<<

RAPID COMMUNICATIONS

PHYSICAL REVIEW C, VOLUME 60, 031602

Bremsstrahlung by a tunneling particle: A time-dependent description

C. A. Bertulani,1,* D. T. de Paula,1,† and V. G. Zelevinsky2,‡ 1Instituto de Fı´sica, Universidade Federal do Rio de Janeiro, 21945-970 Rio de Janeiro, Rio de Janeiro, Brazil 2Department of Physics and and National Superconducting Laboratory, Michigan State University, East Lansing, Michigan 48824-1321 ͑Received 21 December 1998; published 23 August 1999͒ We study the bremsstrahlung radiation of a tunneling charged particle in a time-dependent picture. In particular, we treat the case of bremsstrahlung during ␣ decay and show deviations of the numerical results from the semiclassical estimates. A standard assumption of a preformed ␣ particle inside the well to sharp high-frequency lines in the bremsstrahlung emission. These lines correspond to ‘‘quantum beats’’ of the internal part of the wave function during tunneling arising from the interference of the neighboring resonances in the open well. ͓S0556-2813͑99͒50709-7͔

PACS number͑s͒: 23.60.ϩe, 03.65.Sq, 27.80.ϩw, 41.60.Ϫm

Recent experiments ͓1͔ have triggered great interest in the that a stable numerical solution of the Schro¨dinger equation, phenomena of bremsstrahlung during tunneling processes keeping track simultaneously of fast oscillations in the well which was discussed from different theoretical viewpoints in ͑‘‘escape attempts’’͒ and extremely slow tunneling, is virtu- Refs. ͓2–4͔. This can shed on basic and still controver- ally impossible in this approach: for 210Po it would require sial quantum-mechanical problems such as tunneling times about 1030 time steps in the iteration process. Instead, we ͓5͔, tunneling in a complex and nonstationary environment study the bremsstrahlung in high energy ␣ decays, for which ͓6͔, preformation of a tunneling cluster in a many-body sys- the decay time is treatable numerically. This allows us to pay tem ͓7͔, and so on. It seems that ␣ decay offers a unique attention to qualitatively new aspects of the bremsstrahlung possibility to study these fundamental questions. The discus- in decay processes, and compare with stationary approaches. sion of the bremsstrahlung radiation in ␣ decay, initiated by Although one could envisage other ways of solving the time- a semiclassical calculation of Dyakonov and Gornyi ͓2͔ and dependent Schro¨dinger equation which could possibly continued by a quantum-mechanical calculation in perturba- handle the real situation, this is beyond the scope of the tion theory of Papenbrock and Bertsch ͓3͔ and a more de- present article, and we refer the reader to other methods, e.g., tailed comparison of quantum-mechanical calculations with those discussed in Ref. ͓10͔. classical and semiclassical results by Takigawa et al. ͓4͔, For an ␣ particle being accelerated from the turning point shows a complicated interference pattern arising from the to infinity, the classical bremsstrahlung can be calculated contributions to bremsstrahlung from inner, under the bar- analytically. Using well-known equations, see, for example rier, and outer, in the classically allowed region, parts of the ͓11,12͔, and integrating along the outward branch of the Ru- wave function. Currently the experimental data are not con- therford trajectory in a head-on collision we get for the en- clusive, and more exclusive experiments, with better statis- ergy emitted by bremsstrahlung per frequency interval d␻, tics, are needed to give a clearer understanding of the phe- in the long- approximation, nomena. ͓ ͔ The theoretical approaches of Refs. 3,4 assume a stan- 8␲␻2 dard stationary description of quantum tunneling which is dE͑␻͒ϭ Z2 e2͉p ͑␻͉͒2, ͑1͒ 2 3 eff r very successful for ␣-decay lifetime and probabilities. How- 3m c ever, a time-dependent picture of a decay process ͓8,9͔ may be essential in understanding physics of the bremsstrahlung where and similar processes, in particular, in obtaining the appro- priate bremsstrahlung spectrum. This can be shown, for ex- ma Ϫ␲␯/2 ample, by looking into a case where there is no Coulomb p ͑␻͒ϭ e KЈ␯͑␯͒ ͑2͒ r 2␲ i acceleration after the tunneling. In this Rapid Communica- tion we model the time evolution of a wave function during tunneling using again ␣ decay as an example. We do not is the Fourier transform of the particle momentum, m ϭ ϩ ϭ attempt to compare our results with experimental data. The mN 4A/(A 4) is the reduced mass, and Zeff (2A Ϫ • ϩ ␣ ϩ reason is simple. The ␣ decay time, e.g., the lifetime of 4Z)/(A 4) is the effective charge of the particle the 210Po, is many orders of magnitude larger than typical times daughter nucleus (A,Z) in the dipole approximation. In Eq. ͑ ͒ ϭ ˆ ϭ 2 ␯ϭ ប for an ␣-particle to traverse across the nucleus. This implies 1 , pr p r, a Ze /E␣ , E␣a/ v0 , E␣ is the energy of ␣ • the particle, and v0 its asymptotic velocity. The functions Ki␯(x) are the modified Bessel functions of imaginary order, Ј *Electronic address: [email protected] and Ki␯(x) are their derivatives with respect to the argument. † Electronic address: [email protected] Dividing this equation by the energy E␥ϭប␻,weget ‡ Electronic address: [email protected] the differential probability per unit energy for the brems-

0556-2813/99/60͑3͒/031602͑4͒/$15.0060 031602-1 ©1999 The American Physical Society RAPID COMMUNICATIONS

C. A. BERTULANI, D. T. de PAULA, AND V. G. ZELEVINSKY PHYSICAL REVIEW C 60 031602

͑ FIG. 1. Average momentum ͑in MeV/c) for ␣ energies of 23.1 FIG. 2. Classical bremsstrahlung emission probability dotted ͒ ͑ ͒ MeV ͑a͒ and 20 MeV ͑b͒.In͑a͒ the dotted line is the classical lines , quantum-mechanical spectrum dashed lines using only the momentum for an ␣ particle running away from the closest ap- part of the wave function outside the barrier, and using the full ͑ ͒ proach distance. The dashed line is the average momentum calcu- space wave function solid lines . The upper part of the graph is for ϭ ϭ lated according to Eq. ͑4͒, but using only the part of the wave E␣ 23.1 MeV, while the lower part is for E␣ 20 MeV. function outside the barrier, i.e., the integral in Eq. ͑4͒ extending ϭ 2 1͑b͔͒. The ␣-energies are changed by keeping the number of from Rcl 2Ze /E␣ to infinity. The solid line represents the mo- mentum calculated according to Eq. ͑4͒, but including the full space nodes constant ͑in this case, seven nodes͒, and varying the range of the wave function. depth of the potential from 14 to 18 MeV, respectively. The barrier height is 27.6 MeV. The dotted line is the classical strahlung, i.e., dP/dE␥ϭ(1/E␥)dE(␻)/dE␥ . At low photon momentum for an ␣ particle running away from the closest ␯2 Ϫ␲␯͓ Ј ␯ ͔2 approach distance. The dashed line is the average momentum energies, we use the relation e Ki␯( ) ˜1 to show that calculated according to Eq. ͑4͒, but using only the part of the wave function outside the barrier, i.e., the integral in Eq. ͑4͒ dP 2 1 v 2 ϭ 2 ϭ 2 ␣ͩ 0 ͪ ͑ ͒ extending from Rcl 2 Ze /E␣ to infinity. In this case, the Zeff . 3 ͑ ͒ dE␥ 3␲ E␥ c wave function entering 4 is normalized within this space range. We solve the time-dependent Schro¨dinger equation for The resulting ‘‘quantum-mechanical’’ momentum at the the ␣ particle in a potential well ϩ Coulomb barrier, and later stage is closer to the classical one. However, the calculate the radial momentum from quantum-mechanical momentum increases slower than the classical one, due to the extended nature of the particle’s ץ ប u͑r,t͒ wave function. As a consequence, one expects that the Fou- p ͑t͒ϭ ͵ dru*͑r,t͒ , ͑4͒ -r rier transform of the quantum momentum has its higher freץ r i quencies suppressed, as compared to the classical case. The where u(r,t) is the radial part of the total wave function. A solid line represents the momentum calculated according to similar numerical study of the time evolution for the problem Eq. ͑4͒, but including the full space range of the wave func- of the ␣ decay was performed by Serot et al. ͓9͔. tion. One observes a wiggling pattern associated with the The time-dependent Schro¨dinger equation is solved, start- interference of neighboring quasistationary states of the par- ing with an s-wave wave function for an ␣ particle confined ticle inside the nucleus. The leaking of the inner part of the ϭϪ Ͻ ϭϪ in a radial well V(r) V0 for r R0, and V(r) V0 wave function creates an effective oscillating dipole that ϩ 2 у ϭ 2Ze /R0 for r R0.WesetR0 8.75 fm, realistic for the emits radiation. The same leaking creates the perturbation ␣ϩ210Po case ͓3͔.Attϭ0 we switch the potential to a mixing different resonance states. The Fourier transform of Ϫ Ͻ spherical square well with depth V0 for r R0 and a Cou- the particle momentum should therefore contain appreciable 2 у lomb potential 2Ze /r for r R0 which triggers the tunnel- amplitudes associated with this motion, as observed in Fig. ing. At a given time t the wave function is found by using the 2, solid line. Asymptotically, the momenta calculated in all standard method of inversion of a tridiagonal matrix at each different ways coincide at large times. time step increment ⌬t ͑see, e.g., Ref. ͓15͔͒. The space is In Fig. 2 the dotted lines show the classical bremsstrah- discretized in steps of ⌬xϭ0.05 fm, and the time step used is lung emission probability, the dashed lines show the ⌬tϭmc(⌬x)2/ប. In Fig. 1 we plot the average momentum ‘‘quantum-mechanical’’ momentum using only the part of for ␣ energies of 23.1 MeV ͓Fig. 1͑a͔͒ and 20 MeV ͓Fig. the wave function outside the barrier, and the solid lines are

031602-2 RAPID COMMUNICATIONS

BREMSSTRAHLUNG RADIATION BY A TUNNELING . . . PHYSICAL REVIEW C 60 031602 for the full space range of the wave function, as a function of the photon momentum. The upper part of the graph is for E␣ϭ23.1 MeV, while the lower part is for E␣ϭ20 MeV. Since the quantum-mechanical momentum of the particle in- creases slower than the classical one, the spectrum at larger photon energies is suppressed in comparison with the classi- cal one. Also, when one includes the whole wave function, the bremsstrahlung spectrum is even more suppressed at large photon energies. However, at very large photon ener- gies (E␥Ӎ8.5 MeV͒ the spectrum shows a peak, revealing the interference between the different components of the wave function in the well. The width ⌬E of the peak is related to the lifetime of the quasistationary state. For E␣ ϭ23 MeV we find ⌬Eϭ0.2 MeV while for E␣ϭ20 MeV the width is narrow, ⌬Eϭ0.02 MeV. These values agree with the conventional Gamow formula for the width of the qua- sistationary state of the ␣ particle with the same energy and potential parameters. In order to assess the relevance of the bremsstrahlung during tunneling and to shed light on the nature of the peaks ␣ in Fig. 2 we wil now consider an particle confined within FIG. 3. ͑a͒ Momentum ͑in MeV/c) of a particle in a square well ϭϪ Ͻ a three-dimensional square well with V V0 for r R0 , plus barrier as a function of time. ͑b͒ Corresponding bremsstrahlung ϭ Ͻ Ͻ ϭ V U0 for R0 r R1, and V 0, otherwise. We use the spectrum. ϭ Ϫ ϭ ϭ parameters R0 8.75 fm, R1 R0 1 fm, V0 14 MeV, U0 ϭ27.6 MeV, and E␣ϭ24.2 MeV. For this problem a much u0(r) as the initial state. We obtained a bremsstrahlung spec- simpler time-dependent solution can be found. The time- trum for the soft part of the spectrum which is by far smaller dependent wave function is obtained from the expansion than that displayed in Fig. 3͑b͒. Moreover, the spectrum de- cays much faster than ours. This can be understood as fol- u͑r,t͒ϭ ͵ dE a͑E͒eiEt/ប u ͑r͒, ͑5͒ lows. The soft part of the spectrum is due to the bremsstrah- E lung during tunneling. For a particle in a one-dimensional tunneling motion through a square barrier, Dyakonov and ͑ ͒ where uE(r) is the continuum radial wave function with Gornyi’s approach yields the spectrum given by ␲͐ ϭ␦ Ϫ Ј energy E, normalized to 4 druE*(r)uEЈ(r) (E E ) and dP 4 U0 ϭ Z2 ␣ exp͑Ϫ2k d͒ ␲ 2 eff 1 dE␥ 3 mc E␥ ͑ ͒ϭ ␲ ͵ ͑ ͒ ͑ ͒ ͑ ͒ a E 4 dru0 r uE r , 6 2 E␥d E␥d ϫexpͩ Ϫ2 ͪͫ1ϩexpͩ Ϫ ͪͬ , ͑7͒ បv បv where u0(r) is the radial wavefunction of the initial state. 1 1 For a square well plus barrier u (r), u (r), and a(E) are 0 E ϭͱ Ϫ given analytically. The time-dependent wave function is ob- where U0 is the barrier height, v1 2(U0 E␣)/m is the tained from Eq. ͑5͒ by a simple integration. imaginary velocity of the particle during tunneling, and d is In Fig. 3͑a͒ we plot the momentum of the particle for this the barrier width. system as a function of time. We observe a similar pattern as This formula shows that the semiclassical bremsstrahlung Ϫ1 in Fig. 1͑a͒, solid curve, but with stronger oscillations, due to spectrum of a tunneling particle varies as E␥ , for low pho- Ϫ ប the quantum beats. In Fig. 3͑b͒ we show the corresponding ton energies, and as ͓exp( 2E␥d/ v1)͔/E␥ , for high photon bremsstrahlung spectrum. The resulting pattern is very simi- energies. The slope parameter for high photon energies is ប ͑ ͒ lar to that displayed in Fig. 2. It is important to notice how- given by 2d/ v1 as displayed by the dashed line in Fig. 3 b . ever that there is no Coulomb acceleration for this system. However, the spectrum obtained from the dynamical calcu- The lower part of the energy spectrum is due solely to the lation, solid line in Fig. 3͑b͒, has a smaller slope parameter. tunneling through the barrier, while the peak at higher ener- Additionally, the spectrum shows pronounced peaks at large gies is again due to the interference of the resonances during photon energies. Figure 4, where we show the scattering the tunneling process. phase shift for the system ‘‘well ϩ barrier,’’ clarifies the We have compared the lower part of the spectrum in Fig. origin of these peaks. The resonances at 15.7, 24.2, and 35.4 3͑b͒ with the result of Dyakonov and Gornyi ͓2͔. They have MeV correspond to ͑bound or virtual͒ levels in the well. The obtained the bremsstrahlung spectrum for a tunneling charge amplitudes of Eq. ͑6͒ are presented in Fig. 4͑b͒ where the using perturbation theory and semiclassical wave functions resonance peaks are also evident. The peaks at 9.1 MeV and for the initial and final state of the particle. We have done a 11.1 MeV, shown in Fig. 3͑b͒, are due to interference of similar calculation, but using the bound-state wave function resonances at 15.7 MeV and 35.4 MeV with the initial state

031602-3 RAPID COMMUNICATIONS

C. A. BERTULANI, D. T. de PAULA, AND V. G. ZELEVINSKY PHYSICAL REVIEW C 60 031602

trum. The observation of those peaks would be valuable for inferring the content of the initial wave function of a pre- formed ␣ particle ͑or fission product͒. The coupling to the radiation field can also influence the tunneling process. Such effects require a fully quantum- mechanical approach which can be formulated in spirit of the classical work by Pauli and Fierz ͓13͔. It includes important feedback effects which renormalize the particle trajectory by the coupling to the accompanying radiation field ͓14͔. The exact solution of these equations is crucial in the case where the photon energy is comparable to the particle energy. For example, in the decay of very low energy ␣’s this coupling may reduce the yield of high energy . The reason is that when a particle emits a photon before tunneling, it looses energy and this leads to a reduction of its barrier tun- FIG. 4. ͑a͒ Phase shifts (lϭ0) for the ␣ϩ210Po system, assum- ing a radial square well size of 8.75 fm ϩ a square barrier of 1 fm neling probability. The exact results might be sensitive to the located at the border of the well. ͑b͒ Amplitudes for the overlap of shape of the barrier. We hope to come to this formulation of the initial state (Eϭ24.2 MeV͒ of the closed well ͑barrier of infi- the problem elsewhere. nitely large width͒ with neighboring states of the open well. In conclusion, we have obtained the bremsstrahlung spec- trum of a tunneling particle ͑an ␣ particle in a nucleus͒ by directly solving the time-dependent Schro¨dinger equation. of energy 24.2 MeV. The quantum beats apparently corre- As expected, we have found that there are large deviations spond to the energy differences (24.2Ϫ15.7ϭ8.5) MeV and from the classical bremsstrahlung spectrum. We have also (35.4Ϫ24.2ϭ11.2) MeV, respectively. These values are demonstrated that aproaches based on perturbation theory close to the energies of the peaks appearing in Fig. 3͑b͒. miss an important piece of information, namely, the time- The bremsstrahlung peaks associated with quantum beats dependent modification of the particle wave function in the are present in any dynamical tunneling process, since an ini- well during the decay time. This leads to substantial emis- tial localized state always has some overlap amplitude with sion of photons with frequencies close to those of quantum neighboring states of the open well. The importance of these beats between neighboring resonances. This effect should be peaks, or, equivalently, of the admixture of neighboring reso- relevant in radiation emitted during ␣ decay in nuclei. In a nances, decreases if the initial state is a very sharp resonance more general case, the time dependence of the wave function as in the case of the ␣ decay of 210Po studied in ͓1͔. Until of a tunneling particle seems to deviate substantially from now it was poorly understood how an ␣ particle was pre- the spectrum calculated by using perturbation theory with formed inside a nucleus. The possible manifestations of vir- semiclassical wave functions. More experimental data on tually excited clusters in nuclei are predicted to be seen in bremsstrahlung radiation by a tunneling particle would be knockout reactions induced by protons or ͓7͔. very welcome in learning more about preformation states Studying bremsstrahlung radiation in the tunneling process and dynamics of quantum beats. can provide additional information. Indeed, the initial wave function must be of a localized nature, thus having a nonzero This work was supported in part by the Brazilian funding amplitude for carrying a part of the wave function of an agencies CNPq, FAPERJ, FUJB/UFRJ, and PRONEX, under adjacent resonance. For high-lying states, as shown above, Contract No. 41.96.0886.00, and by NSF Grant No. 96- this leads to pronounced peaks in the bremsstrahlung spec- 05207.

͓1͔ J. Kasagi et al., Phys. Rev. Lett. 79, 371 ͑1997͒. ͓8͔ M.G. Fuda, Am. J. Phys. 52, 838 ͑1984͒. ͓2͔ M.I. Dyakonov and I.V. Gornyi, Phys. Rev. Lett. 76, 3542 ͓9͔ O. Serot, N. Carjan, and D. Strottman, Nucl. Phys. A569, 562 ͑1996͒. ͑1994͒. ͓3͔ T. Papenbrock and G.F. Bertsch, Phys. Rev. Lett. 80, 4141 ͓10͔ S.A. Gurvwitz, Phys. Rev. A 38, 1747 ͑1988͒. ͑1998͒. ͓11͔ J.D. Jackson, Classical Electrodynamics, 2nd ed. ͑Wiley, New ͓4͔ N. Takigawa et al., Phys. Rev. C 59, 593 ͑1999͒. York, 1975͒. ͓ ͔ ͓ ͔ 5 E.H. Hauge and J.A. Sto”vneng, Rev. Mod. Phys. 61, 917 12 L.D. Landau and E.M. Lifshitz, The Classical Theory of Fields ͑1989͒. ͑Pergamon, New York, 1975͒. ͓6͔ A.O. Caldeira and A.J. Leggett, Ann. Phys. ͑N.Y.͒ 149, 374 ͓13͔ W. Pauli and M. Fierz, Nuovo Cimento C 15, 167 ͑1938͒. ͑1983͒. ͓14͔ V.V. Flambaum and V.G. Zelevinsky, Report No. ͓7͔ A.A. Sakharuk and V.G. Zelevinsky, Phys. Rev. C 55, 302 nucl-th/9812076, RMSUCL-1121, 1998. ͑1997͒; A. Sakharuk, V. Zelevinsky and V.G. Neudatchin, ͓15͔ S. Koonin and D.C. Meredith, Computational Physics ibid. 60, 014605 ͑1999͒. ͑Addison-Wesley, Reading, MA, 1990͒.

031602-4