1: \documentclass[12pt]{article}
2: \usepackage{a4}
3: \usepackage{amssymb,amsmath,amsfonts}
4: \usepackage{epsfig}
5: %\usepackage{draftcopy}
6: %\draftcopyLastPage{1}
7: %\setlength{\textwidth}{16.cm}
8: \def\bfit#1{\textbf{\textit{#1}}}
9:
10: \begin{document}
11:
12: \title{On the reduction of hypercubic lattice artifacts}
13: \date{\today}
14: \author{ F.~de~Soto$^{a,b}$ and C.~Roiesnel$^c$ }
15: \par \maketitle
16: \begin{center}
17: $^a$ Dpto. de Sistemas F\'{\i}sicos, Qu\'{\i}micos y Naturales,\\
18: Universidad Pablo de Olavide, Ctra. Utrera Km 1, 41013 Sevilla, Spain.\\
19: $^b$ Laboratoire de Physique Subatomique et Cosmologie, IN2P3, \\
20: 53 av. des Martyrs, 38026 Grenoble, France. \\
21: $^c$ Centre de Physique Th\'eorique, Ecole Polytechnique, CNRS,\\
22: 91128 Palaiseau, France.
23: \end{center}
24:
25: \begin{abstract}
26: This note presents a comparative study of various options to reduce
27: the errors coming from the discretization of a Quantum Field Theory
28: in a lattice with hypercubic symmetry. We show that it is possible
29: to perform an extrapolation towards the continuum which is able to
30: eliminate systematically the artifacts which break the $O(4)$
31: symmetry.
32: \end{abstract}
33: \vfill
34: \begin{flushleft}
35: CPHT RR 016.0307\\
36: \end{flushleft}
37:
38: \enlargethispage{0.5cm}
39:
40: \newpage
41:
42: \section{Introduction}
43: \label{sec:intro}
44:
45: The problem of restoration of rotational invariance was the focus of
46: much work in the early days of numerical simulations of lattice gauge
47: theories, which were performed on very small lattices. Most noteworthy
48: were the attempts to find alternative discretizations which would
49: approach the continuum limit more rapidly than the simple hypercubic
50: lattice. One line of attack \cite{CEL82} was to discretize gauge
51: theories on the most symmetric of all four-dimensional lattices, the
52: four-dimensional body-centered hypercubic lattice, whose point
53: symmetry group is three times as large as the hypercubic group.
54: Another angle of investigation worth mentioning was to formulate gauge
55: theories on random lattices \cite{LEE82}. The interest in these
56: alternate formulations faded away in subsequent years, first because
57: of their inherent complications, but mainly when it was realized that
58: rotational invariance was in fact restored within statistical errors
59: at larger distances on the hypercubic lattice.
60:
61: However, the treatment of discretization errors in numerical
62: simulations of a lattice gauge theory can remain a vexing problem
63: in some data analyses. Indeed, the signal of some lattice
64: observables, such as the two-point Green functions in momentum
65: space, has become so good that the systematic errors become very
66: much larger than the statistical errors. A general method, which
67: we call the H4 method, has been devised quite some time ago
68: \cite{LPTX99,LPTX00} to eliminate hypercubic artifacts from the
69: gluon two-point functions and extrapolate the lattice data towards
70: the continuum. This extrapolation is crucial to succeed in a
71: quantitative description, at least in the ultraviolet regime. Such
72: a method, despite its success in describing other two-point
73: functions as well, as the fermion \cite{LPT03} or the ghost
74: \cite{LPTX05} propagators, has not been widely adopted. Indeed,
75: most other studies of the lattice two-point functions are still
76: using phenomenological recipes \cite{ADELAIDE} which only allow
77: for a qualitative description of the data, since it is usually not
78: possible to make quantitative fits with a reasonable chisquare.
79:
80: The purpose of this note is threefold. First we want to gather some
81: pieces about the H4 technique which are scattered in various sections
82: of previous publications and which may have been overlooked. Our
83: second objective is to stress, on a simple controllable model, that
84: the H4 method can be systematically improved, contrarily to the
85: empirical methods, when the statistical errors decrease. Our last goal
86: is to point out the general applicability of the method, not only to
87: those scalar form factors in momentum space which depend on a single
88: invariant, but also to various other lattice observables.
89:
90: The plan of the paper is as follows. In the next section we recall
91: the general technique of hypercubic extrapolations towards the
92: continuum of any lattice scalar form factor depending upon a
93: single momentum. In the following section we show that a simple
94: model, a free real scalar field in four dimensions, can be used as
95: a testbed for the hypercubic extrapolations. Then we make a
96: detailed comparison of the different strategies to eliminate the
97: hypercubic lattice artifacts. The concluding section is devoted to
98: recommendations about the best usage of the H4 extrapolation
99: method. We also outline some straightforward generalizations.
100:
101: \section{Hypercubic artifacts}
102: \label{sec:artifact}
103:
104: Any form factor $F_L(p)$ which is a scalar invariant on the lattice,
105: is invariant along the orbit $O(p)$ generated by the action of the
106: isometry group $H(4)$ of hypercubic lattices on the discrete momentum
107: $p\equiv\frac{2\pi}{La}\times(n_{1},n_{2},n_{3},n_{4})$ where the
108: $n_{\mu}$'s are integers, $L$ is the lattice size and $a$ the lattice
109: spacing. The general structure of polynomial invariants under a
110: finite group is known from group-invariant theory \cite{WEYL}. In
111: particular, it can be shown that any polynomial function of $p$ which
112: is invariant under the action of $H(4)$ is a polynomial function of
113: the 4 invariants
114: \begin{align}
115: p^{[n]} \equiv \sum_{\mu}p_{\mu}^{n},\quad n = 2, 4, 6, 8
116: \end{align}
117: which index the set of orbits. The appendix contains an elementary
118: derivation.
119:
120: It is thus possible to use these 4 invariants to average the form
121: factor over the orbits of $H(4)$ to increase the statistical accuracy:
122: \begin{align}
123: F_L(p) \equiv F_{L}(p^{[2]},p^{[4]},p^{[6]},p^{[8]}) =
124: \frac{1}{\|O(p)\|} \sum_{p\in O(p)} F_L(p)
125: \end{align}
126: where $\|O(p)\|$ is the cardinal number of the orbit $O(p)$.
127:
128: The orbits of the continuum isometry group $O(4)$ are of course
129: labeled by the single invariant $p^{[2]}\equiv p^2$, and lattice
130: momenta which belong to the same orbit of $O(4)$ do not belong in
131: general to the same orbit of $H(4)$. For instance, as soon as
132: $n^2\equiv\sum_{\mu=1}^{4}n^2_{\mu}=4$ in integer lattice units,
133: the $O(4)$ orbit splits into two distinct H(4) orbits, those of
134: the vectors $(2,0,0,0)$ and $(1,1,1,1)$ respectively. Therefore we
135: can distinguish two kinds of lattice artifacts, those which depend
136: only upon the invariant $p^{2}$, and which produce the scaling
137: violations, and those which depend also upon the higher-order
138: invariants $p^{[n]}$ ($n= 4,6,8$) and which we call hypercubic
139: artifacts. When the difference between the values of $F_L(p)$
140: along one orbit of $O(4)$ become larger than the statistical
141: errors, one needs at least to reduce the hypercubic artifacts from
142: the lattice data before attempting any quantitative analysis.
143:
144: The treatment of these discretization artifacts can be inferred
145: from lattice perturbation theory, as Green functions will depend
146: on some lattice momentum~\footnote{Depending on the discretization
147: scheme, it will be $\widehat{p}_\mu$ or
148: $\overline{p}_\mu=\frac{1}{a}\sin{a p_\mu}$, etc.}
149: \begin{equation}
150: \widehat{p}_\mu\equiv \frac{2}{a} \sin\left(\frac{ap_\mu}{2}\right) \
151: \end{equation}
152: instead of the continuum one, $p_\mu=\frac{2\pi}{La}n_\mu$. By
153: developing the lattice momentum $\widehat{p}^2\equiv \sum_\mu
154: \widehat{p}_\mu^2$ in terms of the lattice spacing $a$, one gets:
155: \begin{equation}
156: \widehat{p}^2 \approx\ p^2 - \frac{a^2}{12} p^{[4]} + \frac{a^4}{360}
157: p^{[6]} - \frac{a^6}{20160} p^{[8]} + \cdots \label{hatkdev}
158: \end{equation}
159: and thus, the lattice momentum differs from the "continuum" one by
160: discretization artifacts that are proportional to the invariants
161: $p^{[4]}$ (of order $a^2$), $p^{[6]}$ (order $a^4$), etc.
162:
163: The strategies to minimize the hypercubic artifacts are based on the
164: fact these artifacts depend on the non O(4) invariants, $p^{[4]}$,
165: $p^{[6]}$, etc. and thus reducing $p^{[4]}$ would also reduce the
166: artifacts. For example, the improved restoration of the rotational
167: symmetry on the four-dimensional body-centered hypercubic lattice can
168: be analyzed in terms of the primitive invariant $p^{[4]}$ \cite{NEU87}
169: \footnote{We thank Ph.~de~Forcrand for pointing out this reference to
170: us.}. These strategies fall into three general groups:
171:
172: \begin{itemize}
173: \item The simplest one is just to keep only the H(4) orbits which
174: minimizes $p^{[4]}$ along each O(4) orbit. As they lay near the
175: diagonal, a more efficient prescription \cite{ADELAIDE} is to
176: impose a "cylindrical" cut on the values of $p$, keeping only
177: those that are within a prescribed distance of the diagonal. This
178: completely empirical recipe has been widely adopted in the
179: literature and we shall refer to it in the sequel as the
180: ``democratic'' method. The main drawbacks are that the information
181: for most of the momenta is lost (for moderate lattices only a
182: small fraction of the momenta is kept) and that although $p^{[4]}$
183: is small for the orbits kept, it is not null, and therefore the
184: systematic errors are still present.
185:
186: \item The other methods try to fully eliminate the contribution of
187: $p^{[4]}$, etc. and we will generically refer to them as the H4
188: methods. By analogy with the free lattice propagators, it is natural
189: to make the hypothesis that the lattice form factor is a smooth
190: function of the discrete invariants $p^{[n]}$, $n\geq 4$, near the
191: continuum limit,
192: \begin{align}
193: \label{eq:invariants}
194: \begin{split}
195: F_{L}(p^{2},p^{[4]},p^{[6]},p^{[8]}) &\approx
196: F_{L}(p^{2},0,0,0) + p^{[4]}\frac{\partial F_{L}}
197: {\partial p^{[4]}}(p^{2},0,0,0) + \\
198: &\quad p^{[6]}\frac{\partial F_{L}}{\partial p^{[6]}}(p^{2},0,0,0)
199: + (p^{[4]})^2\frac{\partial^2 F_{L}}{\partial^2 p^{[4]}}(p^{2},0,0,0)
200: + \cdots
201: \end{split}
202: \end{align}
203: and $F_{L}(p^{2},0,0,0)$ is nothing but the form factor of the
204: continuum in a finite volume, up to lattice artifacts which do not
205: break $O(4)$ invariance and which are true scaling violations. We
206: emphasize that we are merely conjecturing that the restoration of
207: rotational invariance is smooth when taking the continuum limit at
208: fixed $p^2$. When several orbits exist with the same $p^2$, the
209: simplest method \cite{LPTX99} to reduce the hypercubic artifacts is to
210: extrapolate the lattice data towards $F_{L}(p^{2},0,0,0)$ by making a
211: linear regression at fixed $p^{2}$ with respect to the invariant
212: $p^{[4]}$ (note that the contributions of other invariants are of
213: higher order in the lattice spacing).
214:
215: Obviously this method only applies to the O(4) orbits with more than
216: one H(4) orbit. If one wants to include in the data analysis the
217: values of $p^2$ with a single H(4) orbit, one must interpolate the
218: slopes extracted from \eqref{eq:invariants}. This interpolation can be
219: done either numerically or by assuming a functional dependence of the
220: slope with respect to $p^{2}$ based, for example, on dimensional
221: arguments \cite{LPTX00}. For instance, for a massive scalar lattice
222: two-point function, the simplest ansatz would be to assume that the
223: slope has the same leading behavior as for a free lattice propagator:
224: \begin{align}
225: \label{eq:slope}
226: \frac{\partial F_{L}}
227: {\partial p^{[4]}}(p^{2},0,0,0) &=
228: \frac{a^2}{\left(p^{2}+m^2\right)^{2}}\left( c_{1}+ c_{2}a^2p^{2}\right)
229: \end{align}
230: The range of validity of the method can be checked a posteriori
231: from the smoothness of the extrapolated data with respect to
232: $p^2$. The quality of the two-parameter fit to the slopes, and the
233: extension of the fitting window in $p^{2}$, supplies still another
234: independent check of the validity of the extrapolations, although
235: the inclusion of $O(4)$-invariant lattice spacing corrections is
236: usually required to get fits with a reasonable $\chi^{2}$.
237:
238: This strategy based on independent extrapolations for each value
239: of $p^2$ will be referred to as the local H4 method.
240:
241: \item The number of distinct orbits at each $p^{2}$ --in physical
242: units-- increases with the lattice size and, eventually, a linear
243: extrapolation limited to the single invariant $p^{[4]}$ breaks down.
244: But, by the same token, it becomes possible to improve the local H4
245: method by performing a linear regression at fixed $p^2$ in the
246: higher-order invariants as well. Therefore, when the lattice size
247: increases, the H4 technique provides a systematic way to include
248: higher-order invariants and to extend the range of validity of the
249: extrapolation towards the continuum. For those $p^2$ which do not
250: have enough orbits to perform the extrapolation, it is still
251: possible to make use of all available physical information in the
252: modelling of the functional derivatives appearing in
253: \eqref{eq:invariants} and to perform an interpolation.
254:
255: An alternative strategy is based on the fact that the functional
256: derivatives which appear in \eqref{eq:invariants} are functions of
257: $p^2$ only. These functions can be represented by a Taylor
258: development in their domain of analyticity, or, more conveniently,
259: by a Laurent series, as it does not assume analyticity and makes
260: appear all the terms allowed by dimensional arguments. Moreover, it
261: is always possible to use polynomial approximation theory and expand
262: the functional derivatives in terms of, e.g., Chebyshev polynomials
263: or in a fourier series, etc.
264:
265: In any case, these linear expansions allow to perform the
266: continuum extrapolation through a global linear fit of the
267: parameters for all values of $p^2$ inside a window at once. Such
268: a strategy has been developed for the analysis of the quark
269: propagator \cite{LPT03} and we shall refer to it as the global H4
270: method. The global H4 extrapolation is simple to implement since
271: the numerical task amounts to solving a linear system. It
272: provides a systematic way to extend the range of validity of the
273: extrapolation towards the continuum, not only for large lattices
274: (where the inclusion of $O(a^4)$ and even $O(a^6)$ discretization
275: errors becomes possible) but also for small lattices (where the
276: local H4 method for $O(a^2)$ errors is inefficient due to the
277: small number of orbits), by using in the fit all available lattice
278: data points.
279:
280: \end{itemize}
281:
282: \section{The free scalar field}
283:
284: In order to analyze a model simple enough to provide a complete
285: control of the hypercubic errors in four dimensions, we have
286: chosen a free real scalar field, whose dynamics is given by the
287: lagrangian:
288: \begin{align}
289: \mathcal{L} \ =\ \frac{1}{2} m^2 \phi(x) \phi(x) + \frac{1}{2}
290: \partial_\mu\phi(x) \partial^\mu\phi(x)
291: \label{lcon}
292: \end{align}
293: The naive discretization of \eqref{lcon} leads to the lattice action:
294: \begin{align}
295: S \ =\ \frac{a^4}{2} \sum_x \left\{m^2 \phi_x^2 + \sum_{\mu=1}^{4}
296: (\nabla_\mu \phi_x)^2 \right\}
297: \end{align}
298: where $\nabla_{\mu}$ is the forward lattice derivative, or in momentum
299: space,
300: \begin{align}
301: S \ =\ \frac{a^4}{2} \sum_p \left(m^2 + \widehat{p}^2\right)
302: |\widetilde{\phi}_p|^2
303: \end{align}
304: where $p$ is the discrete lattice momentum. Therefore, the field
305: $\widetilde{\phi}_p$ can be produced by means of a gaussian
306: sampling with standard deviation $\sqrt{m^2 +
307: \widehat{p}^2}$. As this is a cheap lattice calculation, we
308: can go to rather big volumes, up to $64^4$ in this work, and we can
309: generate a high number of fully decorrelated configurations. In order
310: to study the effect of statistics over the results, averages will be
311: made over ensembles of $100$ till $1000$ configurations.
312:
313: This lattice model is of course solvable, and the propagator reads:
314: \begin{align}
315: \label{plat} \Delta_L (p)\ =\ \frac{1}{\widehat{p}^2+m^2}
316: \end{align}
317: The lattice artifacts are exactly computable by expanding
318: $\widehat{p}^2$ in terms of the $H(4)$ invariants introduced in the
319: previous section and plugging the development (\ref{hatkdev})
320: into \eqref{plat},
321: \begin{align}
322: \label{devprop}
323: \Delta_L(p^2,p^{[4]},p^{[6]},p^{[8]}) &\approx \frac{1}{p^2+m^2}
324: \ +\ a^2 \left\{ \frac{1}{12} \frac{p^{[4]}}{(p^2+m^2)^2}
325: \right\}
326: \nonumber \\
327: &+ a^4 \left\{ \frac{1}{72} \frac{{p^{[4]}}^2}{(p^2+m^2)^3} -
328: \frac{2}{8!}\frac{p^{[6]}}{(p^2+m^2)^2} \right\}+\cdots
329: \end{align}
330: and the continuum propagator $\Delta_0(p)$ is indeed recovered
331: \bfit{smoothly} in the limit $a\to0$. But as long as we are
332: working at finite lattice spacing, there will be corrections in
333: $a^2$, $a^4$, etc. that are not at all negligible, as can be
334: appreciated in figure \ref{bonefish} which plots the ratio
335: $\Delta_L(p)/\Delta_0(p)$ for a $32^4$ lattice.
336:
337: One could wonder whether such a model is really useful since the
338: lattice artifacts are exactly known. For instance one can recover the
339: continuum propagator from the lattice propagator by merely plotting
340: the lattice data as a function of $\widehat{p}^2$ rather than $p^2$!
341: However this simple recipe is no longer applicable to an interacting
342: theory where the lattice two-point functions \bfit{do depend} upon the
343: independent variables $\widehat{p}^{[n]} =
344: \sum_{\mu}(\widehat{p}_{\mu})^{n},\ n=4,6,8$ (as illustrated in figure
345: 1 of reference \cite{LPTX99}). And there is no systematic way to
346: separate out cleanly the effect of these additional variables because
347: $\widehat{p}^2$ is not an $O(4)$ invariant. Indeed, because
348: $\widehat{p}^2$ takes on different values on every $H(4)$ orbit, there
349: is only one data point per value of $\widehat{p}^2$ and the H4 method,
350: either local or global, is not appropriate for the choice of momentum
351: variable $\widehat{p}$.
352:
353: However one should exercise special attention at using this model
354: without the information provided by expression \eqref{devprop}
355: (except of course the smoothness assumption in the $H(4)$
356: invariants $p^{[n]}$). Under this proviso, the model can serve as
357: a bench test of the different approaches to eliminate hypercubic
358: artifacts. In particular {\it we will not use}
359: Eq.\,\eqref{eq:slope}.
360:
361: \begin{figure}[!ht]
362: \begin{center}
363: \epsfxsize10cm\epsffile{plot_all_N=1000.eps}
364: \end{center}
365: \caption{\small{\it Raw dressing function
366: $\Delta_L(p)/\Delta_0(p)$ as a
367: function of $p^2/m^2$ for a $32^4$ lattice and $am=1$ from a
368: sample size of 1000 configurations.}}
369: \label{bonefish}
370: \end{figure}
371:
372: The model has one mass parameter $m$ which fixes the scale. We
373: will study the worst-case scenario where $m$ cannot be neglected
374: with respect to $p$ when the lattice artifacts are
375: large\footnote{As
376: $p=\frac{2\pi}{La}n$, with $n=0,\cdots,L/2$, a suitable value is $am=1$.}.
377:
378: \textit{The case of QCD is, in fact, simpler, as long as
379: $\Lambda_{QCD}$ and quark masses are negligible in comparison to
380: the momentum scale, which would correspond to the case $am\ll 1$.
381: Then, by dimensional arguments, the artifacts can be modeled at
382: least in the ultraviolet regime, as proposed in \cite{LPTX00} and
383: \cite{LPT03}.}
384:
385: \section{Comparative study of H4 extrapolations}
386: \label{sec:H4}
387:
388: We will now use a free scalar field with $am=1$ to compare the
389: different strategies to extract the continuum behavior from the
390: lattice data. We will use lattice units and set $a\equiv 1$
391: throughout this section. We restrict ourselves to one or two
392: representative methods within each strategy:
393: \begin{itemize}
394: \item The democratic method with a cylindrical cut selecting out the
395: orbits that are within a distance of 2 lattice units from the diagonal
396: $(1,1,1,1)$.
397: \item The local H4 method with independent extrapolations up to
398: $\mathcal{O}(a^2)$ artifacts for every $p^2$ with several orbits
399: within the window $n^2>5$ ($p=\frac{2\pi}{L}n$) up to some
400: $n^2_{max}$:
401: \[\Delta_L(p^2,p^{[4]},p^{[6]},p^{[8]}) = \Delta_L(p^2,0,0,0,0) +
402: c(p^2)p^{[4]} \] The slopes $c(p^2)$ are then fitted with the following
403: functional form
404: \begin{align}
405: c(p^2) = \frac{c_{-1}}{p^2} + c_0 + c_1p^2
406: \end{align}
407: which is used to extrapolate the points with only one orbit inside the
408: window $]5,n^2_{max}]$.
409: \item The global H4 methods with the coefficients of the artifacts up
410: to $\mathcal{O}(a^2)$ or up to $\mathcal{O}(a^4)$ chosen as a
411: Laurent series:
412: \begin{align}
413: \label{eq:global}
414: \nonumber
415: \Delta_L(p^2,p^{[4]},p^{[6]},p^{[8]}) &= \Delta_L(p^2,0,0,0,0) +
416: f_1(p^2)p^{[4]} + f_2(p^2)p^{[6]} + f_3(p^2)(p^{[4]})^2 \\
417: f_n(p^2) &= \sum_{i=-1}^{1} c_{i,n}(p^2)^{-i}\,,\quad n=1,2,3
418: \end{align}
419: With such a choice, a global fit within the window $]5,n^2_{max}]$
420: amounts to solving a linear system of respectively $n^2_{max}-2$ and
421: $n^2_{max}+4$ equations~\footnote{Those variables correspond
422: respectively to the extrapolated propagators, $\Delta_L(p^2,0,0,0,0)$,
423: and the 3 coefficients of each Laurent series.}.
424: \end{itemize}
425: Notice that we do not use the knowledge of the mass, $m=1$, in
426: both the local H4 method and the global H4 method, neither
427: directly nor indirectly (by introducing a mass scale as a
428: parameter). Our purpose is to stress the H4 extrapolation methods
429: to their limits. In practice, of course, all the physical
430: information can be used in order to improve the elimination of the
431: discretization artifacts.
432:
433: In figure~\ref{fig:dem_vs_local} the extrapolated dressing
434: functions $\Delta_E(p^2)/\Delta_0(p^2)$, with the notation
435: $\Delta_E(p^2)\equiv \Delta_L(p^2,0,0,0)$, of the democratic
436: method and of the local H4 method (with $p^2_{max}=3\pi^2/4$), are
437: compared for 1000 configurations generated on a $32^4$ lattice. It
438: can be seen that the dressing function of the democratic method
439: deviates very early from unity whereas the dressing function of
440: the local H4 method is pretty consistent with unity within
441: statistical errors for $p^2$ up to $\approx\pi^2/4$.
442:
443: \begin{figure}[!ht]
444: \begin{center}
445: \epsfxsize10cm\epsffile{plot_dem_N=1000.eps}
446: \end{center}
447: \caption{\small{\it Comparison of the extrapolated dressing function
448: $\Delta_E(p^2)/\Delta_0(p^2)$ as a function of $p^2$ on a
449: $32^4$ lattice ($a=m=1)$, between the democratic method (open
450: squares) and the local H4 method (black circles) -
451: {\bf 1000 configurations}.}}
452: \label{fig:dem_vs_local}
453: \end{figure}
454:
455: Figure~\ref{fig:global} compares the extrapolated dressing
456: functions of the global H4 methods, with respectively up to
457: $\mathcal{O}(a^2)$ and up to $\mathcal{O}(a^4)$ artifacts (again
458: with $p^2_{max}=3\pi^2/4$), for 1000 configurations generated on a
459: $64^4$ lattice. The global H4 method up to $\mathcal{O}(a^2)$
460: performs roughly as the local H4 method. The global H4 method
461: which takes into account $\mathcal{O}(a^4)$ artifacts is able to
462: reproduce the continuum dressing function within statistical
463: errors for $p^2$ up to $\approx \pi^2/2$.
464:
465: \begin{figure}[!ht]
466: \begin{center}
467: \epsfxsize10cm\epsffile{plot_a4_N=1000.eps}
468: \end{center}
469: \caption{\small{\it Comparison of the extrapolated dressing function
470: $\Delta_E(p^2)/\Delta_0(p^2)$ as a function of $p^2$ on a $64^4$
471: lattice ($a=m=1$), between the global methods with
472: $\mathcal{O}(a^2)$ artifacts (open losanges) and $\mathcal{O}(a^4)$
473: (black circles) - {\bf 1000 configurations}.}}
474: \label{fig:global}
475: \end{figure}
476:
477: It is possible to put these qualitative observations on a more
478: quantitative basis, and show precisely the effect of both the lattice
479: size and the sample size on each extrapolation method. Since
480: all components of a free scalar field in momentum space are
481: independent gaussian variables, the statistical distribution of the
482: quantity
483: \begin{align}
484: \label{eq:chi2}
485: \chi^2 = \sum_{p^2=1}^{p^2_{max}}
486: \left(\frac{\Delta_E(p^2) - \Delta_0(p^2)}{\delta\Delta_E(p^2)} \right)^2
487: \end{align}
488: should follow exactly the chisquare law for $n^2_{max}$
489: independent variables, if the systematic errors of an
490: extrapolation method are indeed smaller than the statistical
491: errors. The criterion is exact for the democratic and local H4
492: methods which produce independent extrapolated values.
493: Extrapolations by the global H4 method are correlated and one must
494: include the full covariance matrix of the fit in the definition of
495: the chisquare:
496: \begin{equation}
497: \chi^2 = \sum_{p^2=1}^{p^2_{max}} \sum_{q^2=1}^{p^2_{max}}
498: (\Delta_E(p^2) - \Delta_0(p^2))M(p^2,q^2)(\Delta_E(q^2) -
499: \Delta_0(q^2))\ ,
500: \end{equation}
501: and $M(p^2,q^2)=p^2_{max} (C^{-1})(p^2,q^2)$ is related to the
502: covariance matrix $C(p^2,q^2)$.
503:
504: With these considerations, we compute the $\chi^2/d.o.f.$ of a
505: zero-parameter fit of the extrapolated form factor to its known
506: value $\Delta_0(p^2)=1$ for all $p^2$. Figure~\ref{fig:chi2}
507: displays the evolution of the chisquare per degree of freedom as a
508: function of the fitting window $]5,n^2_{max}]$ on a $32^4$
509: lattice, for each extrapolation method. The local and global H4
510: methods which cure just $\mathcal{O}(a^2)$ artifacts are indeed
511: safe up to $p^2_{max}\approx \pi^2/4$.
512:
513: \begin{figure}[!ht]
514: \begin{center}
515: \epsfxsize10cm\epsffile{chi2_32_1000.eps}
516: \end{center}
517: \caption{\small{\it Evolution of the $\chi^2/d.o.f$ as a function of
518: $p^2_{max}$ on a $32^4$ lattice ($a=m=1)$, for the local $a^2$ method
519: (blue solid line), the global $a^2$ method (red dotted line) and the
520: global $a^4$ method (green dash-dotted line). The smooth curves are
521: the 95\% confidence levels lines - {\bf 1000 configurations}.}}
522: \label{fig:chi2}
523: \end{figure}
524:
525: For the range of lattice sizes and sample sizes considered in this
526: work, the global H4 method which takes into account
527: $\mathcal{O}(a^4)$ artifacts performs best. With such a method it
528: is possible to extend the range of validity of the extrapolation
529: towards the continuum up to $p^2\approx 5-6$, according to the
530: lattice size and at least down to the levels of statistical
531: accuracy studied here.
532:
533: \section{Conclusion}
534:
535: Table~\ref{tab:dem_vs_local} summarizes our findings. For each
536: lattice size, sample size and extrapolation method studied in this
537: work, the table displays the upper bound $p^2_{max}$ of the
538: momentum window $]0,a^2p^2_{max}]$ ($am=1$), inside which the
539: extrapolated dressing function $\Delta_E(p^2)/\Delta_0(p^2)$ is
540: consistent with 1 at a $\chi^2/d.o.f.=2$ level.
541:
542: \begin{table}[ht]
543: \begin{center}
544: \begin{tabular}{|| c || c | c || c | c ||}
545: \hline
546: \hline
547: Lattice size & 32 & 32 & 64 & 64 \\
548: \hline
549: Sample size & 100 & 1000 & 100 & 1000 \\
550: \hline
551: Democratic method & 1.4 (2\%) & 0.54 (0.9\%)& 1.8 (1.9\%) & 1.1 (0.6\%) \\
552: \hline
553: Local $a^2$ method & 6.3 (2\%) & 4.2 (0.8\%) & 4.4 (1.4\%) & 3.4 (0.5\%) \\
554: \hline
555: Global $a^2$ method & 6.3 (0.7\%) & 4.3 (0.3\%) & 4.0 (0.46\%) & 3.1 (0.15\%) \\
556: \hline
557: Global $a^4$ method & $\pi^2$ (0.9\%) & 9.2 (0.4\%) & $\pi^2$ (0.35\%) & 6.7 (0.12\%)\\
558: \hline
559: \hline
560: \end{tabular}
561: \end{center}
562: \caption{\small \it $p^2_{max}$ as a function of the lattice size
563: and the sample size for which $\chi^2/d.o.f.=2$. The statistical
564: error on the extrapolated dressing function is shown between
565: parentheses.} \label{tab:dem_vs_local}
566: \end{table}
567:
568: The limits established in table \ref{tab:dem_vs_local} have been
569: obtained as described in section \ref{sec:H4}. They could be
570: improved by adding more terms to the Laurent's development, or
571: taking into account their perturbative form in the parametrization
572: of the artifacts.
573:
574: Table~\ref{tab:dem_vs_local} is all what is needed to set up an H4
575: extrapolation towards the continuum. Our recommendations are the
576: following. If it is not required to push the extrapolation in
577: $a^2p^2$ above $\approx \pi^2/4$, then it is sufficient to use an
578: H4 method, either local or global, up to $\mathcal{O}(a^2)$
579: artifacts. On larger windows, the global H4 method at least up to
580: $\mathcal{O}(a^4)$ artifacts should be used. The precise tuning of
581: $p^2_{max}$ can be read off the table in each case.
582:
583: The sample sizes used in this study are what is typically achieved
584: in lattice studies of two-point functions with $\mathcal{O}(1-10)$
585: GFlops computers. With sufficient time allocated on
586: $\mathcal{O}(1)$ Tflops computers, it would become possible to
587: increase the statistics by one or two orders of magnitude. Then
588: Table~\ref{tab:dem_vs_local} would no longer be accurate enough
589: and the analysis of this work would need to be repeated, including
590: the global H4 method up to $\mathcal{O}(a^6)$ artifacts in order
591: to keep the extrapolation windows as large. Let us emphasize that
592: such an analysis is straightforward to implement. With adequate
593: statistics, the global H4 extrapolation method can be
594: systematically improved.
595:
596: A one or two order of magnitude increase of statistics would also
597: allow to apply the H4 extrapolation techniques to three-point
598: functions as well. Indeed, with a sample size around 1000
599: configurations, the discretization errors in such lattice
600: observables, although noticeable, are not large enough to be
601: separated from the statistical errors. Three-point functions
602: depend on two momenta. It can be shown that there are now 14
603: \bfit{algebraically} independent symmetric invariants $\phi(p,q)$
604: under the action of the hypercubic group H(4), and among them, we
605: have the three $O(4)$ invariants
606: \begin{align*}
607: \sum_{\mu} p^2_{\mu}\,,\quad\sum_{\mu} q^2_{\mu}\,,\quad
608: \sum_{\mu} p_{\mu}q_{\mu}
609: \end{align*}
610: and 5 algebraically, and functionnally, independent invariants of
611: order $a^2$ which can be chosen as
612: \begin{align*}
613: \sum_{\mu} p^4_{\mu}\,,\quad\sum_{\mu} q^4_{\mu}\,,\quad
614: \sum_{\mu} p^2_{\mu}q^2_{\mu}\,,\quad\sum_{\mu} p^3_{\mu}q_{\mu}\,,\quad
615: \sum_{\mu} p_{\mu}q^3_{\mu}
616: \end{align*}
617: Three-point form factors are usually measured only at special
618: kinematical configurations. Assuming again smoothness of the
619: lattice form factor with respect to these $\mathcal{O}(a^2)$
620: invariants, the global H4 extrapolation method could still be
621: attempted provided that enough lattice momenta and enough H4
622: orbits are included in the analysis.
623:
624: A more straightforward application of the (hyper)cubic extrapolation
625: method is to asymmetric lattices $L^3\times T$ with spatial cubic
626: symmetry. Lattices with $T\gg L$ are produced in large scale
627: simulations of QCD with dynamical quarks at zero temperature, whereas
628: simulations of QCD at finite temperature require lattices with $T\ll
629: L$. For such lattices, the continuum limit can still be obtained
630: within each time slice by applying the techniques described in this
631: note to the cubic group $O_h$.
632:
633: We want to end by pointing out that (hyper)cubic extrapolations
634: methods are not restricted to momentum space but can also be used
635: directly in spacetime. We will sketch one example for
636: illustration, the static potential.
637:
638: Lattice artifacts show up in the static potential at short distances
639: and the standard recipe \cite{MICHAEL} to correct the artifacts is to
640: add to the functional form which fits the static potential a term
641: proportional to the difference $\delta G(R)$ between the lattice
642: one-gluon exchange expression and the continuum expression $1/R$. The
643: technique we advocate is rather to eliminate the cubic artifacts from
644: the raw data measured on the lattice.
645:
646: Indeed the lattice potential extracted from the measurements of an
647: ``off-axis'' Wilson loop connecting the origin to a point at distance
648: $R=\sqrt{x^2+y^2+z^2}$ can be expressed\footnote{at least for L-shaped
649: loops.}, after averaging over the orbits of the cubic group $O_h$, as a
650: function of three invariants:
651: \begin{align*}
652: V_L(x,y,z) \equiv V_L(R^2,R^{[4]},R^{[6]})\,,\quad R^{[n]} = x^n+y^n+z^n
653: \end{align*}
654: An extrapolation towards the continuum can be performed with the
655: methods described in section~\ref{sec:artifact} by making the
656: smoothness assumption with respect to the invariants
657: $R^{[4]},\,R^{[6]}$.
658:
659: \appendix
660: \subsection*{Acknowledgments}
661:
662: We wish to thank our colleagues, Ph.~Boucaud, J.P.~Leroy, A. Le
663: Yaouanc, J.~Micheli, O.~P\`ene, J.~Rodr\'iguez-Quintero, who have
664: collaborated to the development of the H4 extrapolation method.
665: These calculations were performed at the IN2P3 computing center in
666: Lyon. F.S. is specially indebted to J. Carbonell and LPSC for
667: their warm hospitality.
668:
669: \section{H(4) invariants}
670:
671: A general polynomial of degree $N$ in the four components of the
672: momentum $p$ reads:
673: \begin{align*}
674: P_N(p_1,p_2,p_3,p_4) = \sum_{n=0}^{N} \sum_{n_1+n_2+n_3+n_4=n} c_{n_1n_2n_3n_4}
675: \,p_1^{n_1}p_2^{n_2}p_3^{n_3}p_4^{n_4}\ .
676: \end{align*}
677: But any polynomial function of $p$ which is invariant under the
678: action of $H(4)$ must be invariant under every permutation of the
679: components of $p$ and every reflection $p_{\mu}\rightarrow
680: -p_{\mu}$. In particular such a polynomial must be an even
681: function of each component $p_{\mu}$ and contain only symmetric
682: combinations of the components. As there are 4 components, we can
683: construct 4 symmetric combinations that are independent. They are
684: usually chosen as the elementary symmetric polynomials:
685: \begin{align*}
686: \sigma_1 &= p_1^2+p_2^2+p_3^2+p_4^2 \\
687: \sigma_2 &= p_1^2p_2^2+p_1^2p_3^2+p_1^2p_4^2+p_2^2p_3^2+p_2^2p_4^2+p^2_3p^2_4 \\
688: \sigma_3 &= p^2_1p^2_2p^2_3+p^2_1p^2_2p^2_4+p^2_1p^2_3p^2_4+p^2_2p^2_3p^2_4 \\
689: \sigma_4 &= p^2_1p^2_2p^2_3p^2_4
690: \end{align*}
691: Noticing that the variables $p_{\mu}^2$ are the roots of the polynomial
692: \begin{align*}
693: Q(t) = t^4-\sigma_1 t^3+\sigma_2 t^2-\sigma_3 t +\sigma_4
694: \end{align*}
695: the invariant polynomial $P_N$ can be written, after recursive
696: substitution of all fourth powers of the $p^2_{\mu}$'s, as a
697: polynomial $\widetilde{P}_N$ in the four symmetric invariants:
698: \begin{align*}
699: P_N(p_1,p_2,p_3,p_4) = \widetilde{P}_N(\sigma_1,\sigma_2,
700: \sigma_3,\sigma_4)\ .
701: \end{align*}
702:
703: We could have chosen other invariants to represent the polynomial,
704: as the power sums $p^{[n]} \equiv p_1^n+p_2^n+p_3^n+p_4^n$. They
705: can be indeed recovered from the symmetric invariants $\sigma_n$
706: via the recursive formulas:
707: \begin{align*}
708: \sigma_1 &= p^{2} \\
709: 2\sigma_2 &= \sigma_1p^{2} - p^{[4]} \\
710: 3\sigma_3 &= \sigma_2p^{2} - \sigma_1p^{[4]} + p^{[6]} \\
711: 4\sigma_4 &= \sigma_3p^{2} - \sigma_2p^{[4]} +
712: \sigma_1p^{[6]} - p^{[8]}\ .
713: \end{align*}
714: Thus, any polynomial on the four components of $p$ invariant under
715: the action of H(4) can be written as a polynomial in terms of the
716: power sums $p^{[n]}$. A complete, elegant proof can be found in
717: \cite{WEYL}.
718:
719:
720: \begin{thebibliography}{99}
721: \bibitem{CEL82}
722: W.~Celmaster,
723: %``Gauge Theories On The Body-Centered Hypercubic Lattice''
724: Phys.~Rev.~D{\bf 26} (1982) 2955.
725: \bibitem{LEE82}
726: N.H.~Christ, R.~Friedberg, and T.D.~Lee,
727: %``Gauge Theory On A Random Lattice''
728: Nucl.~Phys.~B{\bf 210} (1982) 310.
729: \bibitem{LPTX99}
730: D.~Becirevic, Ph.~Boucaud, J.P.~Leroy, J.~Micheli,
731: O.~P\`ene, J.~Rodr\'iguez-Quintero, C.~Roiesnel,
732: %``Asymptotic behaviour of the gluon propagator from lattice {QCD}''
733: Phys.~Rev.~D{\bf 60} (1999) 094509
734: [arXiv:hep-ph/9903364].
735: \bibitem{LPTX00}
736: D.~Becirevic, Ph.~Boucaud, J.P.~Leroy, J.~Micheli,
737: O.~P\`ene, J.~Rodr\'iguez-Quintero, C.~Roiesnel,
738: % ``Asymptotic scaling of the gluon propagator on the lattice''
739: Phys.~Rev.~D{\bf 61} (2000) 114508
740: [arXiv:hep-ph/9910204].
741: \bibitem{LPT03}
742: Ph.~Boucaud {\it et al.},
743: %``Quark propagator and vertex: Systematic corrections of hypercubic
744: % artifacts from lattice simulations''
745: Phys.~Lett.~B{\bf 575} (2003) 256
746: [arXiv:hep-lat/0307026].
747: \bibitem{LPTX05}
748: Ph.~Boucaud {\it et al.},
749: % Asymptotic behavior of the ghost propagator in SU3 lattice
750: % gauge theory"
751: Phys.~Rev.~D{\bf 72} (2005) 114503
752: [arXiv:hep-lat/0506031].
753: \bibitem{ADELAIDE}
754: D.B.~Leinweber, J.I.~Skullerud, A.G.~Williams and C.~Parrinello,
755: % ``Asymptotic scaling and infrared behavior of the gluon propagator''
756: Phys.~Rev.~D{\bf 60} (1999) 094507
757: [arXiv:hep-lat/9811027].
758: \bibitem{WEYL}
759: H.~Weyl,
760: {\it The Classical Groups},
761: Princeton University Press (1946).
762: \bibitem{NEU87}
763: H.~Neuberger,
764: %``Spinless Fields On F(4) lattice''
765: Phys.~Lett.~B{\bf 199} (1987) 536.
766: \bibitem{MICHAEL}
767: C. Michael,
768: % ''The Running Coupling from Lattice Gauge Theory''
769: Phys.~Lett.~B{\bf 283} (1992) 103
770: [arXiv:hep-lat/9205010].
771: \end{thebibliography}
772:
773: \end{document}
774: