0806.4586/rw.tex
1: %% LyX 1.5.3 created this file.  For more info, see http://www.lyx.org/.
2: %% Do not edit unless you really know what you are doing.
3: \documentclass[english,aps,preprint,nofootinbib]{revtex4}
4: \usepackage[T1]{fontenc}
5: \usepackage[latin1]{inputenc}
6: \usepackage{color}
7: \usepackage{graphicx}
8: \usepackage{amssymb}
9: 
10: \makeatletter
11: 
12: 
13: \linespread{1.13}
14: 
15: 
16: \newcommand{\bee}{\begin{equation}}
17: \newcommand{\ee}{\end{equation}}
18: \newcommand{\beea}{\begin{eqnarray}}
19: \newcommand{\eea}{\end{eqnarray}}
20: \newcommand{\gfive}{\gamma_5}
21: \newcommand{\sign}{{\rm sign}}
22: \newcommand{\rd}{{\rm d}}
23: \newcommand{\nn}{{\nonumber}}
24: \def\Tr{{\rm tr}}
25: \newcommand{\imu}{{\rm i}}
26: 
27: \usepackage{babel}
28: \makeatother
29: 
30: \begin{document}
31: 
32: \title{Low energy chiral constants from epsilon-regime simulations with
33: improved Wilson fermions}
34: 
35: 
36: \preprint{HU-EP-08/23;SFB/CPP-08-35}
37: 
38: 
39: \author{Anna Hasenfratz}
40: 
41: 
42: \email{anna@eotvos.colorado.edu}
43: 
44: 
45: \affiliation{Department of Physics, University of Colorado, Boulder, Colorado-80309-390,USA}
46: 
47: 
48: \author{Roland Hoffmann}
49: 
50: 
51: \email{hoffmann@pizero.colorado.edu}
52: 
53: 
54: \affiliation{Bergische Universit\"at Wuppertal, Gaussstra{\ss}e~20, 42219 Wuppertal,
55: Germany}
56: 
57: 
58: \author{Stefan Schaefer}
59: 
60: 
61: \email{sschaef@physik.hu-berlin.de}
62: 
63: 
64: \affiliation{Institut für Physik, Humboldt Universität, Newtonstra{\ss}e~15, 12489
65: Berlin, Germany}
66: 
67: \begin{abstract}
68: We present a lattice QCD calculation of the low-energy constants of
69: the leading order chiral Lagrangian. In these simulations the epsilon
70: regime is reached by using tree-level improved nHYP Wilson fermions
71: combined with reweighting in the quark mass. We analyze two point
72: functions on two ensembles with lattices of size $(1.85{\rm fm})^{4}$
73: and $(2.8{\rm fm})^{4}$, and at several quark mass values between
74: 4 and 20 MeV. The data are well fitted with next-to-leading order
75: chiral perturbative formulas and predict $F=90(4)$MeV and $\Sigma^{1/3}=248(6)$MeV
76: in the $\overline{{\rm MS}}$ scheme at 2 GeV. 
77: \end{abstract}
78: \maketitle
79: 
80: \section{Introduction}
81: 
82: At low energies, quantum chromodynamics can be described by a low-energy effective theory, chiral perturbation theory\cite{Gasser:1983yg},
83: $\chi$PT. To leading order, it has two {\it a priori} unknown parameters,
84: the chiral condensate and the pion decay constant. In this paper,
85: we compute these constants by an {\it ab initio} computation. Our strategy
86: itself is not new: we compute two-point functions in the epsilon regime
87: and fit them to their $\chi$PT~predictions. However, for chiral
88: perturbation theory to be an accurate description of low-energy phenomena,
89: we need calculations on the QCD side at very low pion masses, probably
90: lower than typically reached in current simulations, and at the same time on large 
91: enough volumes to control higher order corrections.
92: 
93: Basically, there are two regions in parameter space where calculation
94: of chiral perturbation theory can be carried out: one is the so-called
95: $p$-regime where we are essentially at infinite volume. The pion
96: wave length is much smaller than the size of the box, and the small
97: corrections due to its finite extent can be taken into account analytically.
98: 
99: The other region is the $\epsilon$-regime\cite{Gasser:1987ah}.
100: There the pion wave length is much larger than the spatial and temporal
101: size of the lattice. One can carry out the integrals over the (effectively)
102: constant pion modes exactly and arrive at a different power counting.
103: This regime has considerable appeal. First of all, even at next-to-leading order, only the two leading order low-energy constants of
104: chiral perturbation theory enter into observables\cite{Hansen:1990un,Hansen:1990yg}.
105: Second, being fundamentally a finite volume regime, once one has
106: reached it, lowering the pion mass actually improves the validity
107: of $\chi$PT at a given order, whereas in the $p$-regime one needs
108: to increase the volume at the same time.
109: 
110: The problem, however, lies in the task to actually reach this regime
111: in numerical simulations. Because the epsilon expansion is in powers
112: of $1/(FL)^{2}$, it turns out that we actually need a fairly big
113: volume and therefore a very small quark mass to satisfy the condition
114: $m_{\pi}\ll1/L$. This poses many algorithmical problems. In a recent
115: publication \cite{Hasenfratz:2008fg} we have presented a setup with
116: which we can actually reach the regime of small quark masses at moderate
117: cost avoiding many of the problems a more direct approach would face.
118: We propose to use smeared link improved Wilson fermions \cite{Hasenfratz:2007rf,Schaefer:2007dc}
119: to generate an ensemble of gauge configurations above the desired
120: quark mass. From this ensemble, we reweight to the quark mass which
121: we actually want to reach. Using this method, we manage to reduce
122: the sea quark mass by roughly a factor of 2 to 4.
123: 
124: Reweighting has another advantage apart from
125: the actual possibility to go to such light quarks. At very small quark
126: mass statistical fluctuations grow dramatically because the mass no
127: longer provides an infrared cutoff. Therefore, very small eigenvalues
128: of the Dirac operator can appear, which are suppressed by the fermion
129: determinant, but also lead to large values of the measured observables.
130: We thus have an anticorrelation between the weight and the function
131: value which means importance sampling breaks down. Reweighting avoids
132: this problem. The small eigenvalues are not as efficiently suppressed,
133: the region of large signal gets over-sampled and the error is actually
134: reduced. 
135: 
136: One might wonder how the explicit chiral symmetry
137: breaking of Wilson fermions influence simulations done with very light quarks in the epsilon regime.
138: Unlike in the $p$-regime, the finite volume of the $\epsilon$-regime provides 
139: the system  with an effective, dynamical IR cut-off. The chiral
140: symmetry violations will remain managable even in the chiral limit
141: as long as they are small compared to the inverse lattice size.
142: The numerical study of the continuum limit to prove this expectation will
143: be the subject of a future work.
144: 
145: 
146: In the current paper, we apply reweighting to a large volume data
147: set and provide measurements of the low-energy constants $F$ and
148: $\Sigma$. We already mentioned that this type of simulation is not
149: new. A number of quenched studies have been carried out \cite{Bietenholz:2003bj,Giusti:2003iq,Giusti:2004yp,Fukaya:2005yg,Bietenholz:2006fj,Giusti:2008fz}
150: using overlap fermions, which proved the feasibility of the method
151: but also highlighted the need for sufficiently large volume. Recently,
152: there also have been computations with dynamical overlap\cite{Fukaya:2007pn}
153: and twisted mass\cite{Jansen:2007rx} fermions to which we will compare
154: our results in the conclusion. For a recent review of these calculations
155: see Ref.~\cite{Necco:2007pr}. The strength of our calculation is
156: that we compare two different volumes, of which the larger one has
157: not been reached in previous studies. The simulations in this work
158: were fairly inexpensive, and it is within reach to repeat the calculation
159: at a finer lattice spacing to verify scaling.
160: 
161: The outline of this paper is the following: We first describe in Sec.~\ref{sec:lat}
162: the set-up of our simulations, the generation of the ensemble of gauge
163: configurations and the details of our reweighting procedure. The relevant
164: renormalization constants are computed in Sec.~\ref{sec:ren} using
165: the RI-MOM scheme. In Sec.~\ref{sec:eps} we collect the $\epsilon$-regime
166: formulas, the procedure of the extraction of the low-energy constants
167: and give the results.
168: 
169: 
170: \section{Simulations\label{sec:lat}}
171: 
172: The numerical simulations for this project were done with 2 flavors
173: of nHYP smeared Wilson-clover fermions and one-loop Symanzik improved
174: gauge action. The action and the simulation method are described in
175: details in Refs. \cite{Hasenfratz:2007rf,Schaefer:2007dc}. We
176: use tree-level $c_{SW}=1.0$ clover coefficient, so our action is
177: not fully $\mathcal{O}(a)$ improved. Based on our quenched investigation~\cite{Hoffmann:2007nm},
178: we expect that a nonperturbatively improved action would require
179: $c_{SW}\lesssim1.2$, so even with only tree-level improvement the
180: $\mathcal{O}(a)$ corrections are likely small. We have generated
181: two sets of gauge ensembles, both at gauge coupling $\beta=7.2$.
182: The first set consists of $16^{4}$ configurations at $\kappa=0.1278$,
183: the second $24^{4}$ configurations at $\kappa=0.12805$. We have
184: 180 and 154 thermalized configurations, separated by 5 trajectories
185: at the two volumes. The autocorrelation is 3-4 trajectories for the
186: plaquette, and about the same for the two point functions. Preliminary
187: results using the first set were already reported in Ref. \cite{Hasenfratz:2008fg}.
188: 
189: We set the lattice scale from the static quark potential, using $r_{0}=0.49$
190: fm for the Sommer parameter. On both configuration sets we found $r_{0}/a=4.25(2)$
191: (the error is from the larger volume set where we have a better signal
192: for the potential), giving $a=0.1153(5)$ fm. With this value the
193: physical volumes are (1.85 fm)$^{4}$ and (2.77 fm)$^{4}.$ Based
194: on the PCAC quark mass and the pseudoscalar and axialvector renormalization
195: factors (see Sec.~\ref{sec:ren}), we estimate the renormalized quark
196: mass in the $\overline{\rm MS}$ scheme at 2 GeV to be 22 and 8.5MeV, respectively. These values, and some other
197: details of the simulation, are listed in Table \ref{cap:Basics}.
198: %
199: \begin{table}
200: \begin{tabular}{|c|c|c|c|c|c|}
201: \hline 
202: $\kappa$  & $\kappa_{{\rm rew}}$  & $L$  & $N_{{\rm conf}}$  & $a\, m_{{\rm PCAC}}$  & $m${[}MeV]\tabularnewline
203: \hline
204: \hline 
205: 0.1278  & 0.1278  & 16  & 180  & 0.0117(3)  & 22\tabularnewline
206: \hline 
207:  & 0.1279  & 16  & 180  & 0.0088(5)  & 16.5\tabularnewline
208: \hline 
209:  & 0.1280  & 16  & 180  & 0.0058(7)  & 11\tabularnewline
210: \hline 
211:  & 0.12805  & 16  & 180  & 0.0047(8)  & 9\tabularnewline
212: \hline 
213:  & 0.1281  & 16  & 180  & 0.0028(11)  & 5\tabularnewline
214: \hline
215: \hline 
216: 0.12805  & 0.12805  & 24  & 154  & 0.0044(3)  & 8.5\tabularnewline
217: \hline 
218:  & 0.12810  & 24  & 154  & 0.0030(3)  & 5.8\tabularnewline
219: \hline 
220:  & 0.128125  & 24  & 154  & 0.0024(3)  & 4.2\tabularnewline
221: \hline 
222:  & 0.12815  & 24  & 154  & 0.0019(4)  & 3.8\tabularnewline
223: \hline
224: \end{tabular}
225: 
226: \caption{The parameters of the simulation. The first column gives the coupling
227: $\kappa$ of the dynamical simulation, the second the reweighted coupling
228: $\kappa_{{\rm rew}}$. The last column is the renormalized quark mass
229: using $m=m_{{\rm {PCAC}}}Z_{A}/Z_{P}$. \label{cap:Basics}}
230: 
231: \end{table}
232: 
233: 
234: The dynamical simulations were performed at particularly low quark
235: masses, even if we consider the relatively large volumes. This is
236: possible due to the highly improved chiral properties of the nHYP
237: smeared clover fermions. Figure \ref{cap:The-gap-distribution} shows
238: the histogram of the absolute value of the lowest Hermitian eigenmode
239: of the configurations. Simulations with Wilson-like fermions require
240: a well defined gap between zero and the first eigenmode \cite{DelDebbio:2006cn}.
241: As the figure shows, our simulations are safe on both volumes, though
242: very close to the low mass limit. In both cases, the median of the
243: distribution is about four times its width $\sigma$, which is above
244: the $3\sigma$ stability criterion of Ref.~\cite{DelDebbio:2006cn}.
245: However, it also shows that going lower in the quark mass can be very
246: dangerous because of the algorithm becoming unstable. The same paper
247: predicts $\sqrt{V}\sigma/a$ to be a scaling quantity. We measure
248: 0.56(4) and 0.77(5) for the $24^{4}$ and $16^{4}$ ensembles respectively.
249: 
250: %
251: \begin{figure}
252: \includegraphics[scale=0.4]{gap_all}
253: 
254: \caption{The Hermitian gap distribution on the original $16^{4}$ and $24^{4}$
255: configurations. The dashed lines correspond to $am_{{\rm PCAC}}$
256: . \label{cap:The-gap-distribution}}
257: 
258: \end{figure}
259: 
260: 
261: Starting from the original configurations one can explore a range
262: of quark masses in fully dynamical systems by reweighting the configurations.
263: In Ref.~\cite{Hasenfratz:2008fg} we have described an effective
264: technique to calculate the necessary weight factors. It is a stochastic
265: calculation, and one must take care not to introduce significant statistical
266: errors with the stochastic process. We apply three methods, low mode
267: separation, determinant breakup, and ultraviolet (UV) noise reduction
268: to control the statistical fluctuations. In both the $16^{4}$ and
269: $24^{4}$ ensembles we separate 6 low Hermitian eigenmodes. In addition
270: we break up the determinant to the product of 33 and 60 terms for
271: each $\Delta\kappa=0.0001$ shift in reweighting on the $16^{4}$
272: and $24^{4}$ volumes, respectively. To control and remove some of
273: the UV noise we introduce a pure gauge term in the reweighted action
274: . This term is just an nHYP plaquette term and has a very small coefficient.
275: We found that in our system it can be chosen to be proportional to
276: the shift in $\kappa$, \begin{equation}
277: \beta_{{\rm nHYP}}=6.0(\kappa-\kappa_{{\rm rew}})\:\label{eq:beta-nhyp}\end{equation}
278:  on both the $16^{4}$ and $24^{4}$ configuration sets. This value
279: is so small that there is no difference within the errors in the lattice
280: spacings or quark masses between $\beta_{{\rm nHYP}}=0$ and Equation \ref{eq:beta-nhyp}.
281: For further details and the exact definition of the reweighting action
282: we refer to Ref. \cite{Hasenfratz:2008fg}, especially Equation 17 and
283: Figure 4.
284: 
285: %
286: \begin{figure}
287: 
288: 
289: \includegraphics[scale=0.8]{overlap}
290: 
291: \caption{The distribution of the pseudoscalar correlator at $t=8$ on the $16^{4}$
292: ensemble at $\kappa=0.1280$. On both panels the dashed line is the
293: partially quenched distribution and the solid lines correspond to the
294: reweighted distributions. Left panel: $\beta_{{\rm nHYP}}=0$, right
295: panel: $\beta_{{\rm nHYP}}$ as in Equation \ref{eq:beta-nhyp}.\label{fig:overlap}}
296: 
297: \end{figure}
298: In addition to removing the UV fluctuations, the introduction
299: of the nHYP plaquette term also increases the overlap between the
300: original and target ensembles. The largest weight factors are pushed
301: from the edge of the plaquette distribution to the middle, where the
302: statistical sampling is better, and the effect is similar for other
303: observables as well. We illustrate this in Figure \ref{fig:overlap}
304: with the distribution of the pseudoscalar correlator at $t=8$. All
305: data correspond to the $16^{4}$ data set at $\kappa=0.1280$. The
306: left panel shows the reweighted distribution without the nHYP plaquette
307: term, the right one with $\beta_{{\rm nHYP}}$ as in Equation \ref{eq:beta-nhyp}.
308: For reference both panels show the partially quenched (unweighted)
309: distribution. In both cases the overlap between the reweighted and
310: partially quenched distributions is excellent, there is no sign that
311: reweighting would prefer region that is poorly sampled by the original
312: ensemble. The main difference between the reweighted and partially
313: quenched data is the suppression of the long tail of the latter one,
314: a quenching artifact. The apparent increase of the reweighted distributions
315: is mainly due to normalization: the dynamical distribution is narrower,
316: resulting in a higher peak. 
317: It is worthwhile to emphasize that including the nHYP plaquette term does not 
318: introduce any systematic error, rather it improves the overlap between the ensembles, 
319: especially at larger mass difference.
320: 
321: Even though there is no strong difference between
322: the two panels of Figure \ref{fig:overlap}, the introduction of
323: the nHYP plaquette term reduces the statistical errors by up to 40\%
324: for our lightest $16^{4}$ data set, and the results we present in
325: the following reflect that. On the other hand, on the $24^{4}$ volumes
326: within the $\kappa$ range we reweight to there is no difference between
327: the actions with or without the nHYP plaquette term, and the results
328: we present here were obtained with $\beta_{{\rm nHYP}}=0$ .
329: 
330: With reweighting it is possible to reach an eigenvalue that is negative
331: or at least smaller than the typical eigenvalues of the Dirac operator.
332: We approach that case with our last coupling on the $16^{4}$ ensemble
333: where at least one configurations has a negative real Dirac eigenvalue
334: and several has nearly zero eigenvalues, and even more on the lightest
335: reweighted $24^{4}$ configurations where 5\% of the configurations
336: have a negative Dirac eigenvalue and even more have a nearly zero
337: one. %
338: \begin{figure}
339: \includegraphics[scale=0.8]{gap_distr}
340: 
341: \caption{The Hermitian gap distribution on the original and lightest reweighted
342: ensembles for both volumes. The histograms are labeled by the corresponding
343: lattice quark mass which is also indicated by the dashed lines. \label{fig:gap_distr} }
344: 
345: \end{figure}
346: These configurations are suppressed by the weight factor, nevertheless
347: as we will see later one encounters increased statistical errors on
348: these ensembles. In Figure \ref{fig:gap_distr} we compare the Hermitian
349: gap distribution on the original and lightest reweighted ensembles
350: for both volumes. The distribution shifts towards zero but configurations
351: with negative or near-zero eigenmodes are strongly suppressed, and that
352: maintains  the  gap. While the PCAC quark mass approches zero, the median of
353: the gap, controlled by the finite volume,  remains finite. 
354: 
355: 
356: \section{Renormalization factors\label{sec:ren}}
357: 
358: In order to connect the lattice meson correlators to the physical
359: ones we have to determine the corresponding renormalization factors.
360: We used the standard RI-MOM method \cite{Martinelli:1994ty}, where
361: one calculates bilinear quark operators $\langle p|O_{\Gamma}|p\rangle$
362: at specific lattice momentum $p^{2}=\mu^{2}$ and matches them to
363: the corresponding tree-level matrix element. Afterwards the lattice
364: values are connected to the continuum $\overline{{\rm MS}}$ scheme
365: perturbatively \cite{Gimenez:1998ue,Chetyrkin:1999pq}. The renormalization
366: scale $\mu$ has to be much smaller than the lattice cut-off to minimize
367: lattice artifacts but much larger than the QCD scale for continuum
368: perturbation theory to work.
369: 
370: Our code to calculate the lattice matrix elements is based on the
371: one used in Ref. \cite{DeGrand:2005af}. We used 80 propagators from
372: the $16^{4}$ data set to calculate the vector, axialvector, scalar
373: and pseudoscalar matching factors in the chiral limit. While most
374: dynamical calculations do the chiral extrapolations on partial quenched
375: data \cite{Becirevic:2005ta},\cite{Dimopoulos:2007fn}, we can
376: do this extrapolation on fully dynamical configurations on our reweighted
377: ensembles. We used 5 $\kappa$ values, 0.1278, 0.12785, 0.1279, 0.12795
378: and 0.1280, corresponding to quark masses 10-20 MeV. We extrapolated
379: the vector, axial and scalar data linearly in the quark mass, though
380: the data shows no mass dependence within errors. This is not that
381: surprising, since our quark masses are light. The pseudoscalar density
382: couples to the Goldstone boson channel and it develops an $\mathcal{O}(1/m)$
383: singularity in the chiral limit. We subtract this pole assuming a
384: linear mass dependence for the quantity $m/Z_{P}$. Again, with our
385: light mass values we expect this assumption to hold, and our data
386: are indeed consistent with a linear dependence \cite{Becirevic:2004ny,Gattringer:2004iv}.
387: Nevertheless the subtraction introduces fairly large errors at small
388: $\mu$.
389: 
390: %
391: \begin{figure}
392: \includegraphics[scale=0.8]{Z_all2}
393: 
394: \caption{The renormalization factors for the vector, axialvector, pseudoscalar
395: and scalar operators as the function of the lattice energy scale.
396: All values are converted to the continuum $\overline{{\rm MS}}$ scheme
397: at $\mu=2$GeV. \label{cap:The-renormalization-factors}}
398: 
399: \end{figure}
400: 
401: 
402: In Figure \ref{cap:The-renormalization-factors} we show all four
403: renormalization factors converted to the $\overline{{\rm MS}}$ scheme
404: at 2GeV, as the function of the original lattice momentum $p^{2}=\mu^{2}$.
405: The vector and axialvector factors are scale independent, any deviation
406: from a constant is due to lattice artifacts. In our case $Z_{V}$
407: is constant over the whole range, while $Z_{A}$ shows a slight drift
408: at larger $\mu$ values. Calculations with Wilson and Wilson-like
409: improved fermions show similar trends for these quantities \cite{Becirevic:2004ny,Becirevic:2005ta,Gattringer:2004iv}.
410: 
411: The scalar and pseudoscalar operators depend on the energy scale.
412: We connect the lattice data to the continuum one at identical energy,
413: then, using the known 3-loop expression for the running of the coupling,
414: run it to $\mu=2$GeV . We plot these values, therefore $Z_{P}$ and
415: $Z_{S}$ in Figure \ref{cap:The-renormalization-factors} should also
416: be constant. Because of the subtraction of the Goldstone pole the errors
417: are large at small $\mu$ for the pseudoscalar, but we find a long,
418: stable plateau at larger lattice scale values. The scalar operator,
419: on the other hand, shows quite large lattice artifacts at higher scales.
420: Again, this trend has been observed before with other actions. The
421: horizontal lines in Figure \ref{cap:The-renormalization-factors}
422: indicate the range where we extract the renormalization factors. Our
423: final values are \begin{eqnarray}
424: Z_{V}^{\overline{{\rm MS}}} & = & 0.96(1)\nonumber \\
425: Z_{A}^{\overline{{\rm MS}}} & = & 0.99(2)\nonumber \\
426: Z_{P}^{\overline{{\rm MS}}}(2{\rm GeV}) & = & 0.90(2)\label{eq:Z_factors}\\
427: Z_{S}^{\overline{{\rm MS}}}(2{\rm GeV}) & = & 1.01(3)\,.\nonumber \end{eqnarray}
428:  As a simple check we compare the renormalized quark mass as predicted
429: form the bare quark mass $m_{}=Z_{S}^{-1}m_{b}$, $1/(2\kappa)-1/(2\kappa_{cr})$,
430: and from the PCAC mass $m_{r}=Z_{A}Z_{P}^{-1}m_{{\rm PCAC}}$. Fitting
431: $m_{{\rm PCAC}}$ linearly in $1/(2\kappa)$ we predict $\kappa_{cr}=0.12821$
432: and from the slope $Z_{P}Z_{S}^{-1}Z_{A}^{-1}=0.94(3)$. This is consistent
433: from the value obtained from Equation \ref{eq:Z_factors}, 0.90(7). The
434: fact that all four matching factors are close to one indicates small
435: perturbative corrections, as it is usually seen with smeared link
436: actions.
437: 
438: 
439: \section{$\epsilon$-regime analysis\label{sec:eps}}
440: 
441: In the $\epsilon$-regime the pion correlation length is large compared
442: to the linear size of the lattice, the light pseudoscalar mesons
443: dominate the dynamics. Nevertheless, in order to incorporate the massive
444: modes the volume has to be large compared to the QCD scale. One assumes
445: that the quark mass is $m=\mathcal{O}(\epsilon^{4})$ and the inverse
446: size $1/L=\mathcal{O}(\epsilon)$ ( $L^{4}=V=L_{s}^{3}L_{t}$), but
447: $1/L\gg\Lambda_{{\rm QCD}}$. The dimensionless quantity $m\Sigma V$,
448: or equivalently $m_{\pi}^{2}F_{\pi}^{2}V$, is kept order one. Chiral
449: perturbation theory predictions are organized in power of $\epsilon^{2}$
450: or $1/(FL)^{2}$. Predictions for various meson correlators are known
451: to next-to-leading (NLO) order ($\mathcal{O}(\epsilon^{4})$), except
452: for the pseudoscalar that has been calculated up to $\mathcal{O}(\epsilon^{6})$.
453: In our fits we use the NLO predictions as those depends only on two
454: low-energy constants, $\Sigma=\lim_{m\to0}\langle\bar{q}q\rangle$
455: and $F=\lim_{m\to0}F_{\pi}$\cite{Hasenfratz:1989pk,Hansen:1990un,Hansen:1990yg}.
456: These $\chi$PT results are based on a chiral (continuum) action.
457: One expects extra terms, due to the explicit chiral symmetry violation
458: of the Wilson fermion action, in our situation. However these corrections
459: typically show up at the same order as the higher order chiral constants
460: $L_{3},$$L_{4}$, i.e. only at next-to-next-to-leading order in the
461: epsilon regime.
462: 
463: As the quark mass decreases in a large volume (p-regime)
464: simulation, the chiral symmetry breaking effects of Wilson fermions
465: get large compared to the mass, and that can create large lattice
466: artifacts. In practice the continuum limit has to be taken
467: before the chiral limit. The situation is different in the $\epsilon$-regime, where
468: the finite volume of the system creates an infrared cutoff even at
469: vanishing quark mass. This effect is well illustrated by the Hermitian
470: gap distribution in Figure \ref{fig:gap_distr}. While in infinite
471: volume one expects the median of the gap to scale with the mass $\bar{\mu}=Z_{A}m_{{\rm PCAC}}$
472: \cite{DelDebbio:2005qa}, in the $\epsilon-$ regime $\bar{\mu}$,
473: governed by the IR cutoff of the volume, remains finite while $m_{{\rm PCAC}}\to0$.
474: This is clearly the case in our simulations. Therefore one does not need a chiral
475: action to study the epsilon regime, though the explicit symmetry breaking
476: effects should be small compared to the inverse lattice size. As long
477: as the volume is large enough that the NLO relations describe the
478: two-point functions, continuum $\chi$PT results can be used to analyze
479: Wilson fermion data. Since separating topological sectors with Wilson
480: fermions is not always possible, we analyze our data averaged over
481: the topological charge.
482: 
483: For completeness we give the relevant formulas for two degenerate
484: flavors, averaged over the topological charge. The isotriplet pseudoscalar meson correlator up to $\mathcal{O}(\epsilon^{4})$ is \begin{eqnarray}
485: \Gamma_{P}(t) & = & \frac{1}{L_{s}^{3}}\int d^{3}x\langle P(x)P(0)\rangle\nonumber \\
486:  & = & \Sigma^{2}\Big(a_{p}+\frac{L_{t}}{F^{2}L_{s}^{3}}b_{p}h_{1}(\frac{t}{L_{t}})+\mathcal{O}(\epsilon^{4})\Big)\,,\label{eq:eps-pion-corr}\end{eqnarray}
487:  where $P(x,t)=\bar{\psi}\frac{1}{2}\lambda^{i}\gamma_{5}\psi$ is
488: the pseudoscalar density operator and \begin{eqnarray}
489: a_{p} & = & \frac{\rho}{8}I_{1}(u)\,.\label{eq:chipt_coeff}\\
490: b_{p} & = & 1-\frac{1}{8}I_{1}(u)\,,\nonumber \end{eqnarray}
491:  with \[
492: \rho=1+\frac{3\beta_{1}}{2(FL)^{2}}\]
493:  the shape factor ($\beta_{1}=0.14046$ for our symmetric geometry
494: ) and \[
495: u=2\, m\Sigma V\,\rho\,.\]
496:  $I_{1}$ can be expressed in terms of Bessel functions, $I_{1}(u)=8Y'(u)/(uY(u))$.
497: It decreases smoothly from 2 at $u=0$ to 0.68 at $u=10$, the largest
498: value we encounter. The function \begin{equation}
499: h_{1}(\tau)=\frac{1}{2}\big[(\tau-\frac{1}{2})^{2}-\frac{1}{12}\big]\label{eq:h1}\end{equation}
500:  describes the quadratic time dependence. The pseudoscalar correlator
501: is dominated by $\Sigma$, the dependence on $F$ is only through
502: the $\mathcal{O}(\epsilon^{2})$ term.
503: 
504: The flavor triplet axialvector current correlator at NLO is \begin{eqnarray}
505: \Gamma_{A}(t) & = & \frac{1}{L_{s}^{3}}\int d^{3}x\langle A_{0}(x)A_{0}(0)\rangle\nonumber \\
506:  & = & \frac{F^{2}}{V}\Big(a_{a}+\frac{L_{t}}{F^{2}L_{s}^{3}}b_{a}h_{1}\Big(\frac{t}{L_{t}}\Big)+\mathcal{O}(\epsilon^{4})\Big)\,,\label{eq:eps-axial-corr}\end{eqnarray}
507:  with $A_{0}(x,t)=\bar{\psi}\frac{1}{2}\lambda^{i}\gamma_{0}\gamma_{5}\psi$
508: and \begin{eqnarray}
509: a_{a} & = & 1-\frac{1}{4}I_{1}(u)+\frac{\beta_{1}}{(FL)^{2}}\Big(2-\frac{1}{2}I_{1}(u)\Big)-\frac{L_{t}}{F^{2}L_{s}^{3}}\frac{k_{00}}{2}I_{1}(u)\,,\nonumber \\
510: b_{a} & = & \frac{1}{8}u^{2}I_{1}(u)\,.\label{eq:axial_coeff}\end{eqnarray}
511:  In our $L_{s}=L_{t}$ case $k_{00}=\beta_{1}/2$. The axialvector
512: correlator is dominated by $F,$ the dependence on $\Sigma$ enters
513: only through the combination $m\Sigma V$.
514: 
515: The $\epsilon$-expansion formulas are systematic expansions in the
516: parameter $1/(FL)^{2}=\mathcal{O}(\epsilon^{2})$, but depend on the
517: $\mathcal{O}(1)$ quantity $m\Sigma V$. In our simulation we explore
518: the range $m\Sigma V\approx0.7\,-\,5.0$. Large values introduce large
519: NLO and NNLO corrections to the correlators, and at some point one
520: transitions into the large volume $p$-regime. Only by examining the
521: fit results will we be able to decide what range of $m\Sigma V$ values
522: are acceptable in the $\epsilon$-regime.
523: 
524: The lattice correlators have to be multiplied by the renormalization
525: factors $Z_{P}^{2}$ and $Z_{A}^{2}$ to obtain the continuum ones
526: in Equations \ref{eq:eps-pion-corr} and \ref{eq:eps-axial-corr}, while
527: in the product $m\Sigma V$ the quark mass can be expressed in terms
528: of the PCAC mass as $m=m_{{\rm PCAC}}Z_{A}/Z_{P}$. In our fit we
529: use the combinations \begin{equation}
530: \Gamma_{A}=Z_{A}^{2}\Gamma_{A}^{({\rm latt})}\label{eq:axial}\end{equation}
531:  and \begin{equation}
532: m^{2}\Gamma_{P}=Z_{A}^{2}m_{{\rm {PCAC}}}^{2}\Gamma_{P}^{({\rm latt})}\label{eq:pseudo}\end{equation}
533:  that depend only on $F$ and $m\Sigma V$, and do a combined fit
534: to Equations \ref{eq:axial} and \ref{eq:pseudo}. %
535: \begin{figure}
536: \includegraphics[scale=0.7]{plot_16_1}
537: 
538: \caption{The pseudoscalar ( red diamonds) and axialvector (blue bursts) lattice
539: correlators and the combined fit results at $\kappa=0.1278$ and $0.1279$
540: on the $16^{4}$ data set. The axialvector correlators are multiplied
541: by the factor 50 to better match the scale of the pseudoscalar.\label{cap:corr-16-1}}
542: 
543: \end{figure}
544: 
545: 
546: %
547: \begin{figure}
548: \includegraphics[scale=0.7]{plot_16_2}
549: 
550: \caption{Same as Figure \ref{cap:corr-16-1} but at $\kappa=0.1280$ and $0.1281$.\label{cap:corr-16-2}}
551: 
552: \end{figure}
553: 
554: 
555: %
556: \begin{figure}
557: \includegraphics[scale=0.7]{plot_24_1}
558: 
559: \caption{Same as Figure \ref{cap:corr-16-1} but on the $24^{4}$ data set
560: at $\kappa=0.12805$ and $0.1281$.\label{cap:corr-24-1}}
561: 
562: \end{figure}
563: 
564: 
565: %
566: \begin{figure}
567: \includegraphics[scale=0.7]{plot_24_2}
568: 
569: \caption{Same as Figure \ref{cap:corr-16-1} but on the $24^{4}$ data set
570: at $\kappa=0.128125$ and $0.12815$.\label{cap:corr-24-2}}
571: 
572: \end{figure}
573: 
574: 
575: The results of the combined fits on the $16^{4}$ lattices are shown
576: in Figures \ref{cap:corr-16-1} and \ref{cap:corr-16-2}, where we
577: plot both the pseudoscalar and axialvector correlators (the latter
578: is rescaled by a factor 50 to match the scale). We use the time slices
579: $[5,11]$ in the fit. The data are well described by the NLO formulas
580: at all four mass values. The results are summarized in Table \ref{cap:Results}
581: where we list not only the predicted low-energy parameters but the
582: combination $m\Sigma V$ and an estimate for $m_{\pi}L$ as well.
583: We estimate the infinite volume pion mass using the GMOR relation
584: $m_{\pi}^{2}=N_{f}m\Sigma/F^{2}+\mathcal{O}(m^{2})$. In the $\epsilon$-regime 
585: one requires $m_{\pi}L\ll1$ , though according to the analysis
586: of Ref. \cite{Hansen:1990yg}, the $\epsilon$- and $p$-regimes
587: connect smoothly around $m_{\pi}L\approx2$, so values up to that
588: level are also acceptable. While at $\kappa=0.1278$ both $m\Sigma V$
589: and $m_{\pi}L$ are somewhat large, the fit indicates that the data
590: are described well by the $\epsilon$-regime forms at all $\kappa$
591: values. The predicted low-energy constants, especially F, show a slight
592: drift as $m$ decreases, indicating that higher order effects are
593: nevertheless not negligible. Considering that the expansion parameter
594: $1/(FL)^{2}\approx1.45$ is not at all small, this is quite possible.
595: The $\mathcal{O}(\epsilon^{2})$ corrections to the pseudoscalar correlator
596: at $t=N_{t}/2$ in Equation \ref{eq:chipt_coeff} are 34\% at $\kappa=0.1278$,
597: decreasing to 16\% at $\kappa=0.1281$, while the constant term $a_{a}$
598: of the axial correlator in Equation \ref{eq:axial_coeff} has 30-25\% corrections
599: in the same range. A volume of (1.85fm)$^{4}$ is not large enough
600: to suppress finite volume effects. %
601: \begin{table}
602: \begin{tabular}{|c|c|c|c|c|c|}
603: \hline 
604: $\kappa$  & $L$  & $m\Sigma V$  & $m_{\pi}L$ & $F${[}MeV]  & $\Sigma^{1/3}${[}MeV]\tabularnewline
605: \hline
606: \hline 
607: 0.1278  & 16  & 3.1(2)  & 3.14 & 90(3)  & 256(6)\tabularnewline
608: \hline 
609: 0.1279  & 16  & 2.1(1)  & 2.58 & 86(4)  & 254(6)\tabularnewline
610: \hline 
611: 0.1280  & 16  & 1.4(1)  & 2.11 & 83(6)  & 252(7)\tabularnewline
612: \hline 
613: 0.12805  & 16  & 1.0(1)  & 1.78 & 82(7)  & 250(7)\tabularnewline
614: \hline 
615: 0.1281  & 16  & 0.68(5)  & 1.47 & 76(10)  & 251(7)\tabularnewline
616: \hline
617: \hline 
618: 0.12805  & 24  & 5.2(3)  & 2.71 & 90(3)  & 248(6)\tabularnewline
619: \hline 
620: 0.12810  & 24  & 3.4(2)  & 2.19 & 89(4)  & 250(6)\tabularnewline
621: \hline 
622: 0.128125  & 24  & 2.6(1)  & 1.91 & 89(6)  & 248(6)\tabularnewline
623: \hline 
624: 0.12815  & 24  & 2.3(1)  & 1.80 & 92(8)  & 245(8)\tabularnewline
625: \hline
626: \end{tabular}
627: 
628: \caption{Results from the combined fit to the pseudoscalar and axialvector
629: correlators. The values of $F$ and $\Sigma$ are converted to physical
630: units using $r_{0}=0.49$fm. The combination $m\Sigma V$ is predicted
631: by the fit while for  $m_{\pi}L$ we estimate the infinite volume
632: pion mass from the GMOR relation. \label{cap:Results}}
633: 
634: \end{table}
635: 
636: 
637: Our second data set is $24^{4},$ (2.77fm)$^{4}$, considerably larger.
638: The $\mathcal{O}(\epsilon^{2})$ corrections are reduced to $\approx15$\%
639: for the axial correlator, though the corrections are still large,
640: 10-25\% for the pseudoscalar correlator at our mass values. Figures
641: \ref{cap:corr-24-1} and \ref{cap:corr-24-2} show the result of the
642: combined fit. Again, we find good agreement for all correlators in
643: the range of $5\le t\le19,$ though we use only time slices $[8,16]$
644: in the fit. The statistical errors are under control everywhere, though
645: they increase as the reweighting range increases .
646: 
647: The data points for $t<5$ and $t>L_{t}/a-5$ do not follow the $\epsilon$-regime
648: $\chi$PT predictions. A natural explanation is that the heavy excitations
649: that couple to the operators die out only at $t\ge5$ and influence
650: the correlators at small distances. If that is indeed the case, the
651: correlators should show similar behavior on the $16^{4}$ and $24^{4}$
652: sets. Indeed, at $\kappa=0.12805$ and $\kappa=0.1281$, where we
653: have results on both volumes, both the pseudoscalar and axialvector
654: correlators are identical within errors for $t<5$, showing the same
655: transient behavior. It is somewhat puzzling why a recent result using
656: overlap fermions at similar lattice spacing and even smaller quark
657: masses see transient behavior in the pseudoscalar channel up to $t\approx12$
658: \cite{Fukaya:2007pn}. It might be due to the small spatial extent
659: ($L_{s}=16$) or the asymmetric geometry ($L_{t}=32$) used in Ref.
660: \cite{Fukaya:2007pn}, or that the overlap operator is more extended
661: and excited states die out slower.
662: 
663: We also have measurements of the vector current correlator. The data
664: and the fit quality are similar to the axialvector we presented above.
665: Since it does not improve the determination of the low-energy constants,
666: we do not include it in our analysis.
667: 
668: One might ask if the data, especially on the $16^{4}$
669: volumes, could be better fitted with NNLO, $\mathcal{O}(\epsilon^{6})$
670: forms. For the pseudoscalar these two-loop results are known in the
671: continuum \cite{Hasenfratz:1989pk}, and contain two new low-energy
672: constants, $L_{3}$ and $L_{4}$. On the lattice additional terms,
673: due to the chiral symmetry breaking of the Wilson action, also enter.
674: For the axialvector correlator only NLO results are available. Considering
675: the number of unknown parameters, it is not obvious that  a meaningful fit
676: could be done even if full NNLO formulas were available, though it would
677: be very interesting to test.
678: 
679: We close this section with a combined plot of the low-energy constants
680: obtained on the two volumes, as the function of the parameter $m\Sigma V$
681: (Figure \ref{fig:Summary-plot}). %
682: \begin{figure}
683: \includegraphics[scale=0.8]{summary}
684: 
685: \caption{The low-energy constant $F$ and $\Sigma^{1/3}$ as the function of
686: the parameter $m\Sigma V$, predicted by NLO $\chi$PT. \label{fig:Summary-plot}}
687: 
688: \end{figure}
689: 
690: 
691: As is evident both from Figure \ref{fig:Summary-plot} and Table \ref{cap:Results},
692: the different quark mass data on the $24^{4}$ ensemble are consistent
693: for both low-energy constants and the results for the chiral condensate
694: are consistent on the two volumes. $F$, on the other hand, shows
695: a drift as $m\Sigma V$ decreases. Without a large volume data point
696: at $m\Sigma V<2$ we cannot tell if this is due to finite volume effects,
697: or signals the breakdown of the $\epsilon$ expansion for $m\Sigma V>2$.
698: $\chi$PT formulas that connect the $\epsilon$ and $p$ regimes could
699: help to decide this issue. Since the next-to-leading order corrections
700: to $F$ are over 10\% on the $16^{4}$ data set, we prefer using the
701: large volume data to arrive at our final prediction, \begin{eqnarray}
702: F=90(4){\rm MeV,} &  & \Sigma^{1/3}=248(6){\rm MeV}\nonumber \\
703: Fr_{0}=0.224(10), &  & \Sigma^{1/3}r_{0}=0.617(15)\,.\label{eq:results}\end{eqnarray}
704:  The errors only include the statistical uncertainties.
705: 
706: Let us finally compare our results to other recent two flavor computations,
707: even though direct comparisons are problematic due to different systematic
708: errors. In the $p$-regime with maximally twisted mass fermions, the
709: ETM collaboration gets in the continuum limit $Fr_{0}=0.188(2)(7)$
710: and $\Sigma^{1/3}r_{0}=0.597(9)(15)$ and a compatible number for
711: $\Sigma$ in the $\epsilon$ regime~\cite{Dimopoulos:2007qy,Jansen:2007rx}%
712: \footnote{The number for $\Sigma$ is not explicitly given; the quoted errors
713: are obtained using $r_{0}=0.433$fm%
714: }. Another $\epsilon$ regime computation has been performed by JLQCD
715: with dynamical overlap fermions at fixed topology in a $L^{3}\times2L$,
716: $L=1.8{\rm fm}$ box~\cite{Fukaya:2007pn}. They get $Fr_{0}=0.217(14)$
717: and $\Sigma^{1/3}r_{0}=0.596(10)$. Given the statistical and systematic
718: errors, these results nicely agree with our determination.
719: 
720: Another method of extracting the low-energy constant is by looking
721: at the distribution of the lowest eigenvalue of the Dirac operator
722: and comparing it to predictions from random matrix theory. To our
723: knowledge, there are two such results with renormalized $N_{f}=2$
724: results. In Ref.~\cite{Fukaya:2007yv}, JLQCD compute $r_{0}\Sigma^{1/3}=0.624(17)(27)$
725: using these methods. Ref.~\cite{DeGrand:2007tm} finds $r_{0}F=0.213(11)$
726: and $r_{0}\Sigma^{1/3}=0.594(13)$ using nHYP link dynamical overlap
727: fermions. Again, there is good agreement to our findings.
728: 
729: 
730: \section{Conclusion}
731: 
732: The data presented in this paper has been generated with moderate
733: computer resources. This was possible due to the good chiral properties
734: of the action which come at relatively low cost due to the simple
735: nHYP smearing procedure, and the effective reweighting that allowed
736: us to lower the quark mass even further. Obviously, there are still
737: shortcomings of our analysis. The simulation is done at just one lattice
738: spacing, so we are not able to take our results to the continuum limit.
739: However, HYP smearing has improved the scaling properties of a variety
740: of actions. We are confident that also here cut-off effects will
741: be small.
742: 
743: Moreover, setting the scale by $r_{0}=0.49$fm is not satisfactory.
744: The value of $r_{0}$ is not known to high accuracy. We could use
745: $F$ as scale parameter. Apart from that, the $\epsilon$ regime setup
746: makes it problematic to use other widely used scales like the mass
747: of the $\Omega$ baryon.
748: 
749: We still are at finite lattice size and $\epsilon$-regime $\chi$PT
750: is a slowly converging expansion in $1/(FL)^{2}$. Here, our large
751: volume puts us into a good position and the comparison between the
752: $L/a=16$ and $L/a=24$ results shows that the finite volume effects
753: are under control. However, statistical errors due to the limited
754: number of gauge configurations are too large for a more substantiated
755: claim.
756: 
757: For the $\epsilon$ expansion to be valid, the parameter $m\Sigma V$
758: has to be $\mathcal{O}(1)$. Our data span the range 0.7 to 5.2 and
759: might go beyond the validity of the analytical expressions. An expansion
760: that connects the $\epsilon$ and $p$ regimes would be very useful
761: to control this aspect of the calculation.
762: 
763: Nevertheless our results are encouraging. We find that reweighting
764: works on a fairly large volume of $(L/a)^{4}=24^{4}$, $L\approx2.8$fm,
765: and the statistical fluctuations are under control despite quark masses
766: as low as $4$MeV. Repeating the calculation at a smaller lattice
767: spacing would not be prohibitively expensive and could improve on
768: all of the above mentioned issues.
769: 
770: 
771: \section{Acknowledgement}
772: 
773: We have benefited from discussion with O. B\"ar,
774: G. Colangelo, T. DeGrand, P. Hasenfratz, T. Izubuchi, S. Necco, and
775: F. Niedermayer. Most of the numerical work reported in this paper
776: was carried out at the kaon cluster at FNAL. We acknowledge the support
777: of the USQCD/SciDac collaboration.
778: 
779: The renormalization constants calculation was done based on the code
780: developed by T. DeGrand and Z-f. Liu. We are grateful for the permission
781: to use it. This research was partially supported by the US Department
782: of Energy and the Deutsche Forschungsgemeinschaft in the SFB/TR 09.
783: 
784: \bibliographystyle{apsrev} \bibliographystyle{apsrev} \bibliographystyle{apsrev}
785: \bibliography{lattice}
786: 
787: \end{document}
788: