1: \documentclass{article}
2: \sloppy
3: \usepackage{epsfig}
4: \usepackage{float}
5: \begin{document}
6: \title{Allowed Eta-Decay Modes and Chiral Symmetry}
7: \author{Barry R. Holstein$^a$\\[5mm]
8: Department of Physics\\
9: University of Massachusetts\\
10: Amherst, MA 01003\\and\\
11: Institut f\"{u}r Kernphysik\\
12: Forschungszentrum J\"{u}lich\\
13: D-52425 J\"{u}lich, Germany}
14: \begin{titlepage}
15: \maketitle
16: \begin{abstract}
17:
18: Recently, the development of chiral perturbation theory has allowed the
19: generation of rigorous low-energy theorems for various hadronic processes
20: based only on the chiral invariance of the underlying QCD Lagrangian.
21: Such techniques are highly developed and well-tested in the domain of
22: pionic and kaonic reactions. In this note we point out that with the
23: addition of a few additional and reasonable assumptions similar predictive
24: power is available for processes involving the eta meson.
25: \end{abstract}
26: \vfill
27: $^a$ email: holstein@physics.umass.edu
28: \end{titlepage}
29:
30: \section{Introduction}
31: It has long been the holy grail for particle and nuclear knights
32: to generate rigorous
33: predictions from the Lagrangian of QCD
34: \begin{equation}
35: {\cal L}_{\rm QCD}=-{1\over 2}G_{\mu\nu}G^{\mu\nu}+\bar{q}(i\gamma_\mu
36: D^{\mu}-m)q\label{eq:qcd}
37: \end{equation}
38: where
39: \begin{eqnarray}
40: G_{\mu\nu}&=&\partial_\mu A_\nu -\partial_\nu A_\mu -ig[A_\mu ,A_\nu ]\nonumber\\
41: D_\mu q&=& (\partial_\mu -igA_\mu )q.
42: \end{eqnarray}
43: However, despite the ease with which one can write this equation,
44: because of its inherent nonlinearity and the large value of the
45: coupling constant---$g^2/4\pi\sim 1$---progress in this regard has been
46: slow. One approach---lattice gauge theory---holds great promise, but is
47: currently
48: limited by the need for large computational
49: facilities\cite{1}. A second tack, that of
50: perturbative QCD, exploits the feature of asymptotic freedom---the vanishing of
51: the running coupling constant at high momentum
52: transfer\cite{2}. However, such predictions
53: are valid only for the very highest energy processes. It is gratifying then to
54: see that in recent years a third procedure has become available, that of chiral
55: perturbation theory ($\chi$PT) which exploits the chiral symmetry of QCD and
56: allows rigorous predictive power in the case of low energy reactions. This
57: technique, based on a suggestion due to Weinberg\cite{3}, was developed
58: (at one loop level) during the last decade in an important series of papers by Gasser
59: and Leutwyler and others\cite{4}. The idea is based on the feature that the QCD
60: Lagrangian---Eq. \ref{eq:qcd}---possesses a global $SU(3)_L\times SU(3)_R$ (chiral)
61: invariance in the limit of vanishing quark mass. Such invariance is manifested
62: in the real world not in the conventional Wigner-Weyl fashion.
63: Rather, it is spontaneously
64: broken, resulting in eight light pseudoscalar
65: Nambu-Goldstone bosons---$\pi, K,\eta$---which
66: would be massless if the corresponding quark masses also vanished\cite{5}.
67: While the identification of this symmetry is apparent in
68: terms of quark/gluon degrees of
69: freedom, it is not so simple to understand the implications of chiral invariance
70: in the arena of experimental meson/baryon interactions.
71:
72: Early attempts in this direction were based on current algebra/PCAC
73: methods\cite{6}, yielding relationships between processes differing in the
74: number of pions, {\it e.g.}
75: \begin{equation}
76: \lim_{q\rightarrow 0}<B\pi^a_q|{\cal O}|A>={-i\over F_\pi}<B|[F_5^a,{\cal O}]|A>
77: \end{equation}
78: where $F_\pi =92.4$ MeV is the pion decay constant\cite{7}. However, it was
79: soon realized that the most succinct way to present these restrictions is in
80: terms of an effective chiral Lagrangian, the simplest (two-derivative) form of
81: which is, in the Goldstone sector\cite{8},
82: \begin{equation}
83: {\cal L}_{\rm eff}^{(2)}={\bar{F}^2\over 4}{\rm Tr}D_\mu UD^\mu U^\dagger
84: +{\bar{F}^2\over 4}{\rm Tr}2B_0m(U+U^\dagger )+\cdots
85: \end{equation}
86: where
87: \begin{equation}
88: U=\exp\left({i\over \bar{F}}\sum_{j=1}^8\lambda_j\phi_j\right)
89: \end{equation}
90: is a nonlinear function of the pseudoscalar fields, $m=(m_u,m_d,m_s)_{\rm diag}$
91: is the quark mass matrix,
92: \begin{equation}
93: 2B_0={2m_K^2\over m_u+m_s}={2m_\pi^2\over m_u+m_d}
94: \end{equation}
95: is a phenomenological constant, $D_\mu=\partial_\mu-i[V_\mu,
96: ]-i\{A_\mu, \}$ is the covariant derivative coupling to external
97: fields $V_\mu,A_\mu$, and
98: $\bar{F}$ is the pion decay constant in the limit of chiral symmetry. Although
99: these are only two of an infinite number of terms, already at this level
100: there exists
101: predictive power--{\it e.g.}, tree level evaluation of ${\cal L}^{(2)}$
102: yields the familiar Weinberg predictions (at ${\cal O}(p^2,m^2)$) for
103: S-wave $\pi -\pi$ scattering lengths\cite{9}
104: \begin{equation}
105: a_0^0={7m_\pi\over 32\pi F_\pi^2}\qquad a_0^2=-{m_\pi\over 16\pi F_\pi^2}
106: \end{equation}
107: which are approximately borne out experimentally. Loop diagrams, of course,
108: produce terms of ${\cal O}(p^4,p^2m^2,m^4)$ and contain divergences. However,
109: just as in QED such infinities can be absorbed into renormalizing phenomenological
110: chiral couplings, and the most general "four-derivative" Lagrangian has been
111: given by Gasser and Leutwyler\cite{3}
112: \begin{eqnarray}
113: {\cal L}^{(4)}_{\rm eff}&=&L_1({\rm Tr}D_\mu UD^\mu U^\dagger )^2
114: +L_2({\rm Tr} D_\mu UD_\nu U^\dagger )^2\nonumber\\
115: & &+L_3{\rm Tr}(D_\mu UD^\mu U^\dagger)^2 +
116: L_4{\rm Tr}D_\mu UD^\mu U^\dagger ){\rm Tr}(m(U+U^\dagger ))\nonumber\\
117: & &+L_5{\rm Tr}D_\mu UD^\mu U^\dagger m(U+U^\dagger )
118: +L_6({\rm Tr}m(U+U^\dagger ))^2\nonumber\\
119: & &+L_7({\rm Tr}m(U-U^\dagger ))^2+
120: L_8 {\rm Tr}(mUmU+mU^\dagger mU^\dagger )\nonumber\\
121: & &+iL_9({\rm Tr}F^L_{\mu\nu}D^\mu UD^\nu U^\dagger +{\rm Tr}F^R_{\mu\nu}
122: D^\mu UD^\nu U^\dagger )\nonumber\\
123: & &+L_{10}{\rm Tr}F^L_{\mu\nu}UF^{R\mu\nu}U^\dagger
124: \end{eqnarray}
125: where $F^L_{\mu\nu}, F^R_{\mu\nu}$ are external field strength tensors defined
126: via
127: \begin{eqnarray}
128: F^{L,R}_{\mu\nu}&=&\partial_\mu F^{L,R}_\nu -\partial_\nu F^{L,R}_\mu
129: -[F^{L,R}_\mu ,F^{L,R}_\nu ]\nonumber\\
130: F^L_\mu &=&V_\mu +A_\mu \qquad F^R_\mu =V_\mu-A_\mu .
131: \end{eqnarray}
132: Here the bare $L_i$ coefficients are themselves unphysical and are related
133: to empirical quantities $L_i^r (\mu )$ measured at scale $\mu$ via
134: \begin{equation}
135: L_i^r(\mu )=L_i+{\Gamma_i\over 32\pi^2}\left({1\over \epsilon}+{\rm ln}
136: {4\pi\over \mu^2}+1-\gamma\right) ,
137: \end{equation}
138: where $\Gamma_i$ are constants defined in ref. 4b and $\epsilon=4-d$ is the
139: usual parameter arising in dimensional regularization, with $d$ being
140: the number of dimensions. Gasser and Leutwyler
141: have obtained empirical values for the phenomenological constants $L_1^r(\mu),
142: \ldots L_{10}^r(\mu)$, values for some of which are given in Table 1.
143:
144: \begin{table}
145: \begin{center}
146: \begin{tabular}{|lllll|} \hline
147: $L_1^r$ & $L_2^r$ & $L_5^r$ & $L_9^r$ & $L_{10}^r$ \\
148: \hline
149: $0.71\pm 0.28$ & $2.01\pm 0.37$ & $2.7\pm 0.3$ & $7.7\pm 0.2$ & $-5.2\pm 0.3 $\\
150: \hline
151: \end{tabular}
152: \caption{Empirical values of Chiral Expansion Parameters($\times
153: 10^{-3})$ with $\mu =m_\eta$.}
154: \end{center}
155: \end{table}
156:
157: At the four-derivative level it is also necessary to include the contribution
158: of the anomaly, which in the case of coupling to the photon field $A_\mu$ has
159: the form\cite{10}
160: \begin{eqnarray}
161: & &\Gamma_{\rm WZW}(U,A_\mu )=\Gamma_{\rm WZW}(U)\nonumber\\
162: &+&{N_c\over 48\pi^2}\epsilon^{\mu\nu\alpha\beta }\int d^4x\left[eA_\mu{\rm Tr}
163: (Q(R_\nu R_\alpha R_\beta +L_\nu L_\alpha L_\beta))\right. \nonumber\\
164: &-&\left.ie^2F_{\mu\nu}A_\alpha {\rm Tr}\left(Q^2(R_\beta +L_\beta)+{1\over 2}
165: (QU^\dagger QUR_\beta +QUQU^\dagger L_\beta )\right)\right]\label{eq:anom}
166: \end{eqnarray}
167: where $R_\mu \equiv (\partial_\mu U^\dagger)U, L_\mu\equiv
168: U\partial_\mu U^\dagger$ and $\Gamma_{\rm WZW}$ is independent of the photon
169: field and will not be needed for our purposes. A corresponding form involving coupling to
170: a general nonabelian gauge field can also be written, but is lengthy and
171: will not be given here\cite{11}.
172:
173: In a series of recent papers it has been conclusively demonstrated that this
174: chiral effective action formalism provides a succinct and successful description
175: of low energy electroweak interactions of pions and kaons\cite{12}. Specifically, the
176: reactions given in Table 2 are successfully described in terms of
177: GL parameters $L_9(\mu),L_{10}(\mu)$. Clearly, there are far more reactions than parameters, and
178: this overdetermination enables one to construct {\it required} relationships
179: between experimental
180: quantities, the empirical validity of which constitutes a strong test of the
181: chiral formalism and indeed thereby of QCD itself\cite{13}. Such tests are
182: found to be well satisfied, with the possible exception of the
183: relationship between the axial structure function in radiative
184: pion beta decay---$h_A$---and the charged pion
185: polarizability---$\alpha_E^{\pi^+}$\cite{14}.
186: However, recent work indicates that this may not be a problem and in any case
187: a number of experimental efforts are underway to retest this critical
188: stricture\cite{15}.
189:
190: \begin{table}
191: \begin{center}
192: \begin{tabular}{|llll|}
193: \hline
194: {\bf Reaction} & {\bf Quantity} & {\bf Theory} & {\bf Experiment} \\
195: $\gamma\rightarrow\pi^+\pi^-$ & $<r_\pi^2>$({\rm fm}$^2$) & 0.44$^a$ &
196: $0.44\pm 0.02$ \\
197: $\gamma\rightarrow K^+K^-$ & $<r_K^2>$({\rm fm}$^2$) & 0.44 & $0.34\pm 0.05$ \\
198: $\pi^+\rightarrow \pi^+\nu_e\gamma $ & $h_V(m_\pi^{-1})$ & 0.027 & $0.029
199: \pm 0.017$\\
200: {} & $h_A/h_V$ & $0.46^a$ & $0.46\pm0.08$ \\
201: $K^+\rightarrow e^+\nu_e\gamma$ & $(h_V+h_A)(m_\pi^{-1})$ & 0.038 &
202: $0.043\pm 0.003$\\
203: $\pi^+\rightarrow e^+\nu_ee^+e^-$ & $r_A/h_V$& 2.6 & $2.3\pm 0.6$ \\
204: $\gamma\pi^+\rightarrow\gamma\pi^+$ & $(\alpha_E+\beta_M)(10^{-4}({\rm fm}^3)$ &
205: 0 & $1.4\pm 3.1$ \\
206: {} & $\alpha_E(10^{-4}({\rm fm}^3)$ & 2.8 & $6.8\pm 1.4$\\
207: $\gamma\gamma\rightarrow\pi\pi$ & $\alpha_E(10^{-4}({\rm fm}^3)$ & 2.8 &
208: $2.2\pm 1.6$\\
209: $K\rightarrow\pi e^+\nu_e$ & $\xi=f_-(0)/f_+(0)$ & -0.13 & $-0.20\pm 0.08$\\
210: {} & $\lambda_+({\rm fm}^2)$ & 0.067 & $0.065\pm 0.005$\\
211: {} & $\lambda_0({\rm fm}^2)$ & 0.040 & $0.050\pm 0.012$ \\
212: \hline
213: \end{tabular}
214:
215: \caption{Chiral predictions and data in the radiative complex of
216: transitions. The superscript a indicates that this parameter was used
217: as input and is {\it not} a predicted quantity.}
218: \end{center}
219: \end{table}
220:
221: \medskip
222:
223: In the pion and kaon arena then one finds strong evidence for the correctness
224: and utility of chiral perturbative techniques, and it is an obvious next step
225: to attempt to extend this success into the eta sector, which is the
226: subject of this note. This examination of the eta system is important both as
227: a theoretical exercise and because of the existence of
228: high intensity sources of etas\cite{16}.
229: In the next section then we study the ability of the chirally inspired
230: methods to
231: make reliable predictions for eta decay processes. We examine only the
232: "allowed" modes---$\eta\rightarrow 2\gamma , 3\pi , 2\pi\gamma,
233: \pi 2\gamma ,3\pi\gamma$, {\it i.e.} those modes which can occur
234: assuming only isospin violation or the anomaly, eschewing the temptation to
235: analyze ``rare'' processes such as
236: $\eta\rightarrow 3\gamma$, as these have been
237: well-discussed elsewhere\cite{17}.
238: Finally, we summarize our results in a concluding section III.
239:
240: \section{Eta Decay Processes}
241:
242: Strictly from a
243: kinematic perspective, inclusion of the $\eta(547)$ as part of the chiral
244: formalism should not be a problem, as the eta and kaon are
245: roughly degenerate in mass, and as mentioned above the kaon (and pion)
246: predictions
247: obtained in this way are quite successful. Rather the real challenge
248: involves mixing with
249: $\eta '(958)$, which lies outside the simple chiral $SU(3)_L\times SU(3)_R$
250: framework. To lowest order things are simple---in the chiral limit the
251: pseudoscalar mass spectrum would consist of a massless octet of Goldstone
252: bosons plus a massive $SU(3)$ singlet ($\eta_0$). With the breaking of
253: chiral invariance the octet pseudoscalar masses become nonzero and are
254: related, at first order in chiral symmetry breaking,
255: by the Gell-Mann-Okubo formula\cite{gmo}
256: \begin{equation}
257: m^2_{\eta_8}={4\over 3}m_K^2-{1\over 3}m_\pi^2\approx (0.57\,\,{\rm GeV})^2\label{eqn:a}
258: \end{equation}
259: where $\eta_8$ is the eighth member of the octet. At this same order in
260: symmetry breaking the singlet $\eta_0$ will in general mix with $\eta_8$
261: producing the physical eigenstates $\eta,\eta '$ given by
262: \begin{eqnarray}
263: \eta &=&\cos\theta \eta_8-\sin\theta\eta_0\nonumber\\
264: \eta '&=&\sin\theta\eta_8+\cos\theta\eta_0.
265: \end{eqnarray}
266: The mixing angle $\theta$ can be determined via diagonalization of the mass
267: matrix (written in the $\eta_8,\eta_0$ basis)
268: \begin{equation}
269: m^2=\left( \begin{array}{cc}
270: m_{\eta_8}^2 & m_{08}^2 \\
271: m_{08}^2 & m_{\eta_0}^2 \\
272: \end{array} \right)
273: \end{equation}
274: Here $m_{\eta_8}^2 $ is given in Eq. \ref{eqn:a} while $m_{08}^2,m_{\eta_0}^2$,
275: and $\theta$ are unknown. Diagonalizing and
276: fitting these parameters with the two known masses yields the results
277: \begin{equation}
278: \theta = -9.4^\circ ,\quad m_{08}^2=-0.44m_K^2,\quad
279: m_{\eta_0}=0.95\,{\rm GeV}
280: \end{equation}
281: However, there is good reason not to trust this simple and lowest order analysis, since
282: higher order chiral symmetry breaking terms can generate important modifications.
283: For example, taking the leading log correction arising
284: from Figure 1, we find\cite{18}
285: \begin{eqnarray}
286: m_{\eta_8}^2&=&{4\over 3}m_K^2-{1\over 3}m_\pi^2-{2\over 3}{m_K^2\over (4\pi
287: F_\pi )^2}{\rm ln}{m_K^2\over \mu^2}\nonumber\\
288: &\approx & (0.61\,\,{\rm GeV} )^2 \quad {\rm for} \quad \mu =
289: 1\,\,{\rm GeV},
290: \end{eqnarray}
291: for which diagonalization of the mass matrix yields
292: \begin{equation}
293: \theta \approx -19.5^\circ ,\quad m_{08}^2=-0.81m_K^2,\quad
294: m_{\eta_0}=-0.90\,{\rm GeV}\label{eqn:b}
295: \end{equation}
296: suggesting a doubling of the mixing angle.
297: \begin{figure}[htb]
298: \begin{center}
299: \epsfig{file=etafig1.eps,height=3cm,width=6cm}
300: \end{center}
301: \caption{Mass and wavefunction renormalization diagram.}
302: \end{figure}
303: Of course, this is just an approximate result. However, a full one loop
304: calculation using $\chi$PT yields essentially the same result\cite{4}, and
305: consequently in our analysis below we shall use the value in Eq. \ref{eqn:b}, {\it i.e.}
306: \begin{equation}
307: \sin\theta \approx -{1\over 3}\qquad \cos\theta \approx {2\sqrt{2}\over 3}.
308: \end{equation}
309: It is also intriguing that this solution is consistent with the
310: assumptions of simple U(3) invariance wherein $\eta_8,\eta_0$ have the
311: same quark wavefunction, leading to
312: \begin{equation}
313: {m_{08}^2\over m_K^2}\simeq{2\sqrt{2}\over 3}\left(\hat{m}-m_s\over
314: \hat{m}+m_s\right)\simeq -0.9
315: \end{equation}
316:
317: At this same (one-loop) level of symmetry breaking there is generated a shift
318: in the lowest order value of the pseudoscalar decay constant $F_P$.
319: Thus one finds from the diagrams in Figure 2, in
320: leading log approximation,
321: \begin{eqnarray}
322: F_\pi &=&\bar{F}\left[1-{1\over 2}{m_K^2\over (4\pi F_\pi)^2}\ln {m_K^2\over
323: \mu^2}\right] \approx 1.12\bar{F}\nonumber\\
324: F_{\eta_8}&=&\bar{F}\left[1-{3\over 2}{m_K^2\over (4\pi F_\pi )^2}\ln {m_K^2\over
325: \mu^2}\right] \approx 1.25 F_\pi \quad {\rm for} \quad \mu\approx
326: 1\,\,{\rm GeV}.
327: \end{eqnarray}
328: Once again these estimates are in excellent agreement with those given in the full
329: one-loop analysis\cite{4}. (We note in addition that the corresponding prediction
330: \begin{equation}
331: {F_K\over F_\pi}=1-{1\over 4}{m_K^2\over (4\pi F_\pi)^2}\ln {m_K^2\over \mu^2}
332: -{3\over 8}{m_\eta^2\over (4\pi F_\pi)^2}\ln {m_\eta^2\over \mu^2}\approx 1.22
333: \end{equation}
334: is quite consistent with the experimental value $1.22\pm 0.01$)
335: \begin{figure}[htb]
336: \begin{center}
337: \epsfig{file=etafig2.eps,height=3cm,width=8cm}
338: \caption{Loop diagrams leading to renormalization of the pseudoscalar
339: decay constant. Here the symbol $x$ indicates coupling to the axial current.}
340: \end{center}
341: \end{figure}
342:
343: \subsection{\bf $\eta\rightarrow \gamma\gamma$}
344:
345: With this introductory material in hand we can now confront the
346: subject of our report---that of eta decay. First consider the dominant two-photon
347: decay mode, which to leading order arises due to the anomaly. In the
348: analogous $\pi^0\rightarrow \gamma\gamma$ case we find from Eq. \ref{eq:anom}
349: \begin{equation}
350: {\rm Amp}\equiv F_{\pi\gamma\gamma}(0)\epsilon^{\mu\nu\alpha\beta}\epsilon_\mu
351: k_\nu\epsilon'_\alpha k'_\beta \quad{\rm with}\quad F_{\pi\gamma\gamma}^{theo}(0)=
352: {N_c\alpha\over 3\pi F_\pi }=0.025\,\,{\rm GeV}^{-1}.\label{eq:theo}
353: \end{equation}
354: General theorems guarantee that this result is not altered in higher orders
355: of chiral symmetry breaking\cite{19} and, using the experimental value\cite{20}
356: \begin{equation}
357: \Gamma (\pi^0\rightarrow \gamma\gamma )= (7.7\pm 0.7)\,\,{\rm eV},
358: \end{equation}
359: we determine
360: \begin{equation}
361: F_{\pi\gamma\gamma}^{exp}=(0.0250\pm 0.0005)\,\,{\rm GeV}^{-1}
362: \end{equation}
363: in excellent agreement with the theoretically predicted value and eloquently
364: confirming the value $N_c=3$ as the number of colors. Strictly speaking,
365: the prediction of the anomalous four-derivative Lagrangian Eq. \ref{eq:anom}
366: should be in terms of $\bar{F}$ rather than $F_\pi$. Indeed the difference
367: between the value given in Eq. \ref{eq:theo} and the strict four-derivative prediction
368: involves terms of dimension six and is higher order in the chiral expansion.
369: Clearly, however, our prediction for $F_{\pi\gamma\gamma}(0)$, which arises from
370: what we shall term extended-$\chi$PT, is in excellent agreement with experiment.
371: Nevertheless, although very reasonable, this is {\it not} a firm prediction
372: of $\chi$PT itself.
373:
374: The $\eta ,\eta '\rightarrow \gamma\gamma$ couplings
375: also arise from the anomalous
376: component of the effective chiral Lagrangian and, in the extended $\chi$PT
377: approximation, have the values
378: \begin{eqnarray}
379: F_{\eta\gamma\gamma}(0)&=&{F_{\pi\gamma\gamma}(0)\over \sqrt{3}}
380: \left({F_\pi\over F_8}\cos\theta
381: -2\sqrt{2}{F_\pi\over F_0}\sin\theta\right)\nonumber\\
382: F_{\eta '\gamma\gamma}(0)&=&{F_{\pi\gamma\gamma}(0)\over \sqrt{3}}
383: \left( {F_\pi\over F_8}\sin\theta + 2\sqrt{2}{F_\pi\over F_0}\cos
384: \theta )\right) .
385: \end{eqnarray}
386: Using the experimental numbers\cite{fnt1}
387: \begin{eqnarray}
388: \Gamma (\eta\rightarrow\gamma\gamma )&=& (0.51 \pm 0.05)\,\,{\rm keV} \qquad
389: \Gamma (\eta'\rightarrow\gamma\gamma )= (4.7\pm 0.7)\,\,{\rm keV}
390: \end{eqnarray}
391: we find
392: \begin{equation}
393: \quad F_{\eta\gamma\gamma}(0)=0.0249\pm 0.0010\,\,{\rm GeV}^{-1} \qquad
394: F_{\eta'\gamma\gamma}(0)=0.0328\pm 0.0024\,\,{\rm GeV}^{-1}\label{eq:data}
395: \end{equation}
396: In order to solve this system, we require an additional assumption
397: since there are three unknowns---$F_0,F_8,\theta$---but only two pieces
398: of data---Eq. \ref{eq:data}. The standard approach at this point has been
399: to use the leading log prediction from one-loop chiral perturbation
400: theory---
401: \begin{equation}
402: {F_8\over F_\pi}=1-{m_K^2\over (4\pi F_\pi)^2}\ln {m_K^2\over
403: \mu^2}+{m_\pi^2\over (4\pi F_\pi)^2}\ln{m_\pi^2\over \mu^2}\simeq
404: 1.30\quad {\rm at} \quad\mu\sim 1\, {\rm GeV}
405: \end{equation}
406: as input, and then to solve for $F_0,\theta$, yielding
407: \begin{equation}
408: \theta\simeq -20^\circ , \qquad {F_0\over F_\pi}\approx 1.04
409: \end{equation}
410: It is intriguing that these results from two phton decay are quite
411: compatable with those obtained from the one-loop analysis of the mass
412: matrix---{\it i.e.}, $\theta\simeq -20^\circ$ and $F_0/F_\pi$
413: consistent with the value of unity which one would obtain if the
414: singlet state and the pion were to have the same quark wavefunction.
415:
416: Closely related to the above modes are the associated Dalitz
417: decays---$\pi ,\eta ,\eta '\rightarrow \gamma e^+e^- $---which
418: have been
419: studied at DESY\cite{23}. These are traditionally parameterized in terms
420: of a dipole slope parameter $\Lambda_P$ such that
421: \begin{equation}
422: {1\over \Gamma}{d\Gamma\over ds} = (1+ {s\over \Lambda_P^2})^2
423: \quad {\rm with}\quad s=(p_+ +p_-)^2.
424: \end{equation}
425: The experimentally obtained values
426: \begin{equation}
427: \Lambda_\pi = 0.75\pm 0.03\,\,{\rm GeV}\qquad\Lambda_\eta=0.84\pm 0.06
428: \,\,{\rm GeV}\qquad
429: \Lambda_{\eta'}=0.79\pm 0.04\,\,{\rm GeV}
430: \end{equation}
431: are in reasonable agreement with the vector dominance predictions\cite{24}
432: \begin{eqnarray}
433: \Lambda_\pi^2&=&m_{\rho ,\omega}^2\approx (0.77\,\,{\rm GeV})^2\nonumber\\
434: \Lambda_\eta^2&=&m_{\rho ,\omega}^2\left({3\cos\theta-
435: 6\sqrt{2}\sin\theta\over
436: (5-2\xi^2)\cos\theta -(5+\xi^2)\sqrt{2}
437: \sin\theta}\right)\approx (0.75\,\,{\rm GeV})^2\nonumber\\
438: \Lambda_{\eta'}^2&=&m_{\rho ,\omega}^2\left( {3\sin\theta
439: +6\sqrt{2}\cos\theta
440: \over (5-2\xi^2)\sin\theta
441: +(5+\xi^2)\sqrt{2}\cos\theta}\right)\approx (0.83\,\,{\rm GeV})^2
442: \nonumber\\
443: & &{\rm where}\qquad \xi^2=(m_{\rho ,\omega}^2/m_\phi^2 )
444: \end{eqnarray}
445: These results are consistent with the observation that vector/axial dominance
446: provides a remarkably successful representation of the values of the GL
447: parameters $L_i^r(\mu)$ obtained empirically\cite{25} and suggest an additional
448: extension of our (already) extended $\chi$PT to include an effective
449: vector-dominated Lagrangian\cite{26}
450: \begin{equation}
451: {\cal L}_{\rm eff}={\cal L}_{\rm VPP}+{\cal L}_{V\gamma}+{\cal L}_{\rm VVP}
452: +{\cal L}_{PPP\gamma}
453: \end{equation}
454: with
455: \begin{eqnarray}
456: {\cal L}_{\rm VPP}&=&{-ig\over 4}{\rm Tr}V_\mu[\phi, \partial^\mu
457: \phi ]\nonumber\\
458: {\cal L}_{\rm V\gamma}&=&2eF_\pi^2gA^\mu\left(\rho_\mu +{1\over 3}\omega_\mu
459: -{\sqrt{2}\over 3}\phi_\mu \right)\nonumber\\
460: {\cal L}_{\rm VVP}&=&-{\sqrt{3}\over 4}g_{VVP}\epsilon^{\mu\nu\alpha\beta}
461: {\rm Tr}(\partial_\mu V_\nu\partial_\alpha V_\beta \phi )\nonumber\\
462: {\cal L}_{PPP\gamma }&=&{-ieN_c\over 24\pi^2F_\pi^3}
463: \epsilon^{\mu\nu\alpha\beta}A_\mu\partial_\nu\pi^+ \partial_\alpha\pi^-\label{eq:vdl}
464: \partial_\beta\pi^0
465: \end{eqnarray}
466: and
467: \begin{equation}
468: g_{VVP}=-{3g^2\over 8\pi^2F_\pi},
469: \end{equation}
470: in order to understand the six-derivative component of the chiral expansion.
471: Here g is given by the KSRF relation as\cite{27}
472: \begin{equation}
473: g^2={m_\rho^2\over 2F_\pi^2}.
474: \end{equation}
475: In this formalism then the amplitude for $\pi^0\rightarrow \gamma\gamma $
476: is determined from the diagram in Figure 3 to be
477: \begin{equation}
478: F_{\pi\gamma\gamma}(0)={2e^2\over 3m_\omega^2m_\rho^2}(2eF_\pi g)^2
479: g_{\omega\rho\pi}={e^2\over 4\pi^2F_\pi}
480: \end{equation}
481: as the value required by the anomaly.
482:
483: With this background in hand, we
484: can now discuss a second important decay mode of the eta---that of
485: $\eta\rightarrow \pi\pi\gamma$ which also arises from the anomalous
486: component of the Lagrangian.
487:
488: \begin{figure}[htb]
489: \begin{center}
490: \epsfig{file=etafig3.eps,height=4cm,width=5.5cm}
491: \caption{Vector dominance diagram responsible for $\pi^0\rightarrow
492: \gamma\gamma$.}
493: \end{center}
494: \end{figure}
495:
496: \subsection{\bf $\eta\rightarrow\pi^+\pi^-\gamma$}
497: In the previous section we performed an analysis of the QCD anomaly as
498: manifested in the two photon decay of the pseudoscalar mesons and, by
499: use of the one loop leading log value for $F_8/F_\pi$, were able to
500: determine a solution for the $\eta,\eta'$ mixing angle which is close
501: to that found in the mass analysis together with a value for $F_0/F_\pi$ which
502: is near that which results from the assumption that the $\eta_0$
503: and pion have the same wavefunction. While this is somewhat
504: satisfying, it is intriguing to inquire whether one can assess the mixing
505: angle purely phenomenologically. We shall show below that this
506: question can be answered in the affirmative,
507: provided one utilizes the additional information available in
508: the anomalous decays $\eta,\eta'\rightarrow \pi^+\pi^-\gamma$.
509: Such processes involving a photon coupled to three pseudoscalar mesons involve the
510: anomaly and, at zero four-momentum, are completely determined from
511: Eq. \ref{eq:anom}. However, inclusion of higher order effects generates structure and
512: study of such processes requires proper attention to issues of
513: unitarity and final state interactions. Before considering
514: $\eta,\eta'$ decay,
515: however, we consider first the closely related case of
516: $\gamma\rightarrow\pi^+\pi^-\pi^0$.
517: At zero four-momentum the
518: anomaly requires\cite{28}
519: \begin{eqnarray}
520: {\rm Amp}(3\pi -\gamma)&=&A(s_{+-},s_{+0},s_{-0})\epsilon^{\mu\nu\alpha\beta}
521: \epsilon_\mu p_{+\nu}p_{-\alpha}p_{0\beta}\nonumber\\
522: {\rm where}\quad A(0,0,0)&=&{eN_c\over 12 \pi^2F_\pi^3}=9.7\,\,{\rm GeV}^{-3}
523: \quad{\rm and}\quad s_{ij}=(p_i+p_j)^2\label{eq:vv}
524: \end{eqnarray}
525: and one might suspect that vector dominance could reproduce this result
526: directly, as in the case of $\pi^0\rightarrow \gamma\gamma$ discussed above.
527: However, this turns out {\it not} to be the case. Rather, use of the
528: diagram shown in Figure 4a yields
529: \begin{equation}
530: A(s,t,u)={eN_c\over 24\pi^2F_\pi^3}\left( {m_\rho^2\over m_\rho^2 -s}
531: +{m_\rho^2\over m_\rho^2-t}+{m_\rho^2\over m_\rho^2-u}\right)
532: \end{equation}
533: which at zero four-momentum is 50\% larger than the value given by the anomaly.
534: The resolution of this problem is well-known and arises from the a direct
535: $\gamma-3\pi$ coupling---Figure 4b---given in ${\cal L}_{PPP\gamma }$
536: whose origin presumably
537: is from unspecified high-momentum-scale processes\cite{29} .
538: Addition of this contribution to the
539: $\gamma -3\pi$ process yields an amplitude
540: \begin{equation}
541: A(s,t,u)={eN_c\over 12\pi^2F_\pi^3}\left[1+{1\over 2}\left({s\over m_\rho^2-s}
542: +{t\over m_\rho^2-t}+{u\over m_\rho^2-u}\right)\right]\label{eq:edep}
543: \end{equation}
544: which has the structure required by vector dominance, but also agrees with the
545: value required by the chiral anomaly at zero four-momentum. Inclusion of
546: loop corrections modifies Eq. \ref{eq:edep} slightly but does not
547: change the general form.\cite{30}
548:
549: \begin{figure}[htb]
550: \begin{center}
551: \epsfig{file=etafig4.eps,height=4cm,width=11cm}
552: \caption{Vector dominance diagrams responsible for the anomalous
553: process $\gamma\pi\rightarrow\pi\pi$.}
554: \end{center}
555: \end{figure}
556:
557: The $\gamma -3\pi$ reaction has been studied experimentally via
558: pion pair production by the pion in the nuclear Coulomb field
559: and yields a number\cite{31}
560: \begin{equation}
561: A(0,0,0)_{\rm exp}=12.9\pm 0.9 \pm 0.5\,\,{\rm
562: GeV}^{-3}
563: \end{equation}
564: in apparent disagreement with Eq. \ref{eq:vv} and suggesting the
565: value $N_c\approx 4$! This value was obtained, however, assuming no energy
566: dependence of the amplitude and is reduced to
567: \begin{equation}
568: A(0,0,0)_{\rm exp}=11.9\pm 0.9\pm 0.5\,\,{\rm GeV}^{-3}
569: \end{equation}
570: if Eq. \ref{eq:edep} is utilized, but is still too large.
571: The most likely conclusion is that this an experimental
572: problem associated with this difficult-to-measure process, but it has
573: recently been pointed out by Ametller, Knecht, and Talavera
574: that an important electromagnetic
575: effect---the photon exchange diagram connecting $\pi^0\gamma\gamma^*$
576: and $\pi^+\pi^-\gamma$ vertices---can reduce this number by another
577: $1\times 10^{-3}$ or so\cite{akt}. In any
578: case a new high-precision experiment is clearly called for, and
579: this has been accomplished at JLab using the CLAS detector and the
580: reaction $\gamma p\rightarrow \pi^+\pi^- n$.
581: Such a measurement has been also been proposed at Da$\Phi$ne\cite{32}.
582:
583: The JLab experiment is currently being analyzed and this high
584: statistics measurement will require an equally careful theoretical
585: analysis in order to produce the desired extrapolation from the rho
586: resonance region, where most of the data has been obtained, to the
587: zero four-momentum point where the anomaly stricture obtains. This
588: issue has been approached in a number of authors:
589: \begin{itemize}
590: \item [i)] Holstein has used a simple matching of the one loop chiral
591: correction to the rho dominance form\cite{hol}.
592: \item [ii)] Truong has utilized a unitarization based upon the
593: Omnes-Muskhelishvili method\cite{tru}.
594: \item [iii)] Hannah unitarizes the amplitude using the inverse
595: amplitude procedure\cite{han}.
596: \end{itemize}
597: Regardless of the method used, the results are similar and are somewhat
598: robust. The resulting value of the anomaly obtained
599: in a preliminary analysis of the CLAS measurement are consistent, within
600: a significant uncertainty,
601: with the expected number---three. However, a
602: definitive value awaits further analysis.
603:
604: Having warmed up on the $\gamma -3\pi$ process, it is now
605: straightforward to construct the analogous $\eta\rightarrow\pi^+\pi^-\gamma$
606: amplitude. Using the extended $\chi$PT assumption we find\cite{34}
607: \begin{equation}
608: {\rm Amp}(\eta\rightarrow\pi^+\pi^-\gamma )=B(s_{+-},s_{+\gamma},s_{-\gamma})
609: \epsilon^{\mu\nu\alpha\beta}\epsilon^*_\mu p_{+\nu}p_{-\alpha}k_{\gamma\beta}
610: \end{equation}
611: with the anomaly stricture yielding
612: \begin{equation}
613: B_\eta(0,0,0)={eN_c\over 12\sqrt{3}\pi^2F_\pi^3 }\left( {F_\pi\over F_8}\cos\theta
614: -\sqrt{2}{F_\pi \over F_0}\sin\theta \right)\label{eq:vda}
615: \end{equation}
616: However, the physical region for the decay---$4m_\pi^2\leq
617: s_{\pi\pi}\leq m_\eta^2$---is far from the zero-momentum point which
618: is constrained by the anomaly. One indication of this fact is that
619: the decay rate obtained via neglect of momentum
620: dependence---$\Gamma^{(0)}_{\eta\rightarrow\pi\pi\gamma}$=35.7 eV---is
621: significantly different from the experimental
622: value---$\Gamma_{\eta\rightarrow\pi\pi\gamma}^{expt}=64\pm 6$ eV. For
623: the $\eta'$ channel the situation is, of course, much worse. The experimental
624: value---$\Gamma^{expt}_{\eta'\rightarrow\pi\pi\gamma}=61\pm 5$ keV---is a
625: factor of twenty larger than the
626: value $\Gamma^{(0)}_{\eta'\rightarrow\pi^+\pi^-\gamma}$=3 KeV
627: obtained via use of the simple anomaly prediction
628: \begin{equation}
629: B_{\eta'}(0,0,0)={eN_c\over 12\sqrt{3}\pi^2F_\pi^3 }\left( {F_\pi\over F_8}\sin\theta
630: +\sqrt{2}{F_\pi \over F_0}\cos\theta \right)\label{eq:vdb}
631: \end{equation}
632: Thus proper inclusion of momentum dependence is essential.
633: The $\eta\rightarrow\pi^+\pi^-\gamma$ spectrum was measured in the experiment
634: of Gormley et al.\cite{34} and was found to be approximately
635: fit in terms of a pure (width-modified) $\rho$-dominated matrix element.
636: This result is {\it not} in agreement, however, with the simple vector
637: dominance prediction---{\it cf.} Eq. \ref{eq:vdp}---which would require
638: \begin{equation}
639: |B(s,t,u)|_{\rm theo}^2\sim 1+3{s\over m_\rho^2}+\cdots
640: \end{equation}
641: and corresponds instead to
642: \begin{equation}
643: |B(s,t,u)|^2_{\rm exp}\sim 1+2{s\over m_\rho^2}+\cdots .
644: \end{equation}
645: Thus a careful look at the unitarization procedure is called for.
646:
647: We begin by noting that a one-loop chiral perturbation
648: theory calculation gives
649: \begin{eqnarray}
650: B_\eta^{\rm 1-loop}(s,s_{\pi\pi})&=&B_\eta(0,0)[1+{1\over
651: 32\pi^2F_\pi^2}
652: ((-4m_\pi^2+{1\over3}s_{\pi\pi})\ln{m_\pi^2\over m_\rho^2}\nonumber\\
653: &+&{4\over
654: 3}F(s_{\pi\pi})-{20\over 3}m_\pi^2+{3\over 2m_\rho^2}s_{\pi\pi}]
655: \end{eqnarray}
656: where
657: \begin{equation}
658: F(s)=\left\{
659: \begin{array}{ll}
660: (1-{s\over 4m_\pi^2})\sqrt{s-4m_\pi^2\over s}\ln{1+\sqrt{s-4m_\pi^2\over
661: s}\over -1+\sqrt{s-4m_\pi^2\over s}}-2 & s>4m_\pi^2 \\
662: 2(1-{s\over 4m_\pi^2})\sqrt{4m_\pi^2-s\over s}\tan ^{-1}\sqrt{s\over
663: 4m_\pi^2-s}-2& s \leq 4m_\pi^2
664: \end{array}\right.
665: \end{equation}
666: while the vector dominance picture ({\it cf.} Figure 5) yields
667: \begin{equation}
668: B_\eta(s_{\pi\pi})=B_\eta(0,0,0)\left(1+
669: {3\over 2}{s_{\pi\pi}\over m_\rho^2-s_{\pi\pi}}\right)\label{eq:vdp}
670: \end{equation}
671:
672: \begin{figure}[htb]
673: \begin{center}
674: \epsfig{file=etafig5.eps,height=10cm,width=11cm}
675: \caption{Shown are contact (a) and VMD (b,c) contributions to
676: $\eta,\eta'\rightarrow\pi^+\pi^-\gamma$ decay.}
677: \end{center}
678: \label{fig1}
679: \end{figure}
680:
681: Certainly, in order to treat the decay of the $\eta'$, one {\it must}
682: include unitarity effects via final state interactions. One very
683: obvious approach is simply to include the (energy-dependent) width of the
684: rho-meson in the propagator in the vector-dominance form
685: Eq. \ref{eq:vda} via
686: \begin{equation}
687: {s_{\pi\pi}\over m_\rho^2-s_{\pi\pi}}\rightarrow
688: {s_{\pi\pi}\over m_\rho^2-s_{\pi\pi}-im_\rho\Gamma_\rho(s_{\pi\pi})}
689: \end{equation}
690: This use of vector width-modified vector dominance
691: already makes an important difference from the simple anomaly---tree
692: level---results (especially in the case of
693: the $\eta'$), changing the predicted decay widths from the values
694: 35 eV and 3 KeV quoted above to the much more realistic numbers
695: \begin{equation}
696: \Gamma^{theo-VM}_{\eta-\pi\pi\gamma}=62.3\,\,{\rm eV}, \qquad
697: \Gamma^{theo-VM}_{\eta'-\pi\pi\gamma}=67.5\,\,{\rm KeV}
698: \end{equation}
699: if the parameters
700: \begin{equation}
701: F_8/F_\pi=1.3,\qquad F_0/F_\pi=1.04,\qquad \theta=-20^\circ\label{eq:hh}
702: \end{equation}
703: are employed. However, this procedure
704: does not match onto the one-loop chiral form in the low energy limit.
705:
706: In order to determine a form for the final state interactions which
707: matches onto both the one-loop chiral correction {\it and} to the vector
708: dominance result in the appropriate limits, we postulate an N/D
709: structure
710: \begin{equation}
711: B_{\eta-\pi\pi\gamma}(s,s_{\pi\pi})=B_{\eta-\pi\pi\gamma}(0,0)\left[
712: 1-c+c{1+as_{\pi\pi}\over D_1(s_{\pi\pi})}\right]\label{eq:ii}
713: \end{equation}
714: where $D_1(s)$ is the Omnes function and is defined in terms of the
715: p-wave $\pi\pi$ phase shifts via\\cite{fnt2}
716: \begin{equation}
717: D_1(s)=\exp\left(-{s\over
718: \pi}\int_{4m_\pi^2}^\infty{ds'\delta(s')\over s'(s'-s-i\epsilon)}\right)
719: \end{equation}
720: and $a,c$ are free parameters to be determined. In
721: order to reproduce the coefficient of the $F(s_{\pi\pi})$ function, which
722: contains the rho width, we require $c=1$. On the other hand, matching
723: onto the VMD result at ${\cal O}(p^6)$ can be achieved by the choice
724: $a=1/2m_\rho^2$. Thus in the case of the $\eta$ the form is
725: completely determined. Since the $\eta'$ spectrum is closely related
726: and is dominated by the presence of the rho we shall postulate an
727: identical form for the $\eta'$ case. Using these forms we can
728: then calculate the decay widths assuming the theoretical values for
729: the anomaly. Using the parameters given in Eq. \ref{eq:hh}
730: one finds, for example,
731: \begin{eqnarray}
732: i)D_1^{\rm exp}(s) && \Gamma_{\eta-\pi\pi\gamma}=65.7\,\,{\rm eV}, \quad
733: \Gamma_{\eta'-\pi\pi\gamma}=66.2 KeV\nonumber\\
734: ii)D_1^{\rm anal}(s) && \Gamma_{\eta-\pi\pi\gamma}=69.7\,\,{\rm eV}, \quad
735: \Gamma_{\eta'-\pi\pi\gamma}=77.8 KeV
736: \end{eqnarray}
737: There is a tendency then for the numbers obtained via the analytic
738: form of the Omnes function to be somewhat too high.
739:
740: \begin{figure}[htb]
741: \leftline{
742: \begin{minipage}[t]{.47\linewidth}
743: \mbox{\leftline{\epsfig{file=etafig6a.eps,width=6.0cm}}
744: \hspace{0.6cm}
745: \rightline{\epsfig{file=etafig6b.eps,width=6.0cm}}}
746: \end{minipage}}
747: \caption{Shown is the photon spectrum in
748: $\eta\rightarrow\pi^+\pi^-\gamma$ from Gormley et al.\protect\cite{34} as well as
749: various theoretical fits. In the Figure 6a, the dashed line represents the
750: (width-modified) VMD model. The (hardly visible) dotted line and the
751: solid line represent the
752: final state interaction ansatz Eq.\ref{eq:ii} with use of the
753: analytic and experimental version of the Omnes function respectively.
754: Figure 6b shows the experimental Omnes
755: function result (solid line) compared with the one-loop result (dotdash line).}
756: \label{fig2}
757: \end{figure}
758:
759: \begin{figure}[htb]
760: \centering
761: \leavevmode
762: \centerline{\epsfbox{etafig7.eps}}
763: \caption{Shown is the photon spectrum in
764: $\eta'\rightarrow\pi^+\pi^-\gamma$ from Abele et al.\protect\cite{83} as well
765: as various theoretical fits. As in Figure 6a, the dashed line represents the
766: (width-modified) VMD model. The dotted and solid lines represent
767: the final state interaction ansatz Eq.\protect\ref{eq:ii} with use
768: of the analytic and experimental version of the Omnes function, respectively.
769: Here the curves have been normalized to the same number of events.}
770: \label{fig3}\end{figure}
771:
772: We can also
773: compare the predicted spectra with the corresponding experimentally
774: determined values. As shown in Figure 6, we observe that the
775: experimental spectra are well fit in the $\eta$ case
776: in terms of both the N/D or the VMD
777: forms, but that the one-loop chiral expression does not provide an
778: adequate representation of the data. In the case of the corresponding
779: $\eta'$ decay the results are shown in Figure 7, wherein we observe
780: that either the unitarized VMD or the use of $N/D_1^{\rm exp}$ provides
781: a reasonable fit to the data (we get $\chi^2$/dof=32/17 and 20/17,
782: respectively), while the use of the analytic form for the Omnes
783: function yields a predicted spectrum ($\chi^2$/dof=104/17)
784: which is slightly too low on the high energy end. However, for both
785: $\eta$ and $\eta'$ we see that our simple ansatz---Eq.\ref{eq:ii}---provides
786: a very satisfactory representation of the decay spectrum.
787: Our conclusion in the previous section was that if the mixing angle and
788: pseudoscalar coupling constants were assigned values consistent with
789: present theoretical and experimental leanings, then the predicted
790: widths and spectra of
791: both $\eta,\eta\rightarrow\pi^+\pi^-\gamma$ are basically consistent with
792: experimental values. Our goal in this section is to go the other way,
793: however. That is, using the assumed N/D forms for the decay amplitude,
794: and treating the pseudoscalar decay constants $F_8,F_0$ as well as the
795: $\eta-\eta'$ mixing angle $\theta$ as free parameters, we wish to
796: inquire as to
797: how well they can be constrained purely from the experimental data
798: on $\eta,\eta'\rightarrow\gamma\gamma$ and
799: $\eta,\eta'\rightarrow\pi^+\pi^-\gamma$ decays, with reasonable
800: assumptions made about the final state interaction effects in these
801: two channels.
802:
803: On theoretical grounds, one is somewhat more confident about the
804: extraction of the threshold amplitude in the case of the lower energy
805: $\eta\rightarrow\pi^+\pi^-\gamma$ system. Indeed, in this case the
806: physical region extends only slightly into the tail of the rho unlike
807: the related $\eta'$ decay wherein the spectrum extends completely over the
808: resonance so that there exists considerable sensitivity to details
809: of the shape. Thus a first approach might be to utilize only the
810: two-photon decays together with the
811: $\eta\rightarrow\pi^+\pi^-\gamma$ width in order to determine the
812: three desired parameters. In this fashion one finds the results shown
813: in Table 1.
814: \begin{table}
815: \begin{center}
816: \begin{tabular}{|c|c|c|c|} \hline
817: &$F_8/F_\pi$ & $F_0/F_\pi$ & $\theta$\\ \hline
818: VMD & $1.28\pm 0.24$ & $1.07\pm 0.48$ & $-20.3^\circ\pm 9.0^\circ$\\
819: N/D$_1^{\rm anal}$ & $1.49\pm 0.29$ & $1.02\pm 0.42$ &
820: $-22.6^\circ\pm 9.6^\circ$\\
821: N/D$_1^{\rm exp}$ & $1.37\pm 0.26$& $1.02\pm 0.45$& $-21.2^\circ\pm
822: 9.3^\circ$\\ \hline
823: \end{tabular}
824: \caption{Values of the renormalized pseudoscalar coupling
825: constants and the $\eta-\eta'$ mixing angle using the
826: $\eta,\eta'-\gamma\gamma$ and $\eta-\pi\pi\gamma$ amplitudes in a
827: three parameter fit.}
828: \end{center}
829: \end{table}
830: We observe that the results are in agreement, both with each other and
831: with the chiral symmetry expectations---$F_8/F_\pi\sim 1.3$,
832: $F_0/F_\pi\sim 1$, and $\theta\sim -20^\circ$. However, the
833: uncertainties obtained in this way are uncomfortably high.
834:
835: In order to ameliorate this problem, we have also done a maximum likelihood fit
836: including the $\eta'-\pi\pi\gamma$ decay rate, yielding the results
837: shown in Table 2.
838: \begin{table}
839: \begin{center}
840: \begin{tabular}{|c|c|c|c|} \hline
841: &$F_8/F_\pi$& $F_0/F_\pi$& $\theta$\\ \hline
842: VMD & $1.28\pm 0.20$& $1.07\pm 0.04$ & $-20.8^\circ\pm 3.2^\circ$\\
843: N/D$_1^{\rm anal}$& $1.48\pm 0.24$& $1.09\pm 0.03$& $-24.0^\circ\pm
844: 3.0^\circ$\\
845: N/D$_1^{\rm exp}$&$1.38\pm 0.22$&$1.06\pm 0.03$&$-22.0^\circ\pm
846: 3.3^\circ$\\ \hline
847: \end{tabular}
848: \caption{Values of the renormalized pseudoscalar coupling
849: constants and of the $\eta-\eta'$ mixing angle obtained from a maximum
850: likelihood analysis using the $\eta,\eta'-\gamma\gamma$ and
851: $\eta,\eta'-\pi\pi\gamma$ amplitudes.}
852: \end{center}
853: \end{table}
854: We observe that the central values stay fixed but that the error bars
855: are somewhat reduced. The conclusions are the same,
856: however---substantial renormalization for $F_8\sim 1.3 F_\pi$, almost
857: none for $F_0\sim F_\pi$, and a mixing angle $\theta\sim -20^\circ$.
858: These numbers appear nearly invariant, regardless of the approach.
859:
860: An interesting aside here is the recent observation by B\"{a}r and Wiese
861: that the $\pi^0\rightarrow\gamma\gamma$ reaction alone does not verify
862: the three color hypothesis---in a careful analysis the $N_c$-dependence
863: of the Wess-Zumino-Witten term is completely canceled by the
864: $N_c$-dependent part of a Goldstone-Wilczek term, and that it is only the
865: $\eta\rightarrow\pi^+\pi^-\gamma$ measurement which truly confirms the
866: result that $N_c=3$\cite{baw}.
867:
868: Having above confirmed the basic correctness of the predictions of the anomaly
869: (and thereby of this important cornerstone of QCD) we move now to the important
870: three pion decay of the eta, which occurs independent of the anomaly and which
871: rather probes the conventional two- and four-derivative piece of the chiral
872: Lagrangian.
873:
874: \subsection{\bf $\eta\rightarrow\pi\pi\pi$}
875: The decay of the isoscalar eta to the predominantly I=1 final state of the
876: three pion system occurs primarily on account of
877: the d-u quark mass difference\cite{fnt3}, and the result arising from
878: lowest order
879: chiral perturbation theory is well-known\cite{os}
880: \begin{equation}
881: {\rm Amp}(\eta_8\rightarrow\pi^a\pi^b\pi^c)=\delta^{ab}\delta^{c3}C(s_{ab},
882: s_{ac},s_{bc})+{\rm Permutations}
883: \end{equation}
884: where
885: \begin{equation}
886: C(s,t,u)=-{B_0(m_d-m_u)\over
887: 3\sqrt{3}F_\pi^2}\left[1+{3(s-s_0)\over m_\eta^2-m_\pi^2}\right]\label{eq:qq}
888: \end{equation}
889: and we have defined
890: \begin{equation}
891: s_{ab}=(p_a+p_b)^2\qquad{\rm and}\qquad s_0={1\over 3}(m_\eta^2+m_{\pi^+}^2
892: +m_{\pi^-}^2+m_{\pi^0}^2).
893: \end{equation}
894: Equivalently, we can write the prefactor of Eq. \ref{eq:qq} in a form
895: which respects the reparameterization invariance of Kaplan
896: and Manohar\cite{km}
897: \begin{equation}
898: {B_0(m_d-m_u)\over 3\sqrt{3}F_\pi^2}=-{1\over Q^2}{m_K^2\over m_\pi^2}(m_K^2-m_\pi^2)
899: \end{equation}
900: where
901: \begin{equation}
902: Q^2={m_s-\hat{m}\over m_d-m_u}{m_s+\hat{m}\over m_d+m_u}
903: \end{equation}
904: and $\hat{m}={1\over 2}(m_d+m_u)$ is the average $u,d$ quark mass.
905:
906: Thus the decay $\eta\rightarrow 3\pi$ can be used in order to
907: determine the quantity $Q^2$. Alternatively, if $Q^2$ is given from
908: some other process then the $\eta\rightarrow 3\pi$ amplitude is completely
909: determined. The standard approach to such a determination is to utilize
910: the pseudoscalar meson masses via the relation
911: \begin{equation}
912: Q^2={m_K^2\over m_\pi^2}{m_K^2-m_\pi^2\over (m_{K^0}^2-m_{K^+}^2)_{\rm QCD}}
913: \end{equation}
914: where $(m_{K^0}^2-m_{K^+}^2)_{\rm QCD}$ is the nonelectromagnetic
915: component of the $K^0,K^+$ mass difference---
916: \begin{equation}
917: (m_{K^0}^2-m_{K^+}^2)_{\rm QCD}=(m_{K^0}^2-m_{K^+}^2)_{\rm expt}
918: -(m_{K^0}^2-m_{K^+}^2)_{\rm em}\label{eq:da}
919: \end{equation}
920: In order to evaluate the right hand side of Eq. \ref{eq:da} one
921: generally uses Dashen's theorem, which guarantees the identity of the
922: electromagnetic piece of the kaon and pion electromagnetic mass shifts
923: in the chiral symmetric limit\cite{da}
924: \begin{equation}
925: (m_{\pi^+}^2-m_{\pi^0}^2)=(m_{K^+}^2-m_{K^0}^2)_{\rm EM}.
926: \end{equation}
927: This simple assumption gives then
928: \begin{equation}
929: Q^2_{\rm Dashen}={m_K^2\over m_\pi^2}{m_K^2-m_\pi^2\over m_{K^0}^2-m_{K^+}^2+
930: m_{\pi^+}^2-m_{\pi^0}^2}=24.1
931: \end{equation}
932: and results in a prediction
933: \begin{equation}
934: \Gamma (\eta\rightarrow\pi^+\pi^-\pi^0)=66\,\,{\rm eV}
935: \end{equation}
936: in strong contradiction to the experimental result
937: \begin{equation}
938: \Gamma^{exp}(\eta\rightarrow\pi^+\pi^-\pi^0)=281\pm 28\,\,{\rm eV}.
939: \end{equation}
940: At first sight this would appear to be a rather strong and irreparable
941: violation of a lowest order chiral prediction and therefore not salvagable
942: by the expected ${\cal O}(m_\eta^2/(4\pi F_\pi^2)^2)\sim 30\% $ corrections
943: from higher order effects. However, this is not at all the case. In fact
944: the one-loop and counterterm contributions were calculated by Gasser
945: and Leutwyler and were found to enhance the lowest order prediction by a
946: factor 2.6, yielding\cite{gle}
947: \begin{equation}
948: \Gamma^{theo}(\eta\rightarrow\pi^+\pi^-\pi^0 )
949: \approx 167\pm 50\,\,{\rm eV},\label{eq:kk}
950: \end{equation}
951: which is a significant improvement, but still somewhat too low. The origin of
952: such a large correction lies primarily with $\eta_8,\eta_0$ mixing which
953: generates a factor $(\cos\theta -\sqrt{2}\sin\theta )^2\sim 2$ leaving
954: the expected $30\%$ corrections due to conventional higher order loop
955: and counterterm contributions. However, recent work has indicated that
956: Eq. \ref{eq:kk} is probably a considerable underestimate
957: due a significant violation
958: of Dashen's theorem. Indeed, the Dashen requirement was derived in
959: the limit of chiral symmetry---$m_\pi^2=m_K^2=0$---and plausible estimates
960: of chiral breaking effects have yielded the estimate\cite{dhw}
961: \begin{equation}
962: Q^2_{\chi -{\rm broken}}\approx 0.8Q^2_{\rm Dashen},\quad {\it i.e.}\quad
963: Q_{\chi -{\rm broken}}\sim 21.7\label{eq:pp}
964: \end{equation}
965: which corresponds to an additional $\sim$40\% enhancement of the chiral calculation
966: Eq. \ref{eq:kk}, {\it i.e.}
967: $\Gamma^{theo}(\eta\rightarrow\pi^+\pi^-\pi^0)
968: \sim 240\,\,{\rm eV}$,
969: and puts the result in the right ballpark. There are at least two possible
970: sources for the existence of any remaining discrepancy. One is the fact that
971: the estimate for the size of Dashen's theorem violation is just that---an
972: estimate. It is possible that the size of the violation is more significant than
973: that given in Eq. \ref{eq:pp}, leading to an even
974: larger value for $\Gamma (\eta\rightarrow
975: \pi^+\pi^-\pi^0 )$. A second possibility is that the simple one loop estimate
976: given in ref. 4 is not sufficient to include the full impact of final state
977: interaction effects. This has been demonstrated in other
978: processes where the I=0 S-wave $\pi -\pi$ plays an important role, as it does
979: here\cite{tr}. Indeed the closely related $K\rightarrow 3\pi$ reaction is one
980: such case\cite{ns}.
981:
982: In order to decide which---if either---possibility obtains it is necessary
983: to make careful spectral shape measurements in addition to simple
984: lifetime numbers. Also it is necessary to confront such results with
985: precise theoretical calculations. Phenomenologically, we expand the
986: decay amplitude about the center of the Dalitz plot as
987: \begin{equation}
988: C(s,t,u)\equiv \alpha\left[1 +\beta y +\gamma y^2+ \delta
989: x^2+\ldots \right]
990: \end{equation}
991: where
992: \begin{equation}
993: y={3(s-s_0)\over 2m_\eta \Delta_\eta}\qquad{\rm and}\qquad
994: x={\sqrt{3}(t-u)\over 2m_\eta \Delta_\eta}.
995: \end{equation}
996: where $\Delta_\eta=m_\eta-2m_{\pi^\pm}-m_\pi^0$ is the Q-value.
997: These parameters have been determined experimentally
998: to be
999: \begin{eqnarray}
1000: \mbox{Layter et al.}\cite{exp1}: \beta &=& 0.54\pm 0.007\qquad\gamma
1001: = 0.017\pm 0.014\qquad\delta =0.023\pm 0.016\nonumber\\
1002: \mbox{Gormley et al.}\cite{exp2}: \beta&=& 0.585\pm 0.010\qquad \gamma
1003: = 0.105\pm 0.015\qquad\delta
1004: =0.03\pm 0.02\nonumber\\
1005: \mbox{Amsler et al.}\cite{exp3}: \beta&=&0.470\pm
1006: 0.075\qquad\gamma=0.055\pm 0.135
1007: \end{eqnarray}
1008: to be compared to the one-loop chiral prediction
1009: \begin{equation}
1010: \beta = 0.665 \qquad\gamma = 0.21 \qquad\delta = 0.04
1011: \end{equation}
1012: Clearly there is general (though certainly not excellent) agreement,
1013: suggesting the importance of higher order scattering contributions.
1014:
1015: These have been examined by two Swiss collaborations using dispersion
1016: relation treatments in order to address the problem of higher
1017: order three-body scattering effects. The calculation of Anisovich and
1018: Leutwyler quotes only the integrated decay rate which is is agreement
1019: with experiment if the value $Q=22.7\pm 0.8$ is chosen\cite{al}---consistent
1020: with the Dashen theorem violation calculated in \cite{dhw}. Similarly
1021: the integrated decay rate found in the Khuri-Treiman calculation by
1022: Kambor, Wiesendanger, and Wyler agrees with the experimental rate if
1023: the value $Q=22.4\pm 0.9$ is used\cite{kww}. However, these authors
1024: also quote values for the spectral shape
1025: \begin{eqnarray}
1026: \beta&=&0.58\quad\gamma=0.12\quad\delta=0.045\nonumber\\
1027: \beta&=&0.58\quad\gamma=0.115\quad\delta=0.05
1028: \end{eqnarray}
1029: where the two sets of numbers correspond to different ways of
1030: determining the experimental input. Obviously the slope parameter
1031: $\beta$ is in general agreement with experiment. However, the
1032: situation is more complex for the quadratic components. In particular
1033: the calculated values for $\gamma ,\,\delta$ are in good agreement with the
1034: measurement of Gormley et al. (or Amsler et al.) but {\it not} with
1035: that of Layter et al. However, experimental uncertainties are
1036: significant and a new high statistics determination of the spectral
1037: shape is needed.
1038:
1039: Additional information is available by studying the neutral decay mode
1040: $\eta\rightarrow3\pi^0$, for which Bose symmetry determines
1041: the decay amplitude to be
1042: \begin{equation}
1043: {\rm Amp}(\eta\rightarrow 3\pi^0)=N^{000}|1+\epsilon(x^2+y^2)_{\rm sym}|^2
1044: \end{equation}
1045: where
1046: \begin{equation}
1047: (x^2+y^2)_{\rm sym}={1\over 3}\sum_{i=1}^3(x_i^2+y_i^2)=2y_{\rm sym}^2
1048: \end{equation}
1049: Here the dispersive calculation by Kambor et al. predicts
1050: \begin{equation}
1051: \epsilon=-0.028,\quad {\rm or}\quad \epsilon=-0.014
1052: \end{equation}
1053: depending on how the experimental input is handled. On the
1054: experimental side the only measurement of the energy dependence until
1055: recently had been of limited accuracy
1056: \begin{eqnarray}
1057: \mbox{Amsler et al.}\cite{exp4}: \epsilon&=&-0.044\pm 0.046\nonumber\\
1058: \mbox{Baglin et al.}\cite{exp5}: \epsilon&=&-0.64\pm 0.74\nonumber\\
1059: \mbox{Abele et al.}\cite{exp6}: \epsilon&=&-0.104\pm 0.04
1060: \end{eqnarray}
1061: which is consistent with (but with large experimental uncertainty) the
1062: dispersive calculation. However, recently a new result of
1063: uncprecedented precision has been announced from the Crystal Ball
1064: group at BNL\cite{cb}
1065: \begin{equation}
1066: \epsilon=-0.062\pm 0.006\pm 0.004
1067: \end{equation}
1068:
1069: Clearly this number is significantly larger than expected from the
1070: Khuri-Treiman calculation, suggesting that new dynamical input is
1071: involved. This is not unexpected. Indeed the calculation of Kambor
1072: et al. utilized the effective chiral Lagrangian at ${\cal O}(p^4)$ is
1073: input. Clearly from the agreement with experiment this is the main
1074: effect, but one also expects contributions from pieces of the chiral
1075: Lagrangian of ${\cal O}(p^6)$ such as
1076: \begin{equation}
1077: {\cal L}^{(6)}\sim{F_\pi^2\over \Lambda_\chi^2}{\rm tr}[(\chi
1078: U^\dagger+U\chi^\dagger)D^\mu UD_\mu U^\dagger]{\rm tr}(D^\nu UD_\nu U^\dagger)
1079: \end{equation}
1080: where $\chi=2B_0m$, with $B_0$ being a constant and $m$ is the quark
1081: mass matrix, $\Lambda_\chi\sim 4\pi F_\pi\sim 1$ GeV is the chiral
1082: scale. The coefficients of such terms are
1083: unconstrained by the strictures of chiral invariance and are
1084: experimentally undetermined at present, since they arise at two-loop
1085: order. Nevertheless, their presence can lead to peices of the decay
1086: amplitude of the form
1087: \begin{eqnarray}
1088: A&\sim&{\cal A}_1k\cdot q_ca_a\cdot q_b+{\cal A}_2(k\cdot a_aq_b\cdot
1089: q_c+k\cdot q_b q_a\cdot q_c)\nonumber\\
1090: &\simeq&m_\eta^4\left({{\cal A}_1\over 18}+{{\cal A}_2\over
1091: 9}\right)\left[1-{Q_\eta^2\over m_\eta^2}(x^2+y^2)\right]\nonumber\\
1092: &+&{1\over
1093: 12}m_\eta^4({\cal A}_1-{\cal A}_2)\left[{2Q_\eta\over
1094: 3m_\eta}y-{8\over 27}{Q_\eta^2\over m_\eta^2}(y^2-x^2)\right]
1095: \end{eqnarray}
1096: If such a dynamical component is present then its size should be set
1097: by chiral scaling arguments
1098: $${\cal A}_I\sim {m_d-m_u\over \Lambda_\chi^2F_\pi^2}$$
1099: and an isospin relation
1100: \begin{equation}
1101: \gamma^{+-0}(dyn)+\delta^{+-0}(dyn)=\epsilon^{000}(dyn)
1102: \end{equation}
1103: must exist between the quadratic parameters for the charged and
1104: neutral channels. Here the symbol $dyn$ indicates the dynamical ({\it
1105: i.e.}, non-rescattering component of the coefficient in question and is
1106: found by subtracting the theoretical value obtained from the
1107: Khuri-Trieman calculation from the experimental quantity. In this way
1108: we find, using the Gormley numbers for experimental input,
1109: \begin{equation}
1110: \gamma^{+-0}(dyn)=-0.025\pm 0.015\quad\delta^{+-0}(dyn)=-0.02\pm 0.02
1111: \end{equation}
1112: and
1113: \begin{equation}
1114: \epsilon^{000}(dyn)=-0.034\pm 0.007\quad{\rm or}\quad -0.048\pm 0.007
1115: \end{equation}
1116: depending on the dynamical input chosen. The comparison
1117: \begin{equation}
1118: \gamma^{+-0}(dyn)+\delta^{+-0}(dyn)=-0.045\pm 0.03
1119: \end{equation}
1120: vs.
1121: \begin{equation}
1122: \epsilon^{000}(dyn)=-0.034\pm 0.007\quad{\rm or}\quad -0.045\pm 0.007
1123: \end{equation}
1124: is obviously satisfactory within errors but again cries out for a high
1125: precision measurement of the $\eta\rightarrow \pi^+\pi^-\pi^0$
1126: spectrum, such as would be possible using WASA.
1127:
1128: Of course, there is one additional test which we can use. Since
1129: according to isotopic spin invariance the $3\pi^0$ amplitude at the
1130: center of the Dalitz plot must be a factor of three larger than the
1131: corresponding $\pi^+\pi^-\pi^0$ number, the total decay rates should
1132: differ by the factor
1133: \begin{equation}
1134: {\Gamma^{(0)}(000)\over \Gamma^{(0)}(+-0)}={3^2\over 3!}=1.5
1135: \end{equation}
1136: When rescattering corrections are included, the prediction becomes
1137: \begin{equation}
1138: {\Gamma(000)\over \Gamma(+-0)}=\left\{\begin{array}{cc}
1139: 1.43& \mbox{one loop}\\
1140: 1.41\pm 0.03& \mbox{Khuri-Treiman}\end{array}\right.
1141: \end{equation}
1142: Both numbers are consistent with the value quoted by the Particle Data
1143: Group\cite{pdg}
1144: \begin{equation}
1145: \left({\Gamma(000)\over \Gamma(+-0)}\right)^{{\rm exp}}=1.404\pm 0.034
1146: \end{equation}
1147: and are in good agreement with the recent Crystal Ball measurement\cite{exp3}
1148: \begin{equation}
1149: \left({\Gamma(000)\over \Gamma(+-0)}\right)^{\rm exp}=1.44\pm 0.09\pm 0.01
1150: \end{equation}
1151:
1152: Clearly there is plenty of challenge in the three-pion sector for an
1153: eta facility, but there is also interest in examining the remaining
1154: radiative modes $\eta\rightarrow \pi^0\gamma\gamma , 3\pi\gamma$.
1155:
1156: \subsection{\bf $\eta\rightarrow\pi^0\gamma\gamma$}
1157:
1158: For the decay $\eta\rightarrow \pi^0\gamma\gamma$ chiral symmetry does not
1159: play a important role, but vector dominance {\it does}. This can be seen from
1160: the feature that there exists no contribution at all to this process from
1161: the tree level two-derivative Lagrangian. Rather the lowest order
1162: chiral contribution
1163: arises at one-loop level---$\eta\rightarrow 3\pi ,K\bar{K}\pi\rightarrow \pi^0
1164: \gamma\gamma$. However, the $3\pi$ intermediate intermediate state is
1165: suppressed by the factor $m_d-m_u$ while the contribution from $K\bar{K}\pi$
1166: is suppressed kinematically. To see this, we define the general decay
1167: amplitude
1168: \begin{eqnarray}
1169: & &{\rm Amp}(\eta\rightarrow\pi^0\gamma\gamma )=D(s,t,u)\left[\epsilon_1\cdot
1170: \epsilon_2q_1\cdot q_2-\epsilon_1\cdot q_2\epsilon_2\cdot q_1\right]\nonumber\\
1171: &-&E(s,t,u)\left[-\epsilon_1\cdot\epsilon_2p\cdot q_1p\cdot q_2-\epsilon_1
1172: \cdot p\epsilon_2\cdot pq_1\cdot q_2\right.\nonumber\\
1173: &+&\left.\epsilon_1\cdot q_2\epsilon_2\cdot p
1174: p\cdot q_1+\epsilon_1\cdot p\epsilon_2\cdot q_1p\cdot q_2\right]
1175: \end{eqnarray}
1176: Then from the one-pion-loop contributions from ${\cal L}^{(2)}$ we find
1177: \begin{equation}
1178: D_\pi(s,t,u)= {\sqrt{2}\alpha\over \pi}
1179: T(s) F(s,m_\pi^2), \qquad E_\pi (s,t,u)=0
1180: \end{equation}
1181: where $T(s)$ is the lowest order $\eta\rightarrow 3\pi$ amplitude given in
1182: Eq. \ref{eq:qq} and
1183: \begin{equation}
1184: sF(s,m_\pi^2)=1+{4m_\pi^2\over s}\ln^2\left({\beta (s) +1\over
1185: \beta (s) -1}\right) \quad {\rm with} \quad \beta (s)=\sqrt{s-4m_\pi^2\over s}
1186: \end{equation}
1187: with a similar expression obtaining for the kaon loop contribution.
1188: Calculation
1189: of the associated rate yields\cite{44}
1190: \begin{equation}
1191: \Gamma_{\rm loop}(\eta\rightarrow \pi^0\gamma\gamma )\approx
1192: 4\times 10^{-3}\,\,{\rm eV}\quad{\rm vs.}\quad \Gamma_{\rm exp}(\eta\rightarrow
1193: \pi^0\gamma\gamma )=0.84\pm 0.18\,\,{\rm eV}.
1194: \end{equation}
1195: Thus the one-loop chiral contribution plays a very minor role.
1196: Consider, however,
1197: the vector dominance diagram shown in Figure 8, for which
1198: \begin{eqnarray}
1199: D(s,t,u)&=&{2\sqrt{3}\over 9}g_{\omega\rho\pi}^2\left({2eF_\pi^2g\over m_V^2}
1200: \right)^2\left[
1201: {p\cdot q_2-m_\eta^2\over m_V^2-t}+{p\cdot q_1-m_\eta^2\over m_V^2-u}\right]\nonumber\\
1202: &\times&\left({F_\pi\over F_8}\cos\theta -\sqrt{2}{F_\pi\over F_0\sin\theta}\right)
1203: \nonumber\\
1204: E(s,t,u)&=&-{2\sqrt{3}\over 9}g_{\omega\rho\pi}^2\left({2eF_\pi^2g\over m_V^2}
1205: \right)\left[
1206: {1\over m_V^2-t}+{1\over m_V^2-u}\right]\nonumber\\
1207: &\times&\left({F_\pi\over F_8}\cos\theta
1208: -\sqrt{2}{F_\pi\over F_0}\sin\theta\right)
1209: \end{eqnarray}
1210: and yielding for the decay rate
1211: \begin{equation}
1212: \Gamma_{\rm VD}(\eta\rightarrow\pi^0\gamma\gamma )=0.31\,\,{\rm eV}
1213: \end{equation}
1214:
1215: \begin{figure}[htb]
1216: \begin{center}
1217: \epsfig{file=etafig8.eps,height=3cm,width=8cm}
1218: \caption{Vector dominance diagram responsible for
1219: $\eta^0\rightarrow\pi^0\gamma\gamma$.}
1220: \end{center}
1221: \end{figure}
1222:
1223:
1224: Inclusion of other higher order effects such as the contribution from
1225: a pair of anomalous
1226: terms---$\pi\pi\pi\gamma$ and $\eta\pi\pi\gamma$---coupled via a
1227: pion loop increases this estimate to about half the experimental
1228: result, but a considerable discrepancy remains and should be the focus of
1229: future experimental as well as theoretical work.
1230:
1231: A new number from the Brookhaven experiment was first
1232: announced at this meeting by Nefkens
1233: \begin{equation}
1234: \Gamma^{exp}(\eta\rightarrow \pi\gamma\gamma)=(0.38\pm 0.11)\,\,eV
1235: \end{equation}
1236: is about a factor of two smaller than the Particle Data Group value.
1237: The problem with the previous measurements is probably associated with
1238: eliminating the background from other much more probably neutral modes
1239: such as $\eta\rightarrow 3\pi^0$ and confirmation using WASA would
1240: clearly be welcome. In this regard, spectral shape measurements could
1241: be helpful, although this will be difficult, since this is a low
1242: branching ratio---$\sim 7\times 10^{-4}$---experiment.
1243:
1244: \medskip
1245: \subsection{\bf $\eta\rightarrow\pi\pi\pi\gamma$}
1246: \medskip
1247:
1248: The final mode which we shall mention in this report
1249: is $\eta\rightarrow 3\pi\gamma$, for
1250: which on the experimental side there exists at present only an upper
1251: bound\cite{pdg}
1252: \begin{equation}
1253: {\Gamma (\eta\rightarrow 3\pi\gamma )\over \Gamma (\eta\rightarrow
1254: \pi^+\pi^-\pi^0}|_{\rm exp} < 0.0024.
1255: \end{equation}
1256: Ordinarily the dominant component of a radiative mode such as this is due
1257: to the the inner bremsstrahlung process, for which the matrix element is
1258: \begin{eqnarray}
1259: {\rm Amp}(\eta\rightarrow \pi^+\pi^-\pi^0\gamma )&\simeq&{\rm Amp}(\eta\rightarrow
1260: \pi^+\pi^-\pi^0 )\nonumber\\
1261: &\times& ie\epsilon^\mu\left[{(2p_++k)_\mu\over (p_++k)^2
1262: -m_\pi^2}-{(2p_-+k)_\mu\over (p_-+k)^2-m_\pi^2}\right]
1263: \end{eqnarray}
1264: and it is experimentally difficult to distinguish any direct photon emission.
1265: However, an exception occurs when the non-radiative process is suppressed in
1266: some fashion, such as occurs, {\it e.g.},
1267: in the cases of $\pi^+\rightarrow e^+\nu_e\gamma$
1268: and $K_L\rightarrow \pi^+\pi^-\gamma$, wherein the nonradiative process is
1269: small because of helicity suppression and CP violation
1270: respectively\cite{ex}. One might have anticipated the same enhancement
1271: mechanism to apply in our case since
1272: the nonradiative reaction $\eta\rightarrow\pi^+\pi^-
1273: \pi^0$ takes place only due to the relatively small u-d quark mass difference.
1274: For example, the direct emission associated with the vector-dominance
1275: diagrams shown in Figure 9 could be expected to play an important role.
1276: However, a careful analysis by D'Ambrosio et al. has shown that this
1277: unfortunately does not appear to be the case\cite{dam}.
1278: The contribution from
1279: this direct---vector-dominated---mechanism is found to have the form
1280: \begin{eqnarray}
1281: A^\mu_{\rm direct}&(&\eta\rightarrow\pi^+\pi^-\pi^0\gamma )
1282: ={e64h_V\theta_V\over 3\sqrt{3}M_V^2F_\pi^4}\left[
1283: p_\eta\cdot p_0g^\mu_{+-}+p_\eta\cdot p_+g^\mu_{-0}+p_\eta\cdot
1284: p_-g^\mu_{0+}\right]\nonumber\\
1285: \quad
1286: \end{eqnarray}
1287: where
1288: \begin{equation}
1289: G_{ij}^\mu=k\cdot p_ip_j^\mu-k\cdot p_j p_i^\mu
1290: \end{equation}
1291: and the combination of Bose
1292: symmetry and gauge invariance results in a suppression that makes
1293: such pieces even smaller than the pion loop component, which arise to
1294: ${\cal O}(p^4)$ and which are proportional to $m_u-m_d$. In ref. \cite{dam}
1295: it is estimated that the direct emission (DE) component is only a few percent
1296: addition to the inner bremsstrahlung (IB), even at relatively large photon
1297: energies
1298: \begin{equation}
1299: \left[(\Gamma_{\rm IB+DE}-\Gamma_{\rm IB})/\Gamma_{\rm IB}
1300: \right]_{E_\gamma>90\,\,{\rm MeV}}\simeq 3.5\times 10^{-2}
1301: \end{equation}
1302: which appears to be too small to make this a realistic experimental goal.
1303:
1304: \begin{figure}[htb]
1305: \begin{center}
1306: \epsfig{file=etafig9.eps,height=4.3cm,width=5cm}
1307: \caption{Vector dominance diagrams responsible for
1308: $\eta^0\rightarrow\pi^+\pi^-\pi^0\gamma$.}
1309: \end{center}
1310: \end{figure}
1311:
1312: \section{Conclusions}
1313:
1314: We have examined the "non-rare" decay modes of the eta meson---$\eta\rightarrow
1315: \gamma\gamma , \pi^+\pi^-\gamma , 3\pi , \pi^0\gamma\gamma ,
1316: 3\pi\gamma$---in light of current theoretical knowledge and within the
1317: general framework of chiral symmetry. While there exist no
1318: {\it striking} discrepancies observed with respect to any of
1319: these predictions, we
1320: clearly identified problem areas wherein additional experimental
1321: scrutiny, such as would be possible at CELSIUS, would
1322: add significantly to our understanding. These conclusions
1323: can be summarized succinctly as follows:
1324: \medskip
1325:
1326: ${\bf a:\,\, \eta\rightarrow \gamma\gamma}$---a resolution of the discrepancy between
1327: the rates measured via the Primakoff effect\cite{22} and via QED
1328: production\cite{21} is essential to future progress in understanding the eta
1329: system in general.
1330:
1331: \medskip
1332:
1333: ${\bf b:\,\,\eta\rightarrow \pi^+\pi^-\gamma}$---a careful spectral
1334: shape measurement would be useful in order check the spectral shape
1335: with that predicted on fairly solid theoretical grounds from the
1336: anomalous sector of QCD.
1337:
1338: \medskip
1339:
1340: ${\bf c:}\,\,\eta\rightarrow 3\pi$---a precise spectral shape measurement
1341: is called for in order to determine whether the existing disagreement between experimental
1342: findings and (what should be) solid theoretical predictions based on chiral
1343: perturbation theory are cause by an inaccurate value for the d-u quark mass
1344: difference or are due to the importance of higher order final state interaction
1345: effects.
1346:
1347: \medskip
1348:
1349: ${\bf d: \,\,\eta\rightarrow \pi^0\gamma\gamma}$---a precision measurement of the
1350: Dalitz plot distribution is suggested in order to learn the origin of the
1351: existing disagreement between the experimental rate and that predicted from
1352: vector dominance.
1353:
1354: \medskip
1355:
1356: ${\bf e:}\,\, \eta\rightarrow 3\pi\gamma$---a determination of an
1357: actual rate instead
1358: of the existing upper bound would be of interest, but theoretical
1359: indications are that it will be difficult to detect other than the
1360: inner bremsstrahlung component.
1361: \medskip
1362:
1363: Clearly there is lots of interesting physics here and a marriage between
1364: precise and solid new experimental data and careful and well-motivated
1365: theoretical analysis would, I predict, be a happy one, leading to the
1366: offspring of a
1367: new degree of understanding of an important component of low energy
1368: phenomenology.
1369:
1370:
1371:
1372:
1373:
1374:
1375: \medskip
1376:
1377:
1378: {\bf Acknowledgements}: We thank John Donoghue for many clarifying
1379: discussions and the theory group at J\"{u}lich for their warm hospitality.
1380: This work was supported in part by the Alexander von Humboldt Foundation and
1381: by the National Science Foundation under grant PHY98-01875.
1382:
1383: \begin{thebibliography}{99}
1384:
1385: \bibitem{1} See, {\it e.g.}, M. Creutz, ``Quarks, Gluons and Lattices'' (Cambridge
1386: University Press, Cambridge, 1983).
1387: \bibitem{2} H.D. Politzer, Phys. Rep. {\bf 14}, 129 (1974).
1388: \bibitem{3} S. Weinberg, Physica {\bf A96}, 327 (1979).
1389: \bibitem{4} J. Gasser and H. Leutwyler, Ann. Phys. (NY) {\bf 150}, 142 (1984);
1390: Nucl. Phys. {\bf B250}, 465 (1985).
1391: \bibitem{5} J. Goldstone, Nuovo Cim. {\bf 19}, 154 (1961); J. Goldstone, A. Salam
1392: and S. Weinberg, Phys. Rev. {\bf 127}, 965 (1961).
1393: \bibitem{6} See, {\it e.g.}, S.B. Treiman in ``Current Algebras and Anomalies''
1394: (Ed. by S.B. Treiman et al.) (Princeton Univ. Press, Princeton, 1985).
1395: \bibitem{7} B.R. Holstein, Phys. Lett. {\bf B244}, 83 (1990).
1396: \bibitem{8} See, {\it e.g.}, S. Gasiorowicz and D.A. Geffen, Rev. Mod. Phys. {\bf 41},
1397: 531 (1969).
1398: \bibitem{9} S. Weinberg, Phys. Rev. Lett. {\bf 17}, 616 (1966).
1399: \bibitem{10} E. Witten, Nucl. Phys. {\bf B223}, 422 (1983).
1400: \bibitem{11} N.K. Pak and P. Rossi, Nucl. Phys. {\bf B250}, 594 (1985).
1401: \bibitem{12} See, {\it e.g.}, B.R. Holstein, Int. J. Mod. Phys. {\bf A7}, 7873 (1992).
1402: \bibitem{13} J.F. Donoghue and B.R. Holstein, Phys. Rev. {\bf D40}, 2378 and
1403: 3700 (1989).
1404: \bibitem{14} B.R. Holstein, Comm. Nucl. Part. Phys. {\bf 19}, 221 (1990).
1405: \bibitem{15} See, {\it e.g.}, D. Babusci et al., Phys. Lett. {\bf B277}, 158 (1992);
1406: J.F. Donoghue and B.R. Holstein, UMass Preprint UMHEP-383 (1993).
1407: \bibitem{16} See, {\it e.g.}, {\it PILAC Users Group Report on the Physics with
1408: PILAC}, Los Alamos Report LA-UR-92-150.
1409: \bibitem{17} See, {\it e.g.}, P. Herczeg, in ``Rare Decays of Light Mesons''
1410: (ed. B. Mayer) (Editions Frontiers, 1991), p. 97.
1411: \bibitem{gmo} M. Gell-Mann, CalTech Rept {\bf CTSL-20} (1961);
1412: S. Okubo, Prog. Theo. Phys. {\bf 27}, 949 (1962).
1413: \bibitem{18} J.F. Donoghue, B.R. Holstein and Y.-C.R. Lin, Phys. Rev. Lett.
1414: {\bf 55}, 2766 (1985).
1415: \bibitem{19} S.L. Adler and W.A. Bardeen, Phys. Rev. {\bf 182}, 1517 (1969).
1416: \bibitem{20} Particle Data Group, Phys. Rev. {\bf D45} (1992).
1417: \bibitem{21} S. Baru et al., Z. Phys. {\bf C48}, 581 (1990).
1418: \bibitem{22} A. Browman et al., Phys. Rev. Lett. {\bf 32}, 1067 (1974).
1419: \bibitem {fnt1} Here we use the $\eta\rightarrow
1420: \gamma\gamma$ rate arising from QED production $e^+e^-\rightarrow
1421: e^+e^-\gamma^*\gamma^*\rightarrow e^+e^-\eta$\cite{21} rather than
1422: the value $0.324 \pm0.046\,\,{\rm keV}$ from the Primakoff
1423: effect\cite{22}. It is clearly important to resolve this problem.
1424: \bibitem{23} H. Behrend et al., Z. Phys. {\bf C49}, 401 (1991).
1425: \bibitem{24} M. Gell-Mann et al., Phys. Rev. Lett. {\bf 8}, 261 (1962).
1426: \bibitem{25} G. Ecker et al., Nucl. Phys. {\bf B321}, 311 (1989).
1427: \bibitem{26} M. Bando et al., Prog. Theo. Phys. {\bf 73}, 1540 (1985);
1428: T. Fujiwara et al., Prog. Theo. Phys. {\bf 73}, 926 (1985).
1429: \bibitem{27} K. Kawarabayashi and M. Suzuki, Phys. Rev. Lett. {\bf 16},
1430: 255 (1966); Riazuddin and Fayyazuddin, Phys. Rev. {\bf 147}, 1071 (1966).
1431: \bibitem{28} S.L. Adler et al., Phys. Rev. {\bf D4}, 3497 (1971); R. Aviv
1432: and A. Zee, Phys. Rev. {\bf D5}, 2372 (1971); M. Terent'ev, JETP Lett.
1433: {\bf 14}, 94 (1971).
1434: \bibitem{29} T.D. Cohen, Phys. Lett. {\bf B233}, 467 (1989); S. Rudaz,
1435: Phys. Lett. {\bf B145}, 281 (1984).
1436: \bibitem{30} A. Bramon, E. Pallante and R. Petronzio, Phys. Lett. {\bf B271},
1437: 237 (1991).
1438: \bibitem{31} Yu. M. Antipov et al, Phys. Rev. {\bf D36}, 21 (1987).
1439: \bibitem{32} Ll. Ametller et al., Phys. Lett {\bf B276}, 185 (1992);
1440: A. Bramon et al., ``The Da$\Phi$ne Physics Handbook'' (ed. L. Maiani,
1441: G. Pancheri and N. Paver) (INFN, Frascati, 1992), p. 305.
1442: \bibitem{33} P. Ko and T.N. Truong, Phys. Rev. {\bf 43}, R4 (1991).
1443: \bibitem{akt} L. Ametller, M. Knecht, and P talavera, Phys. Rev. {\bf D64},
1444: 094009 (2001).
1445: \bibitem{hol} B.R. Holstein, Phys. Rev. {\bf D53}, 4099 (1996).
1446: \bibitem{tru} T. Truong, in ``Quantum Chromodynamics'' (ed. H.M. Fried
1447: and B. Muller) (World Scientific,
1448: Singapore, 1999), p. 153.
1449: \bibitem{han} T. Hannah, Nucl. Phys. {\bf B593}, 577 (2001).
1450: \bibitem{fnt2} For ease of calculation it
1451: is sometimes useful to employ the simple analytic form
1452: \begin{equation}
1453: D_1(s)=1-{s\over m_\rho^2}-{s\over 96\pi^2F_\pi^2}\ln{m_\rho^2\over
1454: m_\pi^2}-{m_\pi^2\over 24\pi^2F_\pi^2}F(s)
1455: \end{equation}
1456: \bibitem{34} M. Gormley et al., Phys. Rev. {\bf D2}, 501 (1970);
1457: J.G. Layter et al., Phys. Rev. {\bf D7}, 2565 (1973).
1458: \bibitem{83} A. Abele et al., Phys. Lett. {\bf B402}, 195 (1997);
1459: S.I. Bityukov et al., Z. Phys. {\bf C50}, 451 (1991).
1460: \bibitem{baw} O. B\"{a}r and U.-J. Wiese, Nucl. Phys. {\bf B609}, 225 (2001).
1461: \bibitem{fnt3} The transition can also occur due to electromagnetic
1462: effects but it is generally agreed that these are small. See, {\it e.g.},
1463: J.S. Bell and D.G. Sutherland, Nucl. Phys. {\bf B4}, 315
1464: (1968); P. Dittner, P.H. Dondi, and S. Eliezer, Phys. Rev. {\bf D8},
1465: 2253 (1973); R. Baur, J. Kambor, and D. Wyler, Nucl. Phys. {\bf B460},
1466: 127 (1996).
1467: \bibitem{os} H. Osborn and D.J. Wallace, Nucl. Phys. {\bf B20}, 23 (1970);
1468: J.A. Cronin, Phys. Rev. {\bf 161}, 1483 (1967).
1469: \bibitem{km} D. Kaplan and A. Manohar, Phys. Rev. Lett. {\bf 56}, 1994 (1986).
1470: \bibitem{da} R. Dashen, Phys. Rev. {\bf 183}, 1245 (1969).
1471: \bibitem{gle} J. Gasser and H. Leutwyler, Nucl. Phys. {\bf B250}, 539 (1985).
1472: \bibitem{dhw} J.F. Donoghue, B.R. Holstein and D. Wyler, Phys. Rev. Lett.
1473: {\bf 69}, 3444 (1992) and Phys. Rev. {\bf D47}, 2089 (1993);
1474: J. Bijnens, Phys. Lett. {\bf B306}, 343 (1993).
1475: \bibitem{tr} T.N. Truong, Phys. Lett. {\bf B207}, 495 (1988).
1476: \bibitem{ns} A. Neveu and J. Scherk, Ann. Phys. (NY) {\bf 57}, 39 (1970).
1477: \bibitem{exp1} J. Layter et al., Phys. Rev. {\bf D7}, 2565 (1973).
1478: \bibitem{exp2} M. Gormley et al., Phys. Rev. {\bf D2}, 501 (1970).
1479: \bibitem{exp3} C. Amsler et al., Phys. Lett.{\bf B346}, 203 (1995).
1480: \bibitem{al} A.V. Anisovich and H. Leutwyler, Phys. Lett. {\bf B375},
1481: 335 (1996).
1482: \bibitem{kww} J. Kambor, C. Wiesendanger, and D. Wyler,
1483: Nucl. Phys. {\bf B465}, 215 (1996).
1484: \bibitem{cb} W.B. Tippens et al., Phys. Rev. Lett., {\bf 87}, 192001 (2001).
1485: \bibitem{exp4} D. Alde et al., Z. Phys. {\bf C25}, 225 (1984).
1486: \bibitem{exp5} C. Baglin et al., Nucl. Phys. {\bf B22}, 66 (1970).
1487: \bibitem{exp6} A. Abele et al., Phys. Lett. {\bf B417}, 193 (1998).
1488: \bibitem{43} J. Kambor et al., Phys. Rev. Lett. {\bf 68}, 1818 (1992).
1489: \bibitem{pdg} D.E. Groom et al., Eur. Phys. J {\bf C15}, 1 (2001).
1490: \bibitem{44} Ll. Ametller et al., Phys. Lett. {\bf B276}, 185 (1992).
1491: \bibitem{ex} See, {\it e.g.}, G. D'Ambrosio and J. Portoles,
1492: Nucl. Phys. {\bf B533}, 523 (1998); J.F. Donoghue and B.R. Holstein,
1493: Phys. Rev. {\bf D40}, 2378 (1989).
1494: \bibitem{dam} G. D'Ambrosio, G. Ecker, G. Isidori, and H. Neufeld,
1495: Phys. Lett. {\bf B466}, 337 (1999).
1496: \end{thebibliography}
1497:
1498: \end{document}
1499:
1500:
1501:
1502:
1503:
1504:
1505: