hep-ph0601234/text.tex
1: %\documentclass[twocolumn,showpacs,aps,prd,amsfonts,eqsecnum,
2: \documentclass[onecolumn,showpacs,aps,prd,amsfonts,eqsecnum,
3: superscriptaddress,nofootinbib]{revtex4}
4: \usepackage{epsfig}
5: \usepackage{bm}
6: 
7: \begin{document}
8: 
9: \newcommand*{\nl}{\nonumber \\}
10: \newcommand*{\bea}{\begin{eqnarray}}
11: \newcommand*{\eea}{\end{eqnarray}}
12: \newcommand*{\bi}{\bibitem}
13: \newcommand*{\be}{\begin{equation}}
14: \newcommand*{\ee}{\end{equation}}
15: \newcommand*{\rms}{M_\rho^2(s)}
16: \newcommand*{\mrs}{m_\rho^2}
17: \newcommand*{\ra}{\rightarrow}
18: \newcommand*{\die}{e^+e^-}
19: \newcommand*{\eepppp}{e^+e^-\rightarrow\pi^+\pi^-\pi^+\pi^-}
20: \newcommand*{\rpppp}{\rho^0\rightarrow\pi^+\pi^-\pi^+\pi^-}
21: \newcommand*{\arp}{{a_1\rho\pi}}
22: \newcommand*{\eg}{e.g.}
23: \newcommand*{\rp}{{\rho^\prime}}
24: \newcommand*{\rpp}{{\rho^{\prime\prime}}}
25: \newcommand*{\amu}{{\mathbf A}^\mu}
26: \newcommand*{\anu}{{\mathbf A}^\nu}
27: \newcommand*{\amunu}{{\mathbf A}^{\mu\nu}}
28: \newcommand*{\vmu}{{\mathbf V}_\mu}
29: \newcommand*{\vmunu}{{\mathbf V}_{\mu\nu}}
30: \newcommand*{\fai}{\mathbf \phi}
31: \newcommand*{\rf}[1]{(\ref{#1})}
32: \newcommand*{\mas}{m_{a_1}^2}
33: \newcommand*{\lag}{{\mathcal L}}
34: \newcommand*{\p}{{\bm p}}
35: \newcommand*{\opava}{Institute of Physics, Silesian University in Opava,
36: Bezru\v{c}ovo n\'{a}m. 13, 746 01 Opava, Czech Republic}
37: \newcommand*{\praha}{Institute of Experimental and Applied Physics,
38: Czech Technical University, Horsk\'{a} 3/a, 120 00 Prague, Czech Republic}
39: 
40: \title{Electron--positron annihilation into four charged pions and the
41: $\bm{a_1\rho\pi}$ Lagrangian}
42: \thanks{This paper is dedicated to the late Julia Thompson,
43: who drew the attention of one of us (P.~L.) to the experimental program
44: of the Budker Institute of Nuclear Physics at Novosibirsk.}
45: 
46: \author{Peter Lichard}
47: \affiliation{\opava}
48: \affiliation{\praha}
49: \author{Josef Jur\'{a}\v{n}}
50: \affiliation{\opava}
51: 
52: \date{\today}% It is always \today, today,
53:              %  but any date may be explicitly specified
54: 
55: \begin{abstract}
56: The excitation curve of $e^+e^-$ annihilation into four charged pions
57: in the $\rho(770)$ region is calculated using three existing models with
58: $\rho$ mesons and pions in intermediate states supplemented by Feynman
59: diagrams with the $a_1(1260)\pi$ intermediate states. A two-term
60: phenomenological Lagrangian of the $a_1\rho\pi$ interaction is used.
61: The mixing angle is determined by fitting the
62: $e^+e^-\rightarrow\pi^+\pi^-\pi^+\pi^-$ cross section data of the
63: Novosibirsk CMD-2 collaboration and also its combination with the low-energy
64: part of the BaBar collaboration data. It is shown that the inclusion of the
65: $a_1\pi$ intermediate states succeeds in obtaining a good agreement with
66: the data on both cross section and the $\rho^0\rightarrow\pi^+\pi^-\pi^+\pi^-$  
67: decay width. When moving to energies above 1~GeV, the $\rho(1450)$ and
68: $\rho(1700)$ resonances are taken into account to get excellent agreement
69: with the BaBar data over the full energy range up to 4.5~GeV.
70: \end{abstract}
71: 
72: \pacs{13.30.Eg, 13.66.Bc, 12.39.Fe, 13.25.Jx}
73: %14.40.Cs
74: \maketitle
75: 
76: \section{\label{sec:intro}Introduction}
77: Task of describing the excitation curve of the $\eepppp$ reaction
78: at low energies is closely related to the investigation of the energy 
79: dependent decay width of the four pion decay of $\rho(770)$. The validity 
80: of the factorization of the cross section into the $\rho(770)$ production
81: and decay parts is generally assumed on the basis of the vector meson
82: dominance (VMD) hypothesis \cite{vmd}. The assumption of factorization allows
83: experimentalists to determine a particular partial decay width at the nominal 
84: $\rho(770)$ mass on the basis of measurement the $\die$ annihilation cross 
85: section into the corresponding final state at the corresponding energy.
86: The current experimental value $\Gamma(\rpppp)=(2.8\pm1.4\pm0.5)$~keV was 
87: obtained in that way \cite{cmd2}.
88: In this work, we will also assume a one-to-one correspondence between
89: the cross section and the energy dependent decay width at the same energy, 
90: ignoring the complications which may appear if some conditions are not met 
91: \cite{app}. Needless to say that the evaluation of the decay width is 
92: less demanding technically and computationally than that of the cross 
93: section (five-dimensional quadrature instead of eight-dimensional one in 
94: the case of four-body final states).
95: 
96: The cross section of the $\die$ annihilation into the $2\pi^+2\pi^-$ and
97: $\pi^+\pi^-2\pi^0$ final states was considered by Decker, Heiliger, Jonsson,
98: and Finkemeier \cite{decker} in conjunction with the CVC-related decays 
99: of the $\tau$ lepton. The intermediate states of their model contained 
100: $\rho(770)$, $a_1(1260)$, and a scalar-isoscalar two-pion resonance.
101: In the two-charged-two-neutral case also the $\omega(782)$ was included.  
102: The $\arp$ vertex factors were adopted from the study of the three-pion
103: decay of the $\tau$ lepton by Isgur, Morningstar, and Reader \cite{isgur}. 
104: Czy\.{z} and K\"{u}hn \cite{czyz} constructed a Monte Carlo generator of 
105: the reaction
106: $\die\ra\gamma+4\pi$. The hadronic matrix elements were used in the form
107: suggested in \cite{decker}, corrected only for some minor deficiencies. To
108: check the soundness of their approach, Czy\.{z} and K\"{u}hn calculated 
109: also the excitation curves of the nonradiative reactions $\eepppp$ and 
110: $\die\ra\pi^+\pi^-2\pi^0$ in qualitative agreement with available data.
111: Ecker and Unterdorfer \cite{ecker3,ecker4} performed the first calculation
112: of the processes $\die\ra 4\pi$ and $\tau\ra\nu_\tau4\pi$ with the correct 
113: structure to $O(p^4)$ in the low energy expansion of the Standard
114: Model extrapolated to the resonance region. To get a good description of
115: the $\eepppp$ cross section up to 1~GeV, they had to include as additional
116: contribution the $a_1$ exchange. They circumvented the $a_1\rho\pi$ Lagrangian
117: ambiguity by choosing a special relation among the individual coupling
118: constants.
119: 
120: The four-pion decays of the $\rho(770)$ are generally considered
121: a convenient test ground of the low-energy effective theories of the
122: interactions of $\rho$ mesons and pions. In the past, several papers
123: appeared that calculated the corresponding partial decay widths
124: \cite{bramon,eidelman,plant,achasov1,achasov2}. Moreover, Achasov
125: and Kozhevnikov \cite{achasov1,achasov2} argued that the four-pion decay
126: widths of $\rho(770)$ are not experimentally well defined because they
127: require the
128: averaging over a mass interval in which they rise rapidly. They therefore
129: calculated, in addition to the decay widths $\Gamma(\rpppp)$,
130: $\Gamma(\rho^0\ra\pi^+\pi^-\pi^0\pi^0)$, $\Gamma(\rho^+\ra\pi^+3\pi^0)$, and
131: $\Gamma(\rho^+\ra2\pi^+\pi^-\pi^0)$, the $\eepppp$ reaction cross
132: section as a function of incident energy (excitation curve) and compared
133: it to the CMD-2 data \cite{cmd2} from Novosibirsk.
134: 
135: With respect to the role of the axial-vector meson $a_1(1260)$ in the
136: four-pion decays of $\rho^0$, the situation is somewhat controversial. 
137: On one side, the intermediate states containing the $a_1(1260)$ were 
138: either ignored \cite{bramon,eidelman,plant,achasov1,achasov2} or shown 
139: \cite{plant} to have little influence on the four-pion decay widths
140: of $\rho(770)$\footnote{Ecker and Unterdorfer \cite{ecker3,ecker4} also
141: included the $a_1$ contribution when calculated the $\rho^0\ra 4\pi$
142: branching fractions, but it is impossible for us to assess its role because
143: they did not show results without it.}.  
144: On the other side, the analysis of the differential
145: distributions of charged pions coming from the $\die$ annihilation
146: in the energy range 1.05--1.38 GeV demonstrated the dominance
147: of the $a_1(1260)\pi$ intermediate states \cite{akhmetshin1999}.
148: Given the large width of $a_1(1260)$ ($\Gamma_{a_1}$=250 to 600 MeV
149: \cite{pdg2006}) it would be surprising if the
150: role of the $a_1$-meson diminished so fast outside the above range, which is
151: not too far from the $\rho(770)$ mass. Moreover, the $a_1(1260)$ meson
152: was shown to be important in the four-pion decays of the $\tau$-lepton
153: \cite{decker,bondar1999,edwards2000}, which are in a sense isospin
154: counterparts of the four-pion final states in the $\die$ annihilation.
155: Recently, the paper of Achasov and Kozhevnikov has appeared \cite{achasov3}
156: that included the intermediate states with the $a_1$ meson using the
157: generalized hidden local symmetry model \cite{bando1988}. This increased the
158: $\rho^0\rightarrow 2\pi^+2\pi^-$ decay width from 0.94~keV to 1.59~keV
159: assuming the nominal mass of the $a_1$ resonance, see Table~I in
160: \cite{achasov3}. Unfortunately, the authors of \cite{achasov3} do not provide 
161: the comparison with the $e^+e^-\rightarrow 2\pi^+2\pi^-$ cross section,
162: which has a greater discriminatory value than the decay width
163: alone \cite{achasov1,achasov2}.
164: 
165: The present work was triggered by our interest in the electromagnetic 
166: probes in relativistic heavy-ion collisions.
167: The production of prompt dileptons and photons is for a long time considered
168: a powerful tool for investigating the properties of dense systems created
169: in the hadronic and nuclear collisions \cite{feinberg, shuryak}. The interest
170: of the heavy ion community in dilepton production has recently been boosted
171: by very precise dimuon data by the CERN/NA60 collaboration \cite{na60}. The
172: prompt dileptons and photons can, in principle, originate from two sources:
173: (i) quark-gluon plasma, (ii) hadron gas. The theoretical calculations of the
174: dilepton and photon yield from the latter are hampered by not uniquely known
175: Lagrangian of the $\arp$ interaction \cite{song,gaogale}. We suggest one way
176: how to relieve this problem and make the predictions of the dilepton and
177: photon production from hadron gas more reliable. The results of the present
178: work have already been utilized \cite{joerg} in the evaluation of the dimuon 
179: production rate in In-In collisions at 158 A GeV.
180: 
181: A way of narrowing the uncertainty interval in dilepton production from 
182: the hadron
183: gas was advocated a long time ago by Li and Gale \cite{ligale}. They checked
184: the feasibility of the dilepton rate calculation in, \eg, $\omega\pi^0$
185: collisions, by comparing the inverse process ($\die$ annihilation into the
186: $\omega\pi^0$ final state) with the available experimental data. Li and Gale
187: also considered the $\die$ annihilation into four pions in the narrow
188: $a_1$ width approximation, i.e. as process $\die\ra a_1\pi$. In the
189: present work we handle it as a genuine 
190: four-final-pion process, considering not only the $a_1\pi$ intermediate
191: states but also other intermediate states following from various
192: models based on the chiral perturbation theory.
193: 
194: In this paper we investigate the role of the $a_1(1260)$ resonance
195: in the $\die$ annihilation into four charged pions and in the 
196: four-charged-pion decays of $\rho(770)$ in more detail. We supplement
197: three existing models, which consider only $\rho$ and $\pi$ in intermediate
198: states (diagrams (a) and (b) in Fig.~\ref{fig:diagrams}), with the $a_1$
199: contribution (diagrams (d) in Fig.~\ref{fig:diagrams}). Those three models
200: are (1) the model of Eidelman, Silagadze, and Kuraev \cite{eidelman},
201: (2) one of the models considered by Plant and Birse \cite{plant}, and
202: (3) the model of Achasov and Kozhevnikov \cite{achasov1,achasov2}.
203: We will consider only the final state with all charged pions, for which
204: the experimental data are best. The difference between the approach of 
205: Achasov and Kozhevnikov \cite{achasov3} and ours lies mainly in the
206: $\arp$ Lagrangian. In \cite{achasov3}, the generalized hidden
207: local symmetry Lagrangian was chosen, whereas we choose a more
208: phenomenological two-term Lagrangian, which is often used in a different
209: branch of the particle physics. Its individual terms and their specific 
210: combinations appeared in many papers computing the dilepton 
211: and photon production rate from a thermalized meson gas. In this paper 
212: we consider the mixing angle between the two terms as a free parameter and 
213: will determine its value by fitting the excitation curve of the 
214: $\eepppp$ reaction.
215: 
216: \begin{figure}
217: \setlength \epsfxsize{8.6cm}
218: \epsffile{fig1.ps}
219: \caption{\label{fig:diagrams}Selected Feynman diagrams describing
220: decay $\rpppp$}
221: \end{figure}
222: 
223: In order to evaluate the amplitude induced by eight diagrams
224: Fig.~\ref{fig:diagrams}(d) we have to
225: choose a Lagrangian of the $\arp$ interaction. For this choice, there are
226: basically two approaches in the literature. One explores well defined
227: theoretical concepts to build dynamical models, the free parameters of
228: which are then fixed by comparison with observed masses and decay widths.
229: See, \eg, 
230: Refs.~\cite{wess,gomm84,holstein,meissner,kaiser,song,korudaz,li,smejkal97}.
231: 
232: In  other, more phenomenological, approaches the authors simply chose
233: for the $\arp$ Lagrangian various expressions built from the field operators 
234: and compatible with the fundamental conservation laws. Such Lagrangians, 
235: after fixing their coupling constants, were then used to calculate various 
236: observable quantities, see, \eg, \cite{pham,kuhnsanta,janssen,haglin}.
237: In some approaches \cite{isgur,xiong}, directly the vertex factors 
238: were written without showing the corresponding Lagrangian. From the fact 
239: that different authors pick different Lagrangians one might get the 
240: impression that the choice of Lagrangian 
241: is not very important and that various Lagrangians lead to identical, or at
242: least similar, results. This is not true, and the observable
243: quantities may be very sensitive to the choice of the $\arp$ Lagrangian, as
244: was demonstrated, \eg, by Song \cite{song}. The discussion of the
245: $a_1$ phenomenology with emphasis on the photon and dilepton production
246: from a hot meson gas can be found in Ref. \cite{gaogale}.
247: 
248: With so many $\arp$ effective Lagrangians, it is
249: interesting to learn which Lagrangian is preferred experimentally. This
250: would require constructing a general Lagrangian with a set of free
251: parameters and fixing them by comparing all possible observables with
252: the existing data. Of course, such a program is very ambitious. In this
253: paper we are going to do something much simpler. Below, we choose a
254: two-component Lagrangian and determine its two free parameters by requiring
255: that the decay width of $a_1(1260)$ be reproduced and the best possible fit
256: obtained for the excitation curve of the $\eepppp$ reaction. Even
257: this restricted program cannot be accomplished completely. Firstly, the
258: width of $a_1(1260)$ is not known reliably. We will consider three values
259: from the range given in the Particle Data Group tables \cite{pdg2006},
260: namely 250, 400, and 600 MeV. Secondly, the result of the fit will depend
261: also on the basic $\rho$ and $\pi$ intermediate state model to which we add
262: the $a_1\pi$ contribution. Nevertheless, we will show that the inclusion of
263: the $a_1\pi$ intermediate states is necessary for obtaining good agreement
264: with the $\eepppp$ excitation curve.
265: 
266: Following the suggestion of T.~Barnes \cite{barnes} we also include the 
267: ratio of the $D$-wave and $S$-wave amplitudes of the $a_1\ra\rho\pi$ decay 
268: as a fitted quantity. The importance of the $D/S$ ratio for
269: selecting among the $\arp$ Lagrangians was stressed in a different context
270: in \cite{gaogale}. We use the experimental value $D/S=-0.14\pm0.11$
271: \cite{DSratio}, which was obtained by genuine partial wave analysis of the
272: reaction $\pi^- p\ra\pi^+\pi^-\pi^- p$. We consider the other values which 
273: exist in literature strongly model biased. They were obtained by fitting
274: the three-pion mass spectrum in the decay $\tau\ra\nu_\tau 3\pi$ using
275: the model of Isgur, Morningstar, and Reader \cite{isgur} and then
276: calculating the $D/S$ ratio from the optimal parameters of the model.
277: 
278: \section{\label{sec:models}Original models, modifications, and additions}
279: As we already stated, we will complement three existing models of the
280: four-pion decays, which consider only intermediate states with $\rho$ mesons
281: and pions, with the intermediate states containing the axial-vector
282: resonance $a_1(1260)$. We must admit that our choice of models is rather
283: arbitrary. Moreover, we took the models as they appeared in the
284: literature and did not check their compatibility with
285: chiral symmetry and with the constraints on coupling of resonances to
286: the pions \cite{ecker1,ecker2}.
287:  
288: Those three models are characterized below.
289: 
290: \subsection{Model of Eidelman, Silagadze, and Kuraev (ESK)}
291: This model \cite{eidelman} is based on the effective chiral Lagrangian
292: by Brihaye, Pak, and Rossi \cite{brihaye}, which was investigated also
293: in \cite{kuraev}. It follows from that Lagrangian that all (a) and (b)
294: diagrams depicted on Fig.~\ref{fig:diagrams} contribute to the $\rpppp$
295: decay rate. Their amplitudes (in the notation slightly different from ours)
296: are shown in the paper. Our
297: usage of this model will differ from the original paper in three respects:
298: (1) We add $a_1$ diagrams Fig.~\ref{fig:diagrams}(d). (2) We use a
299: different value of the parameter $\alpha_k$, defined in \cite{eidelman}.
300: Instead of 0.55 we set $\alpha_k=0.5$,
301: which follows from the KSRF relation \cite{ksrf}, to be in conformity with
302: other two models. (3) We replace the scalar part of the $\rho$-meson 
303: propagator with fixed mass and fixed width by the prescription
304: \be
305: \label{rhoprop} P_\rho(s)=-\frac{i}{s-\rms+i m_\rho\Gamma_\rho(s)},
306: \ee
307: which uses the running mass squared $\rms$ and the energy dependent total
308: width $\Gamma_\rho(s)$ from Ref. \cite{ratio}.
309: 
310: The last point deserves more comments. The denominator of our propagator
311: \rf{rhoprop} is an analytic function in the $s$-plane with a cut running
312: from $4m_\pi^2$ to infinity, as required by general principles. This
313: property differs \rf{rhoprop} from most of the formulas used in the
314: literature. The real function $\rms$ is calculated from $\Gamma_\rho(s)$
315: using a once subtracted dispersion relation, which guarantees that the
316: condition $M_\rho^2(\mrs)=\mrs$ is satisfied. Further condition
317: \be
318: \label{rmsderiv}
319: \left.\frac{d \rms}{d s}\right|_{s=\mrs}=0
320: \ee
321: is not fulfilled automatically and
322: serves as a test that all important contributions to the total $\rho$-meson
323: width $\Gamma_\rho(s)$ have properly been taken into account. See \cite{ratio}
324: for details.
325: If we replace the $m_\rho$ accompanying $\Gamma_\rho(s)$ in Eq.~\rf{rhoprop}
326: by $\sqrt{s}$, as it is done in some existing formulas, the condition
327: \rf{rmsderiv} cannot be satisfied for any reasonable choice of
328: $\Gamma_\rho(s)$. 
329: 
330: The running mass approach \cite{ratio} takes into account, in addition
331: to the basic two-pion decay channel, several channels 
332: ($\omega\pi^0$,  $K^+K^-$, $K^0\bar K^0$, and $\eta\pi^+\pi^-$) which open 
333: as the $\rho$ resonance goes above its nominal mass. It also considers 
334: structure effects described by the strong form factors. In these two respects 
335: it differs from other approaches that appeared in the literature 
336: \cite{GS,VW,melikhov04}. Gounaris and Sakurai \cite{GS} considered only 
337: the two-pion contribution to the total width of the $\rho^0$ resonance 
338: and ignored structure effects. Vaughn and Wali \cite{VW} took into account 
339: the strong form factor, but again ignored higher decay channels. Melikhov, 
340: Nachtmann, Nikonov, and Paulus \cite{melikhov04} included the $K^+K^-$ 
341: and $K^0\bar K^0$ channels, but did not consider the strong form factors.
342: 
343: We will use the $\rho$ propagator \rf{rhoprop} not only in conjunction with
344: the ESK model, but also with the other ones. This is the main reason why our
345: results calculated within the original models (i.e., without the $a_1\pi$
346: intermediate states) and presented below differ slightly from the results
347: quoted in the original papers.
348: 
349: \subsection{One of the models of Plant and Birse (PB/HG)}
350: Plant and Birse \cite{plant} investigated several models of the four-pion 
351: decays of
352: $\rho^0(770)$. One of them (labeled HG) is a corrected version
353: of the model by Bramon, Grau, and Pancheri \cite{bramon}, which was
354: based on the hidden gauge theory of Bando \textit{et al.}
355: \cite{bando1985}. The 2$\pi$2$\rho$ contact terms, see diagram (b1)
356: in Fig.~\ref{fig:diagrams},
357: are missing in this approach. The amplitude of the (a1) diagram is different
358: from that in work by Eidelman, Silagadze, and Kuraev \cite{eidelman} by
359: a factor $(-1/2)$. The amplitudes of (a2) and (b2) diagrams are equal to
360: their ESK counterparts. 
361: 
362: \subsection{Model of Achasov and Kozhevnikov (AK)}
363: Achasov and Kozhevnikov \cite{achasov1,achasov2} studied the four-pion
364: decays of $\rho(770)$, five-pion decays of $\omega(782)$, and the processes
365: related to them. Namely, the $\die$ annihilation into the four- and five-pion
366: final states and the four-pion decays of the $\tau$ lepton. They
367: used the Weinberg Lagrangian \cite{weinberg} obtained upon the nonlinear
368: realization of chiral symmetry. From their rather extensive work we adopt
369: their prescriptions for the amplitudes (a1), (a2), and (b2) of the
370: $\rpppp$ decay. The contact amplitude (b1) is again vanishing.
371: Achasov and Kozhevnikov used a fixed-mass, variable-width formula for the
372: $\rho$-meson propagator.
373: 
374: \subsection{$\bm{\arp}$ Lagrangian and the amplitude of the diagrams
375: containing the $a_1$ meson}
376: We choose the following interaction among the $a_1$, $\rho$, and $\pi$
377: fields
378: \be
379: \label{genlag} \lag=\frac{g_{\arp}}{\sqrt{2}}
380: \left(\lag_1\cos\theta+\lag_2\sin\theta\right),
381: \ee
382: where $g_{\arp}$ and $\theta$ are yet undetermined parameters,
383: \bea
384: \label{lag1}
385: \lag_1&=& \amu\cdot\left(\vmunu\times\partial^\nu{\fai}\right),\\
386: \label{lag2}
387: \lag_2&=& \vmunu\cdot\left(\partial^\mu\anu\times{\fai}\right),
388: \eea
389: and $\vmunu=\partial_\mu{\mathbf V}_\nu-\partial_\nu\vmu$. The isovector
390: composed of the $\rho$-meson field operators is denoted by $\vmu$, similar
391: objects for $\pi$ and $a_1$ are $\fai$ and $\amu$, respectively. We
392: write
393: \bea
394: \phi_1&=&\frac{1}{\sqrt 2}\left(\phi_c+\phi_c^\dagger\right),\nl
395: \phi_2&=&\frac{i}{\sqrt 2}\left(\phi_c-\phi_c^\dagger\right),\nl
396: \phi_3&=&\phi_n, \nonumber
397: \eea
398: and assume that $\phi_c$ contains the annihilation operators of the
399: positive pion and creation operators of the negative pion. $\phi_n$ is
400: the operator of neutral pion field.
401: 
402: A specific combination of terms \rf{lag1} and \rf{lag2} appeared in the
403: pioneering work by Wess and Zumino \cite{wess}. Term \rf{lag1} alone was
404: used by Xiong, Shuryak, and Brown \cite{xiong} in their study of the photon
405: production from meson gas. Janssen, Holinde, and Speth \cite{janssen}
406: picked the term \rf{lag2} when they evaluated the amplitude of the $\pi\rho$
407: scattering. Another combination of \rf{lag1} and \rf{lag2} appeared in
408: the calculation of dilepton production from meson gas by Song, Ko, and Gale
409: \cite{sokoga}.
410: 
411: Lagrangian \rf{genlag} leads to the following factor for
412: the vertex in which an incoming $a_1^+$ (index $\alpha$), an outgoing
413: $\rho^0$ (index $\mu$), and an outgoing $\pi^+$ meet
414: \bea
415: V^{\alpha\mu}\left(p_{a_1},p_\rho,p_\pi\right)&=&\frac{g_{a_1\rho\pi}}
416: {\sqrt{2}}\left\{\cos\theta
417: \left[p_\rho^\alpha p_\pi^\mu-(p_\pi p_\rho)g^{\alpha\mu}\right]\right.\nl
418: &-&\left.\sin\theta
419: \left[p_\rho^\alpha p_{a_1}^\mu-(p_{a_1}p_\rho)g^{\alpha\mu}\right]\right\}.
420: \label{vertex}
421: \eea
422: The $a_1^-\rho^0\pi^-$ vertex acquires an extra minus sign.
423: The evaluation of the decay rate of $a_1^+\ra\rho^0\pi^+$ using vertex
424: \rf{vertex} is straightforward.
425: \bea
426: \Gamma_{a_1^+\ra\rho^0\pi^+}&=&\frac{g^2_{\arp}}
427: {192\pi m_{a_1}^3}
428: \lambda^{1/2}(m_{a_1}^2,m_\rho^2,m_{\pi^+}^2)\nl
429: &\times&R(m_{a_1}^2,m_\rho^2,m_{\pi^+}^2)\ ,
430: \label{gamrhopiplus}
431: \eea
432: where
433: \[
434: \lambda(x,y,z)=x^2+y^2+z^2-2xy-2xz-2yz
435: \]
436: and
437: \begin{eqnarray*}
438: R(x,y,z)&=&\left[(x-y-z)^2+\frac{y}{2x}(x-y+z)^2\right]\cos^2\theta\\
439: &-&2\left[(x-z)^2+y(x+z-2y)\right]\cos\theta\sin\theta\\
440: &+&\left[(x+y-z)^2+2xy\right]\sin^2\theta\ .
441: \end{eqnarray*}
442: If we assume the charge independent $\arp$ coupling constant and masses
443: of $\rho$ and $a_1$, the width of the decay $a_1^+\ra\rho^+\pi^0$ is
444: obtained from \rf{gamrhopiplus} by changing just the pion mass. 
445: Formula 
446: \be
447: \Gamma_{a_1^+}=\Gamma_{a_1^+\ra\rho^0\pi^+}+\Gamma_{a_1^+\ra\rho^+\pi^0}
448: \label{gamrhopi}
449: \ee
450:  enables us to find the coupling constant
451: $g_{\arp}$ for given $\Gamma_{a_1^+}$ and $\sin\theta$.
452: Because each of the diagrams Fig.~\ref{fig:diagrams}(d) contains two $\arp$
453: vertices, the overall sign of the Lagrangian \rf{genlag} is not important
454: and we can assume a non-negative $\cos\theta$.
455: 
456: A question arises whether the narrow $\rho$-width approximation used above 
457: is accurate enough for the purpose of determination the $g_{\arp}$ coupling 
458: constant. Achasov and Kozhevnikov \cite{achasov3} showed that the $a_1$ 
459: decay width calculated as $\Gamma(a_1\ra\rho\pi)$  came out larger 
460: than that calculated as $\Gamma(a_1\ra 3\pi)$ using the same coupling 
461: constant $g_{\arp}$. We have examined this issue, too, using our two-component 
462: Lagrangian \rf{genlag} and got that the
463: $\Gamma(a_1\ra 3\pi)/\Gamma(a_1\ra\rho\pi)$ ratio is smaller than unity
464: (like in \cite{achasov3}) for $\sin\theta\lesssim 0.2$ and 
465: $\sin\theta\gtrsim 0.65$. In the remaining interval, which contains all the 
466: values of $\sin\theta$ that will be met in our calculations, this ratio
467: is greater than one. Moreover, we have found that if one takes into account 
468: the strong form factors in the $\arp$ and $\rho\pi\pi$ vertices, the results 
469: of both approaches become almost identical. This can be explained as
470: follows. According to Kokoski and Isgur \cite{kokoski} formula, which will 
471: be shown later, the strong form factor in a particular decay vertex is 
472: a decreasing function of the three-momentum of an outgoing particle in the 
473: rest frame of the parent particle. In the $a_1\ra3\pi$ decay the intermediate 
474: mass of $\rho$ is mostly smaller than its nominal mass, what means higher 
475: momenta of particles emerging from the $\arp$ vertex. In addition, the 
476: $a_1\ra3\pi$ width is further reduced by the form factor in the
477: $\rho\pi\pi$ vertex.
478: 
479: Let us now turn to the amplitude of the $a_1$ diagrams
480: Fig.~\ref{fig:diagrams}(d) for the decay
481: $\rho^0(p)\ra\pi^-(p_1)\pi^+(p_2)\pi^+(p_3)\pi^-(p_4)$. We first
482: introduce the notation
483: \bea
484: q_i &=& p-p_i, \nl
485: r_{ij}&=&p_i+p_j, \nl
486: s_{ij}&=&r_{ij}^2,
487: \label{sij}
488: \eea
489: and then write the amplitude in the form
490: \[
491: {\mathcal M}_d^{(\lambda)}=\epsilon_\lambda^\mu J_{d,\mu},
492: \]
493: where $\epsilon_\lambda^\mu$ is the polarization vector of the
494: decaying $\rho^0$ and
495: \begin{eqnarray*}
496: J_{d,\mu}&=&-(1-P_{12}P_{34})(1+P_{14})(1+P_{23})V_{\alpha\mu}(-q_4,-p,p_4)\\
497: &\times&P_{a_1}^{\alpha\beta}(q_4)V_{\beta\nu}(q_4,r_{12},p_3)
498: g_\rho(p_2-p_1)^\nu P_\rho(s_{12}).
499: \end{eqnarray*}
500: Here, $P_{ij}$ denotes the operator that interchange four-momenta $p_i$
501: and $p_j$. The axial-vector meson propagator
502: \be
503: \label{a1propagator}
504: P^{\alpha\beta}_{a_1}(q)=i\frac{-g^{\alpha\beta}+\frac{1}{m_{a_1}^2}
505: q^\alpha q^\beta}
506: {q^2-m_{a_1}^2+i m_{a_1}\Gamma_{a_1}}
507: \ee
508: is chosen in a simple fixed-mass, fixed-width form. Here, we are going
509: to consider the four-pion system with invariant energies less than 1~GeV.
510: The invariant mass of the three-pion system, which is equal to $\sqrt{q^2}$,
511: is thus limited by 0.86~GeV. Achasov and Kozhevnikov \cite{achasov3} showed
512: that the $a_1$ decay width is negligible in that energy range. Referring
513: to their finding we set $\Gamma_{a_1}=0$ in Eq.~\rf{a1propagator}.
514: The scalar part of the $\rho$ propagator is again used in the form
515: \rf{rhoprop}.
516: 
517: \subsection{Technicalities}
518: 
519: The complete amplitude of the $\rpppp$ decay is
520: \be
521: \label{ampl}
522: {\mathcal M}^{(\lambda)}=\epsilon_\lambda^\mu J_\mu,
523: \ee
524: where $\epsilon_\lambda^\mu$ is the polarization vector of the
525: decaying $\rho^0$ and
526: \[
527: J_\mu=J_{a,\mu}+J_{b,\mu}+J_{d,\mu}.
528: \]
529: Four-vectors $J_{a,\mu}$ and $J_{b,\mu}$ describe the contributions from (a)
530: and (b) diagrams in a
531: particular model and $J_{d,\mu}$ is the contribution from (d) diagrams.
532: The sum over the $\rho$-meson polarizations of the amplitude \rf{ampl} squared is given by
533: \be
534: \label{amplsq}
535: \sum_\lambda\left|{\mathcal M}^{(\lambda)}\right|^2
536: =\left(-g^{\mu\nu}+\frac{p^\mu p^\nu}{m_\rho^2}\right)
537: J_\mu J_\nu^*.
538: \ee
539: 
540: This formula is more complicated than that used in \cite{eidelman}, because
541: the four-vectors $J_{a,\mu}$ and $J_{b,\mu}$ of PB/HG and AK models do
542: not satisfy the transversality condition $J_\mu p^\mu=0$. We used the
543: algebraic manipulation program {\textsc REDUCE} \cite{reduce} to express
544: the sum \rf{amplsq} in terms of six invariants $s_{ij}$, $i<j$, $j=2,3,4$
545: defined in \rf{sij}. Of course, only five of them are independent and
546: we used the identity $\sum_{i<j}s_{ij}=m_\rho^2+8m_\pi^2$ for the checks
547: in the process of evaluation of the decay width. When calculating the
548: excitation curve, $m_\rho^2$ is replaced by $s$, the square of the incident
549: energy.
550: 
551: When calculating the decay width of an unpolarized parent particle,
552: we may take advantage of the spherical symmetry of the problem and choose
553: the following kinematic configuration: (1) The parent particle $a$ is at
554: rest. (2) The summed momentum $\p_{12}$ of particles 1 and 2 points in the
555: direction of the $z$ axis. (3) The individual momenta $\p_1$ and $\p_2$ lie
556: in the $xz$ plane. Then the following formula, written in a general case
557: with arbitrary masses and spins, is valid
558: \bea
559: \label{gamma4}
560:  \Gamma &=& \frac{N}{16(2\pi)^6m_a^2}\int_{m_1+m_2}^{m_a-m_3-m_4} d m_{12} \
561:  p_1^*\int_{m_3+m_4}^{m_a-m_{12}} d m_{34}\nl &\times& p_{12} p_3^*
562:  \int_{-1}^1 d\cos\theta_1^*\int_{-1}^1 d\cos\theta_3^* \int_0^{2\pi} d \varphi_3
563:  \overline{\left|{\mathcal M}\right|^2}.
564: \eea
565: The last quantity is the amplitude squared, averaged over the initial
566: spin states, and summed over the final spin states. The factor $N$ takes into
567: account the identity of the final particles and equals 1/4 in our case.
568: The asterisk denotes the momentum in the corresponding rest frame (1-2 or 3-4),
569: $p=|\mathbf{p}|$, and $m_{ij}=\sqrt{s_{ij}}=\sqrt{(p_i+p_j)^2}$.
570: 
571: For evaluation of the integrals in \rf{gamma4} we used a sequence of the
572: five one-dimensional Gauss-Legendre quadratures of the sixteenth order. 
573: We prefer this method to Monte-Carlo integration because we use the result 
574: of the integration in a minimization procedure and therefore we require that
575: the same value of the optimized variable (Lagrangian mixing angle) yield
576: always the same value of the minimized function, which would not be
577: satisfied with the Monte-Carlo integration. Nevertheless, we checked our
578: computer code by evaluating the decay width for a particular value of the
579: mixing angle using a completely independent code based on the Monte-Carlo
580: method.
581: 
582:  To convert the calculated decay width into the cross section, we start
583: with the formula
584: \[
585: \sigma_{4\pi}(s)=\frac{\sigma_{\pi^+\pi^-}(s)}{\Gamma_{\pi^+\pi^-}(s)}
586: \Gamma_{4\pi}(s).
587: \]
588: Using
589: \[
590: \sigma_{\pi^+\pi^-}(s)=\frac{\pi\alpha^2}{3s}
591: \left(1-\frac{4m_\pi^2}{s}\right)^{3/2}\left|F_\pi(s)\right|^2,
592: \]
593: where $F_\pi(s)$ is the contribution of the $\rho$ resonance to the
594: pion form factor, and
595: \[
596: \Gamma_{\pi^+\pi^-}(s)=\frac{g_\rho^2W}{48\pi}
597: \left(1-\frac{4m_\pi^2}{s}\right)^{3/2}
598: \]
599: with $W=\sqrt{s}$, we arrive at
600: \be
601: \sigma_{4\pi}(s)=
602: \left(\frac{4\pi\alpha}{g_\rho}\right)^2\frac{1}{W^3}
603: \left|F_\pi(s)\right|^2\Gamma_{4\pi}(s).
604: \label{convert}
605: \ee
606: We further use the VMD expression for the dielectron
607: decay width of $\rho^0$
608: \be
609: \Gamma_{\die} = \frac{4\pi m_\rho}{3}\left(\frac{\alpha}{g_\rho}\right)^2
610: \label{rhoee}
611: \ee
612: and get
613: \be
614: \label{aksigma}
615: \sigma_{4\pi}(s)=
616: \frac{12\pi\Gamma_{\die}}{m_\rho W^3}
617: \left|F_\pi(s)\right|^2 \Gamma_{4\pi}(s).
618: \ee
619: If we set, following Achasov and Kozhevnikov \cite{achasov2},
620: \be
621: \label{akpiff}
622: F_\pi(s)=\frac{m_\rho^2}{D_\rho(s)},
623: \ee
624: where the inverse $\rho$-meson propagator $D_\rho(s)$ is defined in
625: Eq.~(2.2) of \cite{achasov2}, we reproduce their Eq.~(3.1).
626: 
627: In our opinion, Eq.~\rf{aksigma} overestimates the
628: cross section if the experimental value of $\Gamma_{\die}$ is used.
629: The reason is that
630: the dielectron decay width calculated from \rf{rhoee} is smaller than
631: the experimental value. We will therefore stick with formula \rf{convert}.
632: 
633: We utilize the scalar part of the $\rho$-meson propagator
634: \rf{rhoprop} to write our Ansatz for the $\rho$-meson contribution to the
635: pion form factor
636: \be
637: \label{piff}
638: F_\pi(s)=\frac{M_\rho^2(0)}{M_\rho^2(s)-s-im_\rho\Gamma_\rho(s)}.
639: \ee
640: As shown in \cite{ratio}, this formula gives the correct value of the
641: mean square radius of the pion. The form factor \rf{akpiff}
642: fails in this test.
643: 
644: For the $\rho$ coupling constant we use the same value as in
645: \cite{plant,achasov1,achasov2}, namely $g_\rho=5.89$. This value is 
646: compatible with what follows from the KSRF relation \cite{ksrf} 
647: ($5.900\pm0.011$). Both values are little lower than $g_\rho=6.002\pm0.015$ 
648: calculated from the $\rho$-meson width.
649: 
650: \section{Low-energy results ($\bm{W<1\ \mathrm{GeV}}$)}
651: \label{lowenergy}
652: We deal with the excitation curves of the reaction $\eepppp$ calculated
653: in three different models (ESK, PB/HG, AK) supplemented with the $a_1$
654: diagrams Fig.~\ref{fig:diagrams}(d). We first fit them to the CMD-2 data
655: \cite{cmd2} by varying
656: the sine of the mixing angle $\theta$, defined in \rf{genlag}, for the
657: three fixed values of the width of the $a_1(1260)$ meson. We did not
658: consider the first two points in the CMD-2 data, because they give only
659: upper bounds of the cross section. The ratios of the usually defined 
660: $\chi^2$ to the
661: number of degrees of freedom (NDF), which characterize the quality
662: of the fit, are shown in Table~\ref{tab:chiscmd2}. The last row in
663: Table~\ref{tab:chiscmd2} shows the values of $\chi^2$/NDF that indicate 
664: how well (or badly) the original
665: models without $a_1$ agree with the data. No free parameter is involved
666: in the latter case. Table~\ref{tab:chiscmd2} shows that
667: the ratio $\chi^2/$NDF is always greater than one. In what
668: follows, we shall therefore multiply the statistical errors of the 
669: quantities obtained in the process of minimization by the square root of
670: that ratio. 
671: \begin{table}
672: \caption{\label{tab:chiscmd2}$\chi^2/$NDF of the fits to the CMD-2 cross
673: section data (11 data points)}
674: \begin{ruledtabular}
675: \begin{tabular}{ccccc}
676: $ \Gamma_{a_1}$ & ESK  & PB/HG & AK   & only $a_1$ \\
677:  (MeV) & \cite{eidelman} & \cite{plant} & \cite{achasov1,achasov2} & \\
678: \hline
679: 250 &   1.60   &   1.34   &   1.28   &   1.68   \\
680: 400 &   1.53   &   1.37   &   1.30   &   1.82   \\
681: 600 &   1.61   &   1.41   &   1.31   &   1.94   \\
682: \hline
683:  Only $\rho$, $\pi$& 17.6  & 15.0 & 14.8 &  / \\
684: \end{tabular}
685: \end{ruledtabular}
686: \end{table}
687: 
688:  The inspection of Table~\ref{tab:chiscmd2} shows that the inclusion of
689: the $a_1$ contribution greatly improves the agreement with the data.  The
690: interference between the original diagrams and the new ones is important,
691: the results of the combined model are better than those of the $a_1$ diagrams
692: alone. The best results (lowest $\chi^2$) are obtained with the AK model 
693: supplemented with the $a_1\pi$ intermediate states.
694: 
695: To investigate the sensitivity of our results to the input data,
696: we combine the CMD-2 data \cite{cmd2} and the low-energy ($s<1$~GeV$^2$) 
697: part of the BaBar data \cite{babar} (BaBar-LE in what follows) into a new 
698: set and repeat the calculations.
699: The results are shown in Table~\ref{tab:chiscomb}. Their comparison
700: with the results obtained from the CMD-2 data alone shows two important
701: differences: (i) The agreement of all models with data, characterized 
702: by $\chi^2$/NDF, is now better. It indicates that the CMD-2 and BaBar-LE 
703: data are compatible, so the increased number of data points does not bring
704: proportional increase of $\chi^2$. (ii) Whereas merging of the $a_1$
705: diagrams with the PB/HG or AK model improves the fit, adding the ESK model
706: diagrams to the pure $a_1$ contribution leads to the opposite effect.
707: 
708: \begin{table}
709: \caption{\label{tab:chiscomb}$\chi^2$/NDF of the fits to
710: the combined CMD-2 \& BaBar-LE cross section data (27 data points)}
711: \begin{ruledtabular}
712: \begin{tabular}{ccccc}
713: $ \Gamma_{a_1}$ & ESK   & PB/HG  & AK   & only $a_1$ \\
714:  (MeV)        &\cite{eidelman} &\cite{plant} &\cite{achasov1,achasov2} & \\
715: \hline
716: 250 &   1.42   &   1.20   &   1.19   &   1.32   \\
717: 400 &   1.48   &   1.21   &   1.18   &   1.39   \\
718: 600 &   1.55   &   1.22   &   1.19   &   1.44   \\
719:  \hline
720: Only $\rho$, $\pi$& 9.4 & 10.4 & 10.3 &  / \\
721: \end{tabular}
722: \end{ruledtabular}
723: \end{table}
724: 
725: Next, we add the $D/S$ ratio \cite{DSratio} to the set of fitted
726: experimental values and repeat the calculations. The results are shown 
727: in Table \ref{tab:chiscomb_ds}.
728: \begin{table}
729: \caption{\label{tab:chiscomb_ds}$\chi^2$/NDF of the fits to the CMD-2
730: \& BaBar-LE cross section data and to the $D/S$ ratio (28 data points)}
731: \begin{ruledtabular}
732: \begin{tabular}{ccccc}
733:  $ \Gamma_{a_1}$ & ESK             & PB/HG           &
734: AK              & only $a_1$ \\
735:    (MeV)        & \cite{eidelman} & \cite{plant}     &
736: \cite{achasov1,achasov2} &          \\
737:  \hline
738: 250 &   3.66   &   1.95   &   1.99   &   1.96   \\
739: 400 &   1.98   &   1.34   &   1.33   &   1.48   \\
740: 600 &   1.65   &   1.20   &   1.18   &   1.41   \\
741: %Only $\rho$, $\pi$& 10.5 & 10.4 & 10.3 &  / \\
742: \end{tabular}
743: \end{ruledtabular}
744: \end{table}
745: The salient feature of those results is a clear preference of the
746: highest assumed value (600~MeV) of the total $a_1$ width.
747: 
748: In Figs.~\ref{fig:esk}, \ref{fig:pbhg}, and \ref{fig:ak}
749: we show the comparison of the combined set of data with
750: the excitation curves calculated in all three
751: models supplemented with the $a_1$ diagrams Fig.~\ref{fig:diagrams}(d). 
752: The same comparison for the $a_1$ diagrams alone is depicted in 
753: Fig.~\ref{fig:a1}.
754: \begin{figure}
755: \setlength \epsfxsize{8.6cm}
756: \epsffile{fig2.ps}
757: \caption{\label{fig:esk}Excitation curves calculated in the original
758: (without $a_1$ meson) and expanded ESK model compared to the CMD-2 
759: and BaBar-LE data. The $D/S$ ratio was also used in fit.}
760: \end{figure}
761: 
762: \begin{figure}
763: \setlength \epsfxsize{8.6cm}
764: \epsffile{fig3.ps}
765: \caption{\label{fig:pbhg}Excitation curves calculated in the original
766: (without $a_1$ meson; dash-dotted curve close to the abscissa) and expanded
767: PB/HG model compared to the CMD-2 and BaBar-LE data. The $D/S$ ratio was also
768: used in fit.}
769: \end{figure}
770: 
771: \begin{figure}
772: \setlength \epsfxsize{8.6cm}
773: \epsffile{fig4.ps}
774: \caption{\label{fig:ak}Excitation curves calculated in the
775: original (without $a_1$ meson; dash-dotted curve close to the
776: abscissa) and expanded AK model compared to the CMD-2 and
777: BaBar-LE data. The $D/S$ ratio was also used in fit.}
778: \end{figure}
779: 
780: \begin{figure}
781: \setlength \epsfxsize{8.6cm}
782: \epsffile{fig5.ps}
783: \caption{\label{fig:a1}Excitation curves calculated from
784: the $a_1\pi$ diagrams only, compared to the CMD-2 and
785: BaBar-LE data. The $D/S$ ratio was also used in fit.}
786: \end{figure}
787: 
788: In Table~\ref{tab:sthcombds} we can see the values of
789: $\sin\theta$ together with their errors (defined in the usual way
790: \cite{minuit}) obtained from the fit to the CMD-2 \& BaBar-LE cross
791: section data and to the $D/S$ ratio. As we mentioned above, three
792: different values of the $a_1(1260)$ width are assumed. 
793: Table~\ref{tab:gamcombds} compares the experimental value of the 
794: $\rpppp$ decay width \cite{cmd2} with the results obtained from various 
795: models 
796: under the same conditions. The results for both quantities obtained with
797: other data sets (CMD-2 only, CMD-2 \& BaBar-LE) are very similar.
798: 
799: \begin{table}
800: \caption{\label{tab:sthcombds}Values of $\sin\theta$ from the fit to the
801: CMD-2 \& BaBar-LE data and to the $D/S$ ratio.}
802: \begin{ruledtabular}
803: \begin{tabular}{ccccc}
804: $\Gamma_{a_1}$ & ESK & PB/HG & AK & only $a_1$ \\
805:    (MeV) & \cite{eidelman} & \cite{plant} & \cite{achasov1,achasov2}  & \\
806: \hline
807: 250 &   0.4092(33)   &   0.4278(32)   &   0.4267(32)   &   0.4312(35) \\
808: 400 &   0.4352(24)   &   0.4624(34)   &   0.4608(32)   &   0.4679(39) \\
809: 600 &   0.4659(27)   &   0.5046(44)   &   0.5022(41)   &   0.5132(55) \\
810: \end{tabular}
811: \end{ruledtabular}
812: \end{table}
813: \begin{table}
814: \caption{\label{tab:gamcombds}Decay width $\Gamma(\rpppp)$ (keV)
815: calculated in various models using $\sin\theta$ from the fits to the
816: CMD-2 \& BaBar-LE data and to the $D/S$ ratio. 
817: Experimental value is $(2.8\pm 1.4\pm 0.5)$~keV \cite{cmd2}.}
818: \begin{ruledtabular}
819: \begin{tabular}{ccccc}
820: $ \Gamma_{a_1}$ & ESK  & PB/HG & AK  & only $a_1$ \\
821: (MeV)    & \cite{eidelman} & \cite{plant} & \cite{achasov1,achasov2} & \\
822: \hline
823: 250 &   4.28(01)   &   3.16(25)   &   2.70(23)   &   4.52(30)   \\
824: 400 &   2.81(01)   &   3.55(28)   &   3.03(26)   &   5.08(32)   \\
825: 600 &   1.94(02)   &   3.77(30)   &   3.22(27)   &   5.39(37)   \\
826:  \hline
827: Only $\rho$, $\pi$& 16.2   &   0.59   &   0.89  &  / \\
828: \end{tabular}
829: \end{ruledtabular}
830: \end{table}
831: 
832: \section{High-energy results ($\bm{W}$ up to 4.5~$\bm{\mathrm{GeV}}$)}
833: When we want to get a good description of the data on the $\die$
834: annihilation to four charged pions at energies above 1~GeV, we should
835: consider also the contribution from diagrams where higher $\rho$ 
836: resonances couple to the virtual photon and then convert into four
837: pions. We include two resonances: $\rp=\rho(1450)$ and $\rpp=\rho(1700)$.
838: We assume that the decay of those resonances into four pions is governed
839: by the same Feynman diagrams as that of $\rho(770)$, with all coupling
840: constants scaled by the same factor (different for $\rp$ and $\rpp$).
841: This assumption implies that the four-pion decay widths of $\rp$ and $\rpp$ 
842: have the same shape in $W$ as that of $\rho(770)$. They only differ from it
843: by constant factors. The simplifying assumption we have made allows us to 
844: use the same cross section formula \rf{convert} as in the
845: low energy case, with $F_{\pi}(s)$ replaced by
846: \be
847: \label{heformfactor}
848: F(s)=F_\rho(s)+\delta F_\rp(s)+\epsilon F_\rpp(s),
849: \ee
850: where $F_\rho(s)$ differs from \rf{piff} by including also 
851: $\Gamma_{\rpppp}(s)$ into the total decay width of $\rho(770)$, 
852: \be
853: F_\rp(s)=\frac{m_\rp^2}{m_\rp^2-s-im_\rp\Gamma_\rp},
854: \ee
855: and similar expression holds also for $F_\rpp(s)$. Constants $\delta$ and 
856: $\epsilon$ not only include the coupling constants modification factors 
857: mentioned above,
858: but also account for the fact that the couplings of $\rp$ and $\rpp$ to
859: photon differ from that of $\rho(770)$, which is fixed by VMD. 
860: Unknown complex parameters $\delta$ and $\epsilon$ will be
861: determined, together with the masses and widths of $\rp$ and $\rpp$ 
862: and other two parameters mentioned later on, by fitting the experimental 
863: excitation function\footnote{When we replaced \rf{piff} by \rf{heformfactor}
864: in the low-energy region and kept the form-factor parameters as determined 
865: in the high-energy fit, we got the results that differed slightly from that 
866: in Sec.~\ref{lowenergy}. If we varied also those parameters when fitting 
867: the low-energy data, they acquired unphysical values (masses of $\rp$ and 
868: $\rpp$ around 1~GeV). It may signify that some contribution important at 
869: low energies (scalar resonances) is still missing in our approach.}.
870: 
871: Another effect that has to be taken into account when dealing with higher
872: energies is connected with the structure of the strongly interacting
873: particles. Our decay amplitudes have been derived under the assumption that
874: the pions, $\rho$'s, and $a_1$'s are elementary quanta of the corresponding
875: quantum fields. But this assumption is justified only when their mutual
876: interaction is soft. When the momenta of the mesons which enters a specific
877: interaction vertex get higher, the contribution of that vertex to the
878: amplitude becomes smaller than in the case of the point-like
879: participants. This effect is usually described by strong form factors.
880: Given the present status of the strong interaction theory, we have to
881: turn to models. For example, in the chromoelectric flux-tube 
882: breaking model of Kokoski and Isgur \cite{kokoski}, the vertex describing
883: a two-body decay is modified by the factor $\exp\{-{p^*}^2/(12\beta^2)\}$, 
884: where $p^*$ is the three-momentum magnitude of the decay products in the 
885: parent particle rest frame and $\beta\approx 0.4$~GeV. For decay of the 
886: $\rho$ meson with a (non-nominal) mass $W$ into two on-mass-shell pions, 
887: this form factor can be written as
888: \be
889: \label{kiff}
890: F_{KI}(s)=\exp\left\{-\frac{s-s_0}{48\beta^2}\right\},
891: \ee
892: where $s=W^2$ and $s_0$ is the threshold value of $s$ ($4m_\pi^2$ in
893: the two-pion decay). The complete amplitude of the four-pion 
894: decay of $\rho^0$ contains many vertices, some of them with more than
895: three incoming/outgoing particles. Applying the Kokoski-Isgur factor
896: to each of them would be cumbersome and would require additional assumptions
897: in the case of more complicated vertices. We will therefore assign an 
898: ``effective'' strong form factor of the form \rf{kiff} to the complete 
899: amplitude of the decay $\rpppp$, but with $s_0=16m_\pi^2$. When fitting 
900: the experimental excitation curve, $\beta$ will be considered as another 
901: parameter. With the masses and widths of $\rp$ and $\rpp$, complex 
902: parameters $\delta$ and $\epsilon$, and with the sine of the mixing angle 
903: $\theta$ there are ten real parameters to be determined by fitting the 
904: $\eepppp$ cross section data of BaBar collaboration \cite{babar}. 
905: 
906: In the following, we assume the $a_1$ decay width of 600~MeV,
907: for which the results of all models at energies below 1~GeV were best.
908: The same value was used in the $a_1$ propagator \rf{a1propagator}. 
909: 
910: The resulting optimized values of the parameters listed above
911: and their MINUIT \cite{minuit} errors are shown in 
912: Table~\ref{tabbabar} for the three different models supplemented with 
913: the $a_1\pi$ intermediate states and for the latter alone. Mutual 
914: comparison of the $\chi^2/$NDF ratios clearly
915: shows that the presence of the $a_1\pi$ intermediate states is crucial
916: for obtaining good agreement of the calculated excitation curve with
917: data. They provide a good fit even if taken alone. Adding the diagrams 
918: with $\pi$'s and $\rho$'s in the intermediate states does not change
919: the quality of the fit if their amplitudes are taken from the PB/HG and
920: AK models. On the other hand, the inclusion of the amplitudes of the 
921: ESK model brings some deterioration of the fit.
922: The values of parameter $\beta$ do not differ very much from the value 
923: advocated in \cite{kokoski} for the three-line vertices, what indicates 
924: that the effective-strong-form-factor approach we have chosen \rf{kiff} is
925: reasonable. The values of 
926: the sine of the mixing angle $\theta$ are somewhat lower than those at
927: low energies. The central values of the masses and widths of $\rp$ and
928: $\rpp$ a little different from those listed in \cite{pdg2006}, but 
929: differences are acceptable keeping in mind relatively large errors.
930: 
931: The graphical comparison of the excitation curve calculated from
932: the model containing only the diagrams with the $a_1\pi$ intermediate 
933: states, shown in Fig.~\ref{fig:diagrams}(d), with experimental data 
934: \cite{babar} is presented in Fig.~\ref{fig:babar}.
935: The excitation curves of other models (ESK, PB/HG, AK) combined with the 
936: $a_1\pi$ contribution differ only slightly and are not shown. 
937: 
938: \begin{table}
939: \caption{\label{tabbabar}
940: Results of the fit to the BaBar cross section data and to the $D/S$ ratio
941: ($145$ data points) for $\Gamma_{a_1}=600$~MeV.}
942: \begin{ruledtabular}
943: %\begin{tabular}{lll}
944: \begin{tabular}{ccccc}
945: Model & ESK   & PB/HG  & AK    & only $a_1\pi$ \\
946:       & \cite{eidelman} & \cite{plant} & \cite{achasov1,achasov2}  & \\
947: \hline
948:  $\chi^2/$NDF    & 1.21       & 1.12       & 1.12       & 1.12       \\
949:  $\sin\theta$    & 0.4474(22) & 0.4592(28) & 0.4588(27) & 0.4603(28) \\
950:  $\beta$ (GeV)   & 0.3505(89) & 0.3665(97) & 0.3657(97) & 0.3695(98) \\
951:  $m_{\rp}$ (GeV) & 1.419(12)  & 1.439(13)  & 1.438(13)  & 1.442(13)  \\
952:  $\Gamma_{\rp}$ (GeV)& 0.564(20)& 0.568(21)& 0.568(21)  & 0.566(21)  \\
953:  Re($\delta$)    & 0.1038(41) & 0.1145(51) & 0.1144(51) & 0.1149(53) \\ 
954:  Im($\delta$)    &-0.039(11)  &-0.019(12)  &-0.021(12)  &-0.015(13)  \\ 
955:  $m_{\rpp}$ (GeV) & 1.903(21) & 1.923(24)  & 1.922(24)  & 1.926(24)  \\
956:  $\Gamma_{\rpp}$ (GeV)& 0.247(38)& 0.284(44)& 0.283(44) & 0.290(45)  \\
957:  Re($\epsilon$)&-0.0016(11) &-0.0002(17) &-0.0003(17) & 0.0002(18) \\
958:  Im($\epsilon$)&-0.00373(94) &-0.0054(12) &-0.0054(12) &-0.0056(13)\\
959: \end{tabular}
960: \end{ruledtabular}
961: \end{table}
962: 
963: \begin{figure}
964: \setlength \epsfxsize{8.6cm}
965: \epsffile{fig6.ps}
966: \caption{\label{fig:babar}Theoretical excitation curve compared with 
967: the BaBar data \cite{babar}. The $D/S$ ratio was also used in fit. The 
968: result of the pure $a_1$ model is shown. The other models combined with 
969: $a_1\pi$ intermediate states provide almost identical curves.}
970: \end{figure}
971: 
972: 
973: \section{Conclusions and comments}
974: 
975: Our low-energy results show that the inclusion of the $a_1\pi$ intermediate 
976: states
977: is of vital importance for obtaining a good agreement with the experimental 
978: data on the cross section of the reaction $\eepppp$ as a function of the 
979: incident energy (see Tables~\ref{tab:chiscmd2}, \ref{tab:chiscomb}, and 
980: \ref{tab:chiscomb_ds} or Figs.~\ref{fig:esk}, \ref{fig:pbhg}, and 
981: \ref{fig:ak}). The $\chi^2/$NDF ratio gets much smaller if a particular 
982: model is supplemented with the diagrams containing the $a_1$ resonance 
983: in the intermediate states. Viewing from another perspective, the
984: pure $a_1$ model provides relatively good agreement with the cross
985: section data, much better than each of the three models without the $a_1\pi$
986: intermediate states. See Tables mentioned above and Fig.~\ref{fig:a1}.
987: Adding the diagrams from the original models to the pure $a_1$ model improves
988: the fit in the case of the PB/HG and AK models, but worsens it in the
989: case of the ESK model. Unfortunately, the $\chi^2/$NDF ratio remains
990: greater than one everywhere. It may be the consequence of our ignoring some 
991: important contributions, perhaps those with a scalar resonance considered in 
992: \cite{decker,czyz}. The original models ESK, PB/HG, and AK do not contain 
993: any scalar resonances. We also have not included them as our main concern 
994: was the role of the $a_1$ resonance. It is also possible that the $\arp$
995: Lagrangian should contain more terms than considered in \rf{genlag}.
996: 
997: Using the $D/S$ ratio as an additional data point is important. It can 
998: discriminate among various models. In our case it increases the separation 
999: of the ESK model from the others. It also strongly prefers larger
1000: values of the assumed $a_1$ width. In calculations without the $D/S$ ratio
1001: we were able to find a value of $\sin\theta$ for each $\Gamma_{a_1}$ that 
1002: led to an acceptable fit to the $e^+e^-\rightarrow 2\pi^+ 2\pi^-$
1003: cross section. However, the lowest value of $\Gamma_{a_1}$ is excluded
1004: if we add the $D/S$ ratio to the fitted data (see 
1005: Table~\ref{tab:chiscomb_ds}).
1006: 
1007: As to the partial decay width of $\rpppp$, the conclusion is not so
1008: categorical. Two models (PB/HG and AK) in their original forms provided 
1009: results that were a little smaller, but did not contradict strongly the 
1010: experimental value with its large errors. Only the original ESK model 
1011: gave too large figure, which was in a clear disagreement with the 
1012: experimental value. The inclusion of the $a_1\pi$ intermediate states 
1013: brought all values into the interval given by the experimental value 
1014: and its errors summed linearly, see Table~\ref{tab:gamcombds}. It must
1015: be said that the pure $a_1$ model gives the decay widths that are
1016: close to the one-sigma upper limit or even beyond it. 
1017: 
1018: We originally hoped that our study would tell us the form of the $\arp$
1019: Lagrangian. But with respect to the $\arp$ Lagrangian, no clear picture 
1020: can be inferred from the low-energy results yet. The optimal values of the 
1021: sine of the mixing angle, see Table~\ref{tab:sthcombds}, are from a 
1022: broad interval and depend not only on the choice of the original model
1023: to which the $a_1$ diagrams are added, but also on the assumed value of 
1024: the $a_1$ width. The optimized values of $\sin\theta$ squeeze into interval
1025: $(0.40,0.51)$. 
1026: 
1027: The quality of the fit over the whole energy range of the BaBar experiment 
1028: \cite{babar}, as measured by the $\chi^2/$NDF ratio, seems to be better
1029: than that at low energies. But a more careful investigation in terms of
1030: the confidence level, which takes into account $\chi^2$ and NDF separately,
1031: shows equal quality of those two fits. The values of $\sin\theta$
1032: are compatible with those found at low energies,
1033: $\sin\theta\in(0.41,0.47)$ (lower boundary is obtained from the fits
1034: where $\Gamma_{a_1}=250$ MeV was used). They occupy a narrower
1035: interval, what can be explained by lesser importance of sub-dominant
1036: diagrams with only $\rho$ and $\pi$ mesons in the intermediate states.
1037: 
1038: It is interesting to compare our estimates of the mixing parameter 
1039: $\sin\theta$, defined in Eq.~\rf{genlag}, with its values that have been 
1040: used so far, see Table~\ref{tab:sinus}. The values in the first and fifth 
1041: row simply
1042: reflect the fact that Lagrangians in Refs.~\cite{xiong,haglin,janssen}
1043: contained just one term. The remaining rows refer to various sets
1044: of four fundamental parameters ($m_0$, $g$, $\sigma$, and $\xi$) of the 
1045: model in which the vector and axial-vector mesons were included as massive 
1046: Yang-Mills fields of the SU(2)$\times$SU(2) chiral symmetry \cite{song}. 
1047: The model built on previous works \cite{gomm84,holstein,meissner}. 
1048: Our mixing parameter is related to the parameters $\eta_1$ and $\eta_2$ 
1049: of that model, which can be expressed in terms of the four fundamental 
1050: parameters using Eqs. (2.9) and (2.10) in \cite{song}. The formula is very 
1051: simple
1052: \[
1053: \sin\theta=\frac{\eta_2}{\sqrt{\eta_1^2+\eta_2^2}}.
1054: \] 
1055: In \cite{song}, the fundamental parameters of the massive Yang-Mills model
1056: were determined using the experimental values of the masses and width of
1057: the $\rho$ and $a_1$ mesons. This procedure is not unique, there are
1058: two solutions. The corresponding mixing parameter is shown in the second
1059: and fourth row of Table~\ref{tab:sinus}. Row 3 corresponds to an 
1060: \textit{ad hoc} choice of fundamental parameters made in \cite{turbide1}.
1061: Last two rows show the range of our results obtained from various models
1062: (ESK, PB/HG, AK) and various assumed values of $\Gamma_{a_1}$ (250, 400, 
1063: and 600 GeV). Unfortunately, there is no overlap with the rows above.
1064: The issue definitely requires more attention. The phenomenological models
1065: may be improved by including other intermediate states, perhaps those
1066: with scalar resonances. The parameters of the massive Yang-Mills model
1067: may be tuned by using richer experimental input ($\Gamma(a_1\ra\pi\gamma)$
1068: and $D/S$ ratio as in \cite{korudaz,gaogale}) and more realistic 
1069: formulas for relating theoretical parameters to experimental quantities,
1070: \eg, calculating the $a_1$ width as $a_1\ra 3\pi$ instead of
1071: $a_1\ra\rho\pi$.
1072:  
1073: \begin{table}
1074: \caption{\label{tab:sinus}
1075: Survey of the mixing parameter $\sin\theta$ for various versions of
1076: two-component $\arp$ Lagrangian \rf{genlag} that appeared in
1077: literature.}
1078: \begin{ruledtabular}
1079: \begin{tabular}{ccc}
1080: No. & $\sin\theta$ & Reference \\
1081: \hline
1082: 1 & 0       & \cite{xiong,haglin} \\
1083: 2 & 0.2169  & \cite{song,gaogale,sokoga} \\
1084: 3 & 0.5582  & \cite{turbide1} \\
1085: 4 & 0.6308  & \cite{song,gaogale,turbide2} \\
1086: 5 & 1       & \cite{janssen} \\
1087: \hline
1088:   & 0.40--0.51 & our low-energy fits \\
1089:   & 0.41--0.47 & our all-energy fits \\
1090: \end{tabular}
1091: \end{ruledtabular}
1092: \end{table}
1093: 
1094: Our failure to obtain a more precise value of the mixing angle of the
1095: $\arp$ Lagrangian suggests that it is necessary to make a simultaneous
1096: fit to data about several physical processes. The natural candidates are
1097: the $\die$ annihilation into various four-pion final states, the decay
1098: of the $\tau$ lepton into neutrino and three or four pions, and the
1099: exclusive hadronic reactions of the type investigated, \eg, in
1100: \cite{janssen}.
1101: 
1102: 
1103: \begin{acknowledgments}%
1104: One of us (P.~L.) is indebted to David Kraus for useful discussions. We
1105: thank Prof.~T.~Barnes for useful correspondence. This work was supported 
1106: by the Czech Ministry of Education, Youth and Sports under contracts
1107: MSM6840770029, MSM4781305903, and LC07050.
1108: \end{acknowledgments}
1109: 
1110: 
1111: \begin{thebibliography}{99}
1112: 
1113: \bi{vmd}Y.~Nambu, Phys. Rev. \textbf{106}, 1366 (1957);
1114: W.~R.~Frazer and J.~R.~Fulco, Phys. Rev. Lett. \textbf{2}, 365 (1959);
1115: J.~J.~Sakurai, Ann. Phys. (N.Y.) \textbf{11}, 1 (1960);
1116: Y.~Nambu and J.~J.~Sakurai, Phys. Rev. Lett. \textbf{8}, 79 (1962);
1117: \textbf{8}, 191(E) (1962); M.~Gell-Mann, D.~Sharp, and W.~Wagner, 
1118: \textit{ibid.} \textbf{8}, 261 (1962); 
1119: J.J.~Sakurai, {\it Currents and Mesons} (University of Chicago  
1120: Press, Chicago, IL, 1969).
1121: 
1122: \bi{cmd2}R.~R.~Akhmetshin \textit{et al.}, Phys. Lett. B \textbf{475},
1123:     190 (2000).
1124: 
1125: \bi{app}P.~Lichard, Acta Physics Slovaca \textbf{49}, 215 (1999).
1126: 
1127: \bi{decker}R.~Decker, P.~Heiliger, H.~H.~Jonsson, and M.~Finkemeier,
1128: Z. Phys. C \textbf{70}, 247 (1996).
1129: 
1130: \bi{isgur}N.~Isgur, C.~Morningstar, and C. Reader, Phys. Rev. D
1131: \textbf{39}, 1357 (1989).
1132: 
1133: \bi{czyz}H.~Czy\.{z} and J.~H.~K\"{u}hn, Eur. Phys. J. C (\textbf 18),
1134: 497 (2001).
1135: 
1136: \bi{ecker3}G. Ecker and R. Unterdorfer, Eur. Phys. J. C \textbf{24}, 535
1137: (2002). 
1138: % Four pion production in e+e- annihilation.
1139: 
1140: \bi{ecker4}G. Ecker and R. Unterdorfer, Nucl. Phys. B Proc. Suppl.
1141: \textbf{121}, 175 (2003). 
1142: % Four pion production.
1143: 
1144: \bi{bramon}A.~Bramon, A.~Grau, and G.~Pancheri, Phys. Lett. B
1145: \textbf{317},190 (1993).
1146: 
1147: \bi{eidelman}S.~I.~Eidelman, Z.~K.~Silagadze, and E.~A.~Kuraev,
1148:      Phys. Lett. B \textbf{346}, 186 (1995).
1149: 
1150: \bi{plant}R.~S.~Plant and M.~C.~Birse, Phys. Lett. B \textbf{365}, 292
1151:      (1996).
1152: 
1153: \bi{achasov1}N.~N.~Achasov and A.~A.~Kozhevnikov, Phys. Rev. D \textbf{61},
1154:      077904 (2000).
1155: 
1156: \bi{achasov2}N.~N.~Achasov and A.~A.~Kozhevnikov, Phys. Rev. D \textbf{62},
1157:      056011 (2000).
1158: 
1159: \bi{akhmetshin1999}R.~R.~Akhmetshin \textit{et al.}, Phys. Lett. B
1160:     \textbf{466}, 392 (1999).
1161: 
1162: \bi{pdg2006}W.-M.~Yao \textit{et al.}, Journal of Physics G \textbf{33}, 
1163: 1 (2006).
1164: 
1165: \bi{bondar1999}A.~E.~Bondar, S.~I.~Eidelman, A.~I.~Milstein, and
1166: N.~I.~Root,     Phys. Lett. B \textbf{466}, 403 (1999).
1167: 
1168: \bi{edwards2000}K.~W.~Edwards \textit{et al.},  Phys. Rev. D \textbf{61},
1169:      072003 (2000).
1170: 
1171: \bi{achasov3}N.~N.~Achasov and A.~A.~Kozhevnikov, Phys. Rev. D \textbf{71},
1172:      034015 (2005); Yad. Fiz. \textbf{69}, 314 (2006) [Phys. Atom. Nucl. 
1173:      \textbf{69}, 293 (2006)].
1174: 
1175: \bi{bando1988}M.~Bando, T.~Fujiwara, and K.~Yamawaki, Prog. Theor. Phys.
1176: \textbf{79}, 1140 (1988).
1177: 
1178: \bi{feinberg}E.~L.~Feinberg, Nuovo Cim. A \textbf{34}, 391 (1976).
1179: 
1180: \bi{shuryak}E.~Shuryak, Phys. Lett. \textbf{78B}, 150 (1978);
1181:  Yad. Fiz. \textbf{28}, 796 (1978) [Sov. J. Nucl. Phys. \textbf{28}, 
1182:  1548 (1978)].
1183: 
1184: \bi{na60}R.~Arnaldi \textit{et al.} (NA60 Collaboration),
1185:  Phys. Rev. Lett. \textbf{96}, 162302 (2006).
1186: 
1187: \bi{song}C.~Song, Phys. Rev. C \textbf{47}, 2861 (1993).
1188: 
1189: \bi{gaogale}S.~Gao and C.~Gale,  Phys. Rev. C \textbf{57}, 254 (1998).
1190: 
1191: \bi{joerg}J.~Ruppert, C.~Gale, T.~Renk, P.~Lichard, and J.~I.~Kapusta,
1192: \eprint{arXiv:0706.1934}.
1193: 
1194: \bi{ligale}G.-Q.~Li and C.~Gale, Phys. Rev. Lett. \textbf{81}, 1572
1195: (1998); Phys. Rev. C \textbf{58}, 2914 (1998);
1196: Nucl. Phys. A \textbf{638}, 491C (1998).
1197: 
1198: \bi{wess}J.~Wess and B.~Zumino, Phys. Rev. \textbf{163}, 1727 (1967).
1199: 
1200: \bi{gomm84}H.~Gomm, \"{O}.~Kaymakcalan, and J.~Schechter,
1201:     Phys. Rev. D \textbf{30}, 2345 (1984).
1202: 
1203: \bi{holstein}B.~R.~Holstein, Phys. Rev. D \textbf{33}, 3316 (1986).
1204: 
1205: \bi{meissner}U.~G.~Meissner, Phys. Rept. \textbf{161}, 213 (1988).
1206: 
1207: \bi{kaiser}N.~Kaiser and U.~G.~Meissner, Nucl. Phys. A \textbf{519},
1208: 671 (1990).
1209: 
1210: \bi{korudaz}P.~Ko and S.~Rudaz, Phys. Rev. D \textbf{50}, 6877 (1994).
1211: 
1212: \bi{li}B.~A.~Li, Phys. Rev. D \textbf{52}, 5165 (1995).
1213: 
1214: \bi{smejkal97}J.~Smejkal, E.~Truhl\'{\i}k, and H.~G\"{o}ller,
1215:     Nucl. Phys. A \textbf{624}, 655 (1997).
1216: 
1217: \bi{pham}T.~N.~Pham, C.~Roiesnel, and T.~N.~Truong, Phys. Lett. B
1218: \textbf{78}, 623 (1978).
1219: 
1220: \bi{kuhnsanta}J.~H.~K\"{u}hn and A.~Santamaria, Z. Phys. C \textbf{48},
1221:     445 (1990).
1222: 
1223: \bi{janssen}G.~Janssen, K.~Holinde, and J.~Speth, Phys. Rev. C \textbf{49},
1224:    2763 (1994).
1225: 
1226: \bi{haglin}K.~Haglin, Phys. Rev. C \textbf{50}, 1688 (1994).
1227: 
1228: \bi{xiong}L.~Xiong, E.~Shuryak, and G.~E.~Brown, Phys. Rev. D
1229: \textbf{46},    3798 (1992).
1230: 
1231: \bi{barnes}T.~Barnes, private communication.
1232: 
1233: \bi{DSratio}S.~U.~Chung \textit{et al.} (BNL/E852 Collaboration), Phys. 
1234: Rev.D \textbf{65}, 072001 (2002).
1235: 
1236: \bi{ecker1}G.~Ecker, J.~Gasser, H.~Leutwyler, A.~Pich, and
1237: E. de Rafael, Phys. Lett. \textbf{223}, 425 (1989).
1238: 
1239: \bi{ecker2}G.~Ecker, J.~Gasser,  A.~Pich, and E. de Rafael, Nucl. Phys.
1240: B \textbf{321}, 311 (1989).
1241: 
1242: \bi{brihaye}Y.~Brihaye, N.~K.~Pak, and P.~Rossi,
1243:     Nucl. Phys. B \textbf{254}, 71 (1985);
1244:     Phys. Lett. B \textbf{164}, 111 (1985).
1245: 
1246: \bi{kuraev}E.~A.~Kuraev and Z.~K.~Silagadze, Phys. Lett. B \textbf{292},
1247:     377 (1992).
1248: 
1249: \bi{ksrf}K.~Kawarabayashi and M.~Suzuki, Phys. Rev. Lett \textbf{16},
1250:     255 (1966); \textbf{16}, 384(E) (1966); Riazuddin and Fayyazuddin,
1251:     Phys. Rev. \textbf{147}, 1071 (1966).
1252: 
1253: \bi{ratio}P.~Lichard, Phys. Rev. D \textbf{60}, 053007 (1999).
1254: 
1255: \bi{GS} G.~J.~Gounaris and J.~J.~Sakurai, Phys. Rev. Lett. \textbf{21},
1256: 244 (1968).
1257: 
1258: \bi{VW}M.~T.~Vaughn and K.~C.~Wali, Phys. Rev. Lett. \textbf{21},
1259: 938 (1968).
1260: 
1261: \bi{melikhov04}D.~Melikhov, O.~Nachtmann, V.~Nikonov, and T.~Paulus,
1262: Eur. Phys. J. C \textbf{34}, 345 (2004).
1263: 
1264: \bi{bando1985}M.~Bando, T.~Kugo, S.~Uehara, K.~Yamawaki, and T.~Yanagida,
1265:     Phys. Rev. Lett \textbf{54}, 1215 (1985); M.~Bando, T.~Kugo,
1266:     and K.~Yamawaki, Nucl. Phys. B \textbf{259}, 493 (1985).
1267: 
1268: \bi{weinberg}S.~Weinberg, Phys. Rev. \textbf{166}, 1568 (1968).
1269: 
1270: \bi{sokoga}C.~Song, C.~M.~Ko, and C.~Gale, Phys. Rev. D \textbf{50},
1271:     R1827 (1994).
1272: 
1273: \bi{kokoski}R.~Kokoski and N.~Isgur, Phys. Rev. D \textbf{35}, 907
1274: (1987).
1275: 
1276: \bi{reduce}A.~C.~Hearn, \textit{Reduce User's Manual, Version 3.6},
1277: The Rand Corporation, Santa Monica, July 1995. See also
1278: \url{http://www.reduce-algebra.com/}.
1279: 
1280: \bi{babar}B.~Aubert \textit{et al.} (BaBar Collaboration),
1281:     Phys. Rev. D \textbf{71}, 052001 (2005).
1282: 
1283: \bi{minuit}F.~James and M.~Roos, Comput. Phys. Commun. \textbf{10}, 343
1284:     (1975).
1285: 
1286: \bi{turbide1}S.~Turbide, R.~Rapp, and C.~Gale, Int. J. Mod. Phys. A
1287: \textbf{19}, 5351 (2004).
1288: 
1289: \bi{turbide2}S.~Turbide, R.~Rapp, and C.~Gale, Phys. Rev. C \textbf{69}, 
1290: 014903 (2004).
1291: 
1292: \end{thebibliography}
1293: 
1294: \end{document}
1295: