1: %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %% This file is part of the APS files in the REVTeX 4 distribution.
5: %% Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %% Copyright (c) 2001 The American Physical Society.
9: %%
10: %% See the REVTeX 4 README file for restrictions and more information.
11: %%
12: %
13: % This is a template for producing manuscripts for use with REVTEX 4.0
14: % Copy this file to another name and then work on that file.
15: % That way, you always have this original template file to use.
16: %
17: % Group addresses by affiliation; use superscriptaddress for long
18: % author lists, or if there are many overlapping affiliations.
19: % For Phys. Rev. appearance, change preprint to twocolumn.
20: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21: % Add 'draft' option to mark overfull boxes with black boxes
22: % Add 'showpacs' option to make PACS codes appear
23: % Add 'showkeys' option to make keywords appear
24: %%\documentclass[10pt,aps,prd,groupedaddress,showpacs]{revtex4}%
25: %% ,draft,showpacs,showkeys]{revtex4}
26:
27: %\documentclass[preprint,tightenlines,aps,prd,groupedaddress,showpacs,floatfix]
28: \documentclass[preprint,aps,prd,groupedaddress,showpacs,floatfix,nofootinbib]
29: {revtex4}%
30: \usepackage{graphicx}
31: \usepackage{amsmath}
32: \usepackage{bm}
33: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
34: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
35: % You should use BibTeX and apsrev.bst for references
36: % Choosing a journal automatically selects the correct APS
37: % BibTeX style file (bst file), so only uncomment the line
38: % below if necessary.
39: %%%%%%\bibliographystyle{apsrev}
40:
41: \begin{document}
42:
43: % Use the \preprint command to place your local institutional report
44: % number in the upper righthand corner of the title page in preprint mode.
45: % Multiple \preprint commands are allowed.
46: % Use the 'preprintnumbers' class option to override journal defaults
47: % to display numbers if necessary
48: \preprint{ANL-HEP-PR-06-41}
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: %Title of paper
51: \title{
52: Potential-model calculation of an order-$\bm{v}^2$ \\
53: nonrelativistic QCD matrix element
54: }
55: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56: % repeat the \author .. \affiliation etc. as needed
57: % \email, \thanks, \homepage, \altaffiliation all apply to the current
58: % author. Explanatory text should go in the []'s, actual e-mail
59: % address or url should go in the {}'s for \email and \homepage.
60: % Please use the appropriate macro foreach each type of information
61: % \affiliation command applies to all authors since the last
62: % \affiliation command. The \affiliation command should follow the
63: % other information
64: % \affiliation can be followed by \email, \homepage, \thanks as well.
65: % \altaffiliation{}
66:
67: \author{Geoffrey T.~Bodwin}
68: \affiliation{
69: High Energy Physics Division,
70: Argonne National Laboratory,
71: 9700 South Cass Avenue, Argonne, Illinois 60439}
72:
73: \author{Daekyoung Kang}
74: \affiliation{
75: Department of Physics, Korea University, Seoul 136-701, Korea}
76: \author{Jungil Lee}
77: \altaffiliation{Visiting faculty, Physics Department, Ohio State University,
78: Columbus, Ohio 43210, USA.}
79: \affiliation{
80: Department of Physics, Korea University, Seoul 136-701, Korea}
81: %\email[]{Your e-mail address}
82: %\homepage[]{Your web page}
83: %\thanks{}
84: %\altaffiliation{}
85:
86: %Collaboration name if desired (requires use of superscriptaddress
87: %option in \documentclass). \noaffiliation is required (may also be
88: %used with the \author command).
89: %\collaboration can be followed by \email, \homepage, \thanks as well.
90: %\collaboration{}
91: %\noaffiliation
92:
93: \date{\today}
94: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
95: \begin{abstract}
96: We present two methods for computing dimensionally-regulated NRQCD
97: heavy-quarkonium matrix elements that are related to the second
98: derivative of the heavy-quarkonium wave function at the origin. The
99: first method makes use of a hard-cutoff regulator as an intermediate
100: step and requires knowledge only of the heavy-quarkonium wave function.
101: It involves a significant cancellation that is an obstacle to
102: achieving high numerical accuracy. The second method is more direct
103: and yields a result that is identical to the Gremm-Kapustin relation,
104: but it is limited to use in potential models. It can be generalized to
105: the computation of matrix elements of higher order in the heavy-quark
106: velocity and can be used to resum the contributions to decay and
107: production rates that are associated with those matrix elements.
108: We apply these methods to the Cornell potential model and compute a
109: matrix element for the $J/\psi$ state that appears in the leading
110: relativistic correction to the production and decay of that state
111: through the color-singlet quark-antiquark channel.
112: \end{abstract}
113:
114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
115: % insert suggested PACS numbers in braces on next line
116: \pacs{12.38.-t, 12.39.Pn, 12.38.Bx, 13.20.Gd, 14.40.Gx}
117: % 12.38.-t Quantum chromodynamics
118: % 12.39.Pn Potential models
119: % 12.38.Bx Perturbative calculations
120: % 13.20.Gd Decays of J/psi, Upsilon, and other quarkonia
121: % 14.40.Gx Mesons with S=C=B=0, mass > 2.5 GeV (including quarkonia)
122: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
123: % insert suggested keywords - APS authors don't need to do this
124: %\keywords{}
125:
126: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
127: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
128: \maketitle
129:
130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
131: % body of paper here - Use proper section commands
132: % References should be done using the \cite, \ref, and \label commands
133: %===========================================================
134: \section{Introduction\label{intro}}
135: %===========================================================
136:
137: In the effective field theory nonrelativistic quantum chromodynamics
138: (NRQCD), the leading relativistic corrections to $S$-wave
139: heavy-quarkonium decay and production processes in the color-singlet
140: quark-antiquark channel are proportional to matrix elements that
141: are related to the second derivative of the quarkonium wave function at
142: the origin. These matrix elements are inherently nonperturbative in
143: nature. Their importance in phenomenological calculations has led to a
144: number of attempts to determine their values.
145:
146: Even before the introduction of the NRQCD approach for quarkonium decay
147: and production \cite{Bodwin:1994jh}, these matrix elements appeared in
148: phenomenological studies of charmonium decays \cite{Kwong:1987ak}. Owing
149: to uncertainties that arise from the uncertainty in the charm-quark
150: mass $m_c$ and from uncalculated terms of higher order in the quantum
151: chromodynamic (QCD) strong coupling $\alpha_s$, such phenomenological
152: determinations have not led to accurate values for the matrix elements.
153: There also have been attempts to determine the matrix elements in
154: lattice calculations \cite{bks}. In this case, large uncertainties arise
155: because there is a substantial cancellation that occurs when one
156: converts from lattice to continuum dimensionally-regulated matrix
157: elements. In principle, one can determine these matrix elements
158: approximately by making use of the Gremm-Kapustin relation
159: \cite{Gremm:1997dq}, which expresses the matrix elements in terms of the
160: quarkonium and heavy-quark masses. (See Ref.~\cite{Braaten:2002fi} for
161: an example of this approach.) Unfortunately, this method is plagued by
162: large uncertainties in $m_c$. In both the lattice and
163: Gremm-Kapustin approaches, the uncertainties are so large that even the
164: signs of the matrix elements are in some doubt.
165:
166: A further difficulty that complicates the calculation of the matrix
167: elements that are related to the second derivative of the wave function
168: at the origin is that they contain a linear ultraviolet (UV) divergence,
169: and, hence, must be regulated. Dimensional regularization of these
170: matrix elements is particularly useful because it is consistent with
171: existing calculations of quarkonium decay and production rates at
172: relative order $\alpha_s$ and $\alpha_s^2$.
173:
174: In this paper, we present two methods for calculating the second
175: derivative of the wave function at the origin. In the first method, we
176: initially use a hard-cutoff regulator, with cutoff $\Lambda$, to define
177: the relevant matrix element. Then we compute the difference between the
178: hard-cutoff regularization and dimensional regularization in
179: perturbation theory. We subtract this difference from the hard-cutoff
180: matrix element. There remains a dependence of the matrix element on
181: $\Lambda$ that falls as $1/\Lambda$ in the limit $\Lambda\to\infty$.
182: That dependence can be removed by calculating at a number of values of
183: $\Lambda$ and extrapolating to $\Lambda=\infty$. However, in this
184: method, it is difficult to achieve high numerical accuracy because
185: large, cancelling $\Lambda$-dependent contributions appear. The second
186: method that we present bypasses the hard-cutoff step, but it is
187: applicable only to potential models for the wave function. It yields a
188: result that is identical to the Gremm-Kapustin relation
189: \cite{Gremm:1997dq} for the potential model. That result can be used to
190: resum certain contributions of higher order in $v$ to amplitudes that
191: are computed in NRQCD.
192:
193: Having established a formal procedure for computing the relevant matrix
194: elements, we carry out an explicit computation for the $J/\psi$ (or
195: $\eta_c$) states in the Cornell potential model. We do not distinguish
196: between the $J/\psi$ and $\eta_c$ wave functions, which differ only in
197: corrections of relative order $v^2$, where $v$ is the velocity of the
198: heavy quark or antiquark in the quarkonium. ($v^2\approx 0.3$ in
199: charmonium and $v^2\approx 0.1$ in bottomonium.) In principle, if we
200: know the static heavy-quark-antiquark ($Q\bar Q$) potential exactly,
201: then we can calculate the quarkonium wave function of the leading $Q\bar
202: Q$ quarkonium Fock state up to corrections of relative order $v^2$.
203: Existing lattice data for the static $Q\bar Q$ potential yield values
204: for the string tension. We examine values for the parameters in the
205: Cornell potential that bracket the lattice values for the string
206: tension.
207:
208: In our numerical calculations, the results from our two approaches agree
209: well and give a value for the second derivative of the wave function at
210: the origin that is in agreement with expectations from the $v$-scaling
211: rules of NRQCD \cite{Bodwin:1994jh}. The largest uncertainties in our
212: calculation are of relative order $v^2$. Therefore, our
213: determination of the second derivative of the wave function at the
214: origin is the most accurate to date and should be useful for
215: phenomenological studies of quarkonium production and decay.
216:
217: The remainder of this paper is organized as follows: In
218: Sec.~\ref{model}, we give a brief description of the Cornell potential
219: model. In Sec.~\ref{matrix-els}, we discuss the NRQCD matrix elements
220: that are relevant to this work. Section~\ref{cutoff} contains a description
221: of the hard-cutoff regulator. In Sec.~\ref{difference}, we explain how
222: we compute the difference between hard-cutoff regularization and
223: dimensional regularization. The direct method of calculation, which
224: bypasses the hard-cutoff step, is discussed in Sec.~\ref{direct}, along
225: with its application to resummation of contributions of higher order in
226: $v$. In Sec.~\ref{decomposition}, we decompose the difference
227: between a hard-cutoff matrix element and a dimensionally-regulated
228: matrix element into sums of short-distance coefficients times
229: dimensionally-regulated matrix elements. Such a decomposition is used
230: in existing calculations of the difference between a lattice matrix element
231: and a dimensionally-regulated matrix element.
232: Sec.~\ref{numerical} contains our numerical
233: results and a discussion of them. We summarize our results in
234: Sec.~\ref{conclusions}. In Appendix~\ref{coulomb}, we illustrate our
235: methods for the case of a pure Coulomb potential, and, in
236: Appendix~\ref{integrals}, we list some integrals that are useful in our
237: analyses.
238:
239:
240: %===========================================================
241: \section{Potential model\label{model}}
242: %===========================================================
243: We compute matrix elements for the $J/\psi$ or $\eta_c$ states using a
244: potential model. In this model, we neglect the effects of the heavy-quark
245: spin, which are suppressed as $v^2$. Therefore, we do not distinguish
246: between the $J/\psi$ and $\eta_c$ wave functions or matrix elements. We
247: note that, if we knew the heavy-quark potential exactly, then we could
248: calculate the heavy-quarkonium wave function in a potential model up to
249: corrections of relative order $v^2$ (Ref.~\cite{Brambilla:1999xf}). We
250: make use of the Cornell potential model of Ref.~\cite{Eichten:1978tg}.
251: For appropriate choices of parameters, the Cornell potential
252: provides a reasonably good fit to heavy-quark potentials that are
253: measured in lattice calculations.\footnote{For a recent review that
254: discusses heavy-quark potentials from lattice measurements, see
255: Ref.~\cite{Bali:2000gf}.}
256:
257: Now we summarize the methods that we use to constrain the parameters of
258: the Cornell potential and to solve the Schr\"odinger equation. We refer
259: the reader to Ref.~\cite{bkl} for further details.
260:
261: The Cornell potential \cite{Eichten:1978tg} is given by
262: %--------------------------------------------------------
263: \begin{equation}
264: V(r)=-\frac{\kappa}{r}+\frac{r}{a^2},
265: \label{model-V}
266: \end{equation}
267: %--------------------------------------------------------
268: where the parameters $\kappa$ and $a$ determine the strength of
269: Coulomb and linear potentials, respectively.
270: For a color-singlet $Q\bar{Q}$ pair, the Coulomb-strength
271: parameter $\kappa$ can be expressed in terms of an effective
272: strong coupling $\alpha_s$ as
273: %--------------------------------------------------------
274: \begin{equation}
275: \kappa=\alpha_s C_F,
276: \label{kappa}
277: \end{equation}
278: %--------------------------------------------------------
279: where $C_F=4/3$.
280: The parameter $a$ is related to the string tension $\sigma$ as
281: %--------------------------------------------------------
282: \begin{equation}
283: \sigma=\frac{1}{a^2}.
284: \label{sigma-par}
285: \end{equation}
286: %--------------------------------------------------------
287:
288: Following Ref.~\cite{Eichten:1978tg}, we replace $\kappa$ and the
289: binding energy $\epsilon_{\textrm{B}}$ by dimensionless parameters
290: $\lambda$ and $\zeta$:
291: %--------------------------------------------------------
292: \begin{subequations}
293: \label{lambda-zeta}
294: \begin{eqnarray}
295: \kappa&=&(ma)^{-\frac{2}{3}}\,\lambda,\label{lambda}\\
296: \epsilon_{\textrm{B}}&=& m (ma)^{-\frac{4}{3}}\zeta.\label{zeta}
297: \end{eqnarray}
298: \end{subequations}
299: %--------------------------------------------------------
300: For a given value of $\lambda$, we fix the heavy-quark mass $m$ and the
301: parameter $a$ in the Cornell potential as follows. First, we require
302: that the energies that result from the solutions to the Schr\"odinger
303: equation match the measured values of the difference of the $J/\psi$ and
304: $\psi(2S)$ masses. We use $M_{J/\psi}=3.096916$~GeV and
305: $M_{\psi(2S)}=3.686093$~GeV. Second, we require that the wave function
306: at the origin matches a value that is derived from the measured value of
307: the leptonic width of the $J/\psi$ and the perturbative formula
308: %--------------------------------------------------------
309: \begin{equation}
310: \Gamma[J/\psi\to \ell^+\ell^-]
311: =\frac{4\pi e_c^2 \alpha^2}{m_c^2}|\psi(0)|^2
312: \left(1-\frac{8}{3}\frac{\alpha_s}{\pi}\right)^2.
313: \label{lept-width}
314: \end{equation}
315: %--------------------------------------------------------
316: Here, $\psi(0)$ is the wave function at the origin and $e_c=2/3$ is the
317: fractional electric charge of the charm quark.
318: In Ref.~\cite{Braaten:2002fi}, $\psi(0)$ is estimated by using the
319: the formula (\ref{lept-width}) at both leading order in $\alpha_s$ (LO)
320: and next-to-leading order in $\alpha_s$ (NLO). The results are
321: \begin{equation}
322: \psi(0)=
323: \left\{
324: \begin{array}{lcl}
325: 0.18619~\textrm{GeV}^{3/2}&& \textrm{(LO)}
326: \\
327: 0.23629~\textrm{GeV}^{3/2}&& \textrm{(NLO)},
328: \end{array}
329: \right.
330: \label{psi0-value}
331: \end{equation}
332: where, for convenience, we have taken $\psi(0)$ to be positive and real.
333: In order to estimate the effects of the uncertainty in $\psi(0)$, we
334: carry out our calculations for both the LO and NLO values of $\psi(0)$.
335: For a more detailed discussion of the determination of $m$ and $a$, see
336: Ref.~\cite{bkl}.
337:
338: Values of $m$, $a$, and the scaled energies $\zeta$ of the $1S$ and $2S$
339: states for various values of $\lambda$ are shown in
340: Tables~\ref{table:ma-lo} and \ref{table:ma-nlo}, along with values for
341: $\alpha_s$ from Eq.~(\ref{kappa}), $\sigma$ from Eq.~(\ref{sigma-par}),
342: and $\gamma_{\textrm{C}}$ from Eq.~(\ref{gamma-C}).
343: Table~\ref{table:ma-lo} contains values of the potential-model
344: parameters that correspond to the LO value of $\psi(0)$, while
345: Table~\ref{table:ma-nlo} contains those that correspond to the NLO value
346: of $\psi(0)$.
347: %--------------------------------------------------------
348: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
349: % tables should appear as floats within the text
350: % Here is an example of the general form of a table:
351: % Fill in the caption in the braces of the \caption{} command. Put the label
352: % that you will use with \ref{} command in the braces of the \label{} command.
353: % Insert the column specifiers (l, r, c, d, etc.) in the empty braces of the
354: % \begin{tabular}{} command.
355: % The ruledtabular enviroment adds doubled rules to table and sets a
356: % reasonable default table settings.
357: % Use the table* environment to get a full-width table in two-column
358: % Add \usepackage{longtable} and the longtable (or longtable*}
359: % environment for nicely formatted long tables. Or use the [H]
360: % placement option to break a long table (with less control than
361: % in longtable).
362: \begin{table}%[H] add [H] placement to break table across pages
363: \caption{\label{table:ma-lo}
364: Potential-model parameters and derived quantities as a function of the
365: strength $\lambda$ of the Coulomb potential. The definitions of the
366: parameters and derived quantities are given in the text. The parameters
367: are computed using the inputs $M_{J/\psi}=3.096916$~GeV,
368: $M_{\psi(2S)}=3.686093$~GeV, and the LO value
369: $\psi(0)=0.18619$~GeV$^{3/2}$, as is described in the text.
370: }
371: \begin{ruledtabular}
372: \begin{tabular}{c|lllllllll}
373: $\lambda$&
374: 0& 0.2 & 0.4 & 0.6 & 0.7 & 0.8 & 1.0 & 1.2 & 1.4\\
375: \hline
376: $\zeta_{10}$&
377: 2.33811& 2.16732& 1.98850& 1.80107& 1.70394& 1.60441& 1.39788& 1.18084& 0.95264
378: \\
379: $\zeta_{20}$&
380: 4.08790& 3.97017& 3.85003& 3.72747& 3.66528& 3.60249& 3.47510& 3.34529& 3.21307
381: \\
382: $m$~(GeV)&
383: 1.70670& 1.51548& 1.35120& 1.21003& 1.14710& 1.08877& 0.98458& 0.89501& 0.81796
384: \\
385: $a$~(GeV$^{-1}$)&
386: 1.97932& 2.08520& 2.19805& 2.31833& 2.38139& 2.44648& 2.58295& 2.72816& 2.88253
387: \\
388: $\sqrt{\sigma}$~(GeV)&
389: 0.50522& 0.47957& 0.45495& 0.43134& 0.41992& 0.40875& 0.38715& 0.36655& 0.34692
390: \\
391: $\sigma$~(GeV$^2$)&
392: 0.25525& 0.22999& 0.20698& 0.18606& 0.17633& 0.16708& 0.14989& 0.13436& 0.12035
393: \\
394: $\alpha_s$&
395: 0. & 0.06966& 0.14519& 0.22624& 0.26866& 0.31225& 0.40255& 0.49634& 0.59272
396: \\
397: $\gamma_{\textrm{C}}$~(GeV)&
398: 0. & 0.07037& 0.13079& 0.18250& 0.20545& 0.22664& 0.26423& 0.29615& 0.32322
399: \end{tabular}
400: \end{ruledtabular}
401: \end{table}
402: % Surround table environment with turnpage environment for landscape
403: % table
404: % \begin{turnpage}
405: % \begin{table}
406: % \caption{\label{}}
407: % \begin{ruledtabular}
408: % \begin{tabular}{}
409: % \end{tabular}
410: % \end{ruledtabular}
411: % \end{table}
412: % \end{turnpage}
413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
414:
415: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
416: \begin{table}%[H] add [H] placement to break table across pages
417: \caption{\label{table:ma-nlo} As in Table~\ref{table:ma-lo}, except
418: that the potential-model parameters are computed using the NLO value
419: $\psi(0)=0.23629$~GeV$^{3/2}$.
420: }
421: \begin{ruledtabular}
422: \begin{tabular}{c| lll lll lll}
423: $\lambda$&
424: 0.2 & 0.4 & 0.6 & 0.8 & 0.9 & 1.0 & 1.1 & 1.2 & 1.4 \\
425: \hline
426: $\zeta_{10}$&
427: 2.16732& 1.98850& 1.80107& 1.60441& 1.50242& 1.39788& 1.29071& 1.18084& 0.95264
428: \\
429: $\zeta_{20}$&
430: 3.97017& 3.85003& 3.72747& 3.60249& 3.53910& 3.47510& 3.41050& 3.34529& 3.21307
431: \\
432: $m$~(GeV)&
433: 2.08228& 1.85655& 1.66259& 1.49597& 1.42168& 1.35282& 1.28896& 1.22975& 1.12388
434: \\
435: $a$~(GeV$^{-1}$)&
436: 1.92598& 2.03021& 2.14130& 2.25967& 2.32171& 2.38571& 2.45174& 2.51984& 2.66243
437: \\
438: $\sqrt{\sigma}$~(GeV)&
439: 0.51922& 0.49256& 0.46700& 0.44254& 0.43072& 0.41916& 0.40787& 0.39685& 0.37560
440: \\
441: $\sigma$~(GeV$^2$)&
442: 0.26959& 0.24262& 0.21809& 0.19584& 0.18552& 0.17570& 0.16636& 0.15749& 0.14107
443: \\
444: $\alpha_s$&
445: 0.05942& 0.12387& 0.19301& 0.26638& 0.30448& 0.34342& 0.38310& 0.42343& 0.50566
446: \\
447: $\gamma_{\textrm{C}}$~(GeV)&
448: 0.08249& 0.15331& 0.21393& 0.26567& 0.28859& 0.30972& 0.32920& 0.34714& 0.37887
449: \end{tabular}
450: \end{ruledtabular}
451: \end{table}
452:
453: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
454:
455: Lattice measurements of the heavy-quark potential yield values for
456: effective coupling $\alpha_s$ of 0.22 in the quenched case and
457: approximately 0.26 in the unquenched case \cite{Bali:2000gf}. A lattice
458: measurement of the string tension $K=\sigma$ (Ref.~\cite{Booth:1992bm})
459: gives $Ka_{\rm L}^2=0.0114(2)$ at a lattice coupling $\beta=6.5$, where
460: $a_{\rm L}$ is
461: the lattice spacing. Lattice calculations of the hadron spectrum at
462: $\beta=6.5$ yield values for $1/a_{\rm L}$ of $3.962(127)$~GeV
463: (Refs.~\cite{Gupta:1996sa,Kim:1993gc}) and $3.811(59)$~GeV
464: (Refs.~\cite{Gupta:1996sa,Kim:1996cz}). These yield values of the string
465: tension of $K=0.1790\pm 0.0119$ and $K=0.1656\pm 0.0059$, respectively.
466:
467: Comparing the results of these lattice measurements with the LO
468: parameters in Table~\ref{table:ma-lo}, we see that the values of the
469: string tension at $\lambda=0.7$ and $0.8$ span the range of lattice
470: results for the string tension, while the values of $\alpha_s$ at
471: $\lambda=0.6$ and $0.7$ span the range of lattice results for $\alpha_s$.
472: Comparing the results of the lattice measurements with the NLO
473: parameters in Table~\ref{table:ma-nlo}, we see that the values of
474: the string tension at $\lambda=1.0$ and $1.1$ span the range of lattice
475: results for the string tension. However, the values of $\alpha_s$ at
476: $\lambda=0.9$, $1.0$, and $1.1$ are all larger than the lattice values.
477: It is not clear whether this discrepancy between the lattice and NLO
478: potential-model values for $\alpha_s$ arises from the use of an
479: inaccurate value for $\psi(0)$, from effects due to the running of
480: $\alpha_s$, which are not taken into account in the fits to the lattice
481: data, or from the absence of corrections of relative order $v^2$ in
482: the Cornell potential model. However, we note that the NLO values for
483: $\alpha_s$ at $\lambda=0.9$, $1.0$, and $1.1$ do not differ greatly from
484: the value of the running $\alpha_s$ at the scale of the heavy-quark
485: momentum $m_cv$.
486:
487: Finally, we mention that we obtain the $J/\psi$ wave function by expressing
488: the radial part of the Schr\"odinger equation as a difference equation,
489: which we integrate numerically. See Ref.~\cite{bkl} for details.
490:
491: %===========================================================
492: \section{ NRQCD matrix elements \label{matrix-els}}
493: %===========================================================
494: In this section we describe the NRQCD matrix elements that are relevant
495: to our calculation.
496:
497: In the rest frame of an $S-$wave heavy quarkonium $H$ in a spin-singlet
498: (${}^1S_0$) or spin-triplet (${}^3S_1$) state, one can express the wave
499: function at the origin of the leading $Q\bar Q$ Fock state in terms of
500: the following color-singlet NRQCD matrix elements
501: \cite{Bodwin:1994jh}:
502: %--------------------------------------
503: \begin{subequations}
504: \label{psi0}
505: \begin{eqnarray}
506: \psi(0)&=&
507: \int\frac{d^3p}{(2\pi)^3}
508: \widetilde{\psi}(\bm{p})
509: =\frac{1}{\sqrt{2N_c}}
510: \langle 0|\chi^\dagger\psi|H({}^1S_0)\rangle,
511: \\
512: \bm{\epsilon}\psi(0)&=&\bm{\epsilon}\int\frac{d^3p}{(2\pi)^3}
513: \widetilde{\psi}(\bm{p})
514: =\frac{1}{\sqrt{2N_c}}
515: \langle 0|\chi^\dagger\bm{\sigma}\psi|H({}^3S_1)\rangle.
516: \end{eqnarray}
517: \end{subequations}
518: %--------------------------------------
519: Here $\psi$ and $\chi^\dagger$ are Pauli spinor fields
520: that annihilate a quark and an antiquark, respectively.
521: The bilinear operators involving $\psi^\dagger$ and $\chi$ are evaluated
522: at zero space-time position. $\bm{\sigma}$ is a Pauli matrix, and
523: $\bm{\epsilon}$ is the quarkonium polarization vector.
524: $\widetilde{\psi}(\bm{p})$ is the momentum-space wave function for the
525: leading $Q(\bm{p})\bar{Q}(-\bm{p})$ Fock state in the quarkonium.
526: The wave function is, of course, gauge dependent.
527: Throughout this paper, we work in the Coulomb gauge.
528: The normalization factor $1/\sqrt{2N_c}$ accounts for the traces in the
529: SU(2)-spin and SU(3)-color spaces. In equating the wave functions for the
530: spin-singlet and spin-triplet cases, we are ignoring effects of
531: relative order $v^2$.
532:
533: Relativistic corrections to the production and decay rates for
534: a heavy quarkonium involve matrix elements that are related to the
535: second derivative of the wave function at the origin:
536: \begin{subequations}
537: \label{psi20}
538: \begin{eqnarray}
539: \psi^{(2)}(0)&\equiv&
540: \int\frac{d^3p}{(2\pi)^3}
541: \bm{p}^2
542: \widetilde{\psi}(\bm{p})
543: =\frac{1}{\sqrt{2N_c}}
544: \langle 0|\chi^\dagger\left(-\bm{\nabla}^2\right)\psi|H({}^1S_0)\rangle,
545: \\
546: \bm{\epsilon}\psi^{(2)}(0)&\equiv&\bm{\epsilon}\int\frac{d^3p}{(2\pi)^3}
547: \bm{p}^2
548: \widetilde{\psi}(\bm{p})
549: =\frac{1}{\sqrt{2N_c}}
550: \langle 0|\chi^\dagger\bm{\sigma}\left(-\bm{\nabla}^2\right)
551: \psi|H({}^3S_1)\rangle.
552: \end{eqnarray}
553: \end{subequations}
554: Usually, these operator matrix elements are written in terms of the
555: covariant derivative $\bm{D}$ (Ref.~\cite{Bodwin:1994jh}), rather than
556: $\bm{\nabla}$. However, in the Coulomb gauge, the difference between the
557: $\bm{D}$ and $\bm{\nabla}$ is suppressed as $v$
558: (Ref.~\cite{Bodwin:1994jh}).
559:
560: The quantity $\psi^{(2)}(0)$ is the focus of this paper. It is
561: common in phenomenology to make use of a parameter
562: \begin{equation}
563: \langle v^2\rangle=\frac{\psi^{(2)}(0)}{m_c^2\psi(0)},
564: \label{v-sq}
565: \end{equation}
566: where $m_c$ is the charm-quark pole mass, which we distinguish from the
567: parameter $m$ that appears in our potential model.
568: Note that $\psi^{(2)}(0)$ is different from the expectation value of
569: $\bm{p}^{2}$:
570: %--------------------------------------
571: \begin{equation}
572: \psi^{(2)}(0)\neq
573: \int\frac{d^3p}{(2\pi)^3}
574: \,\bm{p}^{2}
575: \widetilde{\psi}^*(\bm{p})
576: \widetilde{\psi}(\bm{p}).
577: \end{equation}
578: %--------------------------------------
579:
580: Let us investigate the ultraviolet behavior of the matrix elements in
581: Eq.~(\ref{psi20}). At large momentum $|\bm{p}|$, the interaction of the
582: $Q\bar{Q}$ pair in QCD is dominated by the Coulomb potential. In this
583: limit, the bound-state wave function approaches the pure Coulomb wave
584: function:
585: %--------------------------------------
586: \begin{equation}
587: \widetilde{\psi}(\bm{p})
588: \sim \frac{1}{(\bm{p}^2+\gamma^2_{\textrm{C}})^2}.
589: \label{coulomb-wf}
590: \end{equation}
591: %--------------------------------------
592: Here $\gamma_{\textrm{C}}$ is a parameter that is related to the binding
593: energy of Coulomb interaction:
594: %--------------------------------------
595: \begin{equation}
596: \gamma_{\textrm{C}}=\frac{1}{2}\,\alpha_s C_F m,
597: \label{gamma-C}
598: \end{equation}
599: %--------------------------------------
600: where $\alpha_s$ is the effective strong coupling, $C_F=4/3$, and
601: $m$ is the quark mass.
602: Substituting Eq.~(\ref{coulomb-wf}) into Eq.~(\ref{psi20}),
603: we see that the matrix elements in Eq.~(\ref{psi20}) have a linear ultraviolet
604: divergence. Hence, in order to define them, we must impose a regulator.
605:
606:
607: %===========================================================
608: \section{Hard-cutoff regulator\label{cutoff}}
609: %===========================================================
610:
611: Our ultimate goal is to regulate the matrix elements in
612: Eq.~(\ref{psi20}) using dimensional regularization. However, as an
613: intermediate step, we impose a hard-cutoff regulator. In principle, the
614: methods that we use could be employed in lattice calculations of the
615: matrix elements, in which the lattice supplies the hard-cutoff
616: regulator. However, for the purposes of the present work, we make use of
617: a simple, analytic UV regulator in momentum space:
618: %--------------------------------------
619: \begin{equation}
620: \label{psi-lam}
621: \psi^{(2)}_\Lambda(0)=
622: \int\frac{d^3p}{(2\pi)^3}\frac{\Lambda^2}{\bm{p}^2+\Lambda^2}\, \bm{p}^2
623: \widetilde{\psi}(\bm{p}),
624: \end{equation}
625: %--------------------------------------
626: where $\Lambda$ is the cutoff.
627:
628: We introduce the Fourier transform of the wave function to coordinate
629: space:
630: \begin{equation}
631: \psi(\bm{x})=\int\frac{d^3{p}}{(2\pi)^3}e^{i\bm{p}\cdot\bm{x}}
632: \widetilde{\psi}(\bm{p}).
633: \label{fourier-tx-wavefn}
634: \end{equation}
635: For $S$-wave states, we can write $\psi(\bm{x})=R(r)/\sqrt{4\pi}$, where
636: $r=|\bm{x}|$ and $R(r)$ is the radial wave function. Substituting
637: Eq.~(\ref{fourier-tx-wavefn}) into Eq.~(\ref{psi-lam}), we can carry out
638: the angular integration over $\bm{p}$ to obtain an expression in
639: coordinate space:
640: %--------------------------------------
641: \begin{equation}
642: \label{psi-lamr}
643: \psi^{(2)}_\Lambda(0)=
644: \frac{\Lambda^2}{\sqrt{4\pi}}
645: \left[R(0)-\Lambda^2\int_0^\infty rR(r)e^{-\Lambda r}dr
646: \right].
647: \end{equation}
648: %--------------------------------------
649:
650:
651: %===========================================================
652: \section{Difference between hard-cutoff and dimensional
653: regularization\label{difference}}
654: %===========================================================
655: Now we work out the difference between the hard-cutoff matrix
656: element and the dimensionally-regulated matrix element
657: %--------------------------------------
658: \begin{equation}
659: \label{dpsi-lam}
660: \Delta\psi^{(2)}(0)\equiv
661: \psi_{\Lambda}^{(2)}(0) - \psi_{\textrm{DR,}\Lambda}^{(2)}(0).
662: \end{equation}
663: %--------------------------------------
664: Because $\Delta\psi^{(2)}(0)$ involves a change in the ultraviolet
665: cutoff, it is sensitive only to the high-momentum part of the $\bm{p}$
666: integration in the momentum-space definition of $\psi^{(2)}(0)$ in
667: Eq.~(\ref{psi20}). Therefore, we can compute $\Delta\psi^{(2)}(0)$ in
668: perturbation theory. Here, we carry out the computation in lowest
669: (one-loop) order.\footnote{One could, in principle, compute
670: corrections of higher order in $\alpha_s$. Ultimately, the series of
671: these corrections diverges, owing to the renormalon ambiguity that
672: appears in dimensionally-regulated matrix elements. That ambiguity is
673: canceled by a corresponding ambiguity in the NRQCD short-distance
674: coefficients, provided that one computes to the same loop order in both
675: the matrix elements and the short-distance coefficients. See
676: Ref.~\cite{Bodwin:1998mn} for a discussion of this point.} We note,
677: however, that the perturbative calculation of $\Delta\psi^{(2)}(0)$ does
678: not contain all of the corrections that behave as powers of $1/\Lambda$
679: to the difference between $\psi_{\Lambda}^{(2)}(0)$ and
680: $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$. Therefore, it is necessary to
681: take the limit $\Lambda\to\infty$ in order to remove the effects of such
682: power corrections.
683:
684: We begin by using the Bethe-Salpeter equation (or, equivalently, the
685: Schr\"odinger equation) to expose one explicit loop in the wave
686: function. Then, we have
687: \begin{equation}
688: \label{1-loop-renorm}
689: \psi^{(2)}(0)=
690: \int\frac{d^3p}{(2\pi)^3}\,
691: \tilde{I}^{(2)}(\bm{p})
692: \,\widetilde{\psi}(\bm{p}),
693: \end{equation}
694: where $\tilde{I}^{(2)}(\bm{p})$ is the quantity that is
695: represented by the Feynman diagram in Fig.~\ref{fig:Feynman}.
696: %=======================================================================
697: %\begin{figure*}[h]
698: \begin{figure}
699: \includegraphics[width=7cm]{fig1.ps}
700: \caption{Feynman diagram corresponding to $\tilde{I}^{(2)}(\bm{p})$,
701: which is the one-loop renormalization of the operators in
702: Eq.~(\ref{psi20}). The solid lines represent the heavy quark and
703: antiquark, and the dotted line represents the potential between them.}
704: \label{fig:Feynman}
705: \end{figure}
706: %=======================================================================
707: In Fig.~\ref{fig:Feynman}, the dotted line represents the potential.
708: Since the loop integral is dominated by large momenta, we can
709: approximate the potential by the Coulomb part.
710:
711: We emphasize that we can always make this approximation in QCD,
712: irrespective of the use of a potential model. Asymptotic freedom allows
713: one to evaluate the high-momentum loop in Fig.~\ref{fig:Feynman} in
714: perturbation theory. Then, in the Coulomb gauge, the Coulomb-gluon
715: interactions with the heavy quark and antiquark give the result of
716: leading order in $v$, while the transverse-gluon interactions are
717: suppressed as $v^2$ (Ref.~\cite{Bodwin:1994jh}). In the language of
718: NRQCD, $\tilde{I}^{(2)}(\bm{p})$ is the perturbative one-loop
719: renormalization of the operators in Eq.~(\ref{psi20}) in leading order
720: in $\alpha_s$ and $v$.
721:
722: Note that in $\tilde{I}^{(2)}(\bm{p})$ we
723: retain the binding energy of the heavy quarkonium
724: \begin{equation}
725: \epsilon_{\textrm{B}}=-\gamma_{\textrm{B}}^2/m.
726: \label{eps-b}
727: \end{equation}
728: We assign momenta such that the $Q$ and $\bar Q$ each carry half of the
729: binding energy in the rest frame of the heavy quarkonium.
730:
731: $\Delta\psi^{(2)}(0)$ is given by
732: %--------------------------------------
733: \begin{equation}
734: \label{me-corr}
735: \Delta\psi^{(2)}(0)=
736: \int\frac{d^3p}{(2\pi)^3}\,
737: \Delta\tilde{I}^{(2)}(\bm{p})
738: \,\widetilde{\psi}(\bm{p}),
739: \end{equation}
740: %--------------------------------------
741: where
742: \begin{equation}
743: \Delta\tilde{I}^{(2)}(\bm{p})=\tilde{I}_{\Lambda}^{(2)}(\bm{p})
744: -\tilde{I}_{\textrm{DR}}^{(2)}(\bm{p}).
745: \end{equation}
746: Writing out the expression for the diagram in Fig.~\ref{fig:Feynman} for
747: the cases of a hard cutoff and dimensional regularization, we obtain
748: %--------------------------------------
749: \begin{eqnarray}
750: \label{I0p}
751: \Delta\tilde{I}_{\textrm{B}}^{(2)}(\bm{p})&=&
752: 4\pi\alpha_s C_F
753: \int \frac{d^4k}{(2\pi)^4}
754: (\bm{k}+\bm{p})^{2}
755: \frac{i}{\bm{k}^2}
756: \frac{i}{\left[k_0-\frac{\gamma_{\textrm{B}}^2}{2m}
757: -\frac{(\bm{k}+\bm{p})^2}{2m}
758: +i\epsilon\right]}
759: \nonumber\\
760: &&\quad\quad\quad\times
761: \frac{i}{\left[-k_0-\frac{\gamma_{\textrm{B}}^2}{2m}
762: -\frac{(\bm{k}+\bm{p})^2}{2m}
763: +i\epsilon\right]}
764: \left[
765: \frac{\Lambda^2}{(\bm{k}+\bm{p})^2+\Lambda^2}-1_{\rm DR}
766: \right].
767: \end{eqnarray}
768: %--------------------------------------
769: The first term in the numerator brackets in Eq.~(\ref{I0p}) corresponds
770: to the hard cutoff, while the second term ``$1_{\rm DR}$'' corresponds to
771: dimensional regularization. The meaning of the ``$1_{\rm DR}$'' is that it is
772: unity unless it multiplies a scaleless integral (which vanishes in
773: dimensional regularization), in which case it is zero. The subscript
774: ``B'' in $\Delta\tilde{I}^{(2)}_{\rm B}(\bm{p})$ indicates that we have
775: retained the dependence on the binding energy $-\gamma_{\textrm{B}}^2/2m$.
776: Since the integral in Eq.~(\ref{I0p}) is dominated by large momenta, we
777: expect the final result to be insensitive to $\gamma_{\textrm{B}}$. However, in
778: discussions that occur later in this paper, it is illuminating to retain
779: the full $\gamma_{\textrm{B}}$ dependence.
780:
781: The integral over the loop energy $k_0$ in Eq.~(\ref{I0p}) can be carried
782: out by using the residue theorem. The result is
783: %--------------------------------------
784: \begin{equation}
785: \label{dI}
786: \Delta\tilde{I}_{\textrm{B}}^{(2)}(\bm{p})=
787: 4\pi \alpha_s C_F m
788: \int \frac{d^3{k}}{(2\pi)^3}
789: \frac{(\bm{k}+\bm{p})^{2}}
790: {\bm{k}^2 \left[(\bm{k}+\bm{p})^2+\gamma_{\textrm{B}}^2\right]}
791: \left[
792: \frac{\Lambda^2}{(\bm{k}+\bm{p})^2+\Lambda^2}
793: -1_{\rm DR}
794: \right].
795: \end{equation}
796: %--------------------------------------
797: (Note that we also could have reached this expression directly from the
798: momentum-space Schr\"odinger equation.) In the second term in brackets
799: in Eq.~(\ref{dI}), we discard a scaleless integral of the form
800: %--------------------------------------
801: \begin{equation}
802: \int \frac{d^3{k}}{(2\pi)^3}\frac{1_{\rm DR}}{\bm{k}^{2}}.
803: \label{scaless1}
804: \end{equation}
805: %--------------------------------------
806: In the first term in brackets
807: in Eq.~(\ref{dI}), we rewrite the numerator $(\bm{k}+\bm{p})^2$ as
808: $[(\bm{k}+\bm{p})^2+\gamma_{\textrm{B}}^2]-\gamma_{\textrm{B}}^2$ and partial-fraction
809: the $-\gamma_{\textrm{B}}^2$ term. The result of these manipulations is
810: %--------------------------------------
811: \begin{equation}
812: \Delta\tilde{I}^{(2)}_{\textrm{B}}(\bm{p})=
813: \frac{4\pi \alpha_s C_F m}{\Lambda^2-\gamma_{\textrm{B}}^2}
814: \int \frac{d^3{k}}{(2\pi)^3}
815: \frac{1}{\bm{k}^2}
816: \left[
817: \frac{\Lambda^4}{(\bm{k}+\bm{p})^2+\Lambda^2}
818: -
819: \frac{\gamma_{\textrm{B}}^4}{(\bm{k}+\bm{p})^2+\gamma_{\textrm{B}}^2}
820: \right].
821: \label{dIp}
822: \end{equation}
823: %--------------------------------------
824: Evaluation of the integral in Eq.~(\ref{dIp}) is straightforward. The
825: result is
826: %--------------------------------------
827: \begin{equation}
828: \Delta
829: \tilde{I}^{(2)}_{\textrm{B}}(\bm{p})
830: =\frac{\alpha_s C_F m}{|\bm{p}|(\Lambda^2-\gamma_{\textrm{B}}^2)}
831: \left[
832: \Lambda^4 \arctan\left(\frac{|\bm{p}|}{\Lambda}\right)
833: -\gamma_{\textrm{B}}^4
834: \arctan\left(\frac{|\bm{p}|}{\gamma_{\textrm{B}}}\right)
835: \right].
836: \label{Ipm-e}
837: \end{equation}
838: %--------------------------------------
839: As expected, $\tilde{I}^{(2)}_{\textrm{B}}(\bm{p})$ is insensitive to
840: $\gamma_{\textrm{B}}$. Neglecting terms of higher order
841: in $\gamma_{\textrm{B}}^2/\Lambda^2$,
842: we obtain
843: %--------------------------------------
844: \begin{equation}
845: \Delta
846: \tilde{I}^{(2)}_{\textrm{NB}}(\bm{p})
847: \equiv
848: \lim_{\gamma_{\textrm{B}}\to 0}
849: \Delta
850: \tilde{I}^{(2)}_{\textrm{B}}(\bm{p})
851: =
852: \frac{\alpha_s C_F m}{|\bm{p}|}
853: \Lambda^2 \arctan\left(\frac{|\bm{p}|}{\Lambda}\right),
854: \label{Ipm-a}
855: \end{equation}
856: %--------------------------------------
857: where the subscript ``NB'' indicates that we have neglected the binding
858: energy by dropping contributions of higher order in
859: $\gamma_{\textrm{B}}^2/\Lambda^2$.
860: (Note that $\arctan(|\bm{p}|/\gamma_{\textrm{B}})$ is well behaved
861: as $\gamma_{\textrm{B}}\to 0$ and bounded over the entire range of
862: its argument.) In our numerical analyses, we make use of
863: $\Delta\tilde{I}^{(2)}_{\textrm{NB}}(\bm{p})$, rather than
864: $\Delta\tilde{I}^{(2)}_{\textrm{B}}(\bm{p})$, and consistently neglect
865: the binding energy in short-distance (high-momentum) quantities. Of
866: course, the binding-energy dependence in the wave function is retained fully.
867:
868: In our numerical analyses, we solve the Schr\"odinger equation
869: in coordinate space rather than in momentum space. The Fourier
870: transformation of the coordinate-space wave function to momentum space
871: involves an oscillating integrand and, hence, is difficult to evaluate
872: numerically. Therefore, it is convenient to evaluate
873: $\Delta\psi^{(2)}(0)$ in coordinate space:
874: %--------------------------------------
875: \begin{equation}
876: \Delta\psi^{(2)}(0)
877: =\int d^3x\, \Delta I^{(2)}(\bm{x})\psi(\bm{x}),
878: \label{dpsi}
879: \end{equation}
880: %--------------------------------------
881: where
882: \begin{equation}
883: \Delta I^{(2)}(\bm{x})=\int\frac{d^3{p}}{(2\pi)^3}e^{i\bm{p}\cdot\bm{x}}
884: \Delta\tilde{I}^{(2)}(\bm{p})
885: \end{equation}
886: is the Fourier transform of $\Delta\tilde{I}^{(2)}(\bm{p})$. It is a
887: simple matter to evaluate $\Delta I^{(2)}(\bm{x})$ analytically. The
888: results are
889: %--------------------------------------
890: \begin{subequations}
891: \begin{eqnarray}
892: \label{Ix}
893: \Delta I^{(2)}_{\textrm{B}}(\bm{x})
894: &=&
895: \int\frac{d^3{p}}{(2\pi)^3}e^{i\bm{p}\cdot\bm{x}}
896: \Delta\tilde{I}^{(2)}_{\textrm{B}}(\bm{p})
897: = \frac{ \alpha_s C_F m }{4\pi r^2(\Lambda^2-\gamma_{\textrm{B}}^2)}
898: \left(
899: \Lambda^4 e^{-\Lambda r}
900: -\gamma_{\textrm{B}}^4e^{-\gamma_{\textrm{B}} r}
901: \right),\label{Ixb}
902: \\
903: \Delta I^{(2)}_{\textrm{NB}}(\bm{x})
904: &=&
905: \int\frac{d^3{p}}{(2\pi)^3}e^{i\bm{p}\cdot\bm{x}}
906: \Delta\tilde{I}^{(2)}_{\textrm{NB}}(\bm{p})
907: = \frac{ \alpha_s C_F m \Lambda^2}{4\pi r^2}
908: e^{-\Lambda r}.\label{Ixnb}
909: \end{eqnarray}
910: \end{subequations}
911: %--------------------------------------
912: Substituting $\Delta I^{(2)}(\bm{x})$ into Eq.~(\ref{dpsi}), we obtain
913: %--------------------------------------
914: \begin{subequations}
915: \begin{eqnarray}
916: \label{dpsif}
917: \Delta \psi^{(2)}_{\textrm{B}}(0)
918: &=&
919: \frac{\alpha_s C_F m}{\Lambda^2-\gamma_{\textrm{B}}^2}
920: \int_0^\infty
921: \left(
922: \Lambda^4 e^{-\Lambda r}
923: -\gamma_{\textrm{B}}^4e^{-\gamma_{\textrm{B}} r}
924: \right)\frac{R(r)}{\sqrt{4\pi}}\,dr,
925: \\
926: \label{dpsif0}
927: \Delta \psi^{(2)}_{\textrm{NB}}(0)
928: &=&
929: \alpha_s C_F m\Lambda^2
930: \int_0^\infty
931: e^{-\Lambda r}
932: \frac{R(r)}{\sqrt{4\pi}}
933: \,dr.
934: \end{eqnarray}
935: \end{subequations}
936: %--------------------------------------
937: Then, we obtain the dimensionally-regulated matrix element by computing
938: %--------------------------------------
939: \begin{equation}
940: \psi^{(2)}_{\textrm{DR,}\Lambda}(0)=
941: \psi^{(2)}_\Lambda(0)
942: -\Delta\psi^{(2)}_{\textrm{NB}}(0)
943: \label{psiDRL}
944: \end{equation}
945: %--------------------------------------
946: and taking the limit $\Lambda\to\infty$ in order to remove
947: uncompensated power corrections that vanish as $1/\Lambda$:
948: %--------------------------------------
949: \begin{equation}
950: \psi^{(2)}_{\textrm{DR}}(0)=
951: \lim_{\Lambda\to\infty}
952: \psi^{(2)}_{\textrm{DR,}\Lambda}(0).
953: \label{LIMpsiDRL}
954: \end{equation}
955: %--------------------------------------
956:
957: \section{Decomposition of $\bm{\Delta\psi^{(2)}(0)}$}
958: \label{decomposition}
959:
960: It is illuminating to decompose $\Delta\psi^{(2)}(0)$ into products of
961: NRQCD matrix elements times short-distance coefficients.
962: In existing lattice calculations \cite{bks}, the difference between
963: matrix elements in hard-cutoff regularization and those in
964: dimensional regularization is expressed in terms of such a
965: decomposition. We decompose $\Delta\psi^{(2)}(0)$ by applying the
966: method of regions \cite{Beneke:1997zp} to the integration over the
967: wave-function momentum $\bm{p}$ in Eq.~(\ref{me-corr}). We take for
968: $\Delta \tilde{I}^{(2)}(\bm{p})$ the no-binding expression $\Delta
969: \tilde{I}^{(2)}_{\textrm{NB}}(\bm{p})$, which is given in
970: Eq.~(\ref{Ipm-a}).
971:
972: First, we extract the leading term, which is
973: obtained from the $|\bm{p}|^0$ part of
974: $\Delta \tilde{I}^{(2)}_{\textrm{NB}}(\bm{p})$, namely,
975: $\alpha_sC_Fm\Lambda$. It yields a contribution to
976: $\Delta\psi^{(2)}(0)$ that is
977: \begin{equation}
978: \Delta\psi_1^{(2)}(0)=\alpha_s C_F m \Lambda\int \frac{d^3p}{(2\pi)^3}
979: \widetilde{\psi}(\bm{p})
980: =\alpha_sC_Fm\Lambda\psi(0)=2\gamma_{\textrm{C}}\Lambda\psi(0).
981: \end{equation}
982: We identify this as a one-loop short-distance coefficient times
983: the matrix element of leading order in $v$, namely, $\psi(0)$.
984: $\Delta\psi_1^{(2)}(0)$ is linearly divergent in the limit
985: $\Lambda\to\infty$ and cancels the linear divergence in the
986: hard-cutoff matrix element $\psi^{(2)}_{\Lambda}(0)$, resulting in
987: a matrix element $\psi^{(2)}_{\rm DR}(0)$, which, according to the NRQCD
988: $v$-scaling rules, is of order $(m v)^2\psi(0)$. The $v$ scaling of
989: this quantity is verified by the result of an analytic calculation of
990: $\psi^{(2)}_{\textrm{DR}}(0)$ for the case of a pure Coulomb potential
991: in Appendix~\ref{coulomb}.
992:
993: Next we examine the remainder of $\Delta
994: \tilde{I}^{(2)}_{\textrm{NB}}(\bm{p})$, namely,
995: $\alpha_sC_Fm\Lambda[(\Lambda/|\bm{p}|)\,\arctan(|\bm{p}|/\Lambda)-1]$.
996: We decompose the loop integration over $\bm{p}$ into regions of small
997: $|\bm{p}|$, in which $|\bm{p}|\ll \Lambda$, and large $|\bm{p}|$,
998: in which $|\bm{p}|\sim \Lambda$.
999:
1000: In the small-$|\bm{p}|$ region, we expand in powers of $|\bm{p}|$.
1001: The resulting contributions, which we call $\Delta\psi_2^{(2)}(0)$, have
1002: the form
1003: \begin{equation}
1004: \Delta\psi_2^{(2)}(0)=\alpha_s C_F m \int_{\textrm{DR}}
1005: \frac{d^3p}{(2\pi)^3} \widetilde{\psi}(\bm{p}) \sum_{n=3}^\infty
1006: c_n\frac{|\bm{p}|^{n-1}}{\Lambda^{n-2}},
1007: \end{equation}
1008: where $c_n$ is the coefficient of $(|\bm{p}|/\Lambda)^{n}$ in the
1009: power-series expansion of $\arctan(|\bm{p}|/\Lambda)$, and the subscript
1010: $\textrm{DR}$ on the integral indicates that any UV divergence is to be
1011: regulated dimensionally.
1012: $\Delta\psi_2^{(2)}(0)$ corresponds to one-loop short-distance
1013: coefficients times dimensionally-regulated NRQCD matrix elements of
1014: higher order in $v$. The contributions to $\Delta\psi_2^{(2)}(0)$ are of
1015: order $\alpha_s m(mv)^{n-1}/\Lambda^{n-2}$ and, hence, are
1016: suppressed relative to $\psi^{(2)}_{\rm DR}(0)$
1017: by at least one power of $m/\Lambda$ as $\Lambda\to \infty$.
1018:
1019: In the large-$|\bm{p}|$ region, we approximate $\widetilde{\psi}(\bm{p})$
1020: by its asymptotic form at large $|\bm{p}|$, which amounts to neglecting
1021: corrections of higher order in $\alpha_s$ and $mv/\Lambda$. The
1022: asymptotic form of $\widetilde{\psi}(\bm{p})$ is obtained by using the
1023: Bethe-Salpeter equation to expose an explicit loop in the wave function,
1024: as in the derivation of Eq.~(\ref{dI}). The result is
1025: \begin{equation}
1026: \widetilde{\psi}(\bm{p})\sim\int\frac{d^3k}{(2\pi)^3}\widetilde{\psi}(\bm{k})
1027: \frac{8\pi\gamma_{\textrm{C}}}{(\bm{p}-\bm{k})^2\bm{p}^2}
1028: \sim \frac{8\pi\gamma_{\textrm{C}}\psi(0)}{\bm{p}^4}
1029: \equiv \widetilde{\psi}_{\rm asy}(\bm{p}).
1030: \end{equation}
1031: Then, we have for the contribution in the large-$|\bm{p}|$ region
1032: \begin{eqnarray}
1033: \Delta\psi_3^{(2)}(0)
1034: &=&\alpha_s C_F m \Lambda\int \frac{d^3p}{(2\pi)^3}
1035: \widetilde{\psi}_{\rm asy}(\bm{p})
1036: \left[\frac{\Lambda}{|\bm{p}|}\arctan\left(\frac{|\bm{p}|}{\Lambda}\right)
1037: -1\right]\nonumber\\
1038: &=&-(1/2)(\alpha_sC_Fm)^2\psi(0)=-2\gamma_{\textrm{C}}^2\psi(0).
1039: \label{Delta-psi_3}
1040: \end{eqnarray}
1041: $\Delta\psi_3^{(2)}(0)$ has the interpretation of a two-loop short-distance
1042: coefficient times the lowest-order NRQCD matrix element.
1043:
1044: Although the contribution $\Delta\psi_3^{(2)}(0)$ is suppressed as
1045: $\alpha_s/\Lambda$ relative to $\Delta\psi_1^{(2)}(0)$,
1046: $\Delta\psi_3^{(2)}(0)$ can be important numerically because, in
1047: $\psi^{(2)}_{\Lambda,{\rm DR}}(0)$, $\Delta\psi_1^{(2)}(0)$ is canceled
1048: by contributions from $\psi^{(2)}_{\Lambda}(0)$, leaving a small
1049: remainder. As we have remarked, $\psi^{(2)}_{\rm DR}(0)$ itself is
1050: nominally of order $(m v)^2\psi(0)$. Therefore,
1051: $\Delta\psi_3^{(2)}(0)$ is suppressed only as $\alpha_s^2/v^2$ relative
1052: to $\psi^{(2)}_{\rm DR}(0)$. According to the $v$-scaling rules of NRQCD
1053: \cite{Bodwin:1994jh}, $\alpha_s$ is of order $v^2$, and, hence, there is
1054: a slight suppression. It turns out that, for the Cornell potential, the
1055: numerical value of $\Delta\psi_3^{(2)}(0)$ is about $-40\%$ of
1056: $\psi^{(2)}_{\textrm{DR}}(0)$. This suggests that two-loop corrections
1057: to the short-distance coefficients could be important numerically in
1058: converting results of lattice calculations of $\psi^{(2)}(0)$ to
1059: continuum regularization. In the case of a pure Coulomb potential,
1060: $\alpha_s$ is of order $v$, and, so, there is no suppression of
1061: $\Delta\psi_3^{(2)}(0)$ relative to $\psi^{(2)}_{\rm DR}(0)$ at all. We
1062: show in Appendix~\ref{coulomb}, $\psi^{(2)}_{\textrm{DR}}(0)$
1063: satisfies the Gremm-Kapustin relation in the case of a pure Coulomb
1064: potential, while $\psi^{(2)}_{\textrm{DR}}(0)-\Delta\psi_3^{(2)}(0)$ does
1065: not.
1066:
1067: %===========================================================
1068: \section{Direct method of calculation of
1069: $\bm{\psi}^{\bm{(2)}}_{\textrm{DR}}$\label{direct}(0)}
1070: %===========================================================
1071: In the method of calculation that we have outlined, it is necessary to
1072: take the limit of the quantity $\psi^{(2)}_{\textrm{DR,}\Lambda}(0)$ as
1073: $\Lambda$ goes to infinity. $\psi^{(2)}_{\textrm{DR,}\Lambda}(0)$
1074: consists of a difference of $\Lambda$-dependent terms that grow
1075: approximately linearly with $\Lambda$. If one computes at a large enough
1076: value of $\Lambda$ to be near the asymptotic value of
1077: $\psi^{(2)}_{\textrm{DR,}\Lambda}(0)$, then there is a substantial
1078: cancellation between these terms. For example, for the NLO parameters at
1079: $\lambda=1.0$, the ratio
1080: $\psi^{(2)}_{\Lambda}(0)/\psi^{(2)}_{\textrm{DR,}\Lambda}(0)$ takes on
1081: the values 18, 34, and 50 when $\Lambda/m$ equals 10, 20, and 30,
1082: respectively. In numerical calculations, the cancellation in
1083: $\psi^{(2)}_{\textrm{DR,}\Lambda}(0)$ is a significant obstacle to
1084: achieving good accuracy. Therefore, it is desirable to have a direct
1085: method of computation of $\psi^{(2)}_{\textrm{DR}}(0)$ that does not
1086: pass through an intermediate hard-cutoff step. In this section, we
1087: present such a method. We also show how it leads to a simple formula for
1088: the resummation of a class of corrections of higher order in $v$.
1089:
1090: \subsection{Method}
1091:
1092: We begin again by exposing one loop in the matrix element, as in
1093: Fig.~\ref{fig:Feynman}. Now, however, since we are computing the matrix
1094: element itself, not a difference of matrix elements for different
1095: regulators, the loop in Fig.~\ref{fig:Feynman} is not dominated by large
1096: momenta. Therefore, we must retain the complete Cornell potential in the
1097: corresponding expression. Repeating the steps that lead to
1098: Eq.~(\ref{dI}), but for the complete Cornell potential, and using
1099: Eq.~(\ref{1-loop-renorm}), we obtain for the dimensionally-regulated
1100: matrix element
1101: %--------------------------------------
1102: \begin{equation}
1103: \psi^{(2)}_{\textrm{DR}}(0)
1104: =
1105: -m\int\frac{d^3k}{(2\pi)^3}
1106: \frac{\bm{k}^2 1_{\rm DR}}
1107: {\bm{k}^2+\gamma_{\textrm{B}}^2+i\epsilon}
1108: \int\frac{d^3p}{(2\pi)^3}
1109: \widetilde{V}(\bm{k}-\bm{p})\widetilde{\psi}(\bm{p}).
1110: \label{psi2-direct-0}
1111: \end{equation}
1112: It turns out that, in the case of the Cornell potential, the binding
1113: energy is positive. Therefore, it is convenient to define a quantity
1114: $\widetilde{\gamma}_{\textrm{B}}$ as
1115: %--------------------------------------
1116: \begin{equation}
1117: \epsilon_{\textrm{B}}=
1118: -\frac{ \gamma^2_{\textrm{B}} }{m}
1119: =
1120: \frac{ \widetilde{\gamma}^2_{\textrm{B}} }{m}>0.
1121: \label{tilde-gamma-b}
1122: \end{equation}
1123: %--------------------------------------
1124: From the $i\epsilon$ prescription in Eq.~(\ref{psi2-direct-0}), we
1125: see that we can analytically continue from $\gamma_{\textrm{B}}$
1126: positive to $\gamma_{\textrm{B}}$ positive imaginary. Hence, we make the
1127: identification $\gamma_{\textrm{B}}=i\widetilde{\gamma}_{\textrm{B}}$,
1128: where $\widetilde{\gamma}_{\textrm{B}}$ is real and positive.
1129: %--------------------------------------
1130: Discarding the contribution of the scaleless integral
1131: %--------------------------------------
1132: \begin{equation}
1133: \int\frac{d^3k}{(2\pi)^3}
1134: \widetilde{V}(\bm{k}-\bm{p})1_{\rm DR}=
1135: \int\frac{d^3k}{(2\pi)^3}
1136: \widetilde{V}(\bm{k})1_{\rm DR}
1137: \label{scaleless}
1138: \end{equation}
1139: %--------------------------------------
1140: in Eq.~(\ref{psi2-direct-0}), we have
1141: %--------------------------------------
1142: \begin{equation}
1143: \psi^{(2)}_{\textrm{DR}}(0)
1144: =
1145: -m\widetilde{\gamma}^2_{\textrm{B}}
1146: \int\frac{d^3k}{(2\pi)^3}
1147: \int\frac{d^3p}{(2\pi)^3}
1148: \frac{\widetilde{V}(\bm{k}-\bm{p})
1149: \widetilde{\psi}(\bm{p})
1150: }{\bm{k}^2-\widetilde{\gamma}^2_{\textrm{B}}+i\epsilon}
1151: \label{psi2-direct-mom}.
1152: \end{equation}
1153: %--------------------------------------
1154: Now we note that $\widetilde{\psi}(\bm{p})$ satisfies the momentum-space
1155: Schr\"odinger equation
1156: \begin{equation}
1157: (\bm{k}^2-\widetilde{\gamma}^2_{\textrm{B}})\widetilde{\psi}(\bm{k})
1158: =-\int\frac{d^3p}{(2\pi)^3}
1159: m\widetilde{V}(\bm{k}-\bm{p})\widetilde{\psi}(\bm{p}).
1160: \label{schro-mom}
1161: \end{equation}
1162: Therefore, Eq.~(\ref{psi2-direct-mom}) can be written as
1163: \begin{equation}
1164: \psi^{(2)}_{\textrm{DR}}(0)
1165: =\widetilde{\gamma}^2_{\textrm{B}}
1166: \int\frac{d^3k}{(2\pi)^3}
1167: \widetilde{\psi}(\bm{k})
1168: =\widetilde{\gamma}^2_{\textrm{B}}\psi(0).
1169: \label{psi2-direct-final}
1170: \end{equation}
1171: %--------------------------------------
1172:
1173: The Gremm-Kapustin relation \cite{Gremm:1997dq} follows from the NRQCD
1174: equations of motion at leading order nontrivial in $v$. It states that
1175: \begin{equation}
1176: \langle v^2 \rangle=\frac{\psi^{(2)}(0)}{m_c^2\psi(0)}=
1177: \frac{\epsilon_{\textrm{B}}}{m_c}+{\cal O}(v^4).
1178: \label{gremm-kapustin}
1179: \end{equation}
1180: Using the definition of $\widetilde{\gamma}^2_{\textrm{B}}$ in
1181: Eq.~(\ref{tilde-gamma-b}), we see that our result in
1182: Eq.~(\ref{psi2-direct-final}) is precisely the Gremm-Kapustin relation
1183: for the potential model with mass $m=m_c$. This is not surprising, since
1184: a potential model can be derived from NRQCD at leading order in $v$
1185: (Ref.~\cite{Brambilla:1999xf}) and, hence, implicitly respects the
1186: NRQCD equations of motion at leading order in $v$.
1187:
1188: \subsection{Resummation}
1189:
1190: We note that the result in Eq.~(\ref{psi2-direct-final}) can easily be
1191: generalized to the case of NRQCD operator matrix elements involving
1192: additional powers of $\bm{\nabla}^2$, such as
1193: \begin{subequations}
1194: \begin{eqnarray}
1195: \psi^{(2n)}(0)&\equiv&
1196: \int\frac{d^3p}{(2\pi)^3}
1197: \bm{p}^{2n}
1198: \widetilde{\psi}(\bm{p})
1199: =\frac{1}{\sqrt{2N_c}}
1200: \langle 0|\chi^\dagger\left(-\bm{\nabla}^2\right)^n\psi|H({}^1S_0)\rangle,
1201: \\
1202: \bm{\epsilon}\psi^{(2n)}(0)&\equiv&\bm{\epsilon}\int\frac{d^3p}{(2\pi)^3}
1203: \bm{p}^{2n}
1204: \widetilde{\psi}(\bm{p})
1205: =\frac{1}{\sqrt{2N_c}}
1206: \langle 0|\chi^\dagger\bm{\sigma}\left(-\bm{\nabla}^2\right)^n
1207: \psi|H({}^3S_1)\rangle,
1208: \end{eqnarray}
1209: \end{subequations}
1210: where $n$ is an integer. As we have remarked earlier, such operator matrix
1211: elements are usually written in terms of the covariant derivative
1212: $\bm{D}$ (Ref.~\cite{Bodwin:1994jh}), rather than $\bm{\nabla}$.
1213: However, in the Coulomb gauge, the difference between $\bm{D}$ and
1214: $\bm{\nabla}$ is suppressed as $v$ (Ref.~\cite{Bodwin:1994jh}).
1215:
1216:
1217: Repeating the steps that lead to Eq.~(\ref{psi2-direct-0}), we obtain
1218: \begin{equation}
1219: \psi^{(2n)}_{\textrm{DR}}(0)
1220: =
1221: -m\int\frac{d^3k}{(2\pi)^3}
1222: \frac{\bm{k}^{2n} 1_{\rm DR}}
1223: {\bm{k}^2-\widetilde{\gamma}_{\textrm{B}}^2+i\epsilon}
1224: \int\frac{d^3p}{(2\pi)^3}
1225: \widetilde{V}(\bm{k}-\bm{p})\widetilde{\psi}(\bm{p}).
1226: \label{psi2n-direct-0}
1227: \end{equation}
1228:
1229: In Eq.~(\ref{psi2n-direct-0}), we write
1230: \begin{equation}
1231: \frac{\bm{k}^{2n}}{\bm{k}^2-\widetilde{\gamma}_{\textrm{B}}^2+i\epsilon}
1232: =\bm{k}^{2n-2}+\widetilde{\gamma}_{\textrm{B}}^2\bm{k}^{2n-4}+\cdots
1233: +\frac{\widetilde{\gamma}_{\textrm{B}}^{2n}}
1234: {\bm{k}^2-\widetilde{\gamma}_{\textrm{B}}^2+i\epsilon}
1235: \end{equation}
1236: and discard scaleless integrals of the form
1237: \begin{equation}
1238: \int\frac{d^3k}{(2\pi)^3}
1239: \widetilde{V}(\bm{k}-\bm{p})\bm{k}^{2n'}1_{\rm DR}
1240: =\int\frac{d^3k}{(2\pi)^3}
1241: \widetilde{V}(\bm{k})(\bm{k}+\bm{p})^{2n'}1_{\rm DR},
1242: \end{equation}
1243: where $n'$ is an integer. The result is
1244: \begin{equation}
1245: \psi^{(2n)}_{\textrm{DR}}(0)
1246: =-m\widetilde{\gamma}^{2n}_{\textrm{B}}
1247: \int\frac{d^3k}{(2\pi)^3}
1248: \int\frac{d^3p}{(2\pi)^3}
1249: \frac{\widetilde{V}(\bm{k}-\bm{p})
1250: \widetilde{\psi}(\bm{p})
1251: }{\bm{k}^2-\widetilde{\gamma}^2_{\textrm{B}}+i\epsilon}
1252: =\widetilde{\gamma}^{2n}_{\textrm{B}}\psi(0),
1253: \label{psi2n-final}
1254: \end{equation}
1255: where we have used the momentum-space Schr\"odinger equation
1256: [Eq.~(\ref{schro-mom})].
1257: The expression in Eq.~(\ref{psi2n-final}) can be thought of as a
1258: generalized Gremm-Kapustin relation that holds in a potential model.
1259:
1260: Now, suppose that there are contributions to a decay or production
1261: amplitude, computed in NRQCD, of the form
1262: \begin{equation}
1263: A^{(2n)}=\frac{1}{n!}H^{(2n)}(0)\psi_{\textrm{DR}}^{(2n)}(0).
1264: \end{equation}
1265: Here, $(1/n!)H^{(2n)}(0)$
1266: is an NRQCD short-distance coefficient that is defined by
1267: \begin{equation}
1268: H^{(2n)}(0)=\left.
1269: \left(\frac{\partial} {\partial \bm{p}^2}\right)^{n}
1270: H(\bm{p}^2)
1271: \right|_{\bm{p}=0},
1272: \end{equation}
1273: where $H^{(2n)}(\bm{p}^2)$ is the hard-scattering amplitude for the
1274: process, evaluated in the quarkonium rest frame at $Q$ or $\bar Q$
1275: momentum squared $\bm{p}^2$.
1276: We can use Eq.~(\ref{psi2n-final}) to resum the contributions $A^{(2n)}$:
1277: \begin{equation}
1278: \sum_n A^{(2n)}=\sum_n \frac{1}{n!}H^{(2n)}(0)\psi_{\textrm{DR}}^{(2n)}(0)
1279: =\sum_n \frac{1}{n!}H^{(2n)}(0)\widetilde{\gamma}^{2n}_{\textrm{B}}\psi(0)
1280: =H(\widetilde{\gamma}^2_{\textrm{B}})\psi(0).
1281: \end{equation}
1282:
1283: %===========================================================
1284: \section{Numerical results and discussion \label{numerical}}
1285: %===========================================================
1286: We now use the hard-cutoff method of Secs.~\ref{cutoff} and \ref{difference}
1287: and the direct method of Sec.~\ref{direct} to compute
1288: $\psi_{\textrm{DR}}^{(2)}(0)$.
1289:
1290: We begin with the hard-cutoff method. We substitute the $1S$-state
1291: coordinate-space radial wave function that we obtain by integrating the
1292: Schr\"odinger equation numerically into Eqs.~(\ref{psi-lamr}) and
1293: (\ref{dpsif0}) and carry out the integrations over $r$ numerically. We
1294: then use Eq.~(\ref{psiDRL}) to compute
1295: $\psi^{(2)}_{\textrm{DR,}\Lambda}(0)$. The results are shown in
1296: Fig.~\ref{fig:me}.
1297: %=======================================================================
1298: %\begin{figure*}[h]
1299: \begin{figure}
1300: \begin{tabular}{cc}
1301: \includegraphics[width=8.0cm]{fig2a.ps}&
1302: \includegraphics[width=8.0cm]{fig2b.ps}\\[-2.0ex]
1303: \end{tabular}
1304: \caption{\label{fig:me}
1305: $\psi_\Lambda^{(2)}(0)$ and $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)\equiv%
1306: \psi^{(2)}_{\Lambda}(0)-\Delta\psi^{(2)}_{\textrm{NB}}(0)$ as a function
1307: of $\Lambda/m$. The left figure corresponds to the LO
1308: potential-model parameters of Table~\ref{table:ma-lo} for $\lambda=0.6$,
1309: $0.7$, and $0.8$. The right figure corresponds to the NLO
1310: potential-model parameters of Table~\ref{table:ma-nlo} for
1311: $\lambda=0.9$, $1.0$, and $1.1$. In each figure, the three curves that
1312: are nearly linear are $0.1\times \psi_\Lambda^{(2)}(0)$, and the three
1313: curves that reach a plateau at large $\Lambda/m$ are
1314: $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$.
1315: }
1316: \end{figure}
1317: %=======================================================================
1318: As can be seen, $\psi_\Lambda^{(2)}(0)$ grows nearly linearly with
1319: $\Lambda$, as is expected from the linear ultraviolet divergence that it
1320: contains. $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$ is the result of a
1321: substantial cancellation between $\psi^{(2)}_{\Lambda}(0)$ and
1322: $\Delta\psi^{(2)}_{\textrm{NB}}(0)$, at the level of about one part in
1323: 50 at $\Lambda/m=30$ for the NLO parameters. From Fig.~\ref{fig:me},
1324: it is apparent that $\psi_{\textrm{DR},\Lambda}^{(2)}(0)$ reaches a
1325: plateau at large $\Lambda$. $\psi_{\textrm{DR},\Lambda}^{(2)}(0)$
1326: deviates from its asymptotic value by an amount that is of order
1327: $1/\Lambda$. It is the value of $\psi_{\textrm{DR},\Lambda}^{(2)}(0)$ at
1328: the plateau that corresponds to the dimensionally-regulated matrix
1329: element [Eq.~(\ref{LIMpsiDRL})]. A fit of the LO-parameter results for
1330: $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$ at $\lambda=0.6$, $0.7$, and $0.8$
1331: to a constant plus a term that is proportional to $1/\Lambda$ yields
1332: $\psi_{\textrm{DR},\Lambda}^{(2)}(0) =0.123$~GeV$^{7/2}$,
1333: $0.108$~GeV$^{7/2}$, and $0.095$~GeV$^{7/2}$, respectively. A similar fit
1334: to the NLO-parameter results at $\lambda=0.9$, $1.0$, and $1.1$ yields
1335: $\psi_{\textrm{DR},\Lambda}^{(2)}(0) =0.144$~GeV$^{7/2}$, $0.124$~GeV$^{7/2}$,
1336: and $0.107$~GeV$^{7/2}$, respectively.
1337:
1338: Next we consider the direct method for calculating
1339: $\psi_{\textrm{DR}}^{(2)}(0)$. Substituting the LO potential-model
1340: parameters into Eq.~(\ref{psi2-direct-final}), we obtain
1341: $\psi_{\textrm{DR}}^{(2)}(0)=0.124$~GeV$^{7/2}$, $0.109$~GeV$^{7/2}$,
1342: and $0.096$~GeV$^{7/2}$ at $\lambda=0.6$, $0.7$, and $0.8$,
1343: respectively. The results for the NLO-parameters are
1344: $\psi_{\textrm{DR}}^{(2)}(0)=0.146$~GeV$^{7/2}$, $0.127$~GeV$^{7/2}$,
1345: and $0.109$~GeV$^{7/2}$ at $\lambda=0.9$, $1.0$, and $1.1$,
1346: respectively. These results for $\psi_{\textrm{DR}}^{(2)}(0)$ are
1347: in good agreement with those from the extrapolation of
1348: $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$ to $\Lambda=\infty$ in the
1349: hard-cutoff method.
1350:
1351: We consider the NLO value for $\psi(0)$ to be slightly more reliable
1352: than the LO value. Therefore, we use the NLO parameters to compute our
1353: central value for $\psi_{\textrm{DR}}^{(2)}(0)$. We use the LO
1354: parameters to give an indication of the uncertainty in
1355: $\psi_{\textrm{DR}}^{(2)}(0)$ that arises from the uncertainty in
1356: $\psi(0)$. As we have already discussed in Sec.~\ref{model}, the
1357: potential-model values of the string tension that derive from the
1358: NLO parameters for $\lambda=1.0$ and $1.1$ span the range of lattice
1359: values for the string tension. The potential-model values of $\alpha_s$
1360: that derive from the NLO parameters for $\lambda=1.0$
1361: and $1.1$ are somewhat larger than the lattice values for the fixed parameter
1362: $\alpha_s$, but are compatible with the value of the running $\alpha_s$
1363: at the scale of the heavy-quark momentum $m_cv$. We consider the lattice
1364: value of the string tension to be more relevant than the lattice value
1365: of the fixed parameter $\alpha_s$, since the latter does not take into
1366: account the running of $\alpha_s$ in QCD. Therefore, we determine the
1367: central value of $\psi_{\textrm{DR}}^{(2)}(0)$ by taking the average of
1368: the values of $\psi_{\textrm{DR}}^{(2)}(0)$ for the NLO parameters at
1369: $\lambda=1.0$ and $1.1$. We take the difference of these values as the
1370: uncertainty that is attributable to the uncertainty in the
1371: potential-model parameters. We take the difference between the average
1372: of the values of $\psi_{\textrm{DR}}^{(2)}(0)$ for the NLO parameters at
1373: $\lambda=1.0$ and $1.1$ and the average of the values of
1374: $\psi_{\textrm{DR}}^{(2)}(0)$ for the LO parameters at $\lambda=0.7$ and
1375: $0.8$ to be the uncertainty that is attributable to the
1376: uncertainty in $\psi(0)$. We add these uncertainties in quadrature. We
1377: include an additional $30\%$ uncertainty to account for the fact that
1378: our potential model neglects terms of relative order $v^2\approx
1379: 0.3$. Using the direct-method values for $\psi_{\textrm{DR}}^{(2)}(0)$,
1380: we obtain
1381: \begin{equation}
1382: \psi_{\textrm{DR}}^{(2)}(0)= 0.118\pm 0.024\pm 0.035~\hbox{GeV}^{7/2},
1383: \label{psi2-final}
1384: \end{equation}
1385: where the second uncertainty arises from the $v^2$ error. This is the
1386: dominant source of error.
1387:
1388: The effects of the uncertainty in the value of $\psi(0)$ tend to cancel
1389: in the ratio $\psi_{\textrm{DR}}^{(2)}(0)/\psi(0)$. We compute this
1390: ratio for the LO-parameter and NLO-parameter direct-method results and
1391: use the same method for determining the central value and the
1392: uncertainties that we described above for $\psi_{\textrm{DR}}^{(2)}(0)$.
1393: The result is
1394: \begin{equation}
1395: \psi_{\textrm{DR}}^{(2)}(0)/\psi(0)=m_c^2v^2=0.50\pm 0.09\pm 0.15~{\rm
1396: GeV}^2,
1397: \label{final-ratio}
1398: \end{equation}
1399: where, again, the second uncertainty arises from the $v^2$ error.
1400: Taking $m_c=1.4$~GeV in Eq.~(\ref{final-ratio}),
1401: we have $\langle v^2 \rangle=0.25\pm 0.05\pm 0.08 $, which is in good
1402: agreement with expectations from the NRQCD $v$-scaling rules.
1403:
1404: We can now address the numerical importance of the two-loop
1405: contribution to $\Delta\psi^{(2)}(0)$
1406: that is given in Eq.~(\ref{Delta-psi_3}).
1407: Taking the average of the values of $\gamma_{\textrm{C}}$ for
1408: $\lambda=1.0$ and $1.1$ in Table~\ref{table:ma-nlo} and using the NLO
1409: value $\psi(0)=0.23629$~GeV$^{3/2}$, we find that
1410: $\Delta\psi_3^{(2)}(0)=-0.048$~GeV$^{7/2}$. This is about $-40\%$ of the
1411: value of $\psi_{\textrm{DR}}^{(2)}(0)$ in Eq.~(\ref{psi2-final}).
1412:
1413: The $v$-scaling rules of NRQCD state that the hard-cutoff matrix
1414: element $\psi_{\Lambda}^{(2)}(0)$, evaluated at $\Lambda\sim
1415: mv$, should be of order $m_c^2v^2\psi(0)$. Taking $m_c=1.4$~GeV,
1416: $v^2=0.3$, and $\psi(0)=0.23629$~GeV$^{3/2}$, we obtain
1417: $m_c^2v^2\psi(0)\approx 0.14$~GeV$^{7/2}$, which is in reasonably
1418: good agreement with the value of $\psi_{\Lambda}^{(2)}(0)$ at
1419: $\Lambda=m_c v\approx m$ for the NLO results at $\lambda=1.0$ and
1420: $1.1$ in Fig.~\ref{fig:me}.
1421:
1422: In lattice determinations of $\psi_{\textrm{DR},\Lambda}^{(2)}(0)$,
1423: the lattice ultraviolet cutoff, which is of order $\pi$ divided by the
1424: lattice spacing, corresponds approximately to the hard cutoff $\Lambda$.
1425: Existing lattice determinations of $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$
1426: (Ref.~\cite{bks}) have been carried out in the vicinity of
1427: $\Lambda=m_c$. This value of $\Lambda$ is at the boundary of the region
1428: in which asymptotic freedom allows one to evaluate quantities in QCD in
1429: perturbation theory. However, as can be seen from Fig.~\ref{fig:me},
1430: $\Lambda=m_c$ is far from the region in which
1431: $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$ approaches its asymptotic value.
1432: Apparently, at $\Lambda=m_c$, power corrections of order $m_cv/\Lambda$
1433: are still important. Therefore, we expect that the lattice
1434: determinations of $\psi_{\textrm{DR}}^{(2)}(0)$ contain large
1435: $1/\Lambda$ errors.
1436:
1437: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1438: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1439: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1440: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1441: %===========================================================
1442: \section{Summary \label{conclusions}}
1443:
1444: In this paper we have presented two methods for calculating NRQCD matrix
1445: elements that are proportional to $\psi^{(2)}(0)$, the negative of the
1446: second derivative of the wave function at the origin. These matrix
1447: elements enter into the relativistic corrections for the decay and
1448: production of heavy-quarkonium states. The matrix elements are linearly
1449: ultraviolet divergent, and, hence, they must be regulated. We compute
1450: the matrix elements in dimensional regularization, since that is the
1451: regularization that is used most commonly in phenomenology.
1452:
1453: One method that we have presented makes use of a hard-cutoff regulator
1454: as an intermediate step, then employs perturbation theory to compute the
1455: difference between the hard-cutoff matrix element and a
1456: dimensionally-regulated matrix element. This method is quite general, in
1457: that it requires knowledge only of the $Q\bar Q$ wave function in the
1458: quarkonium state. In principle, it could be applied to wave functions
1459: that are determined directly by lattice methods. It involves computing
1460: cancelling quantities in the limit in which the hard cutoff is taken to
1461: infinity. This method has the disadvantage that the limiting procedure and
1462: cancellation make it difficult to achieve adequate numerical accuracy.
1463: In the case of a pure Coulomb potential, we have obtained analytic
1464: expressions for the quantities that enter into the hard-cutoff method.
1465: The result for the matrix elements agrees with the Gremm-Kapustin relation
1466: in that case.
1467:
1468: The second method that we have presented involves a computation of the
1469: matrix elements directly in dimensional regularization. The method is
1470: specific to potential models. It yields the Gremm-Kapustin
1471: relation for the matrix elements. This is not surprising, since the
1472: Gremm-Kapustin relation is based on the NRQCD equations of motion at
1473: leading-order in $v$, and the potential model respects those equations of
1474: motion. This method is easily generalized to the
1475: computation of matrix elements that are proportional to higher powers of
1476: $\bm{\nabla}^2$ acting on the wave function at the origin. We have used
1477: the results for such matrix elements to write a formula that resums
1478: certain contributions of higher order in $v$ to NRQCD amplitudes.
1479:
1480: We have used both the direct method and the hard-cutoff method to
1481: evaluate the dimensionally-regulated quantity
1482: $\psi_{\textrm{DR}}^{(2)}(0)$ for the $J/\psi$ (or the $\eta_c$) in the
1483: Cornell potential model. Since the Cornell potential model contains no
1484: spin dependence, we do not distinguish between the $J/\psi$ and $\eta_c$
1485: matrix elements. If the potential itself were exact, then the
1486: potential model would reproduce QCD up to corrections of relative order
1487: $v^2$. Existing lattice measurements of the static $Q\bar Q$ potentials
1488: yield values for the string tension. In order to estimate errors from
1489: our choice of potential, we have carried out computations for sets of
1490: Cornell-potential parameters that bracket the lattice values of the
1491: string tension.
1492:
1493: The two methods for computing $\psi_{\textrm{DR}}^{(2)}(0)$ yield
1494: results that agree well numerically. Our final result, including
1495: estimates of uncertainties, is given in Eq.~(\ref{psi2-final}). The
1496: first error estimate arises from the uncertainty in the
1497: potential-model parameters and from the uncertainty in the value of the
1498: wave function at the origin that is obtained from the leptonic width of
1499: the $J/\psi$. The second error estimate accounts for a $30\%$
1500: uncertainty that arises from the fact that we have neglected corrections
1501: of relative order $v^2$. This is the dominant source of error. The
1502: effects of the uncertainty in the wave function at the origin tend to
1503: cancel in the ratio of matrix elements
1504: $\psi^{(2)}_{\textrm{DR}}(0)/\psi(0)$ in Eq.~(\ref{final-ratio}). From
1505: the ratio in Eq.~(\ref{final-ratio}), we estimate that $\langle
1506: v^2\rangle=0.25\pm 0.05\pm 0.08$, which is in good agreement with the
1507: $v$-scaling rules of NRQCD.
1508:
1509: From our analysis, it is clear that, in the hard-cutoff method, there
1510: are large corrections of order $m_cv/\Lambda$, even at cutoffs
1511: $\Lambda\approx m_c$. This implies that existing lattice computations of
1512: $\psi^{(2)}_{\textrm{DR}}(0)$ (Ref.~\cite{bks}) contain large errors
1513: that arise from such power corrections.
1514:
1515: We also have identified a large contribution to the difference between
1516: the dimensionally-regulated matrix element $\psi^{(2)}_{\textrm{DR}}(0)$
1517: and the hard-cutoff matrix element $\psi^{(2)}_{\Lambda}(0)$ that arises
1518: from a two-loop contribution to a particular short-distance coefficient
1519: [Eq.~(\ref{Delta-psi_3})]. That contribution is suppressed only as
1520: $\alpha_s^2/v^2$ relative to $\psi^{(2)}_{\textrm{DR}}(0)$ and is about
1521: $-40\%$ of $\psi^{(2)}_{\textrm{DR}}(0)$ in the case of the $J/\psi$.
1522: Therefore, the two-loop contribution to this short-distance coefficient
1523: may be important numerically when one converts a lattice-regulated value
1524: of $\psi^{(2)}(0)$ to a dimensionally-regulated value.
1525:
1526: %===========================================================
1527:
1528:
1529:
1530: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1531: % If in two-column mode, this environment will change to single-column
1532: % format so that long equations can be displayed. Use
1533: % sparingly.
1534: %\begin{widetext}
1535: % put long equation here
1536: %\end{widetext}
1537: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1538:
1539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1540: % Surround figure environment with turnpage environment for landscape
1541: % figure
1542: % \begin{turnpage}
1543: % \begin{figure}
1544: % \includegraphics{}%
1545: % \caption{\label{}}
1546: % \end{figure}
1547: % \end{turnpage}
1548: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1549:
1550: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1551: % tables should appear as floats within the text
1552: % Here is an example of the general form of a table:
1553: % Fill in the caption in the braces of the \caption{} command. Put the label
1554: % that you will use with \ref{} command in the braces of the \label{} command.
1555: % Insert the column specifiers (l, r, c, d, etc.) in the empty braces of the
1556: % \begin{tabular}{} command.
1557: % The ruledtabular enviroment adds doubled rules to table and sets a
1558: % reasonable default table settings.
1559: % Use the table* environment to get a full-width table in two-column
1560: % Add \usepackage{longtable} and the longtable (or longtable*}
1561: % environment for nicely formatted long tables. Or use the [H]
1562: % placement option to break a long table (with less control than
1563: % in longtable).
1564: % \begin{table}%[H] add [H] placement to break table across pages
1565: % \caption{\label{}}
1566: % \begin{ruledtabular}
1567: % \begin{tabular}{}
1568: % Lines of table here ending with \\
1569: % \end{tabular}
1570: % \end{ruledtabular}
1571: % \end{table}
1572: % Surround table environment with turnpage environment for landscape
1573: % table
1574: % \begin{turnpage}
1575: % \begin{table}
1576: % \caption{\label{}}
1577: % \begin{ruledtabular}
1578: % \begin{tabular}{}
1579: % \end{tabular}
1580: % \end{ruledtabular}
1581: % \end{table}
1582: % \end{turnpage}
1583: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1584:
1585: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1586: % Specify following sections are appendices. Use \appendix* if there
1587: % only one appendix.
1588: %\appendix
1589: %\section{}
1590: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1591:
1592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1593: % If you have acknowledgments, this puts in the proper section head.
1594:
1595: \begin{acknowledgments}
1596: % put your acknowledgments here.
1597: We wish to thank Eric Braaten for useful discussions.
1598: The research of G.T.B. in the High Energy Physics Division at
1599: Argonne National Laboratory is supported by the U.~S.~Department of
1600: Energy, Division of High Energy Physics, under Contract No.W-31-109-ENG-38.
1601: The work of J.L. is supported by the Korea Research Foundation
1602: under Grant No. KRF-2004-015-C00092.
1603: The work of D.K. is supported in part by the Seoul Science
1604: Fellowship of Seoul Metropolitan Government.
1605: J.L. thanks the High Energy Theory Group at Argonne National Laboratory
1606: for its hospitality.
1607: \end{acknowledgments}
1608: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1609: \appendix
1610: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1611: %===========================================================
1612: \section{Test of the method with a Coulomb wave function \label{coulomb}}
1613: %===========================================================
1614:
1615: In this appendix, we apply our methods for calculating
1616: $\psi_{\textrm{DR}}^{(2)}(0)$ to the case of the wave function for
1617: a pure Coulomb potential. In this case, we can evaluate all of the
1618: relevant expressions analytically. Ultimately, we compare our results
1619: with the prediction of the Gremm-Kapustin relation.
1620:
1621: Coulomb wave functions in coordinate space and momentum space are
1622: %--------------------------------------
1623: \begin{subequations}
1624: \label{coulomb-wavefunction}
1625: \begin{eqnarray}
1626: \psi_{\textrm{C}}(\bm{x})&=&\psi_{\textrm{C}}(0)e^{-\gamma_{\textrm{C}}r},
1627: \label{coulomb-wavefunction-coor}\\
1628: \widetilde{\psi}_{\textrm{C}}(\bm{p})&=&
1629: \int d^3 x e^{-i\bm{p}\cdot\bm{x}}\psi_{\textrm{C}}(\bm{x})
1630: =8\pi\psi_{\textrm{C}}(0)
1631: \frac{\gamma_{\textrm{C}}}{(\bm{p}^2+\gamma^2_{\textrm{C}})^2},
1632: \label{coulomb-wavefunction-mom}
1633: \end{eqnarray}
1634: \end{subequations}
1635: %--------------------------------------
1636: where $\gamma_{\textrm{C}}$ is defined in Eq.~(\ref{gamma-C}). The
1637: corresponding binding energy is
1638: \begin{equation}
1639: \epsilon_{\textrm{B}}=-\gamma^2_{\textrm{C}}/m.
1640: \label{coul-binding}
1641: \end{equation}
1642: It then follows from Eq.~(\ref{eps-b}) that
1643: \begin{equation}
1644: \gamma_{\textrm{B}}=\gamma_{\textrm{C}}.
1645: \end{equation}
1646:
1647: We first carry out the calculation of $\psi_{\textrm{DR}}^{(2)}(0)$ in
1648: the hard-cutoff method. Substituting the Coulomb wave function into
1649: either Eq.~(\ref{psi-lam}) or Eq.~(\ref{psi-lamr}), we obtain the
1650: hard-cutoff matrix element $\psi^{(2)}_{\Lambda}(0)$:
1651: %--------------------------------------
1652: \begin{equation}
1653: \psi^{(2)}_{\Lambda}(0)
1654: =\Lambda^2 \psi_{\textrm{C}}(0)
1655: \left[ 1-\frac{\Lambda^2}{(\Lambda+\gamma_{\textrm{C}})^2} \right].
1656: \end{equation}
1657: %--------------------------------------
1658: The calculation of $\psi^{(2)}_{\Lambda}(0)$ can be carried out by using
1659: Eqs.~(\ref{dpsi}) and (\ref{Ix}). For the difference between the
1660: hard-cutoff and dimensionally-regulated matrix elements, we have
1661: %--------------------------------------
1662: \begin{equation}
1663: \Delta\psi^{(2)}_{\textrm{B}}(0)=
1664: \psi_{\textrm{C}}(0)
1665: \frac{2\gamma_{\textrm{C}}}{\Lambda^2-\gamma^2_{\textrm{B}}}
1666: \left(\frac{\Lambda^4}{\Lambda+\gamma_{\textrm{C}}}
1667: -\frac{\gamma^4_{\textrm{B}}}
1668: {\gamma_{\textrm{B}}+\gamma_{\textrm{C}}} \right),
1669: \label{dpsi2-C}
1670: \end{equation}
1671: %--------------------------------------
1672: where we have used the definition (\ref{gamma-C}) to replace the
1673: prefactor $\alpha_s C_F m$ with $2\gamma_{\textrm{C}}$. In the limit
1674: $\gamma_{\textrm{B}}\to 0$, Eq.~(\ref{dpsi2-C}) becomes
1675: %--------------------------------------
1676: \begin{equation}
1677: \label{dpsi2-C-approx}
1678: \Delta\psi^{(2)}_{\textrm{NB}}(0)
1679: =
1680: \Lambda^2\psi_{\textrm{C}}(0)
1681: \frac{2\gamma_{\textrm{C}}}
1682: {\Lambda+\gamma_{\textrm{C}}}.
1683: \end{equation}
1684: %--------------------------------------
1685: It follows from Eq.~(\ref{psiDRL}) that
1686: $\psi_{\textrm{DR,}\Lambda}^{(2)}(0)$ is given by
1687: \begin{equation}
1688: \psi_{\textrm{DR,}\Lambda}^{(2)}(0)
1689: =
1690: \psi_{\Lambda}^{(2)}(0)
1691: -
1692: \Delta\psi_{\textrm{NB}}^{(2)}(0)
1693: =-\gamma^2_{\textrm{C}}\psi_{\textrm{C}}(0)
1694: \times \left(\frac{\Lambda}{\Lambda+\gamma_{\textrm{C}}}\right)^2.
1695: \label{psi2-dr-LO}
1696: \end{equation}
1697: %--------------------------------------
1698: Taking the limit $\Lambda \to \infty$,
1699: we obtain the dimensionally-regulated matrix element
1700: $\psi_{\textrm{DR}}^{(2)}(0)$:
1701: %--------------------------------------
1702: \begin{equation}
1703: \psi_{\textrm{DR}}^{(2)}(0)
1704: =\lim_{\Lambda\to \infty}\psi_{\textrm{DR,}\Lambda}^{(2)}(0)
1705: =-\gamma^2_{\textrm{C}}\psi_{\textrm{C}}(0).
1706: \label{psi2-COULOMB-ans}
1707: \end{equation}
1708: %--------------------------------------
1709: We note that the deviations from the asymptotic value go as
1710: $\gamma_{\textrm{C}}/\Lambda\sim mv/\Lambda$ as $\Lambda\to\infty$.
1711: Incidentally, had we retained the effects of the binding energy
1712: by using $\Delta\psi_{\textrm{B}}^{(2)}(0)$ instead of
1713: $\Delta\psi_{\textrm{NB}}^{(2)}(0)$, then we would have obtained
1714: \begin{equation}
1715: \psi_{\textrm{DR,}\Lambda}^{(2)}(0)=\psi_{\Lambda}^{(2)}(0)
1716: -
1717: \Delta\psi_{\textrm{B}}^{(2)}(0)
1718: =-\gamma^2_{\textrm{C}}\psi_{\textrm{C}}(0),
1719: \end{equation}
1720: %--------------------------------------
1721: where we have made use of the Coulomb-potential relation
1722: $\gamma_{\textrm{B}}=\gamma_{\textrm{C}}$. In this case, the $\Lambda$
1723: dependence would have vanished. However, this simplification is a
1724: special property of the pure Coulomb case that arises from the fact that
1725: the Coulomb-gluon exchange in $\Delta\psi_{\textrm{B}}^{(2)}(0)$ is, in
1726: this special case, an interaction of the complete potential.
1727:
1728: The result in Eq.~(\ref{psi2-COULOMB-ans}) agrees with the result from
1729: the direct method in Eq.~(\ref{psi2-direct-final}) and with the
1730: Gremm-Kapustin relation (\ref{gremm-kapustin}). We note that, had we
1731: failed to include the two-loop correction to $\Delta\psi^{(2)}(0)$ in
1732: Eq.~(\ref{Delta-psi_3}), then we would have obtained
1733: $\psi^{(2)}_{\textrm{DR}}(0)=-3\gamma^2_{\textrm{C}}\psi_{\textrm{C}}(0)$,
1734: which does not agree with the Gremm-Kapustin relation.
1735:
1736: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1737: \section{Integrals}\label{integrals}
1738: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1739: Here we record some integrals that are useful in deriving
1740: the expressions in this paper.
1741: \subsection{Loop integrals}
1742: \begin{subequations}
1743: \begin{eqnarray}
1744: \int \frac{d^3k}{(2\pi)^3}
1745: \frac{1}{(\bm{k}-\bm{p})^2(\bm{k}^2+\Lambda^2+i\epsilon)}
1746: &=&
1747: \frac{i}{8\pi|\bm{p}|}
1748: \log
1749: \left(
1750: \frac{\Lambda-i|\bm{p}|}{\Lambda+i|\bm{p}|}
1751: \right)\nonumber\\
1752: &=&
1753: \frac{1}{4\pi|\bm{p}|}\arctan\left(\frac{|\bm{p}|}{\Lambda}\right),
1754: \label{loop-1}
1755: \\
1756: \int \frac{d^3k}{(2\pi)^3}
1757: \frac{1}{(\bm{k}-\bm{p})^2(\bm{k}^2+\Lambda^2+i\epsilon)^2}
1758: &=& \frac{1}{8\pi\Lambda(\bm{p}^2+\Lambda^2)},
1759: \label{loop-2}
1760: \\
1761: \int \frac{d^3k}{(2\pi)^3}
1762: \frac{ \arctan\left(|\bm{k}|/\Lambda\right) }
1763: {|\bm{k}|(\bm{k}^2+\Lambda^2+i\epsilon)^2}
1764: &=& \frac{1}{16\pi\Lambda^2}.
1765: \label{loop-3}
1766: \end{eqnarray}
1767: \end{subequations}
1768: Analytically continuing $\Lambda$ to $i\Lambda$, we obtain the following
1769: formulas:
1770: \begin{subequations}
1771: \begin{eqnarray}
1772: \int \frac{d^3k}{(2\pi)^3}
1773: \frac{1}{(\bm{k}-\bm{p})^2(\bm{k}^2-\Lambda^2+i\epsilon)}
1774: &=&
1775: \frac{1}{4\pi |\bm{p}|}\arctan\left(\frac{|\bm{p}|}{i\Lambda}\right)
1776: \nonumber\\
1777: &=&-\frac{i}{4\pi |\bm{p}|}
1778: \tanh^{-1}\left(\frac{|\bm{p}|}{\Lambda}\right),
1779: \label{loop-1m}
1780: \\
1781: \int \frac{d^3k}{(2\pi)^3}
1782: \frac{1}{(\bm{k}-\bm{p})^2(\bm{k}^2-\Lambda^2+i\epsilon)^2}
1783: &=& \frac{-i}{8\pi\Lambda(\bm{p}^2-\Lambda^2)},
1784: \label{loop-3m}
1785: \\
1786: \int \frac{d^3k}{(2\pi)^3}
1787: \frac{ \arctan\left(-i|\bm{k}|/\Lambda\right) }
1788: {|\bm{k}| (\bm{k}^2-\Lambda^2+i\epsilon)^2}
1789: &=& -\frac{1}{16\pi\Lambda^2}.
1790: \label{loop-4m}
1791: \end{eqnarray}
1792: \end{subequations}
1793: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1794: \subsection{Fourier Transformation}
1795: \begin{subequations}
1796: \begin{eqnarray}
1797: \int \frac{d^3p}{(2\pi)^3}\, e^{i\bm{p}\cdot\bm{x}}
1798: \frac{\arctan(|\bm{p}|/\Lambda)}{|\bm{p}|}
1799: &=& \frac{e^{-\Lambda r}}{4\pi r^2},
1800: \\
1801: \int \frac{d^3p}{(2\pi)^3}\, e^{i\bm{p}\cdot\bm{x}}
1802: \frac{1}{(\bm{p}^2+\Lambda^2)}
1803: &=& \frac{e^{-\Lambda r}}{4\pi r},
1804: \\
1805: \int \frac{d^3p}{(2\pi)^3}\, e^{i\bm{p}\cdot\bm{x}}
1806: \frac{1}{(\bm{p}^2+\Lambda^2)^2}
1807: &=& \frac{e^{-\Lambda r}}{8\pi\Lambda},
1808: \end{eqnarray}
1809: \end{subequations}
1810: where $r\equiv|\bm{x}|$.
1811:
1812: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1813: % Create the reference section using BibTeX:
1814:
1815: \begin{thebibliography}{}
1816: %----------------------------------------------------------------------
1817: %\cite{Bodwin:1994jh}
1818: \bibitem{Bodwin:1994jh}
1819: G.~T.~Bodwin, E.~Braaten, and G.~P.~Lepage,
1820: %``Rigorous QCD analysis of inclusive annihilation and production
1821: % of heavy quarkonium,''
1822: Phys.\ Rev.\ D {\bf 51}, 1125 (1995);
1823: {\bf 55}, 5853(E) (1997)
1824: [arXiv:hep-ph/9407339].
1825:
1826: %----------------------------------------------------------------------
1827: %\cite{Kwong:1987ak}
1828: \bibitem{Kwong:1987ak}
1829: W.~Kwong, P.~B.~Mackenzie, R.~Rosenfeld, and J.~L.~Rosner,
1830: %``Quarkonium Annihilation Rates,''
1831: Phys.\ Rev.\ D {\bf 37}, 3210 (1988).
1832: %%CITATION = PHRVA,D37,3210;%%
1833:
1834: %----------------------------------------------------------------------
1835: \bibitem{bks}
1836: G.~T.~Bodwin, S.~Kim, and D.~K.~Sinclair,
1837: %``Matrix elements for the decays of S and P wave quarkonium:
1838: %An exploratory study,''
1839: Nucl.\ Phys.\ B, Proc.\ Suppl. {\bf 34}, 434 (1994);
1840: %%CITATION = NUPHZ,34,434;%%
1841: %\bibitem{bks-2}
1842: %G.~T.~Bodwin, S.~Kim and D.~K.~Sinclair,
1843: %``Decays rates for S wave and P wave bottomonium,''
1844: %Nucl.\ Phys.\ Proc.\ Suppl.\
1845: {\bf 42}, 306 (1995)
1846: [arXiv:hep-lat/9412011];
1847: %%CITATION = HEP-LAT 9412011;%%
1848: %\bibitem{bks-3}
1849: G.~T.~Bodwin, D.~K.~Sinclair, and S.~Kim,
1850: %``Quarkonium decay matrix elements from quenched lattice QCD,''
1851: Phys.\ Rev.\ Lett.\ {\bf 77}, 2376 (1996)
1852: [arXiv:hep-lat/9605023];
1853: %%CITATION = HEP-LAT 9605023;%%
1854: %\bibitem{bks-4}
1855: %G.~T.~Bodwin, D.~K.~Sinclair and S.~Kim,
1856: %``Lattice calculation of quarkonium decay matrix elements,''
1857: Int.\ J.\ Mod.\ Phys.\ A {\bf 12}, 4019 (1997)
1858: [arXiv:hep-ph/9609371];
1859: %%CITATION = HEP-PH 9609371;%%
1860: %\bibitem{bks-4}
1861: %G.~T.~Bodwin, D.~K.~Sinclair and S.~Kim,
1862: %``Bottomonium decay matrix elements from lattice QCD with two light
1863: %quarks,''
1864: Phys.\ Rev.\ D {\bf 65}, 054504 (2002)
1865: [arXiv:hep-lat/0107011].
1866: %%CITATION = HEP-LAT 0107011;%%
1867:
1868: %----------------------------------------------------------------------
1869: %\cite{Gremm:1997dq}
1870: \bibitem{Gremm:1997dq}
1871: M.~Gremm and A.~Kapustin,
1872: %``Annihilation of S-wave quarkonia and the measurement of alpha(s),''
1873: Phys.\ Lett.\ B {\bf 407}, 323 (1997)
1874: [arXiv:hep-ph/9701353].
1875:
1876: %----------------------------------------------------------------------
1877: %\cite{Braaten:2002fi}
1878: \bibitem{Braaten:2002fi}
1879: E.~Braaten and J.~Lee,
1880: %``Exclusive double-charmonium production in e+ e- annihilation,''
1881: Phys.\ Rev.\ D {\bf 67}, 054007 (2003);
1882: {\bf 72}, 099901(E) (2005)
1883: [arXiv:hep-ph/0211085].
1884: %----------------------------------------------------------------------
1885: %\cite{Brambilla:1999xf}
1886: \bibitem{Brambilla:1999xf}
1887: N.~Brambilla, A.~Pineda, J.~Soto, and A.~Vairo,
1888: %``Potential NRQCD: An effective theory for heavy quarkonium,''
1889: Nucl.\ Phys. {\bf B566}, 275 (2000)
1890: [arXiv:hep-ph/9907240].
1891: %%CITATION = HEP-PH 9907240;%%
1892:
1893: %----------------------------------------------------------------------
1894: %\cite{Eichten:1978tg}
1895: \bibitem{Eichten:1978tg}
1896: E.~Eichten, K.~Gottfried, T.~Kinoshita, K.~D.~Lane and T.~M.~Yan,
1897: %``Charmonium: The Model,''
1898: Phys.\ Rev.\ D {\bf 17}, 3090 (1978);
1899: {\bf 21}, 313(E) (1980).
1900: %%CITATION = PHRVA,D17,3090;%%
1901:
1902: %-----------------------------------------------------------------------------
1903: %\cite{Bali:2000gf}
1904: \bibitem{Bali:2000gf}
1905: G.~S.~Bali,
1906: %``QCD forces and heavy quark bound states,''
1907: Phys.\ Rep.\ {\bf 343}, 1 (2001)
1908: [arXiv:hep-ph/0001312].
1909: %%CITATION = HEP-PH 0001312;%%
1910:
1911: \bibitem{bkl}
1912: G.T.~Bodwin, D.~Kang, and J.~Lee, arXiv:hep-ph/0603185.
1913:
1914:
1915: %-----------------------------------------------------------------------------
1916: %\cite{Booth:1992bm}
1917: \bibitem{Booth:1992bm}
1918: S.~P.~Booth, D.~S.~Henty, A.~Hulsebos, A.~C.~Irving, C.~Michael,
1919: and P.~W.~Stephenson (UKQCD Collaboration),
1920: %``The Running coupling from SU(3) lattice gauge theory,''
1921: Phys.\ Lett.\ B {\bf 294}, 385 (1992)
1922: [arXiv:hep-lat/9209008].
1923: %%CITATION = HEP-LAT 9209008;%%
1924:
1925: %-----------------------------------------------------------------------------
1926: %\cite{Gupta:1996sa}
1927: \bibitem{Gupta:1996sa}
1928: R.~Gupta and T.~Bhattacharya,
1929: %``Light quark masses from lattice QCD,''
1930: Phys.\ Rev.\ D {\bf 55}, 7203 (1997)
1931: [arXiv:hep-lat/9605039].
1932: %%CITATION = HEP-LAT 9605039;%%
1933:
1934: %-----------------------------------------------------------------------------
1935: %\cite{Kim:1993gc}
1936: \bibitem{Kim:1993gc}
1937: S.~Kim and D.~K.~Sinclair,
1938: %``The Light hadron spectrum of quenched QCD on a 32**3 x 64 lattice,''
1939: Phys.\ Rev.\ D {\bf 48}, 4408 (1993).
1940: %%CITATION = PHRVA,D48,4408;%%
1941:
1942: %-----------------------------------------------------------------------------
1943: %\cite{Kim:1996cz}
1944: \bibitem{Kim:1996cz}
1945: S.~Kim and S.~Ohta,
1946: %``Quenched KS light hadron mass at beta = 6.5 on a 64 x 48**3 lattice,''
1947: Nucl.\ Phys.\ B,\ Proc.\ Suppl.\ {\bf 53}, 199 (1997)
1948: [arXiv:hep-lat/9609023].
1949: %%CITATION = HEP-LAT 9609023;%%
1950:
1951: %-----------------------------------------------------------------------------
1952: %\cite{Bodwin:1998mn}
1953: \bibitem{Bodwin:1998mn}
1954: G.~T.~Bodwin and Y.~Q.~Chen,
1955: %``Renormalon ambiguities in NR{QCD} operator matrix elements,''
1956: Phys.\ Rev.\ D {\bf 60}, 054008 (1999)
1957: [arXiv:hep-ph/9807492].
1958: %%CITATION = HEP-PH 9807492;%%
1959:
1960: %\cite{Beneke:1997zp}
1961: \bibitem{Beneke:1997zp}
1962: M.~Beneke and V.~A.~Smirnov,
1963: %``Asymptotic expansion of Feynman integrals near threshold,''
1964: Nucl.\ Phys. {\bf B522}, 321 (1998)
1965: [arXiv:hep-ph/9711391].
1966: %%CITATION = HEP-PH 9711391;%%
1967:
1968: \end{thebibliography}
1969: \end{document}
1970: