hep-ph0605019/reanalysis4.tex
1: \documentclass[prd,nofootinbib,showpacs,showkeys]{revtex4}
2: %\documentclass[twocolumn,prd,nofootinbib,showpacs,showkeys]{revtex4}
3: 
4: \usepackage{hyperref}
5: \usepackage{epic,amsmath,graphicx}
6: \setlength{\topmargin}{-0.8cm}
7: \setlength{\jot}{0.3cm}
8: \def\mb#1         {\mbox{\boldmath $#1$}}
9: \def\vmb#1{\vec{\mbox{\boldmath$#1$}}}
10: 
11: \begin{document}
12: 
13: \title{New Heavy-Light Mesons $Q\bar q$}
14: 
15: %%% Authors list %%%
16: \author{Takayuki Matsuki}
17: \email[E-mail: ]{matsuki@tokyo-kasei.ac.jp}
18: \affiliation{Tokyo Kasei University,
19: 1-18-1 Kaga, Itabashi, Tokyo 173-8602, JAPAN}
20: %
21: \author{Toshiyuki Morii}
22: \email[E-mail: ]{morii@kobe-u.ac.jp}
23: \affiliation{Graduate School of Science and Technology,
24: Kobe University,\\ Nada, Kobe 657-8501, JAPAN}
25: %
26: \author{Kazutaka Sudoh}
27: \email[E-mail: ]{kazutaka.sudoh@kek.jp}
28: \affiliation{Institute of Particle and Nuclear Studies, 
29: High Energy Accelerator Research Organization, \\ 
30: 1-1 Ooho, Tsukuba, Ibaraki 305-0801, JAPAN}
31: 
32: %\date{2005/7/1} % by Sudoh
33: %\date{2006/3/7} % by Morii
34: %\date{2006/3/11} % by Matsuki
35: %\date{2006/3/30} % by Matsuki
36: %\date{2006/4/11} % by Matsuki
37: %\date{2006/4/23} % by Morii
38: %\date{2006/4/28} % by Matsuki
39: %\date{2006/8/23} % by Matsuki
40: %\date{2006/9/11} % by Matsuki
41: %\date{2006/9/18} % by Matsuki
42: %\date{2006/9/30} % by Morii
43: %\date{2006/10/2} % by Matsuki
44: %\date{2006/10/3} % by Sudoh
45: %\date{2006/11/27} % by Matsuki
46: %\date{2007/1/2} % by Matsuki
47: \date{\today} % by Matsuki
48: 
49: \begin{abstract}
50: We succeed in reproducing the $\ell=1$ $B$ mesons, $B_1(5720)$, $B_2^*(5745)$, and
51: $B_{s2}^*(5839)$ that are recently reported by D0 and CDF, by using our
52: semirelativistic quark potential model, which also succeeded in predicting the mass
53: spectra of the narrow $D_{sJ}$ as well as broad $D_0^*(0^+)$ and $D_1'(1^+)$ particles a
54: couple of years ago.
55: 
56: Mass of higher excited states ($\ell=1, 2$) of $B$ and $B_s$ mesons, which are not yet
57: observed, is also predicted at the first order in $p/m_b$ with internal quark momentum
58: $p$ and the $b$ quark mass $m_b$. We find the corresponding
59: $B_{sJ}$ are below $BK/B^*K$ threshold and should have narrow decay widths contrary to
60: most other predictions. Also already established states ($\ell=0$ and $\ell=1$) of
61: $D$, $D_s$, $B$, and $B_s$ heavy mesons are simultaneously reproduced in good
62: agreement with experimental data within one percent of accuracy. To calculate these
63: $D/D_s$ and $B/B_s$ heavy mesons we use different values of strong coupling
64: $\alpha_s$.
65: \end{abstract}
66: %\medskip
67: \preprint{}
68: \pacs{12.39.Hg, 12.39.Pn, 12.40.Yx, 14.40.Lb, 14.40.Nd}
69: \keywords{potential model; spectroscopy; heavy mesons}
70: \maketitle
71: 
72: %--------------------------------------------------------------
73: \section{Introduction}
74: \label{intro}
75: Recently discovered were narrow meson states, $D_{s0}(2317)$ by BaBar
76: \cite{BaBar03} and $D_{s1}(2460)$ by CLEO \cite{CLEO03}, both of which were
77: confirmed by Belle \cite{Belle03}. These are identified as $j^P=0^+$ and $1^+$
78: of $c\bar s$ excited ($\ell=1$) bound states, respectively, and have caused
79: a revival of study on heavy meson spectroscopy. Subsequently, another set of
80: broad heavy mesons, $D_{0}^{*}(2308)$ and $D_{1}'(2427)$ were discovered by
81: Belle collaboration \cite{Belle04}. Those are identified as $c\bar q$ ($q=u/d$)
82: excited ($\ell=1$) bound states and have the same quantum numbers $j^P=0^+$ and
83: $1^+$ as $D_{sJ}$, respectively. 
84: The decay widths of these excited $D_{sJ}$ mesons are narrow since the masses
85: are below $DK/D^{*}K$ threshold and hence the dominant decay modes violate the
86: isospin invariance, whereas those excited $D$ mesons are broad because of no
87: such restriction as in $D_{sJ}$ cases. More recent experiments reported
88: by CDF and D0 \cite{CDF_D006} found narrow $B$ and $B_s$ states of $\ell=1$,
89: $B_1(5720)$, $B_2^*(5745)$, and $B_{s2}^*(5839)$. These are narrow because these
90: decay through the D-waves.
91: 
92: There are some discussions \cite{Barnes03} that a quark potential model is not
93: appropriate to describe these new states. In fact, mass spectra and decay widths
94: of these states could not be reproduced by relativised/relativistic quark potential
95: models \cite{Godfrey85, Godfrey91, Ebert98}.
96: In \cite{Godfrey85, Godfrey91}, for instance,
97: kinetic terms of quarks are taken as positive definite $\sqrt{p^2+m^2}$ and
98: spin-independent linear-rising confining and short-range Coulomb potentials are
99: taken into account together with spin-dependent interaction terms symmetric in $Q$
100: and $\bar q$.
101: In \cite{Ebert98}, even though degeneracy among members of a spin doublet in the
102: limit of heavy quark symmetry is taken into account, they obtained masses more than
103: a hundred MeV larger than the observed $D_{sJ}$ mesons. The nonrelativistic quark
104: potential model applied both to mesons and baryons has been so successful in
105: reproducing low lying hadrons (see the review \cite{Mukherjee93}) that people were
106: puzzled why it did not work for $D_{sJ}$ mesons (atom like mesons).
107: 
108: To understand these heavy meson states, a number of interesting ideas have been
109: proposed so far. One is an effective Lagrangian approach.
110: The authors of \cite{Bardeen03, Bardeen94} have proposed the modified
111: Goldberger-Treiman relation in an effective Lagrangian,
112: by which the mass difference between two spin doublets
113: $(0^-,1^-)$ and $(0^+,1^+)$ are reproduced.
114: This idea is theoretically successful in explaining $D_s$ states but in fail for
115: $D$ states though experimentally the relation holds.
116: Other approaches proposed to understand these new states are the QCD sum rule
117: \cite{Hayashigaki04}, implications of a two-meson molecule
118: \cite{Barnes03, Szczepaniak03}, four quark system
119: \cite{Browder04, Terasaki05, Dmitrasinovic05}, lattice calculation
120: \cite{Bali03} and so on.
121: The Bethe-Salpeter (BS) equation can be an alternative way
122: to take relativistic corrections into account.\cite{Habe87, Gara89, Pierro01}
123: A coupled channel method \cite{Beveren06} is another candidate to
124: interpret the new heavy mesons though underlying physics is unclear.
125: These approaches are not yet completely established even though they may be
126: interesting themselves. Hence, the problem still remains unsolved by these
127: approaches and is challenging. See the detailed review \cite{Swanson06} on
128: heavy mesons.
129: 
130: We believe that a quark potential model still powerfully survives if we treat
131: a bound state equation appropriately, being able to predict not only mass
132: differences but also absolute values of hadron masses. In fact, long time
133: ago before the discovery of $D_{s0}(2317)$ by BaBar, two of us (T.M. and T.M.)
134: \cite{Matsuki97} proposed a new bound state equation for atom-like mesons, i.e.,
135: heavy mesons composed of a heavy quark $Q$ and a light antiquark $\bar q$, in
136: which quarks are treated as four-spinor Dirac particles, and Hamiltonian, wave
137: functions, and eigenvalues are all expanded in $p/m_Q$ so that a light quark is
138: treated as relativistically as possible and nonrelativistic reduction is made for
139: a heavy quark. The heavy quark symmetry is taken
140: into account consistently within a potential model. It is remarkable that the
141: model could predict the levels of $D_{s0}(2317)$, $D_{s1}(2460)$,
142: $D_{0}^{*}(2308)$, and $D_{1}'(2427)$ quite successfully even using rather
143: obscure parameter values due to a small number of data (ground states and only
144: a few excited states) at the time of publication \cite{Matsuki97}.
145: 
146: Recent experiments on $B$ and $B_s$ (discovery of states of $\ell=1$,
147: $B_1(5720)$, $B_2^*(5745)$, and $B_{s2}^*(5839)$) by CDF and D0 \cite{CDF_D006}
148: serves us a good testing ground whether our formulation, semirelativistic quark
149: potential model, or other formulation also works or not. One of the main purposes
150: of this paper is to show that our model can also reproduce these $B/B_s$
151: states within one percent of accuracy compared with the experiments.
152: Another aim of this paper is to reproduce and predict the whole spectrum of $D/D_s$
153: and $B/B_s$ heavy mesons, including higher states, $^3D_1$ and $^3D_2/^1D_2$,
154: which we failed to have reasonal values in \cite{Matsuki97}.
155: 
156: In the previous paper \cite{Matsuki97}, because of the small number of input
157: data in those days, we searched for {\sl a local minimum} which fits with at least
158: the lowest ground states and found that the $\chi^2$ analysis led to a
159: {\em negative} optimal value of $b$ in a scalar potential $S(r)=b+r/a^2$.
160: However, in this paper we have obtained a {\em positive} $b$ owing to enough
161: experimental input data this time using the Minuit.
162: %
163: One difference caused by this sign change
164: appears in light quark mass, $m_q=m_u=m_d$, whose effective mass in the lowest
165: order Hamiltonian becomes $m_q+b$. This effective mass becomes negative,
166: $\sim -38$ MeV for $b<0$ in \cite{Matsuki97} and positive, $\sim 86$ MeV in
167: this paper. Heavy quark masses, $m_c$ and $m_b$, are also affected by this sign
168: change, i.e., we have obtained a few hundred MeV smaller values in this paper
169: than those in \cite{Matsuki97}.
170: Other significant influence of sign change is on the magnitude
171: of the corrections of lower components of a heavy quark, which are comparable
172: to upper component contributions in \cite{Matsuki97} ($b<0$), while
173: in this paper the lower component contributions are comparable to the second
174: order contributions in $p/m_Q$ as we will see later.
175: These are the main difference from \cite{Matsuki97} in the way of data analysis
176: and otherwise we are using the same approach as before.
177: %, but
178: %$changing the sign of $b$ has a big impact on determining optimal values of
179: %parameters to obtain solutions to the Schr\"odinger equation.
180: In addition to a good result for mass levels of $\ell =0$ and $\ell =1$ states of all
181: $D$, $D_s$, $B$ and $B_s$ mesons, a positive $b$ causes not only obtaining plausible
182: values for $k=2~(\ell=2)$ states (corresponding to $^3D_1,\; ^3D_2/^1D_2$)
183: but also obtaining relatively small
184: degenerate mass $M_0$ in a spin multiplet so that the corrections become
185: larger in this paper compared with \cite{Matsuki97}.
186: That is, the sign of $b$ is important to obtain the whole spectra
187: of heavy mesons even though the lower lying spectra are correctly predicted in
188: both cases. Another important and different point from the
189: former paper is that we treat even the light quark masses ($m_u$, $m_d$, and
190: $m_s$) as free parameters in this paper, while they were given as input in
191: \cite{Matsuki97}.
192: 
193: In this paper, using the refined parameters obtained from recent
194: new experimental data, we elaborate computation of the mass spectra of heavy
195: mesons $D$, $D_s$, $B$, and $B_s$ mesons, including $D_{s0}(2317)$,
196: $D_{s1}(2460)$, $D_{0}^{*}(2308)$, $D_{1}'(2427)$, $B_1(5720)$, $B_2^*(5745)$,
197: and $B_{s2}^*(5839)$.  We find that in this new analysis all of the masses of
198: not only above mentioned new excited ($\ell=1$) states $j^P=0^+$, $1^+$, and
199: $2^+$ but also those of other excited ($\ell=1$) states $j^P=1^+$ and $2^+$,
200: and ground
201: ($\ell=0$) states $j^P=0^-$ and $1^-$ of $D$, $D_s$, $B$, and $B_s$ mesons can
202: be reproduced well within one percent of accuracy. Furthermore, we {\em predict}
203: $B_{s0}$ and $B_{s1}$ are below $BK/B^*K$ threshold, which means one can find them
204: as narrow states and is the same results as in \cite{Matsuki05} although we have
205: about 100 MeV smaller values for $0^+$ and $1^+$ states. This is contrary to most
206: other quark potential model predictions as in $D_{sJ}$ particles.
207: 
208: The main difference between our treatment of the quark potential model and others
209: is that we have systematically expanded interaction terms in $p/m_Q$ and that we
210: have taken into account lower components of a heavy quark in a bound state.
211: It turns out that {\sl systematically expanded interaction terms are quite different
212: from those of nonrelativistic and/or relativised models} except for \cite{Pierro01}.
213: In \cite{Pierro01} they used the BS equation and projected heavy quark sector on
214: positive states, otherwise their final formulation is very close to ours.
215: Another large difference is the magnitude of a light quark mass, i.e., other models
216: including \cite{Pierro01} use a constituent quark mass while we got much smaller
217: values, e.g., $m_u, m_d \sim 10$ MeV and $m_s \sim 100$ MeV.
218: %These are the main reasons why other conventional quark models cannot obtain a
219: %right solution, especially for $^{2s+1}\ell_j=^3P_0$ and $^3P_1$ states.
220: 
221: The paper is organized as follows. 
222: In the next section, we briefly review our model, where some necessary formulas
223: are presented for readers' understanding and a discussion on a convenient
224: quantum number $k$ is given. In Sec. \ref{result}, numerical results are
225: presented in Tables \ref{Dmeson} and \ref{Dsmeson} for the mass levels of $D$
226: and $D_s$ mesons including their excited states together with the most optimal
227: values of parameters in Table \ref{parameter}, where we find that calculated
228: masses are in good agreement with all existing experimental data. Those of
229: $B$ and $B_s$ mesons are also given in the same section, Tables \ref{Bmeson} and
230: \ref{Bsmeson}. To show smallness of the second order corrections, we
231: calculate $D$ meson mass spectra with the second order corrections and give them
232: in Table \ref{Dmeson2nd}. Section \ref{summary} is devoted to conclusion and
233: discussion, including the reason why our semirelativistic quark potential
234: model approach gives good results compared with others.
235: 
236: %--------------------------------------------------------------
237: \section{Brief Review of our Semirelativistic Model}
238: \label{formalism}
239: Let us start with an effective two-body Hamiltonian for a system of a heavy
240: quark $Q$ with mass $m_Q$ and a light antiquark $\bar q$ with mass $m_{\bar q}$
241: \cite{Fermi49, Bethe55},\footnote{Miyazawa and Tanaka \cite{Miyazawa92} pointed
242: out that a relativistic bound state equation having this effective Hamiltonian
243: can be derived with a boundary condition in relative time that is different from
244: that of the Bethe--Salpeter equation.}
245: %
246: \begin{eqnarray}
247: H&=&(\vec{\alpha}_{\bar q}\cdot \vec{p}_{\bar q} 
248: + \beta_{\bar q} m_{\bar q}) 
249: + (\vec{\alpha}_Q\cdot \vec{p}_Q 
250: + \beta_Q m_Q) 
251: + \beta_{\bar q} \beta_Q S(r) \nonumber \\
252: &&+ \left\{ 1-\frac{1}{2}\left[\vec{\alpha}_{\bar q}\cdot
253: \vec{\alpha}_Q + 
254: (\vec{\alpha}_{\bar q}\cdot \vec{n})
255: (\vec{\alpha}_Q\cdot \vec{n})
256: \right]\right\}V(r), 
257: \label{total_H}
258: \end{eqnarray}
259: %
260: where $S$ and $V$ are a scalar confining potential and a vector one-gluon
261: exchange Coulomb potential with transverse interaction and $\vec{n}=\vec{r}/r$
262: is a unit vector along the relative coordinate between $Q$ and $\bar q$. $S$
263: and $V$ are given explicitly as
264: %
265: \begin{equation}
266: S(r)=\frac{r}{a^2}+b, ~~~V(r)=-\frac{4}{3}\frac{\alpha_s}{r},
267: \label{potential}
268: \end{equation}
269: %
270: where $a$ and $1/b$ are parameters with length dimension and $\alpha_s$ is a
271: strong coupling constant. Since a heavy quark $Q$ is sufficiently heavier than
272: a light antiquark $\bar q$ which is orbiting $Q$, it is reasonable to apply the
273: Foldy--Wouthuysen--Tani (FWT) transformation \cite{Foldy50} to the heavy quark
274: related operators in $H$ of Eq.(\ref{total_H}). One should notice that the FWT
275: is introduced so that free kinetic terms of a heavy quark become $\beta_Q m_Q$
276: and one can expand the interaction terms in terms of $p/m_Q$ systematically,
277: \footnote{There is another way of expanding the Hamiltonian. See \cite{MS97}
278: where the Bloch method \cite{Bloch} is used to expand the effective Hamiltonian
279: in $p/m_Q$, which is reproduced in Appendix B.} which is equivalent to a
280: nonrelativistic reduction of the
281: original model. By doing this way, we can get an eigenvalue equation for an atom-
282: like bound state $Q\bar q$ \cite{Matsuki97} \footnote{There is a pioneering
283: work of this approach
284: %The similar effective bound
285: %state equation for an atom-like meson was developed in
286: \cite{Morishita88}, where
287: expanded terms of the Hamiltonian in $p/m_Q$ are treated differently from ours.}
288: %
289: \begin{equation}
290: (H_{\rm FWT}-m_{Q}) \otimes\psi_{\rm FWT} =\tilde{E}\psi_{\rm FWT},
291: \label{eigen_FWT}
292: \end{equation}
293: %
294: with
295: %
296: \begin{equation}
297: H_{\rm FWT}-m_{Q}=H_{-1}+H_{0}+H_{1}+H_{2}+\cdots , 
298: \label{H_mQ}
299: \end{equation}
300: %
301: where $\tilde{E}=E-m_Q$ ($E$ being the bound state mass of $Q\bar q$) is the
302: binding energy and a notation $\otimes$ denotes that gamma matrices of a light
303: antiquark is multiplied from left with the wave function while those of a heavy
304: quark from right.  $H_i$ in Eq.(\ref{H_mQ}) denotes the $i$-th order expanded
305: Hamiltonian in $p/m_Q$ and its explicit expressions are presented in
306: \cite{Matsuki97} and also in the Appendix A. Then, Eq.(\ref{eigen_FWT}) can be
307: numerically solved in order by order in $p/m_Q$ according to the standard
308: perturbation method by expanding the eigenvalue and the wave function as
309: %
310: \begin{eqnarray}
311:   \tilde{E} &=& E-m_{Q}=E_{0}^{a}+E_{1}^{a}+E_{2}^{a}+\cdots , 
312:   \label{E_mQ} \\
313: %
314:   \psi_{\rm FWT} &=& \psi_{0}^{a}+\psi_{1}^{a}+\psi_{2}^{a}+\cdots , 
315:   \label{psi_mQ}
316: \end{eqnarray}
317: %
318: where a superscript $a$ represents a set of quantum numbers, a total angular
319: momentum $j$, its $z$ component $m$ and a quantum number $k$ of the spinor
320: operator $K$ defined below in Eq.(\ref{k_quantum}) and a subscript $i$ of
321: $E_i^{a}$ and $\psi_i^{a}$ stands for the $i$-th order in $p/m_Q$.
322: 
323: In practice, it is reasonable to first solve the $0$-th order non-trivial
324: equation by variation to get the $0$-th order eigenvalues and their wave
325: functions and then to estimate the corrections perturbatively in order
326: by order in $p/m_Q$ by evaluating the matrix elements for those terms.
327: The 0-th order eigenvalue $E_0^{a}$ and the wave function $\psi_0^{a}$ for the
328: $Q\bar q$ bound state is obtained by solving the $0$-th order eigenvalue
329: equation,
330: %
331: \begin{equation}
332:   H_0 \otimes\psi _0^{a} = E_0^{a}\psi _0^{a}, \qquad
333: %
334:   H_0 =
335:   \vec{\alpha}_{\bar q}\cdot\vec{p}_{\bar q}+\beta_{\bar q}
336:   \left(m_{\bar q}+S(r)\right)+V(r).
337: \label{0th}
338: \end{equation}
339: %
340: The wave function which has two spinor indexes and is expressed as $4\times 4$
341: matrix form is explicitly described by
342: %
343: \begin{eqnarray}
344:   \psi _0^{a} &=& \left(0~~ \Psi _{j\,m}^k(\vec r) \right),
345:   \label{0thsols1} \\
346:   \Psi _{j\,m}^k(\vec r) &=& \frac{1}{r}\left( 
347: \begin{array}{c}
348: u_k(r)\;y_{j\,m}^k \\ 
349: iv_k(r)\;y_{j\,m}^{-k} 
350: \end{array}
351: \right),
352:   \label{0thsols2}
353: \end{eqnarray}
354: %
355: where $y_{j\,m}^k$ being of $2\times 2$ form are the angular part of the wave
356: functions and $u_k(r)$ and $v_k(r)$ are the radial parts. The total angular
357: momentum of a heavy meson $\vec J$ is the sum of the total angular momentum of
358: the light quark degrees of freedom $\vec S_\ell$ and the heavy quark spin
359: ${1\over 2}\vec \Sigma_Q$:
360: %
361: \begin{equation}
362:   {\vec J} = {\vec S_\ell} +{1\over 2}\vec \Sigma_Q\qquad {\rm with} \quad
363:   {\vec S_\ell}=\vec L + {1\over 2}\vec \Sigma_{\bar q},
364: \end{equation}
365: %
366: where ${1\over 2}\vec\Sigma_{\bar q}\ (={1\over 2}
367: \vec\sigma_{\bar q}\;1_{2\times 2})$ and $\vec L$ are the $4\times 4$ spin and
368: the orbital angular momentum of a light antiquark, respectively. Furthermore,
369: $k$ is the quantum number of the following spinor operator $K$ \cite{Matsuki97}
370: %
371: \begin{equation}
372:   K = -\beta_{\bar q} \left( \vec \Sigma_{\bar q} \cdot \vec L + 1 \right),
373:   \qquad
374:   K\, \Psi_{j\,m}^k = k\, \Psi_{j\,m}^k.
375:   \label{k_quantum}
376: \end{equation}
377: %
378: Note that in our approach $K$ is introduced for a two-body bound system of
379: $Q\bar q$, though the same form of the operator $K$ can be defined in the case
380: of a single Dirac particle in a central potential \cite{JJ}.  One can easily
381: show \cite{Matsuki04} that the operator $K^2$ is equivalent to
382: ${\vec S_\ell}^2$ and it holds $k = \pm \left( s_\ell + \frac{1}{2} \right)$
383: or $s_\ell = \left| k \right| - \frac{1}{2}$. $k$ is also related to $j$ as
384: %
385: \begin{equation}
386: j=|k| ~~{\rm or}~~ |k|-1 ~~~(k\neq 0) . 
387: \end{equation}
388: %
389: 
390: In order to diagonalize the 0-th order Hamiltonian $H_0$ of Eq.(\ref{0th}) in
391: the $k$ representation, it is convenient to introduce a spinor representation
392: of $y_{j\,m}^k$ which is given by the unitary transformation
393: (Eq.(\ref{Unitary}) below) of a set of the spherical harmonics $Y_j^m$ and the
394: vector spherical harmonics defined by 
395: %
396: \begin{eqnarray}
397:   \vec Y_{j\,m}^{(\rm L)} = -\vec n\,Y_j^m, \quad
398:   \vec Y_{j\,m}^{(\rm E)} = {r \over {\sqrt {j(j+1)}}}\vec \nabla Y_j^m,
399:   \quad
400:   \vec Y_{j\,m}^{(\rm M)} = -i\vec n\times \vec Y_{j\,m}^{(\rm E)},
401: \end{eqnarray}
402: %
403: where $\vec n=\vec r/r$. These vector spherical harmonics are nothing but a set
404: of eigenfunctions for a spin-1 particle. 
405: $\vec Y_{j\,m}^{(\rm A)}$ (A=L, M, E) are eigenfunctions of ${\vec J}^2$ and
406: $J_z$, having the eigenvalues $j(j+1)$ and $m$. The parity of these functions
407: is assigned as $(-1)^{j+1}$, $(-1)^j$, $(-1)^{j+1}$ for A=L, M, and E,
408: respectively, since $Y_j^m$ has a parity $(-1)^j$. The unitary transformation
409: is given by 
410: %
411: \begin{equation}
412: \left( 
413: \begin{array}{c}
414: y_{j\,m}^{-(j+1)} \\
415: y_{j\,m}^j
416: \end{array}
417: \right)=U\left( 
418: \begin{array}{c}
419: Y_j^m \\ 
420: \vec \sigma \cdot \vec Y_{j\,m}^{(\rm M)}
421: \end{array}
422: \right), \qquad
423: \left( 
424: \begin{array}{c}
425: y_{j\,m}^{j+1} \\
426: y_{j\,m}^{-j}
427: \end{array}
428: \right)=U\left( 
429: \begin{array}{c}
430: \vec \sigma \cdot \vec Y_{j\,m}^{(\rm L)} \\
431: \vec \sigma \cdot \vec Y_{j\,m}^{(\rm E)}
432: \end{array}
433: \right), 
434: \label{Unitary}
435: \end{equation}
436: %
437: where the matrix $U$ is defined by
438: %
439: \begin{equation}
440: U={1 \over {\sqrt {2j+1}}}\left( 
441: \begin{array}{cc}
442: {\sqrt {j+1}} & {\sqrt j} \\
443: -{\sqrt j} & {\sqrt {j+1}}
444: \end{array}
445: \right). 
446: \label{eq:app:u}
447: \end{equation}
448: %
449: Now, substituting Eqs.(\ref{0thsols1}) and (\ref{0thsols2}) into
450: Eq.(\ref{0th}), one can eliminate the angular part of the wave function,
451: $y_{j\,m}^k$, from the eigenvalue equation and obtain the radial equation as
452: follows,
453: %
454: \begin{equation}
455: \left( 
456: \begin{array}{cc}
457: m_q+S+V & -\partial _r+{k \over r} \\
458: \partial _r+{k \over r} & -m_q-S+V
459: \end{array}
460: \right)\left(
461: \begin{array}{c}
462: u_k(r) \\
463: v_k(r)
464: \end{array}
465: \right)
466: = E^k_0\left(
467: \begin{array}{c}
468: u_k(r) \\
469: v_k(r)
470: \end{array}
471: \right), 
472: \label{radial}
473: \end{equation}
474: %
475: which depends on the quantum number $k$ alone.  This is just the same form as a
476: one-body Dirac equation in a central potential. Since $K$ commutes with $H_0$
477: and the states $\Psi _{j\,m}^k$ have the same energy $E_0^k$ for different
478: values of $j$, these states having a total angular momentum $j=|k|-1$ and $|k|$,
479: a spin doublet, are degenerate with the same value of $k$.
480: 
481: The parity of the eigenfunction $\psi_0^{a}$ is defined by the upper
482: (``large'') component of Eq.(\ref{0thsols2}) and hence taking into account the
483: intrinsic parity of quark and antiquark, the parity of the meson $Q\bar q$ is
484: given by \cite{Matsuki04}
485: %
486: \begin{equation}
487: P=\frac{k}{|k|}(-1)^{|k|+1},
488: \end{equation}
489: %
490: which is equal to the parity $\pi_{\ell}$ of the light degrees of freedom in the
491: heavy quark effective theory (HQET) as shown in Table I. It is remarkable that
492: both the total angular momentum $j$ and the parity $P$ of a heavy meson can be
493: simultaneously determined by $k$ alone. Or the states can be completely
494: classified in terms of $k$ and $j$ as shown in Table \ref{table}.
495: 
496: \begin{table*}[t]
497: \caption{States classified by various quantum numbers}
498: \label{table}
499: \begin{tabular*}{13cm}{c|@{\extracolsep{\fill}}cccccccc}
500: \hline
501: \hline
502:   \makebox[1.7cm]{$j^P$}   &   $0^-$   &   $1^-$   &   $0^+$   &   
503:   $1^+$   &   $1^+$   &   $2^+$   &   $1^-$   &   $2^-$   \\
504: %%%%%%%%%%%%%%%%%%%%%%%%%%%
505:   $k$   &   -1   &   -1   &   1   &   
506:    1    &   -2   &   -2   &   2   &   2   \\
507: %%%%%%%%%%%%%%%%%%%%%%%%%%%
508:   $s_\ell^{\pi_\ell}$ & ${1\over2}^-$ & ${1\over2}^-$ & ${1\over2}^+$ & 
509:   ${1\over2}^+$ & ${3\over2}^+$ & ${3\over2}^+$ & ${3\over2}^-$ & ${3\over2}^-$
510:   \\
511: %%%%%%%%%%%%%%%%%%%%%%%%%%%
512: %%%%%%%%%%%%%%%%%%%%%%%%%%%
513:   $^{2s+1}l_j$ & $^1S_0$ & $^3S_1$  & $^3P_0$  & $^3P_1$, $^1P_1$ &
514:    $^1P_1$, $^3P_1$  & $^3P_2$  & $^3D_1$  & $^3D_2$, $^1D_2$
515:   \\ \hline 
516: %%%%%%%%%%%%%%%%%%%%%%%%%%%
517:   $\Psi_j^k$ & $\Psi_0^{-1}$ & $\Psi_1^{-1}$ & $\Psi_0^1$ & $\Psi_1^1$ &
518:   $\Psi_1^{-2}$ & $\Psi_2^{-2}$ & $\Psi_1^2$ & $\Psi_2^2$ \\
519: \hline
520: \hline
521: \end{tabular*}
522: \end{table*} 
523: 
524: Classification by the quantum number $k$ is convenient to discuss the
525: relation between the symmetry of the system and the structure of mass levels.
526: Now, let us look at how masses of $Q\bar q$ mesons are generated in our model.
527: First, when the light quark mass $m_{\bar q}$ and the scalar potential $S(r)$
528: are neglected, the $0$-th order Hamiltonian $H_0$ has both a chiral symmetry
529: and a heavy quark symmetry and the eigenvalues with the same value of $|k|$ are
530: degenerate in this limit, where all masses of $0^-$, $1^-$, $0^+$, and $1^+$
531: states with $|k|=1$ are degenerate, as
532: $m(0^-)=m(1^-)=m(0^+)=m(1^+)=m_Q$.\cite{Matsuki05} Then, by turning on the light quark mass and the scalar
533: potential, $m_{\bar q}\neq 0$ and $S(r)\neq 0$, the chiral symmetry is broken
534: and the degeneracy between two spin doublets $(0^-, 1^-)$ and $(0^+, 1^+)$ is
535: resolved. At this stage, members in spin multiplets are still degenerate due to
536: the same quantum number $k$. Finally, by adding $p/m_Q$ corrections, the heavy
537: quark symmetry is broken and the hyperfine splitting occurs as shown in
538: Fig. \ref{mass_level}.
539: 
540: \begin{figure*}[t]
541: \begin{center}
542: \begin{picture}(370,90)
543: \setlength{\unitlength}{0.4mm}
544:  \thicklines
545:   \put(5,40){\line(1,0){60}}
546:   \put(130,57.5){\line(1,0){60}} \put(130,22.5){\line(1,0){60}}
547:   \put(245,65){\line(1,0){60}} \put(245,50){\line(1,0){60}}
548:   \put(245,30){\line(1,0){60}} \put(245,15){\line(1,0){60}}
549:  \thinlines
550:   \dottedline{3}(65,40)(130,57.5) \dottedline{3}(65,40)(130,22.5)
551:   \dottedline{3}(190,57.5)(245,65) \dottedline{3}(190,57.5)(245,50)
552:   \dottedline{3}(190,22.5)(245,30) \dottedline{3}(190,22.5)(245,15)
553:   \put(33,46){$|k|=1$}
554:   \put(-15,15){$\left(\begin{array}{c}
555:               m_q\to0,\, S\to0 \\  {\rm no}~ p/m_Q~ {\rm corrections} \\
556:               {\rm Chiral~Symmetric}
557:              \end{array}\right)$}
558:   \put(148,62.5){$k=+1$} \put(148,27.5){$k=-1$}
559:   \put(310,62.5){$1^+$} \put(310,47.5){$0^+$}
560:   \put(310,27.5){$1^-$} \put(310,12.5){$0^-$}
561: %  \put(132,7.5){$\left(\begin{array}{c}
562:   \put(100,5){$\left(\begin{array}{c}
563:   			m_q\neq 0,\, S\neq0 \\
564:   			{\rm Heavy~Quark~Symmetric}
565:              \end{array}\right)$}
566:   \put(242,0){($p/m_Q$ corrections)}
567: \end{picture}
568: \caption{Procedure how the degeneracy is resolved in our model.}
569: \label{mass_level}
570: \end{center}
571: \end{figure*}
572: 
573: %--------------------------------------------------------------
574: \section{Numerical Calculation}
575: \label{result}
576: To obtain masses of $Q\bar q$ states, we first solve the eigenvalue
577: equation Eq.(\ref{radial}) by variation to get the lowest eigenvalues and the
578: wave functions. By taking account of the asymptotic behavior at both
579: $r\rightarrow 0$ and $r\rightarrow \infty$, the trial functions for $u_k(r)$
580: and $v_k(r)$ are given by
581: %
582: \begin{equation}
583: u_{k}(r), ~v_{k}(r)\sim w_{k}(r)\left(\frac{r}{a}\right)^{\lambda}
584: \exp \left[-(m_{q}+b)r-\frac{1}{2}\left(\frac{r}{a}\right)^{2} \right] , 
585: \end{equation}
586: %
587: where
588: %
589: \begin{equation}
590: \lambda=\sqrt{k^{2} - \left(\frac{4\alpha_s}{3}\right)^{2}} , 
591: \end{equation}
592: %
593: and $w_{k}(r)$ is a finite series of a polynomial in $r$ 
594: %
595: \begin{equation}
596: w_{k}(r)=\sum_{i=0}^{N-1}a_{i}^{k}\left(\frac{r}{a}\right)^{i} , 
597: \end{equation}
598: %
599: which takes different coefficients for $u_{k}(r)$ and $v_{k}(r)$.
600: After getting eigenvalues and wave functions for the $0$-th order solutions, we
601: use those wave functions to calculate $p/m_Q$ corrections in order by order.
602: Corrections are estimated by evaluating matrix elements of corresponding terms
603: in the higher order Hamiltonian. Details of the prescription for carrying out
604: numerical calculations and getting reliable solutions are described in
605: \cite{Matsuki97}.
606: 
607: In the work of \cite{Matsuki97}, by fitting the smallest eigenvalues of the
608: Hamiltonian with masses of $D(1867)$ and $D^*(2008)$ for $c\bar q$ ($q=u, d$),
609: $D_s(1969)$ and $D_s^*(2112)$ for $c\bar s$, and $B(5279)$ and $B(5325)$ for
610: $b\bar q$, a strong coupling constant $\alpha_s$ and other potential parameters,
611: $a$ and $b$, were determined. At that time following the paper \cite{Morishita88},
612: we started the search for a set of parameters with $b<0$. Using those parameters
613: obtained this way, other mass levels were calculated and compared with the
614: experimental data for $D_{(s)}/B_{(s)}$ mesons. Light quark ($u$ and $d$) mass
615: was taken to be 10MeV as an input, being close to a value of current quark mass.
616: 
617: Here in this paper since we have now a plenty of the observed data, we follow a
618: different way of analysis.
619: %
620: \def\labelenumi{\theenumi)}
621: \begin{enumerate}
622: \item~First, we determine the six parameters, i.e., a
623: strong coupling constant $\alpha_s^c$, potential parameters $a$ and $b$, quark
624: masses $m_{u,d}$\footnote{In the numerical analysis, we set $m_{u}=m_{d}$ in
625: order to reduce the number of parameters.}, $m_s$, and $m_c$ by fitting
626: calculations to the observed six $D$ meson masses, i.e., those of two $\ell=0$
627: $(0^-, 1^-)$ states and four $\ell=1$ $(0^+, 1^+, 1^+, 2^+)$ states, and
628: similarly six $D_s$ meson masses, using the Minuit $\chi^2$ analysis.
629: The results are given in Table \ref{parameter}.
630: As written in the Introduction, we have found that the optimal value for $b$
631: becomes positive contrary to the former paper \cite{Matsuki97}, which affects
632: all other parameter values as well as the calculated values, $M_{\rm calc}$,
633: $M_0$, $p_i$, $n_i$, and $c_i$ whose meanings are explained below.
634: %
635: \item~Then, using the optimal parameters obtained this way except for a strong
636: coupling,
637: the bottom quark mass $m_b$ and a strong coupling $\alpha_s^b$ which is assumed
638: to be different from $\alpha_s^c$ for $D/D_s$ mesons are
639: determined by fitting the four observed $B$ meson masses and two $B_s$ meson masses.
640: The most optimal values of parameters are presented in Table \ref{parameter} at the
641: first order calculation in $p/m_Q$.
642: \end{enumerate}
643: %
644: 
645: \begin{table*}[t!]
646: \caption{Most optimal values of parameters.}
647: \label{parameter}
648: \begin{tabular}{lcccccccc}
649: \hline
650: \hline
651: Parameters
652: & ~~$\alpha_s^c$ & ~~$\alpha_s^b$ & ~~$a$ (GeV$^{-1}$) & ~~$b$ (GeV) 
653: & ~~$m_{u, d}$ (GeV) & ~~$m_s$ (GeV) & ~~$m_c$ (GeV) & ~~$m_b$ (GeV) \\
654: %%%%%%%%%%%%%%%%%%%%%%%%%%%
655: %First order 
656: & ~~0.261 & ~~0.393 & ~~1.939 & ~~0.0749 
657: & ~~0.0112 & ~~0.0929 & ~~1.032 & ~~4.639 \\
658: %%%%%%%%%%%%%%%%%%%%%%%%%%%
659: \hline
660: \hline
661: \end{tabular}
662: \end{table*}
663: %
664: 
665: The masses with the same value of $k$ degenerate in the $0$-th order for
666: members of each spin doublet, are labeled as $M_0$. $c_i$ denotes the $i$-th
667: order correction in $p/m_Q$ expansion (in this paper $n=1, 2$) and thus, the
668: calculated heavy meson mass $M_{\rm calc}$ is given by the sum of $M_0$ and
669: corrections up to the $n$-th order,
670: %
671: \begin{equation}
672: M_{\rm calc} = M_0 + \sum_{i=1}^{n}\left(p_i+n_i\right) ,
673: \end{equation}
674: %
675: where $p_i$ includes the contributions stemming from only upper
676: components of a heavy quark, $n_i$ from lower, and $c_i=p_i+n_i$ in each
677: order. Here one should notice that $M_0=m_Q+E_0$.
678: The degeneracy in $k$ or for the states in the same spin doublet is resolved by
679: taking account of higher order corrections in $p/m_Q$ expansion. See Fig. \ref{mass_level}.
680: 
681: In our formalism, each state is uniquely classified by two quantum numbers,
682: $k$ and $j$, and
683: also approximately classified by the upper component of the wave function of
684: Eq.(\ref{0thsols2}) in terms of the conventional notation $^{2s+1}\ell_j$.
685: In terms of this notation, the $j^P =1^+$ state is a mixed state of $^3P_1$ and
686: $^1P_1$, while the $j^P =2^-$ state is a mixed state of $^3D_2$ and $^1D_2$. We
687: approximately denote them with double quotations in Tables
688: \ref{Dmeson}--\ref{Bsmeson}. The relations between them are given by
689: \cite{Matsuki97}
690: %
691: \begin{equation}
692: \left(
693: \begin{array}{c}
694: |"^3P_1"\rangle \\ |"^1P_1"\rangle
695: \end{array}
696: \right)
697: =\frac{1}{\sqrt{3}}
698: \left(
699: \begin{array}{cc}
700: \sqrt{2} & 1 \\ -1 & \sqrt{2}
701: \end{array}
702: \right)
703: \left(
704: \begin{array}{c}
705: |^3P_1\rangle \\ |^1P_1\rangle
706: \end{array}
707: \right) ,
708: \end{equation}
709: %
710: %
711: \begin{equation}
712: \left(
713: \begin{array}{c}
714: |"^3D_2"\rangle \\ |"^1D_2"\rangle
715: \end{array}
716: \right)
717: =\frac{1}{\sqrt{5}}
718: \left(
719: \begin{array}{cc}
720: \sqrt{3} & \sqrt{2} \\ -\sqrt{2} & \sqrt{3}
721: \end{array}
722: \right)
723: \left(
724: \begin{array}{c}
725: |^3D_2\rangle \\ |^1D_2\rangle
726: \end{array}
727: \right) .
728: \end{equation}
729: %
730: 
731: \subsection{$D$ and $D_s$}
732: Calculated and optimally fitted values $M_{\rm calc}$ of $D$ and $D_s$ meson
733: masses with the first order corrections are presented in Tables \ref{Dmeson}
734: and \ref{Dsmeson}, respectively, together with experimental data $M_{\rm obs}$.
735: Those mass spectra of $D$ and $D_s$ mesons are also plotted in Figs.
736: \ref{fig-D} and \ref{fig-Ds} and Figure \ref{fig-Ds} also includes
737: $DK/D^{*}K$ threshold lines.
738: As one can see easily, our calculated masses $M_{\rm calc}$ are in good
739: agreement with each observed value $M_{\rm obs}$, and the discrepancies are
740: less than $1\%$. Especially it is remarkable that newly observed levels
741: $^3P_0(0^+)$ and $"^3P_1(1^+)"$ of both $D$ and $D_s$ mesons are reproduced
742: well, where the masses of $D_{sJ}$ mesons are below $DK/D^{*}K$ thresholds.
743: These levels of $D$ and $D_s$ mesons cannot be reproduced by any other quark
744: potential models
745: which predict about $100\sim 200$ MeV higher than our values even though they
746: also succeed in reproducing other levels, i.e., $^1S_0(0^-)$, $^3S_1(1^-)$,
747: $"^1P_1(1^+)"$, and $^3P_2(2^+)$.\footnote{See Table XIV of \cite{Swanson06}.}
748: %
749: This result gives us a great confidence that our framework may give a good solution to the narrow $D_{sJ}$ puzzle.
750: 
751: %
752: \begin{table*}[t!]
753: \caption{$D$ meson mass spectra (units are in MeV).}
754: \label{Dmeson}
755: \begin{tabular}{@{\hspace{0.5cm}}c@{\hspace{0.5cm}}c@{\hspace{1cm}}r@{\hspace{1cm}}r@{\hspace{1cm}}r@{\hspace{1cm}}c@{\hspace{1cm}}c@{\hspace{0.5cm}}}
756: \hline
757: \hline
758: $^{2s+1}L_J (J^P)$ & $M_0$ & 
759: \multicolumn{1}{c@{\hspace{1cm}}}{$c_1 /M_0$} & 
760: \multicolumn{1}{c@{\hspace{1cm}}}{$p_1 /M_0$} & 
761: \multicolumn{1}{c@{\hspace{1cm}}}{$n_1 /M_0$} & 
762: $M_{\rm calc}$ & $M_{\rm obs}$ \\
763: \hline
764: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^1S_0 (0^-)$} 
765: & 1784 & 0.476 $\times 10^{-1}$ 
766: & 0.374 $\times 10^{-1}$ & 1.013 $\times 10^{-2}$ 
767: & 1869 & 1867 \\
768: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3S_1 (1^-)$} 
769: &  & 1.271 $\times 10^{-1}$ 
770: & 1.266 $\times 10^{-1}$ & 0.512 $\times 10^{-3}$ 
771: & 2011 & 2008 \\
772: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_0 (0^+)$} 
773: & 2067 & 1.046 $\times 10^{-1}$ 
774: & 0.959 $\times 10^{-1}$ & 0.874 $\times 10^{-2}$ 
775: & 2283 & 2308 \\
776: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3P_1" (1^+)$} 
777: &  & 1.713 $\times 10^{-1}$ 
778: & 1.689 $\times 10^{-1}$ & 2.444 $\times 10^{-3}$ 
779: & 2421 & 2427 \\
780: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^1P_1" (1^+)$} 
781: & 2125 & 1.415 $\times 10^{-1}$ 
782: & 1.410 $\times 10^{-1}$ & 0.486 $\times 10^{-3}$ 
783: & 2425 & 2420 \\
784: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_2 (2^+)$} 
785: &  & 1.618 $\times 10^{-1}$ 
786: & 1.617 $\times 10^{-1}$ & 1.364 $\times 10^{-4}$ 
787: & 2468 & 2460 \\
788: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3D_1 (1^-)$} 
789: & 2322 & 1.894 $\times 10^{-1}$ 
790: & 1.872 $\times 10^{-1}$ & 2.228 $\times 10^{-3}$ 
791: & 2762 & $-$ \\
792: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3D_2" (2^-)$} 
793: &  & 2.054 $\times 10^{-1}$ 
794: & 2.052 $\times 10^{-1}$ & 1.248 $\times 10^{-4}$ 
795: & 2800 & $-$ \\
796: %%%%%%%%%%%%%%%%%%%%%%%%%%%
797: \hline
798: \hline
799: \end{tabular}
800: \end{table*}
801: %
802: \begin{table*}[t!]
803: \caption{$D_s$ meson mass spectra (units are in MeV).}
804: \label{Dsmeson}
805: \begin{tabular}{@{\hspace{0.5cm}}c@{\hspace{0.5cm}}c@{\hspace{1cm}}
806: r@{\hspace{1cm}}r@{\hspace{1cm}}r@{\hspace{1cm}}c@{\hspace{1cm}}c@{\hspace{0.5cm}}}
807: \hline
808: \hline
809: $^{2s+1}L_J (J^P)$ & $M_0$ & 
810: \multicolumn{1}{c@{\hspace{1cm}}}{$c_1 /M_0$} & 
811: \multicolumn{1}{c@{\hspace{1cm}}}{$p_1 /M_0$} & 
812: \multicolumn{1}{c@{\hspace{1cm}}}{$n_1 /M_0$} & 
813: $M_{\rm calc}$ & $M_{\rm obs}$ \\
814: \hline
815: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^1S_0 (0^-)$} 
816: & 1900 & 0.352 $\times 10^{-1}$ 
817: & 0.270 $\times 10^{-1}$ & 0.816 $\times 10^{-2}$ 
818: & 1967 & 1969 \\
819: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3S_1 (1^-)$} 
820: &  & 1.102 $\times 10^{-1}$ 
821: & 1.098 $\times 10^{-1}$ & 4.076 $\times 10^{-4}$ 
822: & 2110 & 2112 \\
823: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_0 (0^+)$} 
824: & 2095 & 1.101 $\times 10^{-1}$ 
825: & 1.027 $\times 10^{-1}$ & 0.740 $\times 10^{-2}$ 
826: & 2325 & 2317 \\
827: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3P_1" (1^+)$} 
828: &  & 1.779 $\times 10^{-1}$ 
829: & 1.752 $\times 10^{-1}$ & 2.620 $\times 10^{-3}$ 
830: & 2467 & 2460 \\
831: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^1P_1" (1^+)$} 
832: & 2239 & 1.274 $\times 10^{-1}$ 
833: & 1.270 $\times 10^{-1}$ & 3.860 $\times 10^{-4}$ 
834: & 2525 & 2535 \\
835: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_2 (2^+)$} 
836: &  & 1.467 $\times 10^{-1}$ 
837: & 1.466 $\times 10^{-1}$ & 1.035 $\times 10^{-4}$ 
838: & 2568 & 2572 \\
839: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3D_1 (1^-)$} 
840: & 2342 & 2.032 $\times 10^{-1}$ 
841: & 2.008 $\times 10^{-1}$ & 2.382 $\times 10^{-3}$ 
842: & 2817 & $-$ \\
843: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3D_2" (2^-)$} 
844: &  & 2.196 $\times 10^{-1}$ 
845: & 2.195 $\times 10^{-1}$ & 0.989 $\times 10^{-4}$ 
846: & 2856 & $-$ \\
847: %%%%%%%%%%%%%%%%%%%%%%%%%%%
848: \hline
849: \hline
850: \end{tabular}
851: \end{table*}
852: %
853: 
854: The first order corrections for $D$ and $D_s$ mesons are moderately large,
855: whose amount is an order of $10\%$. Since the charm quark mass in our fitting
856: is rather small, second order corrections might be necessary to get more
857: reliable results. However, we have found that they are only an order of $1\%$
858: or less and thus have neglected them in this work.
859: To demonstrate this smallness, we also present the fit for $D$ meson with the
860: second order corrections in Table \ref{Dmeson2nd} as a typical example. When
861: one looks at this Table, one notices that the fit with the experiments does not
862: increase accuracy compared with Table \ref{Dmeson} of the first order.
863: Hence from
864: here on in our computation we use only the first order calculations to obtain
865: any levels including radial excitations which we present in a separate paper
866: \cite{Matsuki06}.
867: 
868: %
869: \begin{table*}[b]
870: \caption{Second order $D$ meson mass spectra (units are in MeV).}
871: \label{Dmeson2nd}
872: \begin{tabular}{@{\hspace{0.5cm}}c@{\hspace{1cm}}c@{\hspace{1cm}}
873: r@{\hspace{1cm}}r@{\hspace{1cm}}c@{\hspace{1cm}}c@{\hspace{0.5cm}}}
874: \hline
875: \hline
876: $^{2s+1}\ell_j~(j^P)$ & $M_0$ & 
877: \multicolumn{1}{@{\hspace{0.6cm}}l}{$c_1 /M_0$} & 
878: \multicolumn{1}{@{\hspace{0.6cm}}l}{$c_2 /M_0$} & 
879: $M_{\rm calc}$ & $M_{\rm obs}$ \\
880: \hline
881: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^1S_0(0^-)$} 
882: & 1783 & 0.678 $\times 10^{-1}$ &-2.091 $\times 10^{-2}$ & 1867 & 1867 \\
883: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3S_1(1^-)$} 
884: & & 1.245 $\times 10^{-1}$ & 2.507 $\times 10^{-3}$ & 2009 & 2008 \\
885: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_0(0^+)$} 
886: & 1935 & 1.694 $\times 10^{-1}$ & 1.567 $\times 10^{-2}$ & 2293 & 2308 \\
887: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3P_1"(1^+)$} 
888: & & 2.132 $\times 10^{-1}$ & 1.665 $\times 10^{-3}$ & 2350 & 2427 \\
889: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^1P_1"(1^+)$} 
890: & 2045 & 1.412 $\times 10^{-1}$ & 2.499 $\times 10^{-3}$ & 2432 & 2420 \\
891: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_2(2^+)$} 
892: & & 1.616 $\times 10^{-1}$ & -1.003 $\times 10^{-3}$ & 2448 & 2460 \\
893: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3D_1(1^-)$} 
894: & 2127 & 1.863 $\times 10^{-1}$ & 2.109 $\times 10^{-4}$ & 2803 & $-$ \\
895: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3D_2"(2^-)$} 
896: & & 2.021 $\times 10^{-1}$ & -1.409 $\times 10^{-3}$ & 2726 & $-$ \\
897: %%%%%%%%%%%%%%%%%%%%%%%%%%%
898: \hline
899: \hline
900: \end{tabular}
901: \end{table*}
902: %
903: 
904: \subsection{$B$ and $B_s$}
905: For $B$ and $B_s$ mesons, unfortunately there are only a few data available
906: although newly discovered states ($B_1(5720)$, $B_2^*(5745)$, and
907: $B_{s2}^*(5839)$) are reported recently by D0 and CDF.\cite{CDF_D006}
908: Calculated masses of $B$ and $B_s$ mesons are given in Tables \ref{Bmeson} and
909: \ref{Bsmeson}, respectively. We have used a new value of a strong coupling
910: $\alpha_s^b$ given by Table \ref{parameter} for calculating $B$ and $B_s$ mesons
911: different from $\alpha_s^c$ for $D$ meson in Table \ref{parameter}.
912: Newly discovered levels are well reproduced by our model, too, as shown in these
913: Tables. We predict the mass of several excited states which are not yet observed.
914: The mass spectra of $B$ and $B_s$ mesons are also plotted in Figs. \ref{fig-B}
915: and \ref{fig-Bs} and Figure \ref{fig-Bs} also includes $BK/B^{*}K$
916: threshold lines. Among these levels, predicted masses of $0^+$ and $1^+$ states
917: for $B_s$ mesons are below $BK/B^{*}K$ threshold the same as the case for
918: $D_{sJ}$ mesons. Therefore, their decay modes are kinematically forbidden, and
919: the dominant decay modes are the pionic decay which is the same results obtained in
920: \cite{Matsuki05} although values of $B_s$ mesons are slightly different from each
921: other:
922: %
923: \begin{subequations}
924: \begin{eqnarray}
925: B_s (0^{+}) &\to& B_s (0^{-})+\pi , \\
926: B_s (1^{+}) &\to& B_s (1^{-})+\pi .
927: \end{eqnarray}
928: \label{Bs-decay}
929: \end{subequations}
930: %
931: The decay widths of these states are expected to be narrow the same as $D_{sJ}$
932: meson cases, since those decay modes of Eq.(\ref{Bs-decay}) violate the isospin
933: invariance. These higher states might be observed in Tevatron/LHC experiments or
934: even in D0 and/or CDF of FNAL in the near future by analyzing the above decay modes.
935: This is because D0 and CDF have recently announced the discovery of
936: $B_1(5720)=B("^1P_1(1^+)")$, $B_2^*(5745)=B(^3P_2(2^+))$, and
937: $B_{s2}^*(5839)=B_s(^3P_2(2^+))$, which have narrow decay width with the same decay
938: products as in Eq.(\ref{Bs-decay}) since only the D-wave decay channel is
939: possible.\cite{CDF_D006} Even though these mass values are reproduced by other
940: models \cite{Eichten93, Pierro01, Ebert98, Orsland99}, but again those models
941: predict larger mass values for $0^+$ and $1^+$ states of $B_s$ mesons
942: than ours, the same situation as in $D_{sJ}$ so that these states have broad decay
943: width contrary to our prediction. We hope that our prediction could be
944: tested in the forthcoming experiments which distinguish ours from other models.
945: 
946: %
947: \begin{table*}[t!]
948: \caption{$B$ meson mass spectra (units are in MeV).}
949: \label{Bmeson}
950: \begin{tabular}{@{\hspace{0.5cm}}c@{\hspace{0.5cm}}c@{\hspace{1cm}}
951: r@{\hspace{1cm}}r@{\hspace{1cm}}r@{\hspace{1cm}}c@{\hspace{1cm}}c@{\hspace{0.5cm}}}
952: \hline
953: \hline
954: $^{2s+1}L_J (J^P)$ & $M_0$ & 
955: \multicolumn{1}{c@{\hspace{1cm}}}{$c_1 /M_0$} & 
956: \multicolumn{1}{c@{\hspace{1cm}}}{$p_1 /M_0$} & 
957: \multicolumn{1}{c@{\hspace{1cm}}}{$n_1 /M_0$} & 
958: $M_{\rm calc}$ & $M_{\rm obs}$ \\
959: \hline
960: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^1S_0 (0^-)$} 
961: & 5277 & -0.161 $\times 10^{-2}$ 
962: & -3.795 $\times 10^{-3}$ & 2.187 $\times 10^{-3}$ 
963: & 5270 & 5279 \\
964: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3S_1 (1^-)$} 
965: &  & 0.981 $\times 10^{-2}$ 
966: & 9.706 $\times 10^{-3}$ & 1.107 $\times 10^{-4}$ 
967: & 5329 & 5325 \\
968: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_0 (0^+)$} 
969: & 5570 & 0.401 $\times 10^{-2}$ 
970: & 1.937 $\times 10^{-3}$ & 2.072 $\times 10^{-3}$ 
971: & 5592 & $-$ \\
972: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3P_1" (1^+)$} 
973: &  & 1.412 $\times 10^{-2}$ 
974: & 1.400 $\times 10^{-2}$ & 1.227 $\times 10^{-4}$ 
975: & 5649 & $-$ \\
976: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^1P_1" (1^+)$} 
977: & 5660 & 1.069 $\times 10^{-2}$ 
978: & 1.066 $\times 10^{-2}$ & 0.289 $\times 10^{-4}$ 
979: & 5720 & 5720 \\
980: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_2 (2^+)$} 
981: &  & 1.364 $\times 10^{-2}$ 
982: & 1.364 $\times 10^{-2}$ & 2.120 $\times 10^{-7}$ 
983: & 5737 & 5745 \\
984: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3D_1 (1^-)$} 
985: & 5736 & 2.203 $\times 10^{-1}$ 
986: & 2.202 $\times 10^{-1}$ & 4.583 $\times 10^{-5}$ 
987: & 6999 & $-$ \\
988: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3D_2" (2^-)$} 
989: &  & 1.430 $\times 10^{-1}$ 
990: & 1.430 $\times 10^{-1}$ & 2.092 $\times 10^{-7}$ 
991: & 6556 & $-$ \\
992: %%%%%%%%%%%%%%%%%%%%%%%%%%%
993: \hline
994: \hline
995: \end{tabular}
996: \end{table*}
997: %
998: 
999: %
1000: \begin{table*}[t!]
1001: \caption{$B_s$ meson mass spectra (units are in MeV).}
1002: \label{Bsmeson}
1003: \begin{tabular}{@{\hspace{0.5cm}}c@{\hspace{0.5cm}}c@{\hspace{1cm}}
1004: r@{\hspace{1cm}}r@{\hspace{1cm}}r@{\hspace{1cm}}c@{\hspace{1cm}}c@{\hspace{0.5cm}}}
1005: \hline
1006: \hline
1007: $^{2s+1}L_J (J^P)$ & $M_0$ & 
1008: \multicolumn{1}{c@{\hspace{1cm}}}{$c_1 /M_0$} & 
1009: \multicolumn{1}{c@{\hspace{1cm}}}{$p_1 /M_0$} & 
1010: \multicolumn{1}{c@{\hspace{1cm}}}{$n_1 /M_0$} & 
1011: $M_{\rm calc}$ & $M_{\rm obs}$ \\
1012: \hline
1013: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^1S_0 (0^-)$} 
1014: & 5394 & -0.302 $\times 10^{-2}$ 
1015: & -0.485 $\times 10^{-2}$ & 0.183 $\times 10^{-2}$ 
1016: & 5378 & 5369 \\
1017: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3S_1 (1^-)$} 
1018: &  & 0.853 $\times 10^{-2}$ 
1019: & 0.844 $\times 10^{-2}$ & 9.249 $\times 10^{-5}$ 
1020: & 5440 & $-$ \\
1021: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_0 (0^+)$} 
1022: & 5598 & 0.350 $\times 10^{-2}$ 
1023: & 0.173 $\times 10^{-2}$ & 0.177 $\times 10^{-2}$ 
1024: & 5617 & $-$ \\
1025: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3P_1" (1^+)$} 
1026: &  & 1.498 $\times 10^{-2}$ 
1027: & 1.444 $\times 10^{-2}$ & 5.396 $\times 10^{-4}$ 
1028: & 5682 & $-$ \\
1029: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^1P_1" (1^+)$} 
1030: & 5775 & 0.978 $\times 10^{-2}$ 
1031: & 0.969 $\times 10^{-2}$ & 8.257 $\times 10^{-5}$ 
1032: & 5831 & $-$ \\
1033: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3P_2 (2^+)$} 
1034: &  & 1.263 $\times 10^{-2}$ 
1035: & 1.261 $\times 10^{-2}$ & 2.014 $\times 10^{-5}$ 
1036: & 5847 & 5839 \\
1037: \multicolumn{1}{@{\hspace{0.6cm}}l}{$^3D_1 (1^-)$} 
1038: & 5875 & 2.949 $\times 10^{-2}$ 
1039: & 2.898 $\times 10^{-2}$ & 5.104 $\times 10^{-4}$ 
1040: & 6048 & $-$ \\
1041: \multicolumn{1}{@{\hspace{0.6cm}}l}{$"^3D_2" (2^-)$} 
1042: &  & 0.564 $\times 10^{-2}$ 
1043: & 0.562 $\times 10^{-2}$ & 1.980 $\times 10^{-5}$ 
1044: & 5908 & $-$ \\
1045: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1046: \hline
1047: \hline
1048: \end{tabular}
1049: \end{table*}
1050: %
1051: 
1052: %--------------------------------------------------------------
1053: \section{Conclusion and Discussion}
1054: \label{summary}
1055: Given recent new data and expanding a parameter space with positive $b$,
1056: we have reanalyzed the mass spectra of heavy mesons, $D$, $D_s$, $B$ and $B_s$
1057: with our semirelativistic quark potential model, in which an effective
1058: Hamiltonian has both a heavy quark symmetry and a chiral symmetry in a certain
1059: limit of parameters. Hamiltonian, wave function, and eigenvalue are all
1060: consistently expanded in $p/m_Q$, and a light antiquark is treated as four-spinor
1061: Dirac particle while a heavy quark is nonrelativistically reduced using the FWT
1062: transformation.
1063: Calculated masses of not only newly observed excited ($\ell=1$) states,
1064: $D_{s0}(2317)$, $D_{s1}(2460)$, $D_{0}^{*}(2308)$ and $D_{1}'(2427)$ which are
1065: not explained by conventional quark potential models, but also already
1066: established ($\ell=0$ and $\ell=1$) states of $D$ and $D_s$ mesons are
1067: simultaneously reproduced in good agreement with experimental data within one
1068: percent of accuracy. Now we also have plausible
1069: %reasonable
1070: values of masses for $^3D_1$
1071: and $"^3D_2"$ which we failed to give in the former paper \cite{Matsuki97}.
1072: This has been achieved by finding the most optimal parameter set with positive
1073: parameter $b$.
1074: 
1075: These results suggest that our model could be a good solution to the narrow
1076: $D_{sJ}$ puzzle, which is considered to be still challenging and people are
1077: still looking for exotic state possibilities. We would say that
1078: the quark potential model still remains powerful enough to understand the heavy
1079: meson spectroscopy. The important thing in a heavy-light system is to expand all
1080: quantities in the model in $p/m_Q$ and set equations consistently order by order
1081: and not to neglect the lower components of a heavy quark, which
1082: has not been taken into account by other quark potential models. We have
1083: also predicted the masses of higher excited states of $B$ and $B_s$ mesons
1084: which are not yet observed, at the first order in $p/m_b$.  Among these, the
1085: newly discovered states by D0 and CDF are also reproduced by our model, too.
1086: There are other models to reproduce these states \cite{Eichten93, Pierro01,
1087: Ebert98, Orsland99} but they cannot predict $0^+$ and $1^+$ states less than
1088: $BK/B^{*}K$ threshold. This is the main difference of the results when
1089: comparing our model with others. Other models for heavy mesons cannot yield the
1090: masses of $^3P_0(0^+)$ and $"^3P_1(1^+)"$ states of $D_s$ and $B_s$ mesons much
1091: lower than $"^1P_1(1^+)"$ and $^3P_2(2^+)$ states like ours.
1092: 
1093: Some comments are in order.
1094: 
1095: (i) As shown in Tables \ref{Dmeson}--\ref{Bsmeson}, the first order corrections
1096: to the $0$-th order mass, $c_{1}/M_0$, are roughly an order of $10\%$ for $D$
1097: and $D_s$ mesons, while $1\%$ for $B$ and $B_s$ mesons. Therefore, $p/m_Q$
1098: expansion works much better for $B$ and $B_s$ mesons. This is quite natural,
1099: since the bottom quark mass is heavier than the charm quark mass. In addition,
1100: it is interesting that the first order correction becomes larger for higher
1101: excited states. It means that the relativistic correction must be more
1102: important in higher excited states, since the higher order corrections in
1103: $p/m_Q$ expansion represent the relativistic effects of a heavy quark. This
1104: is consistent with a naive picture that the inner motion of light as well as
1105: heavy quarks may be larger in excited states, and therefore the effects of
1106: lower components of a heavy quark may become much more important to be included.
1107: 
1108: (ii) The mass difference between members of a spin doublet satisfies the following
1109: relation (given by Eq.(53) of \cite{Matsuki97}):
1110: %
1111: \begin{eqnarray}
1112:   m_{c}(M_{D^{*}}-M_{D}) &=& m_{b}(M_{B^{*}}-M_{B}), \label{spin_doublet1} \\
1113:   m_{c}(M_{D^{*}_s}-M_{D_s}) &=& m_{b}(M_{B^{*}_s}-M_{B_s}),
1114:   \label{spin_doublet2}
1115: \end{eqnarray}
1116: %
1117: in the first order of calculation. This is because each of two states having
1118: the same value of the quantum number $k$ has an equal mass at the $0$-th order
1119: and hence the mass is degenerate at this stage as shown in
1120: Fig. \ref{mass_level}. The splitting occurs by including the first order
1121: ($p/m_Q$) corrections originated from $H_1$ in Eq.(\ref{H_mQ}), which is
1122: proportional to $f^k(m_q)/m_Q$ with the same functional form $f^k(m_q)$ for two
1123: states, and therefore Eqs.(\ref{spin_doublet1}, \ref{spin_doublet2}) are
1124: exactly satisfied in our model. Similar relations hold for higher spin states
1125: with the same value of $k$.
1126: 
1127: (iii) As shown in Table \ref{parameter}, the light quark masses $m_{u, d}$ and
1128: $m_s$ are considerably small, $m_{u, d}\sim 10$ MeV and $m_{s}\sim 90$ MeV,
1129: compared with the constituent quark masses which are adopted in
1130: conventional potential models \cite{Godfrey85}.  It should be noted that these
1131: values are not input but the outcome of this analysis, though in the previous
1132: analysis \cite{Matsuki97}, $m_{u,d}=10$ MeV is taken as an input.
1133: This result is consistent with those obtained in
1134: \cite{Politzer, Matsuki80, Kaburagi}.
1135: 
1136: This situation might be described as follows. A light quark is moving around a
1137: heavy quark in a heavy meson wearing small gluon clouds with large momenta in
1138: our treatment of quarks so that light quarks have so-called current quark mass.
1139: This holds only in heavy-light systems and in the other systems, e.g., quarkonium
1140: and ordinary hadrons, we should use quarks wearing more gluon clouds, i.e.,
1141: constituent quarks as adopted in usual quark potential models. There may be
1142: another interpretation bridging our model with light mass with a constituent quark
1143: model.
1144: 
1145: (iv) Repeatedly saying, the main difference between other conventional
1146: quark potential models and ours is that our model systematically
1147: expands interaction terms in $p/m_Q$, takes into account lower component
1148: contributions of heavy quarks in the intermediate states when calculating
1149: higher order terms, and adopt very light antiquark mass. With regard to $1/m_Q$
1150: expansion, for instance, the model
1151: of \cite{Godfrey85,Godfrey91} has kinetic terms, $\sqrt{p^2+m^2}$, while they
1152: adopted nonrelativistic interaction terms which are symmetric in $Q$ and
1153: $\bar q$ or in $m_Q$ and $m_{\bar q}$.
1154: The model of \cite{Ebert98} somehow manipulated to hold heavy quark symmetry
1155: and expanded interactions in $1/m_Q$ but adopted constituent masses for light
1156: quarks, $m_u$, $m_d$, and $m_s$. Even so, their numerical results give about a
1157: hundred MeV larger values for $D_{sJ}$ particles.
1158: Both models have not taken into account the lower components of a heavy
1159: quark which naturally appear in our semirelativistic potential model.
1160: 
1161: (v) We have been successful in reproducing experimental data for both heavy
1162: mesons $D/D_s$ and $B/B_s$. However we have also left unresolved problems in
1163: our semirelativistic model. When one looks at Table \ref{parameter}, one notices
1164: that values of a strong coupling used for $D/D_s$ and $B/B_s$ are largely
1165: different from each other. Especially that for $D/D_s$ is too small to be
1166: a strong coupling. Also we have not used a running strong coupling although
1167: most of other quark potential models take into account this effect in some way.
1168: 
1169: 
1170: \begin{figure}[t]
1171: \includegraphics[scale=0.5,clip]{D.eps}
1172: \caption{Plot of $D$ meson masses in GeV. 
1173: Solid bars and white circles represent calculated and observed masses, respectively. 
1174: Specific values are given in Table \ref{Dmeson}. }
1175: \label{fig-D}
1176: \end{figure}
1177: 
1178: \begin{figure}[t]
1179: \includegraphics[scale=0.5,clip]{Ds.eps}
1180: \caption{Plot of $D_s$ meson masses. 
1181: The legend is as for Fig. \ref{fig-D}. 
1182: The dashed lines show $DK$ and $D^{*}K$ thresholds, respectively. 
1183: Specific values are given in Table \ref{Dsmeson}.}
1184: \label{fig-Ds}
1185: \end{figure}
1186: 
1187: \begin{figure}[t]
1188: \includegraphics[scale=0.5,clip]{B.eps}
1189: \caption{Plot of $B$ meson masses. 
1190: The legend is as for Fig. \ref{fig-D}. 
1191: Specific values are given in Table \ref{Bmeson}.}
1192: Calculated mass values for $^3D_1$ and "$^3D_2$" are too large to
1193: put them in the same figure with others.
1194: \label{fig-B}
1195: \end{figure}
1196: 
1197: \begin{figure}[t]
1198: \includegraphics[scale=0.5,clip]{Bs.eps}
1199: \caption{Plot of $B_s$ meson masses. 
1200: The legend is as for Fig. \ref{fig-D}. 
1201: The dashed lines show $BK$ and $B^{*}K$ thresholds, respectively. 
1202: Specific values are given in Table \ref{Bsmeson}.}
1203: \label{fig-Bs}
1204: \end{figure}
1205: 
1206: %--------------------------------------------------------------
1207: \begin{acknowledgments}
1208: One of the authors (K.S.) would like to thank S. Yokkaichi for 
1209: helpful discussions about numerical calculations.
1210: \end{acknowledgments}
1211: 
1212: 
1213: %--------------------------------------------------------------
1214: \begin{appendix}
1215: \section{Schr\"odinger equation with an Effective Hamiltonian}
1216: \label{appen_Hamiltonian}
1217: The original Schr\"odinger equation is actually modified by the FWT
1218: transformation in our framework in order to make non-relativistic reduction of a
1219: heavy quark. Now we recall the modified Schr\"odinger
1220: equation of Eq. (\ref{eigen_FWT}), 
1221: %
1222: \begin{equation}
1223: (H_{\rm FWT}-m_{Q}) \otimes\psi_{\rm FWT} =\tilde{E}\psi_{\rm FWT},
1224: \end{equation}
1225: %
1226: The Hamiltonian and the wave function are transformed by the
1227: following FWT and charge conjugate operator on the heavy quark sector:
1228: %
1229: \begin{eqnarray}
1230:   H_{FWT}&=&U_{c}U_{FWT}(p'_{Q})HU^{-1}_{FWT}(p_{Q})U^{-1}_{c} , \\
1231: %
1232:   \psi_{FWT}&=&U_{c}U_{FWT}(p_{Q})\psi , 
1233: \end{eqnarray}
1234: %
1235: where given are
1236: %
1237: \begin{eqnarray}
1238:   U_{FWT}(p)&=&\exp\left(W(p)\vmb{\gamma}_{Q}\cdot
1239:   \vec{\hat{\mbox{\boldmath$p$}}}\right)
1240: %  \nonumber \\
1241: %  &=& 
1242:   =\cos W(p) + \vmb{\gamma}_{Q}\cdot \vec{\hat{\mbox{\boldmath$p$}}}
1243:   \sin W(p) , \\
1244: %
1245:   U_{c}&=& i\gamma^0_{Q}\gamma^2_{Q}=-U^{-1}_{c} , 
1246: \end{eqnarray}
1247: %
1248: and some kinematical variables are defined as 
1249: %
1250: \begin{eqnarray}
1251:   && \vec{\hat{\mbox{\boldmath$p$}}}=\frac{\vmb{p}}{p}, ~~~
1252:   \tan W(p)=\frac{p}{m_{Q}+E}, ~~~E=\sqrt{\vmb{p}^{2}+m^2_{Q}} , \\
1253: %
1254:   && \vmb{p}=\vmb{p}_{q}=-\vmb{p}_{Q}, ~~~\vmb{p}'=\vmb{p}'_{q}=-\vmb{p}'_{Q} ,
1255:   ~~~\vmb{q}=\vmb{p}'-\vmb{p}.
1256: \end{eqnarray}
1257: %
1258: Here we are considering a scattering of a heavy quark $Q$ and a light antiquark
1259: $\bar{q}$. Hence, $\vmb{p}_i$, $\vmb{p}_i'$, and $\vmb{q}$ given above
1260: represent momenta of incoming, outgoing quarks, and gluon which is exchanged
1261: between quarks, respectively. Note that the argument of the FWT transformation
1262: $U_{FWT}$ operating on a Hamiltonian from left is different from the
1263: right-operating one, since an outgoing momentum $\vmb{p}'_{Q}$ is different
1264: from an incoming one $\vmb{p}_{Q}$. However, in our study we work in a
1265: configuration space in which momenta are nothing but the derivative operators.
1266: When we write them differently, for instance as $\vmb{p}_{Q}$ and
1267: $\vmb{p}'_{Q}$, their expressions are reminders of their momentum
1268: representation. 
1269: Therefore, although the arguments of $U_{FWT}$ and $U^{-1}_{FWT}$ look
1270: different, $\vmb{p}_{Q}$ and $\vmb{p}'_{Q}$ are expressed by the same derivative
1271: operator $-i\vmb{\nabla}$. The momentum transfer $\vmb{q}$ operates only on
1272: potentials and provides nonvanishing results. 
1273: 
1274: The charge conjugation operator $U_c$ is introduced to make the wave function
1275: $U_{c}\psi$ a true bi-spinor, i.e., gamma matrices of a light antiquark are
1276: multiplied from left while those of a heavy quark from right, which is
1277: expressed by using a notation $\otimes$. 
1278: 
1279: Then, the Hamiltonian in the modified Schr\"odinger equation given by Eq.
1280: (\ref{eigen_FWT}) is expanded in powers of $p/m_Q$: 
1281: %
1282: \begin{equation}
1283: H_{\rm FWT}-m_{Q}=H_{-1}+H_{0}+H_{1}+H_{2}+\cdots . 
1284: \end{equation}
1285: %
1286: $H_{i}$ stands for the $i$-th order expanded Hamiltonian whose explicit forms
1287: are given by 
1288: %
1289: \begin{widetext}
1290: \begin{subequations}
1291: \begin{eqnarray}
1292: H_{-1}&=&-(1+\beta_{Q})m_{Q} , 
1293: \label{Hm1} \\
1294: H_{0~}&=&\vmb{\alpha}_{q}\cdot\vmb{p} 
1295: + \beta_{q}m_{q} 
1296: - \beta_{q}\beta_{Q}S 
1297: + \left\{1 + \frac{1}{2}\left[\vmb{\alpha}_{q}\cdot \vmb{\alpha}_{Q} +
1298: (\vmb{\alpha}_{q}\cdot \vmb{n})(\vmb{\alpha}_{Q}\cdot \vmb{n}) \right]\right\} V , \\
1299: H_{1~}&=&-\frac{1}{2m_{Q}}\beta_{Q}\vmb{p}^{2} 
1300: + \frac{1}{m_{Q}}\beta_{q}\vmb{\alpha}_{Q}\cdot
1301: \left(\vmb{p}+\frac{1}{2}\vmb{q}\right)S 
1302: + \frac{1}{2m_{Q}}\vmb{\gamma}_{Q}\cdot \vmb{q}V \nonumber \\
1303: &&- \frac{1}{2m_{Q}}\left[\beta_{Q}\left(\vmb{p}+\frac{1}{2}\vmb{q}\right) +
1304: i\vmb{q}\times\beta_{Q}\vmb{\Sigma}_{Q}\right]
1305: \cdot \left[\vmb{\alpha}_{q} + (\vmb{\alpha}_{q}\cdot \vmb{n})\vmb{n}\right] V , \\
1306: H_{2~}&=&\frac{1}{2m_Q^2}\beta_{q}\beta_{Q}
1307: \left(\vmb{p}+\frac{1}{2}\vmb{q}\right)^{2}S 
1308: - \frac{i}{4m_Q^2}\vmb{q}\times \vmb{p}\cdot \beta_{q}\beta_{Q}\vmb{\Sigma}_{Q}S 
1309: - \frac{1}{8m_Q^2}\vmb{q}^{2}V 
1310: - \frac{i}{4m_Q^2}\vmb{q}\times\vmb{p}\cdot \vmb{\Sigma}_{Q}V \nonumber \\
1311: &&- \frac{1}{8m_Q^2}\left\{(\vmb{p}+\vmb{q})(\vmb{\alpha}_{Q}\cdot\vmb{p})
1312: + \vmb{p}\left[\vmb{\alpha}_{Q}\cdot(\vmb{p}+\vmb{q})\right] 
1313: + i\vmb{q}\times\vmb{p}\gamma_Q^5 \right\}
1314: \cdot \left[\vmb{\alpha}_{q}+(\vmb{\alpha}_{q}\cdot\vmb{n})\vmb{n}\right] V , \\
1315: && \vdots \nonumber 
1316: \end{eqnarray}
1317: \label{H_appen}
1318: \end{subequations}
1319: \end{widetext}
1320: %
1321: More details of matrix elements of each order in $1/m_Q$ and properties of
1322: wave functions with several operators are given and evaluated in Ref.
1323: \cite{Matsuki97}.
1324: 
1325: Eq.(\ref{Hm1}) means that the lowest order equation gives
1326: $(1+\beta_Q)\psi=0$, which implies that a heavy quark is regarded as a
1327: static source of color so that projection operator becomes
1328: $(1\pm\beta_Q)/2$, the one at rest system. That is, the solution given by
1329: Eq.(\ref{0thsols1}) includes only upper or positive components with regard
1330: to a heavy quark. In this sense, upper/lower component of a heavy quark has the
1331: same meaning as positive/negative energy component.
1332: 
1333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1334: %% Seo'd definitions
1335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1336: 
1337: \font\singlebf=cmbx10 scaled\magstep1
1338: \font\singlerm=cmr10 scaled\magstep1
1339: \font\singless=cmss10 scaled\magstep1
1340: \font\doublebf=cmbx10 scaled\magstep2
1341: \font\doublerm=cmr10 scaled\magstep2
1342: \font\triplebf=cmbx10 scaled\magstep3
1343: \font\triplerm=cmr10 scaled\magstep3
1344: \font\hsym=cmsy10
1345: \font\singlehsym=cmsy10 scaled\magstep1
1346: \font\sevenit=cmti7
1347: \def\hana#1{\hbox{\hsym #1}}
1348: \def\bighana#1{\hbox{\singlehsym #1}}
1349: \def\yohaku{\hskip1000pt minus 1fill}
1350: \def\tabrule{\noalign{\hrule}}
1351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1352: \section{Bloch Method}
1353: 
1354: This Appendix is based on the paper \cite{MS97}, by which
1355: we explain Bloch's method \cite{Bloch} for a degenerate system. We will see the
1356: Foldy-Wouthuysen-Tani \cite{Foldy50} transformation is automatically carried out
1357: and will obtain the systematic expansion of the heavy-light system
1358: in $1/m_Q$. To distinguish this formulation from the one we have used, we use
1359: different notations in this Appendix.
1360: 
1361:      We expand the eigenvalues of the Hamiltonian with respect to the inverse of
1362: $m_Q$.  First we divide the Hamiltonian as 
1363: ${\hana H}=\beta_Qm_Q+{\hana V}$, and we treat ${\hana V}$ as a perturbation,
1364: where actually ${\hana H}=H$.
1365: The eigenvalues of the unperturbed Hamiltonian are simply $\pm m_Q$, and the
1366: eigenstates of ${\hana H}$ are almost degenerate.  Then we apply the
1367: perturbation method of Bloch \cite{Bloch}.  In general, let us consider the
1368: Hamiltonian 
1369: ${\hana H}=H_0+{\hana V}$.  We denote the projection operator to the subspace
1370: $M$ with an eigenvalue $E$ of $H_0$ as $P$, and the projection operator to the
1371: subspace ${\hana M}$ with eigenvalues $\{{\hana E}_j\}$ of 
1372: ${\hana H}$ as ${\hana P}$, where all ${\hana E}_j$ converge to $E$ in the
1373: vanishing limit of the interaction ${\hana V}$.  Further we choose appropriate
1374: vectors $\{\psi_j\}$ in the subspace $M$ such that their projections on
1375: ${\hana M}$ become the eigenvectors of ${\hana H}$, that is,
1376: ${\hana H}{\hana P}\psi_j={\hana E}_j{\hana P}\psi_j$.
1377: If we define an operator ${\hana U}$ by ${\hana P}P={\hana U}P{\hana P}P$, then
1378: we get the following eigenvalue equation,
1379: %%%
1380: \begin{equation}
1381:  {\hana H}_{eff}(P{\hana P}\psi_j)=
1382: {\hana E}_j(P{\hana P}\psi_j),
1383: \end{equation}
1384: %%\yohaku(2.2)\quad$$
1385: %%%
1386: where the effective Hamiltonian ${\hana H}_{eff}$ is defined by 
1387: ${\hana H}_{eff}=P{\hana H}{\hana U}$.  Here we must notice that this effective
1388: Hamiltonian is not hermitian.  The eigenfunction of ${\hana H}$, ${\hana P}\psi_j$,
1389: is obtained from the eigenfunction of 
1390: ${\hana H}_{eff}$ given by $\phi_j=P{\hana P}\psi_j$, by multiplying
1391: ${\hana U}$ from left, that is,
1392: %%%
1393: \begin{equation}
1394: {\hana U}\phi_j={\hana U}P{\hana P}\psi_j={\hana P}\psi_j.
1395: \end{equation}
1396: %%\yohaku(2.3)\quad$$
1397: %%%
1398: Note that $P\psi_j=\psi_j$.
1399: Again we must notice that ${\hana U}$ is not unitary, therefore, even if
1400: $\phi_j$ is normalized, ${\hana U}\phi_j$ is not.
1401: 
1402:      ${\hana P}$ and ${\hana U}$ are formally expanded as follows.
1403: %%%
1404: \begin{equation}
1405: {\hana P}=P-\sum_{m=1}^{\infty}\sum_{\{k_i\}}S_{k_1}{\hana V}S_{k_2}\cdots
1406: S_{k_m}{\hana V}S_{k_{m+1}},
1407: \end{equation}
1408: %%\yohaku(2.4)\quad $$
1409: %%%
1410: where $S_k$'s are defined by 
1411: %%%
1412: \begin{equation}
1413: S_0=-P\quad {\rm and} \quad S_k={Q\over a^k}\;\; (k=1,2,\cdots) 
1414: \quad {\rm with} \quad Q=1-P,\quad a=E-H_0,
1415: \end{equation}
1416: %%\yohaku (2.5)\quad$$
1417: %%%
1418: and the sum over $\{k_i\}$ extends over non-negative integers satisfying the
1419: condition 
1420: %% \par\noindent
1421: $\sum_{i=1}^{m+1}k_i=m$.
1422: %%%
1423: \begin{equation}
1424: {\hana U}=
1425: \sum_{m=1}^{\infty}{\sum_{\{k_i\}}}'
1426: S_{k_1}{\hana V}S_{k_2}\cdots S_{k_m}{\hana V}P,
1427: \end{equation}
1428: %%\yohaku(2.6)\quad $$
1429: %%%
1430: where the sum over $\{k_i\}$ extends over non-negative integers satisfying the
1431: conditions 
1432: %%% \par\noindent 
1433: $\sum_{i=1}^m k_i=m$ and
1434: \par\noindent $\sum_{i=1}^p k_i\ge p\quad (p=1,2,\cdots m-1)$.  
1435: Making use of the expansion formula for ${\hana U}$, we obtain the effective
1436: Hamiltonian including the second order corrections in $1/a\sim 1/m_Q$ as follows:
1437: %%%
1438: \begin{equation}
1439: {\hana H}_{eff}=EP+P{\hana V}P+P{\hana V}{Q\over a}{\hana V}P+
1440: P{\hana V}{Q\over a}{\hana V}{Q\over a}{\hana V}P-
1441: P{\hana V}{Q\over a^2}{\hana V}P{\hana V}P + \cdots.
1442: \end{equation}
1443: %%\yohaku (2.7)\quad$$
1444: %%%
1445: where $P=(1+\beta_Q)/2$ and $Q=(1-\beta_Q)/2$.
1446: 
1447: \end{appendix}
1448: %--------------------------------------------------------------
1449: 
1450: 
1451: %%%%%%%%%%%%%%%%%% reference %%%%%%%%%%%%%%%%%%%
1452: \def\Journal#1#2#3#4{{#1} {\bf #2}, #3 (#4)}
1453: \def\NIM{Nucl. Instrum. Methods}
1454: \def\NIMA{Nucl. Instrum. Methods A}
1455: \def\NPB{Nucl. Phys. B}
1456: \def\PLB{Phys. Lett. B}
1457: \def\PRL{Phys. Rev. Lett.}
1458: \def\PRD{Phys. Rev. D}
1459: \def\PRO{Phys. Rev.}
1460: \def\ZPC{Z. Phys. C}
1461: \def\EPJ{Eur. Phys. J. C}
1462: \def\PR{Phys. Rept.}
1463: \def\IJM{Int. J. Mod. Phys. A}
1464: \def\PTP{Prog. Theor. Phys.}
1465: 
1466: \begin{thebibliography}{00}
1467: \bibitem{BaBar03} BaBar Collaboration, B. Aubert {\em et al.}, 
1468:              \Journal{\PRL}{90}{242001}{2003}. 
1469: \bibitem{CLEO03} CLEO Collaboration, D. Besson {\em et al.}, 
1470:              \Journal{\PRD}{68}{032002}{2003}. 
1471: \bibitem{Belle03} Belle Collaboration, P. Krokovny {\em et al.}, 
1472:              \Journal{\PRL}{91}{262002}{2003}; 
1473:                   Y. Mikami {\em et al.}, 
1474:              \Journal{\PRL}{92}{012002}{2004}. 
1475: \bibitem{Belle04} Belle Collaboration, K. Abe {\em et al.}, 
1476:              \Journal{\PRD}{69}{112002}{2004}. 
1477: \bibitem{CDF_D006} CDF Collaboration, CDF note 7938 (2005);
1478: 			 D0 Collaboration, D0notes 5026-CONF, 5027-CONF (2006);
1479: 			 I. Kravchenko, hep-ex/0605076.
1480: \bibitem{Barnes03} T. Barnes, F. E. Close, and H. J. Lipkin, 
1481:              \Journal{\PRD}{68}{054006}{2003}. 
1482: \bibitem{Godfrey85} S. Godfrey and N. Isgur, 
1483:              \Journal{\PRD}{32}{189}{1985}. 
1484: \bibitem{Godfrey91} S. Godfrey and R. Kokoski, 
1485:              \Journal{\PRD}{43}{1679}{1991}. 
1486: \bibitem{Ebert98} D. Ebert, V.O. Galkin, and R.N. Faustov,
1487:              \Journal{\PRD}{57}{5663}{1998}.
1488: \bibitem{Mukherjee93} S. N. Mukherjee, R. Nag, S. Sanyal, 
1489:                       T Morii, J. Morishita, and M. Tsuge, 
1490:              \Journal{\PR}{231}{201}{1993}
1491: \bibitem{Bardeen03} W. A. Bardeen, E. J. Eichten, and C. T. Hill, 
1492:              \Journal{\PRD}{68}{054024}{2003}. 
1493: \bibitem{Bardeen94} W. A. Bardeen and C. T. Hill, 
1494:              \Journal{\PRD}{49}{409}{1994}; 
1495:                   M. A. Nowak, M. Rho, and I. Zahed, 
1496:              \Journal{\PRD}{48}{4370}{1993}; 
1497:                   D. Ebert, T. Feldmann, and H. Reinhardt, 
1498:              \Journal{\PLB}{388}{154}{1996}; 
1499:                   A. Deandrea, N. Di Bartolomeo, R. Gatto, G. Nardulli, 
1500:                   and A. D. Polosa, 
1501:              \Journal{\PRD}{58}{034004}{1998}; 
1502:                   A. Hiorth and J. O. Eeg, 
1503:              \Journal{\PRD}{66}{074001}{2002}; 
1504:                   M. Harada, M. Rho, and C. Sasaki, 
1505:              \Journal{\PRD}{70}{074002}{2004}. 
1506: \bibitem{Hayashigaki04} A. Hayashigaki and K. Terasaki, 
1507:              hep-ph/0411285. 
1508: \bibitem{Szczepaniak03} A. P. Szczepaniak, 
1509:              \Journal{\PLB}{567}{23}{2003}. 
1510: \bibitem{Browder04} T. E. Browder, S. Pakvasa, and A. A. Petrov,
1511: 			 \Journal{\PLB}{578}{365}{2004}.
1512: \bibitem{Terasaki05} K. Terasaki, 
1513:              \Journal{\PRD}{68}{011501(R)}{2003}; 
1514:                   K. Terasaki and B. H. J. McKellar, 
1515:              hep-ph/0501188. 
1516: \bibitem{Dmitrasinovic05} V. Dmitra$\check{\rm s}$inovi\'c, 
1517:              \Journal{\PRL}{94}{162002}{2005}. 
1518: \bibitem{Bali03} G. S. Bali, 
1519:              \Journal{\PRD}{68}{071501(R)}{2003}; 
1520:                   UKQCD Collaboration, A. Dougall, R. D. Kenway, 
1521:                   C. M. Maynard, and C. McNeile, 
1522:              \Journal{\PLB}{569}{41}{2003}.
1523: \bibitem{Habe87} C. Habe, K. Iwata, M. Hirano, T. Murota, and D. Tsuruda,
1524: 			 \Journal{\PTP}{77}{917}{1987}.
1525: \bibitem{Gara89} A. Gara, B. Durand, and L. Durand,
1526: 			 \Journal{\PRD}{40}{843}{1989}.
1527: \bibitem{Pierro01} M. Di Piero and E. Eichten,
1528: 			 \Journal{\PRD}{64}{114004}{2001}; see also J.L. Goity and W. Roberts,
1529: 			 \Journal{\PRD}{60}{034001}{1999} and
1530:     J. Zeng, J. W. Van Orden, and W. Roberts, \Journal{\PRD}{52}{5229}{1993}.
1531: \bibitem{Beveren06} E. van Beveren and G. Rupp,
1532: 			 \Journal{\PRL}{97}{202001}{2006}; Mod. Phys. Lett. A19, 1949 (2004);
1533: 			 \Journal{\EPJ}{32}{493}{2004};
1534: 			 \Journal{\PRL}{91}{012003}{2003}.
1535: \bibitem{Swanson06} E. S. Swanson, \Journal{\PR}{429}{243}{2006}, hep-ph/0601110.
1536: \bibitem{Matsuki97} T. Matsuki and T. Morii, 
1537:              \Journal{\PRD}{56}{5646}{1997}; See also T. Matsuki, Mod. Phys.
1538:              Lett. A {\bf 11}, 257 (1996), T. Matsuki and T. Morii, Aust.
1539:              J. Phys. {\bf 50}, 163 (1997), T. Matsuki, T. Morii, and K. Seo,
1540:              Trends in Applied Spectroscopy, {\bf 4}, 127 (2002);
1541: T. Matsuki, T. Morii, and K. Sudoh, AIP Conference Proceedings, Vol. 814, p.533
1542: (2005), "XI International Conference on Hadron Spectroscopy" (Hadron05), held
1543: at Rio De Janeiro, Brazil 21-26 Aug. 2005, hep-ph/0510269 (2005).
1544: \bibitem{Matsuki05} T. Matsuki, K. Mawatari, T. Morii, and K. Sudoh,
1545:              \Journal{\PLB}{606}{329}{2005}.
1546: \bibitem{Fermi49} E. Fermi and C. N. Yang, 
1547:              \Journal{\PRO}{34}{553}{1949}. 
1548: \bibitem{Bethe55} For example, H. Bethe and E. E. Salpeter, 
1549:              {\em Handbuch der Physik}, edited by H. von S. Fl\"uge, 
1550:              Band XXXV ATOMS I (Springer, Berlin, 1955), p. 256. 
1551: \bibitem{Miyazawa92} H. Miyazawa and K. Tanaka,
1552:              \Journal{\PTP}{87}{1457}{1992}.
1553: \bibitem{Foldy50} L. L. Foldy and S. A. Wouthuysen, 
1554:              \Journal{\PRO}{78}{29}{1950}; 
1555:              S. Tani, \Journal{\PTP}{6}{267}{1951}. 
1556: \bibitem{MS97} T. Matsuki and K. Seo, Bulletin of Gifu City Women's College
1557:             {\bf 46}, 7 (1997).
1558: \bibitem{Bloch} C. Bloch, Nucl. Phys. {\bf 6}, 329 (1958); A.
1559:              Messiah, {\it Quantum Mechanics} (North-Holland, Amsterdam),
1560:              p.718 (1962).
1561: \bibitem{Morishita88} J. Morishita, M. Kawaguchi, and T. Morii, 
1562:              \Journal{\PLB}{185}{139}{1987}; \Journal{\PRD}{37}{159}{1988}. 
1563: \bibitem{JJ} M. E. Rose, {\it Relativistic Electron Theory}, John Wiley
1564: 	\& Sons, 1961; 
1565:              J. J. Sakurai, 
1566:              {\it Advanced Quantum Mechanics}, Addison-Wesley, 1967.
1567: \bibitem{Matsuki04} T. Matsuki, K. Mawatari, T. Morii, and K. Sudoh,
1568:              hep-ph/0408326.
1569: \bibitem{Matsuki06} T. Matsuki, T. Morii, and K. Sudoh, a talk given
1570: by T. Matsuki at the
1571: IVth International Conference on Quarks and Nuclear Physics (QNP06), held at
1572: Madrid, Spain 5-10 June 2006, hep-ph/0610186.
1573: \bibitem{Eichten93} E. Eichten, C. Hill, and C. Quigg,
1574:              \Journal{\PRL}{71}{4116}{1993}.
1575: \bibitem{Orsland99} A. Hiorth \"Orsland and H. H\"ogaasen,
1576:              Eur. Phys. J. C9, 503 (1999).
1577: \bibitem{Politzer} H.D. Politzer,
1578:              \Journal{\NPB}{117}{397}{1976}.
1579: \bibitem{Matsuki80} T. Matsuki and N. Yamamoto,
1580:              \Journal{\PRD}{22}{2433}{1980}.
1581: \bibitem{Kaburagi} M. Kaburagi, M. Kawaguchi, T. Morii, T. Kitazoe, and
1582: 			 J. Morishita, \Journal{\ZPC}{9}{213}{1981}.
1583: \end{thebibliography}
1584: 
1585: \end{document}
1586: