0807.1661/ms.tex
1: % ****** Start of file apssamp.tex ******
2: %
3: %   This file is part of the APS files in the REVTeX 4 distribution.
4: %   Version 4.0 of REVTeX, August 2001
5: %
6: %   Copyright (c) 2001 The American Physical Society.
7: %
8: %   See the REVTeX 4 README file for restrictions and more information.
9: %
10: % TeX'ing this file requires that you have AMS-LaTeX 2.0 installed
11: % as well as the rest of the prerequisites for REVTeX 4.0
12: %
13: % See the REVTeX 4 README file
14: % It also requires running BibTeX. The commands are as follows:
15: %
16: %  1)  latex apssamp.tex
17: %  2)  bibtex apssamp
18: %  3)  latex apssamp.tex
19: %  4)  latex apssamp.tex
20: %
21: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
22: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
23: 
24: % Some other (several out of many) possibilities
25: %\documentclass[preprint,aps]{revtex4}
26: %\documentclass[preprint,aps,draft]{revtex4}
27: %\documentclass[prb]{revtex4}% Physical Review B
28: 
29: \usepackage{graphicx}% Include figure files
30: \usepackage{dcolumn}% Align table columns on decimal point
31: \usepackage{bm}% bold math
32: \usepackage{rotating}
33: 
34: %\nofiles
35: 
36: \voffset=+25mm
37: 
38: %%%%% Personal Macros %%%%%%%%%%%%%%%%%%%
39: %\newcommand{\nn}{\nonumber}
40: \def\dfrac#1#2{\displaystyle\frac{#1}{#2}}
41: %\newcommand{\bm}[1]{\mbox{\boldmath $#1$}}
42: \newcommand{\ovl}[1]{\overline{#1}}
43: \newcommand{\wt}[1]{\widetilde{#1}}
44: \newcommand{\eq}[1]{eq.(\ref{#1})}
45: \newcommand{\eqn}[1]{(\ref{#1})}
46: \newcommand{\p}{\partial}
47: \newcommand{\slas}[2]{{{#1}\hspace{-5pt}{/}}_{#2}}
48: \newcommand{\slal}[2]{{{#1}\hspace{-8pt}{/}}_{#2}}
49: \newcommand{\kslash}{k\kern-1ex /}
50: \newcommand{\pslash}{p\kern-1ex /}
51: \newcommand{\qslash}{q\kern-1ex /}
52: \newcommand{\lslash}{l\kern-1ex /}
53: \newcommand{\sslash}{s\kern-1ex /}
54: %\newcommand{\Dslash}{{\cal D}\kern-1.2ex /}
55: \newcommand{\Dslash}{D\kern-1.2ex /}
56: \newcommand{\bpsi}{\overline{\psi}}
57: \newcommand{\bc}{\overline{c}}
58: \newcommand{\tr}{{\rm tr}}
59: %\newcommand{\vev}[1]{\langle #1 \rangle}
60: \newcommand{\VEV}[1]{\left\langle{\rm T} #1\right\rangle}
61: \newcommand{\beqa}{\begin{eqnarray}}
62: \newcommand{\eeqa}{\end{eqnarray}}
63: %\newcommand{\eqn}[1]{(\ref{#1})}
64: \newcommand{\Tr}{{\rm Tr}}
65: %\newcommand{\vev}[1]{\langle #1 \rangle}
66: \newcommand{\vev}[1]{\left\langle #1 \right\rangle}
67: %\newcommand{\be}{\begin{equation}}
68: %\newcommand{\ee}{\end{equation}}
69: \newcommand{\be}{\[}
70: \newcommand{\ee}{\]}
71: \newcommand{\bd}{\begin{description}}
72: \newcommand{\ed}{\end{description}}
73: \newcommand{\la}{\langle}
74: \newcommand{\ra}{\rangle}
75: \newcommand{\ben}{\begin{eqnarray}}
76: \newcommand{\een}{\end{eqnarray}}
77: \newcommand{\nn}{\nonumber}
78: \def\lsim{\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
79: \def\gsim{\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
80: \def\simgt{\rlap{\lower 6.0 pt\hbox{$\mathchar \sim$}}\raise 2.5pt \hbox {$>$}}
81: \def\simlt{\rlap{\lower 6.0 pt\hbox{$\mathchar \sim$}}\raise 2.5pt \hbox {$<$}}
82: \newcommand{\cont}{{\rm cont}}
83: \newcommand{\latt}{{\rm latt}}
84: \newcommand{\tad}{{\rm tad}}
85: \newcommand{\wi}{{\rm WI}}
86: \newcommand{\mf}{{\rm MF}}
87: \newcommand{\msbar}{{\overline {\rm MS}}}
88: \newcommand{\spr}{{s^\prime}}
89: \newcommand{\csw}{{c_{\rm SW}}}
90: \newcommand{\hk}{{\hat k}}
91: \newcommand{\mm}{{\cal M}}
92: \newcommand{\ce}{{\it c}_{\it E}}
93: \newcommand{\cb}{{\it c}_{\it B}}
94: \newcommand{\lqcd}{{\Lambda}_{\rm QCD}}
95: %\newcommand{\pos}{{p^*}}
96: %\newcommand{\qos}{{q^*}}
97: %\newcommand{\qoss}{{q^*_s}}
98: %\newcommand{\qosd}{{q^*_d}}
99: \newcommand{\lo}{{(0)}}
100: \newcommand{\nlo}{{(1)}}
101: \newcommand{\mplo}{{m_p^{(0)}}}
102: \newcommand{\intlat}{{\int_{-\pi}^{\pi}\frac{d^4 k}{(2\pi)^4}}}
103: \newcommand{\mplomf}{{{\tilde m_p}^{(0)}}}
104: \newcommand{\dmplomf}{{{\tilde m_p}^{\tad}}}
105: \newcommand{\mpl}{{m_{p2}}}
106: \newcommand{\mph}{{m_{p1}}}
107: \newcommand{\mpllo}{{m_{p2}^{(0)}}}
108: \newcommand{\mphlo}{{m_{p1}^{(0)}}}
109: \newcommand{\mpllomf}{{{\tilde m_{p2}}^{(0)}}}
110: \newcommand{\mphlomf}{{{\tilde m_{p1}}^{(0)}}}
111: \def\ovec{\partial_\mu\hspace{-0.4cm}\raisebox{1.8ex}{$\rightarrow$}}
112: \def\antivec{\partial_\mu\hspace{-0.4cm}\raisebox{1.8ex}{$\leftarrow$}}
113: \newcommand{\crd}{\color{black}}
114: \newcommand{\crw}{\color{white}}
115: \newcommand{\crb}{\color{blue}}
116: \newcommand{\crr}{\color{red}}
117: \newcommand{\crm}{\color{magenta}}
118: \newcommand{\crg}{\color{green}}
119: \newcommand{\op}{{\cal O}}
120: \newcommand{\ceff}{{c_{\rm eff}}}
121: %%%%%%%%%%%%%% END OF MACROS %%%%%%%%%%%%%%%%%%
122: 
123: 
124: \voffset = -.2 in
125: \begin{document}
126: 
127: \preprint{UTCCS-P-44, UTHEP-567, HUPD-0801, KANAZAWA-08-06}
128: 
129: \title{2+1 Flavor Lattice QCD toward the Physical Point}% Force line breaks with \\
130: 
131: \author{
132:  S.~Aoki${}^{a,b}$,
133: % N.~Ishii${}^{a}$,
134:  K.-I.~Ishikawa${}^{d}$,
135:  N.~Ishizuka${}^{a,c}$,
136:  T.~Izubuchi${}^{b,e}$,
137:  D.~Kadoh${}^{c}$,
138:  K.~Kanaya${}^{a}$,
139:  Y.~Kuramashi${}^{a,c}$,
140:  Y.~Namekawa${}^{c}$,
141:  M.~Okawa${}^{d}$,
142: % K.~Sasaki${}^{a}$,
143:  Y.~Taniguchi${}^{a,c}$,
144:  A.~Ukawa${}^{a,c}$,
145:  N.~Ukita${}^{c}$,
146:  T.~Yoshi\'e${}^{a,c}$\\
147: (PACS-CS Collaboration)
148: }
149: \affiliation{
150:  ${}^a$Graduate School of Pure and Applied Sciences, University of Tsukuba, Tsukuba, Ibaraki 305-8571, Japan\\
151:  ${}^b$Riken BNL Research Center, Brookhaven National Laboratory, Upton,
152: New York 11973, USA\\
153:  ${}^c$Center for Computational Sciences, University of Tsukuba, Tsukuba, Ibaraki 305-8577, Japan\\
154:  ${}^d$Graduate School of Science, Hiroshima University, Higashi-Hiroshima, Hiroshima 739-8526, Japan\\
155:  ${}^e$Institute for Theoretical Physics, Kanazawa University, Kanazawa,
156:        Ishikawa 920-1192, Japan
157: }
158: 
159: %\author{Ann  Author}
160: % \altaffiliation[Also at ]{Physics Department, XYZ University.}%Lines break automatically or can be forced with \\
161: %\author{Second Author}%
162: % \email{Second.Author@institution.edu}
163: %\affiliation{%
164: %Authors' institution and/or address\\
165: %This line break forced with \textbackslash\textbackslash
166: %}%
167: 
168: %\author{Charlie Author}
169: % \homepage{http://www.Second.institution.edu/~Charlie.Author}
170: %\affiliation{
171: %Second institution and/or address\\
172: %This line break forced% with \\
173: %}%
174: 
175: 
176: \date{\today}% It is always \today, today,
177:              %  but any date may be explicitly specified
178: 
179: \begin{abstract}
180: We present the first results of the
181: PACS-CS project which aims to simulate 2+1
182: flavor lattice QCD on the physical point 
183: with the nonperturbatively $O(a)$-improved
184: Wilson quark action and the Iwasaki gauge action.
185: Numerical simulations are carried out at $\beta=1.9$, corresponding to
186: the lattice spacing of $a=0.0907(13)$~fm, 
187: on a $32^3\times 64$ lattice with the use of the domain-decomposed 
188: HMC algorithm to reduce the up-down quark mass. Further 
189: algorithmic improvements make possible the simulation whose up-down
190: quark mass is as light as the physical value. 
191: The resulting pseudoscalar meson masses range from 702~MeV down to
192:  156~MeV, which clearly exhibit the presence of chiral logarithms.
193: An analysis of the pseudoscalar meson sector 
194: with SU(3) chiral perturbation theory reveals that
195: the next-to-leading order corrections are large at the physical strange quark mass.
196: In order to estimate the physical up-down quark mass, we employ the
197: SU(2) chiral analysis expanding the strange quark contributions 
198: analytically around the physical strange quark mass. 
199: The SU(2) low energy constants ${\bar l}_3$ and ${\bar l}_4$ are comparable with
200: the recent estimates by other lattice QCD calculations. 
201: We determine the physical point together with the lattice spacing   
202: employing $m_\pi$, $m_K$ and $m_\Omega$ as input.
203: The hadron spectrum extrapolated to the physical point shows
204: an agreement with the experimental values 
205: at a few \% level of statistical errors,
206: albeit there remain possible cutoff effects.
207: We also find that our results of 
208: $f_\pi=134.0(4.2)$~MeV, $f_K=159.4(3.1)$~MeV and $f_K/f_\pi=1.189(20)$
209: where renormalization is carries out perturbatively at one loop and 
210: the errors are statistical only, are compatible with the experimental values.
211: For the physical quark masses we obtain $m_{\rm ud}^\msbar=2.527(47)$~MeV and
212: $m_{\rm s}^\msbar=72.72(78)$~MeV extracted from the axial-vector
213: Ward-Takahashi identity with the perturbative renormalization factors.
214: We also briefly discuss the results for the static quark potential.
215: \end{abstract}
216: 
217: \pacs{11.15.Ha, 12.38.-t, 12.38.Gc}% PACS, the Physics and Astronomy
218:                              % Classification Scheme.
219: %\keywords{Suggested keywords}%Use showkeys class option if keyword
220:                               %display desired
221: \maketitle
222: 
223: 
224: \section{Introduction}
225: \label{sec:intro}
226: 
227: Lattice QCD is expected to be an ideal tool to understand 
228: the nonperturbative dynamics of strong interactions from first principles.
229: In order to fulfill this promise, the first
230: step should be to establish QCD as the fundamental theory of the
231: strong interaction by reproducing basic physical quantities, {\it e.g.,} the
232: hadron spectrum, with the systematic errors under control.
233: This is about to be attained thanks to
234: the recent progress of simulation algorithms and the 
235: availability of increasingly more powerful computational resources.
236:  
237: Among various systematic errors, the two most troublesome 
238: are quenching effects and chiral extrapolation uncertainties.
239: After the systematic studies on the hadron spectrum in
240: quenched and two-flavor QCD\cite{cppacs_nf0,cppacs_nf2,jlqcd_nf2}, 
241: the CP-PACS and JLQCD collaborations performed 
242: a 2+1 flavor full QCD simulation employing the nonperturbatively 
243: $O(a)$-improved Wilson quark action\cite{csw} and the Iwasaki gauge 
244: action\cite{iwasaki} on a (2 fm$)^3$ lattice 
245: at three lattice spacings\cite{cppacs/jlqcd1,cppacs/jlqcd2}.
246: While the quenching effects were successfully removed, 
247: we were left with a long chiral extrapolation:
248: the lightest up-down quark mass reached with the plain HMC algorithm
249: was about 67 MeV corresponding to $m_\pi/m_\rho\approx 0.6$.
250: 
251: The PACS-CS project, which is based on the PACS-CS (Parallel Array
252: Computer System for Computational Sciences) computer with a
253: peak speed of 14.3 Tflops developed at University of 
254: Tsukuba\cite{ukawa1,ukawa2,boku},
255: aims at calculations on the physical point to remove the ambiguity of 
256: chiral extrapolations. It employs the same quark and gauge actions as the
257: previous CP-PACS/JLQCD work, but uses a different simulation
258: algorithm: the up-down quark mass is reduced by using the
259: domain-decomposed HMC (DDHMC) algorithm 
260: with the replay trick\cite{luscher,kennedy}. At the
261: lightest up-down quark mass, which is about 3 MeV, several algorithmic
262: improvements are incorporated, including 
263: the mass-preconditioning\cite{massprec1,massprec2}, 
264: the chronological inverter\cite{chronological}, 
265: and the deflation technique\cite{deflation}.
266: For the strange quark part we improve the PHMC 
267: algorithm\cite{fastMC,Frezzotti:1997ym,phmc}
268: with the UV-filtering procedure\cite{Alexandrou:1999ii,ishikawa_lat06}.
269: 
270: So far our simulation points cover from 702 MeV to 156 MeV
271: for the pion mass.
272: While we still have to reduce the pion mass by 21 MeV to reach the 
273: real physical point, we consider that 
274: the findings so far already merits a detailed report.   
275: In this paper we focus on the following points: (i) several algorithmic 
276: improvements make possible a simulation with the up-down quark mass
277: as light as the physical value. (ii) The range of pion mass we have 
278: simulated is sufficiently light to deserve chiral
279: analyses with the chiral perturbation theory (ChPT), which
280: reveals that the strange quark mass is not small enough to be treated
281: by the SU(3) ChPT up to the next-to-leading order (NLO).
282: (iii) The SU(2) chiral analysis on the pion sector and the linear chiral
283: extrapolation for other hadron masses yield the hadron spectrum 
284: at the physical point which is compatible with the experimental values
285: at a few \% level of statistical errors.
286: 
287: This paper is organized as follows. In Sec.~\ref{sec:detail} 
288: we present the simulation details.  
289: Measurements of hadron masses, pseudoscalar meson decay constants 
290: and quark masses are described in Sec.~\ref{sec:measurement}.
291: In Sec.~\ref{sec:chiral} we make chiral analyses on the pseudoscalar 
292: meson sector using the SU(3) and SU(2) ChPTs. We present 
293: the values of low energy constants and discuss convergences of 
294: the SU(3) and SU(2) chiral expansions.
295: %We also present the values of low energy constants.
296: The results of hadron spectrum at the physical point are given
297: in Sec.~\ref{sec:physicalpt} together with 
298: the pseudoscalar meson decay constants and the quark masses.
299: %We also invwstigate the $\rho$-$\pi\pi$ mixing effects
300: %on our lattice in Sec.~\ref{sec:rhopipi}.
301: In Sec.~\ref{sec:potential} we show
302: the results for the static quark potential. 
303: Our conclusions are summarized in Sec.~\ref{sec:conclusion}.
304: Appendices are devoted to describe the algorithmic details.
305: Preliminary results have been reported 
306: in Refs.~\cite{kura_lat07,ukita_lat07,kadoh_lat07}.
307: 
308: \section{Simulation details}
309: \label{sec:detail}
310: \subsection{Actions}
311: \label{subsec:action}
312: 
313: We employ the Iwasaki gauge action\cite{iwasaki} and 
314: the nonperturbatively $O(a)$-improved Wilson quark action
315: as in the previous CP-PACS/JLQCD work.
316: The former is composed of a plaquette and a $1\times 2$ rectangle loop:
317: \ben
318: S_{\rm g}=\frac{1}{g^2}\left\{ c_0\sum_{\rm plaquette}\tr U_{pl}
319: +c_1\sum_{\rm rectangle}\tr U_{rtg} \right\}
320: \label{eq:action_g}
321: \een
322: with $c_1=-0.331$ and $c_0=1-8c_1=3.648$.
323: The latter is expressed as
324: \begin{widetext}
325: \ben
326: S_{\rm quark}&=&\sum_{q={\rm  u,d,s}}\left[
327: \sum_n {\bar q}_n q_n 
328:     -\kappa_q \csw \sum_n \sum_{\mu,\nu}\frac{i}{2}
329:               {\bar q}_n\sigma_{\mu\nu}F_{\mu\nu}(n)q_n
330: \right.\nn\\
331: &&\left.
332: -\kappa_q\sum_n
333: \sum_\mu \left\{{\bar q}_n(1-\gamma_\mu)U_{n,\mu} q_{n+{\hat \mu}}
334:                +{\bar q}_n(1+\gamma_\mu)U^{\dag}_{n-{\hat \mu},\mu} q_{n-{\hat \mu}}\right\}
335:  \right],
336: \label{eq:action_q}
337: \een
338: \end{widetext}
339: where we consider the case of a degenerate up and down quark mass
340: $\kappa_{\rm u}=\kappa_{\rm d}$.
341: The Euclidean gamma matrices are defined in terms of
342: the Minkowski matrices in the Bjorken-Drell convention:
343: $\gamma_j=-i\gamma_{BD}^j$ $(j=1,2,3)$, 
344: $\gamma_4=\gamma_{BD}^0$,
345: $\gamma_5=\gamma_{BD}^5$ and 
346: $\sigma_{\mu\nu}=\frac{1}{2}[\gamma_\mu,\gamma_\nu]$.
347: The field strength $F_{\mu\nu}$ in the clover term
348: is given by
349: \ben 
350: F_{\mu\nu}(n)&=&\frac{1}{4}\sum_{i=1}^{4}\frac{1}{2i}
351: \left(U_i(n)-U_i^\dagger(n)\right), \\
352: U_1(n)&=&U_{n,\mu}U_{n+{\hat \mu},\nu}
353:          U^\dagger_{n+{\hat \nu},\mu}U^\dagger_{n,\nu}, \\
354: U_2(n)&=&U_{n,\nu}U^\dagger_{n-{\hat \mu}+{\hat \nu},\mu}
355:          U^\dagger_{n-{\hat \mu},\nu}U_{n-{\hat \mu},\mu}, \\
356: U_3(n)&=&U^\dagger_{n-{\hat \mu},\mu}U^\dagger_{n-{\hat \mu}-{\hat \nu},\nu}
357:          U_{n-{\hat \mu}-{\hat \nu},\mu}U_{n-{\hat \nu},\nu}, \\
358: U_4(n)&=&U^\dagger_{n-{\hat \nu},\nu}U_{n-{\hat \nu},\mu}
359:          U_{n+{\hat \mu}-{\hat \nu},\nu}U^\dagger_{n,\mu}.
360: \een
361: The improvement coefficient $\csw$ for $O(a)$ improvement was 
362: determined nonperturbatively in Ref.~\cite{csw}.
363: 
364: \subsection{Simulation parameters}
365: \label{subsec:params}
366: 
367: 
368: Our simulations are carried out at $\beta=1.90$ on a $32^3\times 64$
369: lattice for which we use $\csw=1.715$~\cite{csw}. This $\beta$ value is one of the three in the previous
370: CP-PACS/JLQCD work, whereas the lattice size 
371: is enlarged from $20^3\times 40$  to investigate the baryon masses.
372: The lattice spacing is found to be 0.0907(14) fm 
373: whose determination is explained later.
374: Table~\ref{tab:param} lists the run parameters of our simulations. 
375: The six combinations of the hopping parameters 
376: $(\kappa_{\rm ud},\kappa_{\rm s})$ are chosen 
377: based on the previous CP-PACS/JLQCD results. 
378: The heaviest combination 
379: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13700,0.13640)$ in this work 
380: corresponds to the lightest one in the previous CP-PACS/JLQCD simulations, 
381: which enable us to make a direct comparison of the two results with
382: different lattice sizes.
383: The physical point of the strange quark at $\beta=1.90$ was
384: estimated as $\kappa_{\rm s}=0.136412(50)$ 
385: in the CP-PACS/JLQCD work\cite{cppacs/jlqcd1, cppacs/jlqcd2}.
386: This is the reason why all our simulations 
387: are carried out with $\kappa_{\rm s}=0.13640$, 
388: the one exception being the run at 
389: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13754,0.13660)$ 
390: to investigate the strange quark mass dependence. 
391: After more than 1000 MD time for thermalization
392: we calculate hadronic observables solving quark propagators at every
393: 10 trajectories for $\kappa_{\rm ud}\ge 0.13770$
394: and 20 trajectories for $\kappa_{\rm ud}=0.13781$,
395: while we measure the plaquette expectation value at every trajectory.
396: 
397: \subsection{Algorithm}
398: \label{subsec:algorithm}
399: 
400: Our base algorithm for penetrating into the small mass region for a 
401: degenerate pair of up and down quarks is the DDHMC algorithm\cite{luscher}.
402: The effectiveness of this algorithm for reducing the quark mass 
403: was already shown in the $N_f=2$ case\cite{luscher,del06,del07}. 
404: We found that it works down to $\kappa_{\rm ud} = 0.13770$ 
405: (or $m_\pi\approx 300$~MeV) on our $32^3\times 64$ lattice.  
406: Moving closer to the physical point, however, we found it necessary to 
407: add further enhancements including mass preconditioning, which we call 
408: mass-preconditioned DDHMC (MPDDHMC).  This is the algorithm we applied 
409: at our lightest point at $\kappa_{\rm ud}=0.13781$. 
410: 
411: The characteristic feature of the DDHMC algorithm is a geometric separation 
412: of the up-down quark determinant into the UV and the IR parts, which 
413: is implemented by domain-decomposing the full lattice into small blocks.
414: We choose $8^4$ for the block size, being less than (1~fm)$^4$ in
415: physical units and small enough to reside 
416: within a computing node of the PACS-CS computer.
417: The latter feature is computationally advantageous since the calculation 
418: of the UV part requires no communication between blocks so that 
419: the inter-node communications are sizably reduced.   
420: 
421: The UV/IR separation enables the application of multiple time scale 
422: integration schemes\cite{sexton}, which reduces the simulation cost 
423: substantially. In our simulation points we find that 
424: the relative magnitudes of the force terms are 
425: \ben   
426: ||F_{\rm g}||:||F_{\rm UV}||:||F_{\rm IR}|| \approx 16:4:1,
427: \label{eq:force_ddhmc}
428: \een
429: where we adopt the convention $||M||^2=-2{\rm tr}(M^2)$
430: for the norm of an element $M$ of the SU(3) Lie algebra, and 
431: $F_{\rm g}$ denotes the gauge part and $F_{\rm UV, IR}$ are for the UV
432: and the IR parts of the up-down quarks.
433: The associated step sizes for the forces are controlled by
434: three integers $N_{0,1,2}$ introduced by  
435: $\delta\tau_{\rm g}=\tau/N_0 N_1 N_2,\ \  \delta\tau_{\rm UV}=\tau/N_1
436: N_2,\ \  \delta\tau_{\rm IR}=\tau/N_2$ with $\tau$ the trajectory
437: length.
438: The integers $N_{0,1,2}$ should be chosen such that 
439: \begin{eqnarray}
440:  \delta\tau_{\rm g} ||F_{\rm g}|| \approx \delta\tau_{\rm UV} ||F_{\rm UV}|| \approx \delta\tau_{\rm IR} ||F_{\rm IR}||.
441: \end{eqnarray} 
442: The relative magnitudes between the forces in Eq.~(\ref{eq:force_ddhmc})
443: tell us that $\delta\tau_{\rm IR}$ may be chosen roughly 16 times as 
444: large as $\delta\tau_{\rm g}$ and 4 times that of $\delta\tau_{\rm UV}$,
445: which means that we need to calculate $F_{\rm IR}$ an order of magnitude 
446: less frequently in the molecular dynamics trajectories.
447: Since the calculation of  $F_{\rm IR}$ contains the quark matrix
448: inversion on the full lattice, which is the most computer 
449: time consuming part, 
450: this integration scheme saves the simulation cost remarkably.
451: 
452: The values for $N_{0,1,2}$ are listed in Table~\ref{tab:param}, where
453: $N_0$ and $N_1$ are fixed at 4 for all the hopping parameters, 
454: while the value of $N_2$ is adjusted taking account of acceptance rate 
455: and simulation stability. The threshold for the replay 
456: trick\cite{luscher,kennedy} 
457: for dealing with instabilities of molecular dynamics 
458: trajectories leading to large values of $dH$ is set to be $\Delta  H>2$.
459: 
460: For the strange quark, we employ the UV-filtered PHMC (UVPHMC) 
461: algorithm\cite{ishikawa_lat06}.
462: The UVPHMC action for the strange quark is obtained through
463: the UV-filtering\cite{Alexandrou:1999ii} applied after 
464: the even-odd site preconditioning for the quark matrix.
465: The domain-decomposition is not used.
466: %The action is properly combined to the DDHMC action and 
467: The polynomial approximation is corrected by the global Metropolis test
468: \cite{NoisyMetropolisMB}.
469: Since we find  $||F_{\rm s}||\approx ||F_{\rm IR}||$, 
470: the step size is chosen as $\delta\tau_{\rm s}=\delta\tau_{\rm IR}$.
471: The polynomial order for UVPHMC, which is denoted 
472: by $N_{\rm poly}$ in Table~\ref{tab:param},
473: is adjusted to yield high acceptance rate for the global Metropolis
474: test at the end of each trajectory.
475: 
476: The inversion of the Wilson-Dirac operator $D$ on the full lattice 
477: is carried out by the SAP (Schwarz alternating procedure) 
478: preconditioned GCR solver.  
479: The preconditioning is accelerated with the single-precision
480: arithmetic\cite{sap+gcr}. 
481: We employ the stopping condition $|Dx-b|/|b|<10^{-9}$ for the force
482: calculation and $10^{-14}$ for the Hamiltonian, which guarantees 
483: the reversibility of the molecular dynamics trajectories to a high
484: precision: $|\Delta U|<10^{-12}$ for the link variables and 
485: $|\Delta H|<10^{-8}$ for the Hamiltonian at 
486: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13781,0.13640)$.
487: We describe the details of the DDHMC algorithm and the solver 
488: implementation used for $\kappa_{\mathrm{ud}}\leq 0.13770$ 
489: in Appendix~\ref{app:DDHMC}.
490: 
491: As we reduce the up-down quark mass, we observe a tendency that
492: the fluctuation of $||F_{\rm IR}||$ during the molecular dynamics 
493: trajectory increases, which results in 
494: a higher replay rate due to the appearance of trajectories with 
495: large $\Delta H$.
496: Since $\Delta  H$ is controlled by the product of
497: $\delta\tau_{\rm IR}$ and $||F_{\rm IR}||$, 
498: a possible solution to suppress the replay rate is to reduce 
499: $\delta\tau_{\rm IR}$. In this case, however, 
500: we find the acceptance becoming unnecessarily close to unity.
501: Another solution would be to tame the fluctuation of $||F_{\rm IR}||$, 
502: and we employ for this purpose the 
503: mass-preconditioner\cite{massprec1,massprec2} to the IR part 
504: of the pseudofermion action.  
505: The quark mass in the preconditioner is controlled by an additional
506: hopping parameter $\kappa_{\rm ud}^\prime=\rho\kappa_{\rm ud}$,
507: where $\rho$ should be less than unity so that calculating with the 
508: preconditioner is less costly than with the original IR part. 
509: The IR force $F_{\rm IR}$ is split into $F_{\rm IR}^\prime$ 
510: and ${\tilde F}_{\rm IR}$. The former is derived from
511: the preconditioner and the latter from the preconditioned action. 
512: 
513: 
514: We employ the mass-preconditioned DDHMC (MPDDHMC) algorithm
515: for the run at the lightest up-down quark mass of
516: $\kappa_{\rm ud}=0.13781$.
517: With our choice of $\rho=0.9995$ 
518: the relative magnitudes of the force terms become
519: \ben   
520: ||F_{\rm g}||:||F_{\rm UV}||:||F_{\rm IR}^\prime||:||{\tilde F}_{\rm IR}|| \approx 16:4:1:1/7.
521: \label{eq:force_mpddhmc}
522: \een
523: According to this result
524: we choose $(N_0,N_1,N_2,N_3)=(4,4,4,6)$ for the associated step sizes.
525: Here the choice of $N_2=4$ does not follow the criterion 
526: $\delta\tau_{\rm IR}^\prime ||F_{\rm IR}^\prime|| \approx 
527: \delta{\tilde \tau}_{\rm IR} ||{\tilde F}_{\rm IR}||$.
528: This is because we take account of the fluctuations of 
529: $||{\tilde F}_{\rm IR}||$.
530: The replay trick is not implemented in the runs at $\kappa_{\rm ud}=0.13781$.
531: For the step size for the strange quark in the UVPHMC algorithm 
532: we choose $\delta\tau_{\rm s}=\delta\tau_{\rm IR}^\prime$
533: as we observe $||F_{\rm s}||\approx ||F_{\rm IR}^\prime||$.
534: 
535: The inversion of $D$ during the molecular dynamics steps 
536: is also improved at $\kappa_{\rm ud}=0.13781$ in three ways.  
537: (i) We employ the chronological guess using the last 16 solutions 
538: to construct the initial solution vector of $D^{-1}$ 
539: on the full lattice\cite{chronological}.
540: In order to assure the reversibility we apply a stringent 
541: stopping condition $|Dx-b|/|b|<10^{-14}$ to the 
542: force calculation.
543: (ii) The inversion algorithm is replaced by a nested BiCGStab solver,
544: which consists of an inner solver
545: accelerated with single precision arithmetic and
546: with an automatic tolerance control ranging from $10^{-3}$ to $10^{-6}$,
547: and an outer solver with a stringent tolerance of $10^{-14}$ operated
548: with the double precision. The approximate solution obtained by the
549: inner solver works as a preconditioner for the outer solver.
550: (iii) We implement the deflation technique 
551: to make the solver robust against possible small eigenvalues 
552: allowed in the Wilson-type quark action.
553: Once the inner BiCGStab solver becomes stagnant during the inversion of $D$, 
554: it is automatically replaced by the GCRO-DR (Generalized Conjugate
555: Residual with implicit inner Orthogonalization and Deflated Restarting)
556: algorithm\cite{deflation}.
557: In our experience the GCRO-DR algorithm is important for calculating 
558: $D^{-1}$ but does not save the simulation time at $\kappa_{\rm ud}=0.13781$.
559: More details of the MPDDHMC algorithm and the improvements are given 
560: in Appendix~\ref{app:MPDDHMC}.
561: 
562: \subsection{Implementation on the PACS-CS computer}
563: 
564: All of the simulations reported in this article have been carried out 
565: on the PACS-CS parallel computer\cite{boku}.  
566: PACS-CS consists of 2560 nodes, each node equipped with a 
567: 2.8GHz Intel Xeon single-core processor ({\it i.e.,} 5.6Gflops of peak 
568: speed) with 2 GBytes of main memory.  The nodes are arranged into a 
569: $16\times 16\times 10$ array and connected by a 3-dimensional 
570: hypercrossbar network made of a dual Gigabit Ethernet in each direction.  
571: The network bandwidth is 750 MBytes/sec for each node. 
572: 
573: The programming language is mainly Fortran 90 with Intel Fortran compiler. 
574: To further enhance the performance we used Intel C++ compiler
575: for the single precision hopping matrix multiplication routines which are
576: the most time consuming parts.
577: The Intel compiler enables us to use the Intel Streaming SIMD 
578: extensions 2 and 3 intrinsics directly without writing assembler language.
579: 
580: We employ a 256 node partition of PACS-CS to execute our $32^3\times 64$ runs. 
581: The sustained performance including communication overhead with our 
582: DDHMC code turns out to be 18\%.  
583: The computer time needed for one MD unit is listed in Table~\ref{tab:param}.  
584:   
585: \subsection{Efficiency of DDHMC algorithms}
586: \label{subsec:efficiency}
587: 
588: The efficiency of the DDHMC algorithm may be clarified in comparison 
589: with that of the HMC algorithm.
590: For $N_f=2$ QCD simulations with the Wilson-clover quark action,  
591: an empirical cost formula suggested for the HMC algorithm
592: based on the CP-PACS and JLQCD $N_f=2$ runs  was as follows~\cite{berlinwall}:
593: \begin{widetext}
594: \ben
595: {\rm cost[Tflops\cdot years]}&=&C\left[\frac{\#{\rm conf}}{1000}\right]\cdot
596: \left[\frac{0.6}{m_\pi/m_\rho}\right]^6\cdot
597: \left[\frac{L}{3{\rm ~fm}}\right]^5\cdot
598: \left[\frac{0.1{\rm ~fm}}{a}\right]^7
599: \label{eq:cost_hmc}
600: \een
601: \end{widetext}
602: with $C\approx 2.8$.
603: A strong quark mass dependence in the above formula 
604: $1/(m_\pi/m_\rho)^6\sim 1/m_{\rm ud}^3$ stems from three factors:
605: (i) the number of iterations for the quark matrix inversion increases 
606: as the condition number which is proportional to $1/m_{\rm ud}$, 
607: (ii)to keep the acceptance rate constant we should take
608: $\delta\tau\propto m_{\rm ud}$ for the step size in
609: the molecular dynamics trajectories, and (iii)
610: the autocorrelation time of the HMC evolution was consistent with 
611: an $1/m_{\rm ud}$ dependence in the CP-PACS runs\cite{cppacs_nf2}.
612: 
613: To estimate the computational cost for
614: $N_f=2+1$ QCD simulations with the HMC algorithm,
615: we assume that the strange quark contribution is given by half of
616: Eq.~(\ref{eq:cost_hmc}) at $m_\pi/m_\rho=0.67$ which
617: is a phenomenologically estimated ratio of the
618: strange pseudoscalar meson ``$m_{\eta_{\rm ss}}$'' and $m_\phi$:
619: \ben
620: \frac{m_{\eta_{\rm
621: ss}}}{m_\phi}=\frac{\sqrt{2m_K^2-m_\pi^2}}{m_\phi}\approx 0.67.\nn
622: \een
623: Since the strange quark is relatively heavy, its computational
624: cost occupies only a small fraction as the up-down quark mass decreases.
625: In Fig.~\ref{fig:berlinwall} we draw the cost formula
626: for the $N_f=2+1$ case as a function of $m_\pi/m_\rho$,
627: where we take \#conf=100, $a$=0.1~fm and $L=3$~fm in Eq.~(\ref{eq:cost_hmc})
628: as a representative case.
629: We observe a steep increase of the computational cost below
630: $m_\pi/m_\rho\simeq 0.5$.
631: At the physical point the cost expected from Eq.~(\ref{eq:cost_hmc}) 
632: would be $O(100)$ Tflops$\cdot$years.
633: 
634: 
635: Let us now see the situation with the DDHMC algorithm.
636: The blue open symbol in Fig.~\ref{fig:berlinwall} denotes the measured cost at
637: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13770,0.13640)$, 
638: which is the lightest point implemented with the DDHMC algorithm.
639: Here we assume that we need 100 MD time separation between 
640: independent configurations.
641: We observe a remarkable reduction in the cost by a factor $20-30$ 
642: in magnitude.
643: The majority of this reduction arises from the multiple time scale
644: integration scheme and the GCR solver
645: accelerated by the SAP preconditioning
646: with the single-precision arithmetic.
647: Roughly speaking, the improvement factor is $O(10)$ for the former
648: and $3-4$ for the latter.
649: The cost of the MPDDHMC algorithm at 
650: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13781,0.13640)$
651: is plotted by the blue closed symbol in Fig.~\ref{fig:berlinwall}. 
652: In this case, the reduction is mainly owing to the multiple time scale
653: integration scheme armored with the mass-preconditioning and 
654: the chronological inverter for $F_{\rm IR}^\prime$ and ${\tilde F}_{\rm IR}$.
655: As we already noted, the GCRO-DR solver does not accelerate the inversion
656: albeit it renders the solver robust against the small eigenvalues
657: of the Wilson-Dirac operator.
658: 
659: %Note that the quark mass dependence is also tamed:
660: Since we find in Table~\ref{tab:param} 
661: that $\tau_{\rm int}[P]$ is roughly independent
662: of the up-down quark mass employed in the DDHMC algorithm,
663: the cost is expected to be proportional to $1/m_{\rm ud}^2$.
664: Assuming this quark mass dependence for the MPDDHMC algorithm,
665: we find that simulations at the physical point is feasible, 
666: at least for $L\approx 3$~fm lattices, with 
667: $O(10)$ Tflops computers, which are already available at present.
668: 
669: 
670: \subsection{Autocorrelations and statistical error analysis}
671: \label{subsec:autoc}
672: 
673: The autocorrelation function $\Gamma(\tau)$ of a time series of 
674: an observable ${\cal O}$ in the course of a numerical simulation is
675: given by
676: \ben
677: \Gamma(\tau)=\la {\cal O}(\tau_0){\cal O}(\tau_0+\tau)\ra-\la{\cal O}(\tau_0)\ra^2
678: \een
679: In Fig.~\ref{fig:PLQ} we show the plaquette history and the normalized
680: autocorrelation function $\rho(\tau)=\Gamma(\tau)/\Gamma(0)$ at
681: $\kappa_{\rm ud}=0.13727$ as an example. 
682: %Note that they are obtained 
683: %in a single run as well as the case of $\kappa_{\rm ud}$=0.13700.
684: The integrated autocorrelation time
685: is estimated as $\tau_{\rm int}[P]=20.9(10.2)$
686: following the definition in Ref.~\cite{luscher}
687: \ben
688: \tau_{\rm int}(\tau)=\frac{1}{2}
689: +\sum_{0< \tau \le W} \rho(\tau),
690: \een
691: where the summation window $W$ is set to the first time lag $\tau$ such that
692: $\rho(\tau)$ becomes consistent with zero within the error bar.
693: In this case we find $W=119.5$.
694: The choice of $W$ is not critical for estimate of
695: $\tau_{\rm int}$ in spite of the long tail observed in Fig.~\ref{fig:PLQ}.
696: Extending the summation window, we find
697: that $\tau_{\rm int}[P]$ saturates at $\tau_{\rm int}[P]\approx 25$
698: beyond $W=200$, which is
699: within the error bar of the original estimate.
700: 
701: Our simulations at $\kappa_{\rm ud}=0.13700$ and $0.13727$ are fast enough 
702: to be executed by a single long run of 2000 MD units. 
703: The simulations at $\kappa_{\rm ud}\ge 0.13754$ become increasingly 
704: CPU time consuming so that we had to execute multiple runs in parallel .   
705: The data obtained from different runs are combined into 
706: a single extended series, for which we define the above autocorrelation
707: function $\Gamma(\tau)$ as if it were a single run.  
708: The results for $\tau_{\rm int}[P]$ are listed in 
709: Table~\ref{tab:param}.
710: Although we hardly observe any systematic quark mass dependence for
711: the integrated autocorrelation time,
712: the statistics may not be
713: sufficiently large to derive a definite conclusion.
714:   
715: In the physics analysis we estimate the statistical errors with the
716: jackknife method in order to take account of the autocorrelation.
717: For the simulations at $\kappa_{\rm ud}\ge 0.13754$ we apply the
718: jackknife analysis after combining the different runs into a single
719: series. The bin size dependence of the statistical error is investigated
720: for each physical observable. 
721: %In order to examine the possibility that 
722: %the auto-correlation estimate is artificially cut off by the length 
723: %of each run, 
724: For a cross-check
725: we also carry out the bootstrap error estimation with 1000 
726: samples.  In all cases we find the two estimates agree for the magnitude of 
727: errors within 10\%.
728: We follow the procedure given in Appendix  B of Ref.~\cite{cppacs_nf2}
729: in estimating the errors for the chiral fit parameters. 
730: 
731: 
732: \section{Measurements of hadronic observables}
733: \label{sec:measurement}
734: 
735: 
736: \subsection{Hadron masses, quark masses and decay constants}
737: \label{subsec:hmass}
738: 
739: We measure the meson and baryon correlators at the unitary
740: points where the valence quark masses are equal to the sea quark masses.
741: For the meson operators we employ
742: \ben
743: M_\Gamma^{fg}(x)={\bar q}_f(x)\Gamma q_g(x),
744: \een
745: where $f$ and $g$ denote quark flavors and $\Gamma$ are 16
746: Dirac matrices $\Gamma={\rm I}$, $\gamma_5$, $\gamma_\mu$, 
747: $i\gamma_\mu\gamma_5$ and  $i[\gamma_\mu,\gamma_\nu]/2$ 
748: $(\mu,\nu=1,2,3,4)$.
749: The octet baryon operators are given by 
750: \ben
751: {\cal O}^{fgh}_\alpha(x)=\epsilon^{abc}((q_f^a(x))^T C\gamma_5 q_g^b(x))
752: q_{h\alpha}^c(x),
753: \een 
754: where $a,b,c$ are color indices, $C=\gamma_4\gamma_2$ is the 
755: charge conjugation matrix and $\alpha=1,2$ labels the $z$-component
756: of the spin 1/2.
757: The $\Sigma$- and $\Lambda$-like octet baryons are distinguished by
758: the flavor structures:
759: \ben
760: \Sigma{\rm -like}\;\; &:& \;\; -\frac{{\cal O}^{[fh]g}+{\cal O}^{[gh]f}}{\sqrt{2}},\\
761: \Lambda{\rm -like}\;\; &:& \;\; \frac{{\cal O}^{[fh]g}-{\cal O}^{[gh]f}-2{\cal O}^{[fg]h}}
762: {\sqrt{6}},
763: \een
764: where $O^{[fg]h}={\cal O}^{fgh}-{\cal O}^{gfh}$.
765: We define the decuplet baryon operators for the four $z$-components of the
766: spin 3/2 as
767: \ben
768: D^{fgh}_{3/2}(x)&=&\epsilon^{abc}((q_f^a(x))^T C\Gamma_+
769: q_g^b(x))q_{h1}^c(x),\\
770: D^{fgh}_{1/2}(x)&=&
771: \epsilon^{abc}[((q_f^a(x))^T C\Gamma_0q_g^b(x))q_{h1}^c(x)\nn\\
772: &&-((q_f^a(x))^T C\Gamma_+q_g^b(x))q_{h2}^c(x)]/3,\\
773: D^{fgh}_{-1/2}(x)&=&
774: \epsilon^{abc}[((q_f^a(x))^T C\Gamma_0q_g^b(x))q_{h2}^c(x)\nn\\
775: &&-((q_f^a(x))^T C\Gamma_-q_g^b(x))q_{h1}^c(x)]/3,\\
776: D^{fgh}_{-3/2}(x)&=&\epsilon^{abc}((q_f^a(x))^T C\Gamma_-
777: q_g^b(x))q_{h2}^c(x),
778: \een
779: where $\Gamma_{\pm}=(\gamma_1\mp\gamma_2)/2$, $\Gamma_0=\gamma_3$ and
780: the flavor structures should be symmetrized.
781: 
782: We calculate the meson and the baryon correlators
783: with point and smeared sources and a local sink.
784: For the smeared source we employ an exponential smearing function 
785: $\Psi(|{\vec x}|)=A_q\exp(-B_q|{\vec x}|)$ $(q={\rm ud,s})$ with
786: $\Psi(0)=1$ for the ud and s quark propagators. The
787: parameters $A_q$ and $B_q$ are adjusted 
788: from a couple of configurations after the beginning of the production run
789: such that the pseudoscalar meson effective masses
790: reach a plateau as soon as possible. 
791: %when averaged over the meson particle types. 
792: Their values are given in Table~\ref{tab:param_smear}.
793: The point and smeared sources allow the hadron propagators 
794: with nonzero spatial momentum, and we calculate them for 
795: ${\vec p}=(0,0,0),(\pi/16,0,0),(0,\pi/16,0),(0,0,\pi/16)$.
796: 
797: 
798: In order to increase the statistics we calculate the hadron correlators
799: with four source points at $(x_0,y_0,z_0,t_0)$=$(17,17,17,1)$, $(1,1,1,9)$,
800: $(25,25,25,17)$, and $(9,9,9,25)$ for $\kappa_{\rm ud}\ge 0.13754$.
801: They are averaged on each configuration before the jackknife analysis. 
802: This procedure reduces the statistical errors 
803: by typically $20-40$\% for the vector meson and the baryon masses
804: and  less than 20\% for the pseudoscalar meson masses 
805: compared to a single source point.
806: For further enhancement of the signal we average zero momentum 
807: hadron propagators over possible spin states on each configuration: 
808: three polarization states for the vector meson and two (four) 
809: spin states for the octet (decuplet) baryons. 
810:  
811: 
812: We extract the meson and the baryon masses from the hadron propagators
813: with the point sink and the smeared source, where all the valence quark
814: propagators in the mesons and the baryons have the smeared sources.
815: Figures~\ref{fig:m_eff_kud54}$-$\ref{fig:m_eff_kud81} show effective
816: mass plots for the meson and the baryon propagators with the smeared
817: source for $\kappa_{\rm ud}\ge 0.13754$. We observe that the excited
818: state contributions are effectively suppressed and good plateaus start  
819: at small values of $t$. 
820: 
821: The hadron masses are extracted by uncorrelated $\chi^2$ fits 
822: to the propagators without
823: taking account of correlations between different time slices,
824: since we encounter instabilities for correlated fits using covariance matrix.
825: We assume a single hyperbolic cosine function for the mesons and a single
826: exponential form for the baryons. 
827: The lower end of the fit range $t_{\rm min}$ is determined by
828: investigating stability of the fitted mass. 
829: On the other hand, the choice of $t_{\rm max}$ 
830: gives little influence on the fit results as far as
831: the effective mass exhibits a plateau and the signal is not lost in the noise. 
832: We employ the same fit range $[t_{\rm min},t_{\rm max}]$ for the same 
833: particle type: $[13,30]$ for pseudoscalar mesons, $[10,20]$ for vector
834: mesons, $[10,20]$ for octet baryons and $[8,20]$ for decuplet baryons. 
835: These fit ranges are independent of the quark masses.
836: Resulting hadron masses are summarized in Table~\ref{tab:hmass}. 
837: 
838: Statistical errors are estimated with the jackknife
839: procedure. In Fig.~\ref{fig:binerr_pi} we show the bin size dependence of
840: the error for $m_\pi$ and $m_{\eta_{\rm ss}}$.  
841: We observe that the magnitude of error 
842: reaches a plateau after 100$-$200 MD time depending on the quark
843: mass. Since similar binsize dependences are found for 
844: other particle types,
845: we employ a binsize of 250 MD time for the jackknife analysis 
846: at $0.13770\ge \kappa_{\rm ud}\ge 0.13754$.  
847: At our lightest point $\kappa_{\rm ud}=0.13781$ with the statistics of 
848: 990 MD units, we had to reduce the bin size to 110 MD units.
849: 
850: We define the bare quark mass based on the axial vector
851: Ward-Takahashi identity (AWI) by the ratio of matrix elements
852: of the pseudoscalar density $P$ and the fourth component of the
853: axial vector current $A_4$: 
854: \ben
855: {\bar m}^{\rm AWI}_f+ {\bar m}^{\rm AWI}_g=\frac{\langle 0 |\nabla_4
856:   A_4^{\rm imp} |{\rm PS} \rangle}{\langle 0| P | PS\rangle},
857: \een
858: where $|{\rm PS}\rangle$ denotes the pseudoscalar meson state
859: at rest and $f$ and $g$ $(f,g={\rm ud},s)$ label the flavors 
860: of the valence quarks.
861: We employ the nonperturbatively $O(a)$-improved 
862: axial vector current $A_4^{\rm imp}=A_4+c_A{\bar \nabla}_4 P$ 
863: with ${\bar \nabla}_4$ the symmetric lattice derivative, and 
864: $c_A=-0.03876106$ as determined in Ref.~\cite{ca}.
865: In practice the AWI quark mass is determined by 
866: \ben
867: {\bar m}^{\rm AWI}_f+{\bar m}^{\rm AWI}_g&=&\frac{m_{\rm PS}}{2}
868: \left\vert \frac{C_A^s}{C_P^s}\right\vert,
869: \een
870: where $m_{\rm PS}$, $C_A^s$ and $C_P^s$ are obtained by
871: applying a simultaneous $\chi^2$ fit to 
872: \ben
873: \langle A_4^{\rm imp}(t) P^s(0)\rangle=2C_A^s
874: \frac{\sinh(-m_{\rm PS}(t-T/2))}{\exp(m_{\rm PS}T/2)}
875: \label{eq:a4p_s}
876: \een
877: with a smeared source and
878: \ben
879: \langle P(t)P^s(0)\rangle=2C_P^s
880: \frac{\cosh(-m_{\rm PS}(t-T/2))}{\exp(m_{\rm PS}T/2)}
881: \label{eq:pp_s}
882: \een
883: with a smeared source, where $T$ denotes the temporal extent of the lattice.
884: We employ the fit range of $[t_{\rm min},t_{\rm max}]=[13,25]$ 
885: for the former and $[13,30]$ for the latter at all the
886: hopping parameters.
887: The renormalized quark mass in the continuum ${\overline{\rm MS}}$
888: scheme is defined as 
889: \ben
890: m^{\overline{\rm MS}}_f&=&\frac{Z_A}{Z_P}m^{\rm AWI}_f,
891: \een
892: with
893: \ben
894: m^{\rm AWI}_f&=&\frac{\left(1+b_A
895:   \frac{m_f^{\rm VWI}}{u_0}\right)}
896: {\left(1+b_P \frac{m_f^{\rm VWI}}{u_0} \right)}{\bar m}^{\rm AWI}_f.
897: \een
898: The renormalization factors
899: $Z_{A,P}$ and the improvement coefficients $b_{A,P}$ are 
900: perturbatively evaluated up to one-loop level\cite{z_pt,z_imp_pt,aokietal}
901: with the tadpole improvement. The VWI quark  masses in the $ma$
902: corrections are perturbatively obtained from the AWI quark masses:
903: \ben
904: \frac{m_f^{\rm VWI}}{u_0} = \frac{Z_A}{Z_P Z_m}{\bar m}^{\rm AWI}_f.
905: \label{eq:mq_vwi}
906: \een
907: In Table~\ref{tab:qmass} we list
908: the values of $m^{\overline{\rm MS}}_{\rm ud}$ and 
909: $m^{\overline{\rm MS}}_{\rm s}$ renormalized at the scale of $1/a$, whose 
910: statistical errors are provided by the jackknife analysis with 
911: the bin size chosen as in the hadron mass measurements.
912: 
913: The bare pseudoscalar meson decay constant on the lattice 
914: is defined by
915: \ben
916: \left\vert\langle 0|A_4^{\rm imp}|{\rm PS}\rangle\right\vert 
917: =f^{\rm bare}_{\rm PS}m_{\rm PS}.
918: \een
919: with $|{\rm PS}\rangle$ the pseudoscalar meson state at rest 
920: consisting of $f$ and $g$ valence quarks.
921: We evaluate $f^{\rm bare}_{\rm PS}$ from the formula
922: \ben
923: f^{\rm bare}_{\rm PS}&=&\left\vert\frac{C_A^s}{C_P^s}\right\vert
924: \sqrt{\frac{2\left\vert C_P^l\right\vert}{m_{\rm PS}}},
925: \een
926: where we extract $m_{\rm PS}$,  $C_A^s$ , $C_{P}^s$ and $C_P^l$
927: from a simultaneous fit of Eqs.~(\ref{eq:a4p_s}), (\ref{eq:pp_s}) and
928: \ben
929: \langle P(t)P^l(0)\rangle=2C_P^l
930: \frac{\cosh(-m_{\rm PS}(t-T/2))}{\exp(m_{\rm PS}T/2)}
931: \label{eq:pp_l}
932: \een
933: with a local source.
934: The fit ranges are $[13,25]$, $[13,30]$ and $[15,25]$, respectively,
935: at all the hopping parameters.
936: The bare decay constant $f^{\rm bare}_{\rm PS}$ is renormalized 
937: perturbatively with
938: \ben
939: f_{\rm PS}&=&u_0 Z_A \left(1+b_A\frac{m^{\rm VWI}_f+m^{\rm VWI}_g}{2u_0}\right)
940: f^{\rm bare}_{\rm PS},
941: \een
942: where $m^{\rm VWI}_f$ is estimated by Eq.~(\ref{eq:mq_vwi}). 
943: Table~\ref{tab:qmass} summarizes
944: the results for $f_{\rm PS}$ with the statistical errors
945: evaluated by the jackknife analysis with the bin size chosen as in the 
946: hadron mass measurements.
947: 
948: 
949: \subsection{Comparison with the previous CP-PACS/JLQCD results}
950: \label{subsec:comparison}
951: 
952: A comparison between the present PACS-CS results 
953: and those with the previous CP-PACS/JLQCD work\cite{cppacs/jlqcd1, cppacs/jlqcd2}
954: obtained with the same gauge and quark actions
955: is possible at $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13700,0.13640)$,
956: except that the lattice sizes are different: $32^3\times 64$ for the former
957: and $20^3\times 40$ for the latter.
958: In Table~\ref{tab:comp} we list the PACS-CS and the CP-PACS/JLQCD
959: results for the hadron masses at 
960: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13700,0.13640)$.
961: While the results for $m_\pi$ are consistent within the errors,
962: we find a $1-2$\% deviation for $m_\rho$ and $m_N$. 
963: This is pictorially confirmed in Fig.~\ref{fig:cmpr_effm}
964: which shows the effective masses for the $\pi$ meson and the nucleon.
965: The pion effective masses  are almost degenerate from  $t=8$ to 17, while   
966: a slight discrepancy is observed for the nucleon results. 
967: The $\rho$ and  nucleon masses may be suffering  from 
968: finite size effects.
969: 
970: \section{Chiral analysis on pseudoscalar meson masses and decay constants}
971: \label{sec:chiral}
972: 
973: The analysis of chiral behavior of pseudoscalar meson masses and decay 
974: constants occupy an important place in lattice QCD.  
975: Theoretically the main points to examine are 
976: the presence of chiral logarithms as predicted 
977: by ChPT and the convergence of the ChPT series itself. 
978: The viability of ChPT is 
979: relevant also for studies of finite-size effects.  The low energy constants 
980: are important from phenomenological points of view.  And finally, the chiral 
981: analysis is required to pin down the physical point in the parameter space of the 
982: simulations. We begin with a discussion of a subtle point in the chiral analysis
983: when Wilson-clover quark action with implicit chiral symmetry breaking is employed. 
984: 
985: 
986: \subsection{Chiral perturbation theory for $O(a)$-improved Wilson-type 
987: quark action }
988: 
989: Our simulations are carried out with a non-perturbatively $O(a)$-improved Wilson 
990: quark action.  At present we correct $O(ma)$ terms in the AWI quark masses 
991: and decay constants by one-loop perturbation theory.  These corrections 
992: are expected to be very small in magnitude, and hence 
993: leading scaling violations in 
994: meson masses and decay constants from our simulations can be 
995: taken as $O(a^2)$.  
996: In this case, the NLO formula of Wilson chiral perturbation theory 
997: for the SU(3) flavor case\cite{wchpt},  
998: which incorporates the leading contributions of the
999: implicit chiral symmetry breaking effects of the Wilson-type quarks, 
1000: are given by
1001: \begin{widetext}
1002: \ben
1003: \frac{m_{\pi}^2}{2 m_{\rm ud}}&=& B_0 \left\{
1004: 1+\mu_\pi-\frac{1}{3}\mu_\eta 
1005: +\frac{2B_0}{f_0^2} \left(
1006: 16 m_{\rm ud}  (2L_{8}-L_5) 
1007: +16 (2 m_{\rm ud} +m_{\rm s}) (2L_{6}-L_4)  
1008: \right) -\frac{2H^{\prime\prime}}{f_0^2}\right\}, 
1009: \label{eq:wchpt_mpi}
1010: \\
1011: \frac{m_K^2}{(m_{\rm ud} +m_{\rm s})}&=&B_0 \left\{
1012: 1+\frac{2}{3}\mu_\eta
1013: +\frac{2B_0}{f_0^2}\left(
1014: 8(m_{\rm ud} + m_{\rm s}) (2L_{8}-L_5)
1015: +16(2 m_{\rm ud}  +m_{\rm s}) (2L_{6}-L_4) 
1016: \right) -\frac{2H^{\prime\prime}}{f_0^2}\right\}, 
1017: \label{eq:wchpt_mk}
1018: \\
1019: f_\pi &=&f_0\left\{
1020: 1-2\mu_\pi-\mu_K
1021: + \frac{2B_0}{f_0^2} \left (
1022: 8 m_{\rm ud} L_5+8(2 m_{\rm ud} +m_{\rm s})L_4
1023: \right) -\frac{2H^{\prime}}{f_0^2}\right\}, 
1024: \label{eq:wchpt_fpi}
1025: \\
1026: f_K &=&f_0\left\{
1027: 1-\frac{3}{4}\mu_\pi-\frac{3}{2}\mu_K-\frac{3}{4}\mu_\eta
1028: +\frac{2B_0}{f_0^2}\left(4(m_{\rm ud} + m_{\rm s}) L_5+8(2 m_{\rm ud} + m_{\rm s})L_4 
1029: \right) -\frac{2H^{\prime}}{f_0^2}\right\}, 
1030: \label{eq:wchpt_fk}
1031: \een
1032: \end{widetext}
1033: where the quark masses are defined by the axial-vector Ward-Takahashi
1034: identities: $m_{\rm ud}=m_{\rm ud}^{\rm AWI}$ and 
1035: $m_{\rm s}=m_{\rm s}^{\rm AWI}$. 
1036: $L_{4,5,6,8}$ are the low energy constants and 
1037: $\mu_{\rm PS}$ is the chiral logarithm defined by 
1038: \ben
1039: \mu_{\rm PS}=\frac{1}{16\pi^2}\frac{{\tilde m}_{\rm PS}^2}{f_0^2}
1040: \ln\left(\frac{{\tilde m}_{\rm PS}^2}{\mu^2}\right),
1041: \label{eq:chlog}
1042: \een
1043: where 
1044: \ben
1045: {\tilde m}_\pi^2 &=& 2 {m_{\rm ud}}B_0,\\
1046: {\tilde m}_K^2 &=& ({m_{\rm ud}} +m_{\rm s})B_0,\\
1047: {\tilde m}_\eta^2 &=& \frac{2}{3}({m_{\rm ud}} +2 m_{\rm s})B_0
1048: \een
1049: with $\mu$ the renormalization scale.
1050: 
1051: The two additional parameters $H^{\prime\prime}$ and $H^\prime$ are associated with
1052: the $O(a^2)$ contributions distinguishing the Wilson ChPT 
1053: from that in the continuum.
1054: Since these parameters are independent of the quark
1055: masses, their contributions can be absorbed into $B_0$ and
1056: $f_0$ by the following redefinitions:
1057: \ben
1058: B_0^\prime=B_0\left(1-\frac{2H^{\prime\prime}}{f_0^2}\right),\\
1059: f_0^\prime=f_0\left(1-\frac{2H^{\prime}}{f_0^2}\right),
1060: \een
1061: Indeed re-expansion of the terms in the curly brackets of 
1062: (\ref{eq:wchpt_mpi}) to (\ref{eq:wchpt_fk}) gives rise only to  
1063: terms of form $O(m_q\cdot a^2)$ and $O(m_q\ln m_q\cdot a^2)$, which are NNLO in
1064: the order counting of WChPT analysis and hence can be ignored. 
1065: Thus, up to NLO, WChPT formula are equivalent to the continuum form. 
1066: Note that the expressions in terms of the VWI
1067: quark masses take different forms
1068: and cannot be reduced to those of the continuum ChPT.
1069: Hereafter we concentrate on the continuum ChPT.
1070: 
1071: \subsection{SU(3) chiral perturbation theory}
1072: 
1073: The SU(3) ChPT formula in the continuum up to NLO\cite{chpt_nf3} is given by
1074: \begin{widetext}
1075: \ben
1076: \frac{m_{\pi}^2}{2 m_{\rm ud}}&=& B_0 \left\{
1077: 1+\mu_\pi-\frac{1}{3}\mu_\eta 
1078: +\frac{2B_0}{f_0^2} \left(
1079: 16 m_{\rm ud}  (2L_{8}-L_5) 
1080: +16 (2 m_{\rm ud} +m_{\rm s}) (2L_{6}-L_4)  
1081: \right) \right\}, 
1082: \label{eq:chpt_mpi}
1083: \\
1084: \frac{m_K^2}{(m_{\rm ud} +m_{\rm s})}&=&B_0 \left\{
1085: 1+\frac{2}{3}\mu_\eta
1086: +\frac{2B_0}{f_0^2}\left(
1087: 8(m_{\rm ud}+m_{\rm s}) (2L_{8}-L_5)
1088: +16(2 m_{\rm ud}  +m_{\rm s}) (2L_{6}-L_4) 
1089:  \right)\right\}, 
1090: \label{eq:chpt_mk}
1091: \\
1092: f_\pi &=&f_0\left\{
1093: 1-2\mu_\pi-\mu_K
1094: + \frac{2B_0}{f_0^2} \left (
1095: 8 m_{\rm ud} L_5+8(2 m_{\rm ud} +m_{\rm s})L_4
1096: \right)\right\}, 
1097: \label{eq:chpt_fpi}
1098: \\
1099: f_K &=&f_0\left\{
1100: 1-\frac{3}{4}\mu_\pi-\frac{3}{2}\mu_K-\frac{3}{4}\mu_\eta
1101: +\frac{2B_0}{f_0^2}\left(4(m_{\rm ud}+m_{\rm s}) L_5+8(2 m_{\rm ud}+ m_{\rm s})L_4 
1102: \right)\right\}, 
1103: \label{eq:chpt_fk}
1104: \een
1105: \end{widetext}
1106: There are six unknown low energy constants $B_0,f_0,L_{4,5,6,8}$ 
1107: in the expressions above.
1108: $L_{4,5,6,8}$ are scale-dependent so as to 
1109: cancel that of the chiral logarithm given by (\ref{eq:chlog}). 
1110: We can determine these parameters                                  
1111: by applying a simultaneous fit to $m_\pi^2/(2 m_{\rm ud})$, 
1112: $m_K^2/(m_{\rm ud} +m_{\rm s})$, $f_\pi$ and $f_K$.
1113: 
1114: 
1115: In order to provide an overview of our data we plot in Fig.~\ref{fig:cmp_chlog}
1116: a comparison of the PACS-CS (red symbols) and the CP-PACS/JLQCD 
1117: results (black symbols) for $m_\pi^2/m_{\rm ud}^{\rm AWI}$ and $f_K/f_\pi$ 
1118: as a function of $m_{\rm ud}^{\rm AWI}$.  
1119: The two data sets show a smooth connection at
1120: $\kappa_{\rm ud}=0.13700$ ($m_{\rm ud}^{\rm AWI}=0.028$). 
1121: More important is the fact that an almost linear quark mass dependence
1122: of the CP-PACS/JLQCD results in heavier quark mass region
1123: changes into a convex behavior,
1124: both for $m_\pi^2/m_{\rm ud}^{\rm AWI}$ and $f_K/f_\pi$, 
1125: as $m_{\rm ud}^{\rm AWI}$ is diminished in the PACS-CS results. 
1126: This is a characteristic feature expected from 
1127: the ChPT prediction in the small quark mass region
1128: due to the chiral logarithm.
1129: This curvature drives up the ratio $f_K/f_\pi$ toward the 
1130: experimental value as the physical point is approached.
1131: 
1132: 
1133: Having confirmed signals for the presence of the chiral logarithm, 
1134: we apply the SU(3) ChPT formulae (\ref{eq:chpt_mpi})$-$(\ref{eq:chpt_fk})
1135: to our results.  We choose the four simulation points at 
1136: $\kappa_{\rm ud}\ge 0.13754$. In Fig.~\ref{fig:cmp_chlog} these four 
1137: points lie to the left and around the turning point of the curvature.  
1138: They also correspond to the region where the $\rho$ meson 
1139: mass satisfies the condition $m_\rho > 2m_\pi$, and hence lie to the left 
1140: of the threshold singularity in the complex energy plane 
1141: for the $\rho$ meson.
1142: The heaviest pion mass at $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13754,0.13640)$
1143: is about 410~MeV with the use of $a^{-1}=2.176(31)$ GeV determined below.
1144: The measured bare AWI quark masses, but corrected for the $O(ma)$ corrections
1145: at one-loop perturbation theory,  
1146: are used for ${m_{\rm ud}}$ and $m_{\rm s}$ 
1147: in Eqs.~(\ref{eq:chpt_mpi})$-$(\ref{eq:chpt_fk}).
1148: 
1149: We present the fit results for the low energy constants 
1150: in Table~\ref{tab:fit_su3chpt}.  
1151: The results are quoted both without (w/o FSE) 
1152: and with (w/ FSE) finite-size corrections in the ChPT formulae 
1153: (see Sec~\ref{subsec:fse}).   We also list 
1154: the phenomenological estimates with the experimental
1155: inputs~\cite{colangelo05,amoros01}, and the results obtained 
1156: by recent 2+1 flavor lattice QCD calculations~\cite{rbcukqcd08,milc07}.  
1157: The renormalization scale is set to be 770~MeV. The MILC results
1158: for the low energy constants quoted at the scale of $m_\eta$ 
1159: are converted according to Ref.~\cite{chpt_nf3}
1160: \begin{widetext}
1161: \ben
1162: L_4(\mu)&=&L_4(m_\eta)-\frac{1}{256\pi^2}
1163: \ln\left(\frac{\mu^2}{m_\eta^2}\right),\\
1164: L_5(\mu)&=&L_5(m_\eta)-\frac{3}{256\pi^2}
1165: \ln\left(\frac{\mu^2}{m_\eta^2}\right),\\
1166: (2L_6-L_4)(\mu)&=&(2L_6-L_4)(m_\eta)-\left(\frac{2}{9}\right)\frac{1}{256\pi^2}
1167: \ln\left(\frac{\mu^2}{m_\eta^2}\right),\\
1168: (2L_8-L_5)(\mu)&=&(2L_8-L_5)(m_\eta)+\left(\frac{4}{3}\right)\frac{1}{256\pi^2}
1169: \ln\left(\frac{\mu^2}{m_\eta^2}\right)
1170: \een   
1171: \end{widetext}
1172: with $\mu$ the renormalization scale.
1173: For $L_4$ and $L_5$ governing the behavior of $f_\pi$ and $f_K$, 
1174: we find that all the results are compatible. 
1175: On the other hand, some discrepancies are observed for the results 
1176: of $2L_6-L_4$ and $2L_8-L_5$ contained in the ChPT formulae 
1177: for $m_\pi^2$ and $m_K^2$.
1178: 
1179:  
1180: For later convenience we convert the SU(3) low energy constants
1181: $B_0,f_0,L_{4,5,6,8}$ to the SU(2) low energy constants $B,f,
1182: l_{3,4}$ defined by
1183: \ben
1184: \frac{m_\pi^2}{2{m_{\rm ud}}}&=&B\left\{1+\mu_\pi(B_0\rightarrow B,
1185: f_0\rightarrow f)+4\frac{{\bar m}_\pi^2}{f^2}l_3\right\},\nn\\
1186: \label{eq:su2chpt_mpi}\\
1187: f_\pi&=&f\left\{1- 2\mu_\pi(B_0\rightarrow B, f_0\rightarrow f)
1188: +2\frac{{\bar m}_\pi^2}{f^2}l_4\right\}\nn\\
1189: \label{eq:su2chpt_fpi}
1190: \een
1191: with ${\bar m}_\pi^2=m_{\rm ud}B$.
1192: The NLO relations are given by\cite{chpt_nf3}
1193: \ben
1194: B&=&B_0\left(1-\frac{1}{3}{\bar \mu}_\eta+
1195: \frac{32 {\bar m}^2_{K}}{f_0^2}(2L_6-L_4) \right),\\
1196: f&=&f_0\left(1-{\bar \mu}_K+
1197: \frac{16 {\bar m}^2_{K}}{f_0^2}L_4 \right),\\
1198: l_3&=&-8L_4-4L_5+16L_6+8L_8-\frac{1}{18}{\bar \nu}_\eta,\\
1199: l_4&=&8L_4+4L_5-\frac{1}{2}{\bar \nu}_K,
1200: \een
1201: where ${\bar \mu}_{K,\eta}$ and ${\bar \nu}_{K,\eta}$ are defined by
1202: \ben
1203: {\bar \mu}_{K,\eta}=\frac{{\bar m}^2_{K,\eta}}{16\pi^2f_0^2}
1204: \ln\left(\frac{{\bar m}^2_{K,\eta}}{\mu^2}\right),\\
1205: {\bar \nu}_{K,\eta}=\frac{1}{32\pi^2}
1206: \left(\ln\left(\frac{{\bar m}^2_{K,\eta}}{\mu^2}\right)+1\right)
1207: \een
1208: with 
1209: \ben
1210: {\bar m}^2_{K}&=&m_{\rm s}B_0,\\
1211: {\bar m}^2_{\eta}&=&\frac{4}{3}m_{\rm s}B_0,
1212: \een
1213: and ${\bar l}_i$ ($i=3,4$) are defined at the renormalization scale
1214: $\mu=m_\pi=139.6$ MeV\cite{chpt_nf2}:
1215: \ben
1216: l_i&=&\frac{\gamma_i}{32\pi^2}\left({\bar l}_i+\ln\frac{m^2_\pi}{\mu^2}\right)
1217: \een
1218: with
1219: \ben
1220: \gamma_3&=&-\frac{1}{2},\\
1221: \gamma_4&=&2.
1222: \een
1223: In Table~\ref{tab:lec_su2_conv} we summarize the results for 
1224: the SU(2) low energy constants obtained by the conversion from the
1225: SU(3) low energy constants.
1226: The vacuum condensations are defined by
1227: \ben
1228: \langle {\bar u}u\rangle_0 &\equiv & \langle {\bar
1229: u}u\rangle\vert_{m_{\rm ud}=m_{\rm s}=0}=-\frac{1}{2}f_0^2 B_0,\\ 
1230: \langle {\bar u}u\rangle &\equiv & \langle {\bar
1231: u}u\rangle\vert_{m_{\rm ud}=0,m_{\rm s}=m_{\rm s}^{\rm physical}}
1232: =-\frac{1}{2}f^2B.
1233: %R_{\rm OZI}&=&\frac{\langle {\bar u}u\rangle}{\langle {\bar u}u\rangle_0}.
1234: \een
1235: These quantities are perturbatively renormalized
1236: at the scale of 2 GeV.
1237: 
1238: In Fig.~\ref{fig:l_34} we compare our results for 
1239: ${\bar l}_{3,4}$ with 
1240: those obtained by other groups 
1241: whose numerical values are listed in Table~\ref{tab:l_34}. 
1242: Black symbols denote the
1243: phenomenological estimates, blue symbols represent
1244: the results obtained by the SU(2) ChPT fit on 
1245: 2 flavor dynamical configurations and red closed (open) symbols
1246: are for those obtained by the SU(3) (SU(2)) ChPT fit
1247: on 2+1 flavor dynamical configurations.
1248: For ${\bar l}_3$ all the results 
1249: reside between 3.0 and 3.5, except for the MILC result which is 
1250: sizably smaller and marginally consistent with others within a large error. 
1251: On the other hand, we find a good consistency
1252: among the results for ${\bar l}_4$. 
1253: 
1254: 
1255: We have found that the SU(3) ChPT fit gives reasonable 
1256: values for the low energy constants.  However, we are concerned with a rather
1257: large value of $\chi^2$/dof=4.2(2.9) (see Table~\ref{tab:fit_su3chpt}).
1258: Figures~\ref{fig:su3fit_mps} and \ref{fig:su3fit_fps}
1259: show how well the
1260: data for $m_\pi^2/m_{\rm ud}$,  $2m_K^2/(m_{\rm ud}+m_{\rm s})$, $f_\pi$
1261: and $f_K$ are described by the SU(3) ChPT up to NLO. 
1262: The filled and open circles are our data, and the fit results are plotted 
1263: by blue triangles.
1264: % 
1265: We note in passing that , 
1266: for the Wilson-clover quark action, $m_s^{\rm AWI}$ 
1267: varies at $O(a^2)$ as $m_{\rm ud}$ varies even if $\kappa_{\rm s}$ is 
1268: held fixed. Thus we are not able to draw a line with a fixed value 
1269: for $m_s^{\rm AWI}$.
1270: %
1271: The blue star symbols represent the extrapolated values at the
1272: physical point whose determination will be explained below
1273: in Sec.~\ref{sec:physicalpt}. 
1274: 
1275: The points around $m_{\rm ud}^{\rm AWI}\approx 0.01$ corresponds to 
1276: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13754,0.13640)$ and $(0.13754,0.13660)$. 
1277: Marked deviations between circles and triangles show that the SU(3) ChPT poorly 
1278: accounts for the strange quark mass dependence of $f_\pi$ and $f_K$.
1279: This flaw is mainly responsible for the large value of $\chi^2$/dof.
1280: 
1281: In order to investigate the origin of discrepancy between the data and the fit 
1282: more closely,  we draw the relative magnitude of the NLO contribution to the 
1283: LO one for $m_\pi^2/m_{\rm ud}$,  $2m_K^2/(m_{\rm ud}+m_{\rm s})$, $f_\pi$ and $f_K$
1284: as a function of $m_{\rm ud}^{\rm AWI}$ in Figs.~\ref{fig:su3nlo2lo_mps} 
1285: and \ref{fig:su3nlo2lo_fps}.  The strange quark 
1286: mass is fixed at the physical value, and the contributions from $\pi$, $K$ and 
1287: $\eta$ loops are separately drawn. 
1288: The relative magnitudes are at most 10\% for
1289: $m_\pi^2/m_{\rm ud}$ and $2m_K/(m_{\rm ud}+m_{\rm s})$.  We find, however, 
1290: significant NLO contributions for the decay constants. 
1291: For $f_\pi$ the relative magnitude rapidly increases from 10\% at 
1292: $m_{\rm ud}=0$ to 40\% at around $m_{\rm ud}=0.01$. 
1293: The situation is worse for $f_K$ for which the NLO contribution is about 40\% of 
1294: the LO one even at $m_{\rm ud}=0$, most of which arises from the $K$ loop. 
1295: 
1296: 
1297: \subsection{SU(2) chiral perturbation theory}
1298: 
1299: The bad convergences of the chiral expansions 
1300: for $f_\pi$ and $f_K$ tell us that the strange quark
1301: mass is not light enough to be appropriately treated by the NLO SU(3)
1302: ChPT. There are two alternative choices for further chiral
1303: analysis. One is to extend SU(3) ChPT to NNLO, and the other is to use 
1304: SU(2) ChPT with the aid of an analytic expansion for the strange
1305: quark contribution around the physical strange quark mass.
1306: 
1307: The former method, which has been employed by the MILC collaboration 
1308: in an incomplete fashion\cite{milc07}, is very demanding: 
1309: we cannot determine the additional low energy constants at NNLO without 
1310: significantly increasing the data points.  There is in addition no guarantee
1311: that the expansion is controlled at NNLO. 
1312: We therefore consider that the latter route is more natural.
1313: This alternative was employed by the RBC/UKQCD collaboration\cite{rbcukqcd08}. 
1314: Since they had data only at a single strange quark mass, they could not study 
1315: the strange quark mass dependence.  This we shall do with our data 
1316: thanks to the second choice of 
1317: the strange quark mass at $\kappa_{\rm ud}=0.13754$.
1318: 
1319: For $m_\pi$ and $f_\pi$ the SU(2) ChPT formulae of (\ref{eq:su2chpt_mpi}) 
1320: and (\ref{eq:su2chpt_fpi}) are employed.  The low energy constants $B$
1321: and $f$ are functions of the strange quark mass.  Assuming that we run 
1322: simulations close enough around the physical point for the strange quark mass 
1323: so that a linear expansion in $m_{\rm s}$ is sufficient, we write 
1324: $B=B_s^{(0)}+m_{\rm s}B_s^{(1)}$ and $f=f_s^{(0)}+m_{\rm s}f_s^{(1)}$, 
1325: where it should be noted that $B_s^{(0)}\ne B_0$ and $f_s^{(0)}\ne f_0$.
1326: 
1327: For the kaon sector we treat the $K$ mesons as matter fields 
1328: in the isospin $1/2$ linear representation, and couple pions 
1329: in SU(2) invariant ways (see, {\it e.g.}, Ref.~\cite{roessl}).
1330: For $m_K$ and $f_K$ this leads to the following fit formulae: 
1331: \ben
1332: m_K^2&=&\alpha_m+\beta_m m_{\rm ud}+\gamma_m m_{\rm s},
1333: \label{eq:su2chpt_mk}\\
1334: %f_K&=&\bar f\nn\\
1335: f_K&=&{\bar f} \left\{1+\beta_f m_{\rm ud} -\frac{3}{4}\frac{{\tilde m}_\pi^2}{16\pi^2 f^2}
1336: \ln\left(\frac{{\tilde m}_\pi^2}{\mu^2}\right) \right\}
1337: \label{eq:su2chpt_fk}
1338: \een
1339: with ${\bar f}={\bar f}_s^{(0)}+m_{\rm s}{\bar f}_s^{(1)}$. 
1340: In these formulae, the linear expansion in $m_{\rm s}$ should be regarded 
1341: as that around the physical strange quark mass.  
1342: 
1343: 
1344: We apply a simultaneous fit to $m_\pi$, $f_\pi$ and $f_K$ 
1345: employing the formulae of Eqs.~(\ref{eq:su2chpt_mpi}), (\ref{eq:su2chpt_fpi})
1346: and (\ref{eq:su2chpt_fk}).  The kaon mass 
1347: $m_K^2$ is independently fitted according to Eq.~(\ref{eq:su2chpt_mk}). 
1348: Calling the four data points corresponding to $\kappa_{\rm ud}\geq 0.13754$
1349: as Range I, the fit results for $B,f,{\bar l}_3,{\bar l}_4$ at the physical 
1350: strange quark mass are presented in
1351: Table~\ref{tab:fit_su2chpt} and Fig.~\ref{fig:l_34} both without and with 
1352: finite-size corrections. 
1353: We find that they are consistent with
1354: those obtained by the NLO conversion from the SU(3) low energy constants
1355: given in Table~\ref{tab:lec_su2_conv}.
1356: Although our result for $\la {\bar u}u\ra$ is about 50\% smaller than that 
1357: of RBC/UKQCD, the difference comes from estimates of 
1358: the renormalization factor: we use one-loop perturbation while they 
1359: employ the nonperturbative RI-MOM scheme.
1360: This is verified by the observation that
1361: the value of $f$ and the renormalization-free quantities $m_{\rm ud} B$
1362: and $m_{\rm s} B$ show consistency
1363: between our results and those of RBC/UKQCD.
1364: 
1365: Figures~\ref{fig:su2fit_mpi} and \ref{fig:su2fit_fps} show that 
1366: the quark mass dependences of $m_\pi^2/m_{\rm ud}^{\rm AWI}$, 
1367: $f_\pi$ and $f_K$
1368: are reasonably described by the SU(2) ChPT formulae of
1369: (\ref{eq:su2chpt_mpi}), (\ref{eq:su2chpt_fpi}) 
1370: and (\ref{eq:su2chpt_fk}). The resulting  $\chi^2$/dof is 0.33(72),
1371: which is an order of magnitude smaller than in the SU(3) case. 
1372: In Fig.~\ref{fig:su2nlo2lo} we illustrate the relative magnitude of  
1373: the NLO contribution against the LO value for 
1374: $m_\pi^2/m_{\rm ud}$, $f_\pi$ and $f_K$ as a function of $m_{\rm ud}^{\rm AWI}$
1375: fixing the strange quark mass at the physical value. 
1376: The convergences for $f_\pi$ and $f_K$ are clearly better than the SU(3) case.
1377: 
1378: In order to investigate the stability of the fit, we try two additional 
1379: choices of the data sets for the SU(2) ChPT fit: 
1380: Range II ($\kappa_{\rm ud}$,$\kappa_{\rm s}$)=(0.13781,0.13640),(0.13770,0.13640),
1381: (0.13754,0.13640),(0.13754,0.13660),(0.13727,0.13640) includes one more data 
1382: at a heavier pion mass added to Range I, and Range III 
1383: ($\kappa_{\rm ud}$,$\kappa_{\rm s}$)=(0.13770,0.13640),
1384: (0.13754,0.13640),(0.13754,0.13660),(0.13727,0.13640) removes the point with 
1385: the lightest pion mass from Range II.  
1386: The results for $B,f,{\bar l}_3,{\bar l}_4$ and corresponding 
1387: $\chi^2$/dof are given in Table~\ref{tab:fit_su2chpt}.
1388: While inclusion of the data at $\kappa_{\rm ud}=0.13727$
1389: increases the value of $\chi^2$/dof, the results for $B,f,{\bar l}_3,{\bar
1390: l}_4$ are consistent among the three cases within the error bars.
1391: 
1392: 
1393: \subsection{Finite size effects based on chiral perturbation theory}
1394: \label{subsec:fse}
1395: 
1396: We evaluate finite-size effects based on the NLO formulae of ChPT.
1397: In the case of SU(3) ChPT the finite size effects defined by 
1398: $R_X=(X(L)-X(\infty))/X(\infty)$ for $X=m_\pi,m_K,f_\pi,f_K$ are given 
1399: by~\cite{colangelo05}:
1400: \ben
1401: R_{m_\pi}%=\frac{M_{\pi}(L)-M_{\pi}}{M_{\pi}}
1402: &=&\frac{1}{4}\xi_{\pi}\tilde g_1(\lambda_\pi)-\frac{1}{12}\xi_{\eta}\tilde g_1(\lambda_\eta),\\
1403: R_{m_K} %=\frac{M_K(L)-M_K}{M_K}
1404: &=&\frac{1}{6}\xi_{\eta}\tilde g_1(\lambda_\eta),\\
1405: R_{f_\pi} %=\frac{F_{\pi}(L)-F_{\pi}}{F_{\pi}}
1406: &=&-\xi_{\pi}\tilde g_1(\lambda_\pi)-\frac{1}{2}\xi_{K}\tilde g_1(\lambda_K),\\
1407: R_{f_K} %=\frac{F_K(L)-F_K}{F_K}
1408: &=&-\frac{3}{8}\xi_{\pi}\tilde g_1(\lambda_\pi)-\frac{3}{4}\xi_{K}\tilde g_1(\lambda_K)
1409: -\frac{3}{8}\xi_{\eta}\tilde g_1(\lambda_\eta)\nn\\
1410: \een
1411: with
1412: \ben
1413: \xi_{\rm PS} &\equiv& \frac{2m_{\rm PS}^2}{(4\pi f_\pi)^2},\\
1414: \lambda_{\rm PS} &\equiv& m_{\rm PS} L, \\
1415: \tilde g_1(x)&=&\sum_{n=1}^{\infty}\frac{4m(n)}{{\sqrt n }x} K_1({\sqrt n} x), 
1416: \een
1417: where $K_1$ is the Bessel function of the second kind and $m(n)$
1418: denotes the multiplicity of the partition $n=n_x^2+n_y^2+n_z^2$. 
1419: The authors in Ref.~\cite{colangelo05} expect that the above formulea
1420: are valid for $m_\pi L>2$, in which our simulation points reside. 
1421: In Figs.~\ref{fig:su3fit_mps} and \ref{fig:su3fit_fps}
1422: we also plotted the ChPT fit results including finite size effects.
1423: The results are almost degenerate with the fit results 
1424: without finite size effects except at the lightest simulation point 
1425: at $\kappa_{\rm ud}=0.13781$ and the
1426: extrapolated values at the physical point. This feature is understood
1427: by looking at Fig.~\ref{fig:fse_su3} where we plot the magnitude of $R_X$ for 
1428: $X=m_\pi,m_K,f_\pi,f_K$ with $L=2.9$ fm as a function of $m_\pi$ 
1429: keeping the strange quark mass fixed at the physical value.
1430: %It should be noted that $R_{m_{\rm PS}}>0$ and $R_{f_{\rm PS}}<0$.
1431: The expected finite size effects are less than 2\% for $m_{\rm PS}$ and
1432: $f_{\rm PS}$ at our simulation points. For $m_{\rm PS}$ this is true 
1433: even at the physical point, while the value of $f_\pi$ is 
1434: decreased by 4\% due to the finite size effects. 
1435: 
1436: We can repeat the above study for the SU(2) case. 
1437: The NLO formulae for $m_\pi$ and $f_\pi$ are given by\cite{colangelo05}
1438: \ben
1439: R^\prime_{m_\pi}%=\frac{M_{\pi}(L)-M_{\pi}}{M_{\pi}}
1440: &=&\frac{1}{4}\xi_{\pi}\tilde g_1(\lambda_\pi),\\
1441: R^\prime_{f_\pi} %=\frac{F_{\pi}(L)-F_{\pi}}{F_{\pi}}
1442: &=&-\xi_{\pi}\tilde g_1(\lambda_\pi).
1443: \een
1444: In Figs.~\ref{fig:su2fit_mpi} and \ref{fig:su2fit_fps}
1445: we hardly detect finite size effects for $m_{\rm ud}^{\rm AWI}>0.001$.   
1446: Figure~\ref{fig:fse_su2} shows $R_X^\prime$ for 
1447: $X=m_\pi,f_\pi$ with $L=2.9$ fm as a function of $m_\pi$.
1448: The situation is similar to the SU(3) case: although finite size
1449: effects increase as $m_\pi$ decreases, their magnitudes are at most 2\%
1450: for $m_\pi$ and 4\% for $f_\pi$
1451: even at the physical point, which is easily expected 
1452: by comparing the expressions of $R$ and $R^\prime$.
1453: 
1454: Let us add a cautionary note that the finite-size formulae analyzed 
1455: here lose viability when $m_\pi L$ becomes too small.  Precisely at 
1456: what values of $m_\pi L$ this takes place is not well controlled 
1457: theoretically, however.  Direct simulations on a larger lattice 
1458: is required to pin down the actual magnitude of finite-size effects at the 
1459: physical point.  The need for such calculations are even more for baryons 
1460: whose sizes are larger than mesons.    
1461: 
1462: 
1463: \section{Results at the physical point}
1464: \label{sec:physicalpt}
1465: 
1466: We need three physical inputs to determine the up-down and 
1467: the strange quark masses and the lattice cutoff.
1468: We choose $m_\pi$, $m_K$ and $m_\Omega$.  The choice of 
1469: $m_\Omega$ has both theoretical and practical advantages:
1470: the $\Omega$ baryon is stable in the strong interactions
1471: and its mass, being composed of three strange quarks, 
1472: is determined with good precision with small finite 
1473: size effects. 
1474: 
1475: For the pseudoscalar meson sector, we employ SU(2) chiral expansion 
1476: as explained in the previous Section.
1477: For the vector mesons and the baryons we use a simple linear formula
1478: $m_{\rm had}=\alpha_h+\beta_h m_{\rm ud}^{\rm AWI}+\gamma_h m_{\rm s}^{\rm
1479: AWI}$, employing the data set in the same range   
1480: $\kappa_{\rm ud}\ge 0.13754$ as for the pseudoscalar meson sector.
1481: We do not rely on heavy meson effective 
1482: theory (HMET)\cite{hmet} or heavy baryon ChPT (HBChPT)\cite{hbchpt}
1483: since they show very poor convergences even 
1484: at the physical point\cite{convergence}.
1485: In Figs.~\ref{fig:chexp_vector},  \ref{fig:chexp_octet} and 
1486: \ref{fig:chexp_decuplet},
1487: we show linear chiral extrapolations
1488: of the vector meson, the octet and the decuplet baryon masses,
1489: respectively. Blue symbols represent the fit results at the measured values of
1490: $m_{\rm ud}^{\rm AWI}$. The extrapolated values at the 
1491: physical point are also denoted by blue star symbols, which should be
1492: compared with the experimental values plotted at $m_{\rm ud}^{\rm AWI}=0$. 
1493: 
1494: Since the linear fit is applied to the data set at $\kappa_{\rm ud}\ge
1495: 0.13754$, blue symbols at  $\kappa_{\rm ud}< 0.13754$ express the
1496: predictions from the fit results.
1497: We observe that the quark mass dependence of $m_\Omega$ 
1498: is remarkably well described by the linear function, which assures
1499: that $m_\Omega$ is a good quantity for the physical input in the sense
1500: that its chiral behavior is easily controlled.
1501: 
1502: The results for the physical quark masses and the lattice cutoff are listed in
1503: Table~\ref{tab:physicalpt_qmass},
1504: where the errors are statistical. They are provided
1505: with and without the finite size corrections based on the NLO SU(2) ChPT
1506: analyses. Both results are almost degenerate.
1507: We find that our quark masses
1508: are smaller than the estimates in the recent 2+1 flavor 
1509: lattice QCD calculations\cite{rbcukqcd08,milc07}.  We note, however, 
1510: that we employ the perturbative renormalization factors at one-loop level 
1511: which should contain an uncertainty.  A nonperturbative calculation 
1512: of the renormalization factor is in progress using the Schr\"odinger 
1513: functional scheme. 
1514: 
1515: In Table~\ref{tab:physicalpt_qmass} we also present
1516: the results for the pseudoscalar meson decay constants
1517: at the physical point
1518: using the physical quark masses and the cutoff determined above, 
1519: which should be compared with the experimental values 
1520: $f_\pi=130.7$~MeV, $f_K=159.8$~MeV, $f_K/f_\pi = 1.223$\cite{pdg}.
1521: We observe a good consistency within the error of 2--3\%.  The
1522: ratio is 3\% smaller than the experimental value in the case of the
1523: SU(2) ChPT fit with the finite size corrections. 
1524: A nonperturbative calculation of $Z_A$ is also in progress. 
1525: 
1526: In Fig.~\ref{fig:spectrum} the light hadron spectrum 
1527: extrapolated to the physical point using SU(2) ChPT with the finite
1528: size corrections are compared with the experimental values.
1529: Numerical values with and without the
1530: finite size corrections are listed in Table~\ref{tab:physicalpt_hmass}.
1531: The largest discrepancy between our results and the
1532: experimental values is at most 3\%, albeit errors are still not small 
1533: for the $\rho$ meson, the nucleon and the $\Delta$ baryon. 
1534: The results are clearly encouraging, but  
1535: further work is needed to remove the cutoff errors of $O((a\Lambda_{\rm QCD})^2)$.
1536: 
1537: 
1538: \section{Static quark potential}
1539: \label{sec:potential}
1540: 
1541: In addition to the hadronic observables presented so far,
1542: we also calculate the Sommer scale which is
1543: a popular gluonic observable.
1544: In order to calculate the static quark potential we measure 
1545: the temporal and the spatial Wilson loops with the use of the
1546: smearing procedure of Ref.~\cite{smear}.
1547: The number of smearing steps is determined to be 20 after
1548: examining the sufficient overlap of the Wilson loops onto the ground state. 
1549: The potential $V(r)$ is extracted from the Wilson loops 
1550: applying a correlated fit of the form 
1551: \ben
1552: W(r,t)=C(r)\exp(-V(r)t),
1553: \een
1554: where the same fitting range $[t_{\rm min},t_{\rm max}]=[5,8]$ is 
1555: chosen for all the simulations after investigating the 
1556: effective potential
1557: \ben
1558: V_{\rm eff}(r,t)=\ln\left[\frac{W(r,t)}{W(r,t+1)}\right].
1559: \een
1560: Figure~\ref{fig:v_eff} shows a typical case of $V_{\rm eff}(r,t)$ 
1561: with $r=4,8,12$ at $\kappa_{\rm ud}=0.13770$.
1562: We find that plateau starts at $t=4$ and signals are lost beyond $t=7$.
1563: A result of $V(r)$ at $\kappa_{\rm ud}=0.13770$ is plotted
1564: in Fig.~\ref{fig:potential} as a representative case. 
1565: Since good rotational symmetry 
1566: and no sign of the string breaking are observed, 
1567: we employ the following fitting form for the potential:
1568: \ben
1569: V(r)=V_0-\frac{\alpha}{r}+\sigma r,
1570: \label{eq:potential}
1571: \een
1572: where $V_0$, $\alpha$, $\sigma$ are unknown parameters.
1573: The fitting range is $[r_{\min},r_{\max}] = [3,16]$.
1574: 
1575: The Sommer scale $r_0$ is a phenomenological quantity 
1576: defined by
1577: \ben
1578: r_0^2=\left.\frac{dV(r)}{dr}\right\vert_{r=r_0}=1.65.
1579: \een
1580: Given Eq.~(\ref{eq:potential}) we obtain
1581: \ben
1582: r_0=\sqrt{\frac{1.65-\alpha}{\sigma}}.
1583: \een
1584: In Table~\ref{tab:potential} we list the results for $r_0$  
1585: including the systematic errors due to the choices of $t_{\rm min}$ and 
1586: $r_{\rm min}$. 
1587: 
1588: At $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13700, 0.13640)$, our 
1589: result is compared with those of CP-PACS/JLQCD \cite{cp-pacs/jlqcd3} in 
1590: Table~\ref{tab:r0comp}.  The two results are in reasonable agreement
1591: given the sizable magnitude of systematic errors caused by 
1592: the shortness of plateau of effective masses for potentials.
1593: 
1594: In order to extrapolate $r_0$ to the physical point
1595: we employ a linear form
1596: $1/r_0=\alpha_r+\beta_r\cdot m_{\rm ud}^{\rm AWI}+\gamma_r\cdot m_{\rm s}^{\rm AWI}$
1597: for the data set at $\kappa_{\rm ud}\ge 0.13754$.
1598: We illustrate the chiral extrapolation in Fig.~\ref{fig:chexp_r0},
1599: where the fit results are plotted by red triangles at the measured 
1600: values of $m_{\rm ud}^{\rm AWI}$. 
1601: The extrapolated result of $r_0$ at the physical point 
1602: is 5.427(51)(+81)($-2$), which is
1603: 0.4921(64)(+74)($-2$) fm in physical units 
1604: with the aid of $a^{-1}=2.176(31)$ GeV. The first error is statistical 
1605: and the second and the third ones are the 
1606:  systematic uncertainties originating from 
1607: the choice of $t_{\rm min}$ and $r_{\rm min}$, respectively.
1608: 
1609: 
1610: \section{Conclusion}
1611: \label{sec:conclusion}
1612: 
1613: We have presented the first results of the PACS-CS project which aims at 
1614: a 2+1 flavor lattice QCD simulation at the physical point
1615: using the $O(a)$-improved Wilson quark action. 
1616: The DDHMC algorithm, coupled with several algorithmic improvements, 
1617: have enabled us to reach $m_{\pi}=156$~MeV,  which corresponds to 
1618: $m_{\rm ud}^{\overline{\rm MS}}(\mu=2{\rm ~GeV})= 3.6$~MeV. 
1619: We are almost on the physical point, except that the strange quark mass
1620: is about 20\% larger than the physical value.
1621: 
1622: We clearly observe the characteristic features of the chiral logarithm 
1623: in the ratios $m_\pi^2/m_{\rm ud}^{\rm AWI}$ and $f_K/f_\pi$. 
1624: We find that our data are not well described by the NLO SU(3) ChPT, 
1625: due to bad convergence of the strange quark contributions. 
1626: We instead employ the NLO SU(2) ChPT for $m_\pi$ and
1627: $f_\pi$, and an analytic expansion around the physical strange quark mass
1628: for $m_K$ and $f_K$ in order to estimate the physical point. 
1629: The low energy constants obtained in this way are compatible
1630: with phenomenological estimates and other recent lattice calculations. 
1631: 
1632: Thanks to the enlarged physical volume 
1633: compared to the previous CP-PACS/JLQCD work,
1634: we obtain good signals not only for the meson masses 
1635: but also for the baryon masses. 
1636: After linear chiral extrapolations of the vector and baryon masses
1637: the hadron spectrum at the physical point 
1638: shows a good agreement with the experimental values,
1639: albeit some of the hadrons have rather large errors and scaling violations 
1640: remain to be examined.
1641: We find smaller values for the physical quark masses 
1642: compared to the recent estimates in the literature.
1643: This may be due to the one-loop estimate of the renormalization factor.
1644: 
1645: At present the simulation at the physical point is under way, and
1646: the statistics of the run at $\kappa_{\rm ud}=0.13781$ is being accumulated.
1647: We are evaluating the nonperturbative renormalization factors
1648: for the quark masses and the pseudoscalar meson decay constants in 
1649: order to remove perturbative uncertainties.
1650: 
1651: Once these calculations are accomplished,
1652: the next step is to investigate the finite size effects 
1653: at the physical point,  and then to reduce the discretization errors by 
1654: repeating the calculations at finer lattice spacings. 
1655: 
1656: 
1657: 
1658: \begin{acknowledgments}
1659: Numerical calculations for the present work have been carried out
1660: on the PACS-CS computer 
1661: under the ``Interdisciplinary Computational Science Program'' of 
1662: Center for Computational Sciences, University of Tsukuba. 
1663: We thank T. Sakurai and H. Tadano for a series of informative discussions 
1664: on single precision acceleration of the solver.  
1665: One of the authors (Y.K.) thank
1666: A.~Kennedy for valuable discussions on the algorithmic improvements.
1667: A part of the code development has been carried out on Hitachi SR11000 
1668: at Information Media Center of Hiroshima University. 
1669: This work is supported in part by Grants-in-Aid for Scientific Research
1670: from the Ministry of Education, Culture, Sports, Science and Technology
1671: (Nos.
1672: %13135204, 
1673: %15540251, 
1674: 16740147,   %Ishikawa
1675: 17340066,   %Kanaya
1676: %17540259, 
1677: 18104005,   %Ukawa
1678: 18540250,   %Kuramashi
1679: 18740130,   %Taniguchi
1680: %18740139,   
1681: 19740134,   %Izubuchi
1682: 20340047,   %Aoki
1683: 20540248,   %Ishizuka
1684: 20740123,   %Ukita
1685: 20740139    %Ishikawa
1686: ).
1687: \end{acknowledgments}
1688: 
1689: \input{appendix}
1690: 
1691: %\newpage %Just because of unusual number of tables stacked at end
1692: \bibliography{apssamp}% Produces the bibliography via BibTeX.
1693: 
1694: \input{refs}
1695: 
1696: \clearpage
1697: %\newpage
1698: 
1699: 
1700: \input{tables}
1701: 
1702: \clearpage
1703: %\newpage
1704: 
1705: \input{figs}
1706: 
1707: \end{document}
1708: