1: \documentclass[12pt,prd,showpacs,tightenlines,nofootinbib]{revtex4}
2: \usepackage{bm}
3: \usepackage{graphics}
4: \usepackage{rotating}
5: \usepackage{epsfig}
6: \begin{document}
7: \title{\begin{flushright}{\rm\normalsize HU-EP-05/73}\end{flushright}
8: Masses and electroweak properties of light mesons in
9: the relativistic quark model}
10: \author{D. Ebert}
11: \affiliation{Institut f\"ur Physik, Humboldt--Universit\"at zu Berlin,
12: Newtonstr. 15, D-12489 Berlin, Germany}
13: \author{R. N. Faustov}
14: \author{V. O. Galkin}
15: \affiliation{Institut f\"ur Physik, Humboldt--Universit\"at zu Berlin,
16: Newtonstr. 15, D-12489 Berlin, Germany}
17: \affiliation{Dorodnicyn Computing Centre, Russian Academy of Sciences,
18: Vavilov Str. 40, 119991 Moscow, Russia}
19:
20: \begin{abstract}
21: The masses, pseudoscalar and vector weak decay constants
22: and electromagnetic form factors of light $S$-wave mesons
23: are studied in the framework of
24: the relativistic quark model based on the quasipotential approach. We
25: use the same model assumptions and parameters as in our
26: previous investigations of heavy meson and baryon properties. The
27: masses and wave functions of the ground state and radially excited
28: $\pi$, $\rho$, $K$, $K^*$ and $\phi$ mesons, obtained by solving
29: numerically the relativistic Schr\"odinger-like equation with the
30: complete relativistic $q\bar q$ potential including both
31: spin-independent and spin-dependent terms, are presented.
32: Novel relativistic expressions
33: for the weak decay constants of the pseudoscalar and vector mesons are
34: derived. It is shown that the intermediate negative-energy quark
35: states give significant contributions which essentially decrease the decay
36: constants bringing them in agreement with experimental data.
37: The electromagnetic form factors of the pion, charged and neutral kaon
38: are calculated in a broad range of the space-like momentum
39: transfer. The corresponding charge radii are determined. All results
40: agree well with available experimental data.
41: \end{abstract}
42:
43: \pacs{14.40.Aq, 13.40.Gp, 12.39.Ki}
44:
45: \maketitle
46:
47:
48: \section{Introduction}
49: \label{sec:int}
50: The theoretical investigation of the properties of light mesons such
51: as $\pi$, $\rho$, $K$, $K^*$ and $\phi$ is a longstanding problem
52: which plays an important role in understanding the low-energy
53: QCD. The description of these mesons within the constituent quark
54: model presents additional difficulties compared to heavy-light mesons
55: and heavy quarkonia. In fact, due to the highly relativistic dynamics
56: of light quarks, the $v/c$ and $1/m_q$ expansions are completely
57: inapplicable in
58: the case of light mesons and the QCD coupling constant $\alpha_s$ at
59: the related scale $\mu$ is rather
60: large. Moreover, the behaviour of $\alpha_s(\mu^2)$ in
61: the infrared region is unknown and thus model dependent (exhibiting, e.g.,
62: freezing behaviour, etc.). The pseudoscalar mesons $\pi$ and $K$
63: produce a special problem,
64: since their small masses originate from their Goldstone nature caused
65: by the broken chiral symmetry.
66: Therefore the reliable description of light mesons requires the completely
67: relativistic approach. It is well known that in the relativistic
68: studies an important role is played by the Lorentz properties of the
69: confining quark-antiquark interaction. The comparison of theoretical
70: predictions with experimental data can provide a valuable information on
71: the form of the confining potential. Such information is of great
72: practical interest, because at present it is not possible to obtain the
73: relativistic $q\bar q$ potential in the whole range of distances from
74: the basic
75: principles of QCD.\footnote{Recent calculations of the
76: nonperturbative $q\bar q$ potential in continuum Yang-Mills theory
77: in Coulomb gauge can be found in Ref.~\cite{fr}.}
78: Most of the main characteristics of light mesons
79: are formed in the infrared (nonperturbative) region, thus providing
80: an important insight in the low-energy properties of $q\bar
81: q$ interaction. Thus investigation of both static (e.g., masses and
82: decay constants) and dynamic (e.g., electroweak decay form factors)
83: properties is of a significant importance.
84:
85: Many different theoretical approaches have been used for studying
86: light mesons, which are based on the
87: relativized quark model \cite{gi}, the
88: Dyson-Schwinger and Bethe-Salpeter equations \cite{mr,k}, chiral quark
89: models with spontaneous symmetry breaking (e.g. the Nambu-Jona-Lasinio
90: model) \cite{erv}, the relativistic Hamilton dynamics \cite{kt,hjd},
91: the finite-energy \cite{kp} and light-cone \cite{bkm} sum rules and
92: lattice QCD \cite{ak}.
93: Here we consider the possibility of investigating light mesons on the
94: basis of the three-dimensional relativistic wave equation with the QCD
95: motivated potential. Our relativistic quark model was originally
96: constructed for the investigation of hadrons with heavy quarks. It was
97: successfully applied for the calculation of their masses and various electroweak
98: decays \cite{efg,egf,fgm,dhb,hbm}. In these studies the
99: heavy quark expansion has been used to simplify calculations. We determined
100: all parameters of our model from few experimental observables (some
101: masses and decay rates) and keep them fixed in all our subsequent
102: calculations, thus ensuring its universality.
103: While describing the properties of heavy-light
104: mesons \cite{egf}, we treated the light quarks in a completely relativistic
105: way. Recently this approach was applied for calculating the masses of light
106: mesons \cite{lmm} and light diquarks inside the heavy baryons
107: \cite{hbm}. Due to the phenomenological character of our model we
108: cannot reveal
109: the origin of the chiral symmetry breaking and thus the model cannot
110: describe the chiral limit and the Goldstone nature of the pion. We
111: consider the pion as the purely bound state of the quark and antiquark
112: with fixed constituent masses.
113:
114: In this paper we extend our previous studies of light mesons and
115: describe their electroweak properties such as the weak decay
116: constants and electromagnetic form factors.
117: The investigation of decay constants and form factors is an
118: important issue since it provides a very sensitive test of the light
119: meson wave functions and, thus, of the quark dynamics in a meson. It
120: requires the completely relativistic
121: consideration of the corresponding decay processes including account
122: for the
123: relativistic transformation of the meson wave functions. The
124: comparison with the available large set of experimental data tests the
125: model
126: predictions in a broad momentum range and helps to discriminate
127: between different model assumptions.
128:
129:
130:
131: The paper is organized as follows. In Sec.~\ref{sec:rqm} we briefly
132: describe our relativistic quark model, formulate our main assumptions
133: and give the values of parameters. Then in Sec.~\ref{sec:lmm} we
134: present our results for the light meson masses \cite{lmm},
135: for selfconsistency. There the procedure of constructing
136: the completely relativistic local potential of the light quark
137: interaction in a meson is described. The obtained
138: potential is applied for calculating the light $S$-wave meson
139: masses and wave functions. In Sec.~\ref{sec:dc} the novel relativistic
140: expressions for the weak decay constants of pseudoscalar and vector
141: meson are derived. Special attention is paid to including all possible
142: intermediate quark states. It is argued that the negative-energy
143: contributions play an essential role. The calculated decay
144: constants are compared with other predictions and experimental data.
145: The electromagnetic form factors of pseudoscalar mesons are studied in
146: Sec.~\ref{em}. The relativistic expressions for these form factors are
147: obtained which take into account the contributions of negative-energy quark
148: states and relativistic transformations of the meson wave functions
149: from the rest frame to the moving one. The calculated form factors are
150: plotted in comparison with experimental data. The charged radii of the
151: pion, charged and neutral kaon are also determined. Our conclusions are
152: given in Sec.~\ref{sec:concl}.
153:
154: \section{Relativistic quark model}
155: \label{sec:rqm}
156:
157: In the quasipotential approach a meson is described by the wave
158: function of the bound quark-antiquark state, which satisfies the
159: quasipotential equation of the Schr\"odinger type~\cite{efg}
160: \begin{equation}
161: \label{quas}
162: {\left(\frac{b^2(M)}{2\mu_{R}}-\frac{{\bf
163: p}^2}{2\mu_{R}}\right)\Psi_{M}({\bf p})} =\int\frac{d^3 q}{(2\pi)^3}
164: V({\bf p,q};M)\Psi_{M}({\bf q}),
165: \end{equation}
166: where the relativistic reduced mass is
167: \begin{equation}
168: \mu_{R}=\frac{E_1E_2}{E_1+E_2}=\frac{M^4-(m^2_1-m^2_2)^2}{4M^3},
169: \end{equation}
170: and $E_1$, $E_2$ are given by
171: \begin{equation}
172: \label{ee}
173: E_1=\frac{M^2-m_2^2+m_1^2}{2M}, \quad E_2=\frac{M^2-m_1^2+m_2^2}{2M}.
174: \end{equation}
175: Here $M=E_1+E_2$ is the meson mass, $m_{1,2}$ are the quark masses,
176: and ${\bf p}$ is their relative momentum.
177: In the center-of-mass system the relative momentum squared on mass shell
178: reads
179: \begin{equation}
180: {b^2(M) }
181: =\frac{[M^2-(m_1+m_2)^2][M^2-(m_1-m_2)^2]}{4M^2}.
182: \end{equation}
183:
184: The kernel
185: $V({\bf p,q};M)$ in Eq.~(\ref{quas}) is the quasipotential operator of
186: the quark-antiquark interaction. It is constructed with the help of the
187: off-mass-shell scattering amplitude, projected onto the positive
188: energy states.
189: Constructing the quasipotential of the quark-antiquark interaction,
190: we have assumed that the effective
191: interaction is the sum of the usual one-gluon exchange term with the mixture
192: of long-range vector and scalar linear confining potentials, where
193: the vector confining potential
194: contains the Pauli interaction. The quasipotential is then defined by
195: \footnote{In our notation, where strong annihilation processes are neglected,
196: antiparticles are described by usual spinors taking into account the
197: proper quark charges.}
198: \begin{equation}
199: \label{qpot}
200: V({\bf p,q};M)=\bar{u}_1(p)\bar{u}_2(-p){\mathcal V}({\bf p}, {\bf
201: q};M)u_1(q)u_2(-q),
202: \end{equation}
203: with
204: $${\mathcal V}({\bf p},{\bf q};M)
205: \equiv{\mathcal V}({\bf p}-{\bf q})=\frac{4}{3}\alpha_sD_{ \mu\nu}({\bf
206: k})\gamma_1^{\mu}\gamma_2^{\nu}
207: +V^V_{\rm conf}({\bf k})\Gamma_1^{\mu}
208: \Gamma_{2;\mu}+V^S_{\rm conf}({\bf k}),$$
209: where $\alpha_s$ is the QCD coupling constant, $D_{\mu\nu}$ is the
210: gluon propagator in the Coulomb gauge
211: \begin{equation}
212: D^{00}({\bf k})=-\frac{4\pi}{{\bf k}^2}, \quad D^{ij}({\bf k})=
213: -\frac{4\pi}{k^2}\left(\delta^{ij}-\frac{k^ik^j}{{\bf k}^2}\right),
214: \quad D^{0i}=D^{i0}=0,
215: \end{equation}
216: and ${\bf k=p-q}$; $\gamma_{\mu}$ and $u(p)$ are
217: the Dirac matrices and spinors
218: \begin{equation}
219: \label{spinor}
220: u^\lambda({p})=\sqrt{\frac{\epsilon(p)+m}{2\epsilon(p)}}
221: \left(
222: \begin{array}{c}1\cr {\displaystyle\frac{\bm{\sigma}
223: {\bf p}}{\epsilon(p)+m}}
224: \end{array}\right)\chi^\lambda,
225: \end{equation}
226: with $\epsilon(p)=\sqrt{p^2+m^2}$.
227: The effective long-range vector vertex is
228: given by
229: \begin{equation}
230: \label{kappa}
231: \Gamma_{\mu}({\bf k})=\gamma_{\mu}+
232: \frac{i\kappa}{2m}\sigma_{\mu\nu}k^{\nu},
233: \end{equation}
234: where $\kappa$ is the Pauli interaction constant characterizing the
235: anomalous chromomagnetic moment of quarks. Vector and
236: scalar confining potentials in the nonrelativistic limit reduce to
237: \begin{eqnarray}
238: \label{vlin}
239: V^V_{\rm conf}(r)&=&(1-\varepsilon)(Ar+B),\nonumber\\
240: V^S_{\rm conf}(r)& =&\varepsilon (Ar+B),
241: \end{eqnarray}
242: reproducing
243: \begin{equation}
244: \label{nr}
245: V_{\rm conf}(r)=V^S_{\rm conf}(r)+V^V_{\rm conf}(r)=Ar+B,
246: \end{equation}
247: where $\varepsilon$ is the mixing coefficient.
248:
249: All the model parameters have the same values as in our previous
250: papers \cite{egf,efg}.
251: The light constituent quark masses $m_u=m_d=0.33$ GeV, $m_s=0.5$ GeV and
252: the parameters of the linear potential $A=0.18$ GeV$^2$ and $B=-0.3$ GeV
253: have the usual values of quark models. The value of the mixing
254: coefficient of vector and scalar confining potentials $\varepsilon=-1$
255: has been determined from the consideration of charmonium radiative
256: decays \cite{efg}.
257: Finally, the universal Pauli interaction constant $\kappa=-1$ has been
258: fixed from the analysis of the fine splitting of heavy quarkonia ${
259: }^3P_J$- states \cite{efg}. In the literature it is widely discussed
260: the 't~Hooft-like interaction between quarks induced by instantons \cite{dk}.
261: This interaction can be partly described by introducing the quark
262: anomalous chromomagnetic moment having an approximate value
263: $\kappa=-0.744$ (Diakonov \cite{dk}). This value is of the same
264: sign and order of magnitude
265: as the Pauli constant $\kappa=-1$ in our model. Thus the Pauli term
266: incorporates at least part of the instanton contribution to the $q\bar q$
267: interaction.\footnote{As is well-known, the instanton-induced 't~Hooft
268: interaction term breaks the axial $U_A(1)$-symmetry, the violation
269: of which is needed for describing the $\eta-\eta'$ mass splitting. We
270: do not consider this issue here.}
271:
272: \section{Light meson masses}
273: \label{sec:lmm}
274: The quasipotential (\ref{qpot}) can be used for arbitrary quark
275: masses. The substitution
276: of the Dirac spinors (\ref{spinor}) into (\ref{qpot}) results in an extremely
277: nonlocal potential in the configuration space. Clearly, it is very hard to
278: deal with such potentials without any additional transformations.
279: In oder to simplify the relativistic $q\bar q$ potential, we make the
280: following replacement in the Dirac spinors:
281: \begin{equation}
282: \label{eq:sub}
283: \epsilon_{1,2}(p)=\sqrt{m_{1,2}^2+{\bf p}^2} \to E_{1,2}
284: \end{equation}
285: (see the discussion of this point in \cite{egf,lmm}). This substitution
286: makes the Fourier transformation of the potential (\ref{qpot}) local.
287: We also limit our consideration only to the $S$-wave
288: states, which further simplifies our analysis, since all terms
289: proportional to ${\bf L}^2$ vanish as well as the spin-orbit
290: ones. Thus we neglect the mixing of states with different values of
291: $L$. Calculating the potential, we keep only operators quadratic
292: in the relative momentum acting on $V_{\rm Coul}$, $V^{V,S}_{\rm conf}$ and
293: replace ${\bf p}^2\to E_{1,2}^2-m_{1,2}^2$ in higher order operators
294: in accord with Eq.~(\ref{eq:sub}) preserving the symmetry under the
295: $(1\leftrightarrow 2)$ exchange.
296:
297: The substitution (\ref{eq:sub})
298: works well for the confining part of the potential. However, it leads to
299: a fictitious singularity $\delta^3({\bf r})$ at the origin arising from the
300: one-gluon exchange part ($\Delta V_{\rm
301: Coul}(r)$), which is absent in the initial potential.
302: Note that this singularity is not important if it is treated
303: perturbatively. Since we are not using the expansion in $v/c$ and
304: are solving the quasipotential equation with the
305: complete relativistic potential, an additional analysis is
306: required. Such singular contributions emerge from the following terms
307: \begin{eqnarray}
308: \label{eq:st}
309: && \frac{{\bf k}^2}{[\epsilon_i(q)(\epsilon_i(q)+m_i)
310: \epsilon_i(p)(\epsilon_i(p)+m_i)]^{1/2}}V_{\rm Coul}({\bf k}^2) ,\cr
311: &&\frac{{\bf k}^2}{[\epsilon_1(q)\epsilon_1(p)
312: \epsilon_2(q)\epsilon_2(p)]^{1/2}}V_{\rm Coul}({\bf k}^2),
313: \end{eqnarray}
314: if we simply apply the replacement (\ref{eq:sub}). However, the Fourier
315: transforms of expressions (\ref{eq:st}) are less singular at $r\to
316: 0$. To avoid such fictitious singularities we note that if the binding effects
317: are taken into account, it is necessary to replace $\epsilon_{1,2}
318: \to E_{1,2}-\eta_{1,2}V$, where $V$ is the quark interaction potential
319: and $\eta_{1,2}=m_{2,1}/(m_1+m_2)$. At small
320: distances $r\to 0$, the Coulomb singularity in $V$ dominates
321: and affords the correct asymptotic behaviour. Therefore, we replace
322: $\epsilon_{1,2} \to E_{1,2}-\eta_{1,2}V_{\rm Coul}$ in the Fourier
323: transforms of terms (\ref{eq:st}) (cf. \cite{bs}). We used
324: the similar regularization of singularities in the analysis of
325: heavy-light meson spectra \cite{egf}. Finally, we ignore the annihilation
326: terms in the quark potential since they contribute only in the
327: isoscalar channels and are suppressed in the $s\bar s$ vector channel
328: \cite{gi}.
329:
330:
331: The resulting $q\bar q$ potential then reads
332: \begin{equation}
333: \label{eq:v}
334: V(r)= V_{\rm SI}(r)+ V_{\rm SD}(r),
335: \end{equation}
336: where the spin-independent potential for $S$-states (${\bf
337: L}^2=0$) has the form
338: \begin{eqnarray}
339: \label{eq:vsi}
340: V_{\rm SI}(r)&=&V_{\rm Coul}(r)+V_{\rm conf}(r)+
341: \frac{(E_1^2-m_1^2+E_2^2-m_2^2)^2}{4(E_1+m_1)(E_2+m_2)}\Biggl\{
342: \frac1{E_1E_2}V_{\rm Coul}(r)\cr
343: && +\frac1{m_1m_2}\Biggl(1+(1+\kappa)\Biggl[(1+\kappa)\frac{(E_1+m_1)(E_2+m_2)}
344: {E_1E_2}\cr
345: &&-\left(\frac{E_1+m_1}{E_1}+\frac{E_1+m_2}{E_2}\right)\Biggr]\Biggr)
346: V^V_{\rm conf}(r)
347: +\frac1{m_1m_2}V^S_{\rm conf}(r)\Biggr\}\cr
348: &&+\frac14\left(\frac1{E_1(E_1+m_1)}\Delta
349: \tilde V^{(1)}_{\rm Coul}(r)+\frac1{E_2(E_2+m_2)}\Delta
350: \tilde V^{(2)}_{\rm Coul}(r)\right)\cr
351: &&-\frac14\left[\frac1{m_1(E_1+m_1)}+\frac1{m_2(E_2+m_2)}-(1+\kappa)
352: \left(\frac1{E_1m_1}+\frac1{E_2m_2}\right)\right]\Delta V^V_{\rm
353: conf}(r)\cr
354: &&+\frac{(E_1^2-m_1^2+E_2^2-m_2^2)}{8m_1m_2(E_1+m_1)(E_2+m_2)}
355: \Delta V^S_{\rm conf}(r),
356: \end{eqnarray}
357: and the spin-dependent potential is given by
358: \begin{eqnarray}
359: \label{eq:vsd}
360: V_{\rm SD}(r)&=&\frac2{3E_1E_2}\Biggl[\Delta \bar V_{\rm Coul}(r)
361: +\left(\frac{E_1-m_1}{2m_1}-(1+\kappa)\frac{E_1+m_1}{2m_1}\right)\cr
362: &&\qquad\quad\times
363: \left(\frac{E_2-m_2}{2m_2}-(1+\kappa)\frac{E_2+m_2}{2m_2}\right)
364: \Delta V^V_{\rm conf}(r)\Biggr]{\bf S}_1{\bf S}_2,
365: \end{eqnarray}
366: with
367: \begin{eqnarray}
368: \label{eq:tv}
369: V_{\rm Coul}(r)&=&-\frac43\frac{\alpha_s}{r},\cr
370: \tilde V^{(i)}_{\rm Coul}(r)&=&V_{\rm Coul}(r)\frac1{\displaystyle\left(1+
371: \eta_i\frac43\frac{\alpha_s}{E_i}\frac1{r}\right)\left(1+
372: \eta_i\frac43\frac{\alpha_s}{E_i+m_i}\frac1{r}\right)},\qquad (i=1,2),\cr
373: \bar V_{\rm Coul}(r)&=&V_{\rm Coul}(r)\frac1{\displaystyle\left(1+
374: \eta_1\frac43\frac{\alpha_s}{E_1}\frac1{r}\right)\left(1+
375: \eta_2\frac43\frac{\alpha_s}{E_2}\frac1{r}\right)},
376: \qquad \eta_{1,2}=\frac{m_{2,1}}{m_1+m_2}.
377: \end{eqnarray}
378: Here we put $\alpha_s\equiv\alpha_s(\mu_{12}^2)$ with $\mu_{12}=2m_1
379: m_2/(m_1+m_2)$. We adopt for $\alpha_s(\mu^2)$ the
380: simplest model with freezing \cite{bvb}, namely
381: \begin{equation}
382: \label{eq:alpha}
383: \alpha_s(\mu^2)=\frac{4\pi}{\displaystyle\beta_0
384: \ln\frac{\mu^2+M_B^2}{\Lambda^2}}, \qquad \beta_0=11-\frac23n_f,
385: \end{equation}
386: where the background mass is $M_B=2.24\sqrt{A}=0.95$~GeV \cite{bvb}, and
387: $\Lambda=413$~MeV was fixed from fitting the $\rho$
388: mass.~\footnote{The definition (\ref{eq:alpha}) of $\alpha_s$ can be
389: smoothly matched with the $\alpha_s$ used for heavy quarkonia
390: \cite{efg} at the scale about $m_c$.} We put the
391: number of flavours $n_f=2$ for $\pi$,
392: $\rho$, $K$, $K^*$ and $n_f=3$ for $\phi$. As a result we obtain
393: $\alpha_s(\mu_{ud}^2)=0.730$, $\alpha_s(\mu_{us}^2)=0.711$ and
394: $\alpha_s(\mu_{ss}^2)=0.731$.
395:
396:
397:
398: \begin{table}
399: \caption{Masses of light $S$-wave mesons (in MeV)}
400: \label{tab:mass}
401: \begin{ruledtabular}
402: \begin{tabular}{ccccccc}
403: Meson& State &
404: \multicolumn{4}{l}{\underline{\hspace{3.5cm}Theory\hspace{3.5cm}}}
405: \hspace{-3.6cm}
406: & Experiment \\
407: & $n^{2S+1}L_J$&this work& \cite{gi}& \cite{mr}& \cite{k}&
408: PDG \cite{pdg}\\
409: \hline
410: $\pi$& $1^1S_0$& 154 &150& 138& 140 &139.57\\
411: $\rho$ & $1^3S_1$ & 776$^\dag$ & 770& 742& 785 & 775.8(5)\\
412: $\pi'$& $2^1S_0$& 1292 &1300& &1331 &1300(100)\\
413: $\rho'$ & $2^3S_1$ & 1486&1450& &1420 & 1465(25)\\
414: $\pi''$&$3^1S_0$& 1788 &1880& &1826& 1812(14)\\
415: $\rho''$ & $3^3S_1$ & 1921 & 2000 & & 1472& \\
416: $K$ & $1^1S_0$ & 482&470& 497& 506& 493.677(16)\\
417: $K^*$&$1^3S_1$ & 897&900& 936& 890& 891.66(26)\\
418: $K'$ & $2^1S_0$ & 1538&1450& &1470& \\
419: ${K^*}'$&$2^3S_1$ & 1675&1580& &1550& 1717(27)\\
420: $K''$ & $3^1S_0$ & 2065 & 2020& &1965& \\
421: ${K^*}''$&$3^3S_1$ &2156 & 2110& & 1588& \\
422: $\phi$& $1^3S_1$& 1038&1020& 1072&990 & 1019.46(2)\\
423: $\phi'$& $2^3S_1$& 1698&1690& &1472 & 1680(20)
424: \end{tabular}
425: \end{ruledtabular}
426: \flushleft${}^\dag$ fitted value
427: \end{table}
428:
429: The quasipotential equation (\ref{quas}) is solved numerically for the
430: complete relativistic potential (\ref{eq:v}) which depends on the
431: meson mass in a complicated highly nonlinear way. The obtained meson
432: masses are presented in Table~\ref{tab:mass} in comparison with
433: experimental data \cite{pdg} and other theoretical results \cite{gi,mr,k}.
434: This comparison exhibits a reasonably good overall agreement of our
435: predictions with experimental mass values.
436: Our results are also consistent with mass formulas derived using
437: the finite-energy sum rules in QCD \cite{kp} and with predictions of
438: lattice QCD \cite{ak}.
439: We consider such agreement to be quite
440: successful, since in evaluating the meson masses we had at our disposal
441: only one adjustable parameter $\Lambda$, which was fixed from fitting
442: the $\rho$ meson mass. All other parameters are kept the same as in
443: our previous papers \cite{egf,efg}. The obtained wave functions of the
444: light mesons are used for the calculation of their decay constants and
445: electromagnetic form factors in the following sections.
446:
447:
448: \section{Decay constants}
449: \label{sec:dc}
450: The decay constants $f_P$ and $f_V$ of the pseudoscalar ($P$) and
451: vector ($V$) mesons parameterize the matrix elements of the weak
452: current $J^W_\mu=\bar q_1{\cal J}^W_\mu q_2=\bar q_1\gamma_\mu(1-\gamma_5)q_2$
453: between the corresponding
454: meson and the vacuum. They are defined by
455: \begin{eqnarray}
456: \label{eq:dc}
457: \left<0|\bar q_1 \gamma^\mu\gamma_5 q_2|P({\bf K})\right>&=& i f_P
458: K^\mu,\\
459: \left<0|\bar q_1 \gamma^\mu q_2|V({\bf K},\varepsilon)\right>&=& f_V
460: M_V \varepsilon^\mu,
461: \end{eqnarray}
462: where ${\bf K}$ is the meson momentum, $\varepsilon^\mu$ and $M_V$ are
463: the polarization vector and mass of the vector meson. This matrix
464: element can be expressed through the two-particle Bethe-Salpeter wave
465: function in the quark loop integral (see Fig.~\ref{fig:diag})
466: \begin{equation}
467: \label{eq:bs}
468: \left<0| J^W_\mu |M({\bf K})\right>=\int\frac{d^4
469: p}{(2\pi)^4}
470: {\rm Tr}\left\{\gamma_\mu(1-\gamma_5)\Psi(M,p)\right\},
471: \end{equation}
472: where the trace is taken over spin indices. Integration
473: over $p^0$ in Eq.~(\ref{eq:bs}) allows one to pass to the single-time
474: wave function in the meson rest frame
475: \begin{equation}
476: \label{eq:stw}
477: \Psi(M,{\bf p})=\int\frac{dp^0}{2\pi}\Psi(M,p).
478: \end{equation}
479: This wave function contains both positive- and negative-energy
480: quark states. Since in the quasipotential approach we use the single-time wave
481: function $\Psi_{M\, {\bf K}}({\bf p})$ projected onto the positive-energy states it
482: is necessary to
483: include additional terms which account for the contributions
484: of negative-energy intermediate states. The weak annihilation amplitude
485: (\ref{eq:bs}) is schematically presented in the left hand side of
486: Fig.~\ref{fig:diag}. The first diagram on the right hand side
487: corresponds to the simple replacing of the single-time wave function
488: (\ref{eq:stw}) $\Psi(M,{\bf p})$ by the quasipotential one
489: $\Psi_{M\, {\bf K}}({\bf p})$.\footnote{The contributions with the exchange by the
490: effective interaction potential ${\mathcal V}$ which contain only
491: positive-energy intermediate states are automatically accounted for
492: by the wave function itself.} The second and third diagrams account for
493: negative-energy contributions to the first and second quark propagators,
494: respectively. The last diagram corresponds to negative-energy
495: contributions from both quark propagators.
496: \begin{figure}%[htb]
497: \centering
498: \includegraphics[width=16cm]{diag1.eps}
499: \caption{Weak annihilation diagram of the light meson. Solid and bold
500: lines denote the positive- and negative-energy part of the quark propagator,
501: respectively. Dashed lines represent the interaction operator
502: ${\mathcal V}$.}
503: \label{fig:diag}
504: \end{figure}
505:
506: Thus in the quasipotential approach
507: this decay amplitude has the form
508: \begin{eqnarray}
509: \label{eq:qpd}
510: \left<0|J^W_\mu |M({\bf K})\right>&=&\sqrt{2M}\Biggl\{\int\frac{d^3
511: p}{(2\pi)^3} \bar u_1(p_1){\cal J}^W_\mu u_2(p_2)
512: \Psi_{M\, {\bf K}}({\bf
513: p})+\Biggl[\int\frac{d^3 p d^3 p'}{(2\pi)^6}\bar u_1(p_1)\Gamma_1\cr
514: &&\!\!\!\!\times
515: \frac{\Lambda^{(-)}_1(p_1')\gamma^0{\cal J}^W_\mu\Lambda^{(+)}_2(p_2')\gamma^0}
516: {M+\epsilon_1(p')-\epsilon_2(p')}\Gamma_2 u_2(p_2)\tilde V(p-p')
517: \Psi_{M\, {\bf K}}({\bf
518: p})
519: +(1\leftrightarrow 2)\Biggr]\cr
520: &&\!\!\!\!+\int\frac{d^3 p d^3 p'}{(2\pi)^6}\bar u_1(p_1)\Gamma_1
521: \frac{\Lambda^{(-)}_1(p_1')\gamma^0{\cal J}^W_\mu\Lambda^{(-)}_2(p_2')\gamma^0}
522: {M+\epsilon_1(p')+\epsilon_2(p')}\Gamma_2 u_2(p_2)\tilde V(p-p')
523: \Psi_{M\, {\bf K}}({\bf
524: p})\Biggr\},\cr&&\!\!\!
525: \end{eqnarray}
526: where ${\bf p}_{1,2}^{(')}={\bf K}/2\pm {\bf p}^{(')}$; matrices
527: $\Gamma_{1,2}$ denote the Dirac structure of the interaction
528: potential (\ref{qpot}) for the first and second quark, respectively, and thus
529: $\Gamma_1\Gamma_2\tilde V(p-p')={\mathcal V}({\bf p}-{\bf p}')$. The
530: factor $\sqrt{2M}$ follows
531: from the normalization of the quasipotential wave function. The
532: positive- and negative-energy projectors have standard definition
533: \[
534: \Lambda^{(\pm)}(p)={\epsilon(p)\pm\bigl( m\gamma ^0+\gamma^0(
535: \bm{\gamma}{\bf p})\bigr) \over 2\epsilon (p)}.
536: \]
537: The quasipotential wave function in the rest frame
538: of the decaying meson $\Psi_M({\bf p})\equiv\Psi_{M\, {\bf 0}}({\bf p})$
539: can be expressed through a product of radial $\Phi_M(p)$,
540: spin $\chi_{ss'}$ and colour $\phi_{q_1q_2}$ wave functions
541: \begin{equation}
542: \label{eq:wff}
543: \Psi_M({\bf p})=\Phi_M(p)\chi_{ss'}\phi_{q_1q_2}.
544: \end{equation}
545:
546: Now the decay constants can be presented in the following form
547: \begin{equation}
548: \label{eq:fpe}
549: f_{P,V}=f_{P,V}^{(1)}+f_{P,V}^{(2+3)}+f_{P,V}^{(4)},
550: \end{equation}
551: where the terms on the right hand side originate from the corresponding
552: diagrams in Fig.~\ref{fig:diag} and parameterize respective terms in
553: Eq.~(\ref{eq:qpd}). In the literature \cite{gi,g,vd,efgdc} usually only
554: the first term is taken
555: into account, since it provides the nonrelativistic limit, while
556: other terms give only relativistic corrections and thus vanish in
557: this limit. Such approximation can be justified for mesons
558: containing heavy quarks. However, as it will be shown below, for
559: light mesons other terms become equally important and their account is
560: crucial for getting the results in agreement with experimental data.
561:
562: The matrix element (\ref{eq:qpd}) and thus the decay constants can be
563: calculated in an arbitrary
564: frame and from any component of the weak current. Such calculation can
565: be most easily performed in the rest frame of the decaying meson from the
566: zero component of the current. The same results will be obtained from the
567: vector component, however, this calculation is more cumbersome since
568: here the rest frame cannot be used and, thus, it is important to take
569: into account the relativistic transformation of the meson wave function
570: from the rest frame to the moving one with the momentum ${\bf K}$ (see
571: Eq.~(\ref{wig}) below). It
572: is also possible to perform calculations in the explicitly covariant
573: way using methods proposed in \cite{efgms}.
574:
575: The resulting expressions for decay constants are given by
576: \begin{eqnarray}
577: \label{eq:fpv1}
578: f^{(1)}_{P,V}&=&\sqrt{\frac{12}{M}}\int \frac{d^3
579: p}{(2\pi)^3}\left(\frac{\epsilon_1(p)+m_1}{2\epsilon_1(p)}\right)^{1/2}
580: \left(\frac{\epsilon_2(p)+m_2}{2\epsilon_2(p)}\right)^{1/2}
581: \cr
582: &&\times \left\{ 1
583: +\lambda_{P,V}\,\frac{{\bf p}^2}{[\epsilon_1(p)+m_1][\epsilon_2(p)+m_2]}\right\}
584: \Phi_{P,V}(p),
585: \end{eqnarray}
586: \begin{eqnarray}
587: \label{eq:fpv3}
588: f^{(2+3)}_{P,V}&=&\sqrt{\frac{12}{M}}\int \frac{d^3
589: p}{(2\pi)^3}\left(\frac{\epsilon_1(p)+m_1}{2\epsilon_1(p)}\right)^{1/2}
590: \left(\frac{\epsilon_2(p)+m_2}{2\epsilon_2(p)}\right)^{1/2}\Biggl[
591: \frac{M-\epsilon_1(p)-\epsilon_2(p)}{M+\epsilon_1(p)-\epsilon_2(p)}\cr
592: &&\times
593: \frac{{\bf p}^2}{\epsilon_1(p)[\epsilon_1(p)+m_1]}
594: \left\{1+\lambda_{P,V}\frac{\epsilon_1(p)+m_1}{\epsilon_2(p)+m_2}\right\}
595: +(1\leftrightarrow 2)\Biggr]
596: \Phi_{P,V}(p),
597: \end{eqnarray}
598: \begin{eqnarray}
599: \label{eq:fpv4}
600: f^{(4)}_{P,V}&=&\sqrt{\frac{12}{M}}\int \frac{d^3
601: p}{(2\pi)^3}\left(\frac{\epsilon_1(p)+m_1}{2\epsilon_1(p)}\right)^{1/2}
602: \left(\frac{\epsilon_2(p)+m_2}{2\epsilon_2(p)}\right)^{1/2}
603: \frac{M-\epsilon_1(p)-\epsilon_2(p)}{M+\epsilon_1(p)+\epsilon_2(p)}\cr
604: &&\times
605: \left\{-\lambda_{P,V}-\frac{{\bf p}^2}{[\epsilon_1(p)+m_1][\epsilon_2(p)+m_2]}\right\} \cr
606: &&\times
607: \left[\frac{(1-\varepsilon)m_1^2m_2^2}{\epsilon_1^2(p)\epsilon_2^2(p)}+
608: \frac{{\bf p}^2}{[\epsilon_1(p)+m_1][\epsilon_2(p)+m_2]}\right]
609: \Phi_{P,V}(p),
610: \end{eqnarray}
611: with $\lambda_P=-1$ and $\lambda_V=1/3$. Here $\varepsilon$ is the
612: mixing coefficient of scalar and vector confining potentials
613: (\ref{vlin}) and the long-range anomalous chromomagnetic quark moment
614: $\kappa$ (\ref{kappa}) is put equal to $-1$. Note that $f_P^{(2+3)}$
615: vanishes for pseudoscalar mesons with equal quark masses, such as the pion.
616: The positive-energy contribution (\ref{eq:fpv1}) reproduces the previously known
617: expressions for the decay constants \cite{gi,g}. The negative-energy
618: contributions (\ref{eq:fpv3}) and (\ref{eq:fpv4}) are new and play a
619: significant role for light mesons (see below).
620:
621: In the nonrelativistic limit $p^2/m^2\to 0$ the expression (\ref{eq:fpv1}) for
622: decay constants gives the well-known formula
623: \begin{equation}
624: \label{eq:fnr}
625: f_{P,V}^{\rm NR}=
626: \sqrt{\frac{12}{M_{P,V}}}\left|\Psi_{P,V}(0)\right|,
627: \end{equation}
628: where $\Psi_{P,V}(0)$ is the meson wave function at the origin
629: $r=0$. All other contributions vanish in the nonrelativistic limit.
630:
631: \begin{table}
632: \caption{Different contributions to the pseudoscalar and vector
633: decay constants of light mesons (in MeV). The notations are taken
634: according to Eqs.~(\ref{eq:fpe}) and (\ref{eq:fnr}).}
635: \label{tab:dc}
636: \begin{ruledtabular}
637: \begin{tabular}{cccccc}
638: Constant& $f_M^{\rm NR}$&$f_M^{(1)}$& $f_M^{(2+3)}+f_M^{(4)}$
639: & $(f_M^{(2+3)}+f_M^{(4)})/f^{(1)}_M $&$f_M$ \\
640: \hline
641: $f_\pi$ & 1290 &515 & $-391$ &$-76\%$ &124 \\
642: $f_K$ & 783 & 353 & $-198$& $-56\%$ & 155 \\
643: $f_\rho$ & 490 & 402 & $-183$&$-46\%$ & 219 \\
644: $f_{K^*}$ & 508 & 410 & $-174$&$-42\%$ & 236\\
645: $f_\phi$ & 511 & 415 &$-170$&$-41\%$ &245
646: \end{tabular}
647: \end{ruledtabular}
648:
649: \end{table}
650:
651: \begin{table}
652: \caption{Pseudoscalar and vector decay constants of light mesons (in
653: MeV).}
654: \label{tab:dce}
655: \begin{ruledtabular}
656: \begin{tabular}{ccccccccc}
657: Constant&this work&\cite{gi} & \cite{mr,mt} &\cite{k} &\cite{hjd}
658: &Lattice \cite{ak} &Lattice \cite{milc}& Experiment \cite{pdg}\\
659: \hline
660: $f_\pi$ & 124& 180 & 131&219&138&$126.6\pm6.4$ &$129.5\pm3.6$ &$130.7\pm0.1\pm0.36$\\
661: $f_K$ & 155& 232& 155& 238&160&$152.0\pm6.1$&$156.6\pm3.7$ & $159.8\pm1.4\pm0.44$\\
662: $f_\rho$ & 219&220&207& &238&$239.4\pm7.3$& & $220\pm2^*$\\
663: $f_{K^*}$ & 236& 267& 241& &241&$255.5\pm6.5$& & $230\pm8^\dag$\\
664: $f_\phi$ &245&336&259&&&$270.8\pm6.5$& & $229\pm3^\ddag$
665: \end{tabular}
666: \end{ruledtabular}
667: \begin{flushleft}
668: ${}^*$ derived from the experimental value for $\Gamma_{\rho^0\to
669: e^+e^-}$.\\
670: ${}^\dag$ derived from the experimental value for the ratio
671: $\Gamma_{\tau\to K^*\nu_\tau}/\Gamma_{\tau\to\rho\nu_\tau}$ and the
672: $f_\rho$ value.\\
673: ${}^\ddag$ derived from the experimental value for $\Gamma_{\phi\to
674: e^+e^-}$.
675: \end{flushleft}
676:
677: \end{table}
678:
679:
680: In Table~\ref{tab:dc} we present our predictions for the
681: light meson decay constants calculated using the meson wave functions
682: which were obtained as the numerical solutions of the quasipotential equation in
683: Sec.~\ref{sec:lmm}. The nonrelativistic values $f_M^{\rm NR}$
684: (\ref{eq:fnr}) as well as the values of different contributions in
685: Fig.~\ref{fig:diag} $f_M^{(1,2,3,4)}$ (\ref{eq:fpv1})--(\ref{eq:fpv4})
686: and the full relativistic results $f_M$ (\ref{eq:fpe}) are given. In
687: Table~\ref{tab:dce} we compare our results for the decay constants $f_M$
688: with predictions of other approaches \cite{gi,mr,mt,k,hjd}, recent values from
689: two- \cite{ak} and three-flavour lattice QCD \cite{milc} and
690: available experimental data \cite{pdg}. It is clearly seen that the
691: nonrelativistic predictions are significantly overestimating all
692: decay constants, especially for the pion (almost by a factor of 10). The
693: account of the part of relativistic corrections by keeping in
694: Eq.~(\ref{eq:fpe}) only the first term $f_M^{(1)}$
695: (\ref{eq:fpv1}), which is usually used for semirelativistic
696: calculations, does not dramatically improve the situation. The
697: disagreement is still large. This is connected with the anomalously
698: small masses of light pseudoscalar mesons exhibiting their chiral
699: nature. In the semirelativistic quark model \cite{gi,g} the pseudoscalar
700: meson mass is replaced by the so-called mock mass $\tilde M_P$, which is equal to
701: the mean total energy of free quarks in a meson, and with our wave
702: functions: $\tilde
703: M_\pi=2\langle\epsilon_q(p)\rangle\approx 1070$~MeV ($\sim 8 M_\pi$)
704: and $\tilde
705: M_K=\langle\epsilon_q(p)\rangle+\langle\epsilon_s(p)\rangle\approx
706: 1232$~MeV ($\sim 2.5 M_K$). Such replacement gives $f_P^{(1)}$ values
707: which are still $\approx 1.4$ times larger than experimental ones
708: (cf. \cite{gi}). As we see from Table~\ref{tab:dc}, in the
709: quasipotential approach it is not justified to neglect contributions of
710: the negative energy intermediate states for light meson decay
711: constants. Indeed, the values of $f_M^{(2+3)}+f_M^{(4)}$ are large and
712: negative (reaching $-76\%$ of $f_\pi^{(1)}$ for the pion)
713: thus compensating the overestimation of decay constants by
714: the positive-energy contribution $f_M^{(1)}$. This is the consequence
715: of the smallness of the
716: light pseudoscalar meson masses compared to the energies of their
717: constituents. The negative-energy contributions (\ref{eq:fpv3}),
718: (\ref{eq:fpv4}) are proportional to the ratio of the meson binding
719: energy $M-\epsilon_1(p)-\epsilon_2(p)$ to its mass. For mesons with
720: heavy quarks this factor leads to the suppression of negative-energy
721: contributions since the binding energies are small on the heavy
722: meson mass scale. This results in the dominance of the positive-energy
723: term $f_M^{(1)}$ since the negative-energy terms give only $1/m_Q$
724: contributions ($m_Q$ is the heavy quark mass).\footnote{For the
725: heavy-heavy $B_c$ meson ($c\bar b$) these negative-energy
726: corrections will be of order $v^4/c^4$ and thus very small. The
727: influence of the
728: negative-energy contributions $f_M^{(2+3,4)}$ on the decay constants
729: of heavy-light $B$ and $D$ mesons will be considered elsewhere.}
730: On the other hand, for light mesons,
731: especially for the pion and kaon, the binding energies are large on the
732: light meson mass scale and, thus, such factor gives no
733: suppression. Taking the complete relativistic expression for decay
734: constants $f_M$ (\ref{eq:fpe}) brings theoretical predictions in good
735: agreement with available experimental data.
736:
737: The comparison of our values of the decay constants with other predictions in
738: Table~\ref{tab:dce} indicate that they are competitive even with the results of
739: more sophisticated approaches \cite{mt,k} which are based on the
740: Dyson-Schwinger and Bethe-Salpeter equations. On the other hand our
741: model is more selfconsistent than some other approaches \cite{hjd,kt,gi,g,vd}.
742: We calculate the meson wave functions by solving the
743: quasipotential equation in contrast to the models based on the
744: relativistic Hamilton dynamics \cite{hjd,kt} where various ad hoc
745: wave function parameterizations are employed. We also do not need to
746: introduce the mock meson mass \cite{gi,g,vd} and to substitute it for the
747: light meson mass as it was discussed above.
748:
749:
750:
751:
752:
753: \section{ Electromagnetic form factors}
754: \label{em}
755:
756: The elastic matrix element of the
757: electromagnetic current $J_\mu$ between the initial and final
758: pseudoscalar meson states is parameterized by the form factor $F_P(Q^2)$
759: \begin{equation}
760: \label{eq:sff}
761: \langle M(P_F)\vert J_\mu \vert M(P_I)\rangle=F_P(Q^2)(P_I+P_F)_\mu,
762: \end{equation}
763: where $Q^2=-(P_F-P_I)^2$.
764:
765: \begin{figure}%[htb]
766: \centering
767: \includegraphics{fig1.eps}
768: \caption{Lowest order vertex function $\Gamma^{(1)}$
769: corresponding to Eq.~(\ref{gam1}). Photon interaction with one quark is shown.}
770: \label{fig:1}
771: \end{figure}
772:
773: \begin{figure}%[htb]
774: \centering
775: \includegraphics{fig2.eps}
776: \caption{ Vertex function $\Gamma^{(2)}$
777: corresponding to Eq.~(\ref{gam2}). Dashed lines represent the interaction
778: operator ${\mathcal V}$. Bold lines denote the
779: negative-energy part of the quark propagator. As on Fig.~\ref{fig:1},
780: photon interaction with one quark is shown.}
781: \label{fig:2}
782: \end{figure}
783:
784:
785: In the quasipotential approach such matrix element has the form \cite{f}
786: \begin{equation}
787: \label{mxet}
788: \langle M(P_F) \vert J_\mu \vert M(P_I)\rangle
789: =\int \frac{d^3p\, d^3q}{(2\pi )^6} \bar \Psi_{M\, {\bf P}_F}({\bf
790: p})\Gamma _\mu ({\bf p},{\bf q})\Psi_{M\, {\bf P}_I}({\bf q}),
791: \end{equation}
792: where $\Gamma _\mu ({\bf p},{\bf
793: q})$ is the two-particle vertex function and $\Psi_{M}$ are the
794: meson wave functions projected onto the positive energy states of
795: quarks and boosted to the moving reference frame.
796: The contributions to $\Gamma$ come from Figs.~\ref{fig:1} and \ref{fig:2}.
797: The term $\Gamma^{(2)}$ includes contributions from the
798: negative-energy quark states. Note that the form
799: of the relativistic corrections resulting from the vertex function
800: $\Gamma^{(2)}$ explicitly depends on the Lorentz structure of the
801: $q\bar q$-interaction. Thus the vertex function is given by
802: \begin{equation}
803: \label{eq:gam}
804: \Gamma_\mu({\bf p},{\bf q})=\Gamma_\mu^{(1)}({\bf p},{\bf q})+
805: \Gamma_\mu^{(2)}({\bf p},{\bf q})+ \cdots ,
806: \end{equation}
807: where
808: \begin{equation}\label{gam1}
809: \Gamma_\mu ^{(1)}({\bf p},{\bf q})=e_1\bar
810: u_1(p_1)\gamma_\mu
811: u_1(q_1)(2\pi)^3\delta({\bf p}_2-{\bf q}_2) +(1\leftrightarrow 2),
812: \end{equation}
813: and
814: \begin{eqnarray}\label{gam2}
815: \Gamma_\mu^{(2)}({\bf p},{\bf q})&=&e_1\bar u_1(p_1)\bar
816: u_2(p_2) \biggl\{{\mathcal V}({\bf p}_2-{\bf q}_2)
817: \frac{\Lambda_1^{(-)}(k_1')}{
818: \epsilon_1(k_1')+ \epsilon_1(q_1)}\gamma_1^0\gamma_{1\mu}
819: \nonumber\\
820: & & + \gamma_{1\mu}
821: \frac{\Lambda_1^{(-)}({k}_1)}{ \epsilon_1(k_1)+\epsilon_1(p_1)}
822: \gamma_1^0{\mathcal V}({\bf p}_2-{\bf q}_2)
823: \biggr\}u_1(q_1) u_2(q_2) +(1\leftrightarrow 2).
824: \end{eqnarray}
825: Here $e_{1,2}$ are the quark charges, ${\bf k}_1={\bf p}_1-{\bf\Delta};
826: \quad {\bf k}_1'={\bf
827: q}_1+{\bf\Delta};\quad {\bf\Delta}={\bf P}_F-{\bf P}_I$;
828: \[
829: \Lambda^{(-)}(p)={\epsilon(p)-\bigl( m\gamma ^0+\gamma^0(
830: \bm{\gamma}{\bf p})\bigr) \over 2\epsilon (p)}, \qquad \epsilon(p)=
831: \sqrt{p^2+m^2},
832: \]
833: and
834: \begin{eqnarray*}
835: p_{1,2}&=&\epsilon_{1,2}(p)\frac{P_{F}}{M}
836: \pm\sum_{i=1}^3 n^{(i)}(P_{F})p^i,\\
837: q_{1,2}&=&\epsilon_{1,2}(q)\frac{P_I}{M} \pm \sum_{i=1}^3 n^{(i)}
838: (P_I)q^i,\end{eqnarray*}
839: where $n^{(i)}$ are three four-vectors given by
840: $$ n^{(i)\mu}(p)=\left\{ \frac{p^i}{M},\ \delta_{ij}+
841: \frac{p^ip^j}{M(E+M)}\right\}, \quad E=\sqrt{{\bf p}^2+M^2},
842: \qquad i,j=1,2,3,$$
843: $P_I=(E_I,{\bf P}_I)$ and $P_F=(E_F,{\bf P}_F)$ are four-momenta of
844: the initial and final mesons.
845:
846:
847:
848: It is important to note that the wave functions entering the current
849: matrix element (\ref{mxet}) cannot be both in the rest frame.
850: In the initial meson rest frame, the final meson is moving
851: with the recoil momentum ${\bf \Delta}$. The wave function
852: of the moving meson $\Psi_{M\,{\bf\Delta}}$ is connected
853: with the wave function in the rest frame
854: $\Psi_{M\,{\bf 0}}\equiv \Psi_{M}$ by the transformation \cite{f}
855: \begin{equation}
856: \label{wig}
857: \Psi_{M\,{\bf\Delta}}({\bf p})
858: =D_1^{1/2}(R_{L_{\bf\Delta}}^W)D_2^{1/2}(R_{L_{\bf\Delta}}^W)
859: \Psi_{M\,{\bf 0}}({\bf p}),
860: \end{equation}
861: where $R^W$ is the Wigner rotation, $L_{\bf\Delta}$ is the Lorentz boost
862: from the rest frame to a moving one, and the rotation matrix
863: $D^{1/2}(R)$ in the spinor representation is given by
864: \begin{equation}\label{d12}
865: {1 \ \ \,0\choose 0 \ \ \,1}D^{1/2}_{1,2}(R^W_{L_{\bf\Delta}})=
866: S^{-1}({\bf p}_{1,2})S({\bf\Delta})S({\bf p}),
867: \end{equation}
868: where
869: $$
870: S({\bf p})=\sqrt{\frac{\epsilon(p)+m}{2m}}\left(1+\frac{\bm{\alpha}
871: {\bf p}} {\epsilon(p)+m}\right)
872: $$
873: is the usual Lorentz transformation matrix of the Dirac spinor.
874:
875: To calculate the matrix element (\ref{eq:sff}) of the electromagnetic
876: current between the pseudoscalar meson states we substitute the vertex
877: functions $\Gamma^{(1)}$ (\ref{gam1}) and $\Gamma^{(2)}$ (\ref{gam2})
878: in Eq.~(\ref{mxet}) and take into account the wave function
879: transformation (\ref{wig}). Then we use the $\delta$ function in
880: $\Gamma^{(1)}$ to perform one of the integrations in the matrix
881: element (\ref{mxet}). For the contribution of $\Gamma^{(2)}$ we use
882: instead the quasipotential equation to replace the integral of the product
883: of the interaction potential and the bound state wave function by the
884: product of the corresponding binding energy and the wave function. To
885: simplify the calculation we explicitly use the value $\kappa=-1$ for
886: the long-range anomalous chromomagnetic quark moment
887: (\ref{kappa}). However, as previously we keep the dependence on the
888: mixing parameter $\varepsilon$ of the vector and scalar confining
889: potentials (\ref{vlin}). As a
890: result we get the following expression for the electromagnetic form
891: factor of the pseudoscalar meson:
892:
893:
894: \begin{equation}
895: \label{eq:eff}
896: F_P({Q}^2)=F_P^{(1)}({Q}^2)+\varepsilon F_P^{(2)S}({Q}^2)
897: +(1-\varepsilon)F_P^{(2)V}({Q}^2),
898: \end{equation}
899: \begin{eqnarray}\label{eq:f1}
900: F_P^{(1)}({Q}^2)&=&\frac{2\sqrt{EM}}{E+M}\Biggl\{e_1
901: \int \frac{d^3p}{(2\pi )^3} \bar\Psi_{M}
902: \left({\bf p}+
903: \frac{2\epsilon_{2}(p)}{E+M}{\bf \Delta } \right)
904: \sqrt{\frac{\epsilon_1(p)+m_1}{\epsilon_1(p+\Delta)+m_1}}
905: \Biggl[\frac{\epsilon_1(p+\Delta)+\epsilon_1(p)}
906: {2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}}\cr
907: &&+\frac{\bf p \Delta}{2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}
908: (\epsilon_1(p)+m_1)} -\frac{\epsilon_1(p+\Delta)-\epsilon_1(p)}
909: {2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}}
910: \frac{{\bf p}_T^2}{\epsilon_1(p)+m_1}\cr
911: &&\times
912: \left(\frac1{\epsilon_1(p)+m_1}+\frac1{\epsilon_2(p)+m_2}\right)
913: \Biggr]\Psi_{M}({\bf
914: p})+(1\leftrightarrow 2)\Biggr\},
915: \end{eqnarray}
916:
917: \begin{eqnarray}\label{eq:f2s}
918: F_P^{(2)S}({Q}^2)&=&\frac{2\sqrt{EM}}{E+M}\Biggl\{e_1
919: \int \frac{d^3p}{(2\pi )^3} \bar\Psi_{M}
920: \left({\bf p}+
921: \frac{2\epsilon_{2}(p)}{E+M}{\bf \Delta } \right)
922: \sqrt{\frac{\epsilon_1(p)+m_1}{\epsilon_1(p+\Delta)+m_1}}
923: \frac{\epsilon_1(p+\Delta)+m_1}
924: {2\epsilon_1(p+\Delta)}\cr
925: &&\times
926: \Biggl[\frac{\epsilon_1(p+\Delta)-\epsilon_1(p)+2m_1}
927: {2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}}
928: -\frac{\bf p \Delta}{2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}
929: (\epsilon_1(p)+m_1)}\cr
930: && -\frac{\epsilon_1(p+\Delta)+m_1}
931: {2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}} \frac{{\bf
932: p}_T^2}{\epsilon_1(p)+m_1}
933: \left(\frac1{\epsilon_1(p)+m_1}+\frac1{\epsilon_2(p)+m_2}\right)\Biggr]\cr
934: &&\times
935: \frac{\epsilon_1(p+\Delta)-\epsilon_1(p)}{\epsilon_1(p+\Delta)
936: [\epsilon_1(p+\Delta)+\epsilon_1(p)]}[M-\epsilon_1(p)-\epsilon_2(p)]
937: \Psi_{M}({\bf
938: p})+(1\leftrightarrow 2)\Biggr\},
939: \end{eqnarray}
940:
941: \begin{eqnarray}\label{eq:f2v}
942: F_P^{(2)V}(Q^2)&=&\frac{2\sqrt{EM}}{E+M}\Biggl\{e_1
943: \int \frac{d^3p}{(2\pi )^3} \bar\Psi_{M}
944: \left({\bf p}+
945: \frac{2\epsilon_{2}(p)}{E+M}{\bf \Delta } \right)
946: \sqrt{\frac{\epsilon_1(p)+m_1}{\epsilon_1(p+\Delta)+m_1}}
947: \frac{\epsilon_1(p+\Delta)+m_1}
948: {2\epsilon_1(p+\Delta)}\cr
949: &&\times
950: \Biggl[\frac{\epsilon_1(p)-m_1}
951: {2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}}
952: +\frac{\bf p \Delta}{2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}
953: (\epsilon_1(p)+m_1)}\cr
954: && +\frac{\epsilon_1(p+\Delta)+m_1}
955: {2\sqrt{\epsilon_1(p+\Delta)\epsilon_1(p)}} \frac{{\bf
956: p}_T^2}{\epsilon_1(p)+m_1}
957: \left(\frac1{\epsilon_1(p)+m_1}+\frac1{\epsilon_2(p)+m_2}\right)\Biggr]\cr
958: &&\times
959: \frac{\epsilon_1(p+\Delta)-\epsilon_1(p)}{\epsilon_1(p+\Delta)
960: [\epsilon_1(p+\Delta)+\epsilon_1(p)]}[M-\epsilon_1(p)-\epsilon_2(p)]
961: \Psi_{M}({\bf
962: p})+(1\leftrightarrow 2)\Biggr\},
963: \end{eqnarray}
964: where $F_P^{(2)S(V)}$ are contributions from scalar (vector) confining
965: potentials and ${\bf p}_T={\bf p}^2-({\bf p\Delta})^2/{\bf \Delta}^2$,
966: $E=\sqrt{M^2+\bf{\Delta}^2}$. As previously, we put $\varepsilon=-1$
967: for further numerical
968: calculations. It is important to note that the above expressions for
969: the electromagnetic form factor of the positively-charged pseudoscalar meson
970: exactly satisfy the normalization condition
971: \begin{equation}
972: \label{eq:nc}
973: F_P(0)=1
974: \end{equation}
975: following from the electric charge conservation.
976:
977: \begin{figure}%[htb]
978: \centering
979: \includegraphics[height=12cm,angle=-90]{fpi.eps}
980:
981: \includegraphics[height=12cm,angle=-90]{plp2.eps}
982:
983: \caption{The charged pion form factor squared in comparison with
984: experimental data from Refs.~\cite{amend} (open circles),
985: \cite{bebek} (solid squares) and \cite{jlab} (crosses).}
986: \label{fig:fpi2}
987: \end{figure}
988:
989: \begin{figure}%[htb]
990: \centering
991: \includegraphics[height=12cm,angle=-90]{fpil.eps}
992: \caption{$Q^2$ times charged pion form factor in comparison with
993: experimental data from Refs.~\cite{amend} (open circles),
994: \cite{bebek} (solid squares) and \cite{jlab} (crosses). }
995: \label{fig:q2fpi}
996: \end{figure}
997:
998:
999: \begin{figure}%[htb]
1000: \centering
1001: \includegraphics[height=12cm,angle=-90]{pk2.eps}
1002:
1003: \caption{The charged kaon form factor squared in comparison with
1004: experimental data from Refs.~\cite{dally} (open circles) and
1005: \cite{amend2} (solid squares).}
1006: \label{fig:fk2}
1007: \end{figure}
1008:
1009: \begin{figure}%[htb]
1010: \centering
1011: \includegraphics[height=12cm,angle=-90]{pkq2.eps}
1012: \caption{$Q^2$ times the charged kaon (solid line) and
1013: neutral kaon (dashed line) form factors. }
1014: \label{fig:q2fk}
1015: \end{figure}
1016:
1017: Now we can use the wave functions of the pseudoscalar light mesons
1018: ($\pi$, $K$), found in
1019: Sec.~\ref{sec:lmm}, for the numerical calculation of their
1020: electromagnetic form factors $F_P(Q^2)$ in the space-like region
1021: $Q^2\ge 0$. The results of such calculations for the charged pion are shown
1022: in Figs.~\ref{fig:fpi2} ($F_\pi^2(Q^2)$) and \ref{fig:q2fpi}
1023: ($Q^2F_\pi(Q^2)$) in comparison with
1024: experimental data from Refs.~\cite{amend,bebek,jlab}. Good agreement
1025: with data both in low and high $Q^2$ regions is found, including recent
1026: JLab data \cite{jlab} which are
1027: plotted with crosses. It is clearly seen
1028: from Fig.~\ref{fig:q2fpi} that the calculated pion form factor at
1029: high $Q^2$ exhibits the asymptotic behaviour $F_\pi(Q^2)\sim
1030: \alpha_s(Q^2)/Q^2$ predicted by the quark counting rule \cite{qcr}
1031: and perturbative QCD \cite{asimpt}. Our results for the pion form
1032: factor can also be compared with QCD based calculations
1033: \cite{bpss} and with recent parameterizations \cite{pp,bkk}
1034: which arise from the constraints of analyticity and unitarity. The
1035: latter form factor models are based on the vector meson dominance and
1036: include a
1037: pattern of radial excitations expected from dual resonance models
1038: \cite{bkk}. The consistency of our results with such parameterizations
1039: (cf. Fig.~\ref{fig:fpi2} with Fig.~2 of Ref.~\cite{bkk}) just means
1040: the manifestation of the quark-hadron duality. Finally, our predictions
1041: agree fairly well with recent lattice computations of the pion form
1042: factor~\cite{beflr,ha}.
1043: The corresponding plots for the charged kaon form factor are given in
1044: Figs.~\ref{fig:fk2} and \ref{fig:q2fk} in comparison with experimental
1045: data from Refs.~\cite{dally,amend2}, which are available only for the low
1046: $Q^2$ region. Again good agreement with experimental data is
1047: found. On Fig.~\ref{fig:q2fk} we also plot the neutral kaon form
1048: factor by the dashed line.
1049:
1050:
1051:
1052:
1053:
1054: \begin{table}
1055: \caption{Charge radii of light pseudoscalar mesons.}
1056: \label{tab:cr}
1057: \begin{ruledtabular}
1058: \begin{tabular}{ccccccc}
1059: charge radii& this work& \cite{gi}& \cite{mt} &\cite{hjd}&Lattice \cite{ha}
1060: & Experiment \cite{pdg}\\
1061: \hline
1062: $\sqrt{\langle r^2\rangle_\pi}$ (fm) & 0.66&0.66&0.67&0.63&$0.63\pm0.1$ & 0.672$\pm$0.08 \\
1063: $\sqrt{\langle r^2\rangle_{K^\pm}}$ (fm) & 0.57&0.59&0.62&0.60& & 0.560$\pm$0.031 \\
1064: $\langle r^2\rangle_{K^0}$ (fm$^2$)& $-0.072$&$-0.09$&$-0.086$&$-0.062$& & $-0.076\pm$0.018
1065: \end{tabular}
1066: \end{ruledtabular}
1067: \end{table}
1068:
1069: The mean-squared charge radius of the pseudoscalar meson ($P=\pi,K$) is defined by
1070: \begin{equation}
1071: \label{eq:chr}
1072: \langle r^2\rangle_P=-6\left[\frac{{\rm d}F_P(Q^2)}{{\rm d}Q^2}\right]_{Q^2=0}.
1073: \end{equation}
1074: The calculated values of the charge radii of light pseudoscalar mesons
1075: are given in Table~\ref{tab:cr} in comparison with predictions of other
1076: approaches \cite{gi,mt,hjd,ha} and experimental data \cite{pdg}. An
1077: overall good agreement with experimental data is
1078: found.
1079:
1080:
1081: \section{Conclusions}
1082: \label{sec:concl}
1083:
1084: The relativistic quark model, which has been
1085: previously developed and successfully used for the comprehensive
1086: investigation of different properties of heavy and heavy-light
1087: hadrons, was applied here for
1088: calculating the masses, weak decay constants and
1089: electromagnetic form factors of the light mesons. The main assumptions
1090: and parameters of the model (such as the Lorentz structure and
1091: parameters of the confining
1092: potential and quark masses) were kept the same as in previous studies. The
1093: only change we made, is the necessary modification of the running coupling
1094: constant $\alpha_s(\mu^2)$ in the infrared region.
1095: Following Ref.~\cite{bvb} we chose
1096: the simplest model with freezing (\ref{eq:alpha}). Therefore only one
1097: additional parameter $\Lambda$ was introduced and it was fixed from
1098: fitting the $\rho$ meson mass. We constructed the local relativistic
1099: quasipotential for the light quarks using the replacement
1100: (\ref{eq:sub}), which was previously tested on the heavy-light mesons.
1101: The resulting relativistic potential (\ref{eq:v}) depends on the meson mass in a
1102: complicated nonlinear way. Solving numerically the quasipotential
1103: equation (\ref{quas}) we got masses of the ground-state and
1104: radially-excited light mesons in a reasonably good overall agreement with
1105: experimental data. Even the masses of the pseudoscalar $\pi$ and $K$
1106: mesons are well reproduced. This is a nontrivial result, since we use the
1107: constituent quark masses in our description and thus the chiral
1108: symmetry is explicitly broken from the very beginning. We determined
1109: the light meson wave functions and used them for studying their
1110: electroweak properties.
1111:
1112: First the weak decay constants of pseudoscalar and vector mesons were
1113: investigated. It was argued that both positive- and negative-energy
1114: parts of the quark propagators in the weak annihilation loop should be
1115: taken into account. Usually in the semirelativistic quark model
1116: \cite{gi,g,vd} only the positive-energy contributions are kept. This
1117: approximation requires to replace in the expression for the
1118: pseudoscalar decay constant (\ref{eq:fpv1}) the meson mass by
1119: the so-called mock meson mass, which is considerably larger, in order
1120: not to get the significant overestimate of the decay constants. We showed
1121: that the negative-energy contributions to the light meson pseudoscalar
1122: decay constants are large and negative. Their account brings
1123: theoretical predictions (with the physical meson masses) in good agreement
1124: with available experimental data.
1125:
1126: Next we studied the electromagnetic form factor of the pseudoscalar
1127: mesons. The corresponding matrix element of the electromagnetic
1128: current was calculated using the quasipotential approach. The additional
1129: contributions of the intermediate negative-energy states (\ref{gam2})
1130: were taken into account as well as the transformation of the meson wave
1131: function from the rest frame to a moving one (\ref{wig}). As a result
1132: the relativistic expression for the electromagnetic form factor was
1133: obtained. We then calculated the pion, charged and neutral kaon form factors
1134: in the space-like region. Good agreement with available experimental
1135: data both in small and large $Q^2$ regions were found. At large
1136: momentum transfer this form factor tends to reproduce the power-law behaviour
1137: predicted by perturtbative QCD \cite{asimpt}. The calculated charge
1138: radii of light pseudoscalar mesons are in good agreement with
1139: experiment.
1140:
1141: In conclusion, we found that the obtained results are
1142: quite competitive with the predictions of other approaches
1143: \cite{gi,mr,k,ak,mt,hjd,kt,milc,beflr,ha} including more sophisticated ones,
1144: which were specially developed for treating light mesons.
1145:
1146: \acknowledgments
1147: The authors are grateful to A. Ali Khan, A. Badalian, M. M\"uller-Preussker,
1148: V. Savrin and Yu. Simonov for useful discussions. Two of us
1149: (R.N.F. and V.O.G.) were supported in part by the {\it Deutsche
1150: Forschungsgemeinschaft} under contract Eb 139/2-3 and by the {\it Russian
1151: Foundation for Basic Research} under Grant No.05-02-16243.
1152:
1153:
1154:
1155: \begin{thebibliography}{00}
1156: \bibitem{fr} C. Feuchter and H. Reinhardt, Phys. Rev. D {\bf 70},
1157: 105021 (2004).
1158: \bibitem{gi} S. Godfrey and N. Isgur, { Phys. Rev. D} {\bf 32}, 189
1159: (1985).
1160: \bibitem{mr} P. Maris and C.D. Roberts, { Int.J. Mod. Phys. E} {\bf
1161: 12}, 297 (2003); P. Maris and P.C. Tandy, { Phys. Rev. C} {\bf
1162: 60}, 055214 (1999).
1163: \bibitem{k} M. Koll et al., {Eur. Phys. J. A} {\bf 9}, 73 (2000).
1164: \bibitem{erv} D. Ebert and H. Reinhardt, Nucl. Phys. B {\bf 271}, 188
1165: (1986); D. Ebert, H. Reinhardt and M.K. Volkov, {
1166: Prog. Part. Nucl. Phys.} {\bf 33}, 1 (1994); M.K. Volkov, D. Ebert
1167: and M. Nagy, { Int. J. Mod. Phys. A} {\bf 13}, 5443
1168: (1998); M.K. Volkov and C. Weiss, Phys. Rev. D {\bf 56}, 221
1169: (1997).
1170: \bibitem{kt} A. F. Krutov and V. E. Troitsky, Phys. Rev. C {\bf 65},
1171: 045501 (2002).
1172: \bibitem{hjd} J. He, B. Julia-Diaz and Y. Dong, Eur. Phys. J A {\bf
1173: 24}, 411 (2005).
1174: \bibitem{kp} N.V. Krasnikov and A.A. Pivovarov, {Phys. Lett. B}
1175: {\bf 112}, 397 (1982);
1176: N.V. Krasnikov, A.A. Pivovarov and N.N. Tavkhelidze, {Z. Phys. C}
1177: {\bf 19}, 301 (1983); A.L. Kataev, hep-ph/9805218.
1178: \bibitem{bkm} V.M. Braun, A. Khodjamirian and M. Maul, Phys. Rev. D
1179: {\bf 61}, 073004 (2000).
1180:
1181: \bibitem{ak} A. Ali Khan et al., Phys. Rev. D {\bf 65}, 054505 (2002);
1182: {\bf 67}, 059901(E) (2003); for a recent review see A. Ali Khan,
1183: hep-lat/0507031 and references therein.
1184: \bibitem{efg} D. Ebert, R.N. Faustov and V.O. Galkin, {
1185: Phys. Rev. D} {\bf 67}, 014027 (2003).
1186: \bibitem{egf} D. Ebert, V.O. Galkin and R.N. Faustov, {
1187: Phys. Rev. D} {\bf 57}, 5663 (1998); {\bf 59}, 019902(E) (1999).
1188: \bibitem{hbm} D. Ebert, R.N. Faustov and V.O. Galkin, {
1189: Phys. Rev. D} {\bf 72}, 034026 (2005).
1190: \bibitem{fgm} R.N. Faustov, V.O. Galkin and A.Yu. Mishurov,
1191: Phys. Rev. D {\bf 53}, 6302 (1996).
1192: \bibitem{dhb} D. Ebert, R.N. Faustov, V.O. Galkin and A.P. Martynenko, {
1193: Phys. Rev. D} {\bf 66}, 014008 (2002); Phys. Atom. Nucl. {\bf 68},
1194: 784 (2005).
1195: \bibitem{lmm} D. Ebert, R.N. Faustov and V.O. Galkin,
1196: Mod. Phys. Lett. A {\bf 20}, 1887 (2005).
1197:
1198:
1199: \bibitem{dk} D. Diakonov, { Progr. Part. Nucl. Phys.} {\bf 51}, 173
1200: (2003); N.I. Kochelev, { Phys. Lett. B} {\bf 426}, 149 (1998).
1201: \bibitem{bs} H.A. Bethe and E.E. Salpeter, {\it Quantum Mechanics of
1202: One-and Two-Electron Atoms} (Springer-Verlag, Berlin, 1957).
1203: \bibitem{bvb} A.M. Badalian, A.I. Veselov and B.L.G. Bakker, {
1204: Phys. Rev. D} {\bf 70}, 016007 (2004); Yu.A. Simonov, {
1205: Phys. Atom. Nucl.} {\bf 58}, 107 (1995).
1206: \bibitem{pdg} Particle Data Group, S. Eidelman et al., {
1207: Phys. Lett. B} {\bf 592}, 1 (2004).
1208:
1209:
1210: \bibitem{g} S. Godfrey, {Phys. Rev. D} {\bf 33}, 1391 (1986).
1211: \bibitem{vd} S. Veseli and I. Dunietz, Phys. Rev. D {\bf 54}, 6803
1212: (1996).
1213: \bibitem{efgdc} D. Ebert, R.N. Faustov and V.O. Galkin,
1214: Mod. Phys. Lett. A {\bf 17}, 803 (2002).
1215: \bibitem{efgms} D. Ebert, R.N. Faustov, V.O. Galkin and A.P. Martynenko, {
1216: Phys. Rev. D} {\bf 70}, 014018 (2004).
1217: \bibitem{mt} P. Maris and P. Tandy, Phys. Rev. C {\bf 62}, 055204
1218: (2000).
1219:
1220:
1221: \bibitem{milc} MILC Collaboration, C. Aubin et al., Phys. Rev. D {\bf
1222: 70}, 114501 (2004); C.T.H. Davies et al., Phys. Rev. Lett. {\bf
1223: 92}, 022001 (2004).
1224:
1225: \bibitem{f} R.N. Faustov, Ann. Phys. (N. Y.) {\bf 78}, 176 (1973);
1226: Nuovo Cimento A {\bf 69}, 37 (1970).
1227: \bibitem{amend} S.R. Amendolia et al., Nucl. Phys. B {\bf 277}, 168
1228: (1986).
1229: \bibitem{bebek} C.J. Bebek et al., Phys. Rev. D {\bf 17}, 1693
1230: (1978).
1231: \bibitem{jlab} J. Volmer et al., Phys. Rev. Lett. {\bf 86}, 1713
1232: (2001).
1233:
1234: \bibitem{qcr} V.A. Matveev, R.M. Muradian and A.N. Tavkhelidze,
1235: Lett. Nuovo Cim. {\bf 7}, 719 (1973); S.J. Brodsky and G.R. Farrar,
1236: Phys. Rev. Lett. {\bf 31}, 1153 (1973).
1237: \bibitem{asimpt} G.P. Lepage and S.J. Brodsky, Phys. Lett. B {\bf
1238: 87}, 359 (1979); G.R. Farrar and D.R. Jackson,
1239: Phys. Rev. Lett. {\bf 43}, 246 (1979); A.V. Efremov and
1240: A.V. Radyushkin, Phys. Lett. B {\bf 94}, 245 (1980).
1241: \bibitem{bpss} A.P.~Bakulev, K.~Passek-Kumericki, W.~Schroers and
1242: N.G.~Stefanis, Phys. Rev. D {\bf 70}, 033014 (2004); N.G.~Stefanis,
1243: W.~Schroers and H.C.~Kim, Phys. Lett. B {\bf 449}, 299 (1999).
1244:
1245: \bibitem{pp} A. Pich and J. Portoles, Phys. Rev. D {\bf 63}, 093005
1246: (2001).
1247: \bibitem{bkk} C. Bruch, A. Khadjamirian and J.H. K\"uhn,
1248: Eur. Phys. J. C {\bf 39}, 41 (2005).
1249: \bibitem{beflr} F.D.R. Bonnet et al., Phys. Rev. D {\bf 72}, 054506
1250: (2005).
1251: \bibitem{ha} S. Hashimoto et al., hep-lat/0510085.
1252:
1253: \bibitem{dally} E.B. Dally et al., Phys. Rev. Lett. {\bf 45}, 232
1254: (1980).
1255: \bibitem{amend2} S.R. Amendolia et al., Phys. Lett. B {\bf 178}, 435
1256: (1986).
1257:
1258: \end{thebibliography}
1259:
1260:
1261:
1262: \end{document}
1263: