hep-ph0102257/Vc.tex
1: %\documentstyle[aps,prl,epsf,floats]{revtex}
2: \documentstyle[preprint,aps,epsf,floats,color]{revtex}
3: 
4: %\usepackage[usenames]{color}
5: 
6: % \input color.sty
7: 
8: \newcommand{\nn}{\nonumber}
9: \newcommand{\ppp}{\mbox{$({\mathbf p'}-{\mathbf p})^2$}}
10: \newcommand{\spp}{\mbox{$i\bsigma\cdot ({\mathbf p'\times p})$}}
11: \newcommand{\two}{{\rm 2}}
12: \newcommand{\alpS}{\alpha_{S}}
13: \newcommand{\alpU}{\alpha_{U}}
14: \newcommand{\lmeas}{\mbox{$d^d\!\!\!\!\!\raisebox{0.12cm}{--}\ $}}
15: 
16: \def\bmag#1{{|{\mathbf #1}|}}
17: 
18: \def\xslash#1{{\rlap{$#1$}/}}
19: \def\Dsl{\hbox{/\kern-.6000em D}} %roman D
20: \def\vabsq#1{\left|{\bf #1}\right|^2}
21: \def\vabs#1{\left|{\bf #1}\right|}
22: \def\vev#1{\left\langle{#1}\right\rangle}
23: \def\dsl{\,\raise.15ex\hbox{/}\mkern-13.5mu D}
24: \def\bsigma{\mbox{\boldmath $\sigma$}}
25: \def\lqcd{\Lambda_{\rm QCD}}
26: \def\psip#1{\psi_{\mathbf{#1}}}
27: \def\chip#1{\chi_{\mathbf{#1}}}
28: \def\bsigma{\mbox{\boldmath $\sigma$}}
29: \def\muus{\mu_{\rm U}}
30: \def\mus{\mu_{\rm S}}
31: \def\abs#1{\left| #1 \right|}
32: \def\ltap{\ \raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}\ }
33: \def\gtap{\ \raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}\ }
34: \def\OMIT#1{}
35: 
36: \def\tb#1{\textcolor{blue}{#1}}
37: \def\tbb#1{\textcolor{black}{#1}}
38: 
39: \def\vect#1{{\bf #1}}
40: \def\msb{{\overline{\rm MS}}}
41: \def\ams#1{\mbox{${\overline\alpha_s^{\,#1}}$}}
42: \def\CL{{\cal L}}
43: 
44: \begin{document}
45: \setlength\baselineskip{20pt}
46: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname
47: 
48: \preprint{\vbox{\tighten  \hbox{UCSD/PTH 00-25} \hbox{MPI-PhT/2001-04}
49: %                          \hbox{February 2001} 
50: \hbox{hep-ph/0102257}
51: }}
52: 
53: \title{The Running Coulomb Potential and Lamb Shift in QCD}
54: 
55: \author{ Andre~H.~Hoang\thanks{ahoang@mppmu.mpg.de} \\[4pt]} 
56: 
57: \address{\tighten
58: Max-Planck-Institut f\"ur Physik\\
59: (Werner-Heisenberg-Institut)\\
60: F\"ohringer Ring 6, 80805 M\"unchen, Germany}
61: 
62: \author{Aneesh V. Manohar\thanks{amanohar@ucsd.edu}, Iain W.\
63: Stewart\thanks{iain@schwinger.ucsd.edu} \\[4pt]} 
64: \address{\tighten Department of Physics, University of California at San
65: Diego,\\[2pt] 9500 Gilman Drive, La Jolla, CA 92093, USA }
66: 
67: \maketitle
68: 
69: {\tighten
70: \begin{abstract}
71: 
72: The QCD $\beta$-function and the anomalous dimensions for the Coulomb
73: potential and the static potential first differ at three loop order. We evaluate
74: the three loop ultrasoft anomalous dimension for the Coulomb potential and give
75: the complete three loop running. Using this result, we calculate the leading
76: logarithmic Lamb shift for a heavy quark-antiquark bound state, which
77: includes all contributions to the 
78: binding energies of the form $m\,\alpha_s^4 (\alpha_s\ln \alpha_s)^k$, $k\ge
79: 0$.
80: 
81: \end{abstract}
82: \pacs{12.39.Hg,11.10.St,12.38.Bx}
83: }%end tighten
84: %]\narrowtext %% end two-column
85: 
86: \section{Introduction}\label{sec:intro}
87: 
88: In this paper, we construct the three-loop anomalous dimension for the Coulomb
89: potential in non-relativistic QCD (NRQCD)~\cite{nrqcd,pNRQCD}. The formalism we
90: use was developed in Refs.~\cite{LMR,amis,amis2,amis3} and will be referred to
91: as vNRQCD, an effective theory for heavy non-relativistic quark-antiquark
92: pairs. Part of our computation is related to the running of the static
93: potential~\cite{PSstat}, however effects associated with motion of the quarks do
94: play an important role. Our final result for the Coulomb potential differs from
95: the static potential at terms beyond those with a single logarithm (i.e. starting
96: at four loops). Combining our Coulomb potential running with previous results
97: for the running of the $1/m$ and $1/m^2$ potentials~\cite{amis,amis3} allows us
98: to compute the next-to-next-to-leading logarithmic (NNLL) corrections in the
99: perturbative energy of a heavy $Q\bar Q$ bound state, which includes the sum of
100: terms $m\, \alpha_s^4 (\alpha_s\ln \alpha_s)^k$, $k\ge 0$, where $m$ is the
101: heavy quark pole mass. This contribution is the QCD analog of the QED
102: $\alpha^5\ln\alpha$ Lamb shift computed by Bethe. In QED, the series $m\,
103: \alpha^4 (\alpha \ln \alpha)^k$ terminates after the $k=1$ term~\cite{amis4}. In
104: QCD, there is an infinite series due to the running QCD coupling as well as the
105: presence of non-trivial anomalous dimensions for other QCD operators.  The
106: results presented here also contribute to the NNLL prediction for the cross
107: section for $e^+e^-\to t\bar t$ near threshold~\cite{hmst}.  Implications for
108: $b\bar b$ sum rules will be addressed in a future publication.
109: 
110: The expansion parameter of the effective theory is the quark velocity $v$. A
111: quark has a momentum of order $mv$ and an energy of order $mv^2$.  We assume
112: that $m$ is large enough that $mv^2\gg \Lambda_{\rm QCD}$ and a perturbative
113: description of the bound state as a Coulombic system is valid. For a Coulombic
114: bound state, $\alpha_s$ is of order $v$ and contributions suppressed by both $v$
115: and $\alpha_s$ are of the same order.  It is useful to distinguish between
116: powers of $\alpha_s$ and $v$ when carrying out the matching and when evolving
117: couplings and operators in the effective theory, and to only take $v\sim
118: \alpha_s$ for the power counting of bound state matrix elements. In the
119: effective theory, the quark-antiquark potentials appear as four-quark
120: operators~\cite{pNRQCD}.  A potential of the form $\alpha_s^r/\bmag{k}^s$ is of
121: order $\alpha_s^r v^{1-s}$, where ${\bf k}$ is the fermion momentum transfer.
122: With this power counting the time-ordered product of a $v^a$ and $v^b$ potential
123: is of order $v^{a+b}$. Up to next-to-next-to-leading order (NNLO) the heavy
124: quark potential has contributions
125: \begin{eqnarray} \label{Vpc} 
126:   V  & \sim & \bigg[\: {\alpha_s \over v} \: \bigg] \ +\ 
127:   \bigg[\: { \alpha_s^2 \over v} \: \bigg] \ + \ \bigg[\: { \alpha_s^3 \over v} 
128:   + {\alpha_s^2 v^0 } + {\alpha_s v } \: \bigg] \ +  \ldots \,. 
129: \end{eqnarray}
130: The order $\alpha_s/v \sim 1$ term in Eq.~(\ref{Vpc}) is the
131: Coulomb potential generated at tree-level. The next-to-leading order (NLO) term
132: is the one-loop correction to the Coulomb potential. The NNLO terms are the
133: two-loop correction to the Coulomb potential, the one-loop value for the $1/(m
134: \bmag{k})$ potential, and the tree-level contribution to the order
135: $\bmag{k}^0/m^2$ potential. Writing the $Q\bar Q$ energy as $E = 2 m + \Delta E$,
136: the terms in Eq.~(\ref{Vpc}) generate contributions of the following order in the
137: binding energy:
138: \begin{eqnarray} \label{Epc} 
139:   {\Delta E} & \sim & \Big[ m\alpha_s^2 \Big] + \Big[ m\alpha_s^3 \Big]  
140:   + \Big[ m\alpha_s^4 \Big] + \ldots \,. 
141: \end{eqnarray}
142: 
143: In Eqs.~(\ref{Vpc}) and (\ref{Epc}) the expansion has been performed at the
144: scale $\mu=m$ so the coupling constants are $\alpha_s = \alpha_s(m)$. A typical
145: perturbative expansion contains logarithms of $\mu$ divided by the various
146: physical scales in the bound state. If the logarithms are large, fixed order
147: perturbation theory breaks down, and one finds a large residual $\mu$
148: dependence. One can minimize the logarithms by setting $\mu$ to a value
149: appropriate to the dynamics of the non-relativistic system.  This is
150: accomplished by summing large logarithms using the renormalization group, and
151: using renormalization group improved perturbation theory. For $Q\bar Q$ bound
152: states, the large logarithms are logarithms of $v\sim \alpha_s$, and can be
153: summed using the velocity renormalization group (VRG)~\cite{LMR}.  For the
154: binding energy this gives the expansion
155: \begin{eqnarray} \label{Ell}
156:   \Delta E &=& \Delta E^{LL}+ \Delta E^{NLL} + \Delta E^{NNLL} + \ldots \,,\\
157:  &\sim & \bigg[m \sum_{k=0}^\infty \alpha_s^{k+2}(\ln\alpha_s)^k\bigg] 
158:  + \bigg[m \sum_{k=0}^\infty \alpha_s^{k+3}(\ln\alpha_s)^k \bigg] 
159:  + \bigg[m \sum_{k=0}^\infty \alpha_s^{k+4}(\ln\alpha_s)^k \bigg] 
160:  + \ldots \,,\nn
161: \end{eqnarray}
162: where the terms are the leading log (LL), next-to-leading log (NLL), and
163: next-to-next-to-leading log (NNLL) results respectively.
164: 
165: In the VRG, one uses a subtraction velocity $\nu$ that is evolved from $1$ to
166: $v$.  This simultaneously lowers the momentum cutoff scale $\mu_S=m\nu$ from $m$
167: to $mv$ and the energy cutoff scale $\mu_U=m\nu^2$ from $m$ to $mv^2$. The VRG
168: properly accounts for the coupling between energy and momentum caused by the
169: equations of motion for the non-relativistic quarks. QED provides a highly
170: non-trivial check of the VRG method. In Ref.~\cite{amis4} it was used to
171: correctly reproduce terms in the subleading series of logarithms, including the
172: $\alpha^3\ln^2\alpha$ corrections to the ortho and para-positronium decay rates,
173: the $\alpha^7\ln^2\alpha$ hyperfine splittings for Hydrogen and positronium, and
174: the $\alpha^8 \ln^3 \alpha$ Lamb shift for Hydrogen. The difference between the
175: VRG, which involves the evolution of the momentum and the energy scale in a
176: single step, and a conventional two stage renormalization group treatment, $m\to
177: mv\to mv^2$, was examined in Ref.~\cite{mss1}.
178: 
179: 
180: In section~\ref{sec:pots} we review the definition of potentials for non-static
181: heavy quarks in the effective theory.  In section \ref{sec:static} we compare
182: these potentials with the Wilson loop definition which is appropriate for static
183: quarks.  In section~\ref{sec:coulomb} we rederive the leading-logarithmic (LL)
184: and next-to-leading-logarithmic (NLL) results for the $Q\bar Q$ binding energy
185: using the effective theory and discuss the two-loop matching for the Coulomb
186: potential using the results in Refs.~\cite{Peter,Schroeder}.  We also discuss
187: some subtleties in the correspondence between diagrams in the static theory and
188: soft diagrams in the effective theory. In section~\ref{sec:threeloop} we compute
189: the three-loop anomalous dimension for the Wilson coefficient of the Coulomb
190: potential. Results for the NNLL energy are given in section~\ref{sec:nnll},
191: followed by conclusions in section~\ref{sec:conclusion}. In Appendix~\ref{UVIR}
192: we give some technical details on the structure of divergences in the effective
193: theory, and in Appendix~\ref{Ns} we list some functions that appear in the
194: energy at NNLL order.
195: 
196: \section{The vNRQCD Potentials}\label{sec:pots}
197: 
198: The effective theory vNRQCD has soft gluons with coupling constant
199: $\alpS(\nu)$, ultrasoft gluons with coupling constant $\alpU(\nu)$, as well as
200: quark-antiquark potentials. The potential is the momentum dependent coefficient
201: of a four-fermion operator:
202: \begin{eqnarray}\label{Vm}
203: {\mathcal L} &=&  - \sum_{\mathbf p,p'}  V\left({\bf p},{\bf p^\prime}\right)\: 
204:   \Big[ {\psip {p^\prime}}^\dagger\: {\psip p}\: 
205:   {\chip {-p^\prime}}^\dagger\:  {\chip {-p}}{}\Big] \,,
206: \end{eqnarray}
207: where spin and color indices are suppressed. The coefficient $V\left({\bf
208: p},{\bf p^\prime} \right)$ has an expansion in powers of $v$, $ V = V_{(-1)}
209: + V_{(0)} + V_{(1)} + \ldots $, where $V_{(-1)}=V_c$ is the Coulomb
210: potential. For equal mass fermions
211: \begin{eqnarray} \label{V}
212:  V_c &=& (T^A \otimes \bar T^A) { \tb{ {\cal V}_c^{(T)}} \over {\mathbf
213:   k}^2} +  (1 \otimes 1)  { \tb{ {\cal V}_c^{(1)} } \over
214:   {\mathbf k}^2} ,\nn\\[10pt]
215:  V_{(0)} &=& (T^A \otimes \bar T^A) { \pi^2 \tb{ {\cal V}_k^{(T)} } \over m\,
216:  |{\mathbf k}| } + (1 \otimes 1) {\pi^2 \tb{ {\cal V}_k^{(1)} } \over 
217:  m\,|{\mathbf k}|} \,, \nn \\[10pt]
218:  V_{(1)} &=& (T^A \otimes \bar T^A) \left[ {\tb{ {\cal
219:   V}_2^{(T)} } \over m^2 } + { \tb{ {\cal V}_s^{(T)} } {\mathbf
220:   S}^2 \over m^2}\, +  { \tb{ {\cal V}_r^{(T)} }\: ({\mathbf p^2 +
221:   p^{\prime 2}}) \over 2\, m^2\, {\mathbf k}^2}  -{i {\tb{ {\cal
222:   V}_\Lambda^{(T)}} }\, {\mathbf S} \cdot ({\bf p\,^\prime \times p} ) 
223:   \over m^2 {\mathbf k}^2 } + { \tb{ {\cal V}_t^{(T)} } T({\mathbf k}) \over 
224:   m^2}\, \right] \nn \\
225:  && + (1 \otimes 1) \left[{ \tb{ {\cal V}_2^{(1)} } \over m^2}\:+ 
226:   { \tb{ {\cal V}_s^{(1)} } \over m^2}\: {\mathbf S}^2 \right] , 
227: \end{eqnarray}
228: where ${\bf k} = {\bf p'} - {\bf p}$, $\mathbf S = { ({\mathbf \bsigma_1 +
229: \bsigma_2}) / 2}$, $T({\mathbf k}) = (\delta^{ij} - {3 {\mathbf k}^i{\mathbf
230: k}^j/{\mathbf k}^2}){ \bsigma_1^i \bsigma_2^j}$.  The Wilson coefficients,
231: ${\cal V}^{(T,1)}$ depend on the subtraction velocity $\nu$. In Eq.~(\ref{V})
232: the color decomposition $V = (T^A \otimes \bar T^A) { V^{(T)} } + (1 \otimes 1)
233: { V^{(1)} } $ has been used and the potential in the color singlet channel is
234: $V^{(s)}=V^{(1)}-C_F {V}^{(T)}$.  (The Casimirs of the adjoint and fundamental
235: representations are denoted by $C_A$ and $C_F$, respectively.) At LL order the
236: running of the coefficients ${\cal V}^{(1,T)}_{2,s,r}$ was computed in
237: Ref.~\cite{amis} and ${\cal V}^{(1,T)}_{\Lambda,t}$ in Ref.~\cite{amis,chen},
238: while the NLL order running of ${\cal V}_k^{(1,T)}$ was computed in
239: Ref.~\cite{amis3}. In this work we compute the running of ${\cal V}_c^{(s)}$ at
240: NNLL order. This allows the computation of the $Q\bar Q$ energy spectrum at NNLL
241: order.
242: 
243: In vNRQCD additional potential-like effects are generated by
244: loops with soft gluons, for which the Feynman rules can be found in
245: Refs.~\cite{LMR,amis2}. Matrix elements of soft gluon diagrams contribute to
246: the energy beginning at NLO. In contrast, matrix elements with ultrasoft gluons
247: start at ${\rm N}^3{\rm LO}$.  The renormalization group improved energies are
248: obtained by computing the anomalous dimensions for these soft interactions and
249: the four fermion operators in Eq.~(\ref{V}).
250: 
251: \section{The static potential versus the Coulomb potential}
252: \label{sec:static}
253: 
254: Parts of our analysis are related to the study of the static limit of QCD which
255: describes heavy quarks in the $m\to\infty$ limit. We therefore briefly
256: review the pertinent results which have been derived in this framework.
257: 
258: In position space the static QCD potential is defined as the expectation value
259: of the Wilson loop operator,
260: \begin{eqnarray} \label{wilson}
261:  V_{\rm stat}(r) = \lim_{T\to\infty} {1\over T} \ln \vev{ {\rm Tr}\, P\exp - 
262:   i g \oint_C A_\mu dx^\mu}, 
263: \end{eqnarray} 
264: where $C$ is a rectangle of width $T$ and fixed height $r$.  This potential is
265: independent of the mass $m$ of the quarks and depends only on $r$.  In QCD
266: perturbation theory the static potential is known at two-loop
267: order~\cite{Peter,Schroeder} .  These calculations use static fermion sources
268: with propagators which are identical to those in Heavy Quark Effective
269: Theory~\cite{MW}. The exponentiation of the static
270: potential~\cite{Gatheral,FrenkelTaylor} guarantees that one can avoid dealing
271: with graphs which have pinch singularities in momentum space. The analysis of
272: Refs.~\cite{Gatheral,FrenkelTaylor} also gives a prescription for the color
273: weight factors for different graphs based on the c-web theorem.
274: 
275: In Ref.~\cite{ADM}, Appelquist, Dine and Muzinich (ADM) pointed out that at
276: three loops the static potential in Eq.~(\ref{wilson}) has infrared (IR)
277: divergences from graphs of the form in Fig.~\ref{fig_ADM}a,b.
278: \begin{figure}
279:  \centerline{ \hbox{\epsfxsize=12cm\epsfbox{fullus.eps} } } \medskip
280:  {\tighten\caption{Graphs contributing to the three-loop IR divergence of
281:  the QCD static potential.} \label{fig_ADM} }
282: \end{figure}
283: In the color singlet channel Fig.~\ref{fig_ADM}a has color factor $C_F C_A^2
284: (C_A-2 C_F)$ while Fig.~\ref{fig_ADM}b is proportional to $C_F^2 C_A^2$.  Taking
285: into account the exponentiation of $V_{\rm stat}$ using the c-web theorem, the
286: color singlet contribution to $V_{\rm stat}$ from Fig.~\ref{fig_ADM}(a,b) is
287: simply Fig.~\ref{fig_ADM}a with the color factor replaced by $C_A^3 C_F$, and is
288: IR divergent. ADM showed that this IR divergence could be avoided by summing a
289: class of diagrams---those of Fig.~\ref{fig_ADM} with the addition of an
290: arbitrary number of gluon rungs.  Summing over these diagrams regulates the IR
291: divergence by building up Coulombic states for the static quark sources. The
292: summation gives a factor $\exp\Big([V^{(s)}(r)-V^{(o)}(r)]T\Big)$ for the
293: propagation of the intermediate color-octet $Q\bar Q$ pair, where $V^{(s)}(r)$
294: and $V^{(o)}(r)$ are the color-singlet and color-octet potentials. The
295: exponential factor suppresses long-time propagation of the intermediate
296: color-octet state, and regulates the IR divergence by introducing an IR cutoff
297: scale of order $[V^{(s)}(r)-V^{(o)}(r)]\sim \alpha_s/r$.
298: 
299: In Ref.~\cite{static1} the ADM divergence in the static potential was studied by
300: Brambilla et al.\ using the effective theory pNRQCD~\cite{pNRQCD}. They made the
301: important observation that along with potential contributions, the definition in
302: Eq.~(\ref{wilson}) contains contributions from ultrasoft gluons, and the latter
303: are responsible for the ADM IR divergence. They showed that the ADM IR
304: divergence in QCD matches with an IR divergence of a pNRQCD graph that describes
305: the selfenergy of a quark-antiquark system due to an ultrasoft gluon with
306: momenta $q^\mu \sim \alpha_s/r$.  Therefore, the static potential in pNRQCD can
307: be defined as a matching coefficient of a four fermion operator, as in
308: Eq.~(\ref{Vm}), in an infrared safe manner. We will refer to this potential as
309: the soft-static potential. The ultrasoft pNRQCD graph also has an ultraviolet
310: divergence.  Brambilla et al.\ computed the coefficient of this divergence and
311: extracted a new $\ln(\mu)$ contribution to the soft-static potential. In
312: Ref.~\cite{PSstat} the three-loop anomalous dimension was computed for the
313: soft-static potential in this framework. In the color singlet channel for scales
314: $\mu \sim \alpha_s(r)/r$ their solution reads
315: \begin{eqnarray}\label{psVc}
316:  V_{\rm stat}(\mu,r) = V_{\rm stat}^{\rm (2 loops)}(r)- 
317:  \frac{1}{4\pi r} \Bigg[\, \frac{2\pi C_F C_A^3}
318:  {3\beta_0}\alpha_s^3(r) \ln\bigg(\frac{\alpha_s(r)}{\alpha_s(\mu)}\bigg) 
319:  \,\Bigg]  \,,
320: \end{eqnarray}
321: where the first term is the two loop static potential derived in
322: Refs.~\cite{Peter,Schroeder}.
323: 
324: For large but finite $m$ the effective theory for $Q\bar Q$ bound states has an
325: expansion in $v$.  The quark potential in this case differs from that in the
326: static case, and in general one cannot obtain the static theory by taking the $m
327: \to \infty$ limit. To illustrate this consider as an example the loop graph
328: involving the iteration of a $1/(m\bmag{k})$ and a Coulomb potential as shown in
329: Fig.~\ref{fig:tproduct}.
330: \begin{figure}
331:  \centerline{\epsfxsize=4cm \epsfbox{fd1.eps}} \medskip
332:  {\tighten\caption{Loop with an insertion of the $1/(m|{\bf k})$ and 
333:  $1/{\bf k^2}$ potentials.}
334:  \label{fig:tproduct}}
335: \end{figure}
336: In the effective theory for non-static fermions, the fermion propagators give a
337: factor proportional to $m$.  The final result for the graph is independent of
338: $m$ and is the same order in $v$ as the Coulomb potential. On the other hand, if
339: we first take the static limit $m \to \infty$, then $1/(m\bmag{k})\to 0$, and
340: there is no such graph.  This example illustrates the general result that the
341: static theory is not obtained as the $m\to\infty$ (or $v\to0$) limit of the
342: non-static effective theory. Loop integrals can produce factors of $m$ or $1/v$,
343: and thereby cause mixing between operators which are of different
344: orders\footnote{\tighten However, we stress that if powers of $\alpha_s$ are
345: also counted as powers of $v$, then operators which are higher order in $v$
346: never mix into lower order operators.}  in $v$. Furthermore, the energy and the
347: momentum transfer in Coulombic states are coupled by the quark equations of
348: motion. We will see that for this reason the anomalous dimension of the static
349: and Coulomb potentials differ at three loops. The three-loop matching for the
350: static and Coulomb potentials can also differ.
351: 
352: \section{The Coulomb potential at one and two loops}\label{sec:coulomb}
353: 
354: In this work dimensional regularization and the $\msb$ scheme will be used.  The
355: $\msb$ QCD running coupling constant will be denoted by $\ams{}(\mu)$, and is
356: determined by the solution of the renormalization group equation
357: \begin{eqnarray}\label{msbar}
358:  \mu {{\rm d \ams{}(\mu) \over {\rm d} \mu}} &=& 
359:   \overline \beta \left( \ams{}(\mu) \right)\nn\\
360:  &=& -2\beta_0 \frac{\ams{2}(\mu)}{4\pi} 
361:   -2 \beta_1 \frac{\ams{3}(\mu)}{(4\pi)^2}
362:   -2 \beta_2 \frac{\ams{4}(\mu)}{(4\pi)^3} + \ldots \,.
363: \end{eqnarray}
364: In a mass-independent subtraction scheme $\beta_0$ and $\beta_1$ are
365: scheme-independent. The notation $\ams{[n]} \left(\mu \right)$
366: will be used to indicate the solution of Eq.~(\ref{msbar}) with coefficients up
367: to $\beta_{n-1}$ (i.e.\ $n$ loop order) kept in the $\beta$-function.
368: 
369: %%% LO %%%
370: 
371: In the VRG, the soft and ultrasoft subtraction scales $\mu_S$ and $\mu_U$ are
372: given by $\mu_S=m \nu$ and $\mu_U=m \nu^2$. We define the soft and ultrasoft
373: anomalous dimensions $\gamma_S$ and $\gamma_U$ as the derivatives with respect
374: to $\ln\mu_S$ and $\ln \mu_U$, respectively. The derivative with respect to $\ln
375: \nu$ gives the total anomalous dimension, $\gamma=\gamma_S+2\gamma_U$.  vNRQCD
376: has three independent but related coupling constants that are relevant for our
377: calculation: the soft gluon coupling $\alpS(\nu)$, the ultrasoft gluon
378: coupling $\alpU(\nu)$, and the coefficient of the Coulomb potential ${\cal
379: V}_c(\nu)$.  The tree-level matching conditions at ($\mu=m \Leftrightarrow
380: \nu=1$) are
381: \begin{eqnarray}
382:  \alpS(1) &=& \alpU (1) = \ams{} (m) \,, \qquad
383:  {\cal V}_c^{(T)}(1) = 4 \pi \ams{} (m)\,, \qquad 
384:  {\cal V}_c^{(1)}(1)=0 \, ,
385: \end{eqnarray} 
386: and the solutions of the one-loop renormalization group equations for the
387: coupling constants in the effective theory are~\cite{Gries,LMR}
388: \begin{eqnarray}\label{loop1}
389:  \alpS\left( \nu \right) &=& \ams{[1]} (m \nu) \,,\qquad\quad\ \
390:  \alpU \left( \nu \right) = \ams{[1]} (m \nu^2) \,, \nn\\
391:  {\cal V}_c^{(T)}(\nu) &=& 4 \pi \ams{[1]}(m \nu) \,,\qquad
392:  {\cal V}_c^{(1)}(\nu) = 0.
393: \end{eqnarray}
394: In deriving the above equations, it has been assumed that any light fermions
395: have masses much smaller than $mv^2$, so that there are no mass thresholds in
396: the renormalization group evolution. If there is a mass threshold larger than
397: $mv^2$ and widely separated from $mv$ and $m$, then it is possible to also
398: include such effects in the effective theory, see Ref.~\cite{LMR}.
399: 
400: At leading order the Hamiltonian for the color singlet $Q\bar Q$ system is
401: \begin{eqnarray}
402:   H_0 = \frac{\bf p^2}{m} + \frac{ \tbb{ {\cal V}_c^{(s)}(\nu)} }{\bf k^2} \,. 
403: \end{eqnarray}
404: To minimize large logarithms in higher order matrix elements we run $\nu$ to
405: the bound state velocity $v_b$, which we define as the solution of the equation
406: \begin{eqnarray}
407:  v_b =  \frac{ \tbb{ a_c(\nu=v_b) }}{ n} = 
408:  \frac{ C_F\: \ams{[1]}(m v_b)}{n} \,,
409: \end{eqnarray}
410: where for convenience we have defined 
411: \begin{eqnarray}
412:    a_c(\nu)=-\: \frac{ {\cal V}_c^{(s)}(\nu)} {4\pi } \,,
413: \end{eqnarray}
414: and $n$ is the principal quantum number.  The LL binding energy is then simply
415: the eigenvalue of the Schr\"odinger equation, $H_0 |\psi_{n,l}\rangle = \Delta E
416: |\psi_{n,l}\rangle$ with the LL solution for the Coulomb potential, ${\cal
417: V}_c^{(s)}(\nu)=-C_F {\cal V}_c^{(T)}(\nu)$ from Eq.~(\ref{loop1}).
418: Thus,
419: \begin{mathletters}
420: \begin{eqnarray} \label{eLL}
421:   \Delta E^{LL} &=& 
422:    -\frac{m }{4n^2}\: \big[ \tb{ a_c(\nu) } \big]^2 \\
423:    &=& - \frac{m}{4\,n^2}\: C_F^2 \,\Big[ \ams{[1]}(m v_b)\Big]^2 
424:      = -\frac{m v_b^2}{4} \,,
425: \end{eqnarray}
426: \end{mathletters}
427: where in the second line we have evaluated the energy at the low scale $\nu=v_b$.
428: Higher order corrections to the energy are all evaluated as perturbative matrix
429: elements with the leading order wavefunctions, $|\psi_{n,l}\rangle$.
430: 
431: %%% NLO %%%
432: 
433: Consider how the results in Eq.~(\ref{loop1}) are extended to higher orders.
434: The graphs for the renormalization of the ultrasoft gluon self coupling have
435: the same rules as for QCD, and those for the renormalization of the lowest
436: order soft gluon vertex have the same rules as for HQET. Since the
437: momenta of soft and ultrasoft gluons are cleanly separated there is no mixing
438: of scales, so the anomalous dimension for $\alpS$ is independent of $\alpU$ and
439: vica-versa.  Thus, one expects that in the $\msb$ scheme $\alpS\left( \nu
440: \right) = \ams{}(m \nu)$ and $\alpU\left( \nu \right) = \ams{}(m \nu^2)$ to all
441: orders.
442: 
443: However, the coefficient of the Coulomb potential can differ from $4 \pi
444: \ams{}(m\nu)$ at higher orders. At one-loop, the only order $\alpha_s^2/v$ graph
445: in the effective theory is the soft diagram~\cite{amis2}\footnote{\tighten Note
446: that the soft loop includes soft gluons, soft light quarks, as well as soft
447: ghosts.}
448: \begin{eqnarray}  \label{eftvc1}
449: \begin{picture}(60,30)(1,1)
450:  \epsfxsize=2.0cm \lower16pt \hbox{\epsfbox{eft_soft.eps}}
451: \end{picture}
452:   &=& \frac{-i \mu_S^{2\epsilon}\alpS^2(\nu)}{\bf k^2} (T^A  \otimes \bar T^A) 
453:   \bigg[ \frac{\beta_0}{\epsilon} + \beta_0
454:    \ln\Big( \frac{\mu_S^2}{\bf k^2}\Big)  + a_1 \bigg] \,, \\[-15pt]\nn
455: \end{eqnarray}
456: where $\beta_0=11/3 C_A-4 T_F n_\ell/3$ and $a_1=31 C_A/9-20 T_F n_\ell/9$ in
457: the $\msb$ scheme, and $n_\ell$ is the number of light soft quarks. The
458: divergence in Eq.~(\ref{eftvc1}) is canceled by a counterterm for ${\cal
459: V}_c^{(T)}$, causing it to run with anomalous dimension
460: $-2\beta_0\alpS^2(\nu)$. The remaining terms in the soft graph are identical to
461: the one-loop soft-static potential calculation and also reproduce the set of
462: $\alpha_s^2/{\bf k^2}$ terms in full QCD with dimensional regularization
463: parameter $\mu$. The one-loop matching correction to ${\cal V}_c^{(T,1)}(1)$ is
464: the difference between the full and effective theory diagrams and therefore
465: vanishes at the matching scale $\mu=\mu_S=m$.
466: 
467: A correspondence between the soft-static potential calculation and soft order
468: $1/v$ diagrams is expected to persist at higher orders in $\alpha_s$ as
469: well. The Feynman rules for the soft vertices are almost identical to the HQET
470: rules used for soft-static potential calculations. There are a few notable
471: differences.  In the effective theory it is not necessary to use the
472: exponentiation theorem~\cite{Gatheral,FrenkelTaylor} to eliminate diagrams with
473: pinch singularities of the form
474: \begin{eqnarray}
475:  \int {\rm d}q^0\ \frac{1}{(q0+i\epsilon)(-q0+i\epsilon)} \,.
476: \end{eqnarray}
477: These are automatically removed in the construction of the tree level soft
478: vertices because the $1/q_0$ factors in the soft Feynman rules do not contain
479: $i\epsilon$'s, and in evaluating diagrams these poles are ignored.  For
480: example~\cite{LMR}, from matching the full theory Compton scattering graphs in
481: Fig.~\ref{fig_compton}
482: \begin{figure}
483:  \centerline{ \hbox{\epsfxsize=10cm\epsfbox{fulls.eps} } } 
484:  {\tighten\caption{Compton scattering graphs that contribute to the soft
485:  vertex.} 
486:  \label{fig_compton} }
487: \end{figure}
488: one obtains the soft vertex in Fig.~\ref{fig_softfr}a which is proportional to
489: \begin{figure}
490:  \centerline{ \hbox{\epsfxsize=8cm\epsfbox{soft_fr.eps} } } 
491:  {\tighten\caption{Examples of vertices involving soft gluons.} 
492:  \label{fig_softfr} }
493: \end{figure}
494: $[T^A, T^B]/q_0$, where terms proportional to $\{T^A, T^B\}$ have canceled.  In
495: the soft-static calculations this cancellation instead takes place at the level
496: of the box and crossed box graphs, and is guaranteed by exponentiation.  For the
497: soft-static potential it is known that at higher orders there are contributions
498: from the $i\pi \delta(q_0)$ terms that originate from the $i\epsilon$'s. It was
499: exactly this type of contribution that was missed in the two-loop calculation by
500: Peter~\cite{Peter}, and was correctly identified by
501: Schr\"oder~\cite{Schroeder}. In the effective theory these delta function
502: contributions belong to the potential regime~\cite{amis3}, and soft-static graphs
503: with this type of contribution are reproduced by operators such as the one shown
504: in Fig.~\ref{fig_softfr}b, where a soft gluon scatters from a
505: potential. Matching induces these operators to account for the difference
506: between the full and effective theory graphs for Compton scattering off two
507: quarks.  Thus, the treatment of $i\epsilon$'s does affect the correspondence
508: between soft-static and soft graphs. The total contribution of the graphs in the
509: static theory with $k^\mu\sim mv$ gluons is reproduced in the effective theory
510: by graphs with soft gluons.
511: 
512: %
513: %Comment on renormalization of soft vertices, \cite{bauer}.
514: %
515: 
516: A real difference between the soft-static and effective theory calculations is
517: the way in which counterterms are implemented. The soft-static potential is
518: defined by local HQET-like Feynman rules and all UV divergent contributions from
519: soft gluons are absorbed into vertex, field, and coupling renormalization.  The
520: renormalization of the four point function is taken care of by the
521: renormalization of the two and three point functions.  Renormalization of the
522: vNRQCD diagrams is quite different because potential gluons are not treated as
523: degrees of freedom.  The effective theory has graphs with soft gluons and in
524: addition the four-quark Coulomb potential operator.  The overall divergence in
525: soft graphs, such as the one in Eq.~(\ref{eftvc1}), are absorbed by ${\cal
526: V}_c$, while subdivergences are taken care of by counterterms for the soft
527: vertices and lower order ${\cal V}_c$ counterterms. The sum of unrenormalized
528: soft-static and purely soft diagrams in the effective theory agree. So if these
529: were the only considerations, then the description of these effects would be
530: basically a matter of convenience.  However, divergences associated with
531: ultrasoft gluons can only be absorbed into ${\cal V}_c$, so at the level that
532: these gluons contribute in the effective theory it is necessary to adopt the
533: four quark operator description from the start (i.e. just below the scale $m$).
534: 
535: To calculate the NLL energy we also need the two-loop anomalous dimension for
536: ${\cal V}_c$, which is obtained from the renormalization of order $\alpha_s^3/v$
537: diagrams.  At this order there are effective theory graphs with iterations of
538: potentials, shown in Fig.~\ref{eft_p2}, and soft effective theory diagrams, as
539: in Fig.~\ref{eft_s2}.  Graphs with ultrasoft gluons do not contribute at this
540: order.
541: \begin{figure}
542:  \centerline{\hbox{\epsfxsize=3cm \epsfbox{eft2.eps}} \qquad 
543:              \hbox{\epsfxsize=3cm \epsfbox{eft3.eps}} \qquad
544:              \hbox{\epsfxsize=3cm \epsfbox{eft4.eps}} \qquad
545: 	     \hbox{\epsfxsize=2.3cm \epsfbox{eft5.eps}} } \medskip\medskip
546: {\tighten\caption{Order $\alpha_s^3/v$ diagrams with potential 
547:  iterations. The $\times$ denotes an insertion of the ${\bf p^4}/8m^3$ 
548:  relativistic correction to the kinetic term.} 
549: \label{eft_p2}}
550: \end{figure}
551: \begin{figure}
552:  \centerline{\hbox{\epsfxsize=2.cm \epsfbox{eft_s1.eps}} \quad
553:              \hbox{\epsfxsize=2.cm \epsfbox{eft_s2.eps}} \quad
554:              \hbox{\epsfxsize=2.cm \epsfbox{eft_s3.eps}} \quad
555:            \lower2pt  \hbox{\epsfxsize=2.3cm \epsfbox{eft_soft3.eps}} 
556:  \raise20pt \hbox{\Large $\ \ \ldots$} } \medskip
557:  {\tighten\caption{Examples of order $\alpha_s^3/v$ diagrams with soft vertices.
558:  The vertex with a cross denotes an insertion of a one-loop counterterm. }
559:  \label{eft_s2}}
560: \end{figure}
561: The potential diagrams are finite in the ultraviolet and reproduce the Coulombic
562: singularities in perturbative QCD. The contributions from the soft diagrams can
563: be determined from the soft-static potential calculations. For the color singlet
564: channel, the UV divergences in the soft-static two-loop diagrams were calculated
565: in Ref.~\cite{Fischler} and the constant terms in
566: Refs.~\cite{Peter,Schroeder}. The sum of unrenormalized soft diagrams has the
567: form
568: \begin{eqnarray}  \label{stat1}
569:   \frac{i \alpS^3(\nu)}{\bf k^2} \frac{C_F}{4\pi}
570:    \bigg[\frac{\beta_0^2}{\epsilon^2} + \frac{\beta_1\!+\!2\beta_0 a_1}
571:    {\epsilon} 
572:    + \frac{2\beta_0^2}{\epsilon}\ln\Big( \frac{\mu_S^2}{\bf k^2}\Big) 
573:    \!+\! \beta_0^2 \ln^2\Big( \frac{\mu_S^2}{\bf k^2}\Big) \!+\!
574:    (\beta_1\!+\!2\beta_0 a_1)\ln\Big( \frac{\mu_S^2}{\bf k^2}\Big) 
575:    + a_2^{(a)} \bigg] \, .
576: \end{eqnarray}
577: The effective theory counterterm graphs give
578: \begin{eqnarray}  \label{eft_ct}
579:   -\frac{i \alpS^3(\nu)}{\bf k^2} \frac{C_F}{4\pi}
580:    \bigg[\frac{2\beta_0^2}{\epsilon^2} 
581:    + \frac{2\beta_0^2}{\epsilon} \ln\Big( \frac{\mu_S^2}{\bf k^2}\Big)
582:    + \frac{2 a_1 \beta_0}{\epsilon} + a_2^{(b)} \bigg] \,.
583: \end{eqnarray}
584: Taking the sum of Eqs.~(\ref{stat1}) and (\ref{eft_ct}) we find that up to
585: two-loops the counterterm for the color singlet Coulomb potential has the form
586: \begin{eqnarray} 
587:  Z_c = 1- \frac{\alpS(\nu)\beta_0}{4\pi\epsilon}  + 
588:  \frac{\alpS^2(\nu)}{(4\pi)^2} \bigg[ 
589:  \frac{\beta_0^2}{\epsilon^2} - \frac{\beta_1}{\epsilon} \bigg]
590:   \,,
591: \end{eqnarray}
592: where $\beta_1 = 34 C_A^2/3-4 C_F T_F n_\ell -20 C_A T_F n_\ell/3$.  The
593: $\alpha_s^2/\epsilon$ divergence is proportional to $\beta_1$, so the two-loop
594: anomalous dimension for ${\cal V}_c^{(s)}$ is determined by the two-loop
595: $\overline{\rm MS}$ $\beta$-function, and the NLL coefficient of the singlet
596: Coulomb potential is ${\cal V}_c^{(s)}(\nu) = 4\pi\ams{[2]}(m \nu)$.
597: 
598: The energy at NLL order involves including the NLL coefficient for the Coulomb
599: potential, ${\cal V}_c^{(s)}(\nu)$ in Eq.~(\ref{eLL}), and calculating the
600: matrix element of the one-loop order $1/v$ soft diagram between Coulombic states
601: [$a_c(\nu)\equiv -{\cal V}_c^{(s)}(\nu)/4\pi$]
602: \begin{eqnarray} \label{s1}
603:  i \: \Bigg\langle \begin{picture}(55,20)(-10,1)
604:   \epsfxsize=1.2cm \lower9pt \hbox{\epsfbox{soft1.eps}}
605:  \end{picture}\Bigg\rangle
606: &=& -C_F\: {\alpS^2(\nu)} \left\langle \frac{1}{\bf k^2}
607:  \bigg[ a_1 + \beta_0 \ln\Big( \frac{\mu_S^2}{\bf k^2} \Big) \bigg]
608: \right\rangle 
609:   \\[5pt]
610: % &&\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! 
611: &=& - \frac{m\,C_F\,\alpS^2(\nu) \tb{ a_c(\nu)} }{8\pi\,n^2}\: 
612:   \Bigg\{ {a_1} + 2 \beta_0\,\bigg[ \ln\Big(\frac{n\: \nu}{\tb{ a_c(\nu) }}\Big) 
613:   +\psi(n+l+1)+\gamma_E\bigg] \Bigg\} \,. \nn
614: \end{eqnarray}
615: As expected, at the low scale $\nu\simeq v_b$, there are no large logarithms in
616: the matrix element. Combining Eq.~(\ref{s1}) with Eq.~(\ref{eLL}) gives the
617: energy valid at NLL order,
618: \begin{eqnarray} \label{eNLL}
619:   \Delta E^{LL}+\Delta E^{NLL} &=&  -\frac{m}{4n^2}\: \big[\tb{ a_c(v_b) }\big]^2
620:  -\frac{m\,C_F\,\alpS^2(v_b) \tb{ a_c(v_b) }}{8\pi\,n^2}\: 
621:  \bigg\{  2\beta_0 \Big[ \psi(n+l+1)+\gamma_E \Big] + {a_1} \bigg\} \,.\nn\\
622: \end{eqnarray}
623: 
624: In the next section, the three-loop running of the Coulomb potential will be
625: derived. We therefore need the two loop matching condition, and so consider the
626: finite parts for the two loop graphs. The sum of renormalized soft diagrams in
627: Fig.~\ref{eft_s2} is
628: \begin{eqnarray}  \label{eftvc2}
629:   \frac{i \alpS^3(\nu)}{\bf k^2} \frac{C_F}{4\pi} \bigg[\beta_0^2
630:    \ln^2\Big( \frac{\mu_S^2}{\bf k^2}\Big) +
631:    (\beta_1+2\beta_0 a_1)\ln\Big( \frac{\mu_S^2}{\bf k^2}\Big) 
632:    + a_2 \bigg] \,,
633: \end{eqnarray}
634: where from Ref.~\cite{Schroeder} the sum of constants in Eqs.~(\ref{stat1}) and
635: (\ref{eft_ct}) is $a_2=a_2^{(a)}+a_2^{(b)} = 456.75-66.354 n_\ell+1.235
636: n_\ell^2$ for $n_\ell$ light flavors.  The matching coefficient for ${\cal V}_c$
637: at the scale $m$ is given by the difference between the $1/{\bf k}^2$ terms in
638: the $Q\bar Q$ scattering amplitude in the full and effective theories.  It is
639: convenient to analyze the two loop result in the full theory by using regions in
640: the threshold expansion~\cite{Beneke}. The soft region exactly reproduces the
641: result from the soft graphs. Furthermore, the potential region exactly
642: reproduces the results for the potential graphs in Fig.~\ref{eft_p2}. Thus, the
643: matching correction for ${\cal V}_c(1)$ is also zero at two-loops. In general, a
644: non-zero matching correction appears when there is a full theory contribution
645: from an off-shell region such as the hard regime or when UV divergences appear
646: in the effective theory graphs.\footnote{\tighten An example where UV
647: divergences in the effective theory affect the matching is the two-loop
648: coefficient for the production current.}  In the full theory at two loops there
649: are no contributions proportional to $1/{\bf k}^2$ from off-shell regions. The
650: soft effective theory graphs are UV divergent, however these divergences are in
651: one-to-one correspondence with UV divergences in the full or static theory.
652: Finally, the graphs with iterations of potentials are UV finite.
653: 
654: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
655: 
656: \section{Three-loop running of ${\cal V}_{\lowercase{c}}$}\label{sec:threeloop}
657: 
658: To compute the three-loop anomalous dimension for the Coulomb potential we need
659: to evaluate the UV divergent graphs in the effective theory that are order
660: $\alpha_s^4/v$. We begin by considering diagrams with an ultrasoft gluon. In
661: Coulomb gauge we have graphs with ${\bf p}\cdot {\bf A}/m$ vertices as well as
662: the coupling of ultrasoft gluons to the Coulomb potential from the
663: operator~\cite{amis3}
664: \begin{eqnarray} \label{Lpu}
665:   {\cal L} &=& {2 i\: \tb{ {\cal V}_c^{(T)} } \, f^{ABC}\over {\mathbf
666:   k}^4}\mu_S^{2\epsilon}\mu_U^\epsilon \:
667:     {\mathbf k}\cdot (g {\bf A}^C) \: \psip{p^\prime}^\dagger\:
668:   T^A {\psip p}\: \chip{-p^\prime}^\dagger\: \bar T^B {\chip {-p}}{} \,.
669: \end{eqnarray}
670: All graphs with two ${\bf p}\cdot {\bf A}$ vertices are UV finite or
671: are canceled by two loop graphs with insertions of the one-loop counterterms
672: for ${\cal V}_2$ and ${\cal V}_r$ computed in Ref.~\cite{amis}.  The remaining
673: diagrams are shown in Figs.~\ref{fig_us1} and \ref{fig_us2}.  Graphs
674: \ref{fig_us1}a through \ref{fig_us1}e have UV subdivergences which are
675: exactly canceled by the diagram with  ${\cal V}_k$ counterterms shown in 
676: Fig.~\ref{fig_us1}f.  These graphs contain subdivergences that were responsible
677: for the running of the $1/(m|{\bf k}|)$ potential at
678: two-loops~\cite{amis3}. Graph \ref{fig_us1}g is UV finite.
679: 
680: \begin{figure}
681:  \centerline{
682:  \hbox{\epsfxsize=3.5cm\epsfbox{usoft2.eps}} \quad
683:  \hbox{\epsfxsize=3.5cm\epsfbox{usoft3.eps}} \quad 
684:  \hbox{\epsfxsize=3.5cm\epsfbox{usoft1.eps}} } \medskip\medskip
685:  \centerline{
686:  \hbox{\epsfxsize=3.5cm\epsfbox{usoft5.eps}} \quad
687:  \hbox{\epsfxsize=3.5cm\epsfbox{usoft6.eps}} \quad
688:  \hbox{\epsfxsize=3.5cm\epsfbox{usoftct.eps}}\quad
689:  \hbox{\epsfxsize=3.5cm\epsfbox{usoft7.eps}}  
690:  } \medskip\medskip
691:  {\tighten \caption{Graphs with ultrasoft gluons which do not contribute to the
692:  running of the Coulomb potential. The divergences in a)-e) are 
693:  canceled by graph f) which has an insertion of the corresponding ${\cal V}_k$ 
694:  counterterm(s) denoted by $\otimes$. Graph g) is UV finite.}
695:  \label{fig_us1} }
696: \end{figure}
697: 
698: The divergent diagrams with an ultrasoft gluon which are not completely
699: canceled by a counterterm diagram are shown in Fig.~\ref{fig_us2}.
700: \begin{figure}
701:  \centerline{
702:  \hbox{ \epsfxsize=4.2cm\epsfbox{eftus2.eps}} \qquad 
703:  \hbox{\epsfxsize=4.2cm\epsfbox{eftus1.eps}} \qquad
704:  \hbox{\epsfxsize=3.8cm\epsfbox{eftus_ct.eps}} 
705:  }
706:  \medskip\medskip
707:  {\tighten \caption{Graphs with an ultrasoft gluon which contributes to the 
708:  three-loop running of the Coulomb potential.} \label{fig_us2} }
709: \end{figure}
710: Consider the three-loop graph in Fig.~\ref{fig_us2}a with momenta $l$ for the
711: loop with the ultrasoft gluon and $k$ and $q$ for the remaining loops.  After
712: performing the $k^0$ and $q^0$ integrals by contours, the loop integration
713: involving $l$ is
714: \begin{eqnarray}
715:    \int d^d l \: \frac{ \delta^{ij}-{\mathbf l^i l^j/l^2}}{
716:   l^2 (l^0+{E}-{\bf k^2}/{m}) (l^0+{E}-{\bf q^2}/{m})    } \,.
717: \end{eqnarray}
718: From this expression we see that the ultrasoft momentum $l$ and potential
719: momentum, ${\bf k}$ and ${\bf q}$, are not completely separable since they
720: appear in the same propagator.  The $l$ integration produces an UV divergence
721: while the remaining integrations are UV and IR finite\footnote{\tighten For
722: static quarks this three-loop graph also has an IR divergence~\cite{static1},
723: but in the non-static case we find that this divergence is regulated by the
724: quark kinetic energy.}. Evaluating the remaining integrals gives
725: \begin{eqnarray} \label{us4}
726:  {\rm Fig.~\ref{fig_us2}a} &=& \frac{4i}{3} ({\cal C}_{\ref{fig_us2}a}) 
727:   \frac{\Big[{\cal V}_c^{(T)}(\nu)\Big]^3 \alpU(\nu)
728:   \mu_S^{2\epsilon}}{(4\pi)^3\: \bf k^2} \bigg[ \frac{1}{\epsilon} 
729:   + \ln\Big(\frac{\mu_U^2}{E^2}\Big) + 2\ln\Big(\frac{\mu_S^2}{\bf k^2}\Big) 
730:   + \ldots \bigg] \,,
731: \end{eqnarray}
732: where the color factor is
733: \begin{eqnarray} 
734:   ({\cal C}_{\ref{fig_us2}a}) &=& C_A C_1 \bigg[ \frac{(C_A\!+\!C_d)}{8} 
735:   1\otimes 1 + T\otimes\bar T \bigg] \,,
736: \end{eqnarray}
737: and for gauge group SU($N_c$), $C_d=N_c-4/N_c$ and $C_1=(N_c^2-1)/(4 N_c^2)$.
738: The graph in Fig.~\ref{fig_us2}c involves the iteration of a $1/{\bf k^2}$
739: potential and a ${\cal V}_k$ counterterm and also has a Coulombic infrared
740: divergence.  This graph cancels the corresponding product of IR and UV
741: divergences arising in Fig.~\ref{fig_us2}b. The sum of graphs in
742: Fig.~\ref{fig_us2}b,c still has an UV divergence, and we find
743: \begin{eqnarray} \label{us3}
744: {\rm Fig.~\ref{fig_us2}b+\ref{fig_us2}c} &=&
745:   -\frac{4i }{3}  ({\cal C}_{\ref{fig_us2}bc}) \frac{\Big[{\cal V}_c^{(T)}(\nu)
746:  \Big]^3 \alpU(\nu)\mu_S^{2\epsilon}}{(4\pi)^3\: \bf k^2} 
747:   \bigg[ \frac{1}{\epsilon} + \ln\Big(\frac{\mu_U^2}{E^2}\Big) 
748:   + 2\ln\Big(\frac{\mu_S^2}{\bf k^2}\Big) + \ldots \bigg] \,,
749: \end{eqnarray}
750: where the color factor is
751: \begin{eqnarray} 
752:   ({\cal C}_{\ref{fig_us2}bc}) &=& C_A \bigg[ \frac{C_1 (C_A\!+\!C_d)}{8} 
753:   1\otimes 1 - \Big( C_1 + \frac{(C_A\!+\!C_d)^2}{32} \Big) T\otimes\bar T
754:   \bigg] \,.
755: \end{eqnarray}
756: The sum of divergences in Eqs.~(\ref{us4}) and (\ref{us3}) are canceled by a
757: three-loop counterterm for ${\cal V}_c$. Differentiating with respect
758: to $\ln\mu_S$ and $\ln\mu_U$ gives the anomalous dimensions
759: \begin{eqnarray}
760:   \gamma_U &=&  \frac{8}{3} \bigg[ 2 C_A C_1+\frac{C_A(C_A+C_d)^2}{32} 
761:    \bigg] \frac{\Big[{\cal V}_c^{(T)}(\nu)\Big]^3}
762:   {(4\pi)^3}\: \alpU(\nu)\  \left(T\otimes\bar T\right) \,,\nn\\
763:   \gamma_S &=&  \frac{16}{3} \bigg[ 2 C_A C_1+\frac{C_A(C_A+C_d)^2}{32} 
764:    \bigg]  \frac{\Big[{\cal V}_c^{(T)}(\nu)\Big]^3}
765:   {(4\pi)^3}\: \alpU(\nu)\ \left(T\otimes\bar T\right) \,.
766: \end{eqnarray}
767: The total anomalous dimension from the ultrasoft diagrams is
768: $\gamma=2\gamma_U+\gamma_S$, so
769: \begin{eqnarray} \label{ad1}
770:   \gamma &=& \frac{32}{3} \bigg[ 2 C_A C_1+\frac{C_A(C_A+C_d)^2}{32} 
771:    \bigg]   \frac{\Big[{\cal V}_c^{(T)}(\nu)\Big]^3}
772:   {(4\pi)^3}\: \alpU(\nu)\ \left(T\otimes\bar T\right) \,.
773: \end{eqnarray}
774: 
775: \OMIT{The UV divergence in the ultrasoft graphs is matched by an IR divergence
776: in the purely soft diagrams at order $\alpha_s^4/v$. This IR divergence is fake;
777: it is not associated with $\lqcd$ physics but originates from the separation of
778: soft and ultrasoft gluon fields. It induces an additional purely soft anomalous
779: dimension (as discussed further in Appendix~\ref{UVIR}):}
780: 
781: The presence of an ultraviolet divergence in the ultrasoft graphs induces  
782: an additional ultraviolet divergence in the soft graphs (as discussed 
783: further in Appendix~\ref{UVIR}).  This divergence induces an additional
784: contribution to the soft anomalous dimension:
785: \begin{eqnarray} \label{ad2}
786:   \gamma_S &=& -{8} \bigg[ 2 C_A C_1+\frac{C_A(C_A+C_d)^2}{32} 
787:    \bigg]\:  \alpS^4(\nu)\  \left(T\otimes\bar T\right) \,.
788: \end{eqnarray}
789: For the Coulomb potential the remaining UV divergences in the soft graphs
790: correspond to divergences which are canceled in the static calculation by field,
791: vertex, and coupling renormalization. As discussed before, these divergences
792: give a contribution proportional to the three-loop $\msb$ $\beta$-function, so
793: for the color singlet channel we have the additional contribution
794: \begin{eqnarray} \label{ad3}
795:   \gamma_S^{(s)} &=& 2 C_F \beta_2 \frac{\alpS^4(\nu)}{(4\pi)^2} \,.
796: \end{eqnarray}
797: For QCD, $\beta_2=2857/2-5033 n_\ell/18 + 325 n_\ell^2 /54$ for $n_\ell$ light flavors.
798: 
799: Combining Eqs.~(\ref{ad1})--(\ref{ad3}) in the color singlet channel and using
800: the LL relation ${\cal V}_c^{(T)}(\nu)=4\pi\alpS(\nu)$ gives the total
801: anomalous dimension for the Coulomb potential to three-loop order
802: \begin{eqnarray} \label{adfin}
803:  \gamma_{\rm total}^{(s)} &=& 2 C_F \bigg[ \beta_0 \alpS^2(\nu) 
804:     +\beta_1 \frac{\alpS^3(\nu)}{4\pi}
805:     +\beta_2 \frac{\alpS^4(\nu)}{(4\pi)^2} \bigg] \nn\\
806:   && - {C _A^3 C_F } 
807:   \bigg[ \frac{4}{3} \alpS^3(\nu)\alpU(\nu)-\alpS^4(\nu) \bigg] \,.
808: \end{eqnarray}
809: Solving this equation with the two-loop boundary condition ${\cal V}_c^{(s)}(1)
810: = -4\pi C_F \alpha_s(m)$, the NNLL result for the running Coulomb
811: potential is
812: \begin{eqnarray} \label{Vcfin}
813:   {\cal V}_c^{(s)}(\nu) &=& -4\pi C_F \ams{[3]}(m\nu) 
814:  +\frac{8\pi C_F C_A^3}{3\beta_0} \: \alpha_s^3(m)\: 
815:  \bigg[ \frac{11}{4}- 2z- \frac{z^2}{2} - \frac{z^3}{4} + 4\ln(w) \bigg] \,, 
816: \end{eqnarray}
817: where 
818: \begin{eqnarray}
819:   z=\frac{ \ams{[1]}(m\nu) }{ \alpha_s(m) }\,,\qquad\quad
820:   w=\frac{ \ams{[1]}(m\nu^2) }{ \ams{[1]}(m\nu) } \,.
821: \end{eqnarray}
822: 
823: To compare to the static potential result from Ref.~\cite{PSstat} we can expand
824: the running couplings in $\alpha_s(m)$. Ignoring the $\ams{[3]}(m\nu)$ term, the
825: remaining logarithms in the ${\cal V}_c^{(s)}(\nu)$ coefficient are
826: \begin{eqnarray} \label{expVc}
827:   -\frac{1}{3}\, {C_F C_A^3}\,  \alpha_s^4(m) \ln(\nu) 
828:   +\frac{2\beta_0}{3\pi}\, {C_F C_A^3}\, \alpha_s^5(m) \ln^2(\nu) 
829:   -\frac{17\beta_0^2}{18\pi^2}\, {C_F C_A^3}\, \alpha_s^6(m) \ln^3(\nu) 
830:   +  \ldots \,.
831: \end{eqnarray} 
832: Expanding the analogous contribution to the static result in  Eq.~(\ref{psVc}) 
833: from Ref.~\cite{PSstat}, taking $\mu=m\nu^2$ and $r=1/m\nu$, gives
834: \begin{eqnarray} \label{expVstat}
835:  -\frac{1}{3}\, {C_F C_A^3}\,  \alpha_s^4(m) \ln(\nu) 
836:   +\frac{3\beta_0}{4\pi}\, {C_F C_A^3}\, \alpha_s^5(m) \ln^2(\nu) 
837:   -\frac{77\beta_0^2}{72\pi^2}\, {C_F C_A^3}\, \alpha_s^6(m) \ln^3(\nu) 
838:   +  \ldots \,.
839: \end{eqnarray}
840: From Eqs.~(\ref{expVc}) and (\ref{expVstat}) we see that the single $\ln(\nu)$
841: terms agree, but the higher order logarithms differ.  
842: 
843: The origin of the difference between the running of the Coulomb and static
844: potentials is the relation between scales for the case of moving versus fixed
845: quarks. For the Coulomb potential the non-relativistic quarks obey the
846: dispersion relation $E={\bf p}^2/(2m)$, which relates the energy and momentum
847: scales, and logarithms of $E$, ${\bf p}$ and ${\bf k}$ cannot be treated
848: independently. The anomalous dimension in Eq.~(\ref{adfin}) generates
849: $\ln^2(\nu)$ terms that reproduce the double logarithms that appear when
850: diagrams such as the ones in Fig.~\ref{fig_usv} are evaluated at the hard scale
851: $m$.  The graph in Fig.~\ref{fig_usv}a includes a vacuum polarization loop for
852: the ultrasoft gluon which only sees the scale $mv^2$, while the soft loop in
853: Fig.~\ref{fig_usv}b sees only the scale $mv$. Even though each of the two
854: logarithms in Fig.~\ref{fig_usv}b comes from a different ratio of low energy
855: scales, the modes in the graph must be included at the same time since in
856: the effective theory the division between soft and ultrasoft modes (and the
857: multipole expansion) occur right at the scale $m$. Furthermore, for all scales
858: below $m$ the couplings for soft and ultrasoft gluons run at different rates.
859: \begin{figure}
860:  \centerline{
861:  \hbox{ \epsfxsize=4.2cm\epsfbox{eftusv1.eps}} \qquad 
862:  \hbox{\epsfxsize=4.2cm\epsfbox{eftusv2.eps}} 
863:  }
864:  \medskip\medskip
865:  {\tighten \caption{Graphs with double logarithms that are determined from 
866: Fig.~\ref{fig_us2}a.} \label{fig_usv} }
867: \end{figure}
868: As mentioned in the introduction, this correlation of $E$, ${\bf p}$ and the
869: momentum transfer ${\bf k}$ has been tested successfully for bound
870: states in QED~\cite{amis4}. 
871:  
872: In contrast, consider the situation with two static quarks where the distance
873: between them, $r\sim 1/|{\bf k}|$, is held fixed externally, and the energy
874: fluctuations are about $E=0$. In this case the scales $r$ and $E$ are not
875: correlated.  Furthermore, operators with powers of $1/m$ play no role in the
876: calculation of the anomalous dimension, unlike the Coulombic case with an
877: expansion in the velocity.  The difference between the static and non-static
878: calculations occurs essentially because neither the $m\to \infty$ nor the $v\to
879: 0$ limit of the effective theory is the same as the static theory.
880: 
881: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
882: 
883:   
884: \section{The NNLL energy for a $Q \overline Q$ bound state}\label{sec:nnll}
885: 
886: The energy at NNLL has contributions from matrix elements of operators of 
887: order
888: \begin{eqnarray}
889:  && \qquad  \Big\{ \frac{\alpha_s^4}{v^2},\ \frac{\alpha_s^3}{v},\ 
890:   {\alpha_s^2}v^0,\ \alpha_s v,\ v^2 \Big\}\ 
891: % \sum_{k=0}^\infty (\alpha_s\ln\alpha_s)^k 
892: \,. \nn  
893: \end{eqnarray}
894: There are contributions from tree level matrix elements, which include the ${\bf
895: p^4}/m^3$ operator ($\sim v^2$), the order $v$ potentials with LL coefficients
896: ($\sim \alpha_s v$), the order $v^0$ potentials with NLL coefficients ($\sim
897: \alpha_s^2 v^0$), and the correction to the energy from the Coulomb potential
898: with the NNLL coefficient ($\sim \alpha_s^3/v$). We also have the matrix element
899: of the order $\alpha_s^3/v$ one loop and two loop soft diagrams in
900: Fig.~\ref{eft_s2}.  Finally, there are the double insertions of two
901: $\alpha_s^2/v$ soft diagrams ($\sim \alpha_s^4/v^2$).  For simplicity all
902: $n_\ell$ light quarks are taken to be massless.
903: 
904: The contributions from tree level matrix elements are 
905: [$a_c(\nu)=-{\cal V}_c^{(s)}(\nu)/4\pi$]
906: \begin{eqnarray}
907: \Big\langle V^{(0)}\Big\rangle &=& 
908:   \frac{m}{4n^3(2l+1)}\:\big[\tb{ a_c(\nu) } \big]^2\:
909:    \tb{ {\cal V}_k^{(s)}(\nu) } \,,  \\[5pt]
910: \Big\langle  V^{(1)}  \Big\rangle 
911:   &=& \frac{m}{4n^3} \big[{\tb{ a_c(\nu) } } \big]^3 
912:  \Bigg\{ \frac{ \delta_{l0}}{2\pi} \:
913:   \Big[ \tb{ {\cal V}_2^{(s)}(\nu) } + s(s+1) \tb{ {\cal V}_s^{(s)}(\nu) } \Big]
914:   - \frac{ (2l+1-4n)}{8\pi n (2l+1)}\: \tb{ {\cal V}_r^{(s)}(\nu) }  \nn\\[3pt]
915:  &+&\frac{ %(j^2\!+\!j\!-\!l^2\!-\!l\!-\!2)
916:   X_{ljs}\,\delta_{s1}\,(1-\delta_{l0})}
917:  {4\pi l (l+1)(2l+1)}\: \tb{ {\cal V}_\Lambda^{(s)}(\nu) } 
918:   + \frac{3 \langle S_{12}\rangle_{ljs}\,\delta_{s1}\,(1-\delta_{l0}) }
919:   {4\pi l (l+1)(2l+1)}\: \tb{ {\cal V}_t^{(s)}(\nu) } \Bigg\} \,, \nn \\[5pt]
920: \Bigg\langle \frac{-\bf p^4}{4 m^3} \Bigg\rangle &=&  
921:   \frac{m}{16n^3}\:\big[\tb{ a_c(\nu) }  \big]^4
922:   \:\bigg[ \frac{3}{4n}- \frac{2}{2l+1} \bigg]  \,, \nn
923: \end{eqnarray}
924: where the Wilson coefficients ${\cal V}^{(s)}_{c,k,2,s,r,\Lambda,t}$ are defined
925: in section~\ref{sec:pots}. The matrix elements of the soft loops are
926: \begin{eqnarray}
927:   && \left\langle\: i\!\!\!\! \begin{picture}(55,20)(-10,1)
928:   \epsfxsize=1.2cm \lower13pt \hbox{\epsfbox{eft_s1.eps}}
929:  \end{picture} \!\!\!+\ldots \right\rangle
930:   =   -\frac{m C_F\alpS^3(\nu)\,\tb{ a_c(\nu) } }{8\pi^2 n^2}\: \Bigg\{ 
931:   \Big( \frac{\beta_1}{2}+\beta_0 a_1 \Big) \Bigg( \ln\bigg(\frac{n\:\nu}
932:   {\tb{ a_c(\nu) }}\bigg) +\Psi(n+l+1) +\gamma_E\Bigg)   \nn\\
933:   && \quad\quad
934:   + \frac{a_2}{4} +\beta_0^2\: \Bigg( \ln^2\bigg(\frac{n\:\nu}{\tb{ a_c(\nu) }}
935:   \bigg) + 2 \ln\bigg(\frac{n\:\nu}{\tb{ a_c(\nu) }}\bigg)
936:  \Big[ \Psi(n+l+1)+\gamma_E \Big]+ N_{2}(n,l)  
937:   \Bigg) \Bigg\} \,, \\[7pt]
938: && 
939:  \Bigg\langle   T\bigg\{  
940: i \begin{picture}(45,20)(-5,1) 
941:  \epsfxsize=1.0cm \lower8pt \hbox{\epsfbox{soft1.eps}} \end{picture} 
942:  \!\!\! ,\  
943: % \!\!\! \mbox{$\sum$} \bigg| \bigg\rangle   \bigg\langle \bigg| \ 
944: i\! \begin{picture}(45,20)(-5,1) 
945:  \epsfxsize=1.0cm \lower8pt \hbox{\epsfbox{soft1.eps}} \end{picture}\!\!\! 
946:  \bigg\}
947:  \Bigg\rangle 
948: = -\frac{m C_F^2\alpS^4(\nu)}{16\pi^2 n^2}\: \left\{\frac{a_1^2}{4}+ \beta_0 a_1 
949:  \Bigg( \ln\bigg(\frac{n\:\nu}{\tb{ a_c(\nu) }}\bigg) \!+\! 2 N_1(n,l) \!+\! 
950:  \gamma_E \Bigg) \right. \nn\\
951:  && \qquad \left. + \beta_0^2\Bigg(\ln^2\bigg(\frac{n\:\nu}{\tb{ a_c(\nu) }}
952:   \bigg) + 4\ln\bigg(\frac{n\:\nu}{\tb{ a_c(\nu) }}\bigg) 
953:  \Big[ N_1(n,l)+\frac{\gamma_E}{2} \Big] + 4 N_0(n,l)+4\gamma_E N_1(n,l)
954:  +\gamma_E^2   \Bigg)  \right\}   \,, \nn  %\\ \phantom{x}
955: \end{eqnarray}
956: where the functions $N_{0,1,2}(n,l)$, $\langle S_{12}\rangle_{ljs}$ and
957: $X_{ljs}$ are obtained from Refs.~\cite{Yndurain,PY} and are summarized in
958: Appendix~\ref{Ns}.  The only additional contribution is the result for the NLL
959: energy in Eq.~(\ref{eNLL}) with the Wilson coefficients evaluated at one higher
960: order.
961: 
962: Again, at the scale $\nu=v_b$ there are no large logarithms in the matrix
963: elements; the large logarithms are summed up into the Wilson coefficients. For
964: $\nu=v_b$ the energy at NNLL order in terms of the pole mass $m$ reads
965: \begin{eqnarray} \label{eNNLL}
966: \Delta E &=& \Delta E^{LL} + \Delta E^{NLL}+ \Delta E^{NNLL} \\[2mm]
967:  &=&  -\frac{m}{4n^2}\: \big[ \tb{ a_c(v_b) } \big]^2 
968:  - \frac{m\,C_F\,\alpS^2(v_b) \tb{ a_c(v_b)} }{(8\pi\,n^2)}
969:  \bigg[\, 2\beta_0\Big(\psi(n+l+1) + \gamma_E \Big) + {a_1} \,\bigg] \nn\\[2mm] 
970:  &+& \frac{m}{4n^3(2l+1)}\:\big[\tb{ a_c(v_b)  } \big]^2\ 
971:    \tb{ {\cal V}_k^{(s)}(v_b) } 
972:  + \frac{m}{16n^3}\: \big[\tb{ a_c(v_b) }  \big]^4
973:  \:\bigg[ \frac{3}{4n}- \frac{2}{2l+1} \bigg]  \nn\\[1mm]
974:   &+& \frac{m}{4n^3} \big[\tb{ a_c(v_b) }  \big]^3 
975:   \Bigg\{ \frac{ \delta_{l0}}{2\pi} \:
976:   \Big[\tb{ {\cal V}_2^{(s)}(v_b) } + s(s+1)\tb{ {\cal V}_s^{(s)}(v_b) } \Big] -
977:   \frac{ (2l+1-4n)}{8\pi n (2l+1)}\: \tb{ {\cal V}_r^{(s)}(v_b) } \nn\\[3pt]
978:  & & +\frac{ 
979:   %(j^2\!+\!j\!-\!l^2\!-\!l\!-\!2)
980:   X_{ljs}\delta_{s1}(1-\delta_{l0})}
981:   {4\pi l (l+1)(2l+1)}\: \tb{ {\cal V}_\Lambda^{(s)}(v_b) } 
982:   + \frac{3 \langle S_{12}\rangle_{ljs}\delta_{s1}(1-\delta_{l0}) }
983:   {4\pi l (l+1)(2l+1)}\: \tb{ {\cal V}_t^{(s)}(v_b) } \Bigg\}  \nn \\[1mm]
984:  &-& \frac{m\,C_F\,\alpS^3(v_b)\,\tb{ a_c(v_b) }}{8\pi^2 n^2} 
985:  \: \Bigg\{ \,\beta_0^2\:  N_{2}(n,l) +  \bigg( \frac{\beta_1}{2}+
986:  \beta_0 a_1 \bigg)
987: \bigg[ \Psi(n+l+1)+\gamma_E \bigg]   
988:   + \frac{a_2}{4}\,   \Bigg\} \nn \\[1mm]
989: &-& \frac{m C_F^2\alpS^4(v_b)}{16\pi^2 n^2}\: \left\{ \beta_0^2\bigg[ 4 N_0(n,l) 
990:  +4\gamma_E N_1(n,l)+\gamma_E^2 \bigg] + \beta_0 a_1 
991:  \bigg[ 2 N_1(n,l)+ \gamma_E 
992:  \bigg] +  \frac{a_1^2}{4}\right\}   \,. \nn 
993: \end{eqnarray}
994: When this expression is expanded in powers of $\alpha_s$ at a given
995: renormalization scale the LL, NLL, and NNLL predictions for the energy become
996: series in $(\alpha_s\ln\alpha_s)$ as in Eq.~(\ref{Ell}). The terms beyond NNLO
997: that are determined unambiguously are those up to $m\alpha_s^4 (\alpha_s
998: \ln\alpha_s)^k$, $k\ge 1$.  From Eq.~(\ref{eNNLL}) we see that up to NLL the
999: series are determined by the running of $\ams{}(\mu)$ since up to NLL order we
1000: have $a_c(\nu)/C_F=\alpS(\nu)=\ams{}(m\nu)$.  At NNLL order the series is no
1001: longer just determined by the QCD $\beta$-function since operators besides the
1002: strong coupling have non-trivial anomalous dimensions.
1003: 
1004: It is well known that the convergence of predictions for $\Delta E$ in terms of
1005: the pole mass are plagued by the presence of infrared renormalons. If
1006: predictions are made in terms of a short distance mass such as the $\msb$ mass
1007: the leading renormalon in the pole mass and $1/{\bf k}^2$ potentials
1008: cancel~\cite{renorm1,renorm2} and the convergence of the perturbation series is
1009: improved.  A phenomenological analysis, which includes the issue of renormalon
1010: cancellation in the presence of resummed logarithms, will be carried out
1011: elsewhere.
1012: 
1013: Past finite order predictions for the $Q\bar Q$ energies have typically been
1014: made using the strong coupling evaluated at the soft scale $mv$.  At LO and NLO
1015: this is the natural choice since all logarithms $\alpha_s^2 (\alpha_s\ln\mu)^k$
1016: and $\alpha_s^3 (\alpha_s\ln\mu)^k$ in higher order matrix elements are
1017: minimized at $\mu=|{\bf k}|$. Effectively this choice turns the LO and NLO
1018: results into the LL and NLL predictions.  However, at NNLO this choice of $\mu$
1019: may not be optimal since matrix elements begin to involve factors of $\alpha_s^4
1020: [\alpha_s\ln(\mu/E)]$.  The NNLL prediction is not generated by a simple
1021: replacement rule and instead involves the non-trivial Wilson coefficients ${\cal
1022: V}_{c,k,2,s,r,\Lambda,t}(\nu)$.
1023: 
1024: In Fig.~\ref{fig:nnllplots} we compare the NNLL energy predictions at
1025: subtraction velocity $\nu$ (thick black lines) with the NNLO predictions with
1026: coupling $\ams{}(m\nu)$ (thin green lines), to illustrate the impact of summing
1027: the logarithms.
1028: \begin{figure}[t] % fig:nnllplots
1029: \begin{center}
1030:  \leavevmode
1031:  \epsfxsize=6cm
1032:  \epsffile[220 580 420 710]{energyvar.eps}
1033: \vskip 4.9cm 
1034: {\tighten 
1035: \caption{\label{fig:nnllplots} Comparison of the NNLL binding energy predictions
1036: (thick black lines) with the NNLO predictions (thin green lines) for the
1037: $1\,{}^3S_1$ (solid), $1\,{}^1S_0$ (dashed), $2\,{}^3S_1$ (dot-dashed), and
1038: $2\,{}^3P_1$ (dotted) states, and for different values of the subtraction
1039: velocity $\nu$.  } }
1040: \end{center}
1041: \end{figure}
1042: The displayed states include $n\,{}^{2S+1}L_J=1\,{}^3S_1$ (solid lines),
1043: $1\,{}^1S_0$ (dashed lines), $2\,{}^3S_1$ (dash-dotted lines), and $2\,{}^3P_1$
1044: (dotted lines).  As input we have chosen $m_b=4.8$~GeV for the bottom quark pole
1045: mass, $\alpha_s^{(n_\ell=4)}(m_b)=0.22$, and have included three-loop running
1046: for the evolution of the strong coupling to lower scales.  The NNLO and the NNLL
1047: energies are equal for $\nu=1$, because no logarithms are summed into the Wilson
1048: coefficients. The summation has the largest impact on the $n=1$ S-wave states,
1049: and in all cases reduces the size of the binding energy.  In
1050: Table~\ref{table_nums} we summarize for $\nu$ of order a typical quark velocity,
1051: $\nu=(0.35,0.4)$, values for the NNLO result and the renormalization group
1052: improved NNLL calculation. Relative to the NNLO results the scale uncertainty in
1053: the NNLL predictions is somewhat reduced.
1054: 
1055: The summation of NNLL logarithms also reduces the size of the ground state
1056: hyperfine splitting, but increases the size of the analog of the QED Lamb shift,
1057: (the $2{}^3S_1\!-\!2{}^3P_1$ splitting).  For $\nu=(0.35,0.4)$ and at NNLO the
1058: hyperfine splitting and the $2{}^3S_1\!-\!2{}^3P_1$ splitting are
1059: \begin{eqnarray}
1060:   E(1{}^3S_1) - E(1{}^1S_0)=(68,52)~{\rm MeV} \,, \qquad
1061:   E(2{}^3S_1)- E(2{}^3P_1) =(39,35)~{\rm MeV} \,,
1062: \end{eqnarray}
1063: while at NNLL order we find
1064: \begin{eqnarray}
1065:   E(1{}^3S_1) - E(1{}^1S_0)=(27,27)~{\rm MeV} \,, \qquad
1066:   E(2{}^3S_1) - E(2{}^3P_1)=(64,47)~{\rm MeV} \,.
1067: \end{eqnarray}
1068: Summing the logarithms reduces the perturbative contribution to the hyperfine
1069: splitting by a factor of two.  The $2{}^3S_1\!-\!2{}^3P_1$ splitting is fairly
1070: sensitive to the value of $\nu$.
1071: 
1072: \begin{table}[t!]
1073: \begin{center}
1074: \begin{tabular}{cll|cccccc}
1075:  & $b\bar b$ state $\: (n{}^{2S+1}L_J)$ & & $1{}^1S_0$ & $1{}^3S_1$ & $2{}^3S_1$ & $2{}^3P_1$ \\ 
1076:  \hline
1077:  & $\Delta E$ at NNLO (MeV) & 
1078:   $\nu=0.35\ \ $ & $-1040$ & $-972$ & $-449$ & $-488$ \\
1079:  & \hspace{0.9cm}\raisebox{0.2cm}{} &  
1080:   $\nu=0.4$      & $-912$  & $-860$ & $-389$ & $-423$  \\ \hline
1081:  & $\Delta E$ at NNLL (MeV) & 
1082:   $\nu=0.35\ \ $ & $-641$ & $-614$ & $-382$ & $-446$ \\
1083:  & \hspace{0.9cm}\raisebox{0.2cm}{} &  
1084:   $\nu=0.4$      & $-705$ & $-678$ & $-356$ & $-403$ 
1085: \end{tabular}
1086: \end{center}
1087: {\tighten \caption{Comparison of the NNLO and NNLL predictions for the binding
1088:  energy corrections (in MeV) for $b\bar b$ states. The NNLL results include the
1089:  summation of logarithms not accounted for by $\ams{}(m\nu)$. Results are
1090:  shown for two values of $\nu$ of order the heavy quark velocity. }
1091: \label{table_nums} }
1092: \end{table}
1093: \OMIT{
1094: \begin{table}[t!]
1095: \begin{center}
1096: \begin{tabular}{cll|cccccc}
1097:  & $b\bar b$ state $\: (n{}^{2S+1}L_J)$ & & $1{}^1S_0$ & $1{}^3S_1$ & $2{}^3S_1$ & $2{}^3P_1$ \\ 
1098:  \hline
1099:  & $\Delta E^{NNLL}-\Delta E^{NNLO}$ & 
1100:   $\nu=0.35\ \ $ & $399$ & $358$ & $67$ & $42$ \\
1101:  & \hspace{0.9cm}\raisebox{0.2cm}{(in MeV)} &  
1102:   $\nu=0.4$ & $207$ & $183$ & $33$ & $20$ 
1103: \end{tabular}
1104: \end{center}
1105: {\tighten \caption{Reduction in the energy (in Mev) for $b\bar b$ states from
1106:  the summation of logarithms not accounted for by $\ams{}(m\nu)$.  The shifts
1107:  are shown for two values of $\nu$ of order the heavy quark velocity. }
1108: \label{table_nums} }
1109: \end{table}
1110: }
1111: 
1112: In order to examine the size of the logarithmic terms that are summed at NNLL
1113: order we can expand Eq.~(\ref{eNNLL}) in powers of $\alpha_s$.  To suppress
1114: logarithms proportional to the QCD $\beta$-function in the energy at NNLO is it
1115: convenient to expand in the coupling $a_s \equiv\ams{} (m v_b)$:
1116: \begin{eqnarray}\label{eNNLLexpanded}
1117:  \Delta E = &-& \, m\,a_s^2\Big[\,\dots\,\Big]
1118:   - m\,a_s^3\Big[\,\dots\,\Big]
1119:   - m\,a_s^4\Big[\,\dots\,\Big]  \\[2mm]
1120:   &-& m\,a_s^5\,\ln a_s\, \frac{C_F^2}{4\,\pi\,n^2}
1121:   \Bigg\{\, \frac{C_A}{3}\,\bigg[\, \frac{C_A^2}{2} + 
1122:   \frac{4\,C_A\,C_F}{n\,(2l+1)} + \frac{2\,C_F^2}{n}\,\bigg(\frac{8}{2l+1}
1123:   -\frac{1}{n}\bigg) \,\bigg] \nn\\
1124: & &\,\, +\, \frac{3\,\delta_{l0}\,C_F^2}{2\,n}(C_A\!+\!2C_F)
1125:  - \frac{7\,C_A\,C_F^2\,\delta_{l0}\,\delta_{s1}}{3\,n}
1126:  - \frac{C_A\,C_F^2\,(1\!-\!\delta_{l0})\,\delta_{s1}}
1127:         {4\,n\,l\,(l\!+\!1)\,(2l+1)}\,\Big(\,4\,X_{ljs} 
1128:    \!+\! \langle S_{12}\rangle_{ljs}\,\Big) \,\Bigg\} \nn\\[2mm]
1129: %
1130:  &-& m\,a_s^6\,\ln^2 a_s\,  \frac{C_F^2}{4\,\pi^2\,n^2}
1131: \Bigg\{\,  \frac{\delta_{l0}\,C_F^2}{6\,n}\,\bigg[\,
1132:   \beta_0\,\bigg(\frac{13\,C_A}{2}-C_F\bigg)
1133:   +\frac{C_A}{3}\,\bigg(25\,C_A+22\,C_F\bigg)
1134: \,\bigg] \nn\\
1135:  & &\,\,- \frac{C_A\,C_F^2\,\delta_{l0}\,\delta_{s1}}{6\,n}\,
1136:   \Big[\, 5\,\beta_0+ 7\,C_A \,\Big] \nn\\
1137:  & &\,\,- \frac{C_A\,C_F^2\,(1-\delta_{l0})\,\delta_{s1}}
1138:         {8\,n\,l\,(l+1)\,(2l+1)}\,
1139:  \bigg[\,
1140:    \beta_0\,\Big(\,2\,X_{ljs} 
1141:               + \frac{1}{2}\,\langle S_{12}\rangle_{ljs}\,\Big)
1142:    + C_A\,\Big(\,2\,X_{ljs} 
1143:               + \langle S_{12}\rangle_{ljs}\,\Big)
1144:  \,\bigg]
1145: \,\Bigg\} \nn\\[2mm]
1146:  &+& \ldots 
1147: \,.\nn
1148: \end{eqnarray} 
1149: The $a_s^2$, $a_s^3$, $a_s^4$ terms are not displayed, but agree with the result
1150: in Ref.~\cite{PY}. The terms proportional to $m\alpha_s^5\ln\alpha_s$ agree with
1151: Ref.~\cite{a5lna}. The $m\alpha_s^6\ln^2\alpha_s$ result is new, as are higher
1152: terms in the series.  Numerically we find
1153: \begin{eqnarray} \label{lnseries}
1154:   \frac{\Delta E(1{}^1S_0)}{m} &=& -0.444\, a_s^2 -1.595\, a_s^3 -9.73\, a_s^4 
1155:      -8.56\, a_s^5\ln a_s - 3.41\, a_s^6\ln^2 a_s -15.5\, a_s^7\ln^3 a_s\nn\\
1156:      && + \ldots \,, \nn\\
1157:   \frac{\Delta E(1{}^3S_1)}{m} &=& -0.444\, a_s^2 -1.595\, a_s^3 -8.68\, a_s^4 
1158:      -6.80\, a_s^5\ln a_s - 0.904\, a_s^6\ln^2 a_s -12.1\, a_s^7\ln^3 a_s\nn\\
1159:      && + \ldots \,, \nn\\
1160:   \frac{\Delta E(2{}^3S_1)}{m} &=& -0.111\, a_s^2 -0.546\, a_s^3 -3.07\, a_s^4 
1161:      -0.961\, a_s^5\ln a_s - 0.113\, a_s^6\ln^2 a_s -1.73\, a_s^7\ln^3 a_s\nn\\
1162:      && + \ldots \,, \nn\\
1163:   \frac{\Delta E(2{}^3P_1)}{m} &=& -0.111\, a_s^2 -0.644\, a_s^3 -3.22\, a_s^4 
1164:      -0.398\, a_s^5\ln a_s - 0.005\, a_s^6\ln^2 a_s -0.752\, a_s^7\ln^3 a_s\nn\\
1165:      && + \ldots \,.
1166: \end{eqnarray} 
1167: We note that for $(1{}^1S_0,1{}^3S_1,2{}^3S_1,2{}^3P_1)$ using $a_s=0.35$, the
1168: $a_s^5\ln a_s$ terms in Eq.~(\ref{lnseries}) give $(93\%,85\%,84\%,80\%)$ of
1169: the complete sum of logarithmic terms.
1170: 
1171: \section{Conclusion} \label{sec:conclusion}
1172:   
1173: The three-loop anomalous dimension for the Coulomb potential was computed in the
1174: presence of non-static quarks.  In terms of the subtraction velocity $\nu$ the
1175: anomalous dimension depends on both $\alpha_s(m \nu)$ and $\alpha_s(m
1176: \nu^2)$. Our result differs from the three-loop anomalous dimension for the
1177: potential for static quarks.  This is due to the fact that the energy and
1178: momentum scales are coupled for non-static quarks.
1179: 
1180: The perturbative energies for $Q\bar Q$ bound states were computed at NNLL
1181: order, including all terms of order $m \alpha_s^4 (\alpha_s\ln\alpha_s)^k$,
1182: $k\ge 0$.  The main effect of summing the logarithms is to reduce the binding
1183: energy for the $1{}^1S_0$ and $1{}^3S_1$ states by an amount of order a few
1184: hundred ${\rm MeV}$.  The effect on states with $n\ge 2$ is substantially
1185: smaller.
1186: 
1187: \section*{Acknowledgment}
1188: 
1189: We would like to thank A.~Pineda and J.~Soto for discussions, and C.~Bauer for
1190: comments on the manuscript.  AM and IS are supported in part by the
1191: U.S.~Department of Energy under contract~DOE-FG03-97ER40546, and IS is also
1192: supported in part by NSERC of Canada.
1193: 
1194: 
1195: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1196: 
1197: \appendix 
1198: 
1199: 
1200: %\section{Static Potential}
1201: %%%%% Static stuff %%%%%%%%%%%%%%%%%
1202: 
1203: 
1204: 
1205: \section{Ultraviolet and Infrared Divergences}
1206: \label{UVIR}
1207: 
1208: In this appendix we discuss the structure of UV and IR divergences in the
1209: effective theory and their relation to the renormalization group running. In
1210: particular we explain why it is correct to treat all $1/\epsilon$ poles in soft
1211: gluon loops as UV divergences. This result follows from the fact that we have
1212: taken the full propagating gluon field and split it into two parts (soft and
1213: ultrasoft) which fluctuate on different length scales ($mv$ and $mv^2$) and have
1214: different Feynman rules. The information that the two gluons came from a single
1215: field is reflected by a relation between the IR divergences in soft gluon
1216: diagrams and the UV divergences in the ultrasoft diagrams.  Here we address how
1217: this correlation is accounted for in the presence of the relation $\mu_U =
1218: \mu_S^2/m$ between ultrasoft and soft subtraction scales.
1219: 
1220: For simplicity only diagrams with single $1/\epsilon$ poles and either purely
1221: soft or ultrasoft gluons will be discussed. Also, real Coulombic IR divergences
1222: will be dropped. A general diagram with a divergent soft gluon loop then gives a
1223: soft amplitude with the divergence structure:
1224: \begin{eqnarray} \label{Asoft}
1225:  i{\cal A}^S &=& {A\over \epsilon_{UV}} + {B \over \epsilon_{IR}} +
1226:    C \left( {1\over \epsilon_{UV}}-{1\over \epsilon_{IR}} \right)
1227:   = {A+B \over \epsilon_{UV}} +
1228:    (C-B) \left( {1\over \epsilon_{UV}}-{1\over \epsilon_{IR}} \right)\,.
1229: \end{eqnarray}
1230: Strictly speaking pure dimensional regularization does not distinguish between
1231: UV and IR divergences; however a distinction can always be made either by
1232: examining the analytic structure before expanding about $\epsilon=0$ or by also
1233: performing the calculation with a different IR regulator.  The notation in
1234: Eq.~(\ref{Asoft}) is slightly redundant so that $C$ represents diagrams
1235: involving scaleless integrals (such as tadpole graphs), and $A$ and $B$
1236: represent all other graphs. For example, for the Lamb shift in QED for Hydrogen
1237: only $C$ is non-zero, while positronium also has non-zero $A$ and $B$. At the
1238: same order in the power counting as Eq.~(\ref{Asoft}) there is an amplitude with
1239: a divergent ultrasoft gluon loop which has the form:
1240: \begin{eqnarray} \label{Ausoft}
1241:  i{\cal A}^U  &=&  { (C-B) \over \epsilon_{UV}} + {D \over \epsilon_{IR}}\,.
1242: \end{eqnarray}
1243: The nontrivial information is that the $(C-B)$ term in Eq.~(\ref{Ausoft}) is
1244: determined by the IR divergence in the soft amplitude in Eq.~(\ref{Asoft}). In
1245: general $D$ is independent of $C-B$ since ultrasoft graphs are not always
1246: proportional to $1/\epsilon_{UV} - 1/\epsilon_{IR}$. The infrared divergence in
1247: Eq.~(\ref{Ausoft}) will match with an infrared divergence in full QCD.
1248: 
1249: The renormalization group running is determined by the UV divergences. Naively,
1250: in Eq.~(\ref{Asoft}) the $\epsilon_{UV}$'s correspond to the scale $m$ and the
1251: $\epsilon_{IR}$'s correspond to the scale $mv$, while in Eq.~(\ref{Ausoft}) the
1252: $\epsilon_{UV}$'s correspond to the scale $mv$ and the $\epsilon_{IR}$'s
1253: correspond to scales of order $mv^2$. However, examining $i{\cal A}^S+i{\cal
1254: A}^U$ we see that the $(C-B)$ term in the soft amplitude simply pulls up or
1255: transports the $1/\epsilon_{UV}$ in the ultrasoft graph to the hard
1256: scale~\footnote{\tighten This addition might seem strange since it involves
1257: canceling an UV and IR pole.  However, for a multiscale problem what is
1258: ultraviolet and what is infrared is always relative; if we label the
1259: $\epsilon$'s by the corresponding scale then the cancellation is just
1260: $1/\epsilon(mv)-1/\epsilon(mv)=0$.}. The scale dependence of the coefficient
1261: $C-B$ in Eq.~(\ref{Ausoft}) does not affect this argument since the scale
1262: dependence of $C-B$ in Eq.~(\ref{Asoft}) can be chosen arbitrary.  Note that the
1263: $1/\epsilon_{IR}$ divergences in soft diagrams do not always correspond to IR
1264: divergences in QCD, which is further evidence of their unphysical nature.  This
1265: is the case if $D\ne B-C$ such as for the two loop graphs contributing to the
1266: running of ${\cal V}_k(\nu)$~\cite{amis3}.  Finally, we see that setting
1267: $\epsilon_{IR} = \epsilon_{UV}$ in the soft amplitude in Eq.~(\ref{Asoft}) and
1268: running the ultrasoft modes from $m$ to $mv^2$ with an anomalous dimension
1269: proportional to $C-B$ and running the soft modes from $m$ to $mv$ with an
1270: anomalous dimension proportional to $A+B$ correctly performs the running between
1271: the scales.  This is the method used here and in Refs.~\cite{amis,amis4}.
1272: 
1273: The correspondence in Eqs.~(\ref{Asoft}) and (\ref{Ausoft}) also provides a
1274: useful calculational tool: if the UV divergences $A+C$ in the soft diagrams are
1275: known, and the combination $B-C$ is determined from the UV divergences in the
1276: ultrasoft calculation, one arrives at $(A+C)+(B-C) = A+B$ which is the
1277: combination needed to determine the soft anomalous dimension from ${\cal A}^S$.
1278: 
1279: \section{Functions that appear in $\Delta E^{\rm NNLL}$}
1280: \label{Ns}
1281: 
1282: The following functions of the principal and orbital quantum numbers ($n,l,j$)
1283: were derived in Refs.~\cite{Yndurain,PY} and appear in our result in
1284: Eq.~(\ref{eNNLL}):
1285: \begin{eqnarray}
1286:   N_0(n,l) &=& \frac{\Psi(n+l+1)}{4}\Big[\Psi(n+l+1)-2\Big] 
1287:   + \frac{n\Gamma(n-l)}{2\Gamma(n+l+1)} \sum_{j=0}^{n-l-2} \frac{
1288:  \Gamma(j+2l+2)}{\Gamma(j+1)(j+l+1-n)^3}
1289:   \nn\\
1290:  && + \frac{n\Gamma(n+l+1)}{2\Gamma(n-l)} \sum_{j=n-l}^{\infty}  \frac{
1291:  \Gamma(j+1)}{\Gamma(j+2l+2)(j+l+1-n)^3} \,, \\[4mm]
1292:  N_1(n,l) &=&  \frac{\Psi(n+l+1)}{2} -\frac12 \,, \\[4mm]
1293:  N_2(n,l) &=& \Big[ \Psi(n+l+1) +\gamma_E \Big]^2 + \Psi'(n+l+1) 
1294:   + \frac{\pi^2}{12} \nn\\
1295:  && +\frac{ 2 \Gamma(n-l)}{\Gamma(n+l+1)} \sum_{j=0}^{n-l-2} 
1296:  \frac{ \Gamma(2l+2+j) }{\Gamma(j+1) (j+l+1-n)^2 }  \,,
1297: \\[4mm]
1298:  \langle S_{12}\rangle_{ljs} & = &
1299: \left\{
1300: \begin{array}{r@{\quad:\quad}l}
1301:  \displaystyle 
1302:  \frac{2(l+1)}{1-2l} & j=l-1 \\
1303:  2 & j=l\\
1304:  \displaystyle
1305:  \frac{-2\,l}{2l+3} & j=l+1
1306: \end{array}
1307: \right.
1308: \,,
1309: \\[4mm]
1310: X_{ljs} & = & \frac{1}{2}\Big(\,j(j+1)-l(l+1)-s(s+1)\,\Big) \,.
1311: \end{eqnarray}
1312: 
1313: 
1314: {\tighten
1315: \begin{references}
1316: 
1317: \bibitem{nrqcd}
1318: W.E.~Caswell and G.P.~Lepage,
1319: %``Effective Lagrangians For Bound State Problems In QED, QCD, And Other Field
1320: %                  Theories,''
1321: Phys. Lett. {\bf 167B}, 437 (1986);
1322: %%CITATION = PHLTA,167B,437;%%
1323: G.T.~Bodwin, E.~Braaten and G.P.~Lepage,
1324: %``Rigorous QCD analysis of inclusive annihilation and production of heavy
1325: %                  quarkonium,''
1326: Phys. Rev. {\bf D51}, 1125 (1995), ibid. {\bf D55}, 5853
1327: (1997);
1328: %%CITATION = PHRVA,D51,1125;%%
1329: P.~Labelle,
1330: %``Effective field theories for QED bound states: Extending  nonrelativistic QED to study retardation effects,''
1331: Phys.\ Rev.\ D {\bf 58}, 093013 (1998);
1332: %[hep-ph/9608491].
1333: %%CITATION = HEP-PH 9608491;%%
1334: M.~Luke and A.V.~Manohar,
1335: %``Bound states and power counting in effective field theories,''
1336: Phys. Rev. {\bf D55}, 4129 (1997);
1337: %  hep-ph/9610534.
1338: %%CITATION = PHRVA,D55,4129;%%
1339: A.V.~Manohar,
1340: %``The HQET / NRQCD Lagrangian to order alpha / m-3,''
1341: Phys. Rev. {\bf D56}, 230 (1997);
1342: %  hep-ph/9701294.
1343: %%CITATION = PHRVA,D56,230;%%
1344: B.~Grinstein and I.Z.~Rothstein,
1345: %``Effective field theory and matching in nonrelativistic gauge theories,''
1346: Phys. Rev. {\bf D57}, 78 (1998);
1347: %  hep-ph/9703298.
1348: %%CITATION = PHRVA,D57,78;%%
1349: M.~Luke and M.J.~Savage,
1350: %``Power counting in dimensionally regularized NRQCD,''
1351: Phys. Rev. {\bf D57}, 413 (1998);
1352: % hep-ph/9707313
1353: %%CITATION = PHRVA,D57,413;%%
1354: 
1355: \bibitem{pNRQCD}
1356: A.~Pineda and J.~Soto,
1357: %``Effective field theory for ultrasoft momenta in NRQCD and NRQED,''
1358: Nucl. Phys. Proc. Suppl. {\bf 64}, 428 (1998);
1359: % hep-ph/9707481.
1360: %%CITATION = NUPHZ,64,428;%%
1361: 
1362: \bibitem{LMR} M.~Luke, A.~Manohar and I.~Rothstein,
1363: %``Renormalization group scaling in nonrelativistic QCD,''
1364: Phys.\ Rev.\  {\bf D61}, 074025 (2000).
1365: % [hep-ph/9910209].
1366: %%CITATION = HEP-PH 9910209;%%
1367: 
1368: \bibitem{amis} A.~V.~Manohar and I.~W.~Stewart,
1369: %``Renormalization group analysis of the QCD quark potential to order  v**2,''
1370: Phys.\ Rev.\ D {\bf 62}, 014033 (2000).
1371: %[hep-ph/9912226].
1372: %%CITATION = HEP-PH 9912226;%%
1373: 
1374: 
1375: \bibitem{amis2} A.V.~Manohar and I.W.~Stewart,
1376: %``The QCD heavy-quark potential to order v**2: One loop matching  conditions,''
1377: Phys.\ Rev.\  {\bf D62}, 074015 (2000).
1378: %%CITATION = HEP-PH 0003032;%%
1379: 
1380: \bibitem{amis3}
1381: A.V.~Manohar and I.W.~Stewart,
1382: %``Running of the heavy quark production current and 1/k potential in QCD,''
1383: Phys.\ Rev.\ {\bf D63}, 54004 (2001).
1384: %hep-ph/0003107.
1385: %%CITATION = HEP-PH 0003107;%%
1386: 
1387: \bibitem{PSstat}
1388: A.~Pineda and J.~Soto,
1389: %``The renormalization group improvement of the QCD static potentials,''
1390: Phys.\ Lett.\ {\bf B495}, 323 (2000).
1391: %[hep-ph/0007197].
1392: %%CITATION = HEP-PH 0007197;%%
1393: 
1394: \bibitem{amis4} A.V.~Manohar and I.W.~Stewart,
1395: %``Logarithms of alpha in QED bound states from the renormalization group,''
1396: Phys.\ Rev.\ Lett.\  {\bf 85}, 2248 (2000).
1397: %[hep-ph/0004018].
1398: %%CITATION = HEP-PH 0004018;%%
1399: 
1400: \bibitem{hmst}
1401: A.~H.~Hoang, A.~V.~Manohar, I.~W.~Stewart and T.~Teubner,
1402: %``A renormalization group improved calculation of top quark production  near threshold,''
1403: %Phys.\ Rev.\ Lett.\  {\bf 86}, ?? (2001).
1404: hep-ph/0011254.
1405: %%CITATION = HEP-PH 0011254;%%
1406: 
1407: \bibitem{mss1} A.V.~Manohar, J. Soto, and I.W.~Stewart,
1408: %``The renormalization group for correlated scales: One-stage versus two-stage running,''
1409: Phys.\ Lett.\ {\bf B486}, 400 (2000).
1410: %%CITATION = HEP-PH 0006096;%%
1411: 
1412: \bibitem{Peter} M.~Peter, 
1413: %``The static quark-antiquark potential in QCD to three loops,''
1414: Phys.\ Rev.\ Lett.\ {\bf 78}, 602 (1997).
1415: %[hep-ph/9610209].
1416: %%CITATION = HEP-PH 9610209;%%
1417: 
1418: \bibitem{Schroeder} Y.~Schr\"oder, 
1419: %``The static potential in {QCD} to two loops,''
1420: Phys.\ Lett.\ B {\bf 447}, 321 (1999).
1421: %[hep-ph/9812205].
1422: %%CITATION = HEP-PH 9812205;%%
1423: 
1424: \bibitem{chen}
1425: Y.~Chen, Y.~Kuang and R.~J.~Oakes,
1426: %``On the spin dependent potential between heavy quark and anti-quark,''
1427: Phys.\ Rev.\  {\bf D52}, 264 (1995).
1428: % [hep-ph/9406287].
1429: %%CITATION = HEP-PH 9406287;%%
1430: 
1431: \bibitem{MW} A.~V.~Manohar and M.~B.~Wise, 
1432: %``Heavy quark physics,'' 
1433: {\it Cambridge Monographs on Particle Physics, Nuclear Physics, and Cosmology,
1434: Vol. 10}.
1435: 
1436: \bibitem{Gatheral}
1437: J.~G.~Gatheral,
1438: %``Exponentiation Of Eikonal Cross-Sections In Nonabelian Gauge Theories,''
1439: Phys.\ Lett.\ {\bf B133}, 90 (1983).
1440: %%CITATION = PHLTA,B133,90;%%
1441: 
1442: \bibitem{FrenkelTaylor}
1443: J.~Frenkel and J.~C.~Taylor,
1444: %``Nonabelian Eikonal Exponentiation,''
1445: Nucl.\ Phys.\ {\bf B246}, 231 (1984).
1446: %%CITATION = NUPHA,B246,231;%%
1447: 
1448: \bibitem{ADM}
1449: T.~Appelquist, M.~Dine and I.~Muzinich, 
1450: %``The Static Potential In Quantum Chromodynamics,''
1451: Phys.\ Lett.\ B {\bf 69}, 231 (1977);
1452: %%CITATION = PHLTA,B69,231;%%
1453: T.~Appelquist, M.~Dine and I.~Muzinich, 
1454: %``The Static Limit Of Quantum Chromodynamics,''
1455: Phys.\ Rev.\ D {\bf 17}, 2074 (1978).
1456: %%CITATION = PHRVA,D17,2074;%%
1457: 
1458: \bibitem{static1}
1459: N.~Brambilla, A.~Pineda, J.~Soto and A.~Vairo,
1460: %``The infrared behaviour of the static potential in perturbative {QCD},''
1461: Phys.\ Rev.\  {\bf D60}, 091502 (1999).
1462: %[hep-ph/9903355].
1463: %%CITATION = HEP-PH 9903355;%%
1464: 
1465: \bibitem{Gries}
1466: H.W.~Griesshammer,
1467: %``Threshold expansion and dimensionally regularized NRQCD,''
1468: Phys. Rev. {\bf D58}, 094027 (1998).
1469: % hep-ph/9712467.
1470: %CITATION = PHRVA,D58,094027;%%
1471: 
1472: \bibitem{Fischler} W.~Fischler, 
1473: %``Quark - Anti-Quark Potential In QCD,''
1474: Nucl.\ Phys.\ B {\bf 129}, 157 (1977).
1475: %%CITATION = NUPHA,B129,157;%%
1476: 
1477: 
1478: \bibitem{Beneke}
1479: M.~Beneke and V.A.~Smirnov,
1480: %``Asymptotic expansion of Feynman integrals near threshold,''
1481: Nucl. Phys. {\bf B522}, 321 (1998).
1482: %  hep-ph/9711391.
1483: %%CITATION = NUPHA,B522,321;%%
1484: 
1485: \bibitem{Yndurain}
1486: S.~Titard and F.J.~Yndurain,
1487: %``Rigorous QCD evaluation of spectrum and ground state properties of heavy q
1488: %                  anti-q systems: With a precision determination of m(b)
1489: %                  M(eta(b)),''
1490: Phys.\ Rev.\ {\bf D49}, 6007 (1994).
1491: %%CITATION = PHRVA,D49,6007;%%
1492: 
1493: \bibitem{PY}
1494: A.~Pineda and F.~J.~Yndurain,
1495: %``Calculation of quarkonium spectrum and m(b), m(c) to order  alpha(s)**4,''
1496: Phys.\ Rev.\ {\bf D 58}, 094022 (1998);
1497: %[hep-ph/9711287].
1498: %%CITATION = HEP-PH 9711287;%%
1499: A.~Pineda and F.~J.~Yndurain,
1500: %``Comment on 'Calculation of quarkonium spectrum and m(b), m(c) to order  alpha(s)**4',''
1501: Phys.\ Rev.\ {\bf D 61}, 077505 (2000).
1502: %[hep-ph/9812371].
1503: %%CITATION = HEP-PH 9812371;%%
1504: 
1505: \bibitem{renorm1}
1506: A.~H.~Hoang, M.~C.~Smith, T.~Stelzer and S.~Willenbrock,
1507: %``Quarkonia and the pole mass,''
1508: Phys.\ Rev.\ D {\bf 59}, 114014 (1999).
1509: %[hep-ph/9804227].
1510: %%CITATION = HEP-PH 9804227;%%
1511: 
1512: \bibitem{renorm2}
1513: M.~Beneke,
1514: %``A quark mass definition adequate for threshold problems,''
1515: Phys.\ Lett.\ B {\bf 434}, 115 (1998).
1516: %[hep-ph/9804241].
1517: %%CITATION = HEP-PH 9804241;%
1518: 
1519: \bibitem{a5lna}
1520: N.~Brambilla, A.~Pineda, J.~Soto and A.~Vairo,
1521: %``The heavy quarkonium spectrum at order m alpha(s)**5 ln(alpha(s)),''
1522: Phys.\ Lett.\  {\bf B470}, 215 (1999); 
1523: %[hep-ph/9910238].
1524: %%CITATION = HEP-PH 9910238;%%
1525: see also 
1526: %\bibitem{kp} 
1527: B.A.~Kniehl and A.A.~Penin, 
1528: %``Ultrasoft effects in heavy quarkonium physics,'' 
1529: Nucl.\ Phys.\ {\bf B563}, 200 (1999).
1530: %[hep-ph/9907489].
1531: 
1532: \end{references}
1533: } %end tighten (references & figure captions)
1534: 
1535: 
1536: 
1537: \end{document}
1538: 
1539: 
1540: 
1541: