1: %\documentclass[aps,prd,preprint,showpacs,nobibnotes]{revtex4}
2: \documentclass[aps,prd,preprint,showpacs,nofootinbib]{revtex4}
3: %\documentclass[aps,prd,final,twocolumn,showpacs,nobibnotes,floatfix]{revtex4}
4: %\documentclass[fleqn,twoside]{article}
5: %\usepackage{espcrc2}
6: \usepackage{epsfig}
7: \usepackage{amsmath}
8: \usepackage{amssymb}
9: \usepackage{amsbsy}
10: \usepackage{amsfonts}
11: \usepackage{graphics}
12: \usepackage{latexsym}
13: \usepackage{mathrsfs}
14: \usepackage{dcolumn}
15: \usepackage{wasysym}
16: %\usepackage{slashed}
17: \usepackage[dvips]{color}
18: %\parskip5pt
19: \parindent15pt
20: \setlength{\textheight}{22.5cm}
21: % change this to the following line for use with LaTeX2.09
22: % \documentstyle[twoside,fleqn,espcrc2]{article}
23:
24: % if you want to include PostScript figures
25: \usepackage{graphicx}
26: % if you have landscape tables
27: \usepackage[figuresright]{rotating}
28:
29: % put your own definitions here:
30: % \newcommand{\cZ}{\cal{Z}}
31: % \newtheorem{def}{Definition}[section]
32: % ...
33: \newcommand{\ttbs}{\char'134}
34: %\newcommand{\AmS}{{\protect\the\textfont2
35: % A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}
36:
37: \newcommand{\sixth}{\mbox{\small $\frac{1}{6}$}} % 1/6
38: \newcommand{\half}{\mbox{\small $\frac{1}{2}$}} % 1/2
39: \newcommand{\quarter}{\mbox{\small $\frac{1}{4}$}} % 1/4
40: \newcommand{\third}{\mbox{\small $\frac{1}{3}$}} % 1/3
41: \newcommand{\twothird}{\mbox{\small $\frac{2}{3}$}} % 2/3
42: \newcommand{\twelfth}{\mbox{\small $\frac{1}{12}$}} % 1/12
43: \newcommand{\msbar}{{\overline{MS}}} % Msbar
44: \newcommand{\ripmom}{{RI'-MOM}} %
45: \newcommand{\mom}{\mbox{\tiny $MOM$}} % MOM
46: \newcommand{\rgi}{\mbox{\tiny $RGI$}} % RGI
47: \newcommand{\plaq}{\Box} % PLAQ
48: \newcommand{\SF}{\mbox{\tiny $SF$}} % SF
49: \newcommand{\born}{\mbox{\tiny $BORN$}} %
50: \newcommand{\NS}{\mbox{\tiny $N\!S$}} % NS
51:
52: \newcommand{\pslash}{\ensuremath{\slashed{p}}}
53: \newcommand{\Dslash}{\ensuremath{\slashed{D}}}
54: \newcommand{\Dlr}{\ensuremath{\stackrel{\leftrightarrow}{D}}}
55: \newcommand{\marginlabel}[1]{\mbox{}\marginpar{\raggedleft\hspace{0pt}#1}}
56: \newcommand{\MeV}{\ensuremath{\;\mathrm{MeV}}}
57: \newcommand{\GeV}{\ensuremath{\;\mathrm{GeV}}}
58: \newcommand{\mev}{\ensuremath{\;\mathrm{MeV}}}
59: \newcommand{\gev}{\ensuremath{\;\mathrm{GeV}}}
60: \newcommand{\Tr}{\ensuremath{\mathrm{Tr}}}
61: \newcommand{\fm}{\ensuremath{\;\mathrm{fm}}}
62: \newcommand{\ei}[2]{\ensuremath{e^{\mathrm{i}\;{#1} \cdot {#2}}}}
63: \newcommand{\emi}[2]{\ensuremath{e^{-\mathrm{i}\;{#1} \cdot {#2}}}}
64: \newcommand{\fouriersump}[3]{\ensuremath{\sum_{#1}{#3}\;e^{\mathrm{i}\;{#2}
65: \cdot {#1}}}}
66: \newcommand{\fouriersumm}[3]{\ensuremath{\sum_{#1}{#3}\;e^{-\mathrm{i}\;{#2}
67: \cdot {#1}}}}
68: \newcommand{\order}[1]{\ensuremath{{\cal{O}}\left({#1}\right)}}
69: % add words to TeX's hyphenation exception list
70:
71: \begin{document}
72:
73: \begin{raggedright}
74: DESY 08-112\\
75: Edinburgh 2008/11
76: \end{raggedright}
77:
78: \title{The electric dipole moment of the nucleon from
79: simulations at imaginary vacuum angle theta}
80:
81: \author{R.~Horsley$^{1}$,
82: T.~Izubuchi$^{2,3}$,
83: Y.~Nakamura$^{4}$, D.~Pleiter$^{4}$, P.E.L.~Rakow$^{5}$,
84: G.~Schierholz$^{4}$ and J.~Zanotti$^{1}$}
85:
86: \affiliation{\vspace*{0.3cm}
87: $^1$ School of Physics, University of Edinburgh,
88: Edinburgh EH9 3JZ, UK\\
89: $^2$ Institute for Theoretical Physics, Kanazawa University,
90: Kanazawa, 920-1192 Japan\\
91: $^3$ RIKEN-BNL Research Center, Brookhaven National Laboratory, Upton, NY
92: 11973, USA\\
93: $^4$ Deutsches Elektronen-Synchrotron DESY, John von Neumann-Institut
94: f\"ur Computing NIC, 15738 Zeuthen, Germany\\
95: $^5$ Theoretical Physics Division, Department of Mathematical
96: Sciences, University of Liverpool, Liverpool L69 3BX, UK}
97:
98: %\author{S.~Aoki$^{1,2}$, R.~Horsley$^{3}$,
99: % T.~Izubuchi$^{1,4}$,
100: % Y.~Nakamura$^{5}$, D.~Pleiter$^{5}$, P.E.L.~Rakow$^{6}$,
101: % G.~Schierholz$^{5}$ and J.~Zanotti$^{3}$}
102: %
103: %\affiliation{\vspace*{0.3cm}
104: %$^1$ RIKEN-BNL Research Center, Brookhaven National Laboratory, Upton, NY
105: % 11973, USA\\
106: %$^2$ Graduate School of Pure and Applied Sciences, University of Tsukuba,
107: %Tsukuba, 305-8571 Japan\\
108: %$^3$ School of Physics, University of Edinburgh,
109: % Edinburgh EH9 3JZ, UK\\
110: %$^4$ Institute for Theoretical Physics, Kanazawa University,
111: % Kanazawa, 920-1192 Japan\\
112: %$^5$ Deutsches Elektronen-Synchrotron DESY, John von Neumann-Institut
113: % f\"ur Computing NIC, 15738 Zeuthen, Germany\\
114: %$^6$ Theoretical Physics Division, Department of Mathematical
115: % Sciences, University of Liverpool, Liverpool L69 3BX, UK}
116:
117: %\author{QCDSF-DIK Collaboration}
118:
119: \vspace*{1cm}
120:
121: \begin{abstract}
122: We compute the electric dipole moment of proton and neutron from lattice QCD
123: simulations with $N_f=2$ flavors of dynamical quarks at imaginary vacuum angle
124: $\theta$. The calculation proceeds via the $CP$ odd form factor $F_3$. A novel
125: feature of our calculation is that we use partially twisted boundary
126: conditions to extract $F_3$ at zero momentum transfer. As a byproduct, we test
127: the QCD vacuum at nonvanishing $\theta$.
128: %We compute the electric dipole moment of proton and neutron in lattice QCD
129: %for $N_f=2$ flavors of dynamical quarks at nonvanishing imaginary vacuum angle
130: %$\theta$. For the neutron we find $\displaystyle d_N^n = -4.9(5) \, \times
131: %10^{-15} \, \theta \,e \,\mbox{cm}$. Combined with the experimental limit on
132: %the electric dipole moment this leads to $|\theta|<6 \times 10^{-12}$.
133: \end{abstract}
134:
135: \pacs{11.30.Er,11.15.Ha,12.38.Gc,13.40.Gp}
136:
137: \maketitle
138:
139: %\newpage
140: \section{Introduction}
141: Current measurements of $CP$ violating processes in the $K$ and $B$ meson
142: sector would suggest that the phase of the CKM matrix provides a complete
143: description. However, the baryon asymmetry of the universe cannot be described
144: by this phase alone, suggesting that there are additional sources of $CP$
145: violation awaiting discovery.
146:
147: QCD allows for a gauge invariant extra term in the action that is odd under
148: $CP$ transformations,
149: \begin{equation}
150: S \rightarrow S + i\,\theta\,Q ,
151: \end{equation}
152: where $Q$ is the topological charge. Hence, there is the possibility of strong
153: $CP$ violation arising from a nonvanishing vacuum angle $\theta$. The presence
154: of $CP$ violating forces implies a permanent electric dipole moment of the
155: proton and neutron. This attribute is also deeply related to the question of
156: baryon asymmetry of the universe~\cite{Trodden}.
157:
158: The current experimental bound on the electric dipole moment of the neutron
159: is~\cite{Baker}
160: \begin{equation}
161: |d_N^n| < 2.9 \,\times \, 10^{-13} \, e\,\mbox{fm}
162: \label{ub}
163: \end{equation}
164: ($\displaystyle 1\, \mbox{fm} = 10^{-13} \, \mbox{cm}$). Combining this bound
165: with theoretical estimates of $d_N/\theta$ allows us to derive an upper bound
166: on the value of $|\theta|$. Current estimates from QCD sum
167: rules~\cite{Pospelov} and chiral perturbation theory~\cite{Borasoy} give
168: $\displaystyle |\theta| \lesssim (1 - 3) \times 10^{-10}$. This anomaly is
169: known as the strong $CP$ problem.
170:
171: With the increasingly precise experimental efforts to observe the
172: electric dipole moment of the neutron~\cite{Harris}, it is important to have a
173: rigorous calculation directly from QCD.
174:
175: In this paper we present a calculation of $d_N$ in units of $\theta$ with
176: $N_f=2$ flavors of dynamical quarks using the lattice regularization. The
177: novel feature of our work is that the simulations are performed directly at
178: nonvanishing vacuum angle $\theta$, in contrast to previous lattice
179: studies~\cite{Shintani1,Shintani2,Berruto,Shintani3} (with the exception
180: of~\cite{Izubuchi}), which rely on reweighting correlation functions with
181: topological charge that would otherwise vanish. The calculation becomes
182: feasible if $\theta$ is rotated to purely imaginary
183: values~\cite{Bhanot,Azcoiti,Imachi}. We expect from our
184: method a much enhanced signal to noise ratio. In addition, we may hope to gain
185: some insight into the dynamics of the $\theta$ vacuum. For a recent review on
186: this matter see~\cite{Vicari}.
187:
188: \section{The action}
189:
190: The vacuum angle $\theta$ can be rotated into the mass term, in the continuum
191: and on the lattice~\cite{Seiler,Kerler}. This results in the fermionic action
192: \begin{equation}
193: S_F=\bar{\psi}\left\{D + [\cos(\theta/N_f) +
194: i\, \sin(\theta/N_f)\, \gamma_5]\,m\right\}\psi,
195: \label{fa}
196: \end{equation}
197: where summation over space-time coordinates $\{\vec{x},t\}$ and quark flavors
198: is understood, and $D$ is the massless Dirac operator. For simplicity,
199: we shall write in the following
200: \begin{equation}
201: S_F=\bar{\psi}\left\{D + \bar{m} +
202: i\, (\bar{\theta}/N_f)\, \gamma_5\, \bar{m} \right\}\psi,
203: \end{equation}
204: with
205: \begin{equation}
206: \begin{split}
207: \bar{m} &= \cos(\theta/N_f)\, m ,\\
208: \bar{\theta} &= \tan(\theta/N_f)\, N_f .
209: \end{split}
210: \label{bar}
211: \end{equation}
212:
213: We use clover fermions with $N_f=2$ flavors of degenerate quarks. Taking into
214: account that chiral symmetry is violated, we then have
215: \begin{equation}
216: S_F=\bar{\psi}\left\{D + \bar{m} + i\, (\theta_R/2)\, Z_m^S Z_P \,
217: \gamma_5 \, \bar{m}\right\}\psi,
218: \label{action}
219: \end{equation}
220: where $Z_m^S$ is the {\em singlet} renormalization constant of
221: the vector Ward identity ($VWI$) quark mass and $Z_P$ that of the pseudoscalar
222: density, and $\theta_R$ is the renormalized vacuum angle,
223: \begin{equation}
224: \theta_R = \left(Z_m^S Z_P\right)^{-1} \, \bar{\theta} .
225: \end{equation}
226: It can be shown that $Z_P$ is the same in both the singlet and nonsinglet case,
227: while $Z_m$ is not~\cite{qcdsf}. Note that $Z_m^S Z_P$ is scale independent,
228: and in the continuum limit $Z_m^S Z_P = 1$ and $\theta_R = \bar{\theta}$.
229:
230: \section{The simulation}
231:
232: The simulations are performed with the Iwasaki gauge action and said clover
233: fermions with $c_{SW}=1.47$ on $16^3 \, 32$ lattices at $\beta=2.1$,
234: $\kappa=0.1357$~\cite{cppacs}. The (massless) Dirac operator $D$ in
235: (\ref{action}) is evaluated at $\kappa_c=0.138984$. The bare mass is taken to
236: be $a\bar{m} = 1/(2\kappa) - 1/(2\kappa_c)$, independent of $\theta$. The
237: resulting pion mass is $m_\pi/m_\rho \approx 0.8$, corresponding to a quark
238: mass of $m\approx m_s$, $m_s$ being the strange quark mass. In~\cite{cppacs}
239: the lattice spacing in the chiral limit was estimated at $a \approx 0.11$ fm,
240: using the $\rho$ mass to set the scale.
241:
242: The vacuum angle $\bar{\theta}$ is taken to be purely imaginary,
243: \begin{equation}
244: \bar{\theta} = -i\, \bar{\theta}^{I} , \quad \bar{\theta}^{I} \in
245: \mathbb{R} ,
246: \label{choice}
247: \end{equation}
248: resulting in the action
249: \begin{equation}
250: S_F=\bar{\psi}\left\{D + \bar{m} + (\bar{\theta}^{I}/2) \,
251: \gamma_5 \, \bar{m}\right\}\psi.
252: \end{equation}
253: The simulations are done at $\bar{\theta}^{I} = 0, 0.2$, $0.4$, $1.0$ and
254: $1.5$, where we have collected $9000$, $9000$, $7000$, $6000$ and $6000$
255: trajectories of length one, respectively.
256:
257: We use the highly optimized HMC algorithm of the QCDSF
258: Collaboration~\cite{AliKhan} for updating the gauge field. After integrating
259: out the Grassmann fields, the action reads
260: \begin{equation}
261: S[U,\phi^\dagger,\phi] = S_{\rm G}[U] + S_{\rm det}[U] + \phi^\dagger
262: (Q^\dagger Q)^{-1}\phi,
263: \label{standard}
264: \end{equation}
265: where $S_G[U]$ is the Iwasaki gauge action, $\phi^\dagger$ and $\phi$ are
266: pseudo\-fermion fields, and
267: \begin{equation}
268: \begin{split}
269: S_{\rm det}[U] &= -2\, {\rm Tr}\, \log\left(1+T_{\rm
270: oo}+\frac{\hat{\theta}^{I}}{2} \, \gamma_5 \right),\\
271: Q &= \left(1+T + \frac{\hat{\theta}^{I}}{2}\, \gamma_5\right)_{\rm ee} -
272: M_{\rm eo}
273: \left(1+T +\frac{\hat{\theta}^{I}}{2}\, \gamma_5\right)^{-1}_{\rm oo}
274: M_{\rm oe}
275: \end{split}
276: \end{equation}
277: with
278: \begin{equation}
279: \hat{\theta}^{I} = \left(1 - \frac{\kappa}{\kappa_c}\right)\,
280: \bar{\theta}^{I}.
281: \end{equation}
282: $M_{\rm eo}$ and $M_{\rm oe}$ are Wilson hopping matrices, which connect
283: even with odd and odd with even sites, respectively, and $T$ is the
284: clover matrix
285: \begin{equation}
286: T=\frac{i}{2} c_{SW}\, \kappa\, \sigma_{\mu\nu} F_{\mu\nu}(x).
287: \end{equation}
288:
289: We apply mass preconditioning \`{a} la Hasenbusch~\cite{Hasenbusch} and split
290: the resulting action into three parts, each of which we put on separate time
291: scales~\cite{Sexton}. We use Omelyan's second order integrator~\cite{Omelyan}
292: to integrate Hamilton's equations of motion.
293:
294: %Following Hasenbusch~\cite{Hasenbusch}, we modify the standard action
295: %(\ref{standard}) by introducing an auxiliary matrix $W=Q+\rho$ and
296: %pseudofermion fields $\chi^\dagger$ and $\chi$:
297: %\begin{equation}
298: %\begin{split}
299: %S[U,\phi^\dagger,\phi] \rightarrow S[U,\phi^\dagger,\phi,\chi^\dagger,\chi]
300: %&= S_{\rm G}[U] + S_{\rm det}[U] \\
301: %&+ \phi^\dagger W(Q^\dagger Q)^{-1} W^\dagger \phi
302: %+ \chi^\dagger (W^\dagger W)^{-1} \chi.
303: %\label{split}
304: %\end{split}
305: %\end{equation}
306: %We now split $S[U,\phi^\dagger,\phi,\chi^\dagger,\chi]$ into one ultraviolet
307: %and two infrared parts,
308: %\begin{equation}
309: %\begin{split}
310: %S_{\rm UV} &= S_{\rm G}, \\
311: %S_{\rm IR-1} &= S_{\rm det} + \chi^\dagger (W^\dagger W)^{-1} \chi, \\
312: %S_{\rm IR-2} &= \phi^\dagger W(Q^\dagger Q)^{-1} W^\dagger \phi.
313: %\end{split}
314: %\end{equation}
315: %We put $S_{\rm UV}$, $S_{\rm IR-1}$ and $S_{\rm IR-2}$ on three separate time
316: %scales~\cite{Sexton}, $t/(n\, m)$, $t/n$ and $t$, respectively, in Hamilton's
317: %equations. Accordingly, the leap-frog integrator takes the form
318: %\begin{equation}
319: %\begin{split}
320: %V(t|n,m) = V_{\rm IR-2}\left(\frac{t}{2}\right)\, \Bigg[V_{\rm
321: % IR-1}\left(\frac{t}
322: % {2n} \right)\, &\Big[V_{\rm UV}\left(\frac{t}{2 m\,n}\right)
323: % V_Q\left(\frac{t}{m\,n}\right) V_{\rm UV}\left(\frac{t}{2 m\,n}\right)
324: % \Big]^m \\ &\times \,V_{\rm IR-1}\left(\frac{t}{2n}
325: % \right)\Bigg]^n \,
326: % V_{\rm IR-2}\left(\frac{t}{2}\right),
327: %\end{split}
328: %\end{equation}
329: %where the effect of the $V$s on the (conjugate) momenta $\{P,Q\}$ is
330: %\begin{equation}
331: %\begin{split}
332: %V_Q(t) &:\, Q \rightarrow Q + t\, P, \\
333: %V_{\rm UV}(t) &:\, P \rightarrow P - t\, \partial S_{\rm UV}, \\
334: %V_{\rm IR-1}(t) &:\, P \rightarrow P - t\, \partial S_{\rm IR-1}, \\
335: %V_{\rm IR-2}(t) &:\, P \rightarrow P - t\, \partial S_{\rm IR-2}.
336: %\end{split}
337: %\end{equation}
338: %The length of the trajectory is taken to be one, $0 \leq t \leq
339: %1$, and the integration range is divided into $1/\delta t$ steps of size
340: %$\delta t$. The computational cost is proportional to $1/\delta t$.
341: %The parameters $\rho$, $n$ and $m$ have to be tuned separately for each
342: %run. An optimal choice is $\rho=0.3$, $n=3$, $m=3$ and $\delta t = 1/40$.
343:
344: \section{Renormalization}
345:
346: Let us now compute $Z_m^S Z_P$ for our action and coupling, so that we can
347: compare the results to phenomenology later on. We demand that the renormalized
348: $VWI$ and axial vector ($AWI$) quark masses are equal,
349: \begin{equation}
350: m_R = Z_m^S \, m = \frac{Z_A}{Z_P} \,\widetilde{m} = \widetilde{m}_R.
351: \end{equation}
352: That gives $Z_m^S Z_P = Z_A \, \widetilde{m}/m$. At the largest $\kappa$ value
353: (smallest quark mass) of~\cite{cppacs} we find $\widetilde{m}/m = 1.28(2)$.
354: The renormalization constant $Z_A$ has been computed nonperturbatively
355: in~\cite{cppacs2} for a variety of couplings. Extrapolating the numbers to
356: $\beta=2.1$ gives $Z_A=0.78(1)$. Multiplying these two pieces of information
357: together, we obtain $Z_m^S Z_P = 1.00(5)$. That means $\theta_R =
358: \bar{\theta}$ to a good precision.
359:
360: Alternatively, we may compute $Z_m^S Z_P$ directly from $Z_m^S/Z_m^{NS}$ and
361: $Z_m^{NS} Z_P = Z_P/Z_S^{NS}$. As a comparison, the QCDSF Collaboration, using
362: nonperturbatively improved clover fermions and the plaquette gauge action,
363: finds at $\beta=5.4$~\cite{qcdsf} $Z_m^S/Z_m^{NS} = 1.25(5)$ and~\cite{qcdsf3}
364: $Z_P/Z_S^{NS} = 0.81(2)$. Altogether, this gives $Z_m^S Z_P = 1.01(5)$, in
365: agreement with the CP-PACS result, attesting clover fermions good chiral
366: properties.
367:
368: \section{Charge distribution and $\theta$ vacuum}
369:
370: Before we compute the electric dipole moment now, let us look at the
371: distribution of topological charge and its dependence on $\theta$. Having
372: found that $\theta_R = \bar{\theta}$, the topological charge that follows from
373: the chirally rotated action (\ref{fa}) is the so-called fermionic charge
374: \begin{equation}
375: Q= \bar{m}\,{\rm Tr}\, \gamma_5\,M^{-1} ,
376: \label{qf}
377: \end{equation}
378: where $M = D + \bar{m}$ is the fermion matrix. The evaluation of (\ref{qf})
379: requires the computation of the $O(100)$ lowest-lying eigenvalues of $M$,
380: which is numerically expensive. It has been
381: demonstrated~\cite{Kovacs,Horsley1,Horsley2} that
382: the fermionic charge and the so-called field theoretic charge, which is
383: computed from the field strength tensor by applying an appropriate number of
384: cooling sweeps and rounding the result to the nearest integer value, give
385: consistent results. In the following we shall employ the field theoretic
386: definition of the topological charge. It turns out that the results are rather
387: independent of the degree of cooling. In fact, the numbers stabilize already
388: after $O(10)$ cooling sweeps. The numbers quoted here refer to $O(100)$ cooling
389: sweeps. For a recent appraisal of the cooling method see~\cite{Ilgenfritz}.
390:
391: \begin{figure}[t]
392: \vspace*{-2cm}
393: \begin{center}
394: \epsfig{file=Qdist_all2.ps,width=10cm,clip=}
395: \caption{The topological charge distribution for $\bar{\theta}^{I}=0$,
396: $0.2$, $0.4$, $1.0$ and $1.5$, from right to left. To guide the eye, the
397: distribution is compared to a Gaussian fit at $\bar{\theta}^{I}=0$
398: and $1.0$.}
399: \end{center}
400: \end{figure}
401:
402: \begin{table}[b]
403: \begin{center}
404: \begin{tabular}{c|c|c} $\bar{\theta}^{I}$ & $\langle Q\rangle$ & $\langle
405: Q^2\rangle_c$ \\ \hline
406: 0\phantom{.0} & \phantom{0}-0.06(31) & 24.9(14) \\
407: 0.2 & \phantom{0}-3.52(46) & 24.1(15) \\
408: 0.4 & \phantom{0}-7.35(36) & 22.7(17) \\
409: 1.0 & -18.38(30) & 21.7(15) \\
410: 1.5 & -27.84(37) & 18.1(13)
411: \end{tabular}
412: \caption{The average topological charge and charge squared for our values
413: of $\bar{\theta}^{I}$.}
414: \end{center}
415: \label{avcharge}
416: \end{table}
417:
418: In Fig.~1 we show the charge distribution for our five different values of
419: $\bar{\theta}^{I}$. To identify the shape of the distribution, and to see how
420: it changes with increasing value of $\bar{\theta}^{I}$, we compare our data to
421: a Gaussian fit at $\bar{\theta}^{I} =0$ and $1.0$. A Gaussian distribution is
422: the most common distribution function for independent, randomly generated
423: variables.
424:
425: In Table~I we present the resulting average topological charge $\langle
426: Q\rangle$ and charge squared,
427: \begin{equation}
428: \langle Q^2\rangle_c \equiv \left\langle (Q - \langle Q\rangle)^2 \right\rangle
429: = \langle Q^2\rangle - \langle Q\rangle^2,
430: \end{equation}
431: which we plot in Figs.~2 and 3. As with any distribution, besides the
432: distributions mean, its skewness and kurtosis coefficients should be calculated
433: in order to determine the type of distribution. We do so in Figs.~4 and 5,
434: where we plot the skewness $S$,
435: \begin{equation}
436: S= \frac{\langle Q^3\rangle_c}{\langle Q^2\rangle_c}\, , \quad \langle
437: Q^3\rangle_c \equiv \left\langle (Q - \langle Q\rangle)^3 \right\rangle,
438: \end{equation}
439: and kurtosis $K$,
440: \begin{equation}
441: K= \frac{\langle Q^4\rangle_c}{\langle Q^2\rangle_c}\, , \quad \langle
442: Q^4\rangle_c \equiv \left\langle (Q - \langle Q\rangle)^4 \right\rangle -
443: 3 \left\langle (Q - \langle Q\rangle)^2 \right\rangle^2.
444: \end{equation}
445: Note that $S$ and $K$ have been normalized to $\langle Q^2\rangle_c$,
446: different from the mathematical literature, so that the volume dependence
447: cancels out.
448:
449: Let us first look at the charge distributions in Fig.~1. At
450: $\bar{\theta}^{I}=0$ the lattice data show a higher probability than a
451: Gaussian distributed charge for intermediate values of $|Q|$ and a slightly
452: thinner tail, while at larger values of $\bar{\theta}^{I}$ the distributions
453: appears to show a sharper peak and fatter tail. On top of that, the right tail
454: becomes longer compared to the left one with increasing value of
455: $\bar{\theta}^{I}$.
456:
457: This behavior is reflected in the skewness and kurtosis coefficients shown in
458: Figs.~4 and 5. Skewness is a measure of the degree of asymmetry of a
459: distribution. If the right tail of the distribution is more pronounced than
460: the left one, the distribution is said to have positive skewness, which is
461: what we observe at $\bar{\theta}^{I}>0$. If the reverse is true, it has
462: negative skewness. Kurtosis is the degree of peakedness of a distribution. A
463: distribution with positive kurtosis is called leptokurtic and has an acute
464: peak around its mean. Examples of leptokurtic distributions include the
465: Laplace distribution. A distribution with negative kurtosis is called
466: platykurtic and has a smaller peak around its mean and a lower probability
467: than a Gaussian distribution at extreme values. Such distributions are termed
468: sub-Gaussian.~\footnote{A distribution $P(x)=\exp[-f(x^2)]$ is called
469: sub-Gaussian if $f^\prime(x^2)$ is strictly increasing on $[0,\infty)$.}
470: The kurtosis starts out negative at $\bar{\theta}^{I}=0$ and rises almost
471: linearly to reach positive values at $\bar{\theta}^{I}\gtrsim 0.2$. For recent
472: quenched results see~\cite{DelDebbio,Durr,Giusti}.
473:
474: \clearpage
475: \begin{figure}
476: \begin{center}
477: \vspace*{-2.0cm}
478: \epsfig{file=Q_exgggg.ps,width=9.75cm,clip=}
479: \caption{The average charge (${\Circle}$) compared to the reweighted numbers
480: ($\times$) and to the predictions of the Gaussian distribution (dotted line)
481: and the dilute instanton gas (\ref{qgas}) (dashed line).}
482: \end{center}
483: \label{Qfig}
484: \end{figure}
485:
486: \begin{figure}
487: \begin{center}
488: \vspace*{-2.0cm}
489: \epsfig{file=Q2_exgggg.ps,width=9.75cm,clip=}
490: \caption{The average charge squared (${\Circle}$) compared to the reweighted
491: numbers ($\times$) and to the predictions of the Gaussian distribution
492: (dotted line) and the dilute instanton gas (\ref{qgas}) (dashed line).}
493: \end{center}
494: \label{Qfig2}
495: \end{figure}
496:
497: \clearpage
498: \begin{figure}
499: \vspace*{-2.0cm}
500: \begin{center}
501: \epsfig{file=Qdist_all_S.ps,width=9.75cm,clip=}
502: \caption{The skewness compared to the predictions of the Gaussian distribution
503: (dotted line) and the dilute instanton gas (dashed line). The errors shown
504: are the naive ones, as our sample of large charges $|Q-\langle
505: Q\rangle|$ is too small to allow for a proper jackknive analysis, and may
506: be underestimated.}
507: %The result for $\bar{\theta}^{I}=1.5$ is not shown because of large errors.}
508: \end{center}
509: \end{figure}
510:
511: \begin{figure}[t]
512: \vspace*{-2.0cm}
513: \begin{center}
514: \epsfig{file=Qdist_all_K.ps,width=9.75cm,clip=}
515: \caption{The Kurtosis, together with its naive error, compared to the
516: predictions of the Gaussian distribution
517: (dotted line) and the dilute instanton gas (dashed line).}
518: %The result for $\bar{\theta}^{I}=1.5$ is not shown because of large errors.}
519: \end{center}
520: \end{figure}
521:
522: \clearpage
523: A popular model of the QCD vacuum is the dilute instanton gas~\cite{cdg},
524: whose charge distribution is given by a convolution of separate Poisson
525: distributions for instantons and anti-instantons. Its free energy per
526: spacetime volume V is
527: \begin{equation}
528: F(\theta)= \chi_t \, (1-\cos \theta), \quad F(\theta) = - \frac{1}{V} \,
529: \ln \, Z(\theta),
530: \end{equation}
531: where
532: \begin{equation}
533: \chi_t=\left.\frac{\langle Q^2\rangle}{V}\right|_{\theta=0} .
534: \end{equation}
535: That leads to~\cite{klssw}
536: \begin{equation}
537: %\begin{split}
538: \langle Q^n\rangle_c = - i^n\, V \frac{\partial^n F(\theta)}{\partial
539: \theta^n} ,
540: %\end{split}
541: \end{equation}
542: which at imaginary $\theta$ gives
543: \begin{equation}
544: \begin{split}
545: \langle Q\rangle &= - V \chi_t\, \sinh \theta^{I}, \\
546: \langle Q^2\rangle_c & = V \chi_t\, \cosh \theta^{I}, \\
547: \langle Q^3\rangle_c & = - V \chi_t\, \sinh \theta^{I}, \\
548: \langle Q^4\rangle_c & = V \chi_t\, \cosh \theta^{I}.
549: \end{split}
550: \label{qgas}
551: \end{equation}
552:
553: In Figs.~2-5 we compare the lattice data for $\langle Q\rangle$, $\langle
554: Q^2\rangle_c$, $S$ and $K$ with the predictions of the Gaussian distribution
555: and the dilute instanton gas. While $\langle Q\rangle$ and $\langle
556: Q^2\rangle_c$ are in reasonable agreement with the results of the Gaussian
557: and Poisson distribution for smaller values of $\bar{\theta}^{I}$, the higher
558: cumulants $S$ and $K$ show a far different trend over the entire range of
559: $\bar{\theta}^{I}$ than the predictions of these simple models.
560:
561: Reweighting appears to be the accepted method for simulations at nonvanishing
562: vacuum angle $\theta$ and chemical potential $\mu$. Having results of a direct
563: simulation at nonvanishing value of $\theta$ at hand, it is instructive to
564: test how reliable the method actually is, given the fact that the simulations
565: are necessarily restricted to a finite volume with limited absolute value of
566: the topological charge.
567:
568: The reweighted charges are given by
569: \begin{equation}
570: \langle Q^n \rangle = \frac{1}{Z(\theta)} \sum_Q Q^n \,P_Q \,{\rm
571: e}^{-i\,\theta\, Q}, \quad \sum_Q P_Q = 1,
572: \end{equation}
573: where $P_Q$ denotes the probability of finding a configuration of
574: charge $Q$ in the ensemble of configurations, and
575: \begin{equation}
576: Z(\theta) = \sum_Q P_Q \,{\rm e}^{-i\,\theta\, Q}.
577: \end{equation}
578: At imaginary $\theta = -i\,\theta^{I}$ this becomes
579: \begin{equation}
580: \langle Q^n \rangle = \frac{1}{Z(\theta)} \sum_Q Q^n \,P_Q \,{\rm
581: e}^{- \theta^{I}\, Q}, \quad Z(\theta) = \sum_Q P_Q \,{\rm
582: e}^{- \theta^{I}\, Q}.
583: \end{equation}
584:
585: In Figs.~2 and 3 we compare $\langle Q\rangle$ and $\langle Q^2\rangle_c$ with
586: the reweighted numbers, where we have converted $\theta$ to $\bar{\theta}$
587: using (\ref{bar}). On the quantitative level, reweighting is not able to
588: describe the data for $\bar{\theta}^{I}\geq 0.4$. The reason is that the
589: reweighted charge distributions have largely the same form as the initial
590: $\bar{\theta}^{I}=0$ distribution, but are merely shifted towards negative $Q$
591: values, while the shape of the true charge distributions changes significantly
592: with increasing value of $\bar{\theta}^{I}$. We may expect to find better
593: agreement on larger volumes, provided $Z(\theta)$ is analytic in $\theta$.
594: %We note that this comparison should not depend on the method used for
595: %calculating the topological charge, since it is done the same way in both
596: %cases.
597:
598: \section{Nucleon form factors at $\mathbf{\theta \neq 0}$}
599:
600: At nonvanishing $\theta$ the electromagnetic current between nucleon states
601: can be decomposed in Euclidean space into
602: \begin{equation}
603: \langle p^\prime,s^\prime|J_\mu|p,s\rangle =
604: \bar{u}_\theta(\vec{p}^{\,\prime},s^\prime)\, \mathcal{J}_\mu\,
605: u_\theta(\vec{p},s),
606: \label{me}
607: \end{equation}
608: with
609: \begin{equation}
610: \mathcal{J}_\mu = \gamma_\mu F_1^\theta(q^2) + \sigma_{\mu\nu} q_\nu
611: \frac{F_2^\theta(q^2)}{2m_N^\theta} + \left[ (\gamma q\, q_\mu -
612: \gamma_\mu \, q^2) \, \gamma_5 \, F_A^\theta(q^2) + \sigma_{\mu\nu} q_\nu \,
613: \gamma_5 \frac{F_3^\theta(q^2)}{2m_N^\theta}\right],
614: \label{current}
615: \end{equation}
616: where $q=p^\prime - p$. The form factors and nucleon mass will generally
617: depend on $\theta$ with $F_{\cdots}^{\theta=0} = F_{\cdots}$ and
618: $m_N^{\theta=0} = m_N$. The Dirac spinors are modified by a phase in the
619: $\theta$ vacuum,
620: \begin{equation}
621: \begin{split}
622: u_\theta(\vec{p},s) &= {\rm e}^{i\alpha(\theta)\gamma_5}\, u(\vec{p},s),\\
623: \bar{u}_\theta(\vec{p},s) &= \bar{u}(\vec{p},s)\, {\rm
624: e}^{i\alpha(\theta)\gamma_5},
625: \end{split}
626: \end{equation}
627: so that the standard spinor relation is modified to
628: \begin{equation}
629: \sum_{s^\prime,s} u_\theta(\vec{p},s^\prime) \bar{u}_\theta(\vec{p},s) = {\rm
630: e}^{i\alpha(\theta)\gamma_5} \left(\frac{-i \gamma p + m_N^\theta}{2
631: E_N^\theta}\right) {\rm e}^{i\alpha(\theta)\gamma_5}.
632: \end{equation}
633: As we are primarily interested in the electric dipole moment in the limit
634: $\theta \rightarrow 0$, it is sufficient to consider the lowest order
635: expansion only. Hence, we may write
636: \begin{equation}
637: \alpha(\theta) = \alpha^\prime \, \theta + O(\theta^3).
638: \label{phase}
639: \end{equation}
640: For our choice of $\theta$ (\ref{choice}), this then becomes to lowest order
641: in $\theta$
642: \begin{equation}
643: \sum_{s^\prime,s} u_\theta(\vec{p},s^\prime) \bar{u}_\theta(\vec{p},s) =
644: \frac{-i \gamma p + m_N (1+2\alpha^\prime \bar{\theta}^{I} \gamma_5)}{2 E_N}.
645: \end{equation}
646: Note that in Euclidean space $q^2 = - (E^\prime - E)^2 + (\vec{p}^{\,\prime} -
647: \vec{p})^2$ and $\gamma p = i E \gamma_4 + \vec{\gamma}\vec{p}$.
648:
649: We denote the two-point function of a nucleon of momentum $\vec{p}$ in the
650: theta vacuum by $G_{NN}^\theta(t,\vec{p})$. The phase factor $\alpha^\prime$
651: of (\ref{phase}) can be obtained from the ratio of two-point functions
652: \begin{equation}
653: \begin{split}
654: {\rm Tr}\, [G_{NN}^\theta(t;0) \Gamma_4] &\simeq \frac{1}{2} |Z_N|^2\, {\rm
655: e}^{-m_N t}, \\
656: {\rm Tr}\, [G_{NN}^\theta(t;0) \Gamma_4 \gamma_5] &\simeq - \alpha^\prime
657: \bar{\theta}^{I}\, \frac{1}{2} |Z_N|^2\, {\rm
658: e}^{-m_N t},
659: \end{split}
660: \end{equation}
661: where $\Gamma_4 = (1+\gamma_4)/2$. In Fig.~6 we show
662: \begin{equation}
663: R(t) = \frac{{\rm Tr}\, [G_{NN}^\theta(t,0) \Gamma_4 \gamma_5]}{{\rm Tr}\,
664: [G_{NN}^\theta(t;0) \Gamma_4]} \simeq - \alpha^\prime \bar{\theta}^{I}
665: \label{ratio}
666: \end{equation}
667: for $\bar{\theta}^{I}=0.4$. Fitting to a constant, we find
668: \begin{equation}
669: \begin{tabular}{c|c}
670: $\bar{\theta}^{I}$ & $\alpha^\prime \bar{\theta}^{I}$ \\ \hline
671: 0.2 & 0.048(3) \\
672: 0.4 & 0.081(4)
673: \end{tabular}
674: \end{equation}
675: %
676: \begin{figure}[t]
677: \vspace*{0.25cm}
678: \begin{center}
679: \epsfig{file=Ratio-f1N.eps,width=10cm,clip=}
680: \caption{The ratio $R(t$) given in (\ref{ratio}) for $\bar{\theta}^{I} = 0.4$.}
681: \end{center}
682: \end{figure}
683: %
684: The form factor $F_3(q^2)$, which is needed for the determination of the
685: nucleon electric dipole moment, can be extracted from the ratio of three-point
686: and two-point functions
687: \begin{equation}
688: \begin{split}
689: R_\mu(t^\prime,t;\vec{p}^{\,\prime},\vec{p}) &= \frac{G_{NJ_\mu
690: N}^{\theta\, \Gamma}(t^\prime,t;\vec{p}^{\,\prime},\vec{p})}{{\rm Tr}\,
691: [G_{NN}^\theta(t^\prime;\vec{p}^{\,\prime})\Gamma_4]} \\[0.5em]
692: &\times \left\{\frac{{\rm
693: Tr}\,[G_{NN}^\theta(t;\vec{p}^{\,\prime})\Gamma_4]\,
694: {\rm Tr}\,[G_{NN}^\theta(t^\prime;\vec{p}^{\,\prime})\Gamma_4]\,
695: {\rm Tr}\,[G_{NN}^\theta(t^\prime-t;\vec{p})\Gamma_4]}
696: {{\rm Tr}\,[G_{NN}^\theta(t;\vec{p})\Gamma_4]\,
697: {\rm Tr}\,[G_{NN}^\theta(t^\prime;\vec{p})\Gamma_4]\,
698: {\rm
699: Tr}\,[G_{NN}^\theta(t^\prime-t;\vec{p}^{\,\prime})\Gamma_4]}\right\}^{1/2}
700: \\[0.75em]
701: &= \sqrt{\frac{E^{\theta\,\prime} \,E^\theta}
702: {(E^{\theta\,\prime}+m_N^\theta) \,
703: (E^\theta+m_N^\theta)}} \, F(\Gamma,\mathcal{J}_\mu),
704: \end{split}
705: \label{rw}
706: \end{equation}
707: where $G_{NJ_\mu N}^{\theta\, \Gamma}((t^\prime,t;\vec{p}^{\,\prime},\vec{p})$
708: is the three-point function, with $t^\prime$ being the time location of
709: the nucleon sink and $t$ the time location of the current insertion, and the
710: function $F(\Gamma,\mathcal{J}_\mu)$ is
711: \begin{equation}
712: \begin{split}
713: F(\Gamma,\mathcal{J}_\mu) = \frac{1}{4}\, {\rm Tr}\, \Gamma &\left[{\rm
714: e}^{i\alpha(\theta)\gamma_5} \frac{E^{\theta\,\prime}\gamma_4
715: -i\vec{\gamma}\vec{p}^{\,\prime} + m_N^\theta}{E^{\theta\,\prime}}\,
716: {\rm e}^{i\alpha(\theta)\gamma_5}\right]\\[0.5em]
717: &\times \mathcal{J}_\mu \left[{\rm
718: e}^{i\alpha(\theta)\gamma_5} \frac{E^{\theta}\gamma_4
719: -i\vec{\gamma}\vec{p} + m_N^\theta}{E^{\theta}}\,
720: {\rm e}^{i\alpha(\theta)\gamma_5}\right]
721: \end{split}
722: \end{equation}
723: with $\mathcal{J}_\mu$ given in (\ref{current}). The three-point functions are
724: calculated for various choices of nucleon polarization, $\Gamma = \Gamma_4$,
725: $i\Gamma_4\gamma_5\gamma_1$ and $i\Gamma_4\gamma_5\gamma_2$. The calculation
726: follows the program of the QCDSF Collaboration~\cite{qcdsf4} for computing
727: nucleon three-point functions. We neglect finite volume corrections to two-
728: and three-point functions~\cite{Liu}.
729: For $J_\mu$ we take the local vector current, which needs to be
730: renormalized. We compute the corresponding renormalization constant
731: $Z_V$~\cite{qcdsf2} from the proton form factor $F_1(0)$ at zero momentum
732: transfer.
733:
734: With conventional (periodic) boundary conditions momenta are quantized in
735: units of $2\pi/L$, where $L$ is the spatial extent of the lattice. For the
736: lattices used in the current simulation, this means that the smallest
737: nonvanishing momentum available is $\approx 700 \, \mbox{MeV}$. Since $F_3$
738: can only be computed at $q^2 \neq 0$, we need to extrapolate to $q^2=0$ to
739: find $d_N$.
740: The momentum resolution of hadron observables can be significantly improved by
741: varying the boundary conditions. It was demonstrated~\cite{Sachrajda} that for
742: processes without final state interactions, such as the form factors studied
743: in this paper, it is sufficient to apply twisted boundary conditions to the
744: valence quarks only.
745:
746: In our study we use partially twisted boundary conditions, {\it i.e.}\
747: combining gauge field configurations generated with sea quarks with periodic
748: spatial boundary conditions with valence quarks with twisted boundary
749: conditions. The boundary conditions of the valence quarks attached to the
750: electromagnetic current are
751: \begin{equation}
752: \psi(x_k + L) = {\rm e}^{i\,\alpha_k} \, \psi(x_k), \quad k=1, 2, 3.
753: \end{equation}
754: %
755: By varying $\vec{\alpha}$ we
756: can tune the momenta of the nucleon continuously. We have chosen the following
757: set of twist angles
758: \begin{equation}
759: \begin{split}
760: \vec{\alpha} &= \frac{2\pi}{L} \, (0,0,0) \\
761: \vec{\alpha} &= \frac{2\pi}{L} \, (0.36,0,0) \\
762: \vec{\alpha} &= \frac{2\pi}{L} \, (0.36,0.36,0) \\
763: \vec{\alpha} &= \frac{2\pi}{L} \, (0.36,0.36,0.36)
764: \end{split}
765: \end{equation}
766: The dispersion relation for the nucleon then reads
767: \begin{equation}
768: E=\sqrt{m_N^2 + (\vec{p}+\vec{\alpha})^2} .
769: \end{equation}
770:
771: We define the electric dipole moment as
772: \begin{equation}
773: d_N^\theta = \frac{e\,F_3^\theta(0)}{2m_N^\theta}.
774: \end{equation}
775: In Fig.~7 we show our results for $F_3$ for proton and neutron on
776: configutaions with $\bar{\theta}^{I}=0.2$ (top) and $\bar{\theta}^{I}=0.4$
777: (bottom). We find a very
778: clean signal for the neutron form factor. The signal for the proton, on the
779: other hand, is somewhat more noisy. The reason is that in this case one has to
780: subtract $F_1^p(0)$ from the matrix element (\ref{current}), which is by far
781: the largest contribution.
782:
783: To obtain the
784: result at $q^2=0$, we first attempt a fit using a dipole ansatz
785: \begin{equation}
786: F_3^\theta(q^2) = \frac{F_3^\theta(0)}{(1+q^2/M^2)^2},
787: \end{equation}
788: which is indicated by the solid lines. At $q^2=0$ we find
789: \begin{equation}
790: \begin{tabular}{c|c|c}
791: $\bar{\theta}^{I}$ & $F_3^p(0)/(2m_N)$ & $F_3^n(0)/(2m_N)$\\ \hline
792: 0.2 & $\phantom{-}0.158(33)$ & $-0.108(17)$ \\
793: 0.4 & $\phantom{-}0.256(25)$ & $-0.193(12)$
794: \end{tabular}
795: \label{dp}
796: \end{equation}
797:
798: Alternatively, if we assume that $F_3$ and $F_1^p$ have similar $q^2$
799: behavior, then by forming the ratio
800: \begin{equation}
801: \frac{F_3^\theta(q^2)}{F_1^{\theta\, p}(q^2)},
802: \end{equation}
803:
804: \clearpage
805: \begin{figure}[t!]
806: \vspace*{0.5cm}
807: \begin{center}
808: %%\epsfig{file=F3P-02.eps,width=10cm,clip=}\\[1cm]
809: %%\epsfig{file=F3P.eps,width=10cm,clip=}
810: \epsfig{file=F3P_0.2.eps,width=11cm,clip=}\\[1cm]
811: \epsfig{file=F3P_0.4.eps,width=11cm,clip=}
812: %\epsfig{file=/home/gsch/dN/james/Th2/F3P.eps,width=11cm,clip=}\\[1cm]
813: %\epsfig{file=/home/gsch/dN/james/Th4/F3P.eps,width=11cm,clip=}
814: \vspace*{0.4cm}
815: \caption{The form factor $F_3(q^2)$ for proton and neutron, together with a
816: dipole fit, for $\bar{\theta}^{I}=0.2$ (top) and $\bar{\theta}^{I}=0.4$
817: (bottom), respectively.}
818: \end{center}
819: \end{figure}
820:
821:
822: \clearpage
823: \begin{figure}[t!]
824: \vspace*{0.5cm}
825: \begin{center}
826: %%\epsfig{file=F3PoF1P-02.eps,width=10cm,clip=}\\[1cm]
827: %%\epsfig{file=F3PoF1P.eps,width=10cm,clip=}
828: \epsfig{file=F3PoF1P_0.2.eps,width=11cm,clip=}\\[1cm]
829: \epsfig{file=F3PoF1P_0.4.eps,width=11cm,clip=}
830: %\epsfig{file=/home/gsch/dN/james/Th2/F3PoF1P.eps,width=11cm,clip=}\\[1cm]
831: %\epsfig{file=/home/gsch/dN/james/Th4/F3PoF1P.eps,width=11cm,clip=}
832: \vspace*{0.4cm}
833: \caption{The form factor ratio $F_3(q^2)/F_1(q^2)$ for proton and neutron, for
834: $\bar{\theta}^{I}=0.2$ (top) and $\bar{\theta}^{I}=0.4$ (bottom),
835: respectively.}
836: \end{center}
837: \end{figure}
838:
839: \clearpage
840: \noindent
841: the renormalization constant $Z_V$ cancels, and we may hope to see a constant
842: behavior as a function of $q^2$. In Fig.~8 we show this ratio. Again, we find
843: a very clean signal for the neutron. After performing a linear extrapolation
844: to $q^2=0$, we obtain
845: \begin{equation}
846: \begin{tabular}{c|c|c}
847: $\bar{\theta}^{I}$ & $F_3^p(0)/(2m_N)$ & $F_3^n(0)/(2m_N)$\\ \hline
848: 0.2 & $\phantom{-}0.141(16)$ & $-0.088(8)\phantom{0}$ \\
849: 0.4 & $\phantom{-}0.257(18)$ & $-0.190(9)\phantom{0}$
850: \end{tabular}
851: \label{ra}
852: \end{equation}
853:
854:
855: \section{Electric dipole moment}
856:
857: \begin{figure}[t]
858: \vspace*{-2cm}
859: \begin{center}
860: \epsfig{file=dN_ex2.ps,width=11cm,clip=}
861: \caption{The electric dipole moment $d_N^\theta$ for proton and neutron,
862: together with a linear plus cubic fit. The data points are horizontally
863: displaced for better legibility.}
864: \end{center}
865: \end{figure}
866:
867: Both values of $F_3(0)$, (\ref{dp}) and (\ref{ra}), are consistent with each
868: other within the error bars. In Fig.~9 we show the results together with a fit
869: of the form
870: \begin{equation}
871: d_N^\theta = \frac{\partial d_N^\theta}{\partial \bar{\theta}^{I}}\,
872: \bar{\theta}^{I} + c\,\bar{\theta}^{I\; 3},
873: \end{equation}
874: as we are mainly interested in the derivative $\partial d_N^\theta/\partial
875: \bar{\theta}^{I}$. The fit gives at $\bar{\theta}^{I} =0$
876: \begin{equation}
877: \begin{split}
878: \frac{\partial d_N^\theta}{\partial \bar{\theta}^{I}} &=
879: \phantom{-}0.080(10)\;\; [e \times \mbox{fm}] \quad \mbox{Proton}, \\[0.35em]
880: \frac{\partial d_N^\theta}{\partial \bar{\theta}^{I}} &=
881: -0.049(5)\phantom{0}\;\; [e \times \mbox{fm}] \quad \mbox{Neutron}.
882: \end{split}
883: \end{equation}
884: Combining the upper experimental bound on the electric dipole moment of the
885: neutron (\ref{ub}) with our result for $\displaystyle \partial
886: d_N^\theta/\partial \bar{\theta}^{I}$, we may derive an upper bound on the
887: vacuum angle $\theta$. Taking our results at face value, we find
888: \begin{equation}
889: |\theta| < 6 \times 10^{-12}.
890: \label{bound}
891: \end{equation}
892: It should be noted, however, that we are working at unphysically large quark
893: mass yet, so that this result has limited phenomenological significance.
894: %For comparison, QCD sum rules give~\cite{Pospelov} $\displaystyle |\theta| < 3
895: %\times 10^{-10}$.
896:
897: \section{Conclusion and outlook}
898:
899: We have performed simulations of QCD with $N_f=2$ flavors of dynamical quarks
900: at imaginary vacuum angle $\theta$. It is the first time this has been done in
901: full QCD. The use of partially twisted boundary conditions has allowed us to
902: compute the proton and neutron form factor $F_3(q^2)$ with high precision over
903: the entire range of momenta down to $(aq)^2 \approx 0.02$, which greatly
904: facilitated the extrapolation to $q^2=0$.
905:
906: A further improvement of our calculation is that it does not require
907: reweighting of the three-point functions (\ref{rw}) with the topological
908: charge. Barring the fact that the lattice definition of topological charge is
909: ambiguous, to some extent, reweighting does not describe the charge
910: distribution accurately beyond $\bar{\theta}^{I} > 0.2$, which casts some
911: doubts on the method, if taken at face value.
912:
913: Having demonstrated the benefit of simulations at imaginary $\theta$, the next
914: step is to extend the calculations to more realistic quark masses and larger
915: lattices. The idea then is to make contact to the predictions of chiral
916: perturbation theory, which allow for the extrapolation of the dipole moment
917: from finite to infinite volume and to the physical quark
918: mass~\cite{Borasoy,Savage}. Chiral perturbation theory also predicts the
919: $\theta$ dependence of physical quantities, such as hadron masses. For the
920: pion mass one finds~\cite{Brower}
921: \begin{equation}
922: m_\pi^2(\theta)=m_\pi^2(0)\,\cos(\theta/N_f).
923: \label{mpitheta}
924: \end{equation}
925: Though our quark mass is rather heavy, it is tempting to compare our results
926: with (\ref{mpitheta}). This is done in Fig.~10. The variation of $m_\pi$ with
927: $\theta$ is found to be significant. Our results do not confirm the
928: predictions of chiral perturbation theory. It will be interesting to see if
929: this behavior persists at smaller quark masses.
930:
931: \begin{figure}[t]
932: \vspace*{-2cm}
933: \begin{center}
934: \epsfig{file=mpi_theta2.ps,width=11cm,clip=}
935: \caption{The pion mass squared as a function of $\theta$ squared. The dashed
936: line shows the prediction of chiral perturbation theory.}
937: \end{center}
938: \end{figure}
939:
940: A caveat of our calculation is that clover fermions, though $O(a)$ improved,
941: break chiral symmetry at finite lattice spacing. It is reassuring that $Z_m^S
942: Z_P \approx 1$, which indicates good chiral properties already. But there is
943: the potential danger that the remaining $O(a^2)$ corrections will interfere
944: with the assumed form of the nucleon matrix element (\ref{rw}) and give rise to
945: systematic errors~\cite{Aoki}. To fully rule that out, we need to repeat the
946: simulations at smaller lattice spacing.~\footnote{As an independent check for
947: lattice artifacts, we have repeated the calculation for zero $\theta$
948: angle of the valence quarks on our $\bar{\theta}^I=0.4$ dynamical background
949: field configurations and found a nonvanishing result for $d_N^\theta$. The
950: result would have been zero in the absence of vacuum insertions of the
951: pseudoscalar density~\cite{Guadagnoli}.}
952:
953: Finally, we plan to explore a far wider range of $\theta$ values, intrigued by
954: the results of a recent simulation of the $O(3)$ nonlinear sigma model in two
955: dimension at imaginary $\theta$~\cite{Alles}. By analytic continuation to real
956: values of $\theta$ it was possible to detect the phase transition of the model
957: at $\theta=\pi$ and show that the mass gap vanishes at this point, in
958: agreement with known results~\cite{Bietenholz}.
959:
960: \begin{acknowledgments}
961: We like to thank Wolfgang Bietenholz for carefully reading the manuscript and
962: %Sinya Aoki and
963: Luigi Del Debbio for useful discussions. The simulations of the
964: background gauge field have been performed on the BlueGene/L at KEK under the
965: {\it Large Scale Simulation Program 07-14}, while the analysis and computation
966: of the form factors have been done on the APE computers at DESY Zeuthen. We
967: thank both institutions for their support. This work is supported in part by
968: DFG under contract FOR 465 (Forschergruppe Gitter-Hadronen-Ph\"anomenologie)
969: and 446JAP113/345/0-1 (DFG-JSPS Cooperation Agreement), and by the EU
970: Integrated Infrastructure Initiative Hadron Physics (I3HP) under contract
971: RII3-CT-2004-506078. JZ is supported by STFC Grant PP/D000238/1.
972: \end{acknowledgments}
973:
974: \begin{thebibliography}{99}
975:
976: \bibitem{Trodden}
977: M.~Trodden,
978: %``Electroweak baryogenesis,''
979: Rev.\ Mod.\ Phys.\ {\bf 71}, 1463 (1999)
980: [arXiv:hep-ph/9803479].
981:
982: \bibitem{Baker}
983: C.~A.~Baker {\it et al.},
984: %``An improved experimental limit on the electric dipole moment of the
985: %neutron,''
986: Phys.\ Rev.\ Lett.\ {\bf 97}, 131801 (2006)
987: [arXiv:hep-ex/0602020];
988: C.~A.~Baker {\it et al.},
989: %``Reply to Comment on ``An Improved Experimental Limit on the Electric
990: %Dipole Moment of the Neutron'',''
991: Phys.\ Rev.\ Lett.\ {\bf 98}, 149102 (2007)
992: [arXiv:0704.1354 [hep-ex]].
993:
994: \bibitem{Pospelov}
995: M.~Pospelov and A.~Ritz,
996: %``Theta induced electric dipole moment of the neutron via {QCD} sum rules,''
997: Phys.\ Rev.\ Lett.\ {\bf 83}, 2526 (1999)
998: [arXiv:hep-ph/9904483].
999:
1000: \bibitem{Borasoy}
1001: B.~Borasoy,
1002: %``The electric dipole moment of the neutron in chiral perturbation
1003: %theory,''
1004: Phys.\ Rev.\ D {\bf 61}, 114017 (2000)
1005: [arXiv:hep-ph/0004011].
1006:
1007: \bibitem{Harris}
1008: P.~G.~Harris,
1009: %``The Neutron EDM Experiment,''
1010: arXiv:0709.3100 [hep-ex].
1011:
1012: \bibitem{Shintani1}
1013: E.~Shintani {\it et al.},
1014: %``Neutron electric dipole moment from lattice QCD,''
1015: Phys.\ Rev.\ D {\bf 72}, 014504 (2005)
1016: [arXiv:hep-lat/0505022].
1017:
1018: \bibitem{Shintani2}
1019: E.~Shintani {\it et al.},
1020: %``Neutron electric dipole moment with external electric field method in
1021: %lattice QCD,''
1022: Phys.\ Rev.\ D {\bf 75}, 034507 (2007)
1023: [arXiv:hep-lat/0611032].
1024:
1025: \bibitem{Berruto}
1026: F.~Berruto {\it et al.},
1027: %``Calculation of the neutron electric dipole moment with two dynamical
1028: %flavors of domain wall fermions,''
1029: Phys.\ Rev.\ D {\bf 73}, 054509 (2006)
1030: [arXiv:hep-lat/0512004].
1031:
1032: \bibitem{Shintani3}
1033: E.~Shintani, S.~Aoki and Y.~Kuramashi,
1034: %``Full QCD calculation of neutron electric dipole moment with the external
1035: %electric field method,''
1036: arXiv:0803.0797 [hep-lat].
1037:
1038: \bibitem{Izubuchi}
1039: T.~Izubuchi {\it et al.},
1040: %``Dynamical QCD simulation with theta terms,''
1041: arXiv:0802.1470 [hep-lat].
1042:
1043: \bibitem{Bhanot}
1044: G.~Bhanot and F.~David,
1045: %``The Phases Of The O(3) Sigma Model For Imaginary Theta,''
1046: Nucl.\ Phys.\ B {\bf 251}, 127 (1985).
1047:
1048: \bibitem{Azcoiti}
1049: V.~Azcoiti {\it et al.},
1050: %``New proposal for numerical simulations of theta-vacuum like systems,''
1051: Phys.\ Rev.\ Lett.\ {\bf 89}, 141601 (2002)
1052: [arXiv:hep-lat/0203017].
1053:
1054: \bibitem{Imachi}
1055: M.~Imachi {\it et al.},
1056: %``The Theta-Term, Cp(N-1) Model And The Inversion Approach In The Imaginary
1057: %Theta Method,''
1058: Prog.\ Theor.\ Phys.\ {\bf 116}, 181 (2006).
1059:
1060: \bibitem{Vicari}
1061: E.~Vicari and H.~Panagopoulos,
1062: %``Theta dependence of SU(N) gauge theories in the presence of a topological
1063: %term,''
1064: arXiv:0803.1593 [hep-th].
1065:
1066: \bibitem{Seiler}
1067: E.~Seiler and I.~O.~Stamatescu,
1068: %``Lattice Fermions And Theta Vacua,''
1069: Phys.\ Rev.\ D {\bf 25}, 2177 (1982)
1070: [Erratum-ibid.\ D {\bf 26}, 534 (1982)].
1071:
1072: \bibitem{Kerler}
1073: W.~Kerler,
1074: %``Axial Vector Anomaly In Lattice Gauge Theory,''
1075: Phys.\ Rev.\ D {\bf 23}, 2384 (1981).
1076:
1077: \bibitem{qcdsf}
1078: M.~G\"ockeler {\it et al.},
1079: %``Determination of light and strange quark masses from full lattice QCD,''
1080: Phys.\ Lett.\ B {\bf 639}, 307 (2006)
1081: [arXiv:hep-ph/0409312].
1082:
1083: \bibitem{cppacs}
1084: A.~Ali Khan {\it et al.},
1085: %``Light hadron spectroscopy with two flavors of dynamical quarks on the
1086: %lattice,''
1087: Phys.\ Rev.\ D {\bf 65}, 054505 (2002)
1088: [Erratum-ibid.\ D {\bf 67}, 059901 (2003)]
1089: [arXiv:hep-lat/0105015].
1090:
1091: \bibitem{AliKhan}
1092: A.~Ali Khan {\it et al.},
1093: %``Accelerating the hybrid Monte Carlo algorithm,''
1094: Phys.\ Lett.\ B {\bf 564}, 235 (2003)
1095: [arXiv:hep-lat/0303026];
1096: M.~G\"ockeler {\it et al.},
1097: %``Simulating at realistic quark masses: Light quark masses,''
1098: PoS {\bf LAT2006}, 160 (2006)
1099: [arXiv:hep-lat/0610071];
1100: M.~G\"ockeler {\it et al.},
1101: %``A status report of the QCDSF $N_f=2+1$ Project,''
1102: PoS {\bf LAT2007}, 041 (2007)
1103: [arXiv:0712.3525 [hep-lat]].
1104:
1105: \bibitem{Hasenbusch}
1106: M.~Hasenbusch,
1107: %``Speeding up the Hybrid-Monte-Carlo algorithm for dynamical fermions,''
1108: Phys.\ Lett.\ B {\bf 519}, 177 (2001)
1109: [arXiv:hep-lat/0107019].
1110:
1111: \bibitem{Sexton}
1112: J.~C.~Sexton and D.~H.~Weingarten,
1113: %``Hamiltonian evolution for the hybrid Monte Carlo algorithm,''
1114: Nucl.\ Phys.\ B {\bf 380}, 665 (1992).
1115:
1116: \bibitem{Omelyan}
1117: I.~P.~Omelyan, I.~M.~Mryglod and R.~Folk,
1118: Comput. Phys. Commun. 151, 272 (2003).
1119:
1120: \bibitem{cppacs2}
1121: S.~Aoki {\it et al.},
1122: %``Non-perturbative renormalization for a renormalization group improved
1123: %gauge action,''
1124: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 106}, 780 (2002)
1125: [arXiv:hep-lat/0110128].
1126:
1127: \bibitem{qcdsf3}
1128: M. G\"ockeler {\it et al.} [QCDSF Collaboration], in preparation.
1129:
1130: \bibitem{Kovacs}
1131: T.~G.~Kovacs,
1132: %``The topological susceptibility with dynamical overlap fermions,''
1133: arXiv:hep-lat/0111021.
1134: %%CITATION = HEP-LAT/0111021;%%
1135:
1136: \bibitem{Horsley1}
1137: R.~Horsley {\it et al.},
1138: %``Low-lying fermion modes of N(f) = 2 improved Wilson fermions,''
1139: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 106}, 569 (2002)
1140: [arXiv:hep-lat/0111030].
1141:
1142: \bibitem{Horsley2}
1143: R.~Horsley {\it et al.},
1144: %``Low-lying fermion modes: Dynamical versus quenched,''
1145: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 119}, 763 (2003)
1146: [arXiv:hep-lat/0211030].
1147:
1148: \bibitem{Ilgenfritz}
1149: E.~M.~Ilgenfritz {\it et al.},
1150: %``Vacuum structure revealed by over-improved stout-link smearing compared
1151: %with the overlap analysis for quenched QCD,''
1152: Phys.\ Rev.\ D {\bf 77}, 074502 (2008)
1153: [arXiv:0801.1725 [hep-lat]].
1154:
1155: \bibitem{DelDebbio}
1156: L.~Del Debbio, H.~Panagopoulos and E.~Vicari,
1157: %``Theta dependence of SU(N) gauge theories,''
1158: JHEP {\bf 0208}, 044 (2002)
1159: [arXiv:hep-th/0204125].
1160:
1161: \bibitem{Durr}
1162: S.~D\"urr {\it et al.},
1163: %``Precision study of the SU(3) topological susceptibility in the
1164: %continuum,''
1165: JHEP {\bf 0704}, 055 (2007)
1166: [arXiv:hep-lat/0612021].
1167:
1168: \bibitem{Giusti}
1169: L.~Giusti, S.~Petrarca and B.~Taglienti,
1170: %``Theta dependence of the vacuum energy in the SU(3) gauge theory from the
1171: %lattice,''
1172: Phys.\ Rev.\ D {\bf 76}, 094510 (2007)
1173: [arXiv:0705.2352 [hep-th]].
1174:
1175: \bibitem{cdg}
1176: C.~G.~Callan, R.~F.~Dashen and D.~J.~Gross,
1177: %``Toward a theory of the strong interactions,''
1178: Phys.\ Rev.\ D {\bf 17}, 2717 (1978).
1179:
1180: \bibitem{klssw}
1181: A.~S.~Kronfeld {\it et al.},
1182: %``THE theta VACUUM IN SU(2) LATTICE GAUGE THEORY,''
1183: Nucl.\ Phys.\ B {\bf 305}, 661 (1988).
1184:
1185: \bibitem{qcdsf4}
1186: M.~G\"ockeler {\it et al.},
1187: %``Polarized and unpolarized nucleon structure functions from lattice QCD,''
1188: Phys.\ Rev.\ D {\bf 53}, 2317 (1996)
1189: [arXiv:hep-lat/9508004];
1190: M.~G\"ockeler {\it et al.},
1191: %``Nucleon form factors: Probing the chiral limit,''
1192: PoS {\bf LAT2006}, 120 (2006)
1193: [arXiv:hep-lat/0610118].
1194:
1195: \bibitem{Liu}
1196: K.~F.~Liu,
1197: %``Neutron Electric Dipole Moment at Fixed Topology,''
1198: arXiv:0807.1365 [hep-ph].
1199:
1200: \bibitem{qcdsf2}
1201: T.~Bakeyev {\it et al.},
1202: %``Non-perturbative renormalisation and improvement of the local vector
1203: %current for quenched and unquenched Wilson fermions,''
1204: Phys.\ Lett.\ B {\bf 580}, 197 (2004)
1205: [arXiv:hep-lat/0305014].
1206:
1207: \bibitem{Sachrajda}
1208: C.~T.~Sachrajda and G.~Villadoro,
1209: %``Twisted boundary conditions in lattice simulations,''
1210: Phys.\ Lett.\ B {\bf 609}, 73 (2005)
1211: [arXiv:hep-lat/0411033].
1212:
1213: \bibitem{Savage}
1214: D.~O'Connell and M.~J.~Savage,
1215: %``Extrapolation formulas for neutron EDM calculations in lattice QCD,''
1216: Phys.\ Lett.\ B {\bf 633}, 319 (2006)
1217: [arXiv:hep-lat/0508009].
1218:
1219: \bibitem{Brower}
1220: R.~Brower {\it et al.},
1221: %``QCD at fixed topology,''
1222: Phys.\ Lett.\ B {\bf 560}, 64 (2003)
1223: [arXiv:hep-lat/0302005].
1224:
1225: \bibitem{Aoki}
1226: S.~Aoki {\it et al.},
1227: %``Calculating the neutron electric dipole moment on the lattice,''
1228: Phys.\ Rev.\ Lett.\ {\bf 65}, 1092 (1990).
1229:
1230: \bibitem{Guadagnoli}
1231: D.~Guadagnoli {\it et al.},
1232: %``Neutron electric dipole moment on the lattice: A theoretical
1233: %reappraisal,''
1234: JHEP {\bf 0304}, 019 (2003)
1235: [arXiv:hep-lat/0210044].
1236:
1237: \bibitem{Alles}
1238: B.~Alles and A.~Papa,
1239: %``Mass gap in the 2D O(3) non-linear sigma model with a theta=pi term,''
1240: arXiv:0711.1496 [cond-mat.stat-mech].
1241:
1242: \bibitem{Bietenholz}
1243: W.~Bietenholz, A.~Pochinsky and U.~J.~Wiese,
1244: %``Meron Cluster Simulation Of The Theta Vacuum In The 2-D O(3) Model,''
1245: Phys.\ Rev.\ Lett.\ {\bf 75}, 4524 (1995)
1246: [arXiv:hep-lat/9505019].
1247:
1248: \end{thebibliography}
1249:
1250: \end{document}
1251:
1252:
1253:
1254: