and Sokal \cite{prl-61-1333} in one of the first
154: papers on multigrid methods for disordered systems, namely on random
155: resistor networks, though it was later assumed that algebraic
156: multigrid would be much more costly than the geometric variant
157: \cite{heplat-9204014}.) Recently, algebraic multigrid was applied to
158: lattice gauge theory by Medeke \cite{medeke-wup-1999}, and conversely
159: research about renormalization-group improved and perfect operators in
160: lattice gauge theory has been used to obtain efficient coarse-graining
161: schemes for partial and stochastic differential equations
162: \cite{condmat-0009449}.
163: 
164: %This reasoning was already mentioned by Edwards, Goodman, and Sokal
165: %\cite{prl-61-1333} in one of the first papers on multigrid methods for
166: %disordered systems and inspired the projective multigrid method of
167: %Brower, Rebbi, and Vicari \cite{prd-43-1965}. 
168: 
169: %and, as we will see, the
170: %idealized multigrid algorithm \cite{heplat-9304004} leads to the same
171: %choice for the coarsening kernel as the incomplete LU decomposition.
172: 
173: One advantage of algebraic multigrid over geometric methods is that it
174: directly gives a prescription for calculating the coarse-lattice
175: operator and the interpolation kernel by a submatrix inversion
176: problem. For this purpose, it is instructive to consider the problem
177: of matrix inversion not in terms of the spectral decomposition, but in
178: terms of the von Neumann series
179: \cite{kuti-neumann-series,vyas-random-walk}.  Combined with the block
180: $LU$ decomposition, this results in a simple computational picture of
181: the method as a noninteracting random walk on a hierarchically
182: decomposed lattice in a disordered background. The resulting algebra
183: of paths on this lattice is the basis for the computational
184: construction of multigrid operators below, where the expansion
185: coefficients are calculated numerically in a process similar to Monte
186: Carlo renormalization. In that way, physical (or rather,
187: computational) information about the fine-lattice operator is
188: harvested to approximate the coarse-lattice operator as closely as
189: desired. 
190: 
191: The lattice fermion matrix is an example of the more general problem
192: of disordered matrices operators which has many applications in
193: different fields of computational physics.  Some recent instances in
194: the literature are e.g.  the transport properties of light particles
195: in solid-state physics \cite{kehr-1996,guo-miller-2000}, or transport
196: in porous media, where multigrid methods have already been employed
197: \cite{moulton-knapek-dendy-1998,SKnapek:99b}. The numerical method
198: investigated here is sufficiently general to be applicable to such
199: problems.
200: 
201: Another factor that might turn out in favor of multigrid algorithms
202: comes from recent advances in computer technology.  Supercomputer
203: architecture has long been dominated by vector-oriented machines that
204: can quickly execute relatively simple and uniform elementary
205: operations on large vectors.  But because processor speed
206: advances much faster than memory access speed, the performance of
207: such operations is now severely limited by memory bandwidth.
208: In computer architecture, this is overcome by using
209: cache-oriented architectures such as in modern microprocessors, from
210: which e.g.~cluster computers \cite{iwcc-melbourne} are built.
211: These architectures are efficient for algorithms that have a high
212: balance, i.e.~number of operations per memory access, and make a
213: frequent reuse of memory locations
214: \cite{iwcc-melbourne-cbest,cs-0007027}, but can tolerate much more
215: complexity and irregularity. Multigrid algorithms usually result in
216: denser, but smaller matrices and might thus be favored by these
217: architectures.
218: 
219: Finally, problems other than matrix inversion have recently come into
220: focus for the Dirac operator. One is the calculation of its low-lying
221: eigenvalues, which was one of the first applications of multigrid
222: \cite{heplat-9408013} and is today used for investigating hadron
223: structure \cite{heplat-9709130} and chiral properties
224: \cite{heplat-0010049,heplat-0003021}.  Another one is the development
225: of chirally-improved Dirac operators such as the Neuberger operator
226: \cite{heplat-9806025,jansen-wup-1999} which requires the calculation
227: of the inverse square root of a matrix.  Algorithms for this problem
228: are still under development, but as they also employ iterative
229: solvers, it can be hoped that multigrid operators can aid in improving
230: performance there.  Multigrid methods have also been cited as a
231: possible improvement to the domain-wall formulation of lattice gauge
232: theory on a five-dimensional lattice \cite{heplat-0007003}.
233: 
234: The layout of the paper is as follows: In section 2, we introduce the
235: lattice Klein-Gordon and Wilson-Dirac operator and their
236: interpretation as diffusion on a disordered lattice, and in section 3,
237: the general algebraic multigrid procedure.  The numerical procedure to
238: construct multigrid operators is discussed in section 4, and some
239: applications to preconditioning and relaxation algorithms are
240: discussed in section 5.
241: 
242: \section{Model}
243: 
244: The model considered here is the Euclidean lattice Klein-Gordon and
245: Dirac operators with a $U(1)$ gauge field, discretized on a square
246: lattice of lattice spacing $a=1$. The formalism will also apply to
247: other gauge groups and can be generalized to other operators and other
248: types of disorder. Numerical calculations are performed in two
249: dimensions and on relatively small lattices ($16\times 16$), so that
250: the complete spectrum of the operators can be analyzed.
251: 
252: %Dirac
253: %operators describe fermion fields in lattice gauge theories, and the
254: %calculation of their Green's functions in a given gauge field consumes
255: %most computer time in modern lattice field theory.  The Klein-Gordon
256: %operator describes a bosonic field and does not have the internal spin
257: %degrees of freedom.  Both are linear differential operators with a
258: %link disorder introduced by a gauge field generated in a Monte Carlo
259: %process.
260: 
261: \subsection{Klein-Gordon operator \label{sec:KleinGordon}}
262: 
263: For simplicity, we start with the Klein-Gordon operator
264: which, without a gauge field, is conventionally written as
265: \begin{equation}
266:   \label{eq:29}
267:   - \Delta(x,y) + m^2 \delta_{x,y}
268: \end{equation}
269: with the next-neighbor
270: lattice Laplacian $\Delta(x,y)$ and the (bare) mass $m$. Its
271: eigenmodes are discrete Fourier modes, and its Green's functions decay 
272: exponentially as $e^{-m |x-y|}$ 
273: with the inverse correlation length $m$. Gauge disorder
274: is introduced as a $U(1)$ gauge phase $U_\mu(x)$ associated with the
275: links 
276: $(x,\mu)$ of the 
277: lattice. The disordered Klein-Gordon operator can be rewritten, up to a
278: normalizing factor, as
279: \begin{eqnarray}
280:   M(x,y) &=& \delta_{x,y}- \kappa Q(x,y) \quad, \nnm\\
281:   Q(x,y) &=& \sum_{\mu=1}^{\pm D}
282:                 \delta_{y,x+e_\mu} U_\mu(x)  \quad.      \label{eq:1}
283: \end{eqnarray}
284: Here $x$, $y$ are sites on the lattice, $\mu$ labels positive and
285: negative directions $\pm 1,\ldots,\pm D$, $e_\mu$ is the unit vector
286: in direction $\mu$, and we use the convention $e_{-\mu} = -e_\mu$ and
287: $U_{-\mu}(x) = U^\conj_\mu(x - e_\mu)$.  The completely hopping
288: parameter $\kappa$ are borrowed from fermion terminology and
289: related to the bare mass $m$ in (\ref{eq:29}) by
290: \begin{equation}
291:   \label{eq:34}
292:   \kappa = \frac{1}{m^2 + 2D} \quad.
293: \end{equation}
294: It determines the spectrum of $M$ and thus the physical properties of
295: the theory. For the free theory, i.e.~$U_\mu(x) \equiv 1$, the
296: eigenvalues of $Q$ lie between $-2D$ (checkerboard configuration) and
297: $2D$ (completely constant configuration). Consequently, the
298: eigenvalues $\lambda_n$ of $M$ lie in a band of width $2D\kappa$
299: around $1$:
300: \begin{equation}
301:   \label{eq:91}
302:   1 - 2D \kappa \le \lambda_n \le 1 + 2D\kappa \quad.
303: \end{equation}
304: For values of $\kappa$ below a critical value $1/2D$,
305: \begin{equation}
306:   \label{eq:36}
307:   \kappa < \frac{1}{2D} = \kappa_{\rm crit}
308: \end{equation}
309: the spectrum of $M$ is strictly positive and $M$ invertible. The
310: situation for a theory in a gauge field will be discussed in the next
311: section. 
312: 
313: Physically, the interesting quantity in $M$ is the location of the
314: lowest eigenvalue that determines the mass gap of the theory and the
315: exponential decay of its Green's functions, i.e.~the correlation
316: functions or propagators. As $\kappa$ approaches $\kappa_{\rm crit}$,
317: the mass gap becomes smaller and the correlation length increases,
318: until at $\kappa = \kappa_{\rm crit}$, a zero eigenvalue appears and
319: the correlation functions, after proper subtraction, decays
320: logarithmically or rationally, depending on the dimension $D$.
321: 
322: The performance of iterative solvers for the inverse of
323: $M$ is linked to the condition number, i.e.~the ratio of the largest
324: to the smallest eigenvalue, that diverges as $\kappa \to
325: \kappa_c$. Unfortunately, this is exactly the region that is of
326: interest in lattice gauge theories, where quarks are much lighter
327: than the typical scales of the theory.
328: 
329: \subsection{Diffusion representation of $M^{-1}$\label{sec:diffusion}}
330: 
331: Conventionally, eq.~(\ref{eq:1}) is interpreted in lattice field theory
332: in terms of its spectrum that determines the dispersion relation and
333: thus the mass of the physical particle.  However, this description is
334: less useful in the presence of a gauge field, as there is no
335: simple expression for the eigenmodes and eigenvalues of $M$.
336: Alternatively, $M$ can also be interpreted as describing the diffusion
337: of noninteracting random walkers, i.e.~Brownian motion,
338: on the lattice. This can most easily
339: be seen by using the von Neumann series to calculate $M^{-1}$:
340: \begin{eqnarray}
341: \label{eq:31}
342: (1-\kappa Q)^{-1} &=& 1 + \kappa Q + (\kappa Q)^2 + (\kappa Q)^3 + 
343:   \ldots 
344: \nnm \\ (1-\kappa Q)_{xy}^{-1} &=&
345: \sum_{n=0}^\infty 
346: \sum_{\mu_1}^{\pm D} \ldots \sum_{\mu_n}^{\pm D} \, 
347:   \delta_{y,x+e_{\mu_1}+ \cdots \, 
348:   +e_{\mu_n}} \, \kappa^n \, {\cal P}(U;x,\mu_1,\ldots,\mu_n) \quad.
349: \end{eqnarray}
350: The sum here runs over all connected paths leading from $x$ to $y$, and
351: each path carries a gauge phase assembled from the links it traverses 
352: of the lattice:
353: \begin{eqnarray}
354:   \label{eq:38}
355:   {\cal P}_{\mu_1,\ldots,\mu_m}(U;x)
356:   &= & {\cal P}(U;x,x+e_{\mu_1},\dots,x+e_{\mu_1}+\ldots+e_{\mu_{n}})
357:   \\
358:   \label{eq:38b}
359:    &=& 
360:    U_{\mu_1}(x) U_{\mu_2}(x + e_{\mu_1}) \ldots
361:    U_{\mu_n}(x+e_{\mu_1}+\cdots+e_{\mu_{n-1}}) \quad.
362: \end{eqnarray}
363: The diffusion interpretation is based of the fact that the sum in
364: eq.~(\ref{eq:31}) can be generated by a random walk starting at $x$
365: and making random moves along the links of the lattice until reaching
366: site $y$.
367: %The probability for a random walker to
368: %reach lattice site $y$ when starting out from site $x$ in $n$ steps is 
369: %\begin{equation}
370: %  \label{eq:93}
371: %  p_n(y|x) = \sum_{(\mu_1,\ldots,\mu_n)}^{\pm D} \,
372: %             \delta_{y,x+e_{\mu_1}+ \cdots +e_{\mu_n}} \,
373: %             (2D)^{-n} \quad,
374: %\end{equation}
375: %and the time-integrated probability, i.e.~the average time spent by
376: %all random walkers at site $y$, is given by $\sum_n p_n(y|x)$.  
377: Since the probability for each path of length $n$ is $(2D)^{-n}$,
378: eq.~(\ref{eq:31}) can be rewritten as an ensemble average over
379: random walkers:
380: \begin{equation}
381:   \label{eq:94}
382:   (1-\kappa Q)^{-1}_{xy} = \sum_n \left<
383:   \left(\frac{\kappa}{2D}\right)^n \, {\cal P}(U;x,\mu_1,\ldots,\mu_n) 
384:   \, \delta_{y,x+e_{\mu_1}+ \cdots +e_{\mu_n}}
385:   \right>_n
386: \end{equation}
387: where the average is over all random paths of length $n$, and each
388: random walker carries a phase factor.
389: 
390: In the free case, and for $\kappa=2D$, eq.~(\ref{eq:94}) is just the
391: time-integrated probability, or average time spent, at site $y$ for a
392: random walker starting out at $x$. Other values of $\kappa$ correspond
393: to spontaneous creation or absorption of walkers.  In the presence of
394: gauge disorder, each walker additionally carries a phase factor ${\cal
395:   P}(U)$ picked up from the links it has travesed.  Their
396: contributions can therefore interfere destructively, leading to a
397: stronger decay of the Green's function than in the free case:
398: \begin{equation}
399:   \label{eq:95}
400:   \left|(1-\kappa Q(U))^{-1}_{xy}\right| \le 
401:   \left|(1-\kappa Q(U=0))^{-1}_{xy}\right| \quad.
402: \end{equation}
403: In particular, this means that for those $\kappa$ for which the free
404: correlation function exists, i.e.~for $\kappa < 1/2D$, the disordered
405: correlation function also converges, and the critical hopping
406: parameter $\kappa_{\rm crit}$ of the disordered theory therefore must
407: be equal or larger than $1/2D$. From the discussion in the preceding
408: section, this reflects an upward shift of the lowest eigenvalue of $M$
409: that is interpreted as a dynamically generated mass. 
410: 
411: The value of the dynamical mass depends on the correlation
412: characteristics of the gauge field. If the gauge field is strongly
413: correlated, similar paths will have similar phase factors, the
414: destructive interference will be less severe and the dynamical mass
415: smaller than for more disordered fields. In this way, the diffusion
416: representation provides a qualitative picture of the spectrum of $M$
417: in a disordered field.
418: 
419: \subsection{Wilson-Dirac operator}
420: 
421: The Dirac operator describes the fermionic quark fields in lattice
422: gauge theory simulations. These fields possess an internal spin degree
423: of freedom and are described by a $\lfloor d/2 \rfloor$-component
424: spinor field. The Dirac equation is a first-order differential
425: equation, and its naive discretization on the lattice suffers from
426: fermion doubling, which is removed in the Wilson-Dirac discretization:
427: \begin{eqnarray}
428:   \label{eq:59}
429:   M_D(x,y) &=& \delta_{x,y}- \kappa Q_D(x,y) \quad,\nnm\\
430:   Q_D(x,y) &=& \sum_{\mu=1}^{\pm D}
431:                 \delta_{y,x+e_\mu} 
432:                 \left( 1 - \gamma_\mu \right)
433:                  U_\mu(x)             \quad.
434: \end{eqnarray}
435: where in two dimensions the Dirac matrices $\gamma_\mu$ are given by
436: \begin{equation}
437:   \label{eq:60}
438:   \gamma_1 = 
439:   \left( \begin{array}{cc}  1 & 0 \\ 0 & -1  \end{array} \right)
440:   \quad,\quad
441:   \gamma_2 = 
442:   \left( \begin{array}{cc}  0 & 1 \\ 1 & 0  \end{array} \right) \quad.
443: \end{equation}
444: Together with the unit matrix and the matrix $\gamma_5 = i \gamma_1
445: \gamma_2$, they form a basis for the linear operators in spinor
446: space. The Wilson-Dirac operator is a non-Hermitian operator and thus
447: has in general complex eigenvalues and different left and right
448: eigenvectors, but a Hermitian Wilson-Dirac operator can by defined by
449: multiplying it with $\gamma_5$:
450: \begin{equation}
451:   \label{eq:61}
452:   M_{HD} = \gamma_5 M_D \quad.
453: \end{equation}
454: For the matter of inversion, this operator is completely equivalent to
455: the original Wilson-Dirac operator. Due to the two-component spinors
456: it acts upon, it has twice as many degrees of freedom as the
457: Klein-Gordon operator. These additional degrees of freedom show up in
458: the spectrum of $M_{HD}$ as a band of negative eigenvalues mirroring
459: the positive band seen in the Klein-Gordon operator. The argument
460: used above about the relation of $\kappa$ to the low-lying
461: eigenvectors and the mass gap remain valid for the Dirac operator.
462: 
463: \section{Algebraic multigrid}
464: 
465: %Traditional multigrid approaches that have been tried in lattice gauge
466: %theory relied on blocking, where several geometrically close lattice
467: %sites are replaced by an effective average field on a coarser lattice.
468: %However, the presence of the gauge field makes this process difficult
469: %in particular with respect to maintaining gauge invariance and not
470: %introducing artifacts.  Here, we instead consider the algebraic
471: %multigrid method which only relies on the algebraic structure of the
472: %matrix $M$.
473: 
474: \subsection{Schur complement}
475: 
476: We consider the solution of a linear system
477: \begin{equation}
478:   \sum_y M(x,y) f(x) = a(y)
479:   \label{eq:2}
480: \end{equation}
481: with the right-hand side $a$ given and $f$ unknown. 
482: In the algebraic multigrid approach, the sites of the lattice ${\cal L}$ are
483: decomposed into a coarse lattice ${\cal L}_1$ and a fine complement lattice 
484: ${\cal L}_2 = {\cal L} \backslash {\cal L}_1$. This decomposition can
485: in principle be performed without reference to the geometric locations 
486: of the sites. In particular, the coarse-lattice sites are not block
487: averages, but retain the same values as they had on the fine lattice.
488: Eq.~(\ref{eq:2}) can then be rewritten as
489: \begin{equation}
490:   \label{eq:3}
491:   \left(\begin{array}{cc} M_{11} & M_{12} \\ M_{21} & M_{22} \end{array}
492:   \right) \,
493:   \left( \begin{array}{c} f_1 \\ f_2 \end{array} \right)
494:   =
495:   \left( \begin{array}{c} a_1 \\ a_2 \end{array} \right)
496: \end{equation}
497: where the column vectors are reordered so that $f_1$ and $a_1$ are
498: defined on the coarse lattice, and $f_2$, $a_2$ on the fine
499: lattice. Applying Gaussian elimination, the fine-lattice field
500: $f_2$ is eliminated completely from the first row:
501: \begin{eqnarray}
502:   \label{eq:4}
503:   (M_{11} - M_{12} M_{22}^{-1} M_{21}) \, f_1 &=&
504:   (a_1 - M_{12} M_{22}^{-1} a_2) \\
505:   \label{eq:5}
506:   M_{22} \, f_2 &=& ( a_2 - M_{21} f_1 )
507: \end{eqnarray}
508: The matrix operator in front of $f_1$ plays the role of an effective
509: operator on the coarse lattice. It is called the {\em Schur
510: complement} $S_M$ of the matrix $M$ with respect to the decomposition
511: ${\cal L}_1$:
512: \begin{equation}
513:   \label{eq:6}
514:   S_M = M_{11} - M_{12} M_{22}^{-1} M_{21} \quad.
515: \end{equation}
516: The simplest example of a
517: Schur complement is the odd-even decomposition of a square matrix with
518: next-neighbor interactions, where $M_{22}$ is diagonal.
519: 
520: Since the Schur complement contains the inverse of the fine-grid
521: submatrix $M_{22}$, algebraic multigrid must strive to find
522: decompositions that make $M_{22}$ as dominated by the diagonal as
523: possible so that a good approximation to its inverse can be found.
524: The method thus provides an immediate heuristic for choosing a
525: multigrid decomposition based only on the content of the matrix $M$
526: which is advantageous e.g.~for irregular discretizations.  In the case
527: of lattice gauge theories, the coefficients fluctuate stochastically
528: on the gauge group, but have the same magnitude. We thus choose a
529: regular decomposition where ${\cal L}_1 = (2\setZ)^D$, i.e.~the coarse
530: lattice is the lattice of points with all-even coordinates (cf.\ Fig.\ 
531: \ref{fig:9}; for a different choice see \cite{medeke-wup-1999}). Since
532: ${\cal L}_1$ is not connected with respect to $M$, $M_{11}$ is
533: diagonal, $M_{12}$ links points of ${\cal L}_1$ to the remaining
534: points, and $M_{22}$ forms a connected lattice on the fine grid.
535: 
536: \begin{figure}[htbp]
537: \begin{center}
538: \includegraphics[width=.5\hsize]{plots/basic-amg.eps}
539: \end{center}
540:     \caption{\label{fig:9}
541:       Two-grid decomposition of a two-dimensional lattice. Filled
542:       points are lattice sites of the coarse grid (${\cal L}_1$), open 
543:       points sites of the fine grid (${\cal L}_2$). Dashed lines
544:       are links that enter in $M_{12}$ and $M_{21}$, while solid lines 
545:       are those in $M_{22}$. The arrows show three possible
546:       contributions of order $2$,
547:       $4$, and $6$ to a general coarse-grid operator of the form (\ref{eq:19}).
548:     }
549: \end{figure}
550: 
551: \subsection{Block $LU$ factorization}
552: 
553: Eq.~(\ref{eq:4}) and (\ref{eq:5}) can be expressed in matrix form as
554: an {\em block $LU$ decomposition} of $M$. Its general form is
555: \begin{equation}
556:   \label{eq:40}
557:   M = 
558:   \left(
559:     \begin{array}{cc}
560:       1 & -Q^\conj \\ 0 & 1 
561:     \end{array}
562:     \right)
563:     \,
564:   \left(
565:     \begin{array}{cc}
566:       S & 0 \\ 0 & T
567:     \end{array}
568:     \right)
569:     \,
570:   \left(
571:     \begin{array}{cc}
572:       1 & 0 \\ -Q & 1
573:     \end{array}
574:     \right) \quad.
575: \end{equation}
576: The advantage of the representation (\ref{eq:40}) is
577: that its inverse can be readily written as
578: \begin{equation}
579:   \label{eq:44}
580:   M^{-1} = 
581:   \left(
582:     \begin{array}{cc}
583:       1 & 0 \\ Q & 1
584:     \end{array}
585:     \right) \,
586:   \left(
587:     \begin{array}{cc}
588:       S^{-1} & 0 \\ 0 & T^{-1}
589:     \end{array}
590:     \right)
591:     \,
592:   \left(
593:     \begin{array}{cc}
594:       1 & Q^\conj \\ 0 & 1 
595:     \end{array}
596:     \right)
597:     \,
598: \end{equation}
599: requiring only the inversion of the smaller matrices $S$ and $Q$. In
600: multigrid terminology, applying the factorization (\ref{eq:44}) to a
601: vector amounts to first applying an {\em restriction operator}
602: $Q^\conj$ that moves information from the fine grid to the
603: coarse grid, then solving the {\em effective coarse-grid operator} $S$
604: along with the residual fine-grid operator $T$, and finally
605: using the {\em interpolation operator} $Q$ to move information back
606: from the coarse grid to the full grid.
607: 
608: Inserting (\ref{eq:40}) into (\ref{eq:3}) leads to the following
609: relations that define $S$, $Q$, and $T$.
610: \begin{eqnarray}
611:   \label{eq:41}
612:   S f_1 &=& a_1 - Q^\conj \, a_2 \quad \\
613:   \label{eq:41b}
614:   T \left( -Q f_1 + f_2 \right) &=& a_2 \quad.
615: \end{eqnarray}
616: Comparing with eq.~(\ref{eq:4}) and (\ref{eq:5}) gives the solution
617: \begin{eqnarray}
618:   \label{eq:44b}
619:   S &=& S_M = M_{11} - M_{12} M_{22}^{-1} M_{21} \quad, \nnm \\
620:   Q &=& -M_{22}^{-1} M_{21} \quad,\nnm\\
621:   T &=& M_{22}  \quad.
622: \end{eqnarray}
623: 
624: The Schur complement is the exact effective coarse-grid operator
625: associated with a given decomposition of the lattice. Unlike the
626: geometrical approach, this decomposition is performed simply by
627: thinning the matrix and thus without a block averaging that might
628: introduce artifacts in the presence of a gauge field, and the
629: resulting coarsened operator is defined solely in terms of the
630: original matrix and thus the original disorder field. In the following
631: sections, we will show that this procedure can be interpreted in terms
632: of a noninteracting random walk leading to an explicit
633: representation of the effective matrix in terms
634: of a generalized path-dependent stencil.
635: 
636: \subsection{Renormalization group and projective multigrid}
637: The Schur complement has a close relation to the renormalization group
638: approach used in statistical mechanics, where the Klein-Gordon and
639: Dirac fields are described by a path integral with a Gaussian weight.
640: In this approach, the partition sum
641: \begin{eqnarray*}
642:   \label{eq:73}
643:   Z &=& \int {\rm d}f \, \exp\left[
644:     -\onehalf (f,Mf) + \onehalf (a,f) + {\rm c.c.} \right] \nnm \\
645:   &=& \const \cdot
646:     e^{{\scriptstyle\frac{1}{4}} (a, M^{-1} a)}
647: \end{eqnarray*}
648: (the integration is over all lattice field configurations $f$)
649: generates the inverse of $M$ by
650: taking the derivative with respect to the source field $a(x)$:
651: \begin{equation}
652:   \label{eq:74}
653:   \left.
654:   \frac{1}{Z} \, \frac{\partial^2}{\partial a^*(x) \, \partial a(y)}
655:   \right|_{a=0}
656:   Z = \onehalf M^{-1}(x,y) \quad.
657: \end{equation}
658: 
659: In renormalization-group inspired multigrid approach, such as proposed
660: by Mack \cite{mack-1988}, Brower et.~al.\ 
661: \cite{brower-edwards-rebbi-1991} or Vyas \cite{prd-vyas}, one
662: introduces new variables in the path integral by means of coarsening.
663: In our terminology, a coarsened lattice field $F_1$ defined on on
664: ${\cal L}_1$ is given by
665: \begin{equation}
666:   \label{eq:80}
667:   F_1 = f_1 + C f_2
668: \end{equation}
669: where the coarsening matrix $C$ performs an averaging over neighboring
670: sites in ${\cal L}_2$. (A more general coarsening would also include
671: an averaging over neighboring sites from ${\cal L}_1$).  The coarsened
672: field $F_1$ is interpolated back to the fine lattice using an
673: interpolation kernel $A$ and subtracted from $f_2$ to form the
674: residual fine-grid field:
675: \begin{equation}
676:   \label{eq:81}
677:   \zeta_2 = f_2 - A F_1
678: \end{equation}
679: In this way, it is hoped that some of the dynamics of the fine
680: lattice ${\cal L}_2$ is moved to the coarse lattice.  The whole
681: transformation reads
682: \begin{equation}
683:   \label{eq:82}
684:   \left(\begin{array}{c} F_1 \\ \zeta_2 \end{array}\right) = 
685:   \left(\begin{array}{cc} 1 & 0 \\ -A & 1 \end{array}\right) \,
686:   \left(\begin{array}{cc} 1 & C \\ 0 & 1 \end{array}\right) \,
687:   \left(\begin{array}{c} f_1 \\ f_2 \end{array}\right) \quad.  
688: \end{equation}
689: As its determinant is unity, $F_1$ and $\zeta_2$ can be used as new
690: integration variables in the path integral. The quadratic form in the
691: exponential is then
692: \begin{eqnarray}
693:   \label{eq:83}
694:   \left( f,M f \right) &=& 
695:   \left(\begin{array}{c} F_1 \\ \zeta_2 \end{array} \right)^\conj
696:   \left(\begin{array}{cc} 1 & A^\conj \\ 0 & 1 \end{array}\right) \,
697:   \tilde M \,
698:   \left(\begin{array}{cc} 1 & 0 \\ A & 1 \end{array}\right) 
699:   \left(\begin{array}{c} F_1 \\ \zeta_2 \end{array} \right)^\conj 
700:   \\
701:   \label{eq:83b}
702:   \tilde M &=& 
703:   \left(\begin{array}{cc} 1 & 0 \\ -C^\conj & 1 \end{array}\right) \,
704:   M \,
705:   \left(\begin{array}{cc} 1 & -C \\ 0 & 1 \end{array}\right)
706: %  \nnm\\
707: %  &=&
708: %  \left(\begin{array}{cc} M_{11} & 
709: %                          M_{12} - M_{11} C \\ 
710: %                          M_{21} - C^\conj M_{11} & 
711: %                          M_{22} - M_{21} C_{12} - C^\conj M_{12} 
712: %                          + C^\conj M_{11} C
713: %  \end{array}\right) 
714:   \quad,
715: \end{eqnarray}
716: and the coupling to the external fields
717: \begin{eqnarray}
718:   \label{eq:84}
719:   (a,f) = 
720:   \left(\begin{array}{c} a_1 + A^\conj a_2 - A^\conj C^\conj a_1 \\ 
721:                          a_2 - C^\conj a_1 \end{array} \right)^\conj \,
722:   \left(\begin{array}{c} F_1 \\ \zeta_2 \end{array} \right)^\conj  \quad.
723: \end{eqnarray}
724: 
725: Algebraic multigrid does not use coarsening; the coarse lattice field
726: is obtained by thinning out the original lattice and thus $F_1=f_1$,
727: $C=0$ and $\tilde M = M$.  Then the block $LU$ decomposition (\ref{eq:40})
728: tells us that we can make the quadratic form (\ref{eq:83}) block
729: diagonal by choosing $A = -M_{22}^{-1} M_{21}$, in which case the partition
730: sum factorizes into a coarse-grid and a fine-grid integral:
731: \begin{eqnarray}
732:   \label{eq:75}
733:   Z &=& \int {\rm d}f_1 \, 
734:         e^{-\onehalf (f_1,S_M f_1) + \onehalf (a_1 + A^\conj a_2,f_1) + {\rm c.c.}} \cdot
735:         \int {\rm d}f_2 \,
736:         e^{-\onehalf (f_2,M_{22} f_2) + 
737:             \onehalf (a_2,f_2) + {\rm c.c.}} \quad.        
738: \end{eqnarray}
739: The coarse scale dynamics is obtained by taking derivatives of $Z$
740: with respect to $a_1$ at $a_2=0$ and is therefore completely contained
741: in the first integral, which is governed by $S_M$, the Schur
742: complement.  Thus the algebraic multigrid choice of the interpolation
743: kernel $A$ can be characterized as that choice that leads to a
744: complete decoupling of fine and coarse degrees of freedom, which makes
745: integrating out the fine-grid degrees of freedom $\zeta_2$ trivial.
746: Note that the Schur complement can be used to obtain effective
747: fermions such as used in the blocked fermion approach to full QCD
748: \cite{prd-lippert}.
749: 
750: If a nonvanishing averaging kernel $C$ is used, one can achieve
751: complete decoupling in the quadratic form by applying the Schur
752: complement to the matrix $\tilde M$, but now $a_1$ also couples to
753: $\zeta_2$ in the source term (\ref{eq:84}), and the integral over
754: $\zeta_2$ results in a nontrivial contribution to the partition sum.
755: This contribution is governed by $\tilde M_{22}^{-1}$, the
756: propagator of the residual modes $\zeta_2$ on the fine lattice, which
757: is given by
758: \begin{equation}
759:   \tilde M_{22} = M_{22} - M_{21} C - C^\conj M_{12}
760:                  +C^\conj M_{11} C   \quad.
761: \end{equation}
762: One therefore strives to choose an averaging kernel $C$ such that
763: $\tilde M_{22}^{-1}$ is as local as possible, indicating that most of
764: the (long-range) dynamics has been moved to the coarse-lattice field
765: $F_1$. 
766: 
767: The decoupling of coarse and fine modes was at the heart of Mack's
768: original multigrid proposal \cite{mack-1988}, which was inspired by
769: constructive quantum field theory.  It was already noticed by Brower,
770: Rebbi, and Vicari \cite{prd-43-1965} that minimizing the coupling
771: between coarse and fine modes amounts in spirit to algebraic
772: multigrid, and using a similar reasoning, the form of the Schur
773: complement is mentioned in eq.~(12) of
774: \cite{brower-edwards-rebbi-1991} by Brower, Edwards, and Rebbi. In
775: practice, in the projective multigrid method the coarsening $C$ was
776: chosen to localize $\tilde M_{22}^{-1}$ as much as possible, and the
777: interpolation $A$ to maximize the decoupling between coarse and fine
778: modes.  An explicit method for calculating $A$ for a given $C$ is
779: specified as idealized multigrid algorithm by Kalkreuter
780: \cite{heplat-9304004,heplat-9408013}, and it can be verified that for
781: a ``lazy'' coarsening kernel $C=0$, the resulting interpolation kernel
782: has the form (\ref{eq:44b}). The actual calculation for nontrivial $C$
783: is numerically infeasible as it requires the solution of a nonlinear
784: equation at each coarse lattice site, so instead the ground-state of a
785: block-truncated matrix was used for the kernels, leading to the method
786: known as ground-state projection multigrid. The effective coarse-grid
787: matrix was then approximated by a Galerkin choice using the rows of
788: $A$ as a basis.
789: 
790: Using the algebraic multigrid choice of $C=0$, we cannot optimize the
791: decay of $M_{22}^{-1}$. However, as there is no induced coupling of
792: $a_1$ to the fine lattice field $\zeta_2$, the Schur complement
793: provides an explicit representation of the complete coarse-lattice
794: dynamics. In the following, we will discuss the interpretation of this
795: quantity based on the diffusion representation.
796: 
797: \subsection{Diffusion on a multigrid}
798: \label{sec:diffmultigrid}
799: 
800: In sec.~\ref{sec:diffusion}, the diffusion representation of the
801: inverse of a matrix $M$ was introduced. The matrix elements of the
802: Schur complement $S_M$ define a similar effective diffusion process
803: involving only coarse lattice sites with transition probabilities and
804: gauge phases determined by the elements of $S_M$.  As $S_M^{-1}$ is
805: identical with the upper left submatrix of $M^{-1}$,
806: \begin{equation}
807:   \label{eq:96}
808:   S_M^{-1}(x,y) = M^{-1}(x,y) \qquad\mbox{for }x,y \in {\cal L}_1
809:   \quad,
810: \end{equation}
811: this diffusion process is equivalent to the original diffusion process
812: when origin and destination of the random walk are restricted to the
813: coarse lattice.  Using (\ref{eq:6}), the matrix elements of $S_M$ can
814: itself be written in a diffusion representation:
815: \begin{eqnarray}
816:   \label{eq:39}
817:   S_M(x,y) &=&
818:   \delta_{x,y} - \sum_n \sum_{\begin{array}{c}
819:       (x,x_1,\ldots,x_n,y) \\
820:       x_1,\ldots,x_n \in {\cal L}_2 \,
821:     \end{array}}
822:   \kappa^n \, {\cal P}(U;x,x_1,\ldots,x_n,y) \quad.
823: \end{eqnarray}
824: Since the sum comes from expanding $M_{22}^{-1}$, the inner parts
825: $(x_1,\ldots,x_n)$ of the paths run exclusively on the fine lattice
826: ${\cal L}_2$; they are connected to the coarse lattice ${\cal L}_1$ at
827: points $x$ and $y$ through the matrix elements of $M_{12}$ and
828: $M_{21}$ (some of these paths are shown in Fig.~\ref{fig:9}).  The matrix
829: elements of the Schur complement can therefore be interpreted as
830: resulting from random
831: walks on the fine complement lattice ${\cal L}_2$, and when the Schur
832: complement is inverted, it defines a random walk on the
833: coarse lattice ${\cal L}_1$, in each of whose steps all possible paths
834: on the fine lattice ${\cal L}_2$ are effectively taken into account.
835: In this way, the Schur complement allows to ``integrate out'' the
836: fine degrees of freedom in the diffusion process much as
837: renormalization group transformations do.
838: 
839: \begin{figure}[htbp]
840: \begin{center}
841: \includegraphics[width=.5\hsize]{plots/oe-amg.eps}
842: \end{center}
843:     \caption{\label{fig:10}
844:       Two-grid decomposition of a odd-even decomposed lattice. As in
845:       Fig.~\ref{fig:9}, filled circles mark sites of the coarse grid,
846:       and open circles those of the fine grid. 
847:       Full lines show
848:       links from $M_{22}$, dashed lines those from $M_{11}$, and
849:       dotted lines those from $M_{12}$. $M_{11}$ and $M_{22}$ here
850:       form two similar grids, with $M_{12}$ connecting the two along
851:       the diagonals.
852:     }
853: \end{figure}
854: 
855: In the two-dimensional case, the situation can be made more
856: symmetrical by performing and even-odd-decomposition before the
857: multigrid decomposition. 
858: %For the even-odd-decomposition, 
859: %the vectors
860: %are reordered such that matrix decomposes as
861: %\begin{equation}
862: %  \label{eq:56}
863: %  \left( \begin{array}{cc} 1 & M_{eo} \\ M_{oe} & 1 \end{array} \right)
864: %  \quad.
865: %\end{equation}
866: %Then the odd sites can be eliminated completely, and the effective
867: %operator on the even sites is given by the Schur complement
868: In an even-odd decomposition, the matrix is replaced by the effective
869: matrix on the even sites of the lattice,
870: \begin{equation}
871:   \label{eq:57}
872:   S_{ee} = 1 - M_{eo} M_{oe} \quad,
873: \end{equation}
874: where $M_{eo}$ and $M_{oe}$ are the
875: matrix elements of $M$ between even and odd sites, resp.  In the resulting
876: lattice topology, each site is linked both to its four next neighbors
877: in the straight direction as well to the four next neighbors in the
878: diagonal (cf.\ Fig.~\ref{fig:10}). Applying now the multigrid
879: decomposition, the lattice decomposes into two similar square
880: lattices. $M_{11}$ and $M_{22}$ are represented by the straight
881: links on each lattice, while $M_{12}$ and $M_{21}$ are carried by the
882: diagonals linking the two lattices. 
883: The effective hopping parameter on each lattice is
884: $\kappa_{\rm eff} = \frac{\kappa^2}{1-4 \kappa^2}$ and thus
885: %for $\kappa < (\sqrt{4D^2+1} - 1)/4D$, 
886: %the effective hopping parameter is 
887: smaller than on the original lattice, and consequently $M_{11}$ and
888: $M_{22}$ are less critical than $M$. In the diffusion interpretation,
889: even if probability was conserved on the original lattice, it is not
890: conserved on the two sublattices individually, as the random walkers
891: can move from one lattice to the other along the diagonal links
892: represented by $M_{12}$ and $M_{21}$.  The criticality of the system
893: is only regained when $M_{11}$ and $M_{22}$ are combined in the Schur
894: complement.
895: 
896: The propagator on the fine lattice $M_{22}^{-1}$ will therefore
897: be short-ranged and can be incorporated approximately by considering
898: only a limited number of hops on the fine lattice in the
899: expansion (\ref{eq:39}). In the following section, we exploit this to
900: construct an approximate Schur complement numerically.
901: 
902: \section{Construction of the multigrid operator}
903: 
904: The central problem in deriving a multigrid algorithm for disordered
905: systems is the construction of a suitable approximation to the Schur
906: complement $S_M$ and the block $LU$ decomposition
907: (\ref{eq:40}). 
908: 
909: %As the disorder is stochastic, this construction must be
910: %performed many times in a typical application. The disorder breaks
911: %translation invariance which should be restored by the stochastic
912: %average. For gauge disorders, the fields obey gauge invariance, and
913: %the approximation should respect this symmetry.
914: 
915: \subsection{Numerically optimized Schur complement}
916: 
917: Our method is based on numerically approximating the Schur complement
918: as a linear combination of suitable basis operators.  
919: The full Schur complement $S$ usually cannot be computed exactly, 
920: but it can be characterized numerically by a sufficiently large 
921: set of pairs $\{ f^{(n)}, a^{(n)}; n=1,\ldots,N \}$ which satisfy
922: \begin{equation}
923:   \label{eq:17}
924:   M(U^{(n)}) f^{(n)} = a^{(n)} \quad, 
925: \end{equation}
926: where it is explicitly indicated that $M$ depends on different
927: realizations $U^{(n)}$ of the disorder field. 
928: Using the block $LU$ decomposition (\ref{eq:40}), this implies that 
929: $S$, $Q$, and $T$ satisfy the relations (\ref{eq:41}), (\ref{eq:41b}). 
930: In particular, if
931: the pairs are chosen such that $a_2=0$, 
932: their coarse-lattice components $\{ f_1^{(n)}, a_1^{(n)} \}$
933: characterize the Schur complement by the relation:
934: \begin{equation}
935:   \label{eq:42}
936:   S f_1^{(n)} = a_1^{(n)} \quad.
937: \end{equation}
938: Similarly, the interpolation operator is characterized by the pairs $\{ 
939: f_1^{(n)}, f_2^{(n)} \}$ using
940: \begin{equation}
941:   \label{eq:43}
942:   Q f_1^{(n)} = f_2^{(n)} \quad.
943: \end{equation}
944: There is no need to approximate $T = M_{22}$, as it does not contain the
945: inverse of $M_{22}$.
946: An approximation $\bar S$ to $S$, and similarly $\bar Q$ to Q, can then be
947: characterized with minimum bias by the mean error norm of (\ref{eq:42}):
948: \begin{equation}
949:   \label{eq:16}
950:   \delta^2 = \sum_n^N \left| \bar S(U^{(n)}) f_1^{(n)} - a_1^{(n)} \right|^2
951: \end{equation}
952: This quantity is an approximation to an operator norm
953: $|\bar S S - 1|^2$ averaged over gauge field configurations.  
954: The choice of pairs (\ref{eq:17}) introduces a
955: weight function in this norm and thus determines which functions are
956: well approximated by $\bar S$. Eq.~(\ref{eq:16}) can be rewritten as
957: \begin{equation}
958:   \label{eq:25}
959:   \delta^2 = \sum_n^N \left | \left( \bar S(U^{(n)}) S - 1 \right)
960:     a_1^{(n)} \right|^2
961:   = \sum_n^N \left | \left( \bar S(U^{(n)} - S(U^{(n)} \right)
962:     f_1^{(n)} \right|^2
963: \end{equation}
964: If $f_1$ is chosen randomly on the whole function space, $\delta^2$
965: thus approximates $|\bar S - S|^2$, while if $a_1$ is chosen randomly, it
966: approximates
967: \begin{equation}
968:   \label{eq:26}
969:   |\bar S S^{-1} -1|^2 = |(\bar S - S) S^{-1}|^2
970: \end{equation}
971: The factor $S^{-1}$ increases the weight of functions that have
972: eigenvalues closer to zero, and these are exactly the long-ranged
973: functions we are interested in. Choosing the right-hand sides $a_1$
974: also allows us to enforce $a_2=0$ and thus remove $Q$ from the
975: (\ref{eq:41}). In practice, we choose $a$ to be a delta function on a
976: randomly chosen coarse-lattice point, and determine $f$ by solving the
977: linear system (\ref{eq:17}). In other words, the $f^{(n)}$ are Green's
978: functions, or propagators in lattice gauge theory language, and we
979: look for an approximate operator $\bar S$ that well reproduces the
980: action of $S$ on these functions. 
981: 
982: If we choose a general linear combination of suitably chosen basis
983: operators $S^{(i)}(U)$ respecting the symmetries of the problem as
984: approximate operators,
985: \begin{equation}
986:   \label{eq:17a}
987:   \bar S(U) = \sum_i \alpha_i \bar S^{(i)}(U) \quad,
988: \end{equation}
989: the approximation error $\delta^2$ is a 
990: quadratic form in the coefficients $\alpha_i$:
991: \begin{eqnarray}
992:   \label{eq:18}
993:   \delta^2 &=&
994:   \sum_{ij} \alpha_i^\ast \, A_{ij} \, \alpha_j
995:        - \sum_i \alpha_i \, B_i^*
996:      - {\rm c.c.} + C \\
997:   \label{eq:18a}
998:   A_{ij} &=& \sum_n^N \, \left( S^{(i)}(U^{(n)}) f_1^{(n)}, 
999:                      S^{(j)}(U^{(n)}) f_1^{(n)} \right)_{{\cal L}_1}
1000:                    \\
1001:   \label{eq:18b}
1002:   B_i &=& \sum_n^N \, \left (S^{(i)}(U^{(n)}) f_1^{(n)}, a_1^{(n)}\right)_{{\cal L}_1} \\
1003:   C &=& \sum_n^N \,\left (a_1^{(n)},a_1^{(n)} \right)_{{\cal L}_1} 
1004: \end{eqnarray}
1005: where $(a_1,b_1)_{{\cal L}_1}$ 
1006: denotes that scalar product on the coarse grid ${\cal
1007: L}_1$. Thus the minimum of the approximation error and thus the 
1008: optimal choice of coefficients is found simply by:
1009: \begin{equation}
1010:   \label{eq:27}
1011:   \alpha_i = \sum_j (A^{-1})_{ij} B_j
1012: \end{equation}
1013: The procedure for finding the coefficients of the numerical
1014: approximation to the Schur complement is therefore as follows:
1015: \begin{enumerate}
1016: \item Generate right-hand sides $a^{(n)}$ by choosing random delta functions
1017:   on the coarse grid.
1018: \item Find the corresponding $f^{(n)}$ by solving (\ref{eq:17}) using
1019:   the original fine-grid operator $M$.
1020: \item Project $a^{(n)}$ and $f^{(n)}$ to the coarse lattice.
1021: \item Calculate the matrix elements (\ref{eq:18a}) and (\ref{eq:18b})
1022:   and find $\alpha_i$ from (\ref{eq:27}).
1023: \end{enumerate}
1024: A similar procedure is used to approximate the interpolation operator
1025: $Q$. This method is similar to Monte Carlo renormalization group as it
1026: calculates the coefficients of the effective coarse-grid operator by
1027: applying a block-spin transformation (in our case, simply a
1028: projection) to an ensemble of configurations on the fine grid.  Note
1029: that the procedure is completely general and can be used to
1030: approximate any operator, e.g.~the Neuberger or a
1031: renormalization-group improved operator, if its action on some sample
1032: fields is known.
1033: 
1034: %In particular, $M(U)$ satisfies gauge
1035: %invariance: Let $\Lambda(x)$ a field defined on $x\in{\cal L}$ with
1036: %values in the gauge group. This field acts on the gauge field $U$ by
1037: %\begin{equation}
1038: %  \label{eq:19}
1039: %  U^\Lambda_\mu(x) = \Lambda(x) U_\mu(x) \Lambda(x+e_\mu)^{-1} \quad.
1040: %\end{equation}
1041: 
1042: \subsection{Choice of the operator basis}
1043: The approximate coarse-grid operator $\bar S(U)$ is a general
1044: operator-valued function of the disorder field $U$, but the symmetries
1045: of the problem greatly restrict the possible basis operators $\bar
1046: S^{(i)}(U)$ from which it can be constructed.
1047: %A gauge transform specified by an arbitrary field $\Lambda(x)$ in the
1048: %gauge group acts in the following way on fields and operators:
1049: %\begin{eqnarray}
1050: %  \label{eq:37}
1051: %  {}^\Lambda f(x) &=& \Lambda(x) f(x) \\
1052: %  \label{eq:37b}
1053: %  {}^\Lambda U_\mu(x) &=& \Lambda(x) U_\mu(x) \Lambda(x+e_\mu)^{-1} \\
1054: %  \label{eq:37c}
1055: %  {}^\Lambda M(x,y) &=& \Lambda(x) M(x,y) \Lambda(y)^{-1}
1056: %\end{eqnarray}
1057: %A gauge covariant operator, such as the original fine-grid operator,
1058: %satisfies the relation
1059: %\begin{equation}
1060: %  \label{eq:32}
1061: %  {}^\Lambda M = M({}^\Lambda U) \quad,
1062: %\end{equation}
1063: %and so must the basis operators $S^{(i)}(U)$.
1064: %From (\ref{eq:37b}), only the product of links along a closed path
1065: %satisfies this relation:
1066: %\begin{eqnarray}
1067: %  \label{eq:85} &&
1068: %  {}^\Lambda U_{\mu_1}(x_1)
1069: %  {}^\Lambda U_{\mu_2}(x_1 + e_{\mu_1}) \ldots
1070: %  {}^\Lambda U_{\mu_n}(y = x_1 + e_{\mu_1} + \cdots + e_{\mu_{n-1}})
1071: %  = \nnm \\ && \kern20pt
1072: %  \Lambda(x) 
1073: %  U_{\mu_1}(x_1)
1074: %  U_{\mu_2}(x_1 + e_{\mu_1}) \ldots
1075: %  U_{\mu_n}(x_1 + e_{\mu_1} + \cdots + e_{\mu_{n-1}})
1076: %  \Lambda(y)^{-1}
1077: %\end{eqnarray}
1078: To maintain gauge covariance, each contribution to the matrix element
1079: $M(x,y)$ can be factored into a gauge-field independent part and a
1080: product of the gauge field links along a connected path from $x$ to
1081: $y$, leading to the following representation of $M$:
1082: \begin{equation}
1083:   \label{eq:45}
1084:   M(x,y) = \sum_n \sum_{P = (\mu_1,\ldots,\mu_n)} \,
1085:     m(P) \,
1086:     \delta_{y,x + e_{\mu_1} + \cdots + e_{\mu_n}} \, 
1087:     {\cal P}_{\mu_1,\ldots,\mu_n}(U;x)
1088: \end{equation}
1089: where the sum is over connected paths from $x$ to $y$, and the reduced
1090: matrix element $m(P)$ does not depend on the gauge field $U$. 
1091: Basis operators $S^{(i)}$ can thus be associated with the possible
1092: relative paths $P_i = \{ \mu_1, \ldots, \mu_n \}$:
1093: \begin{equation}
1094:   \label{eq:19}
1095:   \bar S^{(i)}(U)(x,y) = \Gamma^{(i)}
1096:   \delta_{y,x + \mu_1 + \cdots + \mu_n} \, 
1097:   {\cal P}_{\mu_1,\ldots,\mu_n}(U;x) \quad,
1098: \end{equation}
1099: where $\Gamma^{(i)}$ is independent of $x$, $y$, and $U$, and serves to
1100: contain a possible internal structure of the field, e.g.~the Dirac
1101: $\gamma$ matrices.
1102: 
1103: Other symmetries like translation, rotation, and reflection invariance
1104: as well as hermiticity and charge conjugation generate further
1105: restrictions on the coefficients $\alpha_i$.  In the calculations, our
1106: computer code automatically evaluates discrete rotation and reflection
1107: symmetry for all paths and constructs basis operators that are
1108: invariant under these transformations. This reduces significantly the
1109: number of degrees of freedom in (\ref{eq:18}). Hermiticity places
1110: constraints on the real and imaginary parts of the coefficients; it is
1111: not enforced and can be used as a check of the accuracy of the
1112: constructed operator. The general problem of constructing generalized
1113: Dirac operators was recently brought up in the context of
1114: chirally-enhanced operators in \cite{heplat-0001001,heplat-0003005}
1115: and discussed in general in \cite{heplat-0003013}.
1116: 
1117: The representation (\ref{eq:45}) has the same form as the
1118: diffusion representation (\ref{eq:39}), with the additional
1119: restriction that only paths moving exclusively on the fine lattice
1120: contribute there.  The optimization process (\ref{eq:16}) thus results
1121: in assigning weights to the different paths in the diffusion
1122: representation that reflect how much a path contributes, on average,
1123: to the propagator. If a complete sum over all lengths $n$ is taken,
1124: the correct weights are of course $\kappa^{-n}$, as in the von Neumann
1125: series. However, if the sum is truncated, the weights are modified in
1126: the optimization process and the contributions of the individual paths
1127: renormalized to take the obmitted paths into account.
1128: 
1129: The basis (\ref{eq:19}) can be compared to the situation without a
1130: gauge field.  In the terminology of numerical analysis, a
1131: translation-invariant operator is represented by a {\em stencil}
1132: $S(a)$ using the expression corresponding to (\ref{eq:45})
1133: \begin{equation}
1134:   \label{eq:54}
1135:   M(x,y) = \sum_a S(a) \, \delta_{y,x+a}
1136: \end{equation}
1137: where $a$ runs over the possible offsets between lattice points. The
1138: basis operators in this case can therefore be labeled by the offset $a$,
1139: and the equation corresponding to (\ref{eq:19}) reads
1140: \begin{equation}
1141:   \label{eq:55}
1142:   S^{(i)}(x,y) = \delta_{y,x+a_i} \quad.
1143: \end{equation}
1144: Comparing the two cases, one can see that in the disordered case the
1145: relative paths take over the role of the offsets, and a generalized
1146: stencil is given by assigning each relative path a certain weight. The
1147: optimization procedure corresponds to calculating such a generalized
1148: stencil, much as an ordinary stencil is constructed for differential
1149: equations.
1150: 
1151: Once the disordered stencil is known, the coarse-grid operator $S$ for
1152: any gauge field can be constructed simply by evaluating all paths in
1153: the set $\{S^{(i)}\}$ and combining them according to the weights
1154: $\alpha_i$.  Note that it is not necessary to perform the construction
1155: of the on the actual set of gauge fields, but on a representative
1156: ensemble only, so the optimization can be performed in advance of the
1157: simulation.
1158: 
1159: The resulting coarse-grid operator contains in particular also
1160: contributions beyond direct neighbors on the coarse grid, similar to
1161: improved and perfect operators that lose locality but gain a more
1162: faithful representation of the continuum operator, and the amount of
1163: nonlocality can be tuned by selecting the paths in the basis set. On
1164: the other hand, if the basis set is chosen to reproduce the original
1165: topology on the coarse lattice, the coarse-grid operator defines a new
1166: effective disorder field on the coarse-grid. However, this disorder
1167: field is obtained by a sum and will therefore lie outside the gauge
1168: group, though it can be projected back to the gauge group, as is
1169: customarily done in many renormalization schemes. In this way, the
1170: Schur complement can be used to define a block-spin transformation for 
1171: the gauge field.
1172: 
1173: \subsection{Construction in the diagonal subspace}
1174: The basis (\ref{eq:19}) is the most general set of operators that can
1175: make up a gauge-covariant operator. Its size grows exponentially with
1176: the order of the path set and thus makes both the evaluation of the
1177: coefficients (\ref{eq:18a}) and (\ref{eq:18b}) and the actual
1178: construction of the approximate Schur complement very time consuming.
1179: A smaller and thus more tractable basis can be found by considering
1180: the operators that appear in the von Neumann series (\ref{eq:31})
1181: which are:
1182: \begin{eqnarray}
1183:   \label{eq:20}
1184:   \bar S^{(0)} &=& M_{11} \quad, \nnm\\
1185:   \bar S^{(i+1)} &=& M_{12} M_{22}^{2i} M_{21} \quad. 
1186: \end{eqnarray}
1187: This basis generates exactly the same paths as the diffusion
1188: representation (\ref{eq:19}) restricted to the fine lattice, but uses
1189: coefficients that only depend on the length of the paths.  Obviously,
1190: if an infinite set of basis operators is used, the optimal
1191: coefficients are $\pm 1$ as given by the von Neumann series. However,
1192: in a finite set, the optimization procedure amounts to finding a
1193: polynomial approximation to the inverse function.
1194: 
1195: To see this, let $T$ be an approximation to $M_{22}^{-1}$, and write
1196: the approximate Schur complement as
1197: \begin{equation}
1198:   \label{eq:22}
1199:   \bar S = M_{11} - M_{12} T M_{21}
1200: \end{equation}
1201: Inserting this into (\ref{eq:16}) gives after some algebra
1202: \begin{equation}
1203:   \label{eq:23}
1204:   \left| \bar S f_1 - a_1 \right|^2
1205:   = \left( T \hat f, M_{21} M_{12} T \hat f \right)_2
1206:     - 2 \Re \left( \hat b, T \hat f \right)_2 + \left(b, b\right)_1
1207: \end{equation}
1208: with the notation $(\ldots,\ldots)_2$  for the scalar product over
1209: ${\cal L}_2$ and the definitions
1210: \begin{eqnarray}
1211:   \label{eq:24}
1212:   \hat f &=& M_{21} f_1 \quad,\\
1213:    b &=& M_{11} f_1 - a_1 \quad, \nnm\\
1214:   \hat b &=& M_{21} b \nnm\\
1215:   \label{eq:24b}
1216:   &=& M_{21} M_{12} M_{22}^{-1} M_{21} f_1 \quad,
1217: \end{eqnarray}
1218: where we used (\ref{eq:42}) and (\ref{eq:6}) in the last line.
1219: If $T$ is a polynomial in $M_{22}$, it is diagonal in the eigenbasis
1220: of $M_{22}$ and can be written as
1221: \begin{eqnarray}
1222:   \label{eq:22b}
1223:   M_{22} &=& \sum_k \lambda_k \, v_k \otimes v_k^\conj \\
1224:   T &=& \sum_k T_k \, v_k \otimes v_k^\conj
1225: \end{eqnarray}
1226: where $\lambda_k$, $v_k$ are eigenvalues and -vectors of $M_{22}$, and 
1227: $T_k$ the diagonal elements of $T$ in the eigenrepresentation.
1228: Inserting this into (\ref{eq:23}) gives the following
1229: quadratic form for the
1230: diagonal components $T_k$ of the operator $T$:
1231: \begin{eqnarray}
1232:   \label{eq:46}
1233:   \left| \bar S f_1 - a_1 \right|^2 &=&
1234:   \sum_{kl} T_k^\star \,
1235:             ( \hat f^{\star}_k \hat f_l )
1236:   \, \left( M_{21} M_{12} \right)_{kl} T_l
1237:   \\ && - 2 \Re 
1238:   \sum_k T_k \, 
1239:          ( \hat b^{\star}_k \hat f_k )
1240:   + \const 
1241: \end{eqnarray}
1242: where $\hat f_k = (v_k, \hat f)$, $\hat b_k = (v_k, \hat b)$ are the
1243: representations of $\hat f$ and $\hat b$ in the eigenspace of
1244: $M_{22}$, and similarly for $(M_{21} M_{12})_{kl}$. Eq.~(\ref{eq:24b})
1245: relates these quantities by
1246: \begin{equation}
1247:   \label{eq:51}
1248:   \hat b_k = \sum_l (M_{21} M_{12})_{kl} \, \lambda_l^{-1} \, 
1249:              \hat f_l
1250: \end{equation}
1251: which gives the following expression for (\ref{eq:46}):
1252: \begin{eqnarray}
1253:   \label{eq:52}
1254:   \left| \bar S f_1 - a_1 \right|^2 &=&
1255:   \sum_{kl} \left( T_k - \frac{1}{\lambda_k} \right)^\star \, 
1256:   ( \hat f^{\star}_k \hat f_l ) \,
1257:   \left( M_{21} M_{12} \right)_{kl} \,
1258:   \left( T_l - \frac{1}{\lambda_l} \right) + \const \nnm\\
1259:   &=& 
1260:   \sum_k
1261:   \left|
1262:     \sum_l
1263:     \left(M_{12}\right)_{kl} \hat f_l \,
1264:     \left( T_l - \frac{1}{\lambda_l} \right) 
1265:   \right|^2 + \const \quad.
1266: \end{eqnarray}
1267: Minimizing this quantity thus amounts to finding 
1268: a best approximation $T_k$ to
1269: $\lambda_k^{-1}$ in a weighted square norm specified by $M_{12} \hat f
1270: = M_{12} M_{21} f_1$. 
1271: 
1272: Choosing for $T$ a polynomial form
1273: \begin{equation}
1274:   \label{eq:53}
1275:   T = \sum_{n=0}^N \alpha_{n+1} \, M_{22}^{2n} \qquad,\quad
1276:   T_k = \sum_{n=0}^N \alpha_{n+1} \, \lambda_k^{2n}
1277: \end{equation}
1278: in accordance with the expansion (\ref{eq:20}),
1279: the approximate Schur complement then takes the form
1280: \begin{equation}
1281:   \label{eq:53b}
1282:   \bar S = M_{11} 
1283:          - \sum_{k=0}^N \alpha_{k+1} M_{12} M_{22}^{2k} M_{21}
1284: \end{equation}
1285: Fig.~\ref{fig:1} shows a numerical evaluation of such an approximation
1286: using the Dirac operator on a sample $U(1)$ configuration.
1287: 
1288: An operator of the form (\ref{eq:53b}) can be computed relatively
1289: easily: The coefficient $\bar S(x,y)$ between two sites is given by
1290: summing the phase factors over all paths from $x$ to $y$ on ${\cal
1291:   L}_2$
1292: with weights
1293: depending only on the length of the paths and given by the
1294: coefficients $\alpha_k$.  This is a simplification over the full path
1295: set (\ref{eq:19}), where each path can be assigned its own weight.  If
1296: only a finite number $N$ of coefficients is used, the operator is
1297: local, but contains more than next-neighbor interactions. Note that in
1298: a typical iteration scheme the calculation of the matrix elements of
1299: $\bar S$ has to be done only once, and the complexity of the iteration
1300: depends only on the sparsity of the resulting $\bar S$.
1301: 
1302: \begin{figure}[htbp]
1303: \begin{center}
1304: \includegraphics[width=.8\hsize]{plots/approximating-M22-3.eps}
1305: \end{center}
1306:     \caption{\label{fig:1}
1307:     Numerical evaluation of the optimal polynomial 
1308:     approximation to $M_{22}^{-1}$. The curves show the polynomial
1309:     approximations of order 1, 3, 5, and 7 to the inverse
1310:     function. The actual eigenvalues are marked with crosses.}
1311: \end{figure}
1312: 
1313: \begin{figure}[p]
1314:   \begin{center}
1315:     \centerline{\hss 
1316:       \includegraphics[width=3.8in]{plots/fine-2d-2.eps} \kern-60pt
1317:       \includegraphics[width=3.8in]{plots/coarse-2d-2.eps} \hss}
1318:     \caption{%
1319:       Absolute value of a sample fermionic Green's function on an
1320:       $16\times 16$ lattice in a $U(1)$ background field for the
1321:       original Dirac operator (left) and the approximate coarse-grid
1322:       operator obtained from the Schur complement. The source is
1323:       located at the origin; since the lattice is periodic, it shows
1324:       up at the four corners of the plot.
1325:       }
1326:     \label{fig:12}
1327:   \end{center}
1328: %\end{figure}
1329: 
1330: %\begin{figure}[h]
1331:   \begin{center}
1332:     \includegraphics[width=.8\hsize]{plots/greens-function-2.eps}
1333:     \caption{%
1334:       Cut through the example fermionic Green's function (black
1335:       squares) compared to Green's functions calculated from an
1336:       approximate Schur complement and interpolation operator 
1337:       of order 1, 2, and 4.
1338:       }
1339:     \label{fig:13}
1340:   \end{center}
1341: \end{figure}
1342: 
1343: \subsection{Numerical results}
1344: 
1345: We have numerically evaluated the optimal operator both for the full
1346: path set of eq.~(\ref{eq:19}) and the subset defined in
1347: eq.~(\ref{eq:20}). The calculations were performed for the Dirac
1348: equation on a $16\times 16$ lattice in two dimensions in a $U(1)$
1349: gauge field generated at $\beta=3.0$, which results in relatively
1350: smooth configurations and long-ranged correlation
1351: functions. Fig.~\ref{fig:12} and \ref{fig:13} show sample Green's
1352: functions of the original Dirac matrix and the approximate Schur
1353: complement. The disorder from the gauge field shows up clearly in the
1354: Green's function, but the various approximations to the Schur
1355: complement reproduce the irregularities and the overall form of the
1356: function quite well.
1357: 
1358: % and found that the they indeed produce an
1359: %approximation of the true coarse-lattice operator with an error that
1360: %decreases exponentially with the order of the approximation.
1361: 
1362: \begin{figure}[p]
1363:   \begin{center}
1364:     \includegraphics[width=.8\hsize]{plots/error-by-order.eps}
1365:     \caption{%
1366:       The relative approximation error as a function of the order of the
1367:       approximation, calculated in an ensemble of 10 independent
1368:       $16\times16$ $U(1)$ gauge configurations at $\beta=3.0$ and
1369:       $\kappa=0.265$. The full line gives the result from the
1370:       optimization in the diagonal subspace, the dashed line from the
1371:       full optimization. The dotted line is the relative error of the
1372:       truncated von Neumann series of the same order.
1373:       }
1374:     \label{fig:3}
1375:   \end{center}
1376: %\end{figure}
1377: 
1378: %\begin{figure}[h]
1379:   \begin{center}
1380:     \includegraphics[width=.8\hsize]{plots/error-by-order-2.eps}
1381:     \caption{%
1382:       The relative inversion error as a function of the approximation
1383:       order for both diagonal and full optimization. For comparison, 
1384:       the original relative error from Fig.~\ref{fig:3} is also
1385:       shown.
1386:       }
1387:     \label{fig:3a}
1388:   \end{center}
1389: \end{figure}
1390: 
1391: This is quantified in fig.~\ref{fig:3}, which shows the average relative error
1392: \begin{equation}
1393:   \label{eq:70}
1394:   \sqrt{\frac{\delta^2}{\sum_n^N |a_1^{(n)}|^2}}
1395: \end{equation}
1396: evaluated in an ensemble of ten sample $U(1)$ gauge configurations
1397: generated at $\beta=3.0$ on a $16\times16$-lattice. For each
1398: configuration, five different Green's function at $\kappa=0.265$ on each
1399: configuration were calculated and the coefficients (\ref{eq:18a}) and
1400: (\ref{eq:18b}) that characterize the error ellipsoid were evaluated.
1401: For the diagonal optimization, the basis (\ref{eq:20})
1402: with maximum order $i$ from $1$ to $6$, corresponding to a maximum
1403: path length from $2$ to $12$, was used. For the full optimization, a complete
1404: path basis of paths up to length $2i$ constrained to the discrete
1405: rotation and reflection symmetry was taken. For comparison, the
1406: relative error resulting from the von Neumann series is also shown.
1407: The errors show a very clean logarithmic dependence on the order. As
1408: the number of paths in the operator increases exponentially with
1409: order, this corresponds to a polynomial dependence of the error on the
1410: number of paths.
1411: 
1412: The full path set gives only a slightly better approximation than the
1413: diagonal approximation, though it is significantly more difficult to
1414: evaluate, as the number of coefficients grows exponentially. Its main
1415: advantage is that it allows to fine-tune the number and
1416: characteristics of the paths, e.g.~the maximum distance between end
1417: points.  Fig.~\ref{fig:3} shows the resulting errors when the error
1418: ellipsoid is restricted to a certain number of paths. It turns out
1419: that the relative error has a power law dependence on the number of
1420: paths, with an exponent of approximately $-0.55$ in this case. The
1421: contributions from different operators are relatively uniform except
1422: for the operators in the tails that contribute little or nothing, so
1423: the accuracy of the approximate Schur complement can be fine-tuned by
1424: varying the number of paths in the operator.
1425: 
1426: The relative error measures how well the approximate Schur complement
1427: reproduces a delta function when applied to a Green's function. For
1428: actual applications, the more relevant quantity is the relative
1429: inversion error that looks at how good the inverse of the
1430: approximate Schur complement reproduces the Green's function of the
1431: original matrix. This is quantified by
1432: \begin{equation}
1433:   \label{eq:71}
1434:   \delta_{\rm inv.}^2 = 
1435:   \sum_n^N \left| \bar f_1^{(n)} - S^{-1}(U^{(n)})  a_1^{(n)}
1436:   \right|^2
1437: \end{equation}
1438: with the relative error taken relative to the norm of $f_1^{(n)}$:
1439: %\begin{equation}
1440: %  \label{eq:71a}
1441: %  \sqrt{\frac{\delta^2_{\rm inv.}}{  \sum_n^N \left| \bar f_1^{(n)}
1442: %    \right|^2}}
1443: %  \quad.
1444: %\end{equation}
1445: This quantity was evaluated numerically by computing Green's function
1446: of the approximate Schur complement on the coarse lattice and
1447: comparing the results to the Green's functions of the original
1448: Wilson-Dirac matrix.  Fig.~\ref{fig:3a} shows the relative error as a
1449: function of the approximation order both for the diagonal and the full
1450: optimization. It also decreases exponentially but is about twice the
1451: magnitude of the original error. 
1452: %At larger orders, there is some
1453: %levelling off.
1454: 
1455: \subsection{Spectrum}
1456: 
1457: The low eigenvalues of the original matrix $M$ have a direct relation
1458: to those of the Schur complement $S_M$. If $f$ is an eigenvector of
1459: $M$ with eigenvalue $\lambda$,
1460: \begin{equation}
1461:   \label{eq:86}
1462:   M f = \lambda f \quad,
1463: \end{equation}
1464: this becomes using (\ref{eq:4})
1465: \begin{eqnarray}
1466:   \label{eq:87}
1467:   S f_1 &=& \lambda f_1 - M_{21} M_{22}^{-1} \, \lambda f_2 \nnm\\
1468:   M_{22} f_2 &=& \lambda f_2 - M_{21} f_1 \quad.
1469: \end{eqnarray}
1470: Eliminating $f_2$ in the first line yields a modified eigenvalue
1471: equation for $S$:
1472: \begin{equation}
1473:   \label{eq:88}
1474:   S f_1 = \lambda \, \left(
1475:           1 + M_{12} M_{22}^{-1} (M_{22} - \lambda)^{-1} M_{21}
1476:           \right) \, f_1 \quad.
1477: \end{equation}
1478: If the eigenvalues of $M_{22}$ are large, as argued in
1479: sec.~\ref{sec:diffmultigrid}, the additional term on the right-hand
1480: side multiplying the eigenvalue is approximately
1481: \begin{equation}
1482:   \label{eq:89}
1483:   1 + M_{12} M_{22}^{-2} M_{21} = 1 + Q^\conj Q \quad.
1484: \end{equation}
1485: Assuming further that $M_{22}^{-1}$ is 
1486: short-ranged, the second term can be approximated by a constant
1487: $\alpha$ when acting on a low eigenvector, and (\ref{eq:88}) becomes
1488: \begin{equation}
1489:   \label{eq:90}
1490:   S f_1 \approx \lambda (1+\alpha) f_1 \quad.
1491: \end{equation}
1492: This rescaling of the eigenvalues accounts for the different lattice
1493: constant of the coarse grid.  The appearance of the interpolation
1494: kernel $Q$ in (\ref{eq:89}) shows that it accounts for the parts
1495: of the original eigenvector that were on the fine grid and therefore
1496: dropped. In a geometric multigrid method, this is usually done by the
1497: coarsening prescription which moves information from fine to coarse
1498: degrees of freedom; in algebraic multigrid, it results directly from
1499: the $LU$ decomposition.
1500: 
1501: Fig.~\ref{fig:11} shows the lowest positive eigenvalues of the Dirac
1502: matrix and different forms of the Schur complement for a sample $U(1)$
1503: gauge configuration at $\beta=3.0$. With a rescaling factor of
1504: approximately $1+\alpha \approx 4$, the exact Schur complement $S_M$
1505: reproduces the low eigenvalues of $M$ well, and so does the
1506: order-2 optimized operator. The truncated von Neumann series, however,
1507: deviates significantly from the true results for the lowest
1508: eigenvalues, which are expected to be very important for the
1509: convergence properties when used as a preconditioner.
1510: 
1511: \begin{figure}[p]
1512:   \centerline{\includegraphics[width=.8\hsize]{plots/error-by-paths.eps}}
1513:     \caption{The relative error as a function of the number of 
1514:       paths in the approximate Schur complement for different orders
1515:       of the path set. The paths have been reordered so that the
1516:       relative error decays fastests.}
1517:     \label{fig:2}
1518: %\end{figure}
1519: 
1520: %\begin{figure}[htbp]
1521:   \begin{center}
1522:     \includegraphics[width=.8\hsize]{plots/error-by-dist.eps}
1523:     \caption{%
1524:       The relative error from the full optimization as a function of
1525:       the maximum distance of path end points in the path set, shown
1526:       for different maximum lengths of the path.
1527:       }
1528:     \label{fig:4}
1529:   \end{center}
1530: \end{figure}
1531: 
1532: \begin{figure}[htb]
1533:   \begin{center}
1534:     \includegraphics[width=\hsize]{plots/low-spectrum.eps}
1535:     \caption{%
1536:       Lowest positive eigenvalues of the effective coarse-grid
1537:       operator $S$. The lower curve (filled circles) shows 
1538:       eigenvalues of the full Dirac matrix on a $16\times 16$ lattice
1539:       with a sample $U(1)$ gauge field at $\beta=3.0$. The three
1540:       upper curves are the corresponding eigenvalues on the coarse
1541:       grid, calculated from the exact Schur complement (full squares), 
1542:       the order-2 optimized operator (open squares), and the order-2
1543:       von Neumann series (open diamonds). The inset enlarges low
1544:       eigenvalues. 
1545:       }
1546:     \label{fig:11}
1547:   \end{center}
1548: \end{figure}
1549: 
1550: \section{Applications}
1551: 
1552: While a full discussion of the performance of multigrid operators in
1553: different algorithms is outside of the scope of this paper, we give
1554: here two examples of how they can be applied in numerical algorithms.
1555: Preconditioning uses the block $LU$ decomposition provided by the
1556: multigrid decomposition to improve the condition number and thus
1557: convergence in iterative algorithms. Multigrid relaxation is the
1558: classical multigrid algorithm for solving the system (\ref{eq:2}) and
1559: allows a comparison between a two-grid iteration and the original
1560: fine-grid iteration. In all calculations, a $16 \times 16$ lattice was 
1561: used to allow the calculation of complete spectra. Note that only
1562: two-grid algorithms are considered; for realistic problems, the
1563: procedure should be repeated on several grid levels.
1564: 
1565: \begin{figure}[htbp]
1566:   \begin{center}
1567:     \includegraphics[width=\hsize]{plots/precond-1.eps}
1568:     \includegraphics[width=\hsize]{plots/precond-2.eps}
1569:     \caption{
1570:       Spectra of the multigrid preconditioned Dirac matrix $1 - \bar
1571:       M^{-1} M$ for a sample gauge field configuration.
1572:       Optimized approximations of order 1, 2, and 3 where used from left to
1573:       right for the Schur complement.
1574:       The top row shows the spectra obtained using an approximate
1575:       interpolation operator, the bottom row the spectra using the
1576:       exact interpolation operator. Preconditioning with an exact
1577:       Schur complement and interpolation operator would result in a
1578:       single point at the origin.
1579:       \label{fig:5}
1580:     }
1581: 
1582:     \includegraphics[width=\hsize]{plots/precond-3.eps}
1583:     \caption{
1584:       The preconditioned spectrum of fig.~\ref{fig:5},
1585:       using the von Neumann series as approximate Schur complement.
1586:       \label{fig:6}
1587:     }
1588:   \end{center}
1589: \end{figure}
1590: 
1591: \subsection{Preconditioning}
1592: 
1593: In preconditioning, the linear system
1594: \begin{equation}
1595:   \label{eq:69}
1596:   M f = a
1597: \end{equation}
1598: is replaced by an equivalent system
1599: \begin{equation}
1600:   \label{eq:69b}
1601:   \bar M^{-1} M f = a' = \bar M^{-1} a
1602: \end{equation}
1603: using a preconditioning matrix $\bar M$ that is an easily invertible
1604: approximation to $M$. If $\bar M$ is sufficiently close to $M$, the
1605: spectrum of the preconditioned matrix $\bar M^{-1} M$ is contracted towards
1606: unity and its condition number reduced. For multigrid preconditioning,
1607: one uses an approximate block LU decomposition of the form
1608: (\ref{eq:40}) as the preconditioner. The preconditioned matrix is then
1609: \begin{eqnarray}
1610:   \label{eq:97}
1611:   && \bar M^{-1} M =
1612:   \left( \begin{array}{cc} 1 & 0 \\ \bar Q & 1 \end{array}
1613:   \right) \,
1614:   \left( \begin{array}{cc} \bar S^{-1} & 0 \\ 0 & \bar T^{-1} 
1615:   \end{array} \right) \,
1616:   \left( \begin{array}{cc} 1 & \bar Q^\conj \\ 0 & 1 \end{array}
1617:   \right) \,
1618:   \left( \begin{array}{cc} M_{11} & M_{12} \\ M_{21} & \bar M_{22}
1619:   \end{array} \right) 
1620: %\nnm \\
1621: %  &&\quad = 
1622: %  \left( \begin{array}{cc} 
1623: %    \bar S^{-1} ( M_{11} + \bar Q^\conj M_{21} ) &
1624: %    \bar S^{-1} ( M_{12} + \bar Q^\conj M_{22} ) \\
1625: %    \bar Q \bar S^{-1} ( M_{11} + \bar Q^\conj M_{21} ) 
1626: %       + \bar T^{-1} M_{21} &
1627: %    \bar Q \bar S^{-1} ( M_{12} + \bar Q^\conj M_{22} ) 
1628: %       + \bar T^{-1} M_{22} 
1629: %  \end{array} \right) \quad.
1630: \end{eqnarray}
1631: If $Q$ and $T$ are chosen to be exact, $Q=-M_{22}^{-1} M_{21}$ and
1632: $T=M_{22}$, this reduces to
1633: \begin{equation}
1634:   \label{eq:98}
1635:   \bar M^{-1} M =  
1636:   \left( \begin{array}{cc} 
1637:     \bar S^{-1} S_M &
1638:     0\\
1639:     \bar Q \bar S^{-1} ( \bar S^{-1} S_M - 1 ) &
1640:     1 \end{array} \right)
1641:   \quad.
1642: \end{equation}
1643: The properties of the preconditioner are therefore determined by $1 -
1644: \bar S^{-1} S_M$, which is the quantity that was optimized in the
1645: optimization process (\ref{eq:26}) above.
1646: 
1647: %In an iterative algorithm to solve (\ref{eq:69b}), each elementary
1648: %matrix-vector multiplication then entails the application of the
1649: %$\tilde S^{-1}$ and $M_{22}^{-1}$, which itself must be performed
1650: %using an iterative algorithm.  These inner iterations act separately
1651: %on the fine and coarse sublattices and correspond to the V- and
1652: %W-cycles in classical multigrid relaxation schemes.
1653: 
1654: Fig.~\ref{fig:5} shows sample spectra of $1 - \bar M^{-1} M$ with the
1655: Schur complement approximated to order 1, 2, and 3. If an exact
1656: Schur complement was used, the spectrum would reduce to a single point
1657: at the origin; the radius of the disk on which the eigenvalues lie in the
1658: complex plane is a measure of the quality of the preconditioner. The
1659: figure shows the result for two different choices of the interpolation
1660: matrix $Q$: In the bottom row, the exact interpolation $Q=-M_{22}^{-1}
1661: M_{21}$ was used. Since $Q$ is, as opposed to $S$, never inverted in
1662: the process, this is numerically legitimate and no more complex than
1663: the application of the approximate $S^{-1}$. The top row shows the
1664: result when an approximate interpolation matrix $\bar Q$ was used that 
1665: was calculated in the same way as $S$ in an optimization process. It
1666: shows that only if a high order of approximation for $S$ was chosen, it
1667: makes sense to use the exact interpolation operator, otherwise it just 
1668: leads to a larger concentration of eigenvalues without reducing the
1669: radius of the disk.
1670: 
1671: For comparison, fig.~\ref{fig:6} shows the spectrum of a matrix that
1672: was preconditioned using von Neumann series (and the exact
1673: interpolation operator) to the same order as before. While most modes
1674: are quite well reduced, there remain even at order 3 a few eigenvalues 
1675: far away from the origin. These probably correspond to low eigenvalues of
1676: the original matrix that are not properly approximated by the von
1677: Neumann series, as seen in fig.~\ref{fig:11} above.
1678: 
1679: \begin{figure}[p!]
1680:   \begin{center}
1681:     \includegraphics[width=\hsize]{plots/gs-1.eps}
1682:     \includegraphics[width=\hsize]{plots/gs-2.eps}
1683:     \caption{
1684:       Spectra of the convergence matrix for multigrid relaxation
1685:       in a sample gauge field.
1686:       Optimized approximations of order 1, 2, and 3 where used from left to
1687:       right for the Schur complement.
1688:       The top row shows the convergence of the multigrid cycle if a
1689:       full inversion is performed on the coarse grid, and one relaxation 
1690:       step on the fine grid. The bottom row shows the same spectrum if 
1691:       only a single relaxation step is performed on each grid. The
1692:       convergence rate is given by the radius of the smallest
1693:       disk containing eigenvalues. 
1694:       \label{fig:7}}
1695:   \end{center}
1696:   \vspace*{6cm}  %otherwise is placed at end of paper
1697: \end{figure}
1698: 
1699: \subsection{Multigrid relaxation}
1700: 
1701: Relaxation schemes for solving eq.~(\ref{eq:2}) make use of an
1702: iterative process with an update step that is derived from the residual
1703: %Given an approximate solution $f^{(n)}$ of (\ref{eq:2}), the residual
1704: %is easily calculated as
1705: \begin{equation}
1706:   \label{eq:12}
1707:   r^{(n)} = a - M f^{(n)} \quad,
1708: \end{equation}
1709: where $f^{(n)}$ is the current approximate solution.
1710: The true error, i.e.~the amount by which $f^{(n)}$ has to be updated,
1711: can be calculated from the residual by
1712: \begin{equation}
1713:   \label{eq:13}
1714:   e^{(n)} = f - f^{(n)} = M^{-1} r^{(n)} \quad.
1715: \end{equation}
1716: Given an approximation $\bar M^{-1}$ to $M^{-1}$, 
1717: an approximate update step reads
1718: \begin{equation}
1719:   \label{eq:13a}
1720:   f^{(n+1)} = f^{(n)} + T r^{(n)} = (1 - \bar M^{-1} M) f^{(n)} + T a \quad.
1721: \end{equation}
1722: This iteration always has a fixed point at the true solution $f$, and
1723: the quality of the approximation $\bar M^{-1}$ to $M^{-1}$ only
1724: determines the convergence characteristics of the process.
1725: 
1726: Multigrid relaxation is based on using the Schur complement along with
1727: restriction and interpolation to construct an approximation to
1728: $M^{-1}$ on the coarse lattice.  First, the residual is restricted to
1729: the coarse lattice using the restriction
1730: \begin{equation}
1731:   \label{eq:62}
1732:   \bar r^{(n)} = e_1 + Q^\conj e_2 \quad.
1733: \end{equation}
1734: A coarse lattice approximation $\bar e^{(n)}$ to the true error is
1735: then found by applying the inverse of the approximate Schur
1736: complement:
1737: \begin{equation}
1738:   \label{eq:63}
1739:   \bar e^{(n)} = \bar S^{-1} \bar r^{(n)} \quad.
1740: \end{equation}
1741: Finally, this is interpolated back to the fine lattice:
1742: \begin{equation}
1743:   \label{eq:64}
1744:   \bar e_1^{(n)} = \bar e^{(n)} \quad, \qquad
1745:   \bar e_2^{(n)} = Q \bar e^{(n)}
1746: \end{equation}
1747: and used to update the approximation.  This update step acts on the
1748: coarse degrees of freedom only and must be followed by a compatible
1749: relaxation step on the fine lattice.  In matrix form, the
1750: approximate inverse of $M$ used here reads
1751: \begin{equation}
1752:   \label{eq:66}
1753:   T = 
1754:   \left(
1755:     \begin{array}{cc}
1756:       1 & 0 \\ \bar Q & 1
1757:     \end{array}
1758:   \right) \, \left(
1759:     \begin{array}{cc}
1760:       \bar S^{-1} & 0 \\ 0 & 0
1761:     \end{array}
1762:   \right) \, \left(
1763:     \begin{array}{cc}
1764:       1 & \bar Q \\  0 & 1
1765:     \end{array}
1766:   \right) \quad.
1767: \end{equation}
1768: It differs from the true LU decomposition (\ref{eq:44}) in that the
1769: lower right entry of the matrix in the middle is zero instead of
1770: $M^{-1}_{22}$.  If a perfect interpolation and restriction is chosen,
1771: $\bar Q = -M_{22}^{-1} M_{21}$, the iteration matrix for the coarse-grid
1772: step reads
1773: \begin{equation}
1774:   \label{eq:67}
1775:   1-TM = \left(
1776:     \begin{array}{cc}
1777:       1 - \bar S^{-1} S & 0 \\ Q \bar S^{-1} S & 1
1778:     \end{array}
1779:     \right) \quad.
1780: \end{equation}
1781: %and consequently the update step can be written as
1782: %\begin{eqnarray}
1783: %  \label{eq:68}
1784: %  f_1^{(n+1)} &=& (1 - \bar S^{-1} S) \, f_1^{(n)} +
1785: %                \bar S^{-1} \, (a_1 + Q^\conj a_2) \nnm \\
1786: %  f_2^{(n+1)} &=& f_2^{(n)} + Q \left( f_1^{(n+1)} - f_1^{(n)} \right) \quad.
1787: %\end{eqnarray}
1788: Its convergence properties on the coarse lattice are thus given again
1789: by $1 - \bar S^{-1} S$ as in the case of preconditioning; on the fine
1790: lattice, it performs no relaxation at all, and must be followed by an
1791: ordinary relaxation step using $M_{22}$.
1792: 
1793: \begin{figure}[htb]
1794:   \begin{center}
1795:     \includegraphics[width=.8\hsize]{plots/convergence.eps}    
1796:     \caption{Sample convergence of a two-grid relaxation method 
1797:       compared to the same
1798:       method applied to the base grid, for a $U(1)$ Dirac operator on
1799:       a $16 \times 16$-lattice.  The horizontal axis gives the number
1800:       of floating-point multiplications, the vertical axis the true
1801:       error of the algorithm. For the two-grid algorithm, one sweep on
1802:       the coarse lattice and one sweep on the fine lattice were
1803:       performed alternately.}
1804:     \label{fig:8}
1805:   \end{center}
1806: \end{figure}
1807: 
1808: Fig.~\ref{fig:7} shows the spectra of the convergence matrix for
1809: multigrid relaxation. In the top row, each step consists of performing
1810: a complete relaxation on the coarse lattice, and one relaxation step
1811: on the fine lattice. This leads to a concentration of eigenvalues
1812: around zero, showing modes that are nearly completely reduced on the
1813: coarse grid. The remaining modes must be reduced on the fine grid and
1814: determine the residual convergence rate.  In an actual application,
1815: one would not completely reduce the modes on the coarse grid, but
1816: perform coarse- and fine-grid relaxation alternately. The resulting
1817: convergence spectrum is shown in the bottom row; it has a similar
1818: convergence rate as the complete reduction, but requires much less
1819: operations. Note that increasing the order of the approximation from 2
1820: to 3 does not improve the convergence, but this might be related to
1821: the fact that only one fine-grid relaxation step was used here.
1822: 
1823: Finally, we compare the number of operations required to calculate a
1824: sample Green's function in $U(1)$ lattice gauge theory on a $16\times
1825: 16$ lattice. It must be cautioned that this is only an example
1826: calculation using two-grid relaxation, not a full multigrid algorithm.
1827: The method offers a multitude of choices and tunable parameters: the
1828: order and type of the approximation for the $LU$ decomposition, the
1829: iterative algorithm (used here is for simplicity Jacobi relaxation; in
1830: actual applications one would probable use Gauss-Seidel), the
1831: multigrid cycle, and the convergence criteria on the two grids. The
1832: size and dimensionality of the lattice and the gauge group, as well as
1833: the amount of disorder in the gauge field, have been seen to
1834: decisively influence results in other algorithms. Also, the two-grid
1835: algorithm must be extended to a true multigrid algorithm by
1836: recursively applying the method. And finally relaxation might be less
1837: performant than a preconditioned Krylov solver. The ultimate measure
1838: of performance will be the actual computer time used in the
1839: algorithms, but this will depend on the implementation and on the
1840: architecture of the target machine. For example, the computation of
1841: the approximate Schur complement can either be performed beforehand
1842: and stored in memory, or on the fly in the matrix-vector
1843: multiplication of the Schur complement, depending on the amount of
1844: memory available. Here a lower order of approximation, resulting in a
1845: smaller Schur complement, can be favorable though its convergence rate
1846: per relaxation step might be slower.
1847: 
1848: With these precautions, an example run for calculating a sample
1849: Green's function is shown in Fig.~\ref{fig:8}. As a measure of
1850: computer time used, the approximate number of floating-point
1851: multiplications is given. The convergence is measured as the
1852: error norm against the true result.
1853: 
1854: \section{Conclusions}
1855: 
1856: It was demonstrated how the algebraic multigrid method 
1857: can be applied to disordered linear operators.
1858: Other than in previous multigrid approaches, a thinned lattice for the
1859: coarse degrees of freedom is used and averaging over block spins is
1860: avoided. The interpolation operator is chosen to obtain a block $LU$
1861: decomposition, and the resulting coarse-grid dynamics is completely
1862: described by the Schur complement of the original matrix. Gauge
1863: covariance is automatically taken into account by this procedure.
1864: 
1865: The Schur complement can be approximated to arbitrary precision in a
1866: numerical optimization process by expanding it in a linear basis
1867: constructed from connected paths on the original lattice, forming a
1868: generalized stencil for disordered operators.  Similar to
1869: renormalization group techniques, the numerical procedure for
1870: obtaining the expansion coefficients uses the coarse-grid projection
1871: of an ensemble of systems to obtain the coefficients of the
1872: approximate coarse-grid operator.  In this way, the information
1873: gathered in the optimization process can be used to speed up the
1874: actual calculations as the optimized effective coarse-grid operator
1875: ``learns'' the dynamics of the system.
1876: 
1877: The resulting effective coarse-grid operator is constructed from the
1878: generalized stencil as a weighted sum over paths on the original
1879: lattice. It forms a denser, but smaller matrix with next-neighbor and
1880: higher interactions and can be used to improve performance in
1881: numerical algorithms by preconditioning or similar methods, hopefully
1882: not only for matrix inversion, but also for other problems such as
1883: eigensystem analysis and rational matrix functions.  A variety of
1884: tunable parameters make it possible to adapt the procedure to
1885: different algorithms and architectures.  Whether the method will
1886: actually improve real-life algorithms, remains of course to be seen
1887: until realistic systems and efficient implementations are
1888: investigated.
1889: 
1890: \noindent{\small{\bf Acknowledgements}\\
1891:   The author wishes to thank K.~Schilling, T.~Lippert, B.~Medeke,
1892:   and J.~Negele for useful discussions, and the
1893:   Center for Theoretical Physics at MIT for its hospitality. Computer
1894:   time was provided by the NICse cluster computer at the John von
1895:   Neumann Institute.}
1896: 
1897: \ifcbtex
1898: \raggedright
1899: \bibliographystyle{cb}
1900: \fi
1901: \bibliography{mg}
1902: 
1903: \end{document}
1904: