1: \documentclass[aps,prc,onecolumn,preprint,superscriptaddress,showpacs]{revtex4}
2: \usepackage[dvips]{graphicx}
3: \usepackage{bm}
4: \usepackage{amssymb}
5: \usepackage{amsmath}
6: \usepackage{nicefrac}
7: \usepackage{color}
8:
9: \usepackage{xspace}
10: \newcommand{\pb}{Papenbrock and Bertsch~\cite{PaBe1998}\xspace}
11: \newcommand{\dy}{Dyakonov~\cite{Dy1999}\xspace}
12:
13: \newcommand{\etal}{\textit{et al.\xspace}}
14:
15: %\bibliographystyle{myprsty}
16:
17: \begin{document}
18: \sloppy
19:
20: %
21: % quasiclassical description of bremsstrahlung in $\alpha$ decay
22: %
23: \title{Quasiclassical description of bremsstrahlung accompanying\\
24: $\boldsymbol \alpha$ decay including quadrupole radiation}
25:
26: \author{U. D. Jentschura}
27: \affiliation{Max-Planck-Institut f\"ur Kernphysik,
28: Postfach 103980, 69029 Heidelberg, Germany}
29: \affiliation{Institut f\"ur Theoretische Physik,
30: Philosophenweg 16, 69120 Heidelberg, Germany}
31:
32:
33: \author{A. I. Milstein}
34: \author{I.~S.~Terekhov}
35: \affiliation{Max-Planck-Institut f\"ur Kernphysik,
36: Postfach 103980, 69029 Heidelberg, Germany}
37: \affiliation{Budker Institute of Nuclear Physics, 630090 Novosibirsk, Russia}
38:
39: \author{H. Boie}
40: \author{H. Scheit}
41: \affiliation{Max-Planck-Institut f\"ur Kernphysik,
42: Postfach 103980, 69029 Heidelberg, Germany}
43: \author{D. Schwalm\footnote{present address:
44: Weizmann Institute of Science, 76100 Rehovot, Israel}$^{,}$}
45: \affiliation{Max-Planck-Institut f\"ur Kernphysik,
46: Postfach 103980, 69029 Heidelberg, Germany}
47:
48: \begin{abstract}
49: We present a quasiclassical theory of
50: $\alpha$ decay accompanied by bremsstrahlung with a
51: special emphasis on the case of $^{210}$Po, with the aim of finding
52: a unified description that incorporates both the
53: radiation during the tunneling through the Coulomb wall and the
54: finite energy $E_\gamma$ of the radiated photon up to $E_\gamma\sim
55: Q_\alpha/\sqrt{\eta}$, where $Q_\alpha$ is the $\alpha$-decay
56: $Q$-value and $\eta$ is the Sommerfeld parameter. The corrections
57: with respect to previous quasiclassical investigations are found to
58: be substantial, and excellent agreement with a full quantum
59: mechanical treatment is achieved. Furthermore, we find that a
60: dipole-quadrupole interference significantly changes the
61: $\alpha$-$\gamma$ angular correlation. We obtain good agreement between
62: our theoretical predictions and experimental results.
63: \end{abstract}
64:
65: \pacs{23.60.+e, 03.65.Sq, 27.80.+w, 41.60.-m}
66:
67: \maketitle
68:
69: %
70: % Introduction
71: %
72: \section{Introduction}
73:
74: A characteristic feature of the $\alpha$ decay process is the
75: quantum mechanical tunneling~\cite{Ga1929}
76: through the so-called Coulomb wall generated by the electrostatic
77: interaction of the $\alpha$ particle with the constituent protons of the
78: daughter nucleus. Bremsstrahlung in $\alpha$ decay is intriguing because
79: of the classically incomprehensible character of radiation emission during
80: the tunneling process. Considerable attention has therefore been devoted
81: to both
82: experimental~\cite{DAEtAl1994brems,KaEtAl1997,Eremin2000,KaEtAl2000,KaEtAl1997jpg}
83: as well as theoretical
84: investigations~\cite{DyGo1996,PaBe1998,Dy1999,TaEtAl1999,%
85: Tk1999,Tk1999jetp,BPZ1999,MaOl2003,MaBe2004}, with the aim of elucidating
86: the role of tunneling during the emission process. It is
87: necessary to emphasize, however, that the term ``radiation during the
88: tunneling process'' has a restricted meaning as the wavelength of the
89: photon is much larger than the width of the tunneling region and even
90: larger than the main classical acceleration region. It is therefore not
91: possible to identify the region where the photon was
92: emitted. Besides, it is possible to write the matrix element of
93: bremsstrahlung in different forms using operator identities. As a result,
94: the integrands for the matrix element, as well as the relative
95: contributions of the regions of integration, will be different depending
96: on the operator identities used, although the total answer remains, of
97: course, the same. This was demonstrated, e.g., by Tkalya
98: in Refs.~\cite{Tk1999,Tk1999jetp}.
99:
100:
101: In the present paper, we revisit the theory of bremsstrahlung in the
102: $\alpha$ decay of a nucleus with a special emphasis on the
103: quasiclassical approximation. The applicability of this
104: approximation is ensured by the large value of the Sommerfeld
105: parameter $\eta$ (see below). We investigate the range of validity
106: of the result obtained by \dy and show that it is restricted by the
107: condition $x \ll 1/\sqrt{\eta}$ where $x=E_\gamma/Q_\alpha$
108: (here, $E_\gamma$ is the photon energy, and $Q_\alpha$ is the $\alpha$-decay
109: $Q$-value). Our quasiclassical result has no such a restriction
110: although we assume $x\ll 1$. It is consistent with the results
111: of both \dy and \pb
112: in limiting cases. For the experimentally interesting case
113: of the $\alpha$ decay of ${}^{210}{\rm Po}$, our result is valid with
114: high accuracy up to $x \sim 0.1$.
115:
116: Another subject investigated here is the angular distribution of
117: emitted photons. The $\alpha$ particle, initially in an $S$ state,
118: may undergo a dipole transition to a $P$ final state, or a
119: quadrupole transition to a $D$ state. While the quadrupole
120: contribution is parametrically suppressed for small photon energies,
121: the effective charge prefactor for the quadrupole contribution is
122: large. The dipole-quadrupole interference term vanishes after
123: angular averaging, but gives a significant contribution to the
124: differential photon emission probability, resulting in a substantial
125: deviation from the usually assumed dipole emission characteristics.
126:
127: Very recently, the results of our high-statistics measurement of
128: bremsstrahlung emitted in the $\alpha$ decay of $^{210}$Po have
129: been published, see Ref.~\cite{ourexp}.
130: Due to the limited solid-angle coverage of the detectors used in
131: this experiment, it was necessary to account for the
132: $\alpha$-$\gamma$ angular correlation. Taking into account
133: the contributions of the dipole and
134: quadrupole amplitudes in the data analysis, as derived in the present paper,
135: good overall agreement between theory and experiment is observed.
136:
137: This paper is organized in four sections. In Sec.~\ref{dipole}, we
138: investigate the leading dipole contribution to the differential
139: bremsstrahlung probability and evaluate the corresponding
140: amplitude in the quasiclassical approximation. The quadrupole
141: contribution to the amplitude and its interference with the dipole
142: part is analyzed in Sec.~\ref{quadrupole}. Conclusions are drawn in
143: Sec.~\ref{conclu}. Two appendices provide details on the
144: methods used in the calculations.
145:
146: %
147: % DIPOLE EMISSION
148: %
149: \section{DIPOLE EMISSION}
150: \label{dipole}
151:
152: %
153: % Emission Probability
154: %
155: \subsection{Emission Probability}
156:
157: It was shown in Ref.~\cite{PaBe1998} that the differential
158: bremsstrahlung probability $dP/dE_\gamma$ as a function of the energy
159: $E_\gamma$ of the radiated photon in the dipole approximation has the form
160: %
161: \begin{eqnarray}
162: \label{basic}
163: \frac{dP}{dE_{\gamma}} =
164: \frac{4 \, e^2 Z_{\rm eff}^2}{3\,\mu^2\,E_{\gamma}} \,
165: \left|{\cal M}\right|^2\,,
166: \qquad
167: {\cal M}= \left< R_f \left| \partial_r V \right| R_i\right>\,,
168: \end{eqnarray}
169: %
170: where natural units with $\hbar=c=\epsilon_0=1$ are applied
171: throughout the paper, $e$ is the proton charge and $\mu$ is the
172: reduced mass of the combined system of $\alpha$ particle and
173: daughter nucleus, $V \equiv V(r) = z(Z-z) \alpha/r$ is the
174: potential of the daughter nucleus felt by the $\alpha$ particle. The
175: functions $R_i$ and $R_f$ are the radial wave functions of the
176: initial and final states corresponding to the angular momenta $l=0$
177: and $l=1$, respectively (see App.~\ref{radial}). The effective
178: charge for a dipole interaction between an $\alpha$
179: particle with charge number $z=2$ and mass number $4$ emitted from a
180: parent nucleus with charge number $Z$ and mass number $A$ is (see
181: also App.~\ref{effcharge})
182: %
183: \begin{equation}
184: \label{Z1eff}
185: Z_{\rm eff}=Z^{(1)}_{\rm eff} \approx \frac{2\,A - 4\,Z}{A} = \frac25\,,
186: \end{equation}
187: %
188: where the latter value is relevant for the experimentally
189: interesting case of the $\alpha$ decay of ${}^{210}{\rm Po}$ ($Z = 84$).
190: Evaluating the effective charge with accurate
191: values for the masses of the alpha particle and
192: the daughter nucleus ($^{206}$Pb, $Z = 82$), as given in
193: Ref.~\cite{Fi1996} yields a value of $Z^{(1)}_{\rm eff}= 0.399$.
194:
195: In the present paper, we calculate the matrix element ${\cal M}$ in
196: the quasiclassical approximation taking into account the first
197: correction of the order ${\cal O}(\eta^{-1})$, and the corresponding
198: result is given below in Eq.~(\ref{Our}). However, before presenting
199: and discussing our formula for the matrix element ${\cal M}$,
200: let us briefly review several results for ${\cal M}$ obtained earlier in the
201: quasiclassical approximation. These are illustrative with respect to
202: their range of applicability and with respect to the importance of the
203: tunneling contribution.
204:
205:
206: %
207: % Dipole Transition Matrix Element
208: %
209: \subsection{Dipole Transition Matrix Element}
210:
211: Various approximations have been applied for the evaluation of the matrix
212: element $\cal M$ in Eq.~(\ref{basic})
213: \cite{Dy1999,PaBe1998,TaEtAl1999,Tk1999,Tk1999jetp}.
214: The approximations are intertwined with the identification of particular
215: contributions to the real and imaginary parts of the matrix element
216: $\mathcal M$ due to ``tunneling'' and due to ``classical motion'' of the
217: $\alpha$ particle.
218:
219: We use the convention that the complex phase of the matrix element
220: $\cal M$ should be chosen in such a way that it becomes purely real
221: in the classical limit $E_\gamma \to 0$. Our definition of $\cal M$
222: is consistent with that used in Ref.~\cite{Dy1999} and differs by a
223: factor $\rm{i}$ from the definition used in Ref.~\cite{PaBe1998}.
224:
225: Equation~(5) in the work of \pb contains a fully quantum mechanical
226: result for the photon emission amplitude $\mathcal{M}$, expressed in
227: terms of regular and irregular Coulomb functions, without any
228: quasiclassical approximations. However, the physical interpretation
229: of this result depends on a comparison with a
230: quasiclassical approximation, as only such a comparison clearly
231: displays the importance of the finite photon energy and the emission
232: amplitude during tunneling. \pb therefore present and discuss
233: a quasiclassical
234: expression for the imaginary part of their matrix element (real part
235: for our convention), ignoring the contribution from the tunneling
236: process to the emission amplitude. Note that the quasiclassical
237: expression of \pb provides a very good approximation
238: for the imaginary part of their matrix element up to very large photon
239: energies with $x\sim 0.6$.
240:
241: In contrast, the quasiclassical result of \dy is valid only for very
242: small photon energies ($x\ll 1/\sqrt{\eta}$), but includes
243: contributions from tunneling.
244: Here, we unify the treatments of Refs.~\cite{PaBe1998,Dy1999} and obtain a
245: quasiclassical differential emission probability $dP/dE_\gamma$,
246: which includes the effect of photon emission during the tunneling
247: process and which is substantially more accurate for higher photon
248: energies than that of \dy.
249:
250: The quasiclassical approximation for the wave functions of the
251: system of an $\alpha$ particle and a daughter nucleus in the initial
252: and final states is valid for large values of the Sommerfeld
253: parameters $\eta_{i,f}=z(Z-z)e^2\mu/k_{i,f}$ with $k_i=\sqrt{2\mu
254: Q_\alpha}$ and $k_f=\sqrt{2\mu(Q_\alpha-E_\gamma)}$. The indices $i$
255: and $f$ are reserved for initial and final configurations throughout
256: the paper. The value of $\eta_i$ for the decay of $^{210}$Po, which
257: is the experimentally most interesting nucleus, is $22.0$,
258: while the $Q$-value is $Q_\alpha = 5.40746 \, {\rm MeV}$~\cite{Fi1996}.
259:
260: If one neglects the contribution to the matrix element $\mathcal M$
261: from the region $r\lesssim r_0$, where $r_0$ is the nuclear radius,
262: then one is consistent with the simple assumption for the potential as
263: a square well for $r < r_0$ and a pure repulsive Coulomb potential for
264: $r > r_0$. Implementing this procedure according to \pb, one obtains
265: the following approximation for the
266: real part $M$ of the matrix element of $\mathcal M$,
267: %
268: \begin{align}
269: \label{Pap_matix_el_exact}
270: M = \mbox{Re}\, \cal M \approx &
271: \, \eta_i\,\sqrt{\frac{2\,k_i}{\pi\, k_f}}\, \nonumber\\
272: & \; \times \int_{0}^{\infty}\frac{dr}{r^2} \,
273: F_1(\eta_f,k_f r)\, F_0(\eta_i,k_i r) \,.
274: \end{align}
275: %
276: Here, $F_0$ and $F_1$ are the regular Coulomb radial wave functions
277: corresponding to angular momenta $l=0$ and $l=1$, respectively.
278: The importance of the contribution of the region $r \lesssim r_0$ was
279: discussed in Refs.~\cite{PaBe1998,Dy1999}. For relatively small photon
280: energies, which are interesting from an experimental point of view,
281: this contribution is not significant, and we do not consider it in the
282: present paper. For high photon energies, the contribution of the
283: region $r \sim r_0$ can be important (see Ref.~\cite{BPZ1999}).
284: The explicit form of $M$ is given by Eqs.~(6) and (7) of
285: Ref.~\cite{PaBe1998,PaBe1998-misprint}.
286:
287: We have calculated the integral
288: in Eq.~(\ref{Pap_matix_el_exact})
289: using quasiclassical wave functions, keeping the correction of order
290: ${\cal O}(\bar{\eta}^{-1})$ and ignoring terms of order ${\cal
291: O}(1/\bar{\eta}^2)$ and higher in the expansion for large ``mean''
292: Sommerfeld parameter ${\bar \eta} = (\eta_i + \eta_f)/2$. The result
293: of such a quasiclassical calculation for the real part of $\cal M$
294: reads
295: %
296: \begin{align}
297: \label{Pap_matix_el_q}
298: \widetilde{M} =&
299: \sqrt{\frac{2\,k_i}{\pi\, k_f}}\,
300: \frac{k_i\,k_f}{k_i+k_f}\,
301: \frac{\eta_i}{\bar{\eta}} \nonumber\\
302: & \times
303: \xi \, {\rm e}^{-\pi \xi /2} \,
304: \left[-K_{{\rm i}\xi}'(\xi)-
305: \frac{1}{\bar{\eta}} \,
306: K_{{\rm i}\xi}(\xi)\right]\,,
307: \end{align}
308: %
309: where $\xi=\eta_f-\eta_i$, and
310: $K_{a}(b)$ is the modified Bessel function. The derivative is
311: $K_{a}'(b)= \frac{\partial}{\partial b}K_{a}(b)$. \pb also
312: calculated the real part of the integral (\ref{Pap_matix_el_exact})
313: in the quasiclassical approximation. Note, however, that
314: their term of order ${\cal O}(\bar{\eta}^{-1})$ [see Eq.\ (14) of
315: Ref.~\cite{PaBe1998}]
316: contains an additional factor $\sqrt{3}/2$ in comparison to
317: Eq.~(\ref{Pap_matix_el_q}).
318:
319: Figure~\ref{largex} shows the ratio ${\rm Re} \, {\cal M}/ \widetilde{M} =
320: M/\widetilde{M}$ as a function of $x$ for $\eta_i=22.0$,
321: corresponding to the case of $^{210}$Po. Note that even at $x=0.3$,
322: the deviation of the quasiclassical result with the correction
323: ${\cal O}(\bar{\eta}^{-1})$ taken into account is less than $1\%$
324: (solid line), while without this correction
325: ($\widetilde{M}^0$), the deviation is about $5\%$ (dashed line).
326:
327: \begin{figure}
328: \centering
329: \includegraphics[scale=0.7]{Fig1.eps}
330: \caption{\label{largex} The ratio of $M = \mbox{Re}\, \cal M$,
331: see Eq.~(\ref{Pap_matix_el_exact}), to various
332: approximations, as a function of $x=E_\gamma/Q_\alpha$ for
333: for the case of $^{210}$Po ($\eta_i =22.0$). The solid line shows
334: $M/\widetilde{M}$ with $\widetilde{M}$ taken from
335: Eq.~(\ref{Pap_matix_el_q}), illustrating the excellent agreement of
336: the quasiclassical matrix element with the exact result in the range
337: $x \leq 0.3$. For the dashed line we use
338: $\widetilde{M}^0$,
339: which is obtained from $\widetilde{M}$ by omitting the
340: correction of order ${\cal O}(\bar{\eta}^{-1})$, showing a deviation
341: of less than $5\%$. For the dash-dotted line the asymptotics
342: $\widetilde{M}^A$ as given in Eq.~(\ref{MPBA}) is used. The deviation of the
343: latter curve from the others illustrates that the
344: ``$x \to 0$''-asymptotics is indeed only valid for $x \ll 1/\sqrt{\eta}$.}
345: \end{figure}
346:
347: Strictly speaking, the validity of the evaluation of the transition
348: matrix element (\ref{basic}) with the wave functions taken in the
349: quasiclassical approximation requires special consideration (see \S
350: 51 of Ref.~\cite{LaLi1958}) because of possibly noticeable
351: contributions from the vicinities of the turning points. However,
352: one can show that, for $\xi\ll\bar\eta$ (or $x \ll 1$), the
353: contributions of the vicinities of the classical turning points
354: ($r_{ci} =2\eta_i/k_i$ and $r_{cf} =2\eta_f/k_f$ in our case) are
355: small. For $x\lesssim 1$, these contributions are no longer negligible.
356:
357: For $x\ll 1$, we have $\xi=\eta_i\,(x/2+3x^2/8+\ldots)\,$. If
358: $x^2\,\eta_i\ll 1$ (even if $x \, \eta_i \lesssim 1$), we can
359: replace in (\ref{Pap_matix_el_q}) $\xi$ by $x\eta_i/2$ and make the
360: substitution $\eta_f \to \eta_i$ and $k_f \to k_i$. As a result, we
361: obtain the following asymptotics of $\widetilde{M}$:
362: %
363: \begin{align}
364: \label{MPBA}
365: \widetilde{M}^A &=\frac{k_i}{\sqrt{2 \pi}} \,
366: \left(\frac{x \eta_i}{2}\right) \, {\rm e}^{-\pi x \eta_i/4} \,
367: \nonumber\\
368: & \; \times \left[-K_{{\rm i} x \eta_i/2}'\left(\frac{x
369: \eta_i}{2}\right) -\frac{1}{\eta_i} \, K_{{\rm i} x
370: \eta_i/2}\left(\frac{x \eta_i}{2}\right)\right]\,.
371: \end{align}
372: %
373: From the dash-dotted line of Fig.~\ref{largex}, we see that the
374: ratio $M/\widetilde{M}^A$ deviates
375: substantially from unity, illustrating that the applicability of
376: Eq.~(\ref{MPBA}) is indeed restricted to very small values of $x$.
377:
378: Since
379: %
380: \begin{equation}
381: K_{i\nu}(x) = \exp(-\pi\nu/2) \int_0^\infty \cos(x\sinh t - \nu t)\,dt\,,
382: \end{equation}
383: %
384: the quantity $\widetilde{M}^A$ without the ${\cal
385: O}(\eta_i^{-1})$ correction exactly coincides with the real part of
386: the amplitude ${\widetilde{\cal M}}_{\rm D}$ obtained by \dy,
387: where
388: %
389: \begin{subequations}
390: \label{MD}
391: \begin{align}
392: \label{MDM}
393: \widetilde{\cal M}_{\rm D} =& \;
394: \frac{k_i}{\sqrt{2\pi}} \, J_{\rm D}\left(\frac{x\,\eta_i}{2}\right)\,,\\
395: %
396: \label{MDJ}
397: J_{\rm D}(y) =& \; -{\rm i}\, y \exp(-\pi y)\nonumber\\
398: &\times\int_{0}^{\infty} dt \,\sinh(t) \, \exp[{\rm i} \, y \,
399: (\sinh t-t)]\,,
400: \end{align}
401: \end{subequations}
402: %
403: showing that this result is applicable only to very small photon
404: energies. $\widetilde{\cal M}_{\rm D}$
405: contains both a real and an imaginary part and thus takes
406: photon emission during tunneling into account. It was pointed out by
407: \dy that the imaginary part of $\widetilde{\cal M}_{\rm D}$ at $x\eta_i \sim
408: 1$ is of the same order as the real part. For instance, in the case
409: of the $\alpha$ decay of $^{210}$Po with $\eta_i=22.0$, we have
410: $ x \, \eta_i \sim 1$ for $x\approx 0.05$. This indicates
411: that the imaginary part of $\widetilde{\cal M}_{\rm D}$ is important.
412:
413: Our quasiclassical approximation $\widetilde{\cal M}$ for the dipole
414: transition matrix element $\cal M$ from Eq. (\ref{basic}),
415: which includes the contributions from the tunneling of the $\alpha$
416: particle (we would like to refer to this result as the ``unified
417: result'' in the following sections of the current paper) has the
418: form:
419: %
420: \begin{subequations}
421: \label{Our}
422: \begin{align}
423: \label{OurA}
424: \widetilde{\cal M} =& \sqrt{\frac{2\,k_i}{\pi\, k_f}}\,
425: \frac{k_i\,k_f}{k_i+k_f}\, \frac{\eta_i}{\bar{\eta}}\,\left[J(\xi) +
426: \frac 1{\bar{\eta}} J_1(\xi)\right]\,,\\
427: %
428: \label{OurJ}
429: J(y) =& \; {\rm i}\, y \exp(-\pi y)\nonumber\\
430: &\times\int_{0}^{\infty} dt \,\sinh(t) \, \exp[{\rm i} \, y \,
431: (t-\sinh t)]\,,\\
432: %
433: \label{OurB} J_1(y)= &- y \exp(-\pi y) \,\int_{0}^{\infty} dt
434: \,\exp[{\rm i} \, y \, (t - \sinh t)]\,.
435: \end{align}
436: \end{subequations}
437: %
438: The derivation of Eqs.~\ref{Our} (see App.~\ref{radial} for
439: more details) involves a shift of the integration region into the
440: complex plane, which is needed to take the tunneling region
441: into account in a quasi-classical treatment, as explained
442: in~\cite{DyGo1996,Dy1999}.
443: Note that $J(y)$ is the complex
444: conjugation of the function $J_{\rm D}(y)$ as defined in
445: Eq.~(\ref{MD}). Although this is irrelevant for the calculation of the
446: bremsstrahlung emission probability, it is important for the
447: dipole-quadrupole interference term discussed in
448: Sec.~\ref{quadrupole}. The region of applicability of Eq.~(\ref{Our})
449: is much wider than that of Eq.~(\ref{MD}), because at small $x$ there
450: is no additional restriction $x\ll 1/\sqrt{\eta}$ which may otherwise
451: constitute a strong limitation at large value of $\eta$, as it was
452: shown above. Moreover, Eq.~(\ref{Our}) contains a correction ${\cal
453: O}(\bar\eta^{-1})$, which is also essential.
454:
455: \begin{figure}
456: \centering
457: \includegraphics[scale=0.7]{Fig2.eps}
458: \caption{The ratio $\mbox{Im}\,\widetilde{\cal M}/
459: \mbox{Re}\,\widetilde{\cal M}$ for the case of
460: $^{210}$Po ($\eta_i = 22.0$). The graph illustrates that
461: the imaginary part of the matrix element $\widetilde{\cal M}$
462: is quite substantial even for moderate values of $x$.}
463: \label{ImRe}
464: \end{figure}
465:
466: \begin{figure}
467: \centering
468: \includegraphics[width=8cm]{Fig3.eps}
469: \caption{\label{fig3}
470: Differential branching ratio $dP/dE_\gamma$ for bremsstrahlung emission during
471: $\alpha$ decay of $^{210}$Po in units of inverse keV (top panel)
472: and relative to the fully quantum mechanical calculation (PB)
473: of \pb (bottom panel).
474: The thick solid curve corresponds to our quasiclassical result,
475: as given in Eq.~(\ref{Our}). The other results are based on the
476: quasiclassical treatment of Ref.~\cite{PaBe1998}
477: ($\widetilde{\rm PB}$), the semiclassical treatment of
478: Dyakonov~\cite{Dy1999}, and on Eq.~(\ref{Our})
479: of the present work but neglecting the correction terms of
480: order ${\cal O}(\bar \eta^{-1})$.}
481: \end{figure}
482:
483: Our unified result (\ref{Our}) can be
484: used with high accuracy up to $x\sim 0.3$.
485: The real part of $\widetilde{\cal M}$ is identical
486: to $\widetilde{M}$ given in Eq.~(\ref{Pap_matix_el_q}) and is
487: discussed already in detail above (see also Fig.~\ref{largex}).
488: Figure~\ref{ImRe} shows the ratio
489: $\mbox{Im}\,\widetilde{\cal M}/\mbox{Re}\,\widetilde{\cal M} =
490: \mbox{Im}\,\widetilde{\cal M}/ \widetilde{M}$
491: as a function of $x$ at $\eta_i=22.0$,
492: i.e.~for the case of $^{210}{\rm Po}$. One
493: can see that $\mbox{Im}\,\widetilde{\cal M}$ is not small in
494: comparison to $\mbox{Re}\,\widetilde{\cal M}$. Thus, the imaginary
495: part gives a noticeable
496: contribution to $dP/dE_\gamma$ even for small $x$, and should not be
497: neglected. This point was also emphasized in
498: Refs.~\cite{Dy1999,TaEtAl1999}.
499:
500: %
501: % Quantitative comparison of various quasiclassical results
502: %
503: \subsection{Quantitative comparison of various quasiclassical results}
504:
505: As a last step, we compare in Fig.\ \ref{fig3} the differential
506: bremsstrahlung probability $dP/dE_\gamma$
507: for the bremsstrahlung accompanying $\alpha$ decay of $^{210}$Po
508: obtained with the use of
509: our matrix element $\widetilde{\cal M}$ given in Eq.~(\ref{Our}),
510: the matrix element $\widetilde{M}$ [Eq.~(\ref{Pap_matix_el_q})],
511: corresponding to the quasiclassical approach of
512: Ref.~\cite{PaBe1998}, and
513: the matrix element $\widetilde{\cal M}_{\rm D}$ [Eq.~(\ref{MD})], to the
514: full quantum mechanical formula given by Eq.\ (5) in
515: \cite{PaBe1998}. A detailed comparison is shown in the bottom panel
516: of Fig.\ \ref{fig3} in the $\gamma$ energy range
517: $0<E_\gamma<600$~keV. While the result of \dy (dotted line) deviates
518: roughly by a factor two at $E_\gamma=600$~keV, our result
519: (thick solid line) agrees with the exact quantum mechanical
520: treatment within about 2~\% at this photon energy;
521: the inclusion of the
522: ${\cal O}(\bar \eta^{-1})$ correction is crucial in obtaining
523: this agreement, as is evident from the supplementary curve in the
524: bottom panel, where we omit the $J_1$ term from Eq.~(\ref{OurB}).
525: The quasiclassical approximation of \pb, obtained by neglecting the
526: imaginary part of the matrix element (dashed line), deviates by more
527: than 15~\% at $E_\gamma=600$~keV.
528: As expected, at low photon energies
529: all results agree with each other.
530:
531: \begin{figure}
532: \centering
533: \includegraphics[width=8cm]{Fig4.eps}
534: \caption{\label{fig4} Interference term $\chi(x)$,
535: defined in Eq.~(\ref{chi}), for the bremsstrahlung accompanying
536: $\alpha$ decay of $^{210}$Po.}
537: \end{figure}
538:
539:
540: %
541: % QUADRUPOLE EMISSION
542: %
543: \section{QUADRUPOLE EMISSION}
544: \label{quadrupole}
545:
546: We are now concerned with the quadrupole component of the
547: bremsstrahlung probability and the angular distribution of the
548: radiation due to interference with the dipole components. The
549: outgoing $\alpha$ particle defines an axis of symmetry.
550: We therefore we may use $d\Omega = 2\pi \sin\theta d\theta$
551: in order to describe the solid angle element of the
552: photon spanning an infinitesimal range of polar angles
553: $\theta$ with respect to the direction of the emitted
554: $\alpha$-particle. We assume that a
555: summation with respect to photon polarization is performed. Within
556: the dipole approximation, Eq.~(\ref{basic}) gives rise to an angular
557: distribution of the form
558: %
559: \begin{eqnarray}
560: \left. \frac{d^2P}{dE_{\gamma} d\Omega} \right|_{\rm dip} =
561: \frac{e^2 \left( Z^{(1)}_{\rm eff} \right)^2
562: \sin^2\theta}{\pi\,\mu^2\,E_{\gamma}} \,
563: \left|{\cal M}\right|^2\,,
564: \end{eqnarray}
565: %
566: where the index refers to the dipole approximation.
567: Including the quadrupole term (see App.~\ref{radial}), this
568: formula should be generalized to
569: %
570: \begin{align}
571: \label{genangle}
572: \frac{d^2P}{dE_{\gamma} d\Omega} =&\;
573: \frac{e^2 \sin^2\theta}{\pi\mu^2E_{\gamma}}
574: \left|Z_{\rm eff}^{(1)} \, {\rm e}^{{\rm i}\delta_1} \,
575: {\cal M} \, +
576: Z_{\rm eff}^{(2)} \,
577: {\rm e}^{{\rm i}\delta_2} \,
578: {\cal N} \, \cos\theta
579: \right|^2
580: \nonumber\\[2ex]
581: %
582: \,= & \; \left. \frac{d^2P}{dE_{\gamma} d\Omega} \right|_{\rm dip} \,
583: [1+\chi\,\cos\theta\,] +{\cal O}({\cal N}^2)\, ,\\
584: %
585: \intertext{with}
586: \label{chi} \chi=& \; 2 \, \frac{Z_{\rm eff}^{(2)}}{Z_{\rm eff}^{(1)}}\,
587: {\rm Re} \left( \frac{{\cal M} \, {\cal N}^*}{|{\cal M}|^2} \, {\rm
588: e}^{{\rm i} (\delta_1 - \delta_2)} \right) \,,
589: \end{align}
590: %
591: where $\delta_1$ and $\delta_2$ are the Coulomb phases corresponding
592: to the angular momenta $l=1$ and $l=2$, respectively. The effective
593: quadrupole charge $Z^{(2)}_{\rm eff}$ is approximately given by (see
594: App.~\ref{effcharge})
595: %
596: \begin{equation}
597: \label{Z2eff} Z_{\rm eff}^{(2)}
598: \approx 2 + \frac{16 (Z-A)}{A^2} = 1.954\,
599: \end{equation}
600: %
601: for the case of $^{210}{\rm Po}$, and is roughly five times larger
602: than the dipole effective charge given in Eq.~(\ref{Z1eff}).
603: Exact masses~\cite{Fi1996} lead to the same
604: result $Z_{\rm eff}^{(2)}= 1.954$ (up to the last decimal digit indicated).
605:
606: Calculating the quadrupole matrix element ${\cal
607: N}$ within the quasiclassical approximation and keeping the leading
608: in $1/\bar{\eta}$ term, we obtain
609: %
610: \begin{subequations}\label{OurQ}
611: \begin{align}
612: \widetilde{\cal N} =& \;- \sqrt{\frac{2\,k_i}{\pi\, k_f}}\,
613: \frac{k_i\,k_f}{k_i+k_f}\,
614: \frac{\eta_i}{\bar{\eta}}\,v\,J_1(\xi) \,.
615: \end{align}
616: \end{subequations}
617: %
618: We here neglect a parametrically suppressed
619: next-to-leading order correction to the
620: interference term, discussed in more detail in Appendix~\ref{radial},
621: and we introduce the notation $v = \sqrt{2 Q_\alpha/\mu}$, where $v$
622: approximately equals
623: the final velocity of the $\alpha$ particle for bremsstrahlung emission
624: with $x \ll 1$. For $^{210}{\rm Po}$, we have $v
625: \approx 0.05$. Note that for a large Sommerfeld parameter $\eta_f$ we
626: have
627: %
628: \begin{equation}\label{phases}
629: {\rm e}^{{\rm i}\,(\delta_2 - \delta_1)} \approx {\rm i} +
630: \frac{2}{\eta_f} \, .
631: \end{equation}
632: %
633: Because of this ``${\rm i}$'' in the right-hand side~of
634: Eq.~(\ref{phases}), the imaginary parts of the functions $J(\xi)$
635: and $J_1(\xi)$ become very important for the interference term. The
636: quantity $\chi(x)$ defined in Eq.~(\ref{chi}), which determines the relative
637: magnitude of the interference term, vanishes for $x \to 0$ in the leading
638: in $1/\bar\eta$ approximation as
639: %
640: \begin{equation}
641: \chi\approx \frac{\pi}{2} \,
642: \frac{Z_{\rm eff}^{(2)}}{Z_{\rm eff}^{(1)}} \, v \, \eta_i\,x\,.
643: \end{equation}
644: %
645: For ${}^{210}{\rm Po}$, this evaluates to $\chi\approx 8.64\, x$.
646: Therefore, the value of the coefficient $\chi(x)$ becomes
647: significant already at very small photon energies ($\chi\sim 0.1$ at
648: $E_\gamma\sim 0.06$~MeV). The coefficient $\chi(x)$,
649: calculated with our quasiclassical results
650: $\widetilde{\cal M}$ and $\widetilde{\cal N}$, is plotted in
651: Fig.~\ref{fig4} as a function of $x = E_\gamma/Q_\alpha$ for the
652: experimentally interesting case of ${}^{210}{\rm Po}$.
653:
654: When integrating Eq.~(\ref{genangle}) over the total solid
655: angle, the interference term drops out and the remaining
656: quadrupole contribution to the differential emission
657: probability is of the order ${\cal O}( \widetilde {\cal N}^2)$;
658: for the case of $^{210}$Po this contribution amounts to less than
659: 1.5\,\% of the leading dipole term for photon energies up to
660: 600 keV ($x \approx 0.1$).
661:
662:
663: %
664: % Summary
665: %
666: \section{Summary}
667: \label{conclu}
668:
669: In summary, using a quasiclassical approximation we have obtained an
670: expression for the dipole bremsstrahlung probability during $\alpha$
671: decay which is in agreement with the full quantum mechanical treatment
672: of \pb in a substantially wider region
673: of the variable $x=E_\gamma/Q_\alpha$ in comparison with all
674: previous quasiclassical results. Our amplitude given in
675: Eq.~(\ref{Our}) contains both real and imaginary parts and,
676: thus, includes the contribution from bremsstrahlung during the
677: tunneling process. Our results demonstrate that the latter contribution is
678: not negligible even for rather small $x$. As an illustration of
679: these statements, we have considered the experimentally important
680: case of $^{210}$Po.
681:
682: Furthermore, we find the quasiclassical expression for the
683: contribution of the interference term of the dipole and quadrupole
684: components to the double differential bremsstrahlung probability
685: (with respect to the energy and the solid angle of the photon). This
686: contribution turns out to be significant.
687: Because of obvious limitations to the solid angle that can be
688: covered by detectors in a realistic experiment,
689: the angular distribution needs to be considered in the analysis of the
690: experimental data, even though the quadrupole
691: term makes a negligible contribution to the bremsstrahlung probability
692: after integration over the entire solid angle. Using the expression for the
693: dipole and quadrupole amplitudes presented in this work, the data analysis
694: in our recent experiment was performed as described in Ref.~\cite{ourexp},
695: and good overall agreement of our theoretical and experimental results
696: was obtained (see Fig.~5 of Ref.~\cite{ourexp}). We note, however,
697: that a certain subtle question remains with respect to a
698: next-to-leading order, nuclear model-dependent
699: correction to the dipole-quadrupole interference
700: term, as discussed in Appendix~\ref{radial}. These questions leave room for
701: further interesting investigations in the context
702: of under-the-barrier emission of bremsstrahlung in $\alpha$ decay
703: in the future.
704:
705: \begin{acknowledgments}
706:
707: A.I.M. and I.S.T. gratefully acknowledge the School of Physics at
708: the University of New South Wales, and the Max-Planck-Institute for
709: Nuclear Physics, Heidelberg, for warm hospitality and support during
710: a visit. D.S.~acknowledges support by the
711: Weizmann Institute through the Joseph Meyerhoff program,
712: and U.D.J.~acknowledges support from the Deutsche
713: Forschungsgemeinschaft (Heisenberg program). The work was also
714: supported by RFBR Grant No. 05-02-16079. \end{acknowledgments}
715:
716: \appendix
717:
718: %
719: % Effective charges
720: %
721: \section{Effective charges}
722: \label{effcharge}
723:
724: The purpose of this appendix (see also \cite{Eisenberg}) is to
725: clarify how the effective charges in Eqs.~(\ref{Z1eff}) and
726: (\ref{Z2eff}) for the dipole and the quadrupole terms arise in the
727: interaction of a two-body system (charges $eZ_1$ and $eZ_2$, masses
728: $m_1$ and $m_2$, $e$ is the proton charge) with a photon of
729: polarization $\bm{\epsilon}$ and wave vector $\bm{q}$, as given by
730: the part of the Hamiltonian corresponding to emission of a photon,
731: %
732: \begin{align}
733: \label{HI} H_I =& - \frac{eZ_1}{m_1}\,\bm{\epsilon}^*\cdot\bm{p}_1
734: \,
735: \exp\left( -{\rm i} \, \bm{q}\cdot\bm{r}_1\right)\nonumber\\
736: & - \frac{eZ_2}{m_2}\, \bm{\epsilon}^*\cdot {\bm{p}}_2 \, \exp\left(
737: -{\rm i} \, \bm{q}\cdot\bm{r}_2\right)\, .
738: \end{align}
739: %
740: We define the total mass $M = m_1+m_2$, the reduced mass $\mu = m_1
741: m_2/M$ , the center-of-mass coordinate $\bm{R} = (m_1/M) \bm{r}_1 +
742: (m_2/M) \bm{r}_2$, and the relative coordinate $\bm{r} = \bm{r}_1 -
743: \bm{r}_2$. Let $\bm p$ and $\bm P$ be the momenta corresponding to
744: the coordinates $\bm r$ and $\bm R$, respectively. Then $\bm{p}_1
745: =\bm p+ (m_1/M)\bm P$ and $\bm{p}_2 = -\bm p+ (m_2/M)\bm P$. Writing
746: the Hamiltonian (\ref{HI}) in the center-of-mass frame ($\bm P=0$)
747: and performing its expansion over $|\bm q\cdot\bm r|\ll 1$ up to
748: the first term, we obtain
749: %
750: \begin{equation}
751: \label{HIRES} H_I = -e \frac{\bm{\epsilon}^*\cdot\bm{p}}{\mu}\,
752: \left[ Z^{(1)}_{\rm eff} -{\rm i} \, Z^{(2)}_{\rm eff} \,
753: \bm{q}\cdot\bm{r} \right] {\rm e}^{-{\rm i}\bm{q}\cdot\bm{R}}\,,
754: \end{equation}
755: %
756: where the overall phase factor ${\rm e}^{-{\rm i}\bm{q}\cdot\bm{R}}$
757: can be safely ignored.
758:
759: The effective charges are
760: %
761: \begin{align}
762: Z^{(1)}_{\rm eff}
763: =& \frac{Z_1 \, m_2 - Z_2 \, m_1}{m_1 + m_2} \,,
764: \\[2ex]
765: Z^{(2)}_{\rm eff} =&
766: \frac{Z_1 \, m_2^2 + Z_2 \, m_1^2}{(m_1 + m_2)^2} \,.
767: \end{align}
768:
769: %
770: % Matrix elements
771: %
772: \section{Matrix elements}
773: \label{radial}
774:
775: In this appendix we present some details of the
776: derivation of our quasiclassical dipole (\ref{Our}) and quadrupole
777: (\ref{OurQ}) matrix elements. The wave function of the final state
778: has the form (see, e.~g., \S 136 and \S 137 of~\cite{LaLi1958}):
779: %
780: \begin{align}
781: \psi_f(\bm{r})&\equiv \psi^{(-)}_{\bm{k}_f}(\bm{r})\nonumber\\
782: =&
783: \frac{1}{2k_f}
784: \sum_{l=0}^\infty {\rm i}^l \, {\rm e}^{-{\rm i}\delta_l}\,
785: (2 l + 1)\, R_{k_f,\,l}(r) \, P_l(\bm{n} \cdot
786: \bm{\lambda})\,,
787: \end{align}
788: %
789: where $\bm n=\bm r/r$, $\bm \lambda=\bm k_f/k_f$, $P_l$ are the
790: Legendre polynomials, and
791: %
792: \begin{equation}
793: \label{finalstate}
794: R_{k_f,\,l}(r)= \frac{2}{r}F_l(\eta_f, k_f r)
795: \end{equation}
796: %
797: is the regular radial solution of the
798: Schr\"{o}dinger equation in a Coulomb field.
799:
800: When comparing to Eq.~(5) of Ref.~\cite{PaBe1998},
801: it is evident that the ansatz (\ref{finalstate})
802: for the final-state wave function
803: corresponds to the neglect of the contribution from the
804: irregular solution in the Couloumb field, which
805: given by the term $2 r^{-1} G_l(\eta_f, k_f r) \, \tan\alpha$
806: in the integral in Eq.~(5) of Ref.~\cite{PaBe1998}.
807: The basis for our approximation (\ref{finalstate})
808: is as follows.
809: The asymptotics of the wave function at $kr \gg 1$ is
810: $2 r^{-1} \sin(k r + f_C + \delta_N)$, where $f_C$
811: corresponds to the asymptotics in a pure Coulomb field, and
812: the phase shift
813: $\delta_N$ is due to the nuclear potential,
814: which has a size of the order of $r_0$.
815: So, the asymptotic form of $R_{k_f, l}(r)$ is
816: proportional to
817: $\cos(\delta_N)F_l(\eta_f, k_f r) +
818: \sin(\delta_N)G_l(\eta_f, k_f r)$.
819:
820: It follows from the
821: quasiclassical approximation that the wave function at
822: $r \ll r_{cf} \equiv r_t$,
823: where $r_t=2 \eta_f/(\mu v)$ is the classical turning point,
824: is given by $F_l(\eta_f, k_f r) \propto
825: \exp[-2\sqrt{2\mu v\eta_f} \, (\sqrt{r_t}-\sqrt{r})]$. Therefore,
826: $\delta_N\propto \exp[-8 \eta_f] \ll 1$, and one is therefore
827: led to the tentative conclusion that the term with the $G$ function
828: should be entirely negligible.
829: However, one might still object that the function $G$ is
830: exponentially large at $r \ll r_t$, namely
831: $G_l(\eta_f, k_f r)\propto\exp[2\sqrt{2\mu v\eta_f}\,(\sqrt{r_t}-\sqrt{r})]$.
832: In order to convince ourselves that the contribution
833: from the $G$ functions is indeed small,
834: let us consider as an example the term $\sin(\delta_N)G_0(\eta_i, k_i r) \,
835: G_l(\eta_f, k_f r)$. This term is proportional to
836: $\exp[-4(\eta_f-\eta_i)]\exp(-4\sqrt{2 \mu v \bar\eta}\sqrt{r})$. Therefore,
837: the main contribution to the transition
838: matrix element from this term is given
839: by the region $r\sim 1/(32 \mu \eta v)=1/(32 \mu(zZ\alpha)) \ll r_0$.
840: This contribution is indeed small, and
841: we can safely use the approximation (\ref{finalstate})
842: for the final state of the $\alpha$ particle.
843:
844: The wave function of
845: the initial state is given by an $S$ state which consists of
846: an outgoing wave at $r\to\infty$,
847: %
848: \begin{equation}
849: \psi_i(\bm r) = \frac{R_0(r)}{\sqrt{4\pi}}=\frac{1}{\sqrt{4\pi}\,
850: r}\,[G_0(\eta_i,k_i r)+{\rm i}\,F_0(\eta_i,k_i r)]\, .
851: \end{equation}
852: %
853: Substituting the wave functions into the transition matrix element
854: and taking the integrals over the angular parts of the
855: $\alpha$ particle wave functions, we obtain
856: %
857: \begin{align}
858: &\left< f | H_I | i \right> = - \frac{\sqrt{4 \pi}\,e\,
859: \eta_i\,k_i}{2k_f\,\mu^2
860: \omega}\, \bm{\epsilon}\cdot\bm{\lambda} \nonumber\\
861: &\times\Bigl\{ Z^{(1)}_{\rm eff} e^{{\rm i}\delta_1}
862: \int\limits_0^\infty {\rm d}r R_{k_f,\,1}(r)
863: R_{0}(r)\nonumber\\
864: & + Z^{(2)}_{\rm eff} e^{{\rm i}\delta_2}\,
865: \cos\theta\,\int\limits_0^\infty {\rm d}r R_{k_f,\,2}(r)
866: \left[ r \omega + \frac{9}{\mu r} \right] R_{0}(r) \Bigr\}\, ,
867: \label{vogts}
868: \end{align}
869: %
870: where $\theta$ is the angle between the vectors $\bm q$ and $\bm
871: \lambda$, $\omega = |\bm{q}| = E_\gamma$. Defining ${\cal M}$ and
872: ${\cal N}$ as
873: %
874: \begin{align}
875: {\cal M} =& -{\rm i}\eta_i \sqrt{\frac{k_i}{2\pi k_f}}\int_0^\infty
876: dr R_{k_f,\,1}(r)R_{0}(r)\,,
877: \\
878: {\cal N} =& -{\rm i}\eta_i \sqrt{\frac{k_i}{2\pi k_f}}\,
879: \int_0^\infty dr R_{k_f,\,2}(r)\left(\frac{9}{\mu r} + r \omega
880: \right) R_{0}(r)\,,\label{MN}
881: \end{align}
882: %
883: and taking into account that the sum over the photon polarizations
884: gives $ \sum |\bm \epsilon \cdot \bm \lambda|^2=\sin^2\theta $, we
885: obtain the double-differential bremsstrahlung probability as
886: %
887: \begin{align}
888: \frac{d^2P}{dE_{\gamma} d\Omega} =&
889: \frac{e^2}{\pi\mu^2E_{\gamma}}\sin^2\theta\, |{\cal C}|^2 \,
890: ,\nonumber\\
891: %
892: {\cal C} =& Z^{(1)}_{\rm eff} e^{{\rm i}\delta_1} {\cal M} +
893: Z^{(2)}_{\rm eff} e^{{\rm i}\delta_2} {\cal N} \cos\theta \,,
894: \end{align}
895: %
896: in agreement with Eq.~(\ref{genangle}). Then we use the
897: quasiclassical approximation for the radial part of the wave
898: functions (see, e.g., \S~48, 49 of Ref.~\cite{LaLi1958}). The matrix
899: elements with the quasiclassical radial wave functions have been
900: calculated using methods described in detail in
901: Ref.~\cite{Alder56}. Although these methods are in principle well known,
902: we present here some details of the calculation.
903:
904: Let us consider first the contribution $I_0$ of the classically
905: allowed region to the matrix element
906: %
907: \begin{equation}
908: I_1=\int\limits_0^\infty {\rm d}r R_{k_f,\,1}(r)R_{0}(r)
909: \end{equation}
910: %
911: Using the standard quasiclassical wave function and
912: assuming $x\ll 1$, we obtain
913: %
914: \begin{eqnarray}
915: I_0 &\approx&
916: -\frac{1}{4i\bar\eta}\int\limits_0^\infty\frac{d\vartheta}{\cosh^2(\vartheta/2)}
917: \mbox{e}^{i\xi(\sinh\vartheta+\vartheta)}\nonumber\\
918: &&\left\{1+i\frac{b_1-b_0}{2\bar\eta}\tanh(\vartheta/2)\right\}\nonumber\\
919: &=&\frac{\xi}{2\bar\eta}\Bigg(\int\limits_0^\infty d\vartheta \sinh(\vartheta)
920: \,\mbox{e}^{i\xi(\sinh\vartheta+\vartheta)}\nonumber\\
921: &&-\frac{i}{\bar\eta}\int\limits_0^\infty d\vartheta
922: \,\mbox{e}^{i\xi(\sinh\vartheta+\vartheta)}\Bigg)\,,
923: \end{eqnarray}
924: %
925: where $b_l=(l+1/2)^2$ and the second expression is obtained after an
926: integration by parts.
927: According to \cite{Dy1999}, the classically forbidden
928: part can be incorporated by shifting the integration
929: contour for $\vartheta$ into the complex plane,
930: via the replacement $\vartheta\to\vartheta+i\pi$. As a result we arrive at
931: %
932: \begin{eqnarray}\label{M0f}
933: I_1 &=&-\frac{\xi}{2\bar\eta}\,
934: \mbox{e}^{-\pi\xi}\Bigg(\int\limits_0^\infty d\vartheta \sinh(\vartheta)
935: \,\mbox{e}^{i\xi(\sinh\vartheta-\vartheta)}\nonumber\\
936: &&+\frac{i}{\bar\eta}\int\limits_0^\infty d\vartheta
937: \,\mbox{e}^{i\xi(\sinh\vartheta-\vartheta)}\Bigg)\,.
938: \end{eqnarray}
939: %
940: Similarly, the leading in $1/\bar\eta$ contribution to the quadrupole
941: amplitude is determined by the integral [see the second term in
942: square brackets in Eq.~(\ref{MN})]
943: %
944: \begin{eqnarray}\label{M21}
945: I_{21}&=&\int\limits_0^\infty
946: {\rm d}r R_{k_f,\,2}(r)r \omega R_{0}(r)\nonumber\\
947: &=&\frac{iv\xi}{2\bar\eta}\mbox{e}^{-\pi\xi} \int\limits_0^\infty d\vartheta
948: \,\mbox{e}^{i\xi(\sinh\vartheta-\vartheta)}\,.
949: \end{eqnarray}
950: %
951: The results (\ref{M0f}) and (\ref{M21})
952: immediately verify Eqs.~(\ref{Our}) and (\ref{OurQ}).
953:
954: For the second
955: contribution to the quadrupole amplitude, see Eq.~(\ref{MN}),
956: %
957: \begin{eqnarray}
958: I_{22} &=&
959: \int\limits_0^\infty {\rm d}r R_{k_f,\,2}(r)\frac{9}{\mu r} R_{0}(r),
960: \label{M22}
961: \end{eqnarray}
962: %
963: we find that this term
964: is suppressed by a factor $1/\bar\eta$.
965: However, while $I_{21}$ vanishes in the limit of a small photon energy
966: $x\to 0$, the contribution $I_{22}$ tends to
967: a nonzero constant at $x=0$, i.e., even though $I_{22}$ is
968: parametrically suppressed by a factor $1/{\bar \eta}$,
969: it constitutes the dominant contribution to the
970: dipole-quadrupole interference term at very small photon
971: energies, due to its distinctive asymptotic behaviour.
972:
973: A precise calculation of the $I_{22}$ contribution to the
974: interference term is unfortunately hampered by the
975: fact that the region of integration around $r \sim r_0$
976: gives an important contribution to the value of $I_{22}$
977: due to the inverse power of $r$ in the integrand,
978: which thus makes the value of $I_{22}$
979: nuclear model-dependent. In the data analysis of our recent
980: experiment~\cite{ourexp},
981: we therefore did not include the parametrically suppressed term
982: $I_{22}$. While good overall agreement between experiment and
983: theory was obtained in this way (see Fig.~5 of Ref.~\cite{ourexp}),
984: the agreement is better at
985: photon energies $E_\gamma \geq 250\,{\rm keV}$ as compared to
986: photon energies below this region. As the region of small $E_\gamma$
987: coincides with the region where the $I_{22}$ term might contribute
988: to the dipole-quadrupole
989: interference term, its significance cannot be completely ruled
990: out at present.
991:
992: \begin{thebibliography}{10}
993:
994: \bibitem{Ga1929}
995: G. Gamow, Nature (London) {\bf 123}, 606 (1929).
996:
997: \bibitem{DAEtAl1994brems}
998: A. D'Arrigo, N. Eremin, G. Fazio, G. Giardina, M.~G. Glotova, T.~V. Klochko, M.
999: Sacchi, and A. Taccone, Phys. Lett. B {\bf 332}, 25 (1994).
1000:
1001: \bibitem{KaEtAl1997}
1002: J. Kasagi, H. Yamazaki, N. Kasajima, T. Ohtsuki, and H. Yuki, Phys. Rev. Lett.
1003: {\bf 79}, 371 (1997).
1004:
1005: \bibitem{Eremin2000} N.V. Eremin, G. Fazio, and G. Giardina, Phys. Rev. Lett. {\bf 85}, 3061 (2000).
1006:
1007: \bibitem{KaEtAl2000}
1008: J. Kasagi, H. Yamazaki, N. Kasajima, T. Ohtsuki, and H. Yuki, Phys. Rev. Lett.
1009: {\bf 85}, 3062 (2000).
1010:
1011: \bibitem{KaEtAl1997jpg}
1012: J. Kasagi, H. Yamazaki, N. Kasajima, T. Ohtsuki, and H. Yuki, J. Phys. G
1013: {\bf 23}, 1451 (1997).
1014:
1015: \bibitem{DyGo1996}
1016: M.~I. Dyakonov and I.~V. Gornyi, Phys. Rev. Lett. {\bf 76}, 3542 (1996).
1017:
1018: \bibitem{PaBe1998}
1019: T. Papenbrock and G.~F. Bertsch, Phys. Rev. Lett. {\bf 80}, 4141 (1998).
1020:
1021: \bibitem{Dy1999}
1022: M.~I. Dyakonov, Phys. Rev. C {\bf 60}, 037602 (1999).
1023:
1024: \bibitem{TaEtAl1999}
1025: N. Takigawa, Y. Nozawa, K. Hagino, A. Ono, and D. M. Brink,
1026: Phys. Rev. C {\bf 59}, R593 (1999).
1027:
1028: \bibitem{Tk1999}
1029: E.~V. Tkalya, Phys. Rev. C {\bf 60}, 054612 (1999).
1030:
1031: \bibitem{Tk1999jetp}
1032: E.~V. Tkalya, Zh. \'{E}ksp. Teor. Fiz. {\bf 116}, 390 (1999), [JETP {\bf 89},
1033: 208 (1999)].
1034:
1035: \bibitem{BPZ1999} C.~A. Bertulani, D. T. de~Paula, and
1036: V. G.~Zelevinsky, Phys. Rev. C {\bf 60}, 031602(R) (1999).
1037:
1038: \bibitem{MaOl2003}
1039: S.~P. Maydanyuk and V.~S. Olkhovsky, Prog. Theor. Phys. {\bf 109}, 203
1040: (2003).
1041:
1042: \bibitem{MaBe2004}
1043: S.~P. Maydanyuk and S.~V. Belchikov, Prob. At. Science
1044: and Technology {\bf 5}, 19 (2004).
1045:
1046: \bibitem{ourexp}
1047: H. Boie, H. Scheit, U. D. Jentschura, F. K\"{o}ck,
1048: M. Lauer, A. I. Milstein, I. S. Terekhov, and D. Schwalm,
1049: Phys. Rev. Lett. {\bf 99}, 022505 (2007).
1050:
1051: \bibitem{Fi1996}
1052: R.~B. Firestone, {\em Table of Isotopes} (J. Wiley \& Sons, New York, 1996).
1053:
1054: \bibitem{PaBe1998-misprint}
1055: In the notation of Eq.~(7) of \cite{PaBe1998}, $\eta\eta'/\xi^2$ should be
1056: replaced by $4\eta\eta'/\xi^2$ in the expression for $M_j$ (twice).
1057:
1058: \bibitem{LaLi1958}
1059: L.~D. Landau and E.~M. Lifshitz, {\em Quantum Mechanics {\em (Volume 3 of the
1060: Course of Theoretical Physics)}} (Pergamon Press, London, 1958).
1061:
1062: \bibitem{Eisenberg} J. M. Eisenberg and W. Greiner,
1063: {\em Nuclear Theory Vol. 2: Excitation mechanisms of the nucleus},
1064: (North-Holland, Amsterdam, 1988), 3$^{\rm rd}$ edition.
1065:
1066: \bibitem{Alder56} K. Alder, A. Bohr, T. Huus, B. Mottelson, and A. Winther,
1067: Rev. Mod. Phys., {\bf 28},432 (1956).
1068: \end{thebibliography}
1069:
1070: \end{document}
1071: