1: %\documentclass[aps,preprint,preprintnumbers,12pt]{revtex4}
2: \documentclass[aps,prl,twocolumn]{revtex4}
3: \usepackage{epsfig,psfrag}
4:
5: \def\nn{\nonumber}
6:
7: \begin{document}
8:
9: \title{Precise Quark Mass Dependence of Instanton Determinant}
10: \author{Gerald V. Dunne}
11: %\email{dunne@phys.uconn.edu}
12: \affiliation{Department of Physics, University of Connecticut, Storrs, CT 06269, USA}
13: \author{Jin Hur}
14: % \email{hurjin2@snu.ac.kr}
15: \author{Choonkyu Lee}
16: %\email{cklee@phya.snu.ac.kr}
17: \affiliation{Department of Physics and Center for Theoretical
18: Physics\\ Seoul National University, Seoul 151-742, Korea}
19: \author{Hyunsoo Min}
20: %\email{hsmin@dirac.uos.ac.kr}
21: \affiliation{Department of Physics, University of Seoul, Seoul 130-743, Korea\\ Department of Physics, University of Connecticut, Storrs, CT 06269, USA}
22:
23:
24:
25:
26: \begin{abstract}
27: The fermion determinant in an instanton background for a quark field of arbitrary mass is determined exactly using an efficient numerical method to evaluate the determinant of a partial wave radial differential operator. The bare sum over partial waves is divergent but can be renormalized in the minimal subtraction scheme using the result of WKB analysis of the large partial wave contribution. Previously, only a few leading terms in the extreme small and large mass limits were known for the corresponding effective action. Our approach works for any quark mass and interpolates smoothly between the analytically known small and large mass expansions.
28: \end{abstract}
29: \maketitle
30:
31:
32: To study instanton-related physics in QCD it is of fundamental importance
33: to determine the one-loop tunnelling amplitude given by the
34: Euclidean one-loop effective action in an instanton background \cite{shifman}.
35: This quantity involves the fermion determinant in a nontrivial gauge configuration and so is also of interest for unquenching effects in dynamical quark simulations in lattice QCD \cite{montvay}.
36: In the 1970s, 't Hooft \cite{thooft} calculated analytically this one-loop amplitude for {\it massless} scalar or quark fields, but this exact calculation
37: is no longer possible if the fields have nonzero mass. The small mass limit was extended further by Carlitz and Creamer \cite{carlitz}, and by Kwon et al. \cite{kwon}. This small mass limit is also closely related to the study of zero modes and spectra of the Dirac operator, which have been investigated extensively recently in lattice QCD \cite{bruckmann}. In the other extreme, the large mass
38: limit is naturally obtained from the Schwinger-DeWitt (or the heat-kernel) expansion
39: \cite{dewitt,lee,nsvz,fliegner} within the proper-time representation of
40: the effective action. For phenomenological applications \cite{shifman}, and also for the extrapolation of lattice results \cite{lattice}
41: obtained at unphysically large quark masses to lower physical
42: masses, it is important to be able to connect the large and small mass regimes.
43: In this paper, we provide the first precise bridge between these two
44: extremes of small and large mass, by presenting a method which
45: computes the exact effective action in an instanton background for {\it any} value of the quark mass.
46: The resulting mass dependence interpolates smoothly between the explicitly known results at the opposite ends.
47:
48: Our computational procedure is based on an efficient way to compute one-dimensional determinants, combined with the fact that an instanton background is sufficiently symmetric that the corresponding effective action can be reduced to a sum over partial waves (see details below). This bare sum over partial waves is of course divergent, and the nontrivial part of the calculation is to renormalize this divergent sum in a physically unambiguous manner. We use the minimal subtraction scheme \cite{thooft,kwon} and a WKB analysis \cite{wkb}, which gives the exact counterterms, including finite parts. The result is a simple finite expression for the renormalized one-loop effective action [see (\ref{answer}) below] which can be evaluated numerically.
49:
50:
51: Due to a hidden supersymmetry \cite{thooft,jackiw}, the spinor
52: Dirac operator in an instanton background has the same spectrum
53: (except for zero modes and an overall multiplicity factor of $4$) as
54: that of the scalar Klein-Gordon operator in the same instanton
55: background. This implies that the one-loop effective action of a
56: Dirac spinor field of mass $m$ (and isospin $\frac{1}{2}$),
57: $\Gamma^{F}_{\rm ren}(A;m)$, is directly related to the
58: corresponding scalar effective action (for a complex scalar of mass
59: $m$ and isospin $\frac{1}{2}$) by \cite{thooft,brown,kwon}:
60: $\Gamma^{F}_{\rm ren}(A;m)=
61: -\frac{1}{2}\ln\left(\frac{m^2}{\mu^2}\right) -2\, \Gamma^{S}_{\rm
62: ren}(A;m)$, where $\mu$ is a renormalization scale. The logarithmic contribution
63: corresponds to the existence of a zero eigenvalue in the spectrum of
64: the Dirac operator for a single instanton background. This
65: relationship (which is special to a self-dual background) has the
66: important consequence that it is sufficient to consider the scalar
67: effective action $\Gamma^{S}_{\rm ren}(A;m)$
68: to learn also about the
69: corresponding fermion effective action $\Gamma^{F}_{\rm ren}(A;m)$,
70: for any mass $m$.
71:
72: As in \cite{thooft}, we consider an SU(2) single instanton of scale $\rho$ :
73: $A_{\mu}(x) = \frac{\eta_{\mu\nu
74: a}\tau^{a}x_{\nu}}{r^2+\rho^2}$.
75: Rather than the scalar effective action $\Gamma^{S}_{\rm ren}(A;m)$,
76: it is convenient to consider $\tilde{\Gamma}^{S}_{\rm ren}(m\rho)$,
77: a function of $m\rho$ only,
78: defined by
79: \begin{equation}
80: \Gamma^{S}_{\rm ren}(A;m)=
81: \frac{1}{6}\ln(\mu\rho)+\tilde{\Gamma}^{S}_{\rm ren}(m\rho) .
82: \label{modaction}
83: \end{equation}
84: The factor $\frac{1}{6}$ in (\ref{modaction}) is determined by the
85: one-loop $\beta$ function. We can then concentrate on the $m\rho$
86: dependence of $\tilde{\Gamma}^{S}_{\rm ren}(m\rho)$,
87: and we can set the instanton scale $\rho=1$. In the small and large
88: mass limits, it is known that $\tilde{\Gamma}^{S}_{\rm ren}(m)$
89: behaves as \cite{nsvz,kwon}
90: \begin{eqnarray}
91: \tilde{\Gamma}^{S}_{\rm ren}(m)= &&\label{masslimit}\\
92: &&\hskip -2cm
93: \begin{cases}
94: {\alpha\left(\frac{1}{2}\right)+\frac{1}{2}\left(\ln m+\gamma-\ln
95: 2\right)m^2 +\dots \quad , \quad m\rightarrow 0 \cr
96: -\frac{\ln m}{6}-\frac{1}{75 m^2}-\frac{17}{735 m^4}+\frac{232}{2835 m^6}-\frac{7916}{148225 m^8}+\cdots \cr
97: \hspace{5.8cm} , \quad m\rightarrow \infty}
98: \end{cases}
99: \nonumber
100: \end{eqnarray}
101: where $\alpha(\frac{1}{2})\simeq 0.145873$, and $\gamma\simeq 0.5772\dots$ is Euler's constant.
102: These extreme small and large mass limits
103: (\ref{masslimit}) are shown as dashed curves in Figure \ref{figure1}. There is clearly a significant gap preventing
104: extrapolation between the small and large mass limits, and for mass between $0.5$ and $1$
105: (in units of $1/\rho$) neither of these approximations is
106: particularly accurate. In this paper we provide a simple numerical
107: procedure to determine $\tilde{\Gamma}^{S}_{\rm ren}(m)$,
108: and hence the fermion determinant, for {\it any} value of the mass
109: $m$, not just asymptotically small or large masses. Our result is
110: \begin{eqnarray}
111: \tilde{\Gamma}^S_{\rm ren}(m)&=&\left(\sum_{l=0,\frac{1}{2},\dots}^L \hskip -8pt \Gamma_l^S\right) +2 L^2 + 4 L-\left(\frac{1}{6}+\frac{m^2}{2}\right)\ln L \nonumber\\
112: &&\hskip -1cm +\left[\frac{127}{72}-\frac{1}{3}\ln 2+\frac{m^2}{2}-m^2 \ln 2+\frac{m^2}{2}\ln m \right],
113: \label{answer}
114: \end{eqnarray}
115: where $L$ is a large integer. The sum over partial-wave contributions $\Gamma_l^S$ is done numerically, as described below, in Eq. (\ref{lj1}). The other terms in
116: (\ref{answer}) are renormalization terms, computed using minimal
117: subtraction and WKB. The renormalized effective action
118: $\tilde{\Gamma}^S_{\rm ren}(m)$ in (\ref{answer}) is finite,
119: converges for large $L$, and can be computed for any mass $m$.
120: Figure 1 shows that our numerical results provide a very precise
121: interpolation between the extreme small and large mass limits in
122: (\ref{masslimit}).
123: \begin{figure}[h]
124: \includegraphics[scale=.9]{figure1.eps}
125: \caption{Plot of our numercial results for $\tilde{\Gamma}^{S}_{\rm ren}(m)$ from (\protect{\ref{answer}}), compared with the analytic extreme small and large mass limits [dashed curves] from (\protect{\ref{masslimit}}). The dots are numerical data points from (\protect{\ref{answer}}), and the solid line is a fit through these points.
126: \label{figure1}}
127: \end{figure}
128:
129:
130: To derive (\ref{answer}), note that the regularized one-loop scalar effective action has the proper-time representation
131: \begin{eqnarray}
132: \Gamma_{\Lambda}^{S}(A;m) &=& - \int_{0}^{\infty}
133: \frac{ds}{s}(e^{-m^2 s}-e^{-\Lambda^2 s}) \nonumber\\
134: &&\hskip 1cm \times \int d^4
135: x\;\textrm{tr}\langle x|{e^{-s(-{\rm D}^2
136: )}-e^{-s(-\partial^2 )}}|x\rangle \nonumber \\
137: &\equiv& - \int_{0}^{\infty} \frac{ds}{s}(e^{-m^2 s}-e^{-\Lambda^2
138: s}) F(s),
139: \label{ptaction}
140: \end{eqnarray}
141: From this one obtains the
142: renormalized effective action, in the minimal subtraction scheme \cite{thooft,kwon}:
143: \begin{eqnarray}
144: \Gamma^{S}_{\rm ren}(A;m)
145: &=&
146: \lim_{\Lambda\rightarrow\infty}
147: \left[\Gamma_{\Lambda}^{S}(A;m)-\frac{1}{12}
148: \ln\left(\frac{\Lambda^2}{\mu^2}\right)\right].
149: \label{renaction}
150: \end{eqnarray}
151: Recall that the massive Klein-Gordon (KG) operator, $-D^2+m^2$, for scalars of isospin $\frac{1}{2}$ in the instanton background (with scale $\rho=1$) reduces to the radial form \cite{thooft}
152: \begin{eqnarray}
153: {\mathcal M}_{(l,j)} &=&- \frac{\partial^2}{\partial
154: r^2}-\frac{3}{r}\frac{\partial}{\partial
155: r}+\frac{4l(l+1)}{r^2}+\frac{4(j-l)(j+l+1)}{r^2+1}\nonumber\\
156: &&\hskip 1cm -\frac{3}{(r^2+1)^2} +m^2,
157: \label{instop}
158: \end{eqnarray}
159: where $l=0,\,
160: \frac{1}{2},\,1,\,\frac{3}{2},\dots\;$; $j=| l \pm \frac{1}{2}|$, and with degeneracy factor $d_{(l,j)}=(2l+1)(2j+1)$. Without the instanton
161: background, the free Klein-Gordon operator is
162: \begin{equation}
163: {\mathcal M}^{{\rm free}}_{(l)}=- \frac{\partial^2}{\partial
164: r^2}-\frac{3}{r}\frac{\partial}{\partial r}+\frac{4l(l+1)}{r^2} +m^2,
165: \label{freeop}
166: \end{equation}
167: with degeneracy factor $d_{(l)}=(2l+1)^2$.
168: Thus,
169: it is natural \cite{wkb} to combine the partial waves $(l,j=l+\frac{1}{2})$ and $(l+\frac{1}{2},j=l)$, which have common degeneracy factor $(2l+1)(2l+2)$, so that the sum over $l$ and $j$ reduces to a sum over $l$. This sum is, however, divergent and we must define a consistent regularization and renormalization. To this end, we use the proper-time regularization in (\ref{ptaction}), and split the $l$ sum as follows:
170: \begin{eqnarray}
171: \Gamma_\Lambda^S(A; m)&=& \hskip -5pt \sum_{l=0,\frac{1}{2},\dots}^L \hskip -5pt
172: \Gamma_{\Lambda, (l)}^S(A; m) + \hskip -5pt \sum_{l=L+\frac{1}{2}}^\infty \hskip -5pt \Gamma^S_{\Lambda, (l)}(A; m)
173: \label{actionsplit}
174: \end{eqnarray}
175: where $L$ is a large but finite integer. In the first sum, which is finite, the cutoff $\Lambda$
176: is irrelevant and we can use a numerical method [described below] to
177: evaluate the sum. For the second sum, we use a combination of the
178: WKB approximation, which is good for large $l$, and Euler-Maclaurin
179: summation to perform the sum.
180:
181: To begin, we consider the first sum in (\ref{actionsplit}):
182: \begin{eqnarray}
183: \sum_{l=0,\frac{1}{2},\dots}^L \Gamma_l^S(A; m) & =& \sum_{l=0,\frac{1}{2}, \dots}^L (2l+1)(2l+2) \nonumber\\
184: &&
185: \hskip -3cm \times
186: \left\{ \ln \det \left(\frac{{\mathcal M}_{(l,l+\frac{1}{2})}}{{\mathcal M}^{{\rm free}}_{(l)}}\right)+ \ln \det \left(\frac{{\mathcal M}_{(l+\frac{1}{2},l)}}{{\mathcal M}^{{\rm free}}_{(l+\frac{1}{2})}}\right)\right\}
187: \label{lj}
188: \end{eqnarray}
189: These one-dimensional determinants can be computed efficiently using the following result \cite{levit,coleman,dreyfus,forman,kirsten}: Suppose ${\mathcal M}_1$ and ${\mathcal M}_2$ are two second order ordinary differential operators on the interval $r\in [0,\infty)$, with Dirichlet boundary conditions. Then the ratio of the determinants of ${\mathcal M}_1$ and ${\mathcal M}_2$ is given by
190: \begin{eqnarray}
191: \det\left(\frac{{\mathcal M}_1}{{\mathcal M}_2}\right)=\lim_{R\to\infty}\left(\frac{\psi_1(R)}{\psi_2(R)}\right)
192: \label{theorem}
193: \end{eqnarray}
194: where $\psi_i(r)$ ($i=1,2$) satisfies the initial value problem
195: \begin{eqnarray}
196: {\mathcal M}_i\, \psi_i(r)=0\quad ; \psi_i(r=0)=0\,\, ; \psi^\prime_i(r=0)=1
197: \label{ode}
198: \end{eqnarray}
199: Since an initial value problem is very simple to solve numerically, this theorem provides an efficient way to compute the determinant of an ordinary differential operator. Note in particular that no direct information about the spectrum (either bound or continuum states, or phase shifts) is required in order to compute the determinant.
200:
201: We can simplify the numerical
202: computation further because for the free KG operator, ${\mathcal
203: M}^{{\rm free}}_{(l)} $, the solution to (\ref{ode}) is just the modified Bessel function: $ \psi^{\rm free}_{(l)}(r)=\frac{I_{2l+1}(m
204: r)}{r}$. Since this grows exponentially fast at large $r$, this should also
205: be true of the numerical solutions of (\ref{ode}) for the instanton
206: operators, ${\mathcal M}_{(l,j)} $,
207: in (\ref{instop}). Thus, it is numerically better to consider the
208: ODE satisfied by the {\it ratio} (a similar idea was used by Baacke
209: et al in their analysis of metastable vacuum decay \cite{baacke}) :
210: $T_{(l,j)}(r)=\frac{\psi_{(l,j)}(r)}{\psi^{\rm free}_{(l)}(r)}$.
211: In fact, since we are ultimately interested in the logarithm of the determinant, it is more convenient (and more numerically stable) to consider $S_{(l,j)}=\ln T_{(l,j)}$, which satisfies
212: \begin{eqnarray}
213: &&\frac{d^2 S_{(l,j)}}{dr^2}+\left(\frac{d S_{(l,j)}}{dr}\right)^2+\left(\frac{1}{r}+2m\frac{I^\prime_{2l+1}(m r)}{I_{2l+1}(m r)}\right)\frac{d S_{(l,j)}}{dr}\nonumber\\
214: &&\hskip 1cm = \frac{4(j-l)(j+l+1)}{r^2+1}-\frac{3}{(r^2+1)^2}
215: \label{nonlinear}
216: \end{eqnarray}
217: where $S_{(l,j)}(r)$ satisfies the initial conditions $S_{(l,j)}(0)=0$ and $S^\prime_{(l,j)}(0)=0$.
218: Thus, the first, finite, sum in (\ref{actionsplit}) for the bare effective action can be evaluated as
219: \begin{eqnarray}
220: \sum_{l=0,\frac{1}{2},\dots}^L \Gamma_l^S(A; m)&=&\sum_{l=0,\frac{1}{2}, \dots}^L (2l+1)(2l+2) \nonumber\\
221: &&\hskip -2cm \times \left\{ S_{(l,l+\frac{1}{2})}(r=\infty)+S_{(l+\frac{1}{2},l)}(r=\infty) \right\}.
222: \label{lj1}
223: \end{eqnarray}
224: \begin{figure}[h]
225: \includegraphics[scale=0.5]{figure2.eps}
226: \caption{Plot of the $l$ dependence of $P(l)\equiv S_{(l,l+\frac{1}{2})}(r=\infty)+S_{(l+\frac{1}{2},l)}(r=\infty)$, for $m=1$. The sum of these two asymptotic values is $O(\frac{1}{l})$ for large $l$.
227: \label{figure2}}
228: \end{figure}
229: The large $r$ values of $S_{(l,l+\frac{1}{2})}(r)$ and $S_{(l+\frac{1}{2},l)}(r)$ can be extracted with excellent precision.
230: In fact, the asymptotic values of $S_{(l,l+\frac{1}{2})}(r)$ and
231: $S_{(l+\frac{1}{2},l)}(r)$ very nearly cancel one another, and for a
232: given mass, as a function of $l$, the combination
233: $[S_{(l,l+\frac{1}{2})}(r=\infty)+S_{(l+\frac{1}{2},l)}(r=\infty)]$
234: falls off %lee2, of->off
235: in magnitude like $\frac{1}{l}$ -- see Figure \ref{figure2}. Thus,
236: the sum in (\ref{lj1}) will have terms going like, $L^2$, $L$ and
237: $\ln L$, as well as terms finite and vanishing for large $L$. We now
238: show that these potentially divergent terms are exactly canceled by
239: terms in the second sum in (\ref{actionsplit}).
240:
241:
242: The second sum in (\ref{actionsplit}) can be analyzed using the Euler-Maclaurin method \cite{bender} as follows. Write
243: \begin{eqnarray}
244: \sum_{l=L+\frac{1}{2}}^\infty \hskip -5pt \Gamma^S_{\Lambda, (l)}(A; m)\hskip -2pt =\hskip -4pt \int_0^\infty \hskip -8pt ds
245: \left[-\frac{1}{s}\left(e^{-m^2 s}-e^{-\Lambda^2
246: s}\right)f(s) \right] \label{second}
247: \end{eqnarray}
248: Using WKB, which is good for large $l$, we can write \cite{wkb}
249: \begin{eqnarray}
250: f(s)=\int_0^\infty dr \left(\sum_{l=L+\frac{1}{2}}^\infty
251: f_l(s,r)\right) \label{wkb}
252: \end{eqnarray}
253: where for each $l$, $f_l(s,r)$ has a local expansion in terms of the Langer-modified potential \cite{langer}
254: \begin{eqnarray}
255: V_{(l,j)}^{\rm Lang}\hskip -2pt =\hskip -2pt \frac{(2l+1)^2}{r^2}+\frac{4(j-l)(j+l+1)}{r^2+1}-\frac{3}{(r^2+1)^2}
256: \label{langer}
257: \end{eqnarray}
258: The first three orders of the WKB approximation for $f_l$ were computed in \cite{wkb}. We find $f_l(s,r)=(2l+1)(2l+2)[f_{(l,l+\frac{1}{2})}(s,r)+f_{(l+\frac{1}{2},l)}(s,r)]$, where to first order
259: \begin{eqnarray}
260: f_{(l,j)}^{(1)}(s,r)=\frac{1}{2\sqrt{\pi s}}\, e^{-[s V_{(l,j)}^{\rm Lang}(r)]},
261: \label{wkbfirst}
262: \end{eqnarray}
263: and to second order
264: \begin{eqnarray}
265: f_{(l,j)}^{(2)}(s,r)\hskip -2pt =\hskip -2pt \frac{1}{2\sqrt{\pi s}}\hskip -2pt\left(\frac{s}{4 r^2}-\frac{s^2}{12}\frac{d^2 V_{(l,j)}^{\rm Lang}}{dr^2}\right)\hskip -3pt e^{-[s V_{(l,j)}^{\rm Lang}(r)]}
266: \label{wkbsecond}
267: \end{eqnarray}
268: The sum over $l$ in (\ref{wkb}) can now be performed using the Euler-Maclaurin expansion \cite{bender} :
269: \begin{eqnarray}
270: \sum_{l=L+\frac{1}{2}}^\infty f_l\, = 2 \int_L^\infty dl\, f(l)-\frac{1}{2} f(L)-\frac{1}{24}f^\prime(L)+\dots
271: \label{euler}
272: \end{eqnarray}
273: For each order of the WKB expansion, the $l$ dependence is such that the $l$ integral in
274: (\ref{euler}) can be done exactly. Then in the large $L$ limit, the $r$ and $s$ integrals in (\ref{wkb})
275: and (\ref{second}) can also be done, leading to \cite{inprep}
276: \begin{eqnarray}
277: \sum_{l=L+\frac{1}{2}}^\infty \hskip -5pt \Gamma^S_{\Lambda, (l)}&\sim& \frac{1}{6}\ln \Lambda+2 L^2 + 4 L-\left(\frac{1}{6}+\frac{m^2}{2}\right)\ln L \nonumber\\
278: && \hskip -2.5cm +\left[\frac{127}{72}-\frac{1}{3}\ln 2+\frac{m^2}{2}-m^2 \ln
279: 2+\frac{m^2}{2}\ln m \right]+\dots
280: \label{div}
281: \end{eqnarray}
282: Only the first two orders of the WKB expansion are required for this result; all higher orders contribute terms vanishing in the large $L$ limit. It is important to identify the physical role of the various terms in (\ref{div}). The first term is the expected $\frac{1}{6}\ln \Lambda$ term canceled in (\ref{renaction}). The next three terms give the quadratic, linear and logarithmic divergences in $L$ which cancel the corresponding divergences in the first sum in (\ref{actionsplit}), which were found in our numerical data -- see Fig. \ref{figure2}. It is a highly nontrivial check on the WKB computation that these terms have the correct coefficients to cancel these divergences. Combining (\ref{div}) and (\ref{lj1}), we obtain our result (\ref{answer}).
283:
284: As an interesting application, our result (\ref{answer}) provides a very simple analytic computation of 't Hooft's leading small mass result. When $m=0$, the numerical integration can be done exactly and one finds that \cite{inprep}
285: \begin{eqnarray}
286: S_{(l,l+\frac{1}{2})}(r=\infty)+S_{(l+\frac{1}{2},l)}(r=\infty)=\ln\left(\frac{2l+1}{2l+2}\right)
287: \label{exact}
288: \end{eqnarray}
289: Then it follows from (\ref{answer}) that
290: \begin{eqnarray}
291: \tilde{\Gamma}^S_{\rm ren}(m=0)&=&\sum_{l=0,\frac{1}{2},\dots}^L (2l+1)(2l+2)\ln\left(\frac{2l+1}{2l+2}\right)
292: \nonumber\\
293: &&+2 L^2 + 4 L-\frac{1}{6}\ln L +\frac{127}{72}-\frac{1}{3}\ln 2 \nonumber\\
294: &=& -\frac{17}{72}-\frac{1}{6}\ln 2 +\frac{1}{6}-2\zeta^\prime(-1)+O(\frac{1}{L})\nonumber\\
295: & \buildrel{L\to\infty}\over \longrightarrow &
296: \alpha\left(\frac{1}{2}\right)= 0.145873... \label{masslessanswer}
297: \end{eqnarray}
298: which agrees precisely with the leading term in the small mass limit
299: in (\ref{masslimit}).
300:
301:
302: For other, nonzero, values of the mass $m$ we found excellent convergence in computing $\tilde{\Gamma}^S_{\rm ren}(m)$ in (\ref{answer}) with $L=50$, combined with Richardson extrapolation \cite{bender}. In Figure \ref{figure1} these results are compared to the analytic large and small mass expansions in (\ref{masslimit}) --
303: the agreement is remarkable. Thus, our expression (\ref{answer}) provides a simple and numerically exact interpolation between the large mass and small mass regimes.
304:
305: In conclusion, beyond the instanton problem considered in this paper, our approach should provide a practical numerical scheme for the one-loop effective action in a broad class of nonabelian background fields.
306:
307:
308:
309: \vskip .5cm
310: {\bf Acknowledgments:} GD thanks T. Blum,
311: H. Gies, V. Khemani and K. Kirsten for helpful discussions and correspondence, and the US DOE for support through the grant DE-FG02-92ER40716. The work of CL was supported by the Korea Science Foundation ABRL program (R14-2003-012-01002-0). HM thanks the University of Connecticut for hospitality during sabbatical, and the University of Seoul for support.
312:
313:
314: \begin{thebibliography}{99}
315:
316:
317: \bibitem{shifman}
318: M.~A.~Shifman,
319: {\it Instantons in gauge theories}, (World Scientific, Singapore, 1994);
320: T. Sch\"{a}fer and E. V. Shuryak, Rev. Mod.
321: Phys. \textbf{70}, 323 (1998);
322: M.~A.~Shifman,
323: {\it ITEP lectures on particle physics and field theory}, (World Scientific, Singapore, 1999);
324: %%CITATION = 00327,62,1;%%
325: D. Diakonov, Prog. Part. Nucl. Phys. {\bf 5}, 173 (2003).
326:
327: \bibitem{montvay}
328: I.~Montvay and G.~Munster,
329: {\it Quantum fields on a lattice}, (Cambridge Univ Press, 1994).
330:
331:
332: \bibitem{thooft}
333: G. 't Hooft, Phys. Rev. \textbf{D14}, 3432 (1976).
334:
335: \bibitem{carlitz} R. D. Carlitz and D. B. Creamer, Ann. Phys. (NY) \textbf{118}, 429 (1979).
336:
337: \bibitem{kwon} O-K. Kwon, C. Lee and H. Min, Phys. Rev.
338: \textbf{D62}, 114022 (2000).
339:
340: \bibitem{bruckmann}
341: P.~H.~Damgaard,
342: %%``The microscopic Dirac operator spectrum,''
343: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 106}, 29 (2002);
344: %[arXiv:hep-lat/0110192].
345: %%CITATION = HEP-LAT 0110192;%%
346: F.~Bruckmann, M.~Garcia Perez, D.~Nogradi and P.~van Baal,
347: %``Calorons and fermion zero-modes,''
348: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 129}, 727 (2004);
349: %[arXiv:hep-lat/0308017];
350: %%CITATION = HEP-LAT 0308017;%%
351: C.~Gattringer and R.~Pullirsch,
352: %``Topological lumps and Dirac zero modes in SU(3) lattice gauge theory on the
353: %torus,''
354: Phys.\ Rev.\ D {\bf 69}, 094510 (2004);
355: %[arXiv:hep-lat/0402008].
356: %%CITATION = HEP-LAT 0402008;%%
357: E.~Follana, A.~Hart and C.~T.~H.~Davies,
358: %``The index theorem and universality properties of the low-lying eigenvalues
359: %of improved staggered quarks,''
360: arXiv:hep-lat/0406010.
361: %%CITATION = HEP-LAT 0406010;%%
362:
363:
364: \bibitem{dewitt} B. S. DeWitt, "Dynamical Theory of Groups and
365: Fields" (Gordon and Breach, New York, 1965); Phys. Rep.
366: \textbf{19}, 295 (1975).
367:
368: \bibitem{lee} C. Lee and C. Rim, Nucl. Phys. \textbf{B255},
369: 439(1985); I. Jack and H. Osborn, ibid. \textbf{B249}, 472 (1985);
370: R. D. Ball, Phys. Rep. \textbf{182}, 1 (1989).
371:
372: \bibitem{nsvz}
373: V.~A.~Novikov, M.~A.~Shifman, A.~I.~Vainshtein and V.~I.~Zakharov,
374: %``Calculations In External Fields In Quantum Chromodynamics:. Technical Review
375: %(Abstract Operator Method, Fock-Schwinger Gauge),''
376: Fortsch.\ Phys.\ {\bf 32}, 585 (1985).
377: %%CITATION = FPYKA,32,585;%%
378:
379: \bibitem{fliegner}
380: D.~Fliegner, P.~Haberl, M.~G.~Schmidt and C.~Schubert,
381: %``The higher derivative expansion of the effective action by the string
382: %inspired method. II,''
383: Ann. Phys. (NY) {\bf 264}, 51 (1998)
384: [arXiv:hep-th/9707189].
385: %%CITATION = HEP-TH 9707189;%%
386:
387:
388:
389:
390: \bibitem{lattice}
391: A.~W.~Thomas,
392: %``Chiral extrapolation of hadronic observables,''
393: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 119}, 50 (2003)
394: [arXiv:hep-lat/0208023];
395: C.~Bernard et al,
396: %``Panel discussion on chiral extrapolation of physical observables,''
397: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 119}, 170 (2003)
398: [arXiv:hep-lat/0209086].
399: %%CITATION = HEP-LAT 0209086;%%
400:
401: \bibitem{wkb}
402: G.~V.~Dunne, J.~Hur, C.~Lee and H.~Min, Phys. Lett. B {\bf 600}, 302 (2004),
403: %``Instanton determinant with arbitrary quark mass: WKB phase-shift method and
404: %derivative expansion,''
405: [arXiv:hep-th/0407222].
406: %%CITATION = HEP-TH 0407222;%%
407:
408:
409:
410: \bibitem{jackiw}
411: R.~Jackiw and C.~Rebbi,
412: %``Spinor Analysis Of Yang-Mills Theory,''
413: Phys.\ Rev.\ D {\bf 16}, 1052 (1977);
414: %%CITATION = PHRVA,D16,1052;%%
415: A.~D'Adda and P.~Di Vecchia,
416: %``Supersymmetry And Instantons,''
417: Phys.\ Lett.\ B {\bf 73}, 162 (1978).
418: %%CITATION = PHLTA,B73,162;%%
419:
420: \bibitem{brown}
421: F. R. Ore, Phys. Rev. {\bf D16}, 2577 (1977);
422: L. S. Brown and D. B. Creamer, Phys. Rev.
423: \textbf{D18}, 3695 (1978); E. Corrigan, P. Goddard, H. Osborn, and
424: S. Templeton, Nucl. Phys. \textbf{B159}, 469 (1979); B. Berg and M.
425: L\"{u}scher, Nucl. Phys. \textbf{B160}, 281 (1979);
426: %H. Osborn, Ann. Phys. \textbf{135}, 373 (1981);
427: C.~Lee, H.~W.~Lee and P.~Y.~Pac,
428: %``Calculation Of One Loop Instanton Determinants Using Propagators With
429: %Space-Time Dependent Mass,''
430: Nucl.\ Phys.\ B {\bf 201}, 429 (1982).
431: %%CITATION = NUPHA,B201,429;%%
432:
433:
434:
435:
436:
437: \bibitem{levit} S. Levit and U. Smilansky, Proc. Am. Math. Soc. {\bf 65}, 299 (1977).
438:
439: \bibitem{coleman} S. Coleman, {\it Aspects of Symmetry}, (Cambridge University Press, 1988).
440:
441: \bibitem{dreyfus} T. Dreyfus and H. Dym, Duke Math. J. {\bf 45}, 15 (1978).
442:
443: \bibitem{forman} R. Forman, Invent. Math. {\bf 88}, 447 (1987), Comm. Math. Phys. {\bf 147}, 485 (1992).
444:
445: \bibitem{kirsten}
446: K.~Kirsten and A.~J.~McKane,
447: %``Functional determinants by contour integration methods,''
448: Ann. Phys. (NY) {\bf 308}, 502 (2003),
449: %[arXiv:math-ph/0305010];
450: %%CITATION = MATH-PH 0305010;%%
451: %``Functional determinants for general Sturm-Liouville problems,''
452: J. Phys. A. {\bf 37}, 4649 (2004);
453: %[arXiv:math-ph/0403050];
454: %%CITATION = MATH-PH 0403050;%%
455: K. Kirsten, {\it Spectral Functions in Mathematics and Physics}, (Chapman-Hall, Boca Raton, 2002).
456:
457:
458: \bibitem{baacke}
459: J.~Baacke and G.~Lavrelashvili,
460: %``One-loop corrections to the metastable vacuum decay,''
461: Phys.\ Rev.\ D {\bf 69}, 025009 (2004)
462: [arXiv:hep-th/0307202].
463: %%CITATION = HEP-TH 0307202;%%
464:
465:
466:
467: \bibitem{bender} C. M. Bender and S. A. Orszag, {\it Advanced
468: Mathematical Methods for Scientists and Engineers}, (McGraw-Hill, New York, 1978).
469:
470: \bibitem{langer} R. E. Langer, Phys. Rev. \textbf{51}, 669 (1937).
471: %;J. B. Krieger and C. Rosenzweig, Phys. Rev. \textbf{164},
472: %171 (1967).
473:
474: \bibitem{inprep}
475: G.~V.~Dunne, J.~Hur, C.~Lee and H.~Min, in preparation.
476:
477:
478: \end{thebibliography}
479:
480: \end{document}
481: