1: %%%%%%%%%%%%%%%%%%%%%%%%
2: %\documentstyle[preprint,aps,prd,psfig,side]{revtex}
3: \documentstyle[preprint,aps,prd,psfig]{revtex}
4: %
5: % Start (1999.9.3)
6: % Version 1.0 (1999.9.19)
7: %
8: %
9: %
10: %
11:
12: \newcommand{\la}{\langle}
13: \newcommand{\ra}{\rangle}
14: \newcommand{\figwidth}{3.2in}
15: \newcommand{\figleftmarg}{-5mm}
16: \newcommand{\ttbs}{\char'134}
17: \newcommand{\AmS}{{\protect\the\textfont2
18: A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}
19: \newcommand{\be}{\begin{equation}}
20: \newcommand{\ee}{\end{equation}}
21: \newcommand{\ben}{\begin{eqnarray}}
22: \newcommand{\een}{\end{eqnarray}}
23: \newcommand{\nn}{\nonumber}
24: \newcommand{\tr}{{\rm Tr}}
25: \newcommand{\slas}[2]{{{#1}\hspace{-5pt}{/}}_{#2}}
26: \newcommand{\slal}[2]{{{#1}\hspace{-8pt}{/}}_{#2}}
27: \def\simgt{\rlap{\lower 3.5 pt\hbox{$\mathchar \sim$}}\raise 1pt \hbox {$>$}}
28: \def\simlt{\rlap{\lower 3.5 pt\hbox{$\mathchar \sim$}}\raise 1pt \hbox {$<$}}
29:
30: \newcommand{\mm}{{\cal M}}
31: \newcommand{\ce}{{\it c}_{\it E}}
32: \newcommand{\cb}{{\it c}_{\it B}}
33: \newcommand{\lqcd}{{\Lambda}_{\rm QCD}}
34:
35: \begin{document}
36: \draft
37:
38: \title{
39: \vspace{-5.5cm}
40: %\begin{flushright}
41: \hspace{120mm}{\normalsize KEK-CP-110}\\
42: \vspace{+8mm}
43: %\end{flushright}
44: Relativistic Heavy Quarks on the Lattice}
45:
46: \author{Sinya~Aoki$^{\rm a}$,
47: Yoshinobu~Kuramashi$^{\rm b}$ and
48: Shin-ichi~Tominaga$^{\rm c}$
49: }
50: \address{$^a$Institute of Physics,
51: University of Tsukuba,\\
52: Tsukuba, Ibaraki 305-8571, Japan\\
53: $^b$Institute of Particle and Nuclear Studies,
54: High Energy Accelerator Research Organization(KEK),\\
55: Tsukuba, Ibaraki 305-0801, Japan\\
56: $^c$Center for Computational Physics,
57: University of Tsukuba,\\
58: Tsukuba, Ibaraki 305-8577, Japan\\
59: }
60:
61: \date{\today}
62:
63: \maketitle
64:
65: \vspace{-7mm}
66:
67: %\newpage
68: \begin{abstract}
69: \vspace{-10mm}
70:
71: Lattice QCD should allow quantitative predictions for the
72: heavy quark physics from first principles.
73: Up to now, however, most approaches have based on the
74: nonrelativistic effective theory, with which
75: the continuum limit can not be taken in principle.
76: In this paper we investigate feasibility of relativistic approaches to
77: the heavy quark physics in lattice QCD.
78: We first examine validity of the idea that
79: the use of the anisotropic
80: lattice could be advantageous to control the $m_Q a$ corrections.
81: Our perturbative calculation, however, reveals that this is not true.
82: We instead propose a new relativistic approach to handle
83: heavy quarks on the isotropic lattice.
84: We explain how power corrections of $m_Q a$ can be avoided and
85: remaining uncertainties are reduced to be of order
86: $(a\Lambda_{\rm QCD})^2$.
87:
88:
89: \end{abstract}
90: \pacs{11.15.Ha, 12.38.Gc, 13.30.-a }
91:
92:
93:
94: \section{Introduction}
95:
96: Weak matrix elements associated with $B$ mesons are essential ingredients
97: to determine Cabibbo-Kobayashi-Maskawa matrix
98: elements. In principle lattice QCD provides the opportunity
99: of evaluating these matrix elements from first principles.
100: However it is still difficult to simulate the $b$ quark with
101: high precision on the lattice. The main source of systematic errors
102: originates from the fact that the $b$ quark mass in the
103: lattice unit is large: $m_b a\sim 1-2 $ in the quenched
104: approximation and $m_b a\sim 2-3 $ in full QCD with current
105: accessible computational resources. In order to control
106: large $m_Q a$ errors, several ways have been proposed so far:
107: A static approximation with $m_Q\rightarrow \infty$\cite{static},
108: a nonrelativistic QCD\cite{nrqcd}, a nonrelativistic interpretation
109: applied to results with the
110: Wilson/Sheikholeslami-Wholert (SW) action\cite{fermilab} and an anisotropic
111: lattice with finer temporal lattice spacing $a_t$ while keeping
112: the spatial one $a_s$ modest\cite{aniso}.
113: Although the $b$ quark can be directly simulated with any of the last
114: three approaches, only the last one has the advantage that we can take the
115: continuum limit, which is a fascinating feature stimulating our interest.
116:
117: Practical effectiveness of the anisotropic lattice is transparent:
118: with finer temporal lattice spacing
119: time evolutions of all kinds of correlation
120: functions become milder, which benefits us the better
121: signal-to-noise ratio.
122: On the theoretical side, our interest exists in the use of the
123: anisotropic lattice to control $m_Q a$ errors. If they
124: are restricted to only powers of $m_Q a_t$,
125: they can be made small by the anisotropic lattice with smaller $a_t$.
126: Indeed
127: $m_Q a_s$ corrections can be removed at the tree level\cite{symm_at}.
128: Our main concern, however, is whether
129: $m_Q a_s$ corrections could revive perturbatively or nonperturbatively.
130: Up to now no one has successfully
131: eliminated the possibility that $m_Q a_s$
132: corrections can appear beyond the tree level.
133: Without the proof given the anisotropic lattice is no better
134: than the isotropic one, where one can also
135: eliminate $m_Q a$ corrections at the tree level,
136: in terms of controlling $m_Q a$ corrections.
137:
138: The first part of this paper is devoted to a one-loop
139: calculation of the quark self energy on the anisotropic
140: lattice. We analyze its $O(g^2 a)$ terms to examine
141: the possibility of appearance of $O(g^2 m_Q a_s)$
142: contributions.
143: Our results strongly suggest that
144: one-loop radiative corrections allow the revival of
145: $m_Q a_s$ corrections.
146: Since we find that the anisotropic lattice is not theoretically advantageous
147: any more,
148: in the second part of this paper
149: we propose a new relativistic way to deal with the heavy quarks on the
150: isotropic lattice,
151: analyzing $m_Q a$ corrections carefully.
152: We discuss cutoff effects of the heavy quark system
153: following the on-shell improvement programme\cite{sym1,sym2}.
154: An important finding is that leading cutoff effects of
155: order $(m_Q a)^n$ can be absorbed
156: in the definition of renormalization factors for the
157: quark wave function and mass, so that the leading $m_Q a$ correction
158: is reduced to be of order $ (m_Q a)^n a \Lambda_{\rm QCD} $.
159: After removing remaining leading cutoff effects of
160: $O((m_Q a)^n a\Lambda_{\rm QCD})$ with parameters in the
161: quark action properly adjusted,
162: we are left with
163: only $O((a\Lambda_{\rm QCD})^2)$ errors, which are
164: expected to be fairly small.
165: We also provide a nonperturbative method to control the $O((m_Q a)^n)$
166: corrections involved in renormalization factors.
167:
168: This paper is organized as follows.
169: In Sec.~\ref{sec:aniso}
170: we present a one-loop calculation of the quark self energy
171: on the anisotropic lattice
172: and discuss the possibility of the
173: revival of $m_Q a_s$ corrections.
174: In Sec.~\ref{sec:iso} we propose a new relativistic
175: approach to handle the heavy quarks on the isotropic lattice
176: avoiding large $m_Q a$ corrections.
177: Our conclusions are summarized in Sec.~\ref{sec:conclusion}.
178:
179: \section{Anisotropic lattice}
180: \label{sec:aniso}
181:
182: \subsection{On-shell improvement on the anisotropic lattice}
183: \label{subsec:onshell}
184:
185: In order to obtain a generic form of the quark action
186: allowed on the anisotropic lattice, let us make the
187: operator analysis according to the Symanzik's
188: improvement programme\cite{sym1,sym2}.
189: The lattice theory is described by a local
190: effective theory as
191: \be
192: S_{\rm eff}=S_0+\sum_{k\ge 1,i} a^k \int d^4 x c_{4+k,i}(g){\cal O}_{4+k,i} (x),
193: \ee
194: where $S_0$ denotes the continuum action.
195: ${\cal O}_{k,i} (x)$ is a
196: local composite operator with $k$ dimensions, which consist of
197: quark mass, quark fields and link variables.
198: These higher dimensional operators must respect
199: symmetries on the lattice such as the invariance under
200: gauge, parity and charge-conjugation transformations and
201: discrete rotations.
202: The coefficient $c_{k,i}$
203: is a function of the gauge coupling $g$
204: to be determined perturbatively or nonperturbatively.
205:
206: Symanzik's improvement programme was originally
207: designed to reduce cutoff effects order by order for on-shell and
208: off-shell Green functions.
209: However, we are interested in
210: only on-shell quantities such as hadron masses and matrix elements
211: which require correlation functions at non-zero
212: physical distances.
213: Here it would be better to consider the on-shell improvement
214: procedure that is much simpler but restricted to
215: on-shell quantities\cite{onshell}.
216:
217: Under the requirement of various symmetries on the lattice,
218: we find the following set of operators with dimension up to
219: five:
220: \ben
221: {\rm dim.3:}&& {\cal O}_3^\prime(x)={\bar q}(x)q(x), \\
222: {\rm dim.4:}&& {\cal O}_{4a}^\prime(x)={\bar q}(x)\gamma_0 D_0 q(x),\\
223: && {\cal O}_{4b}^\prime(x)=
224: \sum_i {\bar q}(x)\gamma_i D_iq(x), \\
225: {\rm dim.5:}&& {\cal O}_{5a}^\prime(x)={\bar q}(x)D_0^2 q(x),\\
226: && {\cal O}_{5b}^\prime(x)=\sum_i {\bar q}(x) D_i^2 q(x), \\
227: && {\cal O}_{5c}^\prime(x)=
228: i\sum_i {\bar q}(x)\sigma_{0i}F_{0i} q(x),\\
229: && {\cal O}_{5d}^\prime(x)=
230: i\sum_{i,j} {\bar q}(x)\sigma_{ij}F_{ij} q(x),\\
231: && {\cal O}_{5e}^\prime(x)=
232: \sum_i {\bar q}(x) [\gamma_0 D_0,\gamma_i D_i] q(x),
233: \een
234: where $D_\mu$ is the covariant derivative on the lattice and
235: $\sigma_{\mu\nu}$ and $F_{\mu\nu}$ are defined as
236: $\sigma_{\mu\nu}\equiv[\gamma_\mu,\gamma_\nu]/2$
237: and $ig F_{\mu\nu}\equiv[D_\mu,D_\nu]$.
238: The subscript $0$ denotes the time component, while $i,j=1,2,3$
239: space components.
240: In terms of these operators, a general form of the quark
241: action on the anisotropic lattice is given by
242: \ben
243: S_q^{\rm imp}&=&\sum_x\left[
244: c_3^\prime{\cal O}_3^\prime(x)
245: +\sum_{i=a,b}c_{4i}^\prime{\cal O}_{4i}^\prime(x)
246: +\sum_{i=a,\dots ,e}c_{5i}^\prime{\cal O}_{5i}^\prime(x)
247: \right],
248: \label{eq:qaction_g}
249: \een
250: where $c_3^\prime,\dots,c_{5e}^\prime$ are functions of the bare gauge
251: coupling $g$, the bare quark mass $m_0$ and the time and
252: space lattice spacings $a_t$, $a_s$.
253: Since degrees of freedom of $c_3^\prime$ and $c_{4a}^\prime$ can be
254: absorbed in the renormalization of the quark mass and the
255: wave function respectively, we choose $c_3^\prime=m_0$ and
256: $c_{4a}^\prime=1$ for
257: convenience. We also find ${\cal O}_{5a}^\prime$ and
258: ${\cal O}_{5e}^\prime$ are related to
259: other operators by using the classical field
260: equation, and hence $c_{5a}^\prime$ and $c_{5e}^\prime$ can be
261: set by hand. We eliminate ${\cal O}_{5e}^\prime$ by choosing
262: $c_{5e}^\prime=0$ and employ ${\cal O}_{5a}^\prime$ to avoid
263: species doubling with $c_{5a}^\prime$ finite.
264: After all remaining parameters are $c_{4b}^\prime$,
265: $c_{5b}^\prime$, $c_{5c}^\prime$ and $c_{5d}^\prime$, which should be
266: tuned (nonperturbatively) in order to remove
267: $O(a_{t,s})$ discretization errors,
268: so that we are left with only discretization ambiguities of $O(a_{t,s}^2)$.
269: This point should be stressed, since previous
270: papers\cite{symm_at,pp} claim that only three parameters
271: $c_{4b}^\prime$, $c_{5c}^\prime$ and $c_{5d}^\prime$ are enough to be tuned
272: for the $O(a_{t,s})$ improvement.
273:
274:
275: \subsection{Quark and gauge actions}
276:
277: According to the discussion in the above subsection
278: a general form of the
279: quark action on the anisotropic lattice is given by:
280: \ben
281: S_q&=& a_t a_s^3 \sum_x {\bar q}(x)\left[ \gamma_0 D_0
282: +\nu \sum_i \gamma_i D_i +m_0
283: -\frac{a_t}{2}r ( D_0^2+\eta \sum_i D_i^2)\right.\nn\\
284: &&\;\;\;\;\left.-a_t \frac{ig}{4}r
285: \left\{ \ce(1+\eta)\sum_i \sigma_{0i} F_{0i}
286: +\cb \eta \sum_{ij} \sigma_{ij} F_{ij}
287: \right\}\right]q(x),
288: \label{eq:qaction_d}
289: \een
290: where parameters in eq.~(\ref{eq:qaction_g}) are rewritten as
291: \ben
292: c_{4b}^\prime&=& \nu,\\
293: c_{5a}^\prime&=& -\frac{r}{2},\\
294: c_{5b}^\prime&=& -\frac{r \eta}{2},\\
295: c_{5c}^\prime&=& -\frac{g}{4} r \ce (1+\eta),\\
296: c_{5d}^\prime&=& -\frac{g}{4} r \cb \eta.
297: \een
298: Here the Wilson parameter $r$ is taken arbitrary by
299: hand as mentioned in the previous subsection.
300: Note that the quark fields $q$, the covariant derivatives
301: $D_0$ and $D_i$, the bare quark mass $m_0$, the gauge field
302: strengths $F_{0i}$ and $F_{ij}$ are dimensionful.
303: They are transformed into
304: dimensionless quantities by
305: \ben
306: {\tilde q}&=&a_s^{3/2} q,\\
307: {\tilde D}_0&=& a_t D_0,\\
308: {\tilde D}_i&=& a_s D_i,\\
309: {\tilde m}_0&=& a_t m_0,\\
310: {\tilde F}_{0i}&=& a_t a_s F_{0i},\\
311: {\tilde F}_{ij}&=& a_s^2 F_{ij}.
312: \een
313: With this transformation, eq.~(\ref{eq:qaction_d}) is
314: rewritten as
315: \ben
316: S_q(x)&=& \sum_x {\bar {\tilde q}}(x)\left[ \gamma_0 {\tilde D}_0
317: +\frac{\nu}{\xi}\sum_i \gamma_i {\tilde D}_i +{\tilde m}_0
318: -\frac{r}{2}( {\tilde D}_0^2
319: +\frac{\eta}{\xi^2} \sum_i {\tilde D}_i^2)\right.\nn\\
320: &&\;\;\;\;\left.- \frac{ig}{4}r
321: \left\{ \ce \frac{(1+\eta)}{\xi}\sum_i \sigma_{0i} {\tilde F}_{0i}
322: +\cb \frac{\eta}{\xi^2}
323: \sum_{ij} \sigma_{ij} {\tilde F}_{ij} \right\}\right]
324: {\tilde q}(x)
325: \label{eq:qaction}
326: \een
327: with $\xi=a_s/a_t$ the anisotropy parameter.
328:
329: From eq.~(\ref{eq:qaction_d})
330: we find
331: the inverse free quark propagator
332: \ben
333: a_t S_q^{-1}(p)&=&i \gamma_0 {\rm sin}(p_0 a_t)
334: +\frac{\nu}{\xi}
335: i \sum_i \gamma_i {\rm sin}(p_i a_s) +m_0 a_t\nn\\
336: &&+r (1-{\rm cos}(p_0 a_t))+\frac{r\eta}{\xi^2}
337: \sum_i(1-{\rm cos}(p_i a_s)),
338: \label{eq:qp_tree}
339: \een
340: and relevant vertices for the present calculation
341: \ben
342: V_{10}^{A}(q,p)&=&-g T^A\left\{i \gamma_0
343: {\rm cos}\left(\frac{p_0 a_t + q_0 a_t}{2}\right)
344: +r {\rm sin}\left(\frac{p_0 a_t + q_0 a_t}{2}\right)\right\},\\
345: V_{1i}^{A}(q,p)&=&-g T^A\left\{\nu i \gamma_i
346: {\rm cos}\left(\frac{p_i a_s + q_i a_s}{2}\right)
347: +\frac{r\eta}{\xi}
348: {\rm sin}\left(\frac{p_i a_s + q_i a_s}{2}\right)\right\},\\
349: V_{200}^{AB}(q,p)&=&\frac{a_s}{2\xi} g^2 \frac{1}{2}
350: \left\{T^A,T^B\right\}\left\{i \gamma_0
351: {\rm sin}\left(\frac{p_0 a_t + q_0 a_t}{2}\right)
352: -r {\rm cos}\left(\frac{p_0 a_t + q_0 a_t}{2}\right)\right\},\\
353: V_{2ii}^{AB}(q,p)&=&\frac{a_s}{2} g^2 \frac{1}{2}
354: \left\{T^A,T^B\right\} \left\{\nu i \gamma_i
355: {\rm sin}\left(\frac{p_i a_s + q_i a_s}{2}\right)
356: -\frac{r\eta}{\xi}
357: {\rm cos}\left(\frac{p_i a_s + q_i a_s}{2}\right)\right\},\\
358: V_{c0}^{A}(q,p)&=&-\ce g T^A \frac{r(1+\eta)}{4\xi}
359: \sum_i \sigma_{0i}{\rm sin}\left(p_i a_s - q_i a_s \right)
360: {\rm cos}\left(\frac{p_0 a_t - q_0 a_t}{2}\right),\\
361: V_{ci}^{A}(q,p)&=&-\ce g T^A \frac{r(1+\eta)}{4}
362: \sigma_{i0}{\rm sin}\left(p_0 a_t - q_0 a_t \right)
363: {\rm cos}\left(\frac{p_i a_s - q_i a_s}{2}\right)\nn\\
364: &&-\cb g T^A \frac{r \eta}{2\xi}
365: \sum_{j} \sigma_{ij}{\rm sin}\left(p_j a_s - q_j a_s \right)
366: {\rm cos}\left(\frac{p_i a_s - q_i a_s}{2}\right),
367: \een
368: where $p$ is for the incoming momentum into the vertex and
369: $q$ for the outgoing momentum. $T^A$ $(A=1,\dots,N_c^2-1)$
370: is a generator of color SU($N_c$).
371:
372: For the gauge part we take the standard Wilson action on the
373: anisotropic lattice:
374: \ben
375: S_g&=&\frac{2 N_c}{g^2}\sum_n \left[
376: \xi \sum_{i}\left(1-\frac{1}{2N_c} {\rm Tr}(U_{0i}(n)
377: +U_{0i}^\dagger(n))\right)\right. \nn\\
378: &&\;\;\;\;\left. +\frac{1}{\xi} \sum_{i<j}\left(1-
379: \frac{1}{2N_c} {\rm Tr}(U_{ij}(n)+U_{ij}^\dagger(n))\right)\right]
380: \een
381: with
382: \be
383: U_{\mu\nu}(n)=U_{\mu}(n)
384: U_{\nu}(n+{\hat \mu})U_{\mu}^\dagger(n+{\hat \nu})
385: U_{\nu}^\dagger(n).
386: \ee
387: The gluon propagator is given by
388: \be
389: G_{\mu\nu}^{AB}(k)=a_s^2\frac{\delta_{\mu\nu}\delta_{AB}}
390: {\xi^2 4 {\rm sin}^2\left(\frac{k_0 a_t}{2}\right)
391: +4\sum_i {\rm sin}^2\left(\frac{k_i a_s}{2}\right)
392: +\lambda^2a_s^2}
393: \ee
394: in the Feynman gauge with the fictitious mass $\lambda^2$,
395: which is introduced to work as the infrared cutoff.
396:
397: \subsection{Tree-level analysis on the quark propagator}
398:
399: Expanding the inverse free quark propagator in
400: eq.~(\ref{eq:qp_tree}) up to $O(a_t)$ we obtain
401: \be
402: S_q^{-1}(p)=i \gamma_0 p_0+\nu i \sum_i \gamma_i p_i + m_0
403: +\frac{r a_t}{2} p_0^2 + \frac{r\eta a_t}{2} \sum_i p_i^2
404: +O((p_0 a_t)^2,(p_i a_s)^2) ,
405: \ee
406: which yields the following expression for the quark propagator:
407: \be
408: S_q(p)=\frac{1}{1+m_0 r a_t}
409: \frac{-i \gamma_0 p_0-\nu i \sum_i \gamma_i p_i + m_0
410: +\frac{r a_t}{2}\left(
411: p_0^2 + \eta \sum_i p_i^2
412: \right)}
413: {p_0^2+\frac{\nu^2+m_0 r\eta a_t}{1+m_0 r a_t}\sum_i
414: p_i^2 + \frac{1}{1+m_0 r a_t}m_0^2 } + O(a_{t,s}^2) .
415: \label{eq:qprop}
416: \ee
417: We consider
418: the tree-level on-shell improvement for this quark
419: propagator requiring that
420: $S_q(p)$ should reproduce the form
421: \be
422: S_q(p)=
423: \frac{1}{Z_q}
424: \frac{ -i \gamma_0 p_0- i \sum_i \gamma_i p_i + m_R}
425: {p_0^2+\sum_i p_i^2 +m_R^2}
426: +{\rm (no\; pole\; terms)}+O(a_{t,s}^2)
427: \label{eq:qprop_r}
428: \ee
429: with the appropriate choice for $Z_q$, $Z_m$,
430: $\nu$ and $\eta$, where
431: $Z_q$ and $Z_m$ denote renormalization factors
432: for the quark wave function and mass defined by
433: \ben
434: q_R&=&Z_q^{1/2} q,\\
435: m_R&=&Z_m m_0,
436: \een
437: where $m_R$ is the pole mass.
438: It should be reminded that $r$ is a free parameter.
439: We remark that terms without the pole in the quark
440: propagator of eq.~(\ref{eq:qprop_r}) yield only contact
441: terms in the configuration space, which do not contribute to
442: on-shell quantities in Green functions.
443: In terms of the inverse quark propagator,
444: the condition eq.~(\ref{eq:qprop_r}) is equivalent to
445: \ben
446: S^{-1}_q(p) &=& \left[ Z_q - 2 C m_0 a_t \right]( i \slas{p}{} + m_R)
447: + C a_t (p^2+m_R^2) +O( a_t(i\slas{p}{} + m_R)^2) + O(a_t^2)
448: \een
449: with $C$ constant.
450: Therefore ``(no pole terms)'' in eq.~(\ref{eq:qprop_r})
451: are not necessary to be $O(a_t^2)$.
452: Comparing the expressions of
453: eqs.~(\ref{eq:qprop}) and (\ref{eq:qprop_r}), we find at the tree level
454: \ben
455: Z_q^{-1/2}&=&1-\frac{r}{2}m_0 a_t,\\
456: Z_m&=&1-\frac{r}{2}m_0 a_t,\\
457: \nu&=&1,\\
458: \eta&=&1.
459: \label{eq:condition}
460: \een
461:
462: Up to now three types of quark actions with different choices for
463: $r$ and $\eta$ have been proposed:
464: (i) $r=1$ and $\eta=1$\cite{symm_at},
465: (ii) $r=\xi$ and $\eta=1$\cite{symm_as} and
466: (iii) $r=1$ and $\eta=\xi$\cite{pp}.
467: Although the action in the case of (iii) has been most extensively studied
468: numerically, the choice of the parameter $\eta$ does not
469: meet the condition of eq.~(\ref{eq:condition}) that is
470: required from the on-shell improvement at the tree-level.
471: This primitive failure makes us consider that it is not worthwhile to
472: work on the case (iii) in this paper.
473: We focus on only cases (i) and (ii) hereafter.
474:
475:
476: Let us first derive the relation between the bare quark mass $m_0$ and
477: the pole mass $m_p$.
478: Putting $p_i=0$ and $p_0=i m_p$
479: into the inverse free quark propagator of eq.~(\ref{eq:qp_tree}),
480: the on-shell condition yields
481: \be
482: m_p a_t = {\rm log}\left| \frac{m_0 a_t +r
483: +\sqrt{(m_0 a_t)^2+2 r m_0 a_t +1}}{1+r} \right|.
484: \ee
485: While in the case (i) with $r=1$ we can expand
486: $m_p a_t$ in powers of
487: $m_0$ under the condition $m_0 a_t \ll 1$,
488: in the case (ii) with $r=\xi$ the condition $\xi m_0 a_t=m_0 a_s\ll 1$
489: is necessary. To avoid any confusions we assume
490: $m_0 a_s\ll 1$ from now on. We remark that this
491: assumption does not affect any conclusions in this section.
492:
493: On the anisotropic lattice we have to be careful about
494: contributions of space doublers.
495: Pole masses of space doublers are written as
496: \be
497: m_p^d a_t = {\rm log}\left| \frac{m_0^d a_t +r
498: +\sqrt{(m_0^d a_t)^2+2 r m_0^d a_t +1}}{1+r} \right|
499: \ee
500: with
501: \be
502: m_0^d=m_0+\frac{2 r }{\xi^2} \frac{N_d}{a_t},
503: \ee
504: where $N_d$ components of spatial momentum
505: $p_i$ are equal to $\pi/a_s$ at the edge of the Brillouin
506: zone. Although doubler pole masses are always heavier than
507: the physical one irrespective of the value of $m_0$, $r$ and
508: $\xi$, their differences in the large limit of $\xi$ are
509: given by
510: \ben
511: (m_p^d-m_p)a_s &\rightarrow&
512: \frac{2}{\xi} N_d \frac{1}{1+m_0 a_t}+O\left(\frac{1}{\xi^2}\right)
513: \;\;\;\; {\rm case\; (i)},\\
514: (m_p^d-m_p)a_s &\rightarrow&
515: \sqrt{1+2 m_0 a_s+4 N_d}-\sqrt{1+2 m_0 a_s}+O\left(\frac{1}{\xi}\right)
516: \;\;\;\; {\rm case\; (ii)}.
517: \een
518: For the case (i) we find the gap $(m_p^d-m_p)a_s$ diminishes as
519: $\xi$ becomes larger. This brings a
520: practical problem in numerical studies of heavy quarks:
521: contributions of doublers could contaminate
522: signals of hadron states.
523: On the other hand, we are free from this problem
524: in the case (ii).
525:
526:
527: \subsection{One-loop quark self energy}
528:
529: The inverse full quark propagator is written as
530: \be
531: S_q^{-1}(p)=i \gamma_0 p_0+\nu i \sum_i \gamma_i p_i + m_0
532: +\frac{a}{2} p_0^2 + \eta \frac{a}{2} \sum_i p_i^2-\Sigma(p,m_0),
533: \label{eq:qp_inv}
534: \ee
535: where we take $a=r a_t$(;$a=a_t$ for (i) while $a=a_s$ for (ii)).
536: One-loop contributions to the quark self-energy
537: $\Sigma(p,m_0)$ consist of two types of diagrams depicted
538: in Figs.~\ref{fig:qse} (a) and (b), which are expressed by
539: \ben
540: \Sigma_a(p,m)&=&\int \frac{d^4 k}{(2\pi)^4}\sum_A\sum_{\mu}
541: \left\{
542: V_{1\mu}^A(p,p+k) S_q(p+k) V_{1\mu}^A(p+k,p)\right. \nn\\
543: &&\;\;\;\;\left.+V_{c\mu}^A(p,p+k) S_q(p+k) V_{1\mu}^A(p+k,p)\right. \nn\\
544: &&\;\;\;\;\left.+V_{1\mu}^A(p,p+k) S_q(p+k) V_{c\mu}^A(p+k,p)\right. \nn\\
545: &&\;\;\;\;\left.+V_{c\mu}^A(p,p+k) S_q(p+k) V_{c\mu}^A(p+k,p)
546: \right\}G_{\mu\mu}^{AA}(k)
547: \een
548: and
549: \be
550: \Sigma_b(p)=\int \frac{d^4 k}{(2\pi)^4}\sum_A\sum_{\mu}
551: V_{2\mu\mu}^{AA}(p,p) G_{\mu\mu}^{AA}(k)
552: \ee
553: with
554: \ben
555: &&-\frac{\pi}{a_t}\le k_0 \le \frac{\pi}{a_t},\\
556: &&-\frac{\pi}{a_s}\le k_i \le \frac{\pi}{a_s}.
557: \een
558: Expanding $\Sigma(p,m_0)=\Sigma_a(p,m_0)+\Sigma_b(p)$ in terms
559: of $p$ and $m_0$, we obtain
560: \ben
561: \Sigma(p,m_0)&=&\frac{g^2}{16\pi^2} C_F\left[ \frac{\Sigma_0}{a}
562: +i \gamma_0 p_0(-L+\Sigma_1^t)+i \sum_i \gamma_i p_i(-L+\Sigma_1^s)
563: +m_0(-4L+\Sigma_2)\right. \nn\\
564: &&\;\;\;\;\left. +ap_0^2((1-3c_{\rm SW})L/2+\sigma_1^t)
565: +a\sum_i p_i^2((1-3c_{\rm SW})L/2+\sigma_1^s)\right.\nn\\
566: &&\;\;\;\;\left. +am_0 i \gamma_0 p_0((5+3c_{\rm SW})L/2+\sigma_2^t)
567: +am_0 i \sum_i \gamma_i p_i((5+3c_{\rm SW})L/2+\sigma_2^s)\right.\nn\\
568: &&\;\;\;\;\left. +am_0^2((5-3c_{\rm SW})L+\sigma_3)\right]+O(a_{t,s}^2)
569: \label{eq:qse}
570: \een
571: with $L=-{\rm log}(\lambda^2 a_s^2)$
572: the contribution of the infrared divergence.
573: Here we take $c_B = c_E \equiv c_{\rm SW}$ for the clover coefficient.
574: $\Sigma_0$, $\Sigma_1^{t,s}$, $\Sigma_2$ and
575: $\sigma_1^{t,s}$, $\sigma_2^{t,s}$, $\sigma_3$ are independent of
576: $a$ but functions of $\xi$, $\eta$ and $r$ parameters.
577: We evaluate these quantities numerically using the Monte
578: Carlo integration routine BASES\cite{bases}.
579:
580:
581:
582: From eqs.~(\ref{eq:qp_inv}) and (\ref{eq:qse}) we obtain
583: \ben
584: Z_q^{-1/2}&=&\left\{1+\frac{g^2}{16\pi^2}
585: \frac{C_F}{2}(-L+\Delta_q^{(0)})\right\}
586: \left\{1+a_t m\left(-\frac{r}{2}+\frac{g^2}{16\pi^2}C_F\Delta_q^{(1)}
587: \right)\right\},\\
588: Z_m&=&\left\{1+\frac{g^2}{16\pi^2} C_F(3L+\Delta_m^{(0)}\right\}
589: \left\{1+a_t m\left(-\frac{r}{2}+\frac{g^2}{16\pi^2} C_F\Delta_m^{(1)}
590: \right)\right\},\\
591: \nu&=&1-\frac{g^2}{16\pi^2} C_F\Delta_\nu^{(0)}
592: -\frac{g^2}{16\pi^2} C_F a_t m \Delta_\nu^{(1)},\\
593: \eta&=&1-2 \frac{g^2}{16\pi^2} C_F \Delta_\eta^{(0)},
594: \een
595: where
596: \ben
597: m&=&m_0-\frac{g^2}{16\pi^2}C_F\frac{\Sigma_0}{a},\\
598: \Delta_q^{(0)}&=&\Sigma_1^t,\\
599: \Delta_q^{(1)}&=&r\left(\sigma_1^t+
600: \frac{\sigma_2^t}{2}-\Sigma_1^t+\frac{\Sigma_2}{2}
601: +\frac{3(1-c_{\rm SW})}{4}L
602: \right),\\
603: \Delta_m^{(0)}&=&\Sigma_1^t-\Sigma_2,\\
604: \Delta_m^{(1)}&=&r\left(\sigma_1^t+
605: \sigma_2^t-\sigma_3-\Sigma_1^t+\frac{\Sigma_2}{2}
606: +3(c_{\rm SW}-1)L
607: \right),\\
608: \Delta_\nu^{(0)}&=&\Sigma_1^t-\Sigma_1^s,\\
609: \Delta_\nu^{(1)}&=&r \left(\sigma_2^t-\sigma_2^s\right),\\
610: \Delta_\eta^{(0)}&=&\sigma_1^t-\sigma_1^s.
611: \een
612: We find for the anisotropic case that
613: the $g^2a {\log} a$ terms disappear in $Z_q$, $Z_m$ for $c_{\rm SW}$ =1
614: as well as $\nu$ and $\eta$ for an arbitrary values of $c_{\rm SW}$,
615: as observed in $Z_q$ and $Z_m$ for the isotropic case\cite{sw_pt}.
616: Thus $c_{\rm SW}=1$ gives the tree level estimate for $c_B$ and $c_E$,
617: and we take this value for the latter numerical calculation in this section.
618: With this choice for $Z_q$, $Z_m$, $\nu$ and $\eta$ the
619: quark propagator is given by
620: \ben
621: S_q(p)&=&\frac{Z_q^{-1}}{p_0^2+\sum_i p_i^2 +m_R^2}
622: \left[-i \gamma_0 p_0 -i \sum_i \gamma_i p_i + m_R\right.\nn\\
623: &&\;\;\;\;\left. +a(p_0^2+\sum_i p_i^2 +m_R^2)\left\{\frac{1}{2}
624: -\frac{g^2}{16\pi^2}C_F
625: \left(-\frac{L}{2}+\sigma_1^t-\frac{\Sigma_1^t}{2}\right)\right\}\right]\\
626: &=&Z_q^{-1} S_q^R(p)+Z_q^{-1}a\left\{\frac{1}{2}
627: -\frac{g^2}{16\pi^2}C_F
628: \left(-\frac{L}{2}+\sigma_1^t-\frac{\Sigma_1^t}{2}\right)\right\},
629: \een
630: where $S_q^R(p)$ is the renormalized quark propagator.
631: We again remark that the term without the pole in the quark
632: propagator does not contribute to
633: on-shell quantities in Green functions.
634:
635: We show $\xi$ dependences of one-loop coefficients
636: for $Z_q$, $Z_m$, $\nu$ and $\eta$ in Figs.~\ref{fig:zwf},
637: \ref{fig:zm}, \ref{fig:nu} and \ref{fig:eta}, respectively.
638: Here we consider both $r=1$ and $r=\xi$ cases with $\eta=1$, which satisfy
639: the tree-level on-shell condition in eq.~(\ref{eq:condition}).
640: Although our main concern is whether $m a_s$ corrections
641: could revive at the one-loop level
642: for the $r=1$ case, we also present the results of the
643: $r=\xi$ case for comparison.
644: Results for $\Delta_q^{(0)}$, $\Delta_q^{(1)}$, $\Delta_m^{(0)}$ and
645: $\Delta_m^{(1)}$ in the isotropic case are already given in
646: Refs.~\cite{pt_g2,pt_g2a}.
647: They show an agreement with our results with the choice of $\xi=1$.
648: For the $r=1$ case, Figs.~\ref{fig:zm}(b),
649: \ref{fig:nu}(b) and \ref{fig:eta} show
650: approximate linear dependences on $\xi$ for
651: $\Delta_m^{(1)}$, $\Delta_\nu^{(1)}$ and
652: $\Delta_\eta^{(0)}$, which tells us that
653: $O(g^2 a)$ contributions to $Z_m$, $\nu$ and $\eta$
654: are effectively of order $g^2 m
655: a_s=g^2 m a_t\xi$.
656: We observe similar linear dependences for the $r=\xi$ case
657: in Figs.~\ref{fig:zwf}(b),
658: \ref{fig:zm}(b) and \ref{fig:eta}.
659: From these observations we conclude that $m a_s$
660: corrections are allowed to revive at the one-loop level.
661:
662: This is a reasonable conclusion in view of the on-shell
663: improvement. As far as we know there is no symmetry on the
664: anisotropic lattice to prohibit the higher dimensional operators
665: with the form of $(m a_s)^n {\cal O}_{4+k}$ ($n,k\ge 1$),
666: where ${\cal O}_{4+k}$ denotes $4+k$ dimensional operators.
667: Unless such symmetry is uncovered, the theoretical advantage of
668: the anisotropic lattice over the isotropic one would
669: be never confirmed.
670:
671:
672:
673: \section{Isotropic lattice}
674: \label{sec:iso}
675:
676: In this section we propose a new relativistic approach to
677: control $m_Q a$ corrections for the heavy quarks on the
678: lattice. The idea is based on the on-shell improvement
679: programme applied to the heavy quarks
680: on the isotropic lattice.
681: This method allows us to obtain the physical quantities
682: in the continuum limit
683: without requiring harsh condition $m_Q a \ll 1$ that is not
684: achievable in near future.
685:
686: Let us consider the general cutoff effects for the heavy quarks on
687: the lattice.
688: Here we assume that the heavy quark mass $m_Q$ is much
689: heavier than $\Lambda_{\rm QCD}$, while the light quark mass $m_q$
690: is lighter than $\Lambda_{\rm QCD}$.
691: Under this condition we assume that the leading cutoff
692: effects are
693: \be
694: f_0(m_Q a)>f_1(m_Q a)a\lqcd>f_2(m_Q a)(a\lqcd)^2>\cdots,
695: \ee
696: where $f_i(m_Q a)$ ($i\geq 0$) are smooth and continuous
697: all over the range of
698: $m_Q a$ and have Taylor expansions at $m_Q a=0$ with
699: sufficiently large convergence radii beyond $m_Q a=1$.
700: To control the scaling
701: violation effects we want to remove the cutoff effects up to
702: $f_1(m_Q a)a\lqcd$ by adding the counter terms
703: to the lattice quark action with the on-shell improvement.
704: If $m_Q a$ is small enough,
705: the remaining $f_2(m_Q a)(a\lqcd)^2$ contributions can be removed by
706: extrapolating the numerical data at several lattice spacings
707: to the continuum limit. Otherwise, in case of
708: sufficiently small lattice spacing,
709: the $O((a\Lambda_{\rm QCD})^2)$ errors can be neglected.
710: Our aim in this section is to search for the relevant counter terms
711: required in the on-shell improvement and propose a
712: nonperturbative method to determine their coefficients.
713: We also show that the behavior of $f_i(m_Q a)$ ($i\geq 1$) in
714: the large $m_Q a$ region can be discussed by investigating the
715: static limit.
716:
717:
718: \subsection{On-shell improvement on the isotropic lattice}
719: \label{subsec:onshell_iso}
720:
721: We first list the allowed operators under the
722: requirement of the gauge, axis permutation and other various
723: discrete symmetries on the lattice, where
724: the chiral symmetry is not imposed.
725: According to the work of Ref.~\cite{sw}, all the operators
726: with dimension up to six are given by
727: \ben
728: {\rm dim.3:}&& {\cal O}_3(x)={\bar q}(x)q(x), \\
729: {\rm dim.4:}&& {\cal O}_{4}(x)={\bar q}(x)\slal{D}{} q(x),\\
730: {\rm dim.5:}&& {\cal O}_{5a}(x)={\bar q}(x)D_\mu^2 q(x),\\
731: && {\cal O}_{5b}(x)=i{\bar q}(x)
732: \sigma_{\mu\nu}F_{\mu\nu} q(x),\\
733: {\rm dim.6:}&& {\cal O}_{6a}(x)={\bar q}(x)\gamma_\mu D_\mu^3 q(x),\\
734: && {\cal O}_{6b}(x)={\bar q}(x)D_\mu^2 \slal{D}{}
735: q(x),\\
736: && {\cal O}_{6c}(x)={\bar q}(x)\slal{D}{}
737: D_\mu^2 q(x),\\
738: && {\cal O}_{6d}(x)=i{\bar q}(x)\gamma_\mu[D_\nu,
739: F_{\mu\nu}] q(x),\\
740: && {\cal O}_{6e}(x)={\bar q}(x){\slal{D}{}}^{\,3} q(x),\\
741: && {\cal O}_{6f}(x)={\bar q}(x)\Gamma q(x)
742: {\bar q}(x)\Gamma q(x),
743: \een
744: where $\Gamma=1,\gamma_5,\gamma_\mu,\gamma_\mu\gamma_5,\sigma_{\mu\nu}$.
745: The higher dimensional operators are
746: related to the lower dimensional operators with additional
747: $(m_Q a)^n$ corrections with and without classical field
748: equations, or otherwise they have the contributions
749: of order $(a\Lambda_{\rm QCD})^2$ or less as the cutoff effects.
750:
751:
752: These operators lead to a following generic form of the quark
753: action on the isotropic lattice:
754: \be
755: S_q^{\rm imp}=\sum_x\left[
756: c_3{\cal O}_3(x)+c_{4}{\cal O}_{4}(x)
757: +\sum_{i=a,b} c_{5i}{\cal O}_{5i}(x)
758: +\sum_{i=a,\dots,f} c_{6i}{\cal O}_{6i}(x)
759: \right],
760: \label{eq:qimp_iso}
761: \ee
762: where $c_3,\dots,c_{6f}$ are functions of the bare gauge
763: coupling $g$ and the power corrections of $ma$.
764: We first remark that the $m a$ corrections to the quark mass term
765: and the kinetic term can be absorbed
766: in the renormalizations of the quark mass $Z_m$ and the
767: wave function $Z_q$.
768: For the sake of convenience we choose $c_3=m_0$ and $c_4=1$.
769:
770: In the next step we reduce the number of basis
771: operators with the aid of the classical field equations.
772: It is easily found that ${\cal O}_{5a}$, ${\cal O}_{6b}$,
773: ${\cal O}_{6c}$ and ${\cal O}_{6e}$ can be
774: related to the quark mass term
775: or the kinetic term. In the on-shell improvement these
776: operators are redundant and can be eliminated
777: from the action of eq.~(\ref{eq:qimp_iso}).
778: The operator ${\cal O}_{5a}$, however,
779: is used to avoid the species doubling and the value of its
780: coefficient $c_{5a}$ is given by hand.
781:
782: The remaining operators are
783: ${\cal O}_{5b}$, ${\cal O}_{6a}$, ${\cal O}_{6d}$ and
784: ${\cal O}_{6f}$,
785: whose contributions as the cutoff effects
786: are estimated in Table~\ref{tab:action}.
787: We use the classical field equation for
788: ${\bar q}(x)\gamma_0 D_0^3 q(x)$ in ${\cal O}_{6a}$.
789: Here it should be noted that $\lqcd$ means
790: the order $\Lambda_{\rm QCD}$ or less throughout this paper.
791: In some cases the actual contribution may become smaller.
792: For example,
793: the contributions of the bilinear terms
794: whose Dirac matrices consist of off-diagonal components
795: are suppressed by the extra $\Lambda_{\rm QCD}/m_Q$
796: for the heavy quarks compared to the light quarks.
797: The operator ${\cal O}_{5b}$ is the so-called clover term,
798: for which
799: the nonperturbative method to determine the coefficient
800: $c_{5b}$ in the massless limit is already
801: established\cite{impr}.
802: However, the contributions of $(ma)^n{\cal O}_{5b}$
803: ($n\geq 1$) cannot be neglected in the
804: present condition that allows $m_Q a\sim O(1)$.
805: For $O(a\Lambda_{\rm QCD})$ improvement
806: the coefficient $c_{5b}$ has to be adjusted in the mass dependent way.
807: The differences in magnitude between the time and space components in
808: ${\cal O}_{6a}$
809: originate from the violation of rotational
810: symmetry on the lattice with finite lattice spacing.
811: While the contributions of the space components
812: are found to be negligible,
813: those of the time components should be removed.
814: We also find the contributions of the four-quark operators
815: in ${\cal O}_{6f}$ are negligible.
816:
817: The generalization of
818: the above argument to any operators with higher
819: dimensions makes the discussion more transparent.
820: Let us consider an arbitrary operator with
821: $4+k$ dimension,
822: $a^k {\cal O}_{4+k}$, where we write the lattice spacing $a$ explicitly.
823: The operator ${\cal O}_{4+k}$ contains $l$ pairs of $\bar q$ and $q$ and
824: $n$ covariant derivatives $D_\mu$ with $4+ k = 3 \times l + n$.
825: Using the classical field equation,
826: some (but not all) of covariant derivatives
827: can be replaced by the quark mass $m$. For $l\ge 2$
828: the largest possible power of the scaling violation is
829: $ (m a )^n (a\Lambda_{\rm QCD})^{3l-4}$.
830: Therefore the operators which contain four or more quarks
831: are irrelevant for the $O(a\Lambda_{\rm QCD})$ improvement.
832: All the relevant contributions come from the quark bilinear operators.
833: With the aid of the classical field equations, they can be
834: reduced to
835: \ben
836: && (m a)^n a^{-1} {\bar q}(x)q(x) \\
837: && (m a)^{n-1} {\bar q}(x)\gamma_0 D_0 q(x),\;\;
838: (m a)^{n-1} \sum_i {\bar q}(x)\gamma_i D_i q(x)\\
839: && (m a)^{n-2}a {\bar q}(x) D_0^2 q(x),\;\;
840: (m a)^{n-2}a \sum_i {\bar q}(x) D_i^2 q(x)\\
841: && (m a)^{n-2} a i\sum_i {\bar q}(x)\sigma_{0i}F_{0i} q(x),\;\;
842: (m a)^{n-2} a i\sum_{ij} {\bar q}(x)\sigma_{ij}F_{ij} q(x),
843: \een
844: for $n \geq 0$. The time and space components of
845: ${\cal O}_{4}$ and ${\cal O}_{5a,5b}$ should be treated separately
846: in case of finite $ma$, where
847: the space-time asymmetry reflects the contributions of the higher
848: dimensional operators that break the rotational symmetry.
849: Now we know that the seven operators are
850: needed for the $O(a\lqcd)$ improvement.
851: Since three coefficients among these seven operators
852: can be absorbed in $Z_m$, $Z_q$ and the Wilson parameter $r_t$
853: for the time derivative as already explained,
854: the remaining four coefficients have to be actually tuned.
855:
856: In conclusion, at all order of $m a$,
857: the generic quark action is
858: written as
859: \ben
860: S_q^{\rm imp}&=&\sum_x\left[ m_0{\bar q}(x)q(x)
861: +{\bar q}(x)\gamma_0 D_0q(x)
862: +\nu \sum_i {\bar q}(x)\gamma_i D_i q(x)
863: -\frac{r_t a}{2} {\bar q}(x)D_0^2 q(x)\right.\nn\\
864: &&\;\;\;\;\;\;\;\;\left.-\frac{r_s a}{2} \sum_i {\bar q}(x)D_i^2 q(x)
865: -\frac{ig a}{2}c_E \sum_i {\bar q}(x)\sigma_{0i}F_{0i} q(x)
866: -\frac{ig a}{4}c_B \sum_{i,j} {\bar q}(x)\sigma_{ij}F_{ij} q(x)
867: \right],
868: \label{eq:qaction_iso}
869: \een
870: where we are allowed to choose $r_t=1$ and the four parameters
871: $\nu$, $r_s$, $c_E$ and $c_B$ are to be adjusted.
872: In general these parameters have the form that
873: $X=\sum_n X_n(g^2) (m a)^n$
874: with $X=\nu$, $r_s$, $c_E$ and $c_B$, and
875: $X_0$ should agree with the one in the massless $O(a)$ improved
876: theory: $\nu_0=1$, $(r_s)_0=r_t=1$, $(c_E)_0=(c_B)_0 = c_{\rm SW}$\cite{impr}.
877: Note that $\nu = 1 + O( (ma)^2)$ and $r_s = r_t + O( ma)$ since
878: the space-time asymmetry arises from Lorentz non-covariant terms such as
879: ${\cal O}_{6a}$ via the on-shell reduction,
880: accompanied by extra $(ma)^2$ factors.
881:
882:
883: {}From the above consideration,
884: the leading scaling violation in the massive theory,
885: except for
886: $\sum_{n=1}^\infty C_n^{q,m}(g^2, \log a) (m_Q a)^n$ in $Z_q$ and $Z_m$,
887: is $\sum_{n=0}^\infty C_n^{\rm W}(g^2, \log a) (m_Q a)^n a\Lambda_{\rm QCD} $
888: for the Wilson quark action, or
889: $\sum_{n=1}^\infty C_n^{\rm SW}(g^2, \log a) (m_Q a)^n
890: a\Lambda_{\rm QCD} $ for the (massless) $O(a)$ improved SW quark action.
891: Here we should notice that the contribution of
892: $(\nu-1) \sum_i {\bar q}(x)\gamma_i D_i q(x)$
893: is of order $m_Q^2 a^2\Lambda_{\rm QCD}/m_Q\sim m_Q a^2\Lambda_{\rm QCD}\sim
894: a\Lambda_{\rm QCD}$ in the heavy quark region.
895: This implies that once we fix the pole mass from some spectral
896: quantity, the cutoff effects in other spectral quantities are
897: at most of order $a\Lambda_{\rm QCD}$, not $(m_Q a)^n$,
898: for the Wilson and the
899: (massless) $O(a)$ improved SW quark actions.
900: It should be noted that the quark wave function does not affect on the
901: spectral quantities.
902: If $\nu$, $r_s$, $c_E$ and $c_B$ are
903: properly adjusted in the mass dependent way,
904: the remaining scaling violations are reduced to
905: $\sum_{n=0}^\infty C_n^{\rm ours}(g^2, \log a) (m_Q a)^n
906: (a \Lambda_{\rm QCD} )^2$.
907:
908: This relativistic argument about the on-shell improvement
909: on the massive quarks helps us understand some numerical
910: results for heavy quark physics previously obtained by using
911: the Wilson and SW quark actions under the condition $m_Q
912: a\sim O(1)$.
913: We find a good example in Fig.~2 of Ref.~\cite{jlqcd96},
914: which compares
915: the difference between twice of the heavy-light meson mass and
916: the heavy-heavy one, $2m_{HL}-m_{HH}$, obtained by using the pole mass
917: or the kinetic mass defined by $\partial^2
918: E_{HH}/\partial p_i^2$.
919: We observe that $2m_{HL}-m_{HH}$ with the pole mass is
920: consistent with the experimental values within the
921: ambiguities of order $a\Lambda_{\rm QCD}$ as expected, while
922: $2m_{HL}-m_{HH}$ with the kinetic mass gets deviated from the
923: experimental values further and further as $m_Q a$ becomes larger.
924: Recall that the authors of Ref.~\cite{fermilab} suggest from a
925: nonrelativistic point of view that
926: the kinetic mass should be used for analyses on the
927: heavy quark quantities.
928: Our relativistic argument, however, tells us
929: that the use of the pole mass makes the remaining cutoff
930: effects $O(a\Lambda_{\rm QCD})$.
931: Since the difference between the kinetic mass and the pole
932: mass is of order $(m_Q a)^2$ at the tree-level,
933: the use of the kinetic mass eventually yields unwanted
934: additional $(m_Q a)^2$ errors. This is the reason why the results
935: of $2m_{HL}-m_{HH}$ with the kinetic mass show considerable
936: deviation from the
937: experimental values.
938:
939: At the end of this subsection we have to remark one point.
940: The $m a$ corrections in $Z_m$ and $Z_q$, though they are irrelevant
941: for spectral quantities, become important,
942: together with other renormalization factors
943: in case of calculating the quark
944: masses and the matrix elements of various composite operators.
945: In Sec.~\ref{subsec:npr} we will show how to
946: calculate these renormalization factors nonperturbatively
947: including the $(m_Q a)^n$ corrections.
948:
949: \subsection{Improvement of the axial current}
950: \label{subsec:axial}
951:
952: The cutoff effects in the correlation functions of local
953: composite fields are originated from not only the action but
954: also the composite fields themselves.
955: In this subsection we demonstrate the on-shell improvement
956: on the axial current, which is relevant for calculation
957: of the heavy-light pseudoscalar meson
958: decay constants like $f_B$ and $f_D$.
959:
960: The axial current in the $N_f$ flavor space is given by
961: \be
962: {\cal A}_\mu^a (x)={\bar q}(x)\gamma_\mu\gamma_5\frac{\lambda^a}{2} q(x),
963: \ee
964: where $\lambda^a/2$ ($a=1,\dots,N_f^2-1$)
965: are generators of SU($N_f$).
966: The improvement of this operator is performed in the same
967: way as the improvement of the quark action.
968: After some consideration we find that it is sufficient to
969: consider only the dimension four operators
970: for the $O(a\lqcd)$ improvement:
971: the higher dimensional operators can be reduced to ${\cal
972: A}_\mu^a$ or the
973: dimension four operators multiplied by $(ma)^n$
974: using the classical field equations, or otherwise
975: their contributions are of order $(a\lqcd)^2$.
976: The requirement of various
977: symmetries on the lattice allows the following
978: dimension four operators:
979: \ben
980: {\rm dim.4:}&& ({\cal A}_{4a})_\mu^a(x)={\bar q}(x)\gamma_5 D_\mu
981: \frac{\lambda^a}{2} q(x)+
982: {\bar q}(x){\overleftarrow D}_\mu \gamma_5
983: \frac{\lambda^a}{2} q(x),\\
984: && ({\cal A}_{4b})_\mu^a(x)={\bar q}(x)
985: \gamma_5 \sigma_{\mu\nu}D_\nu
986: \frac{\lambda^a}{2} q(x)
987: -{\bar q}(x){\overleftarrow D}_\nu
988: \sigma_{\mu\nu}\gamma_5
989: \frac{\lambda^a}{2} q(x),
990: \een
991: where we do not take the sum on the index $\mu$.
992: We find that $({\cal A}_{4b})_\mu^a$ is related to
993: $({\cal A}_{4a})_\mu^a$ and ${\cal A}_\mu^a$
994: with the aid of the classical field equations,
995: which means the operator $({\cal A}_{4b})_\mu^a$
996: is redundant for the $O(a\lqcd)$ improvement.
997:
998: After all the improved axial current is written as
999: \be
1000: ({\cal A}_\mu^a)^{\rm imp}(x)=
1001: {\cal A}_\mu^a(x)+d_\mu(g^2,ma)({\cal A}_{4a})_\mu^a(x).
1002: \ee
1003: The parameter $d_\mu$ has been already calculated in the massless case.
1004: Perturbative estimate gives
1005: $d_\mu(g^2,ma=0)=-0.00756g^2$\cite{dmu}, which is fairly
1006: small in magnitude.
1007: From this fact we think that
1008: it would be sufficient to evaluate $d_\mu$ perturbatively
1009: in the massive case.
1010: The lattice operator $({\cal A}_\mu^a)^{\rm imp}$ is related to
1011: the continuum operator with
1012: the renormalization factor $Z_A$:
1013: \be
1014: ({\cal A}_\mu^a)^{\rm con}(x)
1015: =Z_A(g^2,{\rm log}a,ma)({\cal A}_\mu^a)^{\rm imp}(x).
1016: \ee
1017: The power corrections of $ma$ in $Z_A$ need to be under
1018: control to obtain
1019: the heavy-light pseudoscalar meson decay constants defined
1020: in the continuum regularization scheme.
1021: In Sec.~\ref{subsec:npr} we will show how to remove
1022: the $ma$ corrections in $Z_A$ with a nonperturbative
1023: method.
1024:
1025: \subsection{Benefit of chiral symmetry}
1026: \label{subsec:chiral}
1027:
1028: Although the above discussions are free from the chiral symmetry,
1029: it is also interesting to look into what can be changed by
1030: the presence of the chiral symmetry.
1031: For the convenience we treat the quark mass matrix in the
1032: $N_f$ flavor space
1033: \be
1034: M=\left( \begin{array}{cccc}
1035: m_1 & 0 & \cdots & 0 \\
1036: 0 & m_2 & \cdots & 0 \\
1037: \vdots & \vdots & \ddots & \vdots \\
1038: 0 & 0 & \cdots & m_{N_f}
1039: \end{array}
1040: \right)
1041: \ee
1042: as a spurious field $\mm$
1043: which transforms like
1044: \ben
1045: \mm \rightarrow V_R \mm V_L^\dagger,\\
1046: \mm^\dagger \rightarrow V_L \mm^\dagger V_R^\dagger,
1047: \een
1048: under SU$(N_f)_L\times$SU$(N_f)_R$, where $V_L$ and $V_R$ are
1049: elements of the fundamental representation of
1050: SU$(N_f)_L$ and SU$(N_f)_R$ respectively.
1051: In terms of quark fields $q$ and ${\bar q}$,
1052: SU$(N_f)_L$ and SU$(N_f)_R$ act on the left and right handed
1053: components,
1054: \ben
1055: q_L=\frac{1-\gamma_5}{2}q,\\
1056: {\bar q}_L={\bar q}\frac{1+\gamma_5}{2},\\
1057: q_R=\frac{1+\gamma_5}{2}q,\\
1058: {\bar q}_R={\bar q}\frac{1-\gamma_5}{2},
1059: \een
1060: whose transformation properties are given by
1061: \ben
1062: &&q_L \rightarrow V_L q_L,\\
1063: &&{\bar q}_L \rightarrow {\bar q}_L V_L^\dagger,\\
1064: &&q_R \rightarrow V_R q_R,\\
1065: &&{\bar q}_R \rightarrow {\bar q}_R V_R^\dagger.
1066: \een
1067: We work all the calculations assuming this symmetry and
1068: at the end of calculations we can choose $\mm=\mm^\dagger=M$.
1069:
1070:
1071: With this artifice let us consider which operators among
1072: ${\cal O}_3,\dots,{\cal O}_{6f}$ in the quark action are allowed
1073: under SU$(N_f)_L\times$SU$(N_f)_R$ symmetry.
1074: We easily find that the dimension three and five
1075: operators are not allowed.
1076: As for the dimension six operators, some four-fermi
1077: operators are excluded.
1078: We also observe that the power corrections of $m_Q a$ emerge
1079: as the form
1080: \be
1081: M^{2n}\cdot ({\rm SU}(N_f)_L\times{\rm SU}(N_f)_R
1082: \;\;{\rm invariant}\;\;{\rm operators})
1083: \label{eq:ma_power}
1084: \ee
1085: with $n\geq 0$, which means
1086: \ben
1087: &&{\bar q}(x)M^{2n+1}q(x),\\
1088: &&{\bar q}(x)M^{2n}\gamma_\mu D_\mu q(x),\\
1089: &&i{\bar q}(x)M^{2n+1}\sigma_{\mu\nu}F_{\mu\nu} q(x)
1090: \een
1091: are allowed, while
1092: \ben
1093: &&{\bar q}(x)M^{2n}q(x),\\
1094: &&{\bar q}(x)M^{2n+1}\gamma_\mu D_\mu q(x),\\
1095: &&i{\bar q}(x)M^{2n}\sigma_{\mu\nu}F_{\mu\nu} q(x)
1096: \een
1097: are forbidden.
1098: This may be advantageous in controlling the cutoff
1099: effects as $m_Q a$ becomes smaller away from one.
1100: Even for the chiral non-invariant quark action such as
1101: the SW quark action, however,
1102: the leading cutoff effects except for $Z_q$ and $Z_m$
1103: are $ (m_Q a)^n a\lqcd $ with
1104: $n\not=0$, which are of the same order as those in
1105: the chirally symmetric actions,
1106: once the coefficient of ${\cal O}_{5b}$ in
1107: the quark action is nonperturbatively tuned in the
1108: massless limit.
1109:
1110: As for the improvement of the axial current,
1111: the similar argument can be applied. The dimension four operators
1112: $({\cal A}_{4a})_\mu$ and $({\cal A}_{4b})_\mu$
1113: are not allowed by the chiral
1114: symmetry and the power corrections of $m_Q a$ are restricted
1115: to the form of eq.~(\ref{eq:ma_power}).
1116: As an example,
1117: \be
1118: {\bar q}(x)M^{2n+1}\gamma_5 D_\mu \frac{\lambda^a}{2} q(x)+
1119: {\bar q}(x)M^{2n+1}{\overleftarrow D}_\mu \gamma_5
1120: \frac{\lambda^a}{2} q(x),\\
1121: \ee
1122: with $n\geq 0$ are allowed.
1123:
1124: \subsection{$m_Q a$ corrections at tree-level}
1125: \label{subsec:tree}
1126:
1127: In order to control the $m_Q a$ corrections
1128: it should be essential to nonperturbatively
1129: determine the renormalization factors $Z_m$
1130: and $Z_q$ and
1131: the four parameters $\nu$,
1132: $r_s$, $c_E$ and $c_B$.
1133: However, we think it is instructive to first
1134: investigate the $m_Q a$ corrections at the tree-level.
1135:
1136: $Z_q$, $Z_m$, $\nu$ and $r_s$ can be determined by
1137: demanding that the tree-level quark propagator
1138: $S_q(p)$ derived from eq.~(\ref{eq:qaction_iso}) should reproduce
1139: the relativistic form
1140: \be
1141: S_q(p_0,p_i)=
1142: \frac{1}{Z_q}
1143: \frac{ -i \gamma_0 p_0- i \sum_i \gamma_i p_i + m_p}
1144: {p_0^2+\sum_i p_i^2 +m_p^2}
1145: +{\rm (no\;\; pole\;\; terms)}+O((p_i a)^2)
1146: \label{eq:qprop_r_iso}
1147: \ee
1148: around the pole.
1149: Imposing $p_i=0$ we first obtain
1150: \ben
1151: m_p&=&{\rm log}\left|\frac{m_0+r_t+\sqrt{m_0^2+2r_t m_0+1}}
1152: {1+r_t}\right|,\\
1153: Z_m&=&\frac{m_p}{m_0},\\
1154: Z_q&=&{\rm cosh}(m_p)+r_t {\rm sinh}(m_p).
1155: \een
1156: We then find with finite spatial momenta
1157: \ben
1158: \nu&=&\frac{{\rm sinh}(m_p)}{m_p},
1159: \label{eq:nu}\\
1160: r_s&=&\frac{{\rm cosh}(m_p)+r_t {\rm sinh}(m_p)}{m_p}
1161: -\frac{{\rm sinh}(m_p)}{m_p^2}.
1162: \label{eq:rs}
1163: \een
1164: We should notice that the $m_Q a$ corrections start at
1165: $O((m_Q a)^2)$, not $O(m_Q a)$, in the $\nu$ parameter as expected.
1166: Figure~\ref{fig:tree_ma} illustrates the $m_p a$ dependences of
1167: $Z_q$, $Z_m$, $\nu$ and $r_s$ in case of $r_t=1$. We observe that
1168: the $m_p a$ dependences of $\nu$ is relatively
1169: mild compared to those of $Z_q$, $Z_m$ and $r_s$.
1170:
1171: To fix the $c_E$ and $c_B$ parameters we
1172: consider the quark-quark scattering amplitude
1173: depicted in Fig.~\ref{fig:scatt}.
1174: The improvement condition is that
1175: $c_E$ and $c_B$ should
1176: be chosen to reproduce the following form of
1177: scattering amplitude at the on-shell point
1178: removing the $m_Q a$ corrections,
1179: \ben
1180: T &=&-g^2{\bar u}(p^\prime)\gamma_\mu u(p)D_{\mu\nu}(p-p^\prime)
1181: {\bar u}(q^\prime)\gamma_\nu u(q)\nn\\
1182: &&-g^2{\bar u}(q^\prime)\gamma_\mu u(p)D_{\mu\nu}(p-q^\prime)
1183: {\bar u}(p^\prime)\gamma_\nu u(q)
1184: +O((p_i a)^2,(q_i a)^2,(p^\prime_i a)^2,(q_i^\prime a)^2),
1185: \een
1186: where $D_{\mu\nu}$ is the gluon propagator on the lattice.
1187: Notice that with the use of the Gordon identity
1188: (; on-shell condition for external spinors $u$, ${\bar u}$)
1189: the quark-gluon interaction induced by the
1190: clover term can be transformed into the ordinary quark-gluon
1191: vertex:
1192: \ben
1193: &&{\bar u}(p^\prime)\sum_l i \sigma_{0l}
1194: ({\rm sin}(p_l^\prime)-{\rm sin}(p_l)) u(p) \nn\\
1195: &=&{\bar u}(p^\prime)\frac{1}{\nu}[
1196: i{\rm sin}(p_0^\prime)+i{\rm sin}(p_0)
1197: +\gamma_0\{ 2(m_0+r_t)-r_t({\rm cos}(p^\prime_0)+{\rm
1198: cos}(p_0))\nn\\&&
1199: +r_s\sum_l (2-{\rm cos}(p^\prime_l)-{\rm cos}(p_l))\}] u(p),
1200: \label{eq:gordon_t}\\
1201: &&{\bar u}(p^\prime) i \sigma_{i0}
1202: ({\rm sin}(p_0^\prime)-{\rm sin}(p_0)) u(p)
1203: +{\bar u}(p^\prime)\sum_{l}\nu i \sigma_{il}
1204: ({\rm sin}(p_l^\prime)-{\rm sin}(p_l)) u(p) \nn\\
1205: &=&{\bar u}(p^\prime)[{\nu}
1206: (i{\rm sin}(p_i^\prime)+i{\rm sin}(p_i))
1207: +\gamma_i\{ 2(m_0+r_t)-r_t({\rm cos}(p^\prime_0)+{\rm cos}(p_0))\nn\\&&
1208: +r_s\sum_l (2-{\rm cos}(p^\prime_l)-{\rm cos}(p_l))\}] u(p).
1209: \label{eq:gordon_s}
1210: \een
1211: This improvement procedure with the finite quark mass is
1212: an extension of the previous work\cite{c_sw} that
1213: determined the $c_{\rm SW}=c_E=c_B$ parameter up to
1214: one-loop level in the massless case.
1215: After some algebra with the aid of eqs.~(\ref{eq:gordon_t})
1216: and (\ref{eq:gordon_s}), we obtain
1217: \ben
1218: c_E&=&r_t \nu,
1219: \label{eq:ce}\\
1220: c_B&=&r_s,
1221: \label{eq:cb}
1222: \een
1223: where $\nu$ and $r_s$ are already determined from the
1224: on-shell improvement on the quark propagator.
1225:
1226: You may have already noticed that our values for $\nu$, $r_s$ and
1227: $c_E$ are different from those derived in Ref.~\cite{fermilab}.
1228: This difference originates from whether the on-shell
1229: improvement is implemented in the relativistic way or in the
1230: nonrelativistic way. (more precisely, in case that the
1231: Lagrangian does not retain the
1232: rotational invariance on the Euclidean space-time,
1233: we need both the momentum operator and the Hamiltonian
1234: to discuss the rotational symmetry.)
1235: For example, both methods give the same
1236: relation for the $\nu$ and $r_s$ parameters from the
1237: dispersion relation:
1238: \be
1239: \nu^2+r_s{\rm sinh}(m_p)=\frac{{\rm sinh}(m_p)}{m_p}
1240: \left\{{\rm cosh}(m_p)+r_t {\rm sinh}(m_p)\right\}.
1241: \ee
1242: In the nonrelativistic approach, however, $\nu$ and $r_s$
1243: are not distinguishable due to the lack of relativistic
1244: informations.
1245: Generally speaking, we do not have sufficient number of
1246: nonrelativistic conditions
1247: to fix the coefficients of the
1248: relativistic operators with higher dimensions because the
1249: degrees of freedom of quarks are
1250: smaller
1251: in the nonrelativistic approximation compared to the
1252: relativistic case.
1253: Although the nonrelativistic approach with the Wilson type
1254: quark action\cite{fermilab} has been considered to work better
1255: than the NRQCD in the charm quark region where
1256: the sizable relativistic effects are expected,
1257: this is not necessarily assured because this approach
1258: does not meet the relativistic on-shell improvement.
1259:
1260: \subsection{Large $m_Q a$ and static limit}
1261: \label{subsec:static}
1262:
1263: Although we have restricted ourselves to the case of finite
1264: $m_Q a$ so far,
1265: it is worthwhile to show that we can
1266: derive the static quark action from
1267: eq.~(\ref{eq:qaction_iso}) by taking $m_Q a\rightarrow \infty$.
1268: In terms of the heavy quark field $h(x)$ defined by
1269: \ben
1270: q(x) &=& \frac{{\rm e}^{iE t}}{\sqrt{m_0}} h(x),
1271: \een
1272: where $E = \pm i m_p$,
1273: the lattice action in eq.~(\ref{eq:qaction_iso}) becomes
1274: \ben
1275: S_q^{\rm imp} &=& \sum_x \left[
1276: \left({1+ \frac{r_t}{m_0}+\frac{3r_s}{m_0}}\right) \bar h(x) h(x) -
1277: \frac{{\rm e}^{iE}}{2m_0}\bar h(x) (r_t -\gamma_0)
1278: U_0(x) h(x+\hat 0)\right. \nn \\
1279: &&\left.-\frac{{\rm e}^{-iE}}{2m_0}\bar
1280: h(x+\hat 0) (r_t +\gamma_0) U_0(x)^\dagger h(x)\right]
1281: + O\left(\frac{(\nu, c_E, r_s, c_B)}{m_0}\right) .
1282: \een
1283: Taking $m_0\rightarrow \infty$ and setting $r_t=1$ for simplicity,
1284: we obtain
1285: \ben
1286: S_q^{\rm imp} &\simeq& \sum_x \left[\bar h(x) h(x) -
1287: \bar h(x+\hat 0) \frac{1 +\gamma_0}{2} U_0(x)^\dagger h(x) \right]
1288: \een
1289: for the heavy quark($E = i m_p$), or
1290: \ben
1291: S_q^{\rm imp} &\simeq& \sum_x \left[\bar h(x) h(x) -
1292: \bar h(x) \frac{1 -\gamma_0}{2} U_0(x) h(x+\hat 0) \right]
1293: \een
1294: for the heavy anti-quark($E = - i m_p$),
1295: since
1296: \ben
1297: \frac{\nu}{m_0}=\frac{\ce}{m_0}&\sim& O\left(\frac{1}{m_p}\right),
1298: \label{eq:nu/m0}\\
1299: \frac{r_s}{m_0}=\frac{\cb}{m_0}&\sim& O\left(\frac{1}{m_p}\right),
1300: \label{eq:rs/m0} \\
1301: \frac{{\rm e}^{m_p}}{m_0} &\simeq& 1,
1302: \een
1303: where $m_p={\rm log}|1+m_0|$ is used.
1304: We replace $\frac{1\pm\gamma_0}{2} h(x) = h(x)$
1305: from the property that $ \gamma_0 h(x) = h(x)$ for the quark and
1306: $\gamma_0 h(x) = -h(x)$ for the anti-quark.
1307: Thus we exactly obtain the static quark action, where the quark moves forward,
1308: or the static anti-quark action, where the anti-quark moves backward
1309: in time.
1310:
1311:
1312: The existence of the static limit may give some constraints on the
1313: mass dependent scaling violation for $m_Q a \gg 1$.
1314: As an explicit example, we take the heavy-light pseudoscalar
1315: meson decay constant $f_{HL}$.
1316: One can extract $f_{HL}$ from the correlation function of the
1317: heavy-light current
1318: ${\cal A}^{HL}_0(x) = \bar q_H(x)\gamma_0\gamma_5 q_L(x)$
1319: with zero spatial momentum following
1320: \ben
1321: Z_q^{(0)}\langle {\cal A}^{HL}_0 (t) {\cal A}^{HL}_0(0)\rangle
1322: &\simeq& f_{HL}^2 m_{HL} e^{-m_{HL} t},
1323: \een
1324: where $Z_q^{(0)}$ denotes the renormalization factor of the
1325: quark wave function at the tree-level.
1326: On the other hand, in the static limit, the current is given by
1327: $ \bar h(x) \gamma_0 \gamma_5 q_L(x) =
1328: \sqrt{m_0}{{\rm e}^{m_Qt}}{\cal A}^{HL}_0(x)$, and
1329: the correlation function
1330: \ben
1331: m_0 \langle {\cal A}^{HL}_0 (t) {\cal A}^{HL}_0(0)\rangle &\simeq &
1332: C^2 e^{-\Delta E t}
1333: \een
1334: has a well-defined limit.
1335: From the relation $ Z_q^{(0)}={\rm e}^{m_p}\simeq m_0$ with
1336: $r_t=1$ and
1337: $ m_{HL} \simeq m_Q + \Delta E$, we obtain
1338: \ben
1339: f_{HL} &=& \frac{ C}{ \sqrt{m_{HL}}} .
1340: \een
1341: Since the continuum limit can be taken in the static theory,
1342: $ C$ should behave as
1343: \ben
1344: \frac{C}{\lqcd^{3/2}} &=& w_0 + w_k (a\lqcd)^k +O( (a\lqcd)^{k+1})
1345: \een
1346: where $k=1$ for the Wilson light quark action or
1347: $k=2$ for the (nonperturbatively tuned) SW light quark action.
1348: Therefore the following relation holds
1349: for $m_Q a\rightarrow \infty$:
1350: \ben
1351: \frac{f_{HL}}{\lqcd} &=& \sqrt{\frac{\lqcd}{m_{HL}}}
1352: \left[ w_0 + w_k (a\lqcd)^k +O( (a\lqcd)^{k+1})\right].
1353: \label{eq:fhl_sta}
1354: \een
1355: On the other hand, the decay constant should behave
1356: \ben
1357: \frac{f_{HL}}{\lqcd} &=& v_0(\lqcd/m_Q)\nn\\
1358: &&\times\left[1 + v_k^L ( a\lqcd )^k
1359: + v_n^H (m_Q a) ( a\lqcd )^n +
1360: O( (a\lqcd )^{k+1},(a\lqcd )^{n+1})\right]
1361: \label{eq:fhl_rel}
1362: \een
1363: from our consideration in the previous sections, where
1364: $v_k^L$ is the scaling violation caused by the light quark action,
1365: and $v_n^H(m_Q a)$ comes mainly from the Wilson/SW heavy quark action
1366: for $ n=1$ or from the action of eq.~(\ref{eq:qaction_iso}) with
1367: the nonperturbatively tuned $\nu$, $r_s$, $c_E$ and $c_B$ for
1368: $n=2$.
1369: Comparing eqs.~(\ref{eq:fhl_sta}) and (\ref{eq:fhl_rel}), we find
1370: \ben
1371: v_0(\lqcd/m_Q) &\rightarrow& \sqrt{\frac{\lqcd}{m_{HL}}} w_0
1372: \een
1373: and
1374: \ben
1375: v_k^L &\rightarrow& w_k, \qquad v_n^H (m_Q a) \rightarrow 0
1376: \qquad \mbox{for $ n < k$} \\
1377: v_k^L+v_n^H (m_Q a) &\rightarrow& w_k, \qquad \mbox{for $ n = k$} \\
1378: v_k^L &\rightarrow& w_k, \qquad v_n^H (m_Q a) \rightarrow w_n
1379: \qquad \mbox{for $ n > k$}
1380: \een
1381: in the limit $m_Q a \rightarrow\infty$,
1382: where $w_k$ and $w_n$ are universal constants independent of the choice of
1383: the heavy quark action.
1384: Note that we expect $v_0$ can be expanded
1385: in terms of of $m_q/\lqcd$ near the chiral limit.
1386:
1387: The important point is that the function $v_n^H ( m_Q a)$
1388: becomes constant (or even vanishes)
1389: in the limit $m_Q a \rightarrow\infty$. On the other hand,
1390: we also know that the behavior of $v_n^H ( m_Q a)$ is
1391: benign under the condition $m_Q a\simlt 1$.
1392: From these facts it would be reasonable to assume that
1393: the function $v_n^H ( m_Q a)$ behaves modestly all over the
1394: range of $m_Q a$ from the
1395: massless limit to the static limit.
1396: This assumption is supported by previous numerical
1397: results for the heavy-light decay constants $f_{D_s}$ and $f_{B_s}$
1398: which are obtained by using the pole mass under the
1399: condition $m_Q a\simgt 1$
1400: with the Wilson quark action.
1401: Figures 4 and 5 in
1402: Ref.\cite{jlqcd96} show that the naive continuum extrapolation with
1403: a linear form for
1404: $f_{D_s}$ and $f_{B_s}$
1405: gives ``reasonable'' values in the
1406: continuum limit.
1407: This outcome is understood as follows:
1408: Since $v_n^H ( m_Q a)$ is a modest function in terms of
1409: $m_Q a$, the leading scaling violation effects
1410: $v_1^H ( m_Q a)a\lqcd$ for $f_{D_s}$ and $f_{B_s}$
1411: are effectively removed by the naive linear extrapolation.
1412: However, the $O(a\lqcd)$ cutoff
1413: effects cannot be completely removed. The remnant is
1414: estimated to be
1415: $|v_1^H (m_Q a_{\rm max})-v_1^H (m_Q a_{\rm min})|a_{\rm av}\lqcd$, where
1416: $a_{\rm max}$ and $a_{\rm min}$ are the maximum
1417: and minimum of the lattice spacing used for the continuum
1418: extrapolation,
1419: and $a_{\rm av} = a_{\rm max} a_{\rm min}/
1420: (a_{\rm max}-a_{\rm min})$.
1421:
1422:
1423:
1424: Although we have focused on the case of heavy-light
1425: pseudoscalar meson decay constants,
1426: the above arguments on scaling violation effects
1427: can be easily generalized to any observables which
1428: can be defined in the static limit.
1429:
1430: \subsection{Nonperturbative renormalization}
1431: \label{subsec:npr}
1432:
1433: Let us turn to a nonperturbative determination of
1434: $Z_q$, $Z_m$, $\nu$, $r_s$, $\cb$ and $\ce$.
1435: We first consider the $\nu$ and $r_s$ parameters.
1436: Since the rotational symmetry breaking due to the
1437: $m_Q a$ corrections deviates the $\nu$ and $r_s$ parameters
1438: from one, it would be a reasonable way to adjust them such that
1439: the correct dispersion relations are reproduced for some
1440: hadronic states. We think
1441: a set of dispersion relations for the
1442: heavy-heavy and heavy-light mesons is a good choice.
1443: A previous study demonstrated a clear distinction
1444: between the the kinetic
1445: masses defined by $\partial^2 E_{\rm HH}/\partial p_i^2$
1446: for the heavy-heavy meson and
1447: $\partial^2 E_{\rm HL}/\partial p_i^2$ for the heavy-light meson
1448: (see Fig.~1 in Ref.~\cite{jlqcd96}),
1449: which tells us that the dispersion relations for the
1450: heavy-heavy and heavy-light mesons give two independent
1451: conditions to fix both $\nu$ and $r_s$ parameters.
1452: We point out that to avoid the ambiguities
1453: coming from the clover term
1454: it would be better to consider the dispersion relations of
1455: the spin averaged meson states over the pseudoscalar and
1456: vector channels. This is motivated by an
1457: observation that the clover term causes the hyperfine splitting.
1458:
1459: Nonperturbative determination of $\cb$ and $\ce$ is a little bit troublesome.
1460: Although in the massless case the clover coefficient
1461: can be determined with the
1462: aid of the PCAC relation,
1463: this method does not work with the massive case.
1464: The reason is that
1465: the chiral symmetry allows
1466: the clover terms with the odd power of $ma$ corrections
1467: $(ma)^{2n+1}i{\bar q}(x)\sigma_{\mu\nu}F_{\mu\nu} q(x)$
1468: $(n\ge 0)$
1469: as discussed in Sec.~\ref{subsec:chiral}.
1470: This implies that even the chirally symmetric quark actions,
1471: {\it e.g.,} the domain wall and the overlap quark actions,
1472: suffer from this difficulty.
1473: However, we can at least evaluate
1474: $\cb$ and $\ce$ perturbatively up to one-loop level
1475: by extending the calculation in Sec.~\ref{subsec:tree}.
1476: In this case the remaining
1477: cutoff effects are of order $\alpha^2 a\lqcd$, which might be
1478: small enough for numerical studies.
1479:
1480: As for a nonperturbative determination of
1481: renormalization factors,
1482: we consider the use of the Schr\"{o}dinger
1483: functional(SF) method\cite{sf}.
1484: The renormalization of the quark mass is made through the
1485: renormalizations of the axial current and the pseudoscalar
1486: density using the PCAC relation:
1487: \be
1488: {\bar m}(\mu=1/L)=
1489: \frac{Z_A({\bar g},{\bar m}L,g,ma)
1490: \partial_\mu {\bar q}(x)\gamma_\mu\gamma_5 q(x)}
1491: {Z_P({\bar g},L,{\bar m}L,g,ma){\bar q}(x)\gamma_5 q(x)}.
1492: \label{eq:pcac}
1493: \ee
1494: where $L$ is the physical box size and
1495: ${\bar g}$ and ${\bar m}$ are the renormalized coupling and
1496: quark mass in the SF scheme.
1497: $Z_A$ and $Z_P$ with the finite quark mass are defined by
1498: \ben
1499: Z_A({\bar g},{\bar m}L,g,ma)&=&
1500: \frac{\sqrt{c_m f_{1}}}{f_A(x_0=L/2)},\\
1501: Z_P({\bar g},L,{\bar m}L,g,ma)&=&
1502: \frac{\sqrt{c_m f_{1}}}{f_P(x_0=L/2)},
1503: \een
1504: where $c_m$ is a mass-dependent constant, and
1505: $f_A$,
1506: $f_P$ and $f_{1}$ are
1507: the correlation functions given by
1508: \ben
1509: f_A(x_0)&=&-\frac{1}{3}\int d^3{\vec y}d^3{\vec z}
1510: \la {\bar q}(x)\gamma_\mu\gamma_5 q(x)
1511: {\bar \zeta}({\vec y})\gamma_5 \zeta({\vec z}) \ra ,\\
1512: f_P(x_0)&=&-\frac{1}{3}\int d^3{\vec y}d^3{\vec z}
1513: \la {\bar q}(x)\gamma_5 q(x)
1514: {\bar \zeta}({\vec y})\gamma_5 \zeta({\vec z}) \ra ,\\
1515: f_{1}&=&-\frac{1}{3L^6}\int d^3{\vec u}d^3{\vec v} d^3{\vec y}d^3{\vec z}
1516: \la {\bar \zeta}^\prime({\vec u})\gamma_5 \zeta^\prime({\vec v})
1517: {\bar \zeta}({\vec y})\gamma_5 \zeta({\vec z}) \ra
1518: \een
1519: with $\zeta$, $\zeta^\prime$, ${\bar \zeta}$ and
1520: ${\bar \zeta}^\prime$ the boundary quark fields.
1521: We
1522: illustrate the function $f_{A,P}$ and $f_{1}$ in Fig.~\ref{fig:sf}.
1523: As for the boundary conditions we refer to the description
1524: in Ref.~\cite{qmass}.
1525:
1526: Although each of $Z_A$, $Z_P$,
1527: $\la {\bar q}(x)\gamma_\mu\gamma_5 q(x)\ra$ and
1528: $\la {\bar q}(x)\gamma_5 q(x)\ra$ has the power corrections
1529: of $ma$, they are canceled out in the combinations
1530: $Z_A\la {\bar q}(x)\gamma_\mu\gamma_5 q(x)\ra$ and
1531: $Z_P\la {\bar q}(x)\gamma_5 q(x)\ra$ and
1532: the $O((a\Lambda_{\rm QCD})^2)$ uncertainties are
1533: left.
1534: This assures us to take the continuum limit for
1535: the renormalized matrix elements and also for the
1536: renormalized quark mass of eq.~(\ref{eq:pcac}).
1537: These quantities, however, are
1538: defined in the SF scheme with the finite quark mass
1539: and hence different from those
1540: renormalized in the SF scheme at
1541: the massless point, which are exactly what we want.
1542: Therefore, we still need the finite renormalization factor
1543: to make the conversion between the two renormalization descriptions.
1544: This can be obtained by taking the continuum limit of the ratio
1545: \ben
1546: \frac{Z_A({\bar g},{\bar m}L=0,g^\prime,ma^\prime=0)}
1547: {Z_A({\bar g},{\bar m}L,g^\prime,ma^\prime)},\\
1548: \frac{Z_P({\bar g},L,{\bar m}L=0,g^\prime,ma^\prime=0)}
1549: {Z_P({\bar g},L,{\bar m}L,g^\prime,ma^\prime)},
1550: \een
1551: for several $g^\prime$ chosen to satisfy $ma^\prime\ll 1$.
1552: We remark that the physical size $L$ for the SF scheme can
1553: be taken to be much smaller than that for the measurement of
1554: the spectral quantities and the various matrix elements,
1555: allowing us to access the finer lattice spacing with $g^\prime$.
1556:
1557:
1558:
1559:
1560:
1561: \section{Conclusions}
1562: \label{sec:conclusion}
1563:
1564: In this paper we have first examined the validity
1565: of the idea of anisotropic lattice by making a perturbative
1566: calculation of the quark self energy up to $O(g^2 a)$.
1567: Our results show that the $m_Q a_s$ corrections revive
1568: through one-loop diagram.
1569: We also find that on the anisotropic lattice
1570: the four parameters must be adjusted to
1571: remove all the terms of order $a$ even in the massless case.
1572: From a theoretical point of view
1573: the anisotropic lattice is not necessarily advantageous
1574: over the isotropic one.
1575:
1576: In the second part of this paper we have presented a new
1577: relativistic approach to the heavy quarks on the lattice.
1578: The idea is based on the relativistic on-shell improvement with
1579: the finite $m_Q a$ corrections.
1580: We have shown that the cutoff effects can be reduced to
1581: $O((a\Lambda_{\rm QCD})^2)$ putting
1582: the $(m_Q a)^n$ corrections
1583: on the renormalization factors of the quark mass $Z_m$ and
1584: wave function $Z_q$.
1585: As far as the spectral quantities such as hadron masses are concerned,
1586: the $(m_Q a)^n$ corrections in $Z_m$ and $Z_q$
1587: do not matter:
1588: $Z_q$ does not affects on the spectral
1589: quantities and the $(m_Q a)^n$ corrections
1590: in $Z_m$ can be handled by employing the pole
1591: mass fixed from some spectral quantity.
1592: On the other hand, in case of calculating
1593: the quark mass or the various hadron matrix elements
1594: we can control them by determining
1595: the renormalization factors nonperturbatively.
1596:
1597:
1598: The leading scaling violation for
1599: various types of actions are summarized as follows.
1600: $f_1(m_Q a) a\lqcd $ for the Wilson quark
1601: ($f_1(0)\not=0$) and the $O(a)$ improved SW quark ($f_1(0)=0$), while
1602: $f_2(m_Q a) (a\lqcd)^2$ for our proposed action with mass-dependently
1603: tuned $\nu$, $r_s$, $c_E$ and $c_B$. Therefore,
1604: if the magnitude of $f_1( m_Q a)$ is $O(1)$,
1605: the scaling violation for heavy hadron masses
1606: with the $O(a)$ improved SW quark might
1607: be as bad as that for light hadron masses
1608: with the ordinary Wilson quark.
1609: For sufficiently small $m_Q a$,
1610: we can remove the leading scaling violations by
1611: extrapolating the data at several lattice spacings
1612: to the continuum limit.
1613: Even if $m_Q a\sim O(1)$,
1614: $f_2(m_Q a) (a\lqcd)^2$ in our proposed action is expected
1615: to be negligibly small in case of $(a\lqcd)^2\sim 0.01$
1616:
1617: Our relativistic approach has the strong point over the
1618: nonrelativistic ones:
1619: the finer the lattice spacing becomes, the better the
1620: approach works. This is a desirable feature because
1621: we can take the full advantage of configurations with
1622: finer lattice spacing generated to control the cutoff
1623: effects on the light hadron physics.
1624: We are going to perform a numerical simulation to test the
1625: ideas presented in this paper.
1626:
1627: \acknowledgements
1628:
1629: One of us (Y.K.) thanks M.~Okawa for useful discussions.
1630: This work is supported in part by the Grants-in-Aid of the Ministry of
1631: Education (Nos. 13740169,12014202,12640253).
1632:
1633: %\section*{appendix}
1634:
1635:
1636: %\input{ref.tex}
1637:
1638: %\end{document}
1639:
1640: %%%%%%%%%%%%%%%%%% ref.tex %%%%%%%%%%%%%%%%
1641: \begin{thebibliography}{99}
1642:
1643: \bibitem{static} E.~Eichten,
1644: Nucl. Phys. {\bf B} (Proc. Suppl.) {\bf 4} (1988) 170.
1645:
1646: \bibitem{nrqcd} G.~P.~Lepage and B.~A.~Thacker,
1647: Nucl. Phys. {\bf B} (Proc. Suppl.) {\bf 4} (1988) 199;
1648: B.~A.~Thacker and G.~P.~Lepage,
1649: Phys. Rev. D{\bf 43} (1991) 196;
1650: G.~P.~Lepage, L.~Magnea, C.~Nakhleh, U.~Magnea and K.~Hornbostel,
1651: {\it ibid.} {\bf 46} (1992) 4052.
1652:
1653: \bibitem{fermilab} A.~X.~El-Khadra, A.~S.~Kronfeld and P.~B.~Mackenzie,
1654: Phys. Rev. D{\bf 55} (1997) 3933.
1655:
1656: \bibitem{aniso} F.~Karsch,
1657: Nucl. Phys. {\bf B205} (1982) 285;
1658: G.~Burgers, F.~Karsch, A.~Nakamura and I.~O.~Stamatescu,
1659: Nucl. Phys. {\bf B304} (1988) 587.
1660:
1661: \bibitem{symm_at} M.~Alford, T.~R.~Klassen and G.~P.~Lepage,
1662: Nucl. Phys. {\bf B496} (1997) 377;
1663: J.~Harada, A.~S.~Kronfeld, H.~Matsufuru,
1664: N.~Nakajima and T.~Onogi,
1665: hep-lat/0103026.
1666:
1667: \bibitem{sym1} K.~Symanzik, in
1668: {\it Mathematical Problems In Theoretical Physics},
1669: eds. R.~Schrader {\it et al.},
1670: Lecture Notes in Physics, Vol. 153 (Springer, New York, 1982).
1671:
1672: \bibitem{sym2} K.~Symanzik,
1673: Nucl. Phys. {\bf B226} (1983) 187, 205.
1674:
1675: \bibitem{onshell} M.~L\"{u}scher and P.~Weisz,
1676: Commun. Math. Phys. {\bf 97} (1985) 59.
1677:
1678: \bibitem{pp} T.~R.~Klassen,
1679: Nucl. Phys. {\bf B} (Proc. Suppl.) {\bf 73} (1999) 918.
1680:
1681: \bibitem{symm_as} S.~Groote and J.~Shigemitsu,
1682: Phys. Rev. D{\bf 62} (2000) 014508.
1683:
1684: \bibitem{bases} S.~Kawabata, Comput. Phys. Commun. {\bf 41}
1685: (1986) 127; {\it ibid.} {\bf 88} (1995) 309.
1686:
1687: \bibitem{sw_pt} G.~Heatlie, C.~T.~Sachrajda, G.~Martinelli,
1688: C.~Pittori and G.~C.~Rossi,
1689: Nucl. Phys. {\bf B352} (1991) 266.
1690:
1691: \bibitem{pt_g2} E.~Gabrielli, G.~Martinelli,
1692: C.~Pittori, G.~Heatlie and C.~T.~Sachrajda,
1693: Nucl. Phys. {\bf B362} (1991) 475;
1694: S.~Aoki, K.~Nagai, Y.~Taniguchi and A.~Ukawa,
1695: Phys. Rev. D{\bf 58} (1998) 074505.
1696:
1697: \bibitem{pt_g2a} Y.~Taniguchi and A.~Ukawa,
1698: Phys. Rev. D{\bf 58} (1998) 114503.
1699:
1700: \bibitem{sw} B.~Sheikholeslami and R.~Wohlert,
1701: Nucl. Phys. {\bf B259} (1985) 572.
1702:
1703: \bibitem{impr} M.~L\"{u}scher, S.~Sint, R.~Sommer and P.~Weisz,
1704: Nucl. Phys. {\bf B478} (1996) 365.
1705:
1706: \bibitem{jlqcd96} JLQCD Collaboration, S.~Aoki {\it et al.},
1707: Nucl. Phys. {\bf B} (Proc. Suppl.) {\bf 53} (1997) 355.
1708:
1709: \bibitem{dmu} M.~L\"{u}scher and P.~Weisz,
1710: Nucl. Phys. {\bf B479} (1996) 429.
1711:
1712: \bibitem{c_sw} R.~Wohlert,
1713: Improved continuum limit lattice action for quarks,
1714: DESY preprint 87-069 (1987), unpublished.
1715:
1716:
1717: \bibitem{sf} K.~Jansen {\it et al.},
1718: Phys. Lett. {\bf B372} (1996) 275.
1719: \bibitem{qmass} ALPHA Collaboration,
1720: S.~Capitani, M.~L\"{u}scher, R.~Sommer and H.~Wittig,
1721: Nucl. Phys. {\bf B544} (1999) 669.
1722:
1723:
1724: \end{thebibliography}
1725:
1726: \begin{figure}[h]
1727: \centering{
1728: \hskip -0.0cm
1729: \psfig{file=fig1.eps,width=140mm,angle=-90}
1730: \vskip +3mm
1731: }
1732: \caption{One-loop diagrams for the quark self energy.}
1733: \label{fig:qse}
1734: %\vspace{-8mm}
1735: \end{figure}
1736:
1737: \newpage
1738:
1739: \begin{figure}[h]
1740: \centering{
1741: \hskip -0.0cm
1742: \psfig{file=fig2.eps,width=120mm,angle=0}
1743: \vskip +3mm
1744: }
1745: \caption{$\xi$ dependences of (a) $\Delta_q^{(0)}$
1746: and (b) $\Delta_q^{(1)}$ in
1747: the renormalization constant of the quark wave function.
1748: $\ce$ and $\cb$ are chosen to be 1. Errors are within symbols.}
1749: \label{fig:zwf}
1750: %\vspace{-8mm}
1751: \end{figure}
1752:
1753: \newpage
1754:
1755: \begin{figure}[h]
1756: \centering{
1757: \hskip -0.0cm
1758: \psfig{file=fig3.eps,width=120mm,angle=0}
1759: \vskip +3mm
1760: }
1761: \caption{$\xi$ dependences of (a) $\Delta_m^{(0)}$
1762: and (b) $\Delta_m^{(1)}$ in
1763: the renormalization constant of the quark mass.
1764: $\ce$ and $\cb$ are chosen to be 1. Errors are within symbols.}
1765: \label{fig:zm}
1766: %\vspace{-8mm}
1767: \end{figure}
1768:
1769: \newpage
1770:
1771: \begin{figure}[h]
1772: \centering{
1773: \hskip -0.0cm
1774: \psfig{file=fig4.eps,width=120mm,angle=0}
1775: \vskip +3mm
1776: }
1777: \caption{$\xi$ dependences of (a) $\Delta_\nu^{(0)}$
1778: and (b) $\Delta_\nu^{(1)}$ in
1779: the $\nu$ parameter.
1780: $\ce$ and $\cb$ are chosen to be 1. Errors are within symbols.}
1781: \label{fig:nu}
1782: %\vspace{-8mm}
1783: \end{figure}
1784:
1785: \newpage
1786:
1787: \begin{figure}[h]
1788: \centering{
1789: \hskip -0.0cm
1790: \psfig{file=fig5.eps,width=120mm,angle=0}
1791: \vskip +3mm
1792: }
1793: \caption{$\xi$ dependences of $\Delta_\eta^{(0)}$
1794: in the $\eta$ parameter.
1795: $\ce$ and $\cb$ are chosen to be 1. Errors are within symbols.}
1796: \label{fig:eta}
1797: %\vspace{-8mm}
1798: \end{figure}
1799:
1800: \newpage
1801:
1802: \begin{figure}[h]
1803: \centering{
1804: \hskip -0.0cm
1805: \psfig{file=fig6.eps,width=140mm,angle=-90}
1806: \vskip +3mm
1807: }
1808: \caption{Tree-level values for $Z_q$, $Z_m$, $\nu$ and
1809: $r_s$ as functions of $m_p a$. We choose $r_t=1$.}
1810: \label{fig:tree_ma}
1811: %\vspace{-8mm}
1812: \end{figure}
1813:
1814: \newpage
1815:
1816: \begin{figure}[h]
1817: \centering{
1818: \hskip -0.0cm
1819: \psfig{file=fig7.eps,width=140mm,angle=-90}
1820: \vskip +3mm
1821: }
1822: \caption{Tree-level diagrams for the quark-quark scattering.}
1823: \label{fig:scatt}
1824: %\vspace{-8mm}
1825: \end{figure}
1826:
1827: %\newpage
1828: \vspace{10mm}
1829:
1830: \begin{figure}[h]
1831: \centering{
1832: \hskip -0.0cm
1833: \psfig{file=fig8.eps,width=140mm,angle=-90}
1834: \vskip +3mm
1835: }
1836: \caption{Quark diagrams contributing to (a) $f_{A,P}(x)$ and
1837: (b) $f_{1}$. ${\cal O}_{A,P}$ are inserted at the point
1838: $x$. $C$ and $C^\prime$ denote the boundary conditions for
1839: the gauge fields at $t=0$ and $T$.}
1840: \label{fig:sf}
1841: %\vspace{-8mm}
1842: \end{figure}
1843:
1844: \newpage
1845:
1846: \begin{table}[h]
1847: \begin{center}
1848: \caption{\label{tab:action}Expected magnitude of the cutoff
1849: effects due to the higher
1850: dimensional operators in the quark action.
1851: %$m_q$ and
1852: $m_Q$ denote the %light and
1853: heavy quark masses. }
1854: \begin{tabular}{llll}
1855: \multicolumn{2}{c}{operator} & light & heavy \\
1856: \hline
1857: ${\cal O}_{5b}:$ & $i\sum_i {\bar q}(x)\sigma_{0i}F_{0i} q(x)$
1858: & $a \Lambda_{\rm QCD}$
1859: & $a \Lambda_{\rm QCD}^2/m_Q$ \\
1860: & $i\sum_{ij} {\bar q}(x)\sigma_{ij}F_{ij} q(x)$
1861: & $a \Lambda_{\rm QCD}$
1862: & $a \Lambda_{\rm QCD}$ \\
1863: %${\cal O}_{6d}:$ & $mi\sum_i {\bar q}(x)\sigma_{0i}F_{0i} q(x)$
1864: % & $m_q a^2 \Lambda_{\rm QCD}$
1865: % & $(a \Lambda_{\rm QCD})^2$ \\
1866: % & $mi\sum_{ij} {\bar q}(x)\sigma_{ij}F_{ij} q(x)$
1867: % & $m_q a^2 \Lambda_{\rm QCD}$
1868: % & $m_Q a^2 \Lambda_{\rm QCD}$ \\
1869: ${\cal O}_{6a}:$ & ${\bar q}(x)\gamma_0 D_0^3 q(x)$
1870: & $(a \Lambda_{\rm QCD})^2$
1871: &
1872: $(m_Q a)^3 {\bar q}(x) q(x)$,
1873: $(m_Q a)^2 {\bar q}(x)\gamma_i D_i q(x)$,
1874: $(m_Q a) {\bar q}(x)D_i^2 q(x)$ \\
1875: & ${\bar q}(x)\gamma_i D_i^3 q(x)$
1876: & $(a \Lambda_{\rm QCD})^2$
1877: & $(a \Lambda_{\rm QCD})^2$\\
1878: ${\cal O}_{6d}:$ & $i\sum_i {\bar q}(x)\gamma_i[D_0,F_{i0}] q(x)$
1879: & $(a \Lambda_{\rm QCD})^2$
1880: & $a^2 \Lambda_{\rm QCD}^3/m_Q$ \\
1881: & $i\sum_i {\bar q}(x)\gamma_0[D_i,F_{0i}]q(x)$
1882: & $(a \Lambda_{\rm QCD})^2$
1883: & $(a \Lambda_{\rm QCD})^2$ \\
1884: & $i\sum_i {\bar q}(x)\gamma_i[D_j,F_{ij}]q(x)$
1885: & $(a \Lambda_{\rm QCD})^2$
1886: & $a^2 \Lambda_{\rm QCD}^3/m_Q$ \\
1887: ${\cal O}_{6f}:$ &
1888: & $(a \Lambda_{\rm QCD})^2$
1889: & $(a \Lambda_{\rm QCD})^2$ \\
1890: %${\cal O}_{7 }:$ & ${\bar q}(x)D_0^4 q(x)$
1891: % & $(a \Lambda_{\rm QCD})^3$
1892: % &
1893: % $(m_Q a)^4 {\bar q}(x) q(x)$,
1894: % $(m_Q a)^2 {\bar q}(x)D_i^2 q(x)$ \\
1895: % & ${\bar q}(x)D_i^4 q(x)$
1896: % & $(a \Lambda_{\rm QCD})^3$
1897: % & $(a \Lambda_{\rm QCD})^3$
1898: \end{tabular}
1899: \end{center}
1900: \end{table}
1901:
1902: %\begin{table}[h]
1903: %\begin{center}
1904: %\caption{\label{tab:axial}Expected magnitude of the cutoff
1905: %effects due to the higher
1906: %dimensional operators in the axial current.
1907: %$m_q$ and $m_Q$ denote the light and
1908: %heavy quark masses. }
1909: %\begin{tabular}{lll}
1910: %operator & light-light & heavy-light \\
1911: %\hline
1912: %$({\cal A}_{4a})_0$ & $a \Lambda_{\rm QCD}$
1913: % & $a \Lambda_{\rm QCD}$ \\
1914: %$({\cal A}_{4a})_i$ & $a \Lambda_{\rm QCD}$
1915: % & $a \Lambda_{\rm QCD}$ \\
1916: %$({\cal A}_{5a})_0$ & $(a \Lambda_{\rm QCD})^2$
1917: % & $(a \Lambda_{\rm QCD})^2$ \\
1918: %$({\cal A}_{5a})_i$ & $(a \Lambda_{\rm QCD})^2$
1919: % & $(a \Lambda_{\rm QCD})^2$ \\
1920: %$({\cal A}_{5e})_0$ & $(a \Lambda_{\rm QCD})^2$
1921: % & $(a \Lambda_{\rm QCD})^2$ \\
1922: %$({\cal A}_{5e})_i$ & $(a \Lambda_{\rm QCD})^2$
1923: % & $(a \Lambda_{\rm QCD})^2$ \\
1924: %\end{tabular}
1925: %\end{center}
1926: %\end{table}
1927:
1928:
1929: \end{document}
1930:
1931:
1932:
1933:
1934:
1935:
1936:
1937:
1938:
1939:
1940: