0711.3938/ms.tex
1: % Please use the skeleton file you have received in the 
2: % invitation-to-submit email, where your data are already
3: % filled in. Otherwise please make sure you insert your 
4: % data according to the instructions in PoSauthmanual.pdf
5: \documentclass{PoS}
6: 
7: 
8: %%%%% Personal Macros %%%%%%%%%%%%%%%%%%%
9: %\newcommand{\nn}{\nonumber}
10: \def\dfrac#1#2{\displaystyle\frac{#1}{#2}}
11: %\newcommand{\bm}[1]{\mbox{\boldmath $#1$}}
12: \newcommand{\ovl}[1]{\overline{#1}}
13: \newcommand{\wt}[1]{\widetilde{#1}}
14: \newcommand{\eq}[1]{eq.(\ref{#1})}
15: \newcommand{\eqn}[1]{(\ref{#1})}
16: \newcommand{\p}{\partial}
17: \newcommand{\slas}[2]{{{#1}\hspace{-5pt}{/}}_{#2}}
18: \newcommand{\slal}[2]{{{#1}\hspace{-8pt}{/}}_{#2}}
19: \newcommand{\kslash}{k\kern-1ex /}
20: \newcommand{\pslash}{p\kern-1ex /}
21: \newcommand{\qslash}{q\kern-1ex /}
22: \newcommand{\lslash}{l\kern-1ex /}
23: \newcommand{\sslash}{s\kern-1ex /}
24: %\newcommand{\Dslash}{{\cal D}\kern-1.2ex /}
25: \newcommand{\Dslash}{D\kern-1.2ex /}
26: \newcommand{\bpsi}{\overline{\psi}}
27: \newcommand{\bc}{\overline{c}}
28: \newcommand{\tr}{{\rm tr}}
29: %\newcommand{\vev}[1]{\langle #1 \rangle}
30: \newcommand{\VEV}[1]{\left\langle{\rm T} #1\right\rangle}
31: \newcommand{\beqa}{\begin{eqnarray}}
32: \newcommand{\eeqa}{\end{eqnarray}}
33: %\newcommand{\eqn}[1]{(\ref{#1})}
34: \newcommand{\Tr}{{\rm Tr}}
35: %\newcommand{\vev}[1]{\langle #1 \rangle}
36: \newcommand{\vev}[1]{\left\langle #1 \right\rangle}
37: %\newcommand{\be}{\begin{equation}}
38: %\newcommand{\ee}{\end{equation}}
39: \newcommand{\be}{\[}
40: \newcommand{\ee}{\]}
41: \newcommand{\bd}{\begin{description}}
42: \newcommand{\ed}{\end{description}}
43: \newcommand{\la}{\langle}
44: \newcommand{\ra}{\rangle}
45: \newcommand{\ben}{\begin{eqnarray}}
46: \newcommand{\een}{\end{eqnarray}}
47: \newcommand{\nn}{\nonumber}
48: \def\lsim{\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
49: \def\gsim{\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
50: \def\simgt{\rlap{\lower 6.0 pt\hbox{$\mathchar \sim$}}\raise 2.5pt \hbox {$>$}}
51: \def\simlt{\rlap{\lower 6.0 pt\hbox{$\mathchar \sim$}}\raise 2.5pt \hbox {$<$}}
52: \newcommand{\cont}{{\rm cont}}
53: \newcommand{\latt}{{\rm latt}}
54: \newcommand{\tad}{{\rm tad}}
55: \newcommand{\wi}{{\rm WI}}
56: \newcommand{\mf}{{\rm MF}}
57: \newcommand{\msbar}{{\overline {\rm MS}}}
58: \newcommand{\spr}{{s^\prime}}
59: \newcommand{\csw}{{c_{\rm SW}}}
60: \newcommand{\hk}{{\hat k}}
61: \newcommand{\mm}{{\cal M}}
62: \newcommand{\ce}{{\it c}_{\it E}}
63: \newcommand{\cb}{{\it c}_{\it B}}
64: \newcommand{\lqcd}{{\Lambda}_{\rm QCD}}
65: %\newcommand{\pos}{{p^*}}
66: %\newcommand{\qos}{{q^*}}
67: %\newcommand{\qoss}{{q^*_s}}
68: %\newcommand{\qosd}{{q^*_d}}
69: \newcommand{\lo}{{(0)}}
70: \newcommand{\nlo}{{(1)}}
71: \newcommand{\mplo}{{m_p^{(0)}}}
72: \newcommand{\intlat}{{\int_{-\pi}^{\pi}\frac{d^4 k}{(2\pi)^4}}}
73: \newcommand{\mplomf}{{{\tilde m_p}^{(0)}}}
74: \newcommand{\dmplomf}{{{\tilde m_p}^{\tad}}}
75: \newcommand{\mpl}{{m_{p2}}}
76: \newcommand{\mph}{{m_{p1}}}
77: \newcommand{\mpllo}{{m_{p2}^{(0)}}}
78: \newcommand{\mphlo}{{m_{p1}^{(0)}}}
79: \newcommand{\mpllomf}{{{\tilde m_{p2}}^{(0)}}}
80: \newcommand{\mphlomf}{{{\tilde m_{p1}}^{(0)}}}
81: \def\ovec{\partial_\mu\hspace{-0.4cm}\raisebox{1.8ex}{$\rightarrow$}}
82: \def\antivec{\partial_\mu\hspace{-0.4cm}\raisebox{1.8ex}{$\leftarrow$}}
83: \newcommand{\crd}{\color{black}}
84: \newcommand{\crw}{\color{white}}
85: \newcommand{\crb}{\color{blue}}
86: \newcommand{\crr}{\color{red}}
87: \newcommand{\crm}{\color{magenta}}
88: \newcommand{\crg}{\color{green}}
89: \newcommand{\op}{{\cal O}}
90: \newcommand{\ceff}{{c_{\rm eff}}}
91: %%%%%%%%%%%%%% END OF MACROS %%%%%%%%%%%%%%%%%%
92: 
93: \title{$N_f=2+1$ dynamical Wilson quark simulation toward the physical point}
94: 
95: \ShortTitle{$N_f=2+1$ dynamical Wilson quark simulation toward the physical point}
96: 
97: \author{Yoshinobu Kuramashi\thanks{E-mail: 
98: kuramasi@het.ph.tsukuba.ac.jp}\hspace{3mm}for the PACS-CS Collabolation \\
99: Center for Computational Sciences and
100: %University of Tsukuba, Tsukuba, Ibaraki 305-8577, Japan\\
101: Graduate School of Pure and Applied Sciences,\\ 
102: University of Tsukuba, Tsukuba, Ibaraki 305-8571, Japan}
103: \abstract{
104: We present preliminary results of the
105: PACS-CS project which simulates 2+1
106: flavor lattice QCD toward the physical point 
107: with the nonperturbatively $O(a)$-improved
108: Wilson quark action and the Iwasaki gauge action.
109: Calculations are carried out at $\beta=1.9$ on a
110: $32^3\times 64$ lattice with the use of the domain-decomposed HMC algorithm
111: to reduce the up-down quark mass.
112: The resulting pseudoscalar meson masses range from 730~MeV down to 210~MeV.
113: We discuss the physical results including the chiral analysis in
114: the pseudoscalar meson sector and the hadron spectrum.
115: Some algorithmic issues are also discussed.
116: %Other recent unquenched results  
117: %using the Wilson-type quarks are briefly reviewed.
118: }
119: 
120: 
121: \FullConference{The XXV International Symposium on Lattice Field Theory\\
122: 		 July 30-4 August 2007\\
123: 		 Regensburg, Germany}
124: 
125: \begin{document}
126: 
127: \section{Introduction}
128: 
129: The first lattice QCD calculation of hadron masses in 1981
130: revealed its potential ability to nonperturbatively evaluate the
131: physical quantities in the strong interaction from  first principles.
132: Since then, the history of lattice QCD has been a succession of enduring efforts
133: to control the major systematic errors due to finite lattice size,
134: finite lattice spacing, quenching and chiral extrapolation.  
135: Thanks to recent progress of simulation algorithms and increasing
136: availability of computational resources, we are about to
137: bring all the above systematic errors under control. 
138: This will allow us to establish whether or not QCD is the fundamental theory of the strong
139: interaction by investigating the hadron spectrum, and further proceed to elucidate 
140: the fundamental issues of the strong interactions and the Standard Model.
141:  
142: The previous CP-PACS/JLQCD project\cite{cppacs/jlqcd1,
143: cppacs/jlqcd2} aimed at a full removal of the quenching effects 
144: by performing $N_f=2+1$ lattice QCD simulations 
145: with the nonperturbatively $O(a)$-improved Wilson quark action\cite{csw}
146: and the Iwasaki gauge action\cite{iwasaki} 
147: on a $(2\rm ~fm)^3$ lattice at three lattice spacings.
148: While we have succeeded in incorporating the dynamical strange quark
149: effects by the Polynomial Hybrid Monte Carlo (PHMC) algorithm\cite{phmc},  
150: the lightest up-down quark mass reached with the HMC algorithm was 
151: 64~MeV  corresponding to $m_{\pi}/m_{\rho}\approx0.6$, which required
152: a long chiral extrapolation to the physical point at $m_{\pi}/m_{\rho}\approx0.18$.
153:   
154: The PACS-CS(Parallel Array Computer System for Computational
155: Science) project\cite{ukawa1, ukawa2, kura_lat06}
156: is the successor to the CP-PACS/JLQCD project, which 
157: takes up the task that the latter has left off, namely simulation at the
158: physical point to remove the ambiguity of chiral extrapolation.  
159: It employs the same quark and gauge actions 
160: as the CP-PACS/JLQCD project,  
161: %at $\beta=1.9$ on a $32^3\times 64$ lattice 
162: but uses the PACS-CS computer with 
163: a total peak speed of 14.3 TFLOPS 
164: developed and installed at University of Tsukuba on 1 July 2006.
165: The up-down quark masses are reduced by employing
166: the domain-decomposed HMC (DDHMC) algorithm 
167: with the replay trick proposed by L\"uscher\cite{luscher, kennedy}.
168: So far, we have reached the up-down quark mass of 6~MeV
169: which yields the pion mass of about 210~MeV.
170: We also improve the simulation of the strange quark part with 
171: the UV-filtered PHMC (UV-PHMC) algorithm\cite{ishikawa_lat06}. 
172: 
173: 
174: In this report we present simulation details and preliminary results
175: which include the chiral analysis on 
176: the pseudoscalar meson masses and the decay constants with
177: chiral perturbation theory,
178: the light hadron spectrum and the $\rho$-$\pi\pi$ mixing effects.
179: Some algorithmic issues are also discussed. 
180: Selected topics on the light hadron spectrum 
181: and the ChPT analysis on the pseudoscalar meson masses 
182: and the decay constants are also reported in Refs.\cite{ukita_lat07,kadoh_lat07}.
183: 
184: \section{Simulation details}
185: 
186: 
187: \subsection{Simulation parameters}
188: 
189: \begin{table}[b!] 
190: \setlength{\tabcolsep}{10pt}
191: \renewcommand{\arraystretch}{1.2}
192: \centering
193: %\begin{tabular}{ccccccccc}% \hline
194: % $\beta=1.90$, $L^3\times T=32^3\times64$, $c_{\rm SW}=1.715$\\ %\hline
195: %\end{tabular}
196: \begin{tabular}{cccccccc} \hline
197: $\kappa_{\rm ud}$ & $\kappa_{\rm s}$ &$\tau$& $(N_0,N_1,N_2)$ & $N_{\rm poly}$ & MD time & $\tau_{\rm int}[P]$ \\ \hline \hline
198: 0.13700 & 0.13640 &0.5& (4,4,10) &180& 2000 & 38.2(17.3)\\
199: 0.13727 & 0.13640 &0.5& (4,4,14) &180& 2000 & 20.9(10.2)\\
200: 0.13754 & 0.13640 &0.5& (4,4,20) &180& 2500 & 19.2(8.6)   \\
201:         & 0.13660 &0.5& (4,4,28) &220& 900 & 10.3(2.9)   \\ 
202: 0.13770 & 0.13640 &0.25& (4,4,16) &180& 2000& 38.4(25.2)\\
203: 0.13781 & 0.13640 &0.25& (4,4,48) &180& 350 &9.1(6.1)   \\ \hline
204: \end{tabular}
205: \caption{Simulation parameters. MD time is the number of
206: trajectories multiplied by the trajectory length $\tau$.
207: $\tau_{\rm int}[P]$ denotes the integrated autocorrelation time for 
208: the plaquette.}
209: \label{tab:param}
210: \end{table} 
211: 
212: We employ the $O(a)$-improved Wilson quark action with a nonperturbative
213: improvement coefficient $c_{\rm SW}=1.715$\cite{csw} and 
214: the Iwasaki gauge action at $\beta=1.90$ on a $32^3\times64$ lattice
215: which is enlarged from $20^3\times40$ in the CP-PACS/JLQCD 
216: project to investigate the baryon masses.
217: %The lattice spacing is about $0.09$~fm. 
218: Simulation parameters are summarized in Table~\ref{tab:param}.
219: We choose six combinations of 
220: the hopping parameters $(\kappa_{\rm ud}, \kappa_{\rm s})$
221: based on the previous CP-PACS/JLQCD results.
222: Among them the heaviest combination
223: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13700,
224: 0.13640)$ in this work is 
225: the lightest one in the previous CP-PACS/JLQCD simulations, 
226: which enable us to make a 
227: direct comparison of the two results with different lattice sizes.
228: As for the strange quark,  the hopping parameter $\kappa_{\rm s}=0.13640$ 
229: corresponds to the physical point $\kappa_{\rm s}=0.136412(50)$ as estimated 
230: in the CP-PACS/JLQCD work\cite{cppacs/jlqcd1, cppacs/jlqcd2}.
231: This is the reason why all our simulations are carried out with $\kappa_{\rm s}=0.13640$, 
232: the one exception being the run at $\kappa_{\rm s}=0.13660$ and $\kappa_{\rm ud}=0.13754$
233: to investigate the strange quark mass dependence. 
234:  
235: In order to simulate the degenerate up-down quarks
236: we employ the DDHMC algorithm,
237: whose effectiveness for small quark mass region has already
238: been shown in the $N_f=2$ case\cite{luscher,del06,del07}.
239: The characteristic feature of this algorithm is a geometric separation 
240: of the up-down quark determinant into the UV and the IR parts
241: as a preconditioner of HMC, which  
242: is implemented by domain-decomposing the full lattices with small blocks.
243: We choose $8^4$ for the block size being less than (1~fm)$^4$ in
244: physical units and small enough to reside within a computing node of the PACS-CS computer.
245: There are two prominent points in the DDHMC algorithm. Firstly,
246: communication between the computing nodes is not required in calculating the UV part, 
247: which is a preferable feature for alleviating the problem of 
248: a widening gap between the processor floating point performance and 
249: the network communication bandwidth with parallel computers.
250: Secondly, we can incorporate the multiple time scale 
251: integration scheme\cite{sexton} 
252: to reduce the simulation cost efficiently.
253: The relative magnitudes of the force terms are found to be 
254: \begin{eqnarray}   
255: ||F_{\rm g}||:||F_{\rm UV}||:||F_{\rm IR}|| \approx 16:4:1,
256: \label{eq:force}
257: \end{eqnarray}
258: where we adopt the convention $||M||^2=-2{\rm tr}(M^2)$
259: for the norm of an element $M$ of the SU(3) Lie algebra.
260: $F_{\rm g}$ denotes the gauge part and $F_{\rm UV, IR}$ are for the UV
261: and the IR parts of the up-down quarks.
262: The associated step sizes for the forces are controlled by
263: three integers $N_{0,1,2}$: 
264: $\delta\tau_{\rm g}=\tau/N_0 N_1 N_2,\ \  \delta\tau_{\rm UV}=\tau/N_1
265: N_2,\ \  \delta\tau_{\rm IR}=\tau/N_2$ with $\tau$ the trajectory
266: length.
267: The integers $N_{0,1,2}$ are chosen such that 
268: \begin{eqnarray}
269:  \delta\tau_{\rm g} ||F_{\rm g}|| \approx \delta\tau_{\rm UV} ||F_{\rm UV}|| \approx \delta\tau_{\rm IR} ||F_{\rm IR}||.
270: \end{eqnarray} 
271: Taking account of the relative magnitudes of the forces in eq.(\ref{eq:force})
272: we find a larger value is allowed for $\delta\tau_{\rm IR}$ compared
273: to $\delta\tau_{\rm g}$ and $\delta\tau_{\rm UV}$,
274: which means that we need to calculate $F_{\rm IR}$ less frequently
275: in the molecular dynamics trajectories.
276: Since the calculation of  $F_{\rm IR}$ contains the quark matrix
277: inversion on the full lattice, which is the most time consuming part, 
278: this integration scheme 
279: saves the simulation cost remarkably.
280: The values for $N_{0,1,2}$ are listed in Table~\ref{tab:param}, where
281: $N_0$ and $N_1$ are fixed at 4 for all the hopping parameters, 
282: while the value of $N_2$ is adjusted taking account of acceptance rate 
283: and simulation stability. 
284: 
285: 
286: For the UV-PHMC algorithm for the strange quark,  
287: the domain-decomposition is not implemented.
288: Since we have found $||F_{\rm s}||\approx ||F_{\rm IR}||$, 
289: the step size is chosen as $\delta\tau_{\rm s}=\delta\tau_{\rm IR}$.
290: The polynomial order for UV-PHMC, which is denoted 
291: by $N_{\rm poly}$ in Table~\ref{tab:param},
292: is adjusted to yield high acceptance rate for the global Metropolis
293: test at the end of each trajectory.
294: 
295: The inversion of the Wilson-Dirac operator $D$ on the full lattice 
296: is carried out by 
297: the SAP (Schwarz alternating procedure) preconditioned GCR solver, where 
298: the preconditioning can be accelerated with the single-precision
299: arithmetic whereas the
300: GCR solver is implemented with the double precision\cite{sap+gcr}. 
301: We employ the stopping condition $|Dx-b|/|b|<10^{-9}$ for the force
302: calculation and $10^{-14}$ for the Hamiltonian, which guarantees 
303: the reversibility of the molecular dynamics trajectories to high
304: precision: $|\Delta U|<10^{-12}$ for the link variables and 
305: $|\Delta H|<10^{-8}$ for the Hamiltonian at 
306: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13781,0.13640)$.
307: 
308:  
309: \subsection{Plaquette history and autocorrelation time}
310: \begin{figure}[t!]
311: \vspace{3mm}
312: \begin{center}
313: \begin{tabular}{cc}
314: \includegraphics[width=71mm,angle=0]{figs/ukita/PLQ_his_b190kud013727ks013640b.eps}  &
315: \includegraphics[width=69mm,angle=0]{figs/ukita/PLQ_autocorr.eps} 
316: \end{tabular}
317: \end{center}
318: \vspace{-.5cm}
319: \caption{Plaquette history (left) and normalized autocorrelation 
320: function (right) for $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13727,
321: 0.13640)$.  Horizontal lines in the left denote the average value of
322: the plaquette with an error band.}
323: \label{fig:PLQ}
324: \end{figure}
325: 
326: In Fig.~\ref{fig:PLQ} we show the plaquette history and the normalized 
327: autocorrelation function $\rho(\tau)$ at
328: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13727, 0.13640)$ as a
329: representative case.
330: The integrated autocorrelation time 
331: is estimated as $\tau_{\rm int}[P]=20.9(10.2)$
332: following the definition in Ref.~\cite{luscher}  
333: \ben
334: \tau_{\rm int}(\tau)=\frac{1}{2}
335: +\sum_{0< \tau \le W} \rho(\tau),
336: \een
337: where the summation window $W$ is set to the first time lag $\tau$ that
338: $\rho(\tau)$ becomes consistent with zero within the error bar. 
339: In this case we find $W=119.5$. 
340: The choice of $W$ is not critical for estimate of 
341: $\tau_{\rm int}$ in spite of the long tail observed in Fig.~\ref{fig:PLQ}.
342: Extending the summation window, we find 
343: that $\tau_{\rm int}[P]$ saturates at $\tau_{\rm int}[P]\approx 25$ 
344: beyond $\tau=200$, which is
345: within the error bar of the original estimate.
346: For other hopping parameters 
347: we have found similar behaviors for the plaquette history and
348: the normalized autocorrelation function. 
349: We hardly observe the quark mass dependence for 
350: $\tau_{\rm int}[P]$ listed in Table~\ref{tab:param}. 
351: The statistics may not be
352: sufficiently large to derive a definite conclusion, however.
353: 
354: \section{Algorithmic issues}
355: 
356: 
357: \subsection{Efficiency of DDHMC algorithm}
358: \label{subsec:efficiency}
359: 
360: \begin{figure}[t!]
361: \vspace{3mm}
362: \begin{center}
363: \begin{tabular}{cc}
364: \includegraphics[width=75mm,angle=0]{figs/kura/cost_ddhmc.eps}
365: \end{tabular}
366: \end{center}
367: \vspace{-.5cm}
368: \caption{Cost estimate of $N_f=2+1$ QCD simulations 
369: with the HMC (solid line) and the
370: DDHMC (red circle) algorithms at $a=0.1$~fm with $L=3$~fm 
371: for 100 independent configurations. Vertical dotted 
372: line denotes the physical point.}
373: \label{fig:berlinwall}
374: \end{figure}
375: 
376: 
377: 
378: In order to discuss the efficiency of the DDHMC algorithm,
379: it is instructive to compare with that of the HMC algorithm.
380: We first recall an empirical cost formula for
381: $N_f=2$ QCD simulations with the HMC algorithm 
382: based on the CP-PACS results\cite{berlinwall}:
383: \ben
384: {\rm cost[Tflops\cdot years]}&=&C\left[\frac{\#{\rm conf}}{1000}\right]\cdot
385: \left[\frac{0.6}{m_\pi/m_\rho}\right]^6\cdot
386: \left[\frac{L}{3{\rm ~fm}}\right]^5\cdot
387: \left[\frac{0.1{\rm ~fm}}{a}\right]^7\nn
388: \label{eq:cost_hmc}
389: \een
390: with $C\approx 2.8$.
391: A strong quark mass dependence is found in the above formula:
392: $1/(m_\pi/m_\rho)^6$ behaves as $1/m_{\rm ud}^3$ 
393: in the leading term for the small quark mass region. 
394: This quark mass dependence is owing to the following three factors:
395: The number of the quark matrix inversion is governed by the
396: condition number which should be proportional to $1/m_{\rm ud}$;
397: to keep the acceptance rate fixed we should take 
398: $\delta\tau\propto m_{\rm ud}$ for the step size in 
399: the molecular dynamics trajectories;
400: The autocorrelation time of the HMC evolution shows 
401: $1/m_{\rm ud}$ dependences in the CP-PACS run\cite{cppacs_nf2}.
402: 
403: To estimate the computational cost for
404: $N_f=2+1$ QCD simulations with the HMC algorithm,
405: we assume that the strange quark contribution is given by half of
406: eq.(\ref{eq:cost_hmc}) at $m_\pi/m_\rho=0.67$ which
407: is a phenomenologically estimated ratio of the
408: strange pseudoscalar meson ``$m_{\eta_{\rm s}}$'' and $m_\phi$:
409: \ben
410: \frac{m_{\eta_{\rm
411: s}}}{m_\phi}=\frac{\sqrt{2m_K^2-m_\pi^2}}{m_\phi}\approx 0.67.
412: \een
413: Since the strange quark is relatively heavy, its computational
414: cost occupies only a small fraction as the up-down quark masses decrease. 
415: In Fig.~\ref{fig:berlinwall} we draw the cost formula 
416: for the $N_f=2+1$ case as a function of $m_\pi/m_\rho$,
417: where we take \#conf=100, a=0.1~fm and $L=3$~fm in eq.(\ref{eq:cost_hmc})
418: as a representative case.
419: We observe a steep increase of the computational cost below 
420: $m_\pi/m_\rho\simeq 0.5$.
421: At the physical point the expected cost is $O(100)$ Tflops$\cdot$years.
422: 
423: Now let us turn to the case of the DDHMC algorithm.
424: The red symbol denotes the measured cost at 
425: $\kappa_{\rm ud}$=0.13781, 0.13770 with $\kappa_{\rm s}=0.13640$,
426: which are the lightest two points in our simulation.
427: The DDHMC algorithm show a remarkable improvement
428: reducing the cost by $30-50$ times in magnitude. 
429: The majority of this reduction arises from the multiple time scale
430: integration scheme and the GCR solver
431: accelerated by the SAP preconditioning
432: with the single-precision arithmetic.
433: Roughly speaking, the improvement factor is $O(10)$ for the former
434: and $3-4$ for the latter.
435: Note that the quark mass dependence is also tamed: 
436: Since we find that $\tau_{\rm int}[P]$ is independent 
437: of the quark masses, 
438: the cost is proportional to $1/m_{\rm ud}^2$.
439: Our results show a feasibility of simulations at the physical point
440: with the $O(10)$ Tflops computer which is available at present.
441: 
442: 
443: \subsection{$\tau$ dependence of DDHMC algorithm}
444: 
445: \begin{table}[t!]
446: \centering
447: \begin{tabular}{ccccccc}  \hline
448:  $\tau$  & $(N_0,N_1,N_2)$  &  $N_{\rm poly}$ & trajs. & MD time & 
449: $\tau_{\rm int}[P]$ & $\tau_{\rm int}[{\rm \#mult}]$\\ \hline
450: 0.5   & (4,4,6) & 130 & 6000  & 3000 & 23.7(9.2) & 92(56) \\ 
451: 0.5/3 & (4,4,2) & 130 & 18000 & 3000 & 18.6(5.9) & 42(21) \\ \hline
452: \end{tabular}
453: \caption{Parameters for $\tau$-dependence study. \#mult denotes
454: the number of multiplications
455: of the Wilson-Dirac quark matrix on the full lattice.}
456: \label{tab:tau_dep}
457: \end{table}
458: 
459: 
460: \begin{figure}[b!]
461: \vspace{3mm}
462: \begin{center}
463: \begin{tabular}{cc}
464: \includegraphics[width=65mm,angle=0]{figs/kura/plaq_ac_tau.eps} &
465: \includegraphics[width=65mm,angle=0]{figs/kura/mult_ac_tau.eps}
466: \end{tabular}
467: \end{center}
468: \vspace{-.5cm}
469: \caption{Normalized autocorrelation functions for the plaquette (left) 
470: and the \#mult (right) at $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13700,0.13640)$.
471: Black (red) symbols denote the $\tau=0.5$ ($\tau=0.5/3$) case.}
472: \label{fig:tau_dep}
473: \end{figure}
474: 
475: In the DDHMC algorithm a subset of of all link variables,
476: which are referred to as the active link variables, are 
477: updated during the molecular dynamics evolution, while keeping
478: other field variables fixed\cite{luscher}. 
479: The fraction of the active link variables 
480: depends on the block size we choose. In our case of $8^4$ it is only
481: 37\%. To ensure that all the link variables on the lattice should be 
482: updated equally on average we implement random gauge field 
483: translations at the end of every molecular dynamics trajectories
484: following Ref.~\cite{luscher}.
485: 
486: Our concern is that the DDHMC algorithm might have 
487: a long autocorrelation time due to the existence of fixed link variables 
488: during the molecular dynamics evolution.
489: A possible way to reduce the effects of the fixed link variables
490: is more frequent random gauge field translations.
491: This is easily realized by making $\tau$ shorter with
492: $\delta\tau$ fixed.
493: We investigate the $\tau$ dependence of the DDHMC algorithm
494: employing a smaller lattice size of $16^3\times 32$ at
495: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13700,0.13640)$.
496: Other parameters are summarized in Table~\ref{tab:tau_dep}.
497:  
498: In Fig.~\ref{fig:tau_dep} we show the normalized autocorrelation
499: functions for the plaquette and the number of multiplications
500: of the Wilson-Dirac quark matrix on the full lattice
501: during a molecular dynamics trajectory. 
502: Black symbols denote the $\tau=0.5$ case, while red ones are for
503: the $\tau=0.5/3$ case.
504: We observe that the normalized autocorrelation functions 
505: for $\tau=0.5$ have longer tails than $\tau=0.5/3$
506: before becoming consistent with zero:
507: This is quantitatively checked by the integrated autocorrelation times
508: $\tau_{\rm int}[P]$ and $\tau_{\rm int}[{\rm \#mult}]$
509: in Table~\ref{tab:tau_dep}: The $\tau=0.5/3$ case has
510: shorter autocorrelation times.
511: Since the computational cost is the same for
512: the $\tau=0.5$ and the $\tau=0.5/3$ cases in terms of MD time, 
513: we can conclude that 
514: the $\tau=0.5/3$ case shows a better efficiency than 
515: the $\tau=0.5$ case. 
516: Based on this study, albeit conducted at a relatively heavy quark mass
517: $(m_\pi/m_\rho\approx 0.6)$, we employ $\tau=0.25$
518: for the production run at $(\kappa_{\rm ud},\kappa_{\rm s})
519: =(0.13781,0.13640)$ and $(0.13770,0.13640)$, which is half
520: of the trajectory length at other hopping parameters.
521: 
522: \subsection{Simulation stability}
523: 
524: In Refs.~\cite{del06,del07} simulation stability was discussed
525: based on the spectral gap distribution of the Wilson-Dirac operator
526: for two-flavor lattice QCD simulations.
527: The spectral gap is defined as
528: \ben
529: \mu={\rm min}\{|\lambda|\; |\;
530: \lambda\;\mbox{is an eigenvalue of}\;Q\},
531: \een
532: where $Q$ is the hermitian Wilson-Dirac operator
533: $Q=\gamma_5 D_W$ with $D_W=(1/2)\{\gamma_\mu(\nabla_\mu^*+\nabla_\mu)
534: -a\nabla_\mu^*\nabla_\mu\}+m_0$.
535: Important indices to characterize the distribution are
536: its median ${\bar \mu}$ and width $\sigma$. 
537: The latter is defined as
538: $(v-u)/2$, where $[u,v]$ is the smallest range of $\mu$ which contains
539: more than 68.3\% of the data. This is to avoid potentially large
540: statistical uncertainties which might occur when data are not sufficiently
541: sampled.
542: Their chief findings are two points: The first one is that
543: the median ${\bar \mu}$ shows a good 
544: linear dependence on the current up-down quark mass $m_{\rm ud}^{\rm AWI}$
545: and the magnitude of the slope is well described by $Z_A$ empirically.
546: The second one is that the width $\sigma$ scales as
547: \ben
548: \sigma\frac{\sqrt{V}}{a}\simeq 1
549: \een
550: with $V$ the four-dimensional volume in physical units. 
551: They also observe that the width $\sigma$ is roughly independent of the
552: quark mass for the unimproved Wilson quark action, while it shows 
553: a trend to decrease with the mass for the improved one.
554: 
555: This study was also applied to the $N_f=2+1$ case 
556: in Ref.~\cite{kura_lat06}
557: where we reported on our preliminary run on a $16^3\times 32$ lattice 
558: preparing for the PACS-CS project.  
559: We observed ${\bar \mu}\propto m_{\rm ud}^{\rm AWI}$ and 
560: found $0.5\simlt \sigma({\sqrt{V}}/{a})\simlt 0.76$ for
561: $15{\rm ~MeV}<m_{\rm ud}^{\rm AWI}<64{\rm ~MeV}$, where
562: $\sigma$ diminishes as the up-down quark mass decreases.
563: 
564: The existence of a gap in the spectrum of the Wilson-Dirac operator
565: allows us to simulate the light quarks efficiently.
566: The authors in Ref.~\cite{del06} propose a stability condition requiring
567: ${\bar \mu}\ge 3\sigma$ to assure the existence of the gap. 
568: Let us apply this condition to our case.
569: Assuming $\sigma({\sqrt{V}}/{a})= 1$ we estimate
570: $\sigma=2.26$~MeV using $a=0.09$~fm and $V=(2.8{\rm ~fm})^4$ which
571: will be obtained later.
572: By using the empirical relation 
573: ${\bar \mu}\simeq Z_A m_{\rm ud}^{\rm AWI}$ 
574: we find $Z_A m_{\rm ud}^{\rm AWI}\simgt 6.8$~MeV for the
575: stability condition, which is heavier than the physical point.  
576: On the other hand, we found $\sigma({\sqrt{V}}/{a})< 1$ in 
577: Ref.~\cite{kura_lat06}, indicating that the actual bound will be lower.  
578: Our runs toward the physical point should shed light on the actual bound 
579: of stability for our lattice parameters.
580:  
581: 
582: \subsection{Comparison of DDHMC and mass-preconditioned HMC}
583: 
584: \begin{table}[t!]
585: \centering
586: \begin{tabular}{ccccccc}  \hline
587: prec. & block size  & $\rho$ & $(N_0,N_1,N_2)$  & $\tau$  & MD time & $P_{\rm acc}$ \\ \hline
588: DD   & $8^4$  & $-$  & (4,8,12) & 1 & 3000 & 0.857(8) \\ 
589: mass & $-$    & 0.09 & (4,8,12) & 1 & 3000 & 0.794(8) \\ \hline
590: \end{tabular}
591: \caption{Simulation parameters for the DDHMC 
592: and the mass-preconditioned HMC algorithms. $P_{\rm acc}$ denote the
593: acceptance rate.}
594: \label{tab:dd-mass}
595: \end{table}
596: 
597: 
598: \begin{figure}[b!]
599: \vspace{3mm}
600: \begin{center}
601: \begin{tabular}{cc}
602: \includegraphics[width=65mm,angle=0]{figs/kura/plaq_ac_prec.eps} &
603: \includegraphics[width=65mm,angle=0]{figs/kura/mult_ac_prec.eps}
604: \end{tabular}
605: \end{center}
606: \vspace{-.5cm}
607: \caption{Normalized autocorrelation functions for the plaquette (left) 
608: and the number of multiplications of the Wilson-Dirac quark matrix on
609: the full lattice (right).
610: Black (red) symbols denote the DDHMC (mass-preconditioned HMC) case.}
611: \label{fig:dd-mass_autoc}
612: \end{figure}
613: 
614: As discussed in Sec.~\ref{subsec:efficiency}, 
615: it is essential for the efficiency of the DDHMC algorithm to incorporate
616: the multiple time scale integration scheme.  
617: It is well known that this scheme is also 
618: applicable to the mass-preconditioned HMC algorithm\cite{massprec1,massprec2}.
619: We have made a direct comparison of the two algorithms
620: in $N_f=2$ QCD on a $16^3\times 32$ lattice 
621: employing the $O(a)$-improved
622: Wilson quark action with the nonperturbative improvement coefficient
623: $c_{\rm SW}=2.0171$\cite{csw_nf2} 
624: and the plaquette gauge action at $\beta=5.2$.
625: The lattice spacing is 0.1~fm and the physical pseudoscalar meson mass
626: is about 600~MeV at $\kappa_{\rm ud}=0.1355$.
627: For the mass-preconditioned HMC algorithm we employ two set of 
628: the pseudofermion fields which decompose the fermion determinant as
629: \ben
630: \det Q^2=\det (W^\dagger W)\det\left(\frac{Q^2}{W^\dagger W}\right),
631: \een
632: where $Q$ is the hermitian Wilson-Dirac operator 
633: and the preconditioning operator is
634: given by $W=Q+\rho$.
635: For convenience we refer to $\det (W^\dagger W)$ as the UV part
636: and the $\det({Q^2}/(W^\dagger W))$ as the IR part in an analogy with
637: the DDHMC algorithm.
638: The step sizes are chosen with the
639: three integers $N_{0,1,2}$ in exactly the same way as the DDHMC algorithm.
640: Simulation parameters are summarized in Table~\ref{tab:dd-mass}.
641: The block size for the DDHMC algorithm and the $\rho$ parameter for the
642: mass-preconditioned HMC algorithm are chosen such that $||F_{0,1,2}||$
643: are roughly the same between these two algorithms.
644: This condition yields comparable acceptance ratios with $N_{0,1,2}$ in common. 
645: We employ the BiCGStab algorithm for the quark matrix inversion 
646: in both the UV and the IR parts. 
647: 
648: 
649: \begin{table}[t!]
650: \centering
651: \begin{tabular}{cccccc}  \hline
652: prec. & $\tau_{\rm int}[P]$  & $\tau_{\rm int}[{\rm \#mult}]$  & 
653: \#mult  & cost[$P$]  & cost[\#mult]  \\ \hline
654: DD   & 27(10) & 22(7)  & 45530(280) & 1.2(5) & 1.0(3) \\ 
655: mass & 28(12) & 35(16) & 67160(380) & 1.9(8) & 2.4(1.0) \\ \hline
656: \end{tabular}
657: \caption{Integrated autocorrelation time and cost estimate 
658: for the DDHMC and the mass-preconditioned HMC algorithms.}
659: \label{tab:dd-mass_cost}
660: \end{table}
661: 
662: 
663: \begin{figure}[b!]
664: \vspace{3mm}
665: \begin{center}
666: \begin{tabular}{cc}
667: \includegraphics[width=65mm,angle=0]{figs/kura/mult_dd.eps} &
668: \includegraphics[width=65mm,angle=0]{figs/kura/mult_mass.eps}
669: \end{tabular}
670: \end{center}
671: \vspace{-.5cm}
672: \caption{History  
673: for number of multiplications of the Wilson-Dirac quark matrix 
674: on the full lattice for the DDHMC algorithm (left) 
675: and the mass-preconditioned HMC algorithm (right).}
676: \label{fig:dd-mass_hist}
677: \end{figure}
678: 
679: In Fig.~\ref{fig:dd-mass_autoc} we plot the normalized autocorrelation 
680: function for the plaquette as a function of MD time. 
681: The results of both algorithms show
682: quite similar behaviors and the integrated autocorrelation time  
683: $\tau_{\rm int}[P]$ in Table~\ref{tab:dd-mass_cost} 
684: are consistent within the errors.
685: Figure~\ref{fig:dd-mass_hist} shows the MD-time history of the 
686: number of multiplications of the Wilson-Dirac quark matrix 
687: on the full lattice.
688: The total number of multiplications is the sum of those required to
689: calculate the UV and the IR forces and the Hamiltonian.
690: Their contributions are denoted by black, red, green, blue lines in
691: order. 
692: Comparing the results of the DDHMC and the mass-preconditioned HMC,
693: we observe a clear difference in the UV part contribution:
694: the mass-preconditioned HMC needs more than twice of the
695: multiplication number for the DDHMC.  
696: This ends up in a 50\% difference in the total number of multiplications.   
697: In Fig.~\ref{fig:dd-mass_autoc} we also plot the normalized autocorrelation 
698: function for the total number of multiplications.
699: Although the DDHMC result seems to show a slightly steeper fall-off,
700: both results are consistent within the error bars.
701: This is confirmed by the integrated autocorrelation time 
702: $\tau_{\rm int}[{\rm \#mult}]$ in  Table~\ref{tab:dd-mass_cost}. 
703:   
704: Now let us compare the efficiencies of both algorithms.
705: We define the machine-independent cost formula by
706: \ben
707: {\rm cost}[{\cal O}]=\#{\rm mult(total)/MD\;\;time}\times 
708: \tau_{\rm int}[{\cal O}]/10^6, 
709: \een
710: where the observable $\cal O$ is the plaquette or the total number of
711: multiplications.
712: In Table~\ref{tab:dd-mass_cost} we summarize the results of
713: cost[${\cal O}$]. For both observables the DDHMC algorithm shows
714: better efficiency than the mass-preconditioned HMC algorithm
715: albeit the errors are rather large.
716: 
717: There remains a couple of concerns in this study.
718: The first one is the quark mass dependence,
719: because our results are obtained at only one hopping parameter.
720: The second one is the optimization.
721: While we choose $8^4$ block size for the DDHMC algorithm
722: and $\rho=0.09$ for the mass-preconditioned HMC algorithm since 
723: $||F_{0,1,2}||$ are roughly the same, these parameters may not be the optimal
724: values for each of the algorithms. We leave these issues to future studies.
725: 
726: 
727: 
728: \section{Physical results}
729: 
730: \subsection{Measurement of hadron masses, quark masses, decay constants}
731: 
732: 
733: \begin{figure}[t!]
734: \vspace{3mm}
735: \begin{center}
736: \begin{tabular}{cc}
737: \includegraphics[width=65mm,angle=0]{figs/ukita/msn_13727.eps} &
738: \includegraphics[width=65mm,angle=0]{figs/ukita/msn_13781.eps} 
739: \vspace*{5mm}
740: \end{tabular}
741: \begin{tabular}{cc}
742: \includegraphics[width=65mm,angle=0]{figs/ukita/brn_13727.eps} &
743: \includegraphics[width=65mm,angle=0]{figs/ukita/brn_13781.eps}
744: \end{tabular}
745: \end{center}
746: \vspace{-.5cm}
747: \caption{Effective masses for the mesons (top) and the baryons
748:  (bottom) at $\kappa_{\rm ud}=0.13727$ (left) and 0.13781 (right).
749: Horizontal lines represent the fitting results with an error band.}
750: \label{fig:Meff}
751: \end{figure}
752: 
753: 
754: 
755: \begin{figure}[t!]
756: \vspace{3mm}
757: \begin{center}
758: \begin{tabular}{cc}
759: \includegraphics[width=65mm,angle=0]{figs/ukita/binsize_meson_13727.eps} &
760: \includegraphics[width=65mm,angle=0]{figs/ukita/binsize_baryon_13727.eps}
761: \end{tabular}
762: \end{center}
763: \vspace{-.5cm}
764: \caption{Binsize dependence of magnitude of error for mesons (left) and 
765: baryons (right) at $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13727,0.13640)$.}
766: \label{fig:binsize}
767: \end{figure}
768: 
769: We measure both the meson and the baryon correlators at every 10 trajectories
770: at the unitary points where 
771: the valence quark masses are equal to the sea quark masses.
772: Light hadron masses are extracted from a single exponential 
773: $\chi^2$ fit to the correlators
774: with an exponentially smeared source and a local sink. 
775: Figure~\ref{fig:Meff} shows the hadron effective masses  
776: at $\kappa_{\rm ud}=0.13727$ and 
777: $0.13781$ as representative cases. 
778: We observe clear plateau for the mesons except for the $\rho$
779: meson and also good signal for the baryons thanks to a large volume.
780: Especially, the $\Omega$ baryon has a stable signal, which we use
781: as a physical input to determine the cutoff scale later.   
782: The horizontal lines denote the fitting results with an error band
783: of one standard deviation. Their widths represent the fitting ranges.
784: Statistical errors are estimated by the jackknife method.
785: In Fig.~\ref{fig:binsize} we plot binsize dependence of magnitude
786: of error for the mesons and the baryons at 
787: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13727,0.13640)$.
788: For the pseudoscalar mesons we observe that the
789: magnitude of error gradually increases as the bin size is enlarged up
790: to about 40 MD time, beyond which it stabilizes. 
791: For other hadrons we do not find any clear binsize dependence.
792: The data at other hopping parameters show similar behaviors.
793: Based on this observation
794: we choose 50 molecular dynamics time for the jackknife analysis
795: at all the hopping parameters.  
796: 
797: We extract the bare quark mass through the axial vector
798: Ward-Takahashi identity (AWI)  by 
799: \begin{eqnarray}
800: am^{\rm AWI}_q =\lim_{t\rightarrow \infty} \frac{\langle\nabla_4
801:   A_4^{\rm imp}(t) P(0)\rangle}{2\langle P(t)P(0)\rangle}
802: \end{eqnarray}
803: with $P$ the pseudoscalar density and $A_4^{\rm imp}$ 
804: the nonperturbatively $O(a)$-improved axial vector current\cite{ca}.
805: The renormalized quark mass and the
806: pseudoscalar meson decay constant in the continuum ${\overline{\rm MS}}$
807: scheme are defined as follows:
808: \begin{eqnarray}
809: m^{\overline{\rm MS}}_q&=&\frac{Z_A \left(1+b_A
810:   \frac{m^{\rm AWI}}{u_0}\right)}
811: {Z_P \left(1+b_P \frac{m^{\rm AWI}}{u_0} \right)}m^{\rm AWI}_q,\\
812: f_{\rm PS}&=&2\kappa u_0 Z_A \left(1+b_A\frac{m^{\rm AWI}_q}{u_0}\right) \frac{C_A^s}{C_P^s}\sqrt{\frac{2C_P^l}{m_{PS}}}.
813: \end{eqnarray}
814: Here $C_{A,P}^s$ are the amplitudes extracted from the correlators
815: $\langle A_4^{\rm imp}(t) P(0)\rangle$ and $\langle P(t)P(0)\rangle$
816: with an exponentially smeared source and a local sink, while
817: $C_P^l$ is from $\langle P(t)P(0)\rangle$ with a local source and
818: a local sink. The renormalization factors
819: $Z_{A,P}$ and the improvement coefficients $b_{A,P}$ are evaluated
820: perturbatively up to 
821: one-loop level\cite{z_pt,z_imp_pt}with the
822: tadpole improvement. 
823: 
824: 
825: \subsection{Comparison with the previous CP-PACS/JLQCD results}
826: 
827: 
828: \begin{figure}[t!]
829: \vspace{3mm}
830: \begin{center}
831: \begin{tabular}{cc}
832: \includegraphics[width=65mm,angle=0]{figs/kura/e.msn_k22_s11_pi.eps} &
833: \includegraphics[width=65mm,angle=0]{figs/kura/e.msn_k22_s11_rho.eps}
834: \end{tabular}
835: \end{center}
836: \vspace{-.5cm}
837: \caption{Effective masses for the $\pi$ (left) and the $\rho$ (right) 
838: at $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13700,0.13640)$.
839: Black and red symbols denote the PACS-CS and the CP-PACS/JLQCD results, 
840: respectively.}
841: \label{fig:comp_eff}
842: \end{figure}
843: 
844: \begin{table}[h!]
845: \centering
846: \begin{tabular}{ccccc}  \hline
847:                                  & lattice size           &  $am_{\pi}$ & $am_{\rho}$ & $am_{\rm N}$\\ \hline
848: PACS-CS     &$32^3\times 64$& 0.3220(6) & 0.506(2) & 0.726(3) \\
849: $[t_{\rm min},t_{\rm max}]$ & & [13,30] & [10,20] & [10,20] \\ 
850: CP-PACS/JLQCD&$20^3\times 40$ &0.3218(8)& 0.516(3) & 0.733(4)\\ 
851: $[t_{\rm min},t_{\rm max}]$ & & [8,20] & [9,15] & [11,17] \\ \hline
852: \end{tabular}
853: \caption{PACS-CS and CP-PACS/JLQCD results for 
854: $\pi$, $\rho$ and nucleon masses at 
855: $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13700, 0.13640)$.
856: [$t_{\rm min}$,$t_{\rm max}$] denotes the fitting range.}
857: \label{tab:comp}
858: \end{table}
859: 
860: 
861: \begin{figure}[b!]
862: \vspace{3mm}
863: \begin{center}
864: \begin{tabular}{cc}
865: \includegraphics[width=65mm,angle=0]{figs/ukita/kud.PS_LL2.eps}
866: \hspace*{2mm} &
867: \includegraphics[width=65mm,angle=0]{figs/ukita/kud.V_LL.eps}
868: \end{tabular}
869: \end{center}
870: \vspace{-.5cm}
871: \caption{$(am_{\pi})^2$ (left) and $am_{\rho}$ (right) 
872: as a function of $1/\kappa_{\rm ud}$. 
873: Red and black symbols denote the PACS-CS and the CP-PACS/JLQCD results, 
874: respectively.}
875: \label{fig:comp}
876: \end{figure}
877: 
878: We first compare the PACS-CS results on $32^3\times 64$ 
879: with the previous CP-PACS/JLQCD results 
880: on $20^3\times 40$\cite{cppacs/jlqcd1, cppacs/jlqcd2} 
881: at $(\kappa_{\rm ud}, \kappa_{\rm s})=(0.13700, 0.13640)$.
882: In Fig.~\ref{fig:comp_eff} we plot the effective masses for the $\pi$
883: and the $\rho$ mesons. The PACS-CS and the CP-PACS/JLQCD
884: results are consistent for the $\pi$ meson, while a slight deviation
885: is observed for the $\rho$ meson. 
886: This is numerically confirmed by
887: the fitting results listed in Table~\ref{tab:comp}, where 
888: we employ a single exponential $\chi^2$ fit. 
889: The nucleon mass is also given in Table~\ref{tab:comp}.
890: We find 1$-$2\% deviation 
891: for the $\rho$ meson and nucleon masses, which
892: could be due to possible finite size effects. 
893: 
894: Figure~\ref{fig:comp} shows the up-down quark mass dependence of
895: $(am_{\pi})^2$ and $am_{\rho}$ with $\kappa_{\rm s}$ fixed at 0.13640.
896: For the pion mass we observe that the PACS-CS and 
897: the CP-PACS/JLQCD results are smoothly connected 
898: as a function of $1/\kappa_{\rm ud}$.
899: On the other hand, the quark mass dependence is not so smooth  
900: for the $\rho$ meson.
901: Although this may be attributed to finite size effects,
902: further studies are needed in the $\rho$ channel. 
903: 
904: 
905: \subsection{Chiral analysis on pseudoscalar meson masses and decay constants}
906: 
907: We examine the chiral behaviors of the pseudoscalar meson masses
908: and decay constants in comparison with the prediction of chiral
909: perturbation theory (ChPT). Our interest exist in the following 
910: points: (i) signals for chiral logarithms, 
911: (ii) determination of low energy
912: constants in the chiral lagrangian, (iii) determination of the
913: physical point with the ChPT fit, (iv) estimate of the magnitude 
914: of finite size effects based on one-loop calculations of ChPT.
915: 
916: 
917: We first recall the one-loop expressions of ChPT 
918: for the pseudoscalar meson masses 
919: and the decay constants\cite{gasser}\footnote{
920: $f_\pi$ is normalized as 92.4~MeV in these expressions, while
921: our results are presented in the $f_\pi=130.7$~MeV normalization.}:
922: \begin{eqnarray}
923: \qquad 
924: m_{\pi}^2&=& 2 \hat m B_0 \left\{
925: 1+\mu_\pi-\frac{1}{3}\mu_\eta 
926: +\frac{B_0}{f_0^2} \left(
927: 16 \hat m  (2L_{8}-L_5) 
928: +16 (2 \hat m +m_{\rm s}) (2L_{6}-L_4)  
929: \right) \right\}, 
930: \label{eq:chpt_mpi}
931: \\
932: m_K^2&=&(\hat m +m_{\rm s})B_0 \left\{
933: 1+\frac{2}{3}\mu_\eta
934: +\frac{B_0}{f_0^2}\left(
935: 8(\hat m+m_{\rm s}) (2L_{8}-L_5)
936: +16(2 \hat m  +m_{\rm s}) (2L_{6}-L_4) 
937:  \right)\right\}, 
938: \label{eq:chpt_mk}
939: \\
940: f_\pi &=&f_0\left\{
941: 1-2\mu_\pi-\mu_K
942: + \frac{B_0}{f_0^2} \left (
943: 8 \hat m L_5+8(2 \hat m +m_{\rm s})L_4
944: \right)\right\}, 
945: \label{eq:chpt_fpi}
946: \\
947: f_K &=&f_0\left\{
948: 1-\frac{3}{4}\mu_\pi-\frac{3}{2}\mu_K-\frac{3}{4}\mu_\eta
949: +\frac{B_0}{f_0^2}\left(4(\hat m+m_{\rm s}) L_5+8(2 \hat m+ m_{\rm s})L_4 
950: \right)\right\}, 
951: \label{eq:chpt_fk}
952: \end{eqnarray}
953: where ${\hat m}=(m_{\rm u}+m_{\rm s})/2$ and 
954: $L_{4,5,6,8}$ are the low energy constants, and 
955: $\mu_{\rm PS}$ is the chiral logarithm defined by 
956: \begin{eqnarray}
957: \mu_{\rm PS}=\frac{1}{32\pi^2}\frac{m_{\rm PS}^2}{f_0^2}
958: \ln\left(\frac{m_{\rm PS}^2}{\mu^2}\right)
959: \label{eq:chlog}
960: \end{eqnarray}
961: with $\mu$ the renormalization scale.
962: There are six unknown low energy constants $B_0,f_0,L_{4,5,6,8}$ 
963: in the expressions above.
964: The low energy constants are scale-dependent so as to 
965: cancel that of the chiral logarithm (\ref{eq:chlog}). 
966: We determine these parameters                                  
967: by making a simultaneous fit for $m_\pi^2$, $m_K^2$, $f_\pi$ and $f_K$.
968: 
969: 
970: We also consider the contributions of the finite size effects based on ChPT.
971: At the one-loop level the finite size effects defined by 
972: $R_X=(X(L)-X(\infty))/X(\infty)$ for $X=m_\pi,m_K,f_\pi,f_K$ are given 
973: by~\cite{colangelo05}:
974: \begin{eqnarray}
975: && \ \ \ R_{m_\pi}%=\frac{M_{\pi}(L)-M_{\pi}}{M_{\pi}}
976: =\frac{1}{4}\xi_{\pi}\tilde g_1(\lambda_\pi)-\frac{1}{12}\xi_{\eta}\tilde g_1(\lambda_\eta),\\
977: && \ \ \ R_{m_K} %=\frac{M_K(L)-M_K}{M_K}
978: =\frac{1}{6}\xi_{\eta}\tilde g_1(\lambda_\eta),\\
979: && \ \ \ R_{f_\pi}%=\frac{F_{\pi}(L)-F_{\pi}}{F_{\pi}}
980: =-\xi_{\pi}\tilde g_1(\lambda_\pi)-\frac{1}{2}\xi_{K}\tilde g_1(\lambda_K),\\
981: && \ \ \ R_{f_K} %=\frac{F_K(L)-F_K}{F_K}
982: =-\frac{3}{8}\xi_{\pi}\tilde g_1(\lambda_\pi)-\frac{3}{4}\xi_{K}\tilde g_1(\lambda_K)
983: -\frac{3}{8}\xi_{\eta}\tilde g_1(\lambda_\eta)
984: \end{eqnarray}
985: with
986: \begin{eqnarray}
987: \hspace{2.5cm} && \xi_{\rm PS} \equiv \frac{m_{\rm PS}^2}{(4\pi f_\pi)^2},
988: \quad \ \lambda_{\rm PS} \equiv m_{\rm PS} L, \quad
989: \tilde g_1(x)=\sum_{n=1}^{\infty}\frac{4m(n)}{{\sqrt n }x} K_1({\sqrt n} x), 
990: \end{eqnarray}
991: where $K_1$ is the Bessel function of the second kind and $m(n)$
992: denotes the multiplicities in the expression of $n=n_x^2+n_y^2+n_z^2$. 
993: With the use of these formulae
994: we estimate the possible finite size effects in our results.
995: 
996: Before presenting our fitting results, it is instructive to
997: compare the PACS-CS and the CP-PACS/JLQCD results for 
998: $(a m_\pi)^2/(am_{\rm ud}^{\rm AWI})$ and $f_K/f_\pi$.
999: In Fig.~\ref{fig:comparison} we plot them 
1000: as a function of $am_{\rm ud}^{\rm AWI}$ with
1001: $\kappa_{\rm s}$ fixed at 0.13640. The PACS-CS and the CP-PACS/JLQCD
1002: results are denoted by the red and the black symbols, respectively.
1003: The two sets of data together show a smooth behavior as 
1004: a function of $am_{\rm ud}^{\rm AWI}$, and at $\kappa_{\rm ud}=0.13700$  
1005: ($am_{\rm ud}^{\rm AWI}=0.028$) they show good consistency.
1006: It is important to observe that an almost linear quark mass dependence of the 
1007: CP-PACS/JLQCD results for heavier up-down quark masses changes into a convex behavior,  
1008: both for $(a m_\pi)^2/(am_{\rm ud}^{\rm AWI})$ and $f_K/f_\pi$, 
1009: as the quark mass is lowered in the PACS-CS runs. 
1010: This is a characteristic feature expected from 
1011: the ChPT prediction in the small quark mass region
1012: due to the chiral logarithm.
1013: This curvature drives up the ratio $f_K/f_\pi$ toward the 
1014: experimental value as the physical point is approached.
1015: 
1016: 
1017:   
1018: 
1019: \begin{figure}[t!]
1020: \begin{center}
1021: \hspace{-0.6cm}
1022: \includegraphics[width=7cm,keepaspectratio,clip]{figs/kadoh/f.mpi2bmud.cp.eps}
1023: \includegraphics[width=7cm,keepaspectratio,clip]{figs/kadoh/f.fKfpi.cp.eps}
1024: \caption{Comparison of the PACS-CS (red) and the CP-PACS/JLQCD (black)
1025:   results for $(am_\pi)^2/(am_{\rm ud}^{\rm AWI})$ (left) and
1026:   $f_K/f_\pi$ (right) as a function of $am_{\rm ud}^{\rm AWI}$. 
1027: $\kappa_{\rm s}$ is fixed at 0.13640.
1028: Vertical lines denote the physical point and star symbol represents
1029: the experimental value.}
1030: \label{fig:comparison}
1031: \end{center}
1032: \end{figure}
1033: 
1034: 
1035: \begin{table}[b!]
1036: \begin{center}
1037: \hspace{1cm}
1038: \begin{tabular}{ccccc}\hline
1039: $\kappa_{\rm ud}$ & $\kappa_{\rm s}$  & $am_{\pi}$  & $am_{\rm ud}^{\rm AWI}$ & 
1040: $am_{\rm s}^{\rm AWI}$ \\
1041: \hline
1042: \hline
1043: 0.13700  &  0.13640  &  0.32196(62)  &  0.02800(20) & 0.04295(30)  \\
1044: 0.13727  &  0.13640  &  0.26190(66)  &  0.01895(13) & 0.04061(18)  \\
1045: 0.13754  &  0.13640  &  0.18998(56)  &  0.01020(11) & 0.03876(18)  \\ 
1046:          &  0.13660  &  0.17934(78)  &  0.00908(7)  & 0.03257(17)  \\
1047: 0.13770  &  0.13640  &  0.13591(88)  &  0.00521(9)  & 0.03767(10)  \\
1048: 0.13781  &  0.13640  &  0.08989(291) &  0.00227(16) & 0.03716(20)  \\
1049: \hline
1050: \end{tabular}
1051: \caption{Pion masses and unrenormalized AWI quark masses.}
1052: \label{tab:quarkmass}
1053: \end{center}
1054: \end{table}
1055:   
1056: 
1057: 
1058:   
1059: 
1060: Let us apply the ChPT formulae (\ref{eq:chpt_mpi})$-$(\ref{eq:chpt_fk})
1061: to our results at four points 
1062: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13781,0.13640)$,
1063: (0.13770,0.13640), (0.13754,0.13640), (0.13754,0.13660).  For these points, 
1064: the $\rho$ meson mass satisfies the condition that $m_\rho > 2m_\pi$.
1065: The measured bare AWI quark masses 
1066: are used for ${\hat m}$ and $m_{\rm s}$ 
1067: in eqs.(\ref{eq:chpt_mpi})$-$(\ref{eq:chpt_fk}).
1068: The heaviest pion mass at $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13754,0.13640)$
1069: is about 430~MeV with the use of the cutoff determined below.
1070: We summarize the pion masses and the unrenormalized AWI quark masses
1071: in Table~\ref{tab:quarkmass}.
1072: The fit results are shown in Fig.~\ref{fig:fit}, 
1073: where the black solid lines
1074: are drawn with $\kappa_{\rm s}$ fixed at 0.13640 and the black dotted 
1075: lines are for $\kappa_{\rm s}=0.13660$. 
1076: The red solid symbols represent the extrapolated values at the
1077: physical point whose determination is explained 
1078: in Sec.~\ref{sec:physicalpt} below. 
1079: The heaviest point at 
1080: $(\kappa_{\rm ud},\kappa_{\rm s})=(0.13754,0.13640)$
1081: is not well described by ChPT both for 
1082: $(a m_\pi)^2/(a m_{\rm ud}^{\rm  AWI})$ and $f_K/f_\pi$, and 
1083: $\chi^2$/d.o.f. is rather large (see  Table~\ref{tab:fit}).
1084: 
1085: 
1086: \begin{figure}[t!]
1087: \begin{center}
1088: \hspace{-0.6cm}
1089: \includegraphics[width=7cm,keepaspectratio,clip]{figs/kadoh/f.mpi2bmud.eps}
1090: \includegraphics[width=7cm,keepaspectratio,clip]{figs/kadoh/f.fKfpi.eps}
1091: \caption{Fitting results for $(am_\pi)^2/(am_{\rm ud}^{\rm AWI})$
1092:   (left) and $f_K/f_\pi$ (right). Red solid (open) symbols denote the
1093:  extrapolated values at the physical point by the ChPT formulae
1094: without (with) the finite size effects. }
1095: \label{fig:fit}
1096: \end{center}
1097: \end{figure}
1098: 
1099: 
1100: \begin{table}[h!]
1101: \centering
1102: \begin{tabular}{ccccc}\hline
1103: $L_i(\mu=m_\eta)$ &PACS-CS  & PACS-CS with FSE      & exp. value\cite{amoros01}  
1104: & MILC\cite{bernard06}
1105:     \\
1106: \hline
1107: \hline
1108: $L_4$        & 0.25(11)  &  0.23(12)      & 0.27 $\pm$ 0.8  & 0.1(2)(2)           \\
1109: $L_5$        &  2.28(13) &  2.29(14)      & 2.28 $\pm$ 0.1  & 2.0(3)(2)           \\
1110: $2L_6-L_4$ & 0.16(4)   &  0.16(4)          & 0 $\pm$ 1.0     & 0.5(1)(2)    \\ 
1111: $2L_8-L5$  & $-$0.59(5) & $-$0.60(5)         & 0.18 $\pm$ 0.5  & $-$0.1(1)(1)    \\
1112: \hline
1113: $\chi^2$/d.o.f. & 2.1(1.4) &  2.1(1.4) & &  \\
1114: \hline
1115: \end{tabular}
1116: \caption{Results for the low energy constants together with the
1117: phenomenological estimates\cite{amoros01} and the MILC results\cite{bernard06}.}
1118: \label{tab:fit}
1119: \end{table}
1120: 
1121: 
1122: 
1123: The results for the low energy constants are presented
1124: in Table~\ref{tab:fit} where the
1125: phenomenological values with the experimental
1126: inputs\cite{amoros01} and the MILC results\cite{bernard06}
1127: are also given for comparison.
1128: The renormalization scale is chosen to be $m_\eta=0.547$~GeV.   
1129: For $L_4$ and $L_5$ governing the behavior of $f_\pi, f_K$, 
1130: our results show good agreement with both the
1131: phenomenological estimates and the MILC results. 
1132: On the other hand, some discrepancies are observed between three results 
1133: for $2L_6-L_4$ and $2L_8-L_5$ which enter into the ChPT formulae 
1134: for $m_\pi^2$ and $m_K^2$.
1135: 
1136: In Fig.~\ref{fig:fit} we also draw the ChPT fit results 
1137: including the finite size effects. The green solid
1138: lines are drawn for $\kappa_{\rm s}=0.13640$ and the green dotted ones
1139: for $\kappa_{\rm s}=0.13660$. The fit curves with and without 
1140: the finite size effects are almost degenerate for 
1141: $a m_{\rm ud}^{\rm  AWI}>0.003$, but deviations appear 
1142: closer to the physical point,  for which the extrapolated values are 
1143: plotted by the open and solid red symbols.
1144: %by the red open symbols in Fig.~\ref{fig:fit}.
1145: This feature is understood by Fig.~\ref{fig:ratio} where
1146: we plot the magnitude of $R_X$ for $X=m_\pi,m_K,f_\pi,f_K$ 
1147: with $L=2.8$~fm as a function of $m_\pi$
1148: ( we note that $R_{m_{\rm PS}}>0$ and $R_{f_{\rm PS}}<0$).  
1149: The finite size effects are less than 2\% for 
1150: $m_{\rm PS}$ and $f_{\rm PS}$ at our simulation points.    
1151: For $m_{\rm PS}$ this is true even at the physical point, while 
1152: for $f_\pi$ the finite size effects cause the value to decrease by 4\%.
1153: 
1154: \begin{figure}[t!]
1155: \vspace{3mm}
1156: \begin{center}
1157: \begin{tabular}{cc}
1158:   \includegraphics[width=65mm]{figs/kadoh/f.ratio.eps}
1159: \end{tabular}
1160: \end{center}
1161: \vspace{-.5cm}
1162: \caption{$|R_X|$ ($R_{m_{\rm PS}}>0$ and $R_{f_{\rm
1163:       PS}}<0$) for $X=m_\pi,m_K,f_\pi,f_K$ with $L=2.8$~fm 
1164: as a function of $m_\pi$.
1165: Solid vertical line denotes the physical point and the dotted ones
1166: are for our simulation points.}
1167: \label{fig:ratio}
1168: \end{figure}
1169: 
1170: %\begin{figure}
1171: %\begin{minipage}{20zw}
1172: %\begin{center}
1173: %\includegraphics[width=5cm,keepaspectratio,clip]{ratio2.eps}
1174: %\end{center}
1175: %\caption{Contribution of the finite size effect}
1176: %\end{minipage}
1177: %\end{figure}
1178: %
1179: %\makeatletter
1180: %\def\@captype{table}
1181: %\makeatother
1182: %
1183: %\begin{minipage}{20zw}
1184: %\begin{tabular}{ccc}\hline
1185: %$L_i$         &             & exp. value \\
1186: %\hline
1187: %\hline
1188: %$L_4$        &  0.23(11)  &  0.27 $\pm$ 0.8          \\
1189: %$L_5$        &  2.29(14)  &  2.28 $\pm$ 0.1           \\
1190: %$2L_6-L_4$ & 0.16(4)    &  0.16 $\pm$ 1.0      \\ 
1191: %$2L_8-L5$  & -0.60(5)   &  0.18 $\pm$ 0.5      \\
1192: %\hline
1193: %\end{tabular}
1194: %\caption{Low energy constants}
1195: %\end{minipage}
1196: 
1197: %\newpage
1198: 
1199: 
1200: \subsection{Physical point and light hadron spectrum}
1201: \label{sec:physicalpt}
1202: 
1203: \begin{figure}[b!]
1204: \vspace{3mm}
1205: \begin{center}
1206: \begin{tabular}{cc}
1207: \includegraphics[width=65mm,angle=0]{figs/kadoh/f.Del_SSS.eps} &
1208: \includegraphics[width=65mm,angle=0]{figs/kadoh/f.V_SS.eps}
1209: \end{tabular}
1210: \end{center}
1211: \vspace{-.5cm}
1212: \caption{Linear chiral extrapolation for $am_\Omega$ (left) 
1213: and $am_{\phi}$ (right). 
1214: Solid (dotted) lines are drawn with $\kappa_{\rm s}=0.13640$
1215: (0.13660).
1216: Red open symbols denote the extrapolated values at the physical point
1217: with a linear form.}
1218: \label{fig:phiomega}
1219: \end{figure}
1220: 
1221: 
1222: In order to determine the up-down and the strange quark masses and the
1223: lattice cutoff we need three physical inputs.
1224: We try the following two cases: 
1225: $m_\pi, m_K, m_\Omega$ and $m_\pi, m_K, m_\phi$. 
1226: The choice of $m_\Omega$ has theoretical and practical advantages:
1227: the $\Omega$ baryon is stable in the strong interactions
1228: and its mass, being composed of three strange quarks, 
1229: is determined with good precision with small finite 
1230: size effects. 
1231: We also choose $m_\phi$ for comparison.
1232: We employ the NLO ChPT formulae for the chiral extrapolations
1233: of $m_\pi$, $m_K$, $f_\pi$ and $f_K$.  A simple linear formula
1234: $m_{\rm had}=c_0+c_1\cdot m_{\rm ud}^{\rm AWI}+c_2\cdot m_{\rm s}^{\rm AWI}$ is
1235: used for the other hadron masses, employing data in the same range   
1236: $\kappa_{\rm ud}\ge 0.13754$ as for the pseudoscalar mesons.
1237: In Fig.~\ref{fig:phiomega} we show the linear chiral extrapolations
1238: for $m_\phi$ and $m_\Omega$. The solid lines are drawn with 
1239: $\kappa_{\rm s}$ fixed at 0.13640 and the dotted ones are for 
1240: $\kappa_{\rm s}=0.13660$. 
1241: We observe that the quark mass dependences for $m_\phi$ and $m_\Omega$ 
1242: at $\kappa_{\rm ud}\ge 0.13754$ are well described by the linear function.
1243: 
1244: 
1245: 
1246: \begin{table}[t!]
1247: \centering
1248: \begin{tabular}{ccccccc}  \hline
1249:   input  & $a^{-1}$~[GeV]  &  $m^{\overline{\rm MS}}_{ud}$~[MeV] & 
1250: $m^{\overline{\rm MS}}_{s}$~[MeV] & $f_\pi$ & $f_K$ & $f_K/f_\pi$\\ \hline
1251: $m_\Omega$ & 2.256(81)  & 2.37(11) & 69.1(25) & 144(6) & 175(6) & 1.219(22) \\
1252: $m_\phi$   & 2.248(76)  & 2.38(11) & 69.4(25) & 143(6) & 175(5) & 1.219(21) \\
1253: \hline
1254: \end{tabular}
1255: \caption{Cutoff, renormalized 
1256: quark masses, pseudoscalar meson decay constants determined with 
1257: $m_\Omega$ and $m_\phi$ inputs.}
1258: \label{tab:phypoint}
1259: \end{table}
1260: 
1261: 
1262: The results for the quark masses and the lattice cutoff are listed in
1263: Table~\ref{tab:phypoint},
1264: where the errors are statistical. The two sets of results are
1265: consistent within the error.  The quark masses
1266: are smaller than the recent estimates in the literature.  We note, however, 
1267: that we employed the perturbative renormalization factors to one-loop level 
1268: which may contain  a sizable uncertainty.  A nonperutrbative calculation 
1269: of the renormalization factor is in progress using the Schr\"odinger 
1270: functional scheme. 
1271: 
1272: In Table~\ref{tab:phypoint} we also present
1273: predictions for the pseudoscalar meson decay constants
1274: at the physical point
1275: using the physical quark masses and the cutoff determined above, 
1276: which should be compared with the experimental values 
1277: $f_\pi=130.7$~MeV, $f_K=159.8$~MeV, $f_K/f_\pi = 1.223$.
1278: A 10\% discrepancy in the magnitude of $f_\pi$ and $f_K$ 
1279: might be due to use of one-loop perturbative $Z_A$ 
1280: since the ratio shows a good agreement.  
1281: A nonperturbative calculation of $Z_A$ 
1282: is also in progress. 
1283: 
1284: 
1285: \begin{figure}[t!]
1286: \vspace{3mm}
1287: \begin{center}
1288: \begin{tabular}{cc}
1289: \includegraphics[width=70mm,angle=0]{figs/kadoh/hyo.phiomega.mod.eps}
1290: %\includegraphics[width=75mm,angle=0]{f.lighthadron.eps}
1291: \end{tabular}
1292: \end{center}
1293: \vspace{-.5cm}
1294: \caption{Light hadron spectrum extrapolated at the physical point 
1295: with  $\Omega$-input (red) and $\phi$-input (blue).
1296: Horizontal bars denote the experimental values.}
1297: \label{fig:spectrum}
1298: \end{figure}
1299: 
1300: 
1301: 
1302: %\begin{figure}{r}{70mm}
1303: % \begin{center}
1304: %  \includegraphics[width=70mm]{hyo.phiomega.4.eps}
1305: %  \caption{Light hadron spectrum extrapolated at the physical point
1306: %    with $\Omega$-input (red) and $\phi$-input (blue).
1307: %Horizontal bars denote the experimental values.}
1308: %\label{fig:spectrum} 
1309: %\end{center}
1310: %\end{wrapfigure}
1311: 
1312: In Fig.~\ref{fig:spectrum} we compare the light hadron spectrum 
1313: extrapolated to the physical point with the experiment.
1314: The results for the $\Omega$-input and the $\phi$-input  
1315: are consistent with each other, and both are in agreement 
1316: with the experiment albeit errors are still not small 
1317: for some of the hadrons. This is an encouraging result.  
1318: However, further work is needed since 
1319: cutoff errors of $O((a\Lambda_{\rm QCD})^2)$ are present 
1320: in our results.
1321: 
1322: 
1323: 
1324: \section{$\rho$-$\pi\pi$ mixing}
1325: \begin{figure}[t!]
1326: \vspace{3mm}
1327: \begin{center}
1328: \begin{tabular}{cc}
1329: \includegraphics[width=65mm,angle=0]{figs/kura/E-level_p0.eps}  
1330: \hspace*{2mm}&
1331: \includegraphics[width=65mm,angle=0]{figs/kura/E-level_p1.eps}
1332: \end{tabular}
1333: \end{center}
1334: \vspace{-.5cm}
1335: \caption{Energy levels of the $\rho$ meson and the two pion states
1336: without the total momentum $p=0$ (left) and 
1337: with $p=p_{\rm min}\equiv 2\pi/L$ (right) as a function of the
1338: up-down quark mass. $\kappa_{\rm s}$ is fixed at 0.13640.}
1339: \label{fig:rhopipi_energy}
1340: \end{figure}
1341: 
1342: \begin{figure}[t!]
1343: \vspace{3mm}
1344: \begin{center}
1345: \begin{tabular}{cc}
1346: \includegraphics[width=55mm,angle=0]{figs/kura/mixing.eps}  &
1347: \includegraphics[width=65mm,angle=0]{figs/ukita/log.msn_p1_k1_s11_rho_cpca_sy.eps}
1348: \end{tabular}
1349: \end{center}
1350: \vspace{-.5cm}
1351: \caption{Schematic view of $\rho$ and $\pi\pi$ energy levels 
1352: due to mixing effects (left)
1353: and time dependence of the $R$ function (right).}
1354: \label{fig:rhopipi_r}
1355: \end{figure}
1356: 
1357: Since our simulations are carried out at sufficiently 
1358: small up-down quark masses, 
1359: %to satisfy the condition $m_\rho < 2m_\pi$,
1360: it would be interesting to investigate the $\rho$-$\pi\pi$ mixing effects.
1361: We find that the rest mass $m_\rho$ is always smaller than 
1362: the two-pion energy $2\sqrt{m_\pi^2+(2\pi/L)^2}$ for all the hopping parameters,
1363: and hence the $\rho$ meson at rest cannot decay into two pions. 
1364: However, as illustrated in Fig.~\ref{fig:rhopipi_energy}, for 
1365: a moving $\rho$ with a unit of momentum, {\it i.e.,}, its energy  
1366: $\sqrt{m_\rho^2+(2\pi/L)^2}$ becomes larger than the energy of a moving 
1367: pion and a pion at rest given by 
1368: $\sqrt{m_\pi^2+(2\pi/L)^2}+m_\pi$ when the up-down quark mass is sufficiently reduced. 
1369: 
1370: Let us consider two types of the $\rho$ meson propagator with the
1371: momentum $2\pi/L$: $\rho_\|(2\pi/L)$ with polarization parallel
1372: to the spatial momentum and $\rho_\bot(2\pi/L)$ with polarization 
1373: perpendicular to the spatial momentum.
1374: Phenomenologically the $\rho$-$\pi\pi$ coupling is described by
1375: $g_{\rho\pi\pi}\epsilon_{abc}\rho_\mu^a\pi^b\partial_\mu\pi^c$,
1376: which favors $\rho_\|(2\pi/L)\rightarrow \pi(2\pi/L)\pi(0)$
1377: to $\rho_\bot(2\pi/L)\rightarrow \pi(2\pi/L)\pi(0)$.
1378: We expect that the $\rho_\|(2\pi/L)$ propagator is more
1379: strongly affected by the mixing effects than
1380: the $\rho_\bot(2\pi/L)$ correlator.
1381: Since the mixing effects push up the upper energy level further and
1382: push down the lower energy level as shown in Fig.~\ref{fig:rhopipi_r},
1383: they could be detected by measuring the $R$ function defined by
1384: \begin{eqnarray}
1385: R(t)=\frac{\langle\rho_\|({\vec p},t)\rho_\|^\dagger({\vec p},0)\rangle}
1386: {\langle\rho_\bot({\vec p},t)\rho_\bot^\dagger({\vec p},0)\rangle}
1387: \stackrel{{\rm large\;\;}t}{\longrightarrow}
1388: Z{\rm e}^{-(E_{\rho_\|}-E_{\rho_\bot})t}.
1389: \label{eq:ratio}
1390: \end{eqnarray} 
1391: In Fig.~\ref{fig:rhopipi_r} we plot $\log|R(t)|$ as a function of $t$.
1392: The dotted horizontal lines denote $R(t)=(E/m_\rho)^2$ , which is determined
1393: kinematically in the mixing-free case.
1394: The solid lines represent the fitting
1395: results with a single exponential form over $5\le t\le 11$.
1396: The data show clear positive slopes which indicate 
1397: $E_{\rho_\|}<E_{\rho_\bot}$. 
1398: We also observe that the magnitude of the
1399: energy difference is rather small for $\kappa_{\rm ud}\le 0.13754$, while
1400: it grows rapidly as the up-down quark mass is reduced for  
1401: $\kappa_{\rm ud} > 0.13754$.
1402: This feature may suggest that the 
1403: $\langle\rho_\|({\vec p},t)\rho_\|^\dagger({\vec p},0)\rangle$ 
1404: correlator is getting dominated by the $\pi\pi$ state toward
1405: the smaller up-down quark masses.
1406: In order to obtain a definite conclusion, we 
1407: need more detailed investigations with increased statistics.
1408: 
1409: \section{Summary}
1410: We have presented a status report of the PACS-CS project which aims at 
1411: a 2+1 flavor lattice QCD simulation toward the physical point. 
1412: With the aid of the DDHMC algorithm for the up-down quarks we have 
1413: reached $m_{\pi}=210$~MeV,  which roughly corresponds to 
1414: $m_{\rm ud}^{\overline{\rm MS}}(\mu=2{\rm ~GeV})=5.6$~MeV, 
1415: on a $32^3\times 64$ lattice using the $O(a)$-improved Wilson quarks. 
1416: Thanks to the enlarged volume 
1417: compared to the previous CP-PACS/JLQCD work,
1418: we obtain good signals not only for the meson masses 
1419: but also for the baryon masses. 
1420: Our results for the hadron spectrum at the physical point
1421: show a good agreement with the experimental values.
1422:   
1423: At present we have just started the simulation at the physical point.
1424: We are also calculating  the nonperturbative renormalization factors
1425: for the quark masses and the pseudoscalar meson decay constants in 
1426: order to remove perturbative uncertainties in these important quantities.
1427: Once these calculations are accomplished,
1428: the next step is to investigate the finite size effects 
1429: at the physical point,  and then to reduce the discretization errors by 
1430: carrying out calculations at finer lattice spacings. 
1431: 
1432: \vspace{5mm}
1433: \noindent
1434: {\bf \large Acknowledgment}
1435: 
1436: We would like to thank all the collaboration members for discussions
1437: and, in particular, A.~Ukawa for a careful reading of this report.
1438: Numerical calculations for the present work have been carried out
1439: under the ``Interdisciplinary Computational Science Program'' in
1440: Center for Computational Sciences, University of Tsukuba.
1441: This work is supported in part by Grants-in-Aid for Scientific Research
1442: from the Ministry of Education, Culture, Sports, Science and Technology
1443: (Nos.~13135204, 15540251, 17340066, 17540259, 18104005, 18540250, 18740139).
1444: 
1445: \input{refs.tex}
1446: 
1447: \end{document}
1448: 
1449: \begin{table}[b!] 
1450: \begin{center}
1451: %\setlength{\tabcolsep}{10pt}
1452: \renewcommand{\arraystretch}{1.2}
1453: \begin{tabular}{c||ccccc|c} \hline
1454: $\kappa_{\rm ud}$& 0.13700&0.13727&0.13754&0.13770&0.13781&0.13754\\
1455: $\kappa_{\rm s}$&0.13640&0.13640&0.13640&0.13640&0.13640&0.13660 \\ \hline \hline
1456: $am^{\rm AWI}_{\rm ud}$&0.02800(20)&0.01895(13)&0.01020(11)&0.00521(9)&0.00227(16)&0.00908(7) \\
1457: $am^{\rm AWI}_{\rm s}$&0.04295(30)&0.04061(18)&0.03876(18)&0.03767(10)&0.03716(20)&0.03257(17)  \\
1458: $am_{\pi}$&0.32196(62)&0.26190(66)&0.18998(56)&0.13591(88)&0.08989(291)&0.17934(78) \\
1459: $am_{\rm K}$&0.36221(58)&0.32785(66)&0.29289(48)&0.27303(60)&0.25974(153)&0.26834(92) \\
1460: $am_{\rho}$&0.5064(15)&0.4541(37)&0.4155(58)&0.3912(119)&0.3524(428)&0.4030(117) \\
1461: $am_{\Phi}$&0.5555(13)&0.5322(19)&0.5219(23)&0.5109(17)&0.5049(25)&0.4908(42) \\
1462: $am_{\rm N}$&0.7259(34)&0.6487(45)&0.5657(63)&0.5144(112)&0.4681(472)&0.5471(66) \\
1463: $am_{\Omega}$&0.9123(40)&0.8754(44)&0.8559(52)&0.8323(35)&0.8247(29)&0.8022(56)  \\ 
1464: $af_{\pi}$&0.0909(13)&0.0915(17)&0.0806(15)&0.0745(10)&0.0722(18)&0.0750(10) \\
1465: $af_{\rm K}$&0.0950(12)&0.0983(14)&0.0905(13)&0.0867(12)&0.0829(19)&0.0836(9) \\  \hline
1466: \end{tabular}
1467: \caption{Results for bare AWI quark masses, meson masses, baryon masses and decay constants.}
1468: \end{center}
1469: \end{table} 
1470: