This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

M1M1 resonance in 208Pb within the self-consistent phonon-coupling model

V. Tselyaev [email protected]    N. Lyutorovich St. Petersburg State University, St. Petersburg, 199034, Russia    J. Speth Institut für Kernphysik, Forschungszentrum Jülich, D-52425 Jülich, Germany    P.-G. Reinhard Institut für Theoretische Physik II, Universität Erlangen-Nürnberg, D-91058 Erlangen, Germany
Abstract

The main goal of the paper is to investigate theoretically the experimentally observed fragmentation of the isovector M1M1 resonance in 208Pb within a self-consistent model based on an energy-density functional (EDF) of the Skyrme type. This fragmentation (spread of the M1M1 strength) is not reproduced in a conventional one-particle–one-hole (1p1h1p1h) random-phase approximation (RPA) and thus has to be investigated in the framework of more complicated models. However, previously applied models of this type were not self-consistent. In the present work, we use a recently developed renormalized version of the self-consistent time blocking approximation (RenTBA) in which the 1p1h1p1h\otimesphonon configurations are included on top of the RPA 1p1h1p1h configurations. We have determined several sets of the parameters of the modified Skyrme EDF fitted within the RenTBA and RPA and have found the necessary condition of producing the fragmentation of the M1M1 resonance in 208Pb in our model. We present also the results of the RenTBA and RPA calculations for the first excited states of the natural parity modes in 208Pb obtained with these modified parametrizations.

I Introduction

Magnetic dipole (M1M1) excitations in the 208Pb nucleus are the object of numerous experimental and theoretical investigations for several decades. From the theoretical point of view, one of the reasons is the possibility of determining the spin-related parameters of the residual interaction in the calculations of these excitations within the random-phase approximation (RPA) or its extended versions. The calculated energies of the unnatural parity excitations are very sensitive to the values of the underlying model parameters, in particular, of the parameters of the Skyrme energy-density functional. The comparison of these energies with experimental data is the only reliable method to estimate the spin-related parameters of the model. Another reason is the fragmentation (spread) of the isovector M1M1 resonance in 208Pb which is observed in the experiment (see Laszewski et al. (1988)) but which is absent in RPA where the isovector M1M1 strength in this nucleus is concentrated in one state. The description of this fragmentation requires application of more complicated models going beyond the RPA framework (see, e.g., Ref. Kamerdzhiev et al. (2004) for more details).

Most of the early calculations of the M1M1 excitations in 208Pb (see, e.g., Refs. Vergados (1971); Ring and Speth (1973); Tkachev et al. (1976); Speth et al. (1980); Borzov et al. (1984)) were performed within the RPA, the Tamm-Dancoff approximation or within the Migdal’s Theory of Finite Fermi Systems (TFFS, Ref. Migdal (1967)) which in its simplest form used in the applications is equivalent to the RPA with the zero-range residual interaction. In the following, the M1M1 modes were investigated within the generalized models in which the one-particle–one-hole (1p1h1p1h) RPA configuration space is enlarged by adding 2p2h2p2h, 1p1h1p1h\otimesphonon or two-phonons configurations (see, e.g., Lee and Pittel (1975); Dehesa et al. (1977); Kamerdzhiev and Tkachev (1984); Cha et al. (1984); Khoa et al. (1986); Kamerdzhiev and Tkachev (1989); Tselyaev (1989); Kamerdzhiev and Tselyaev (1991); Kamerdzhiev et al. (1993)). However, the fully self-consistent calculations of the M1M1 excitations in 208Pb have been performed so far only within the RPA (see Refs. Cao et al. (2009); Vesely et al. (2009); Nesterenko et al. (2010); Cao et al. (2011); Wen et al. (2014); Tselyaev et al. (2019)).

In a broad sense, self-consistency means the use of the same energy-density functional (EDF) E[ρ]E[\rho] (where ρ\rho is the single-particle density matrix) for the mean field as well as for the RPA residual interaction. This decreases the number of the free parameters of the theory and, in principle, increases its predictive power. Here we use an EDF of Skyrme type Bender et al. (2003). In a recent paper Tselyaev et al. (2019), we have shown that the adequate description of the low-energy M1M1 excitations in 208Pb within the self-consistent RPA based on the Skyrme EDF is possible only if the spin-related parameters of the known EDF are modified. By re-tuning these parameters we managed to reproduce within the RPA the experimental key quantities: energy and the strength of the 11+1_{1}^{+} state as well as the mean energy and the summed strength of the M1M1 resonance in 208Pb in the interval 6.6-8.1 MeV. However, as mentioned above, the observed fragmentation of the isovector M1M1 resonance and its total width in this model are not yet reproduced.

The aim of the present paper is to study the possibility to describe this fragmentation within the extended self-consistent model including the 1p1h1p1h\otimesphonon configurations on top of the RPA 1p1h1p1h configurations. This extended model is treated within the time blocking approximation (TBA) which we use actually in its renormalized version (RenTBA, Tselyaev et al. (2018)). Full self-consistency is maintained also for the extended treatment. The method of re-tuning the spin-related parameters of the Skyrme EDF developed in Ref. Tselyaev et al. (2019) is used also for the RenTBA.

The paper is organized as follows. In Section II the formalism of RPA and RenTBA is briefly described. Section III contains the numerical details and the calculation scheme. The main results of the paper are presented in Section IV. In Section V the fine structure of the M1M1 strength distributions in 208Pb and the impact of the single-particle continuum on this structure are analyzed. In Section VI the problem of the fragmentation of the isovector M1M1 resonance in 208Pb is discussed in detail and the necessary condition of the description of this fragmentation is determined. In Section VII we present the results of the RenTBA and RPA calculations of the low-energy electric excitations in 208Pb obtained with the use of the modified parametrizations of the Skyrme EDF. The conclusions are given in the last section.

II The model

Let us start with the RPA eigenvalue equation

34Ω12,34RPAZ34n=ωnZ12n,\sum_{34}\Omega^{\mbox{\scriptsize RPA}}_{12,34}\,Z^{n}_{34}=\omega^{\vphantom{*}}_{n}\,Z^{n}_{12}\,, (1)

where ωn\omega^{\vphantom{*}}_{n} is the excitation energy, Z12nZ^{n}_{12} is the transition amplitude, and the numerical indices (1,2,3,1,2,3,\ldots) stand for the sets of the quantum numbers of some single-particle basis. In what follows the indices pp and hh are used to label the states of the particles and holes in the basis which diagonalizes the single-particle density matrix ρ\rho and the single-particle Hamiltonian hh in the ground state [see Eq. (5) below]. The transition amplitudes are normalized by the condition

Zn|MRPA|Zn=sgn(ωn),\langle\,Z^{n}\,|\,M^{{}_{\mbox{\scriptsize RPA}}}_{\vphantom{1}}|\,Z^{n}\rangle=\mbox{sgn}(\omega^{\vphantom{*}}_{n})\,, (2)

where

M12,34RPA=δ13ρ42ρ13δ42M^{\mbox{\scriptsize RPA}}_{12,34}=\delta^{\vphantom{*}}_{13}\,\rho^{\vphantom{*}}_{42}-\rho^{\vphantom{*}}_{13}\,\delta^{\vphantom{*}}_{42} (3)

is the metric matrix in the RPA.

In the self-consistent RPA based on the EDF E[ρ]E[\rho] the RPA matrix ΩRPA\Omega^{\mbox{\scriptsize RPA}} is defined by

Ω12,34RPA=h13δ42δ13h42+56M12,56RPAV56,34,\Omega^{\mbox{\scriptsize RPA}}_{12,34}=h^{\vphantom{*}}_{13}\,\delta^{\vphantom{*}}_{42}-\delta^{\vphantom{*}}_{13}\,h^{\vphantom{*}}_{42}+\sum_{56}M^{\mbox{\scriptsize RPA}}_{12,56}\,{V}^{\vphantom{*}}_{56,34}\,, (4)

where the single-particle Hamiltonian hh and the amplitude of the residual interaction VV are linked by the relations

h12=δE[ρ]δρ21,V12,34=δ2E[ρ]δρ21δρ34.h^{\vphantom{*}}_{12}=\frac{\delta E[\rho]}{\delta\rho^{\vphantom{*}}_{21}}\,,\qquad{V}^{\vphantom{*}}_{12,34}=\frac{\delta^{2}E[\rho]}{\delta\rho^{\vphantom{*}}_{21}\,\delta\rho^{\vphantom{*}}_{34}}\,. (5)

In the TBA, the counterpart of Eq. (1) has the form

34Ω12,34TBA(ων)z34ν=ωνz12ν,\sum_{34}\Omega^{\mbox{\scriptsize TBA}}_{12,34}(\omega^{\vphantom{*}}_{\nu})\,z^{\nu}_{34}=\omega^{\vphantom{*}}_{\nu}\,z^{\nu}_{12}\,, (6)

where

Ω12,34TBA(ω)\displaystyle\Omega^{\mbox{\scriptsize TBA}}_{12,34}(\omega) =\displaystyle= Ω12,34RPA+56M12,56RPAW¯56,34(ω),\displaystyle\Omega^{\mbox{\scriptsize RPA}}_{12,34}+\sum_{56}M^{\mbox{\scriptsize RPA}}_{12,56}\,\bar{W}^{\vphantom{*}}_{56,34}(\omega)\,, (7a)
W¯12,34(ω)\displaystyle\bar{W}^{\vphantom{*}}_{12,34}(\omega) =\displaystyle= W12,34(ω)W12,34(0).\displaystyle{W}^{\vphantom{*}}_{12,34}(\omega)-{W}^{\vphantom{*}}_{12,34}(0)\,. (7b)

The matrix ΩTBA(ω)\Omega^{\mbox{\scriptsize TBA}}(\omega) is energy-dependent due to the matrix W(ω)W(\omega) which represents the induced interaction generated by the intermediate 1p1h1p1h\otimesphonon configurations. The subtraction of W(0)W(0) in Eq. (7b) serves to avoid changing the mean-field ground state Toepffer and Reinhard (1988); Gütter et al. (1993) and to ensure stability of solutions of the TBA eigenvalue equation (see Tselyaev (2013)). The matrix W(ω)W(\omega) is defined by the equations

W12,34(ω)\displaystyle{W}^{\vphantom{*}}_{12,34}(\omega) =\displaystyle= c,σσF12c(σ)F34c(σ)ωσΩc,\displaystyle\sum_{c,\;\sigma}\,\frac{\sigma\,{F}^{c(\sigma)}_{12}{F}^{c(\sigma)*}_{34}}{\omega-\sigma\,\Omega^{\vphantom{*}}_{c}}\,, (8a)
Ωc\displaystyle\Omega^{\vphantom{*}}_{c} =\displaystyle= εpεh+ων,ων>0,\displaystyle\varepsilon^{\vphantom{*}}_{p^{\prime}}-\varepsilon^{\vphantom{*}}_{h^{\prime}}+\omega^{\vphantom{*}}_{\nu}\,,\quad\omega^{\vphantom{*}}_{\nu}>0\,, (8b)

where σ=±1\sigma=\pm 1, c={p,h,ν}\,c=\{p^{\prime},h^{\prime},\nu\} is a combined index for the 1p1h1p1h\otimesphonon configurations, ν\nu is the phonon’s index, εp\varepsilon^{\vphantom{*}}_{p^{\prime}} and εh\varepsilon^{\vphantom{*}}_{h^{\prime}} are the particle’s and hole’s energies, ων\omega^{\vphantom{*}}_{\nu} is the phonon’s energy. The amplitudes F12c(σ){F}^{c(\sigma)}_{12} have only particle-hole matrix elements Fphc(σ){F}^{c(\sigma)}_{ph} and Fhpc(σ){F}^{c(\sigma)}_{hp}. They are defined by the equations

F12c()=F21c(+),Fphc()=Fhpc(+)=0,{F}^{c(-)}_{12}={F}^{c(+)*}_{21},\qquad{F}^{c(-)}_{ph}={F}^{c(+)}_{hp}=0\,, (9a)
Fphc(+)=δppghhνδhhgppν,{F}^{c(+)}_{ph}=\delta^{\vphantom{*}}_{pp^{\prime}}\,g^{\nu}_{h^{\prime}h}\!-\!\delta^{\vphantom{*}}_{h^{\prime}h}\,g^{\nu}_{pp^{\prime}}\,, (9b)

where g12νg^{\nu}_{12} is an amplitude of the quasiparticle-phonon interaction.

In the conventional TBA, the phonon’s energies ων\omega^{\vphantom{*}}_{\nu} in Eq. (8b) and the amplitudes g12νg^{\nu}_{12} in Eq. (9b) are determined within the framework of the RPA. In the non-linear version of the TBA developed in Ref. Tselyaev et al. (2018), the phonon’s energies ων\omega^{\vphantom{*}}_{\nu} are the solutions of the TBA equation (6), while the amplitudes g12νg^{\nu}_{12} are expressed through the transition amplitudes z12ν{z}^{\nu}_{12} which are also the solutions of Eq. (6), namely

g12ν=34V12,34z34ν.g^{\nu}_{12}=\sum_{34}{V}^{\vphantom{*}}_{12,34}\,{z}^{\nu}_{34}\,. (10)

The normalization condition for the transition amplitudes z12ν{z}^{\nu}_{12} has the form

(zν)RPA2+(zν)CC2=1,(z^{\nu})^{2}_{\mbox{\scriptsize RPA}}+(z^{\nu})^{2}_{\mbox{\scriptsize CC}}=1\,, (11)

where

(zν)RPA2\displaystyle(z^{\nu})^{2}_{\mbox{\scriptsize RPA}} =\displaystyle= sgn(ων)zν|MRPA|zν,\displaystyle\mbox{sgn}(\omega^{\vphantom{*}}_{\nu})\,\langle\,{z}^{\nu}\,|\,M^{{}_{\mbox{\scriptsize RPA}}}_{\vphantom{1}}|\,{z}^{\nu}\rangle\,, (12a)
(zν)CC2\displaystyle(z^{\nu})^{2}_{\mbox{\scriptsize CC}} =\displaystyle= sgn(ων)zν|Wν|zν,\displaystyle-\mbox{sgn}(\omega^{\vphantom{*}}_{\nu})\,\langle\,{z}^{\nu}\,|\,W^{\prime}_{\nu}\,|\,{z}^{\nu}\rangle\,, (12b)
Wν\displaystyle W^{\prime}_{\nu} =\displaystyle= (dW(ω)dω)ω=ων.\displaystyle\biggl{(}\frac{d\,W(\omega)}{d\,\omega}\biggr{)}_{\omega\,=\,\omega_{\nu}}. (12c)

The terms (zν)RPA2(z^{\nu})^{2}_{\mbox{\scriptsize RPA}} and (zν)CC2(z^{\nu})^{2}_{\mbox{\scriptsize CC}} represent the contributions of the 1p1h1p1h components (RPA) and of the complex configurations (CC) to the norm (11). The model includes only those TBA phonons that satisfy the condition

(zν)RPA2>(zν)CC2,(z^{\nu})^{2}_{\mbox{\scriptsize RPA}}>(z^{\nu})^{2}_{\mbox{\scriptsize CC}}\,, (13)

which together with Eq. (11) means that

(zν)RPA2>12.(z^{\nu})^{2}_{\mbox{\scriptsize RPA}}>\frac{1}{2}\,. (14)

The condition (13) confines the phonon space to the RPA-like phonons in agreement with the basic model approximations.

The feedback described above renders the phonon space of RenTBA fully self-consistent. In the present paper we use the version of this non-linear model in which the energies ων\omega^{\vphantom{*}}_{\nu} and the amplitudes z12ν{z}^{\nu}_{12} entering Eqs. (8b) and (10) (and only in these equations) are determined from the solutions of the TBA equation (6) in the diagonal approximation. This model is what we call the renormalized TBA (RenTBA, see Tselyaev et al. (2018) for more details).

III Numerical details and the calculation scheme

The equations of RPA and RenTBA were solved within the fully self-consistent scheme as described in Refs. Lyutorovich et al. (2015, 2016); Tselyaev et al. (2016). Wave functions and fields were represented on a spherical grid in coordinate space. The single-particle basis was discretized by imposing the box boundary condition with the box radius equal to 18 fm. The particle’s energies εp\varepsilon^{\vphantom{*}}_{p} were limited by the maximum value εpmax=100\varepsilon^{{}_{\mbox{\scriptsize max}}}_{p}=100 MeV. The non-linear RenTBA equations were solved by means of the iterative procedure. The phonon space of the first iteration included the RPA phonons with the energies ωn\omega^{\vphantom{*}}_{n}\leqslant 50 MeV and multipolarities LL\leqslant 15 of both the electric and magnetic types which have been selected according to the criterion of collectivity

Zn|V2|Zn/ωn20.05,\langle\,Z^{n}\,|\,V^{2}|\,Z^{n}\rangle/\omega^{2}_{n}\geqslant 0.05\,, (15)

see Tselyaev et al. (2018).

The field operator 𝑸Q in the case of the M1M1 excitations was taken in the form

𝑸Q =\displaystyle= μN316π{(γn+γp)𝝈+𝒍\displaystyle\mu^{\vphantom{*}}_{N}\,\sqrt{\frac{3}{16\pi}}\,\Bigl{\{}(\gamma_{n}+\gamma_{p}\,)\,\mbox{\boldmath$\sigma$}+\mbox{\boldmath$l$} (16)
+\displaystyle+ [(12ξs)(γnγp)𝝈(12ξl)𝒍]τ3}\displaystyle\bigl{[}\,(1-2\xi_{s})\,(\gamma_{n}-\gamma_{p}\,)\,\mbox{\boldmath$\sigma$}-(1-2\xi_{\,l})\,\mbox{\boldmath$l$}\,\bigr{]}\,\tau_{3}\Bigr{\}}

where 𝒍l is the single-particle operator of the angular momentum, 𝝈\sigma and τ3\tau_{3} are the spin and isospin Pauli matrices, respectively (with positive eigenvalue of τ3\tau_{3} for the neutrons), μN=e/2mpc\mu^{\vphantom{*}}_{N}=e\hbar/2m_{p}c is the nuclear magneton, γp=2.793\gamma_{p}=2.793 and γn=1.913\gamma_{n}=-1.913 are the spin gyromagnetic ratios, ξs\xi_{s} and ξl\xi_{\,l} are the renormalization constants. The nonzero ξs\xi_{s} and ξl\xi_{\,l} correspond to the effective operator 𝑸Q, however in the present calculations we used ξl=0\xi_{\,l}=0. Thus, the reduced probability of the M1M1 excitations B(M1)B(M1) is defined as |Zn|𝑸|2|\langle\,Z^{n}\,|\,\mbox{\boldmath$Q$}\rangle|^{2} in the RPA and as |zν|𝑸|2|\langle\,z^{\nu}\,|\,\mbox{\boldmath$Q$}\rangle|^{2} in the RenTBA.

The Skyrme EDF with the basis parametrizations SKXm Brown (1998) and SV-bas Klüpfel et al. (2009) was used both in RPA and RenTBA. The nuclear matter parameters for these parametrizations are listed in Table 1.

Table 1: Nuclear matter parameters: effective mass m/mm^{*}/m, incompressibility KK_{\infty}, Thomas-Reiche-Kuhn sum rule enhancement factor κTRK\kappa_{{}_{\mbox{\scriptsize TRK}}}, and symmetry energy asyma_{{}_{\mbox{\scriptsize sym}}} for two Skyrme-EDF parametrizations: SKXm Brown (1998) and SV-bas Klüpfel et al. (2009).
EDF m/mm^{*}/m KK_{\infty} κTRK\kappa_{{}_{\mbox{\scriptsize TRK}}} asyma_{{}_{\mbox{\scriptsize sym}}}
(MeV) (MeV)
SKXm 0.97 238 0.34 31
SV-bas 0.90 233 0.40 30

There are four experimental characteristics of the M1M1 excitations in 208Pb which serve as a benchmark in our calculations: energy and excitation probability of the isoscalar 11+1^{+}_{1} state (E1=5.84E_{1}=5.84 MeV with B1(M1)=2.0μN2B_{1}(M1)=2.0\;\mu^{2}_{N}, see Shizuma et al. (2008)) and the mean energy and the summed strength of the isovector M1M1 resonance in the interval 6.6–8.1 MeV (E2=7.4E_{2}=7.4 MeV with B2(M1)=B(M1)B_{2}(M1)=\sum B(M1) = 15.3 μN2\mu^{2}_{N}). The latter two quantities have been deduced by combining the data from Refs. Shizuma et al. (2008); Köhler et al. (1987).

To reproduce these key characteristics, the spin-related EDF parameters W0W_{0} (spin-orbit strength), xWx_{W} (proton-neutron balance of the spin-orbit term), gg (Landau parameter for isoscalar spin mode), and gg^{\prime} (Landau parameter for isovector spin mode) were refitted as explained in Tselyaev et al. (2019) while the remaining spin-related parameters of the functional were switched off. The values of all other parameters of the functional were kept at the values of the original parametrizations. The form of the EDF containing all the parameters mentioned above is given in Ref. Tselyaev et al. (2019). The spin-orbit parameters xWx_{W} and W0W_{0} were refitted to reproduce the experimental value of B1(M1)B_{1}(M1). The parameters gg and gg^{\prime} enter the terms of the modified Skyrme EDF which yield the term VsV^{s} of the residual interaction VV having the form of the Landau-Migdal ansatz

Vs=CN(g𝝈𝝈+g𝝈𝝈𝝉𝝉),V^{s}=C^{\vphantom{*}}_{\mbox{\scriptsize N}}\bigl{(}\,g\;\mbox{\boldmath$\sigma$}\cdot\mbox{\boldmath$\sigma^{\prime}$}+g^{\prime}\,\mbox{\boldmath$\sigma$}\cdot\mbox{\boldmath$\sigma^{\prime}$}\;\mbox{\boldmath$\tau$}\cdot\mbox{\boldmath$\tau^{\prime}$}\,\bigr{)}\,, (17)

where CNC^{\vphantom{*}}_{\mbox{\scriptsize N}} is the normalization constant. These parameters allow us to change the calculated energies of the isoscalar and isovector 1+1^{+} states.

Note that the parameter xWx_{W} was introduced in Refs. Reinhard and Flocard (1995); Sharma et al. (1995) (with the use of slightly different notations in Reinhard and Flocard (1995)) to regulate the isospin dependence of the spin-orbit potential. In the most parametrizations of the Skyrme EDF the value xW0x_{W}\geqslant 0 is used (see, e.g., last two lines of Table 2). In particular, the value xW=1x_{W}=1 (frequently used implicitly) corresponds to the usual Hartree-Fock approximation in the EDF for the two-body spin-orbit zero-range interaction. However, in all these cases the value of B1(M1)B_{1}(M1) in 208Pb calculated within the fully self-consistent RPA is much larger than its experimental value B1(M1)expB_{1}(M1)_{\mbox{\scriptsize exp}} = 2.0 μN2\mu^{2}_{N}. For instance, B1(M1)RPA10B1(M1)expB_{1}(M1)_{\mbox{\scriptsize RPA}}\approx 10\,B_{1}(M1)_{\mbox{\scriptsize exp}} for the SLy5 set Chabanat et al. (1998) even with the use of the effective M1M1 operator (16). In Ref. Tselyaev et al. (2019) we have shown that one should use the negative values of xWx_{W} to decrease the calculated B1(M1)B_{1}(M1) up to B1(M1)expB_{1}(M1)_{\mbox{\scriptsize exp}}. The values of the parameter W0W_{0} should be simultaneously increased because from the set of the refitted parameters xWx_{W}, W0W_{0}, gg, and gg^{\prime}, only the isoscalar combination of the spin-orbit parameters C0J=14(2+xW)W0C_{0}^{\nabla J}=-{\textstyle\frac{1}{4}}(2+x_{W})W_{0} have an impact on the ground-state characteristics of spherical nuclei (see Tselyaev et al. (2019) for more details). This combination remains approximately unchanged in our refitting procedure, so the quality of the description of the ground-state properties with the use of the original and modified parametrizations of the Skyrme EDF is approximately the same.

IV The main results for the 𝑴𝟏M1 resonance in Pb208{}^{208}\mbox{Pb}

The parametrizations obtained in the result of the refitting procedure described above are SKXm-0.54 and SV-bas-0.50 for the RPA and SKXm-0.49 and SV-bas-0.44 for the RenTBA (here and in the following the numerical subindex of the modified parametrization indicates the value of the parameter xWx_{W}). The values of the refitted parameters for these sets are shown in Table 2 along with several sets discussed in Sec. VI. In addition, we used the renormalization constant ξs\xi_{s} in the field operator of the M1M1 excitations (16) to fit the isovector M1M1 strength. The values of this constant for the RPA and RenTBA are also shown in Table 2.

Table 2: Parameters xWx_{W}, W0W_{0}, gg, and gg^{\prime} of the modified Skyrme EDFs determined on the basis of the parametrizations SKXm Brown (1998) and SV-bas Klüpfel et al. (2009). The Landau-Migdal parameters gg and gg^{\prime} are normalized to CN=300C_{\mbox{\scriptsize N}}=300 MeV\cdotfm3. The renormalization constants ξs\xi_{s} of the field operator of the M1M1 excitations corresponding to the each parametrization are shown in the last column. The parameters of the original sets are shown in the last two lines.
EDF xWx_{W} W0W_{0} gg gg^{\prime} ξs\xi_{s}
(MeV\cdotfm5)
SKXm-0.54 -0.54 226.0 -0.078 0.430 0.156
SV-bas-0.50 -0.50 213.0 -0.028 0.516 0.156
SKXm-0.49 -0.49 218.5  0.108 0.930 0.085
SKXm0.49{}^{\prime}_{-0.49} -0.49 218.5  0.108 0.900 0.085
SKXm0.49′′{}^{\prime\prime}_{-0.49} -0.49 218.5 -0.067 0.435 0.151
SV-bas-0.44 -0.44 204.7  0.177 1.030 0.085
SV-bas0.44{}^{\prime}_{-0.44} -0.44 204.7  0.177 1.460 0.085
SKXm 0 155.9 0 0
SV-bas  0.55 124.6 0 0

Note that the set of the phonons in the RenTBA after the renormalization procedure with the use of the condition (14) included 123 electric and 83 magnetic phonons for the parametrization SKXm-0.49 and 121 electric and 85 magnetic phonons for the parametrization SV-bas-0.44.

Most of the calculations presented below have been performed within the discrete versions of RPA and RenTBA that means that the model equations are solved in the discrete basis representation with the use of the box boundary conditions for all functions entering these equations. It is convenient to present these results as well as the experimental data in the form of the strength functions S(E)S(E) obtained by folding the discrete spectra with a Lorentzian of half-width Δ\Delta:

S(E)=Δπνsgn(ων)Bν(M1)(Eων)2+Δ2.S(E)=\frac{\Delta}{\pi}\sum_{\nu}\frac{\mbox{sgn}(\omega^{\vphantom{*}}_{\nu})B^{\vphantom{*}}_{\nu}(M1)}{(E-\omega^{\vphantom{*}}_{\nu})^{2}+\Delta^{2}}. (18)

The results for the modified SKXm parametrizations SKXm-0.49 (RenTBA) and SKXm-0.54 (RPA) obtained with Δ\Delta = 20 keV are shown on the upper panel of Fig. 1. The experimental spectra were taken from Refs. Shizuma et al. (2008) [208Pb(γ,γ)\,(\gamma,\gamma^{\prime}) reaction, data below the neutron separation energy S(n)=S(n)= 7.37 MeV] and Köhler et al. (1987) [207Pb(n,γ)\,(n,\gamma) reaction, data above S(n)S(n)].

Refer to caption
Figure 1: Upper panel: strength distributions of the M1M1 excitations in 208Pb calculated within the RenTBA with parametrization SKXm-0.49 (red solid line) and within the RPA with parametrization SKXm-0.54 (blue dashed line). The black dotted line represents the strength function (18) obtained from the experimental data Shizuma et al. (2008); Köhler et al. (1987). The smearing parameter Δ\Delta = 20 keV was used. See text for more details. Lower panel: the partial M1M1 cross section σM1\sigma_{M1} of the 208Pb(p,p)\,(p,p^{\prime}) reaction from Ref. Poltoratska et al. (2012).

The RenTBA, in contrast to the RPA, reproduces the experimental splitting of the M1M1 resonance into two components separated by the dip near 7.4 MeV. The quantitative characteristics of this splitting are given in Table 3 in comparison with the experiment.

Table 3: The summed strengths B(M1)\sum B(M1) and the mean energies E¯\bar{E} of the M1M1 excitations calculated within the RenTBA with parametrization SKXm-0.49 in two energy intervals. The last column contains the Gaussian width Γ\Gamma of the M1M1 strength distribution calculated in the interval 6.6–8.1 MeV. The experimental data are taken from Refs. Shizuma et al. (2008); Köhler et al. (1987).
interval 6.60–7.37 MeV 7.37–8.10 MeV
B(M1)<\sum B(M1)_{<} E¯<\bar{E}_{<} B(M1)>\sum B(M1)_{>} E¯>\bar{E}_{>} Γ\Gamma
(μN2)(\mu^{2}_{N}) (MeV) (μN2)(\mu^{2}_{N}) (MeV) (MeV)
theory 7.6 7.32 7.8 7.46 0.20
experiment 9.2 7.26 6.2 7.57 0.44

The experimental summed M1M1 strength in the energy interval below the neutron threshold B(M1)<\sum B(M1)_{<} is greater than the strength above the threshold B(M1)>\sum B(M1)_{>} by about 50%, while the respective theoretical values are approximately equal to each other. Nevertheless, the total theoretical summed M1M1 strength in the interval 6.6–8.1 MeV is equal to the experimental one according to the conditions of construction of our modified parametrizations. The absolute values of the calculated mean energies E¯<\bar{E}_{<} and E¯>\bar{E}_{>} are close to the experimental values, however the differences ΔE¯=E¯>E¯<\Delta\bar{E}=\bar{E}_{>}-\bar{E}_{<} are different: the theoretical value ΔE¯theor\Delta\bar{E}_{{}_{\mbox{\scriptsize theor}}} = 0.14 MeV is less than the experimental one ΔE¯exp\Delta\bar{E}_{{}_{\mbox{\scriptsize exp}}} = 0.31 MeV by a factor of two. To estimate the fragmentation of the M1M1 resonance we have also calculated the equivalent Gaussian width Γ\Gamma in the interval 6.6–8.1 MeV both for the experimental and for the theoretical strength distributions. The results presented in last column of Table 3 show that the total width of the resonance is still underestimated.

Existence of the dip near the neutron separation energy in the experimental distribution of the M1M1 strength in 208Pb is generally an uncertain point because the reliability of the experimental data Shizuma et al. (2008); Köhler et al. (1987) goes down in this region. To some extent, the possible existence of this dip is supported by the more recent data of the 208Pb(p,p)\,(p,p^{\prime}) experiment Poltoratska et al. (2012). The partial M1M1 cross section σM1\sigma_{M1} of this reaction is shown on the lower panel of Fig. 1. The dip in energy dependence of σM1\sigma_{M1} near 7.4 MeV exists though it is less pronounced than for the strength function obtained from the data Shizuma et al. (2008); Köhler et al. (1987). Note, however, the following. First, the direct comparison of the M1M1 strength functions S(E)S(E) and the cross section σM1(E)\sigma_{M1}(E) is hindered by the fact that they are determined by the different reaction mechanisms. The distribution of the B(M1)B(M1) values can be obtained from the cross section of the (p,p)(p,p^{\prime}) reaction only within the framework of some model assumptions, see, e.g., Ref. Birkhan et al. (2016). Second, the dip near 7.4 MeV is absent in the distribution of dB(M1)/dEdB(M1)/dE deduced in Birkhan et al. (2016) from the data of Ref. Poltoratska et al. (2012) and shown in Fig. 3(b) of Ref. Birkhan et al. (2016). But this fact can be explained by the different (and quite large) widths of the used energy bins that corresponds to the large and energy-dependent values of the smearing parameter Δ\Delta of the strength function (18).

V The fine structure of the 𝑴𝟏M1 resonance and the impact of the single-particle continuum

To show the fine structure of the theoretical and experimental strength distributions and to study the role of the single-particle continuum (which in principle can manifest itself above the neutron separation energy) we have calculated the M1M1 strength functions in 208Pb within the continuum RenTBA with Δ\Delta = 1 keV and 0.1 keV. The single-particle continuum was included within the response function formalism according to the method developed in Ref. Tselyaev et al. (2016). In this approach the strength function S(E)S(E) is expressed through the response function, and the right-hand side of Eq. (18) is supplemented with the contribution of the continuum part of the spectrum. The parametrizations SKXm-0.49 and SV-bas-0.44 (the latter is discussed in more detail in Sec. VI) were used.

Refer to caption
Figure 2: Upper panel: strength distributions of the M1M1 excitations in 208Pb calculated within the RenTBA with parametrizations SKXm-0.49 (red solid line) and SV-bas-0.44 (blue dashed line). The smearing parameter Δ\Delta = 1 keV was used. See text for more details. Lower panel: experimental distribution of the excitation probabilities B(M1)B(M1) in 208Pb in the interval 7–8 MeV from Refs. Shizuma et al. (2008) [208Pb(γ,γ)\,(\gamma,\gamma^{\prime}) reaction, red vertical lines] and Köhler et al. (1987) [207Pb(n,γ)\,(n,\gamma) reaction, green vertical lines]. The error bars are indicated by the black lines.

The results for Δ\Delta = 1 keV are shown on the upper panel of Fig. 2 in terms of the function B~M1(E)\tilde{B}_{M1}(E) defined as

B~M1(E)=πΔS(E).\tilde{B}_{M1}(E)=\pi\Delta S(E)\,. (19)

Here we use this function because, as follows from Eq. (18),

Bν(M1)=limΔ+0B~M1(ων).B^{\vphantom{*}}_{\nu}(M1)=\lim_{\Delta\rightarrow+0}\tilde{B}_{M1}(\omega^{\vphantom{*}}_{\nu})\,. (20)

So, if the Δ\Delta is small, the peak values of the function B~M1(E)\tilde{B}_{M1}(E) are close to the excitation probabilities at the peak energies. Note that Eq. (20) makes sense only for the states of the discrete spectrum. However, if the Δ\Delta is greater than the escape width of the quasidiscrete state in the continuum, the peak value of the function B~M1(E)\tilde{B}_{M1}(E) allows us to estimate the integrated strength of the single resonance.

In the RenTBA calculation with the SKXm-0.49 set and Δ\Delta = 1 keV, the fragmentation of two main peaks shown in Fig. 1 for the strength distributions with Δ\Delta = 20 keV is very small. This picture does not match the detailed fragmentation structure of the experimental distribution composed from data of Refs. Shizuma et al. (2008); Köhler et al. (1987) and shown on the lower panel of Fig. 2. The M1M1 strength in the interval 7–8 MeV obtained in the RenTBA with the parametrization SV-bas-0.44 is concentrated in one state without visible fragmentation, as in the case of the RPA.

The lack of fragmentation in the presented RenTBA calculations can be explained by the limited (though extended as compared to the RPA) kinds of the correlations included in the model. There are two natural generalizations of the RenTBA which enable one to include the additional correlations. First is the model taking into account the so-called ground-state correlations beyond the RPA. In Refs. Kamerdzhiev and Tselyaev (1991); Kamerdzhiev et al. (1993), it was shown that the inclusion of the correlations of this type increases the fragmentation of the M1M1 resonance in 208Pb. The second generalization is the replacement of the intermediate 1p1h1p1h\otimesphonon configurations by two-phonon configurations according to the scheme suggested in Tselyaev (2007) and in analogy with the first versions of the quasiparticle-phonon model Soloviev (1992). Note that the relative importance of these additional correlations increases at low energies due to the low level densities as compared to higher energies.

To analyze the effect of the single-particle continuum we first note that the theoretical neutron separation energies are equal to 7.30 MeV for the parametrization SKXm-0.49 and 7.64 MeV for the parametrization SV-bas-0.44. So, the single peak of the RenTBA strength distribution for the SV-bas-0.44 set shown on the upper panel of Fig. 2 (blue dashed line) is in the discrete spectrum, while the main strength of the distribution for the SKXm-0.49 set (red solid line) lies in the continuum.

The effect of the continuum is determined by the values of the escape widths of the resonances. The full width at half maximum (FWHM) of the single peak of the strength distribution corresponding to the one or several overlapping resonances is formed by the escape and spreading widths and by the artificial width of 2Δ2\Delta introduced by the smearing parameter. Thus, the FWHM can serve as an upper bound of the escape width. The distribution for the parametrization SKXm-0.49 shown on the upper panel of Fig. 2 contains three main peaks with the energies 7.313 MeV, 7.325 MeV, and 7.457 MeV. These peaks correspond to four states of the discrete RenTBA spectrum with the energies 7.313 MeV, 7.326 MeV, 7.457 MeV, and 7.459 MeV which exhaust 92% of the summed strength of the M1M1 resonance in the interval 6.6–8.1 MeV. So, we can confine ourselves to analyzing the widths of only these peaks. The respective values of the FWHM are equal to 2.1 keV for the quasidiscrete states with E=E= 7.313 MeV and 7.325 MeV and to 3.4 keV for the resonance with E=E= 7.457 MeV. The last FWHM value is appreciably greater than 2Δ2\Delta. This is explained by the fact that the peak with E=E= 7.457 MeV is formed by two overlapping resonances which correspond to two states of the discrete spectrum mentioned above.

In the calculation with Δ\Delta = 0.1 keV, the widths of the main peaks decrease. The values of the FWHM for the quasidiscrete states with E=E= 7.313 MeV and 7.325 MeV become less than 0.3 keV. The peak with E=E= 7.457 MeV is split into two peaks separated by the small interval of 2 keV and having the widths which are less than 1 keV. Thus the escape widths of the main peaks of the distribution for the SKXm-0.49 set are safely less than 1 keV. These results show that the inclusion of the single-particle continuum has no appreciable impact in the calculations with Δ=\Delta= 20 keV presented in the paper.

VI The problem of the fragmentation

The splitting of the isovector M1M1 resonance in 208Pb into two main peaks obtained in RenTBA with the use of the parametrization SKXm-0.49 is not a common result for the self-consistent calculations in our approach. In the typical case, if the EDF parameters gg and gg^{\prime} are fitted to reproduce the experimental energy of the 11+1_{1}^{+} state and the mean energy of the M1M1 resonance in 208Pb, the fragmentation of the isovector M1M1 resonance is reduced to the quenching of the main peak without appreciable broadening. This quenching is compensated by decreasing the renormalization constant ξs\xi_{s} after which the forms of the RenTBA and RPA M1M1 distributions become close to each other. This is illustrated in Fig. 3 where we show results for the modified SV-bas parametrizations.

Refer to caption
Figure 3: Strength distributions of the M1M1 excitations in 208Pb calculated within the RenTBA with parametrizations SV-bas-0.44 (red solid line) and SV-bas0.44{}^{\prime}_{-0.44} (green dashed-dotted line) and within the RPA with parametrization SV-bas-0.50 (blue dashed line) in comparison with the experiment (black dotted line). The smearing parameter Δ\Delta = 20 keV was used. See text for more details.

To clarify the problem, we note that the effects of the fragmentation of the RPA states in TBA and RenTBA are determined by the energy-dependent induced interaction W(ω){W}(\omega), Eq. (8a). The fragmentation of the RPA state with the energy ωRPA\omega_{{}_{\mbox{\scriptsize RPA}}} is strong if (i) one or more energies Ωc\Omega^{\vphantom{*}}_{c} of the 1p1h1p1h\otimesphonon configurations in Eqs. (8) are close to the shifted energy ω~RPA\tilde{\omega}_{{}_{\mbox{\scriptsize RPA}}} (shifted due to the regular contribution of the remaining 1p1h1p1h\otimesphonon configurations) and (ii) the respective amplitudes Fphc(+){F}^{c(+)}_{ph} are non-negligible. In the case of the nucleus 208Pb, the isovector M1M1 strength in the RPA is concentrated as a rule in one state with the energy ωRPA(12+)\omega_{{}_{\mbox{\scriptsize RPA}}}(1^{+}_{2}) (the 11+1^{+}_{1} RPA state is isoscalar) which is formed by two 1p1h1p1h configurations: π(1h9/21h11/21)\pi(1h_{9/2}\otimes 1h^{-1}_{11/2}) and ν(1i11/21i13/21)\nu(1i_{11/2}\otimes 1i^{-1}_{13/2}). So, the phph indices of the amplitudes Fphc(+){F}^{c(+)}_{ph} producing appreciable fragmentation of the 12+1^{+}_{2} RPA state should be one of these two combinations. Under this condition and according to the selection rules for the M1M1 excitations, the minimum value of Ωc\Omega^{\vphantom{*}}_{c} in 208Pb is determined by the configuration c={π(1h9/23s1/21)51}c=\{\pi(1h_{9/2}\otimes 3s^{-1}_{1/2})\otimes 5^{-}_{1}\}, that is

Ωcmin=εphπ+ω(51),\Omega^{\mbox{\scriptsize min}}_{c}=\varepsilon^{\pi}_{ph}+\omega(5^{-}_{1})\,, (21a)
where
εphπ=εpπ(1h9/2)εhπ(3s1/2).\varepsilon^{\pi}_{ph}=\varepsilon^{\pi}_{p}(1h_{9/2})-\varepsilon^{\pi}_{h}(3s_{1/2})\,. (21b)

It turns out that for most Skyrme EDF parametrizations the value of Ωcmin\Omega^{\mbox{\scriptsize min}}_{c} is substantially greater than the mean energy of the isovector M1M1 resonance in 208Pb, that is Ωcmin>7.4\Omega^{\mbox{\scriptsize min}}_{c}>7.4 MeV. Thus, if the parameters of the EDF are fitted to reproduce this mean energy, the fragmentation of the isovector M1M1 resonance is reduced to its quenching as mentioned above. The parametrization SKXm-0.49 is an exception because the value of ωRenTBA(51)\omega_{{}_{\mbox{\scriptsize RenTBA}}}(5^{-}_{1}) comes close the experimental value which, in turn, yields an Ωcmin\Omega^{\mbox{\scriptsize min}}_{c} close to 7.4 MeV. This is shown in Table 4 in comparison with the case of the SV-bas-0.44 parametrization.

Table 4: The values of the particle-hole energies εphπ=εpπ(1h9/2)εhπ(3s1/2)\varepsilon^{\pi}_{ph}=\varepsilon^{\pi}_{p}(1h_{9/2})-\varepsilon^{\pi}_{h}(3s_{1/2}), the energies of 515^{-}_{1} phonon, and their sums Ωcmin\Omega^{\mbox{\scriptsize min}}_{c}, Eqs. (21), in the RenTBA for the parametrizations SKXm-0.49 and SV-bas-0.44. The experimental values are given in the last line.
EDF εphπ\varepsilon^{\pi}_{ph} ω(51)\omega(5^{-}_{1}) Ωcmin\Omega^{\mbox{\scriptsize min}}_{c}
(MeV) (MeV) (MeV)
SKXm-0.49 4.14 3.24 7.38
SV-bas-0.44 4.27 3.55 7.82
experiment 4.21 3.20 7.41

Note that the splitting of the isovector M1M1 resonance shown in Fig. 1 is achieved only in the RenTBA. In conventional TBA, the energies of the phonons in Eqs. (8) are calculated within the RPA. In the case of the parametrization SKXm-0.49, the energy ωRPA(51)\omega_{{}_{\mbox{\scriptsize RPA}}}(5^{-}_{1}) = 3.64 MeV that increases the energy Ωcmin\Omega^{\mbox{\scriptsize min}}_{c} and leads to the RPA-like result in the TBA similar to shown in Fig. 3 by the red solid line.

On the other hand, the fragmentation of the isovector M1M1 resonance in 208Pb in itself can be obtained also in the case Ωcmin>7.4\Omega^{\mbox{\scriptsize min}}_{c}>7.4 MeV if the isovector M1M1 strength is shifted to higher energies by increasing the EDF parameter gg^{\prime}. This is shown in Fig. 3 for the parametrization SV-bas0.44{}^{\prime}_{-0.44} which is constructed from the set SV-bas-0.44 by changing the parameter gg^{\prime} from 1.03 for SV-bas-0.44 to 1.46 for SV-bas0.44{}^{\prime}_{-0.44} (however, the set of the phonons in this illustrative RenTBA calculation for SV-bas0.44{}^{\prime}_{-0.44} was used the same as for SV-bas-0.44). Thus, the simultaneous description of the mean energy of the isovector M1M1 resonance in 208Pb and of the fragmentation of this resonance in the self-consistent calculation is seemingly possible only in rare circumstances as, e.g., in case of the parametrization SKXm-0.49.

Note that the fragmentation of the isovector M1M1 resonance in 208Pb was obtained in the early calculations within the shell model in the 1p1h+2p2h1p1h+2p2h space Lee and Pittel (1975) and within the models based on the TFFS Migdal (1967) and including the particle-phonon interaction on top of the RPA (see, e.g., Dehesa et al. (1977); Kamerdzhiev and Tkachev (1984, 1989); Kamerdzhiev and Tselyaev (1991); Kamerdzhiev et al. (1993)). This result is explained by two reasons. First, the mean energy of the isovector M1M1 resonance in these calculations was greater than the experimental value. The shift to higher energies increases the spreading of the M1M1 strength as was noted in Ref. Lee and Pittel (1975) and is demonstrated in Fig. 3. Second, the phonon energies in the calculations of Refs. Kamerdzhiev and Tkachev (1984, 1989); Kamerdzhiev and Tselyaev (1991); Kamerdzhiev et al. (1993) were fitted to their experimental values that makes the value of Ωcmin\Omega^{\mbox{\scriptsize min}}_{c} more close to the mean energy of the isovector M1M1 resonance, see Table 4.

To demonstrate the role of the intermediate 1p1h1p1h\otimesphonon configuration π(1h9/23s1/21)51\pi(1h_{9/2}\otimes 3s^{-1}_{1/2})\otimes 5^{-}_{1} in the effect of the fragmentation under discussion we show in Fig. 4 the results of three RenTBA calculations with the use of parametrizations SKXm-0.49, SKXm0.49{}^{\prime}_{-0.49}, and SKXm0.49′′{}^{\prime\prime}_{-0.49} (see Table 2).

Refer to caption
Figure 4: Strength distributions of the M1M1 excitations in 208Pb calculated within the RenTBA with the full set of the phonons and with parametrization SKXm-0.49 (red solid line), with the set of all the phonons except for the 515^{-}_{1} phonon and with parametrization SKXm0.49{}^{\prime}_{-0.49} (blue dashed line) and with the set of the phonons including only the 515^{-}_{1} phonon and with parametrization SKXm0.49′′{}^{\prime\prime}_{-0.49} (green dashed-dotted line). The smearing parameter Δ\Delta = 20 keV was used. See text for more details.

The RenTBA calculation with the use of parametrization SKXm-0.49 coincides with one shown in Fig. 1. In the calculation with the use of SKXm0.49{}^{\prime}_{-0.49}, the 515^{-}_{1} phonon was excluded and the EDF parameter gg^{\prime} was slightly changed to fit the mean energy of the isovector M1M1 resonance to the experiment. The calculation with SKXm0.49′′{}^{\prime\prime}_{-0.49} represents the opposite case: only the 515^{-}_{1} phonon was included in the phonon basis of the RenTBA and the EDF parameters gg and gg^{\prime} were changed to fit the energy of the 11+1^{+}_{1} state and the mean energy of the isovector M1M1 resonance to the experiment. The renormalization constant ξs\xi_{s} was also changed to compensate decreasing of the quenching of the M1M1 strength. However, the characteristics of the same phonons (energies, etc.) were the same in all three calculations. These results show that the splitting of the isovector M1M1 resonance in 208Pb is determined in the considered model practically exclusively by the configuration π(1h9/23s1/21)51\pi(1h_{9/2}\otimes 3s^{-1}_{1/2})\otimes 5^{-}_{1}. The other 1p1h1p1h\otimesphonon configurations produce only the shift of the M1M1 resonance and the quenching of the M1M1 strength.

VII Results for the low-energy electric excitations in Pb208{}^{208}\mbox{Pb}

In section VI, we have shown that the RenTBA using the modified Skyrme EDF SKXm-0.49 gives an energy of the first 55^{-} state in 208Pb close to its experimental value. Here we consider the results of the RenTBA and RPA calculations for the first excited states of natural parity in 208Pb with the multipolarity LL from 2 to 6 both for the SKXm-0.49 and the SV-bas-0.44 parametrizations. The results are presented in Tables 5 and 6.

Table 5: The energies (in MeV) of the first excited states of the natural parity in 208Pb calculated within the RenTBA and the RPA with the use of the modified Skyrme EDFs SKXm-0.49 and SV-bas-0.44. The experimental data are taken from Ref. Martin (2007).
SKXm-0.49 SV-bas-0.44
LπL^{\pi} RenTBA RPA RenTBA RPA experiment
21+2^{+}_{1} 4.01 4.45 4.00 4.42 4.09
313^{-}_{1} 2.69 2.91 2.88 3.10 2.61
41+4^{+}_{1} 4.29 4.81 4.30 4.80 4.32
515^{-}_{1} 3.19 3.64 3.49 3.93 3.20
61+6^{+}_{1} 4.43 5.02 4.53 5.13 4.42

Note that the RenTBA results have been obtained without use of the diagonal approximation which is used in the model only for the phonons entering the intermediate 1p1h1p1h\otimesphonon configurations. It explains the small difference between the energies of the 515^{-}_{1} state listed in Tables 4 (where the diagonal approximation is used) and 5.

The RenTBA energies calculated with the parametrization SKXm-0.49 agree fairly well with the experiment. The deviations for SV-bas-0.44 are slightly greater (except for the 41+4^{+}_{1} state). The RPA gives too large energies for both parametrizations. The energy shift ω(RPA)ω(RenTBA)\omega(\mbox{RPA})-\omega(\mbox{RenTBA}) is between 0.2 MeV for the 313^{-}_{1} state and 0.6 MeV for the 61+6^{+}_{1} state.

The situation is the opposite for the excitation probabilities shown in Table 6.

Table 6: The same as in Table 5 but for the excitation probabilities B(EL)B(EL) (in units of e2e^{2}fm2L).
SKXm-0.49 SV-bas-0.44
LπL^{\pi} RenTBA RPA RenTBA RPA experiment
21+2^{+}_{1} 2.6×\times103 3.2×\times103 2.5×\times103 3.0×\times103 3.2×\times103
313^{-}_{1} 5.6×\times105 6.4×\times105 5.8×\times105 6.4×\times105 6.1×\times105
41+4^{+}_{1} 1.1×\times107 1.5×\times107 9.6×\times106 1.3×\times107 1.6×\times107
515^{-}_{1} 1.9×\times108 2.9×\times108 2.3×\times108 3.6×\times108 4.5×\times108
61+6^{+}_{1}   2.6×\,\,2.6\times1010   3.6×\,\,3.6\times1010   1.4×\,\,1.4\times1010   2.2×\,\,2.2\times1010   6.7×\,\,6.7\times1010

The RPA results are closer to the experiment as compared to the RenTBA results (and are in a good agreement with the experiment for 21+2^{+}_{1}, 313^{-}_{1}, and 41+4^{+}_{1} states). The decrease of the B(EL)B(EL) values in RenTBA is caused by the quenching as in the case of the M1M1 excitations.

By construction, the modified parametrizations SKXm-0.49 and SV-bas-0.44 describe the nuclear ground-state properties within the Skyrme EDF approach (with approximately the same accuracy as the original parametrizations SKXm and SV-bas) and reproduce the basic experimental characteristics of the M1M1 excitations in 208Pb within the RenTBA. The results of this section show that the RenTBA with the use of these modified parametrizations is applicable also to the description of the low-energy electric excitations in this nucleus.

VIII Conclusions

The present paper is a continuation of our recent work Tselyaev et al. (2019) in which we investigated the low-energy M1M1 excitations in 208Pb within the self-consistent RPA based on the Skyrme energy-density functionals (EDF). Here we use the extended self-consistent model including the particle-phonon coupling within the renormalized time blocking approximation (RenTBA, Tselyaev et al. (2018)). As in the case of the self-consistent RPA, the description of the basic experimental characteristics of the M1M1 excitations in 208Pb (energy and strength of the 11+1_{1}^{+} state as well as mean energy and summed strength of the isovector M1M1 resonance) requires refitting some of the spin-related parameters of the Skyrme EDF within the self-consistent RenTBA. We have determined several sets of these parameters from this condition. It has been shown that the observed fragmentation of the isovector M1M1 resonance in 208Pb which is absent in all the RPA calculations can be to a certain extent described within the self-consistent RenTBA. However, this description is not fully quantitative and is attained only in some cases of the modified functionals of the Skyrme type. We have found that the necessary condition to obtain this fragmentation in our model is the proximity of the energy of the intermediate 1p1h1p1h\otimesphonon configuration π(1h9/23s1/21)51\pi(1h_{9/2}\otimes 3s^{-1}_{1/2})\otimes 5^{-}_{1} to the mean energy of the isovector M1M1 resonance in 208Pb, i.e. the proximity of the energy of 515^{-}_{1} phonon to the experimental excitation energy of the 515^{-}_{1} state in 208Pb. We have also shown that the modified parametrizations of the Skyrme EDF presented in the paper can be used in the description of the low-energy electric excitations within the RenTBA.

Acknowledgements.
V.T. is grateful to Prof. V.Yu. Ponomarev for discussions. Research was carried out using computational resources provided by Resource Center “Computer Center of SPbU”.

References

  • Laszewski et al. (1988) R. M. Laszewski, R. Alarcon, D. S. Dale,  and S. D. Hoblit, Phys. Rev. Lett. 61, 1710 (1988).
  • Kamerdzhiev et al. (2004) S. Kamerdzhiev, J. Speth,  and G. Tertychny, Phys. Rep. 393, 1 (2004).
  • Vergados (1971) J. D. Vergados, Phys. Lett. B 36, 12 (1971).
  • Ring and Speth (1973) P. Ring and J. Speth, Phys. Lett. B 44, 477 (1973).
  • Tkachev et al. (1976) V. N. Tkachev, I. N. Borzov,  and S. P. Kamerdzhiev, Sov. J. Nucl. Phys. 24, 373 (1976).
  • Speth et al. (1980) J. Speth, V. Klemt, J. Wambach,  and G. E. Brown, Nucl. Phys. A 343, 382 (1980).
  • Borzov et al. (1984) I. N. Borzov, S. V. Tolokonnikov,  and S. A. Fayans, Sov. J. Nucl. Phys. 40, 732 (1984).
  • Migdal (1967) A. B. Migdal, Theory of Finite Fermi Systems and Application to Atomic Nuclei (Wiley, New York, 1967).
  • Lee and Pittel (1975) T.-S. H. Lee and S. Pittel, Phys. Rev. C 11, 607 (1975).
  • Dehesa et al. (1977) J. S. Dehesa, J. Speth,  and A. Faessler, Phys. Rev. Lett. 38, 208 (1977).
  • Kamerdzhiev and Tkachev (1984) S. P. Kamerdzhiev and V. N. Tkachev, Phys. Lett. B 142, 225 (1984).
  • Cha et al. (1984) D. Cha, B. Schwesinger, J. Wambach,  and J. Speth, Nucl. Phys. A 430, 321 (1984).
  • Khoa et al. (1986) D. T. Khoa, V. Y. Ponomarev,  and A. I. Vdovin, Preprint JINR E4-86-198 (1986).
  • Kamerdzhiev and Tkachev (1989) S. P. Kamerdzhiev and V. N. Tkachev, Z. Phys. A 334, 19 (1989).
  • Tselyaev (1989) V. I. Tselyaev, Sov. J. Nucl. Phys. 50, 780 (1989).
  • Kamerdzhiev and Tselyaev (1991) S. P. Kamerdzhiev and V. I. Tselyaev, Bull. Acad. Sci. USSR, Phys. Ser. 55, 45 (1991).
  • Kamerdzhiev et al. (1993) S. P. Kamerdzhiev, J. Speth, G. Tertychny,  and J. Wambach, Z. Phys. A 346, 253 (1993).
  • Cao et al. (2009) L.-G. Cao, G. Colò, H. Sagawa, P. F. Bortignon,  and L. Sciacchitano, Phys. Rev. C 80, 064304 (2009).
  • Vesely et al. (2009) P. Vesely, J. Kvasil, V. O. Nesterenko, W. Kleinig, P.-G. Reinhard,  and V. Y. Ponomarev, Phys. Rev. C 80, 031302(R) (2009).
  • Nesterenko et al. (2010) V. O. Nesterenko, J. Kvasil, P. Vesely, W. Kleinig, P.-G. Reinhard,  and V. Y. Ponomarev, J. Phys. G: Nucl. Part. Phys. 37, 064034 (2010).
  • Cao et al. (2011) L.-G. Cao, H. Sagawa,  and G. Colò, Phys. Rev. C 83, 034324 (2011).
  • Wen et al. (2014) P. Wen, L.-G. Cao, J. Margueron,  and H. Sagawa, Phys. Rev. C 89, 044311 (2014).
  • Tselyaev et al. (2019) V. Tselyaev, N. Lyutorovich, J. Speth, P.-G. Reinhard,  and D. Smirnov, Phys. Rev. C 99, 064329 (2019).
  • Bender et al. (2003) M. Bender, P.-H. Heenen,  and P.-G. Reinhard, Rev. Mod. Phys. 75, 121 (2003).
  • Tselyaev et al. (2018) V. Tselyaev, N. Lyutorovich, J. Speth,  and P.-G. Reinhard, Phys. Rev. C 97, 044308 (2018).
  • Toepffer and Reinhard (1988) C. Toepffer and P.-G. Reinhard, Ann. Phys. (N.Y.) 181, 1 (1988).
  • Gütter et al. (1993) K. Gütter, P.-G. Reinhard, K. Wagner,  and C. Toepffer, Ann. Phys. (N.Y.) 225, 339 (1993).
  • Tselyaev (2013) V. I. Tselyaev, Phys. Rev. C 88, 054301 (2013).
  • Lyutorovich et al. (2015) N. Lyutorovich, V. Tselyaev, J. Speth, S. Krewald, F. Grümmer,  and P.-G. Reinhard, Phys. Lett. B 749, 292 (2015).
  • Lyutorovich et al. (2016) N. Lyutorovich, V. Tselyaev, J. Speth, S. Krewald,  and P.-G. Reinhard, Phys. At. Nucl. 79, 868 (2016).
  • Tselyaev et al. (2016) V. Tselyaev, N. Lyutorovich, J. Speth, S. Krewald,  and P.-G. Reinhard, Phys. Rev. C 94, 034306 (2016).
  • Brown (1998) B. A. Brown, Phys. Rev. C 58, 220 (1998).
  • Klüpfel et al. (2009) P. Klüpfel, P.-G. Reinhard, T. J. Bürvenich,  and J. A. Maruhn, Phys. Rev. C 79, 034310 (2009).
  • Shizuma et al. (2008) T. Shizuma, T. Hayakawa, H. Ohgaki, T. Toyokawa, T. Komatsubara, N. Kikuzawa, A. Tamii,  and H. Nakada, Phys. Rev. C 78, 061303(R) (2008).
  • Köhler et al. (1987) R. Köhler, J. A. Wartena, H. Weigmann, L. Mewissen, F. Poortmans, J. P. Theobald,  and S. Raman, Phys. Rev. C 35, 1646 (1987).
  • Reinhard and Flocard (1995) P.-G. Reinhard and H. Flocard, Nucl. Phys. A 584, 467 (1995).
  • Sharma et al. (1995) M. M. Sharma, G. Lalazissis, J. König,  and P. Ring, Phys. Rev. Lett. 74, 3744 (1995).
  • Chabanat et al. (1998) E. Chabanat, P. Bonche, P. Haensel, J. Meyer,  and R. Schaeffer, Nucl. Phys. A 635, 231 (1998).
  • Poltoratska et al. (2012) I. Poltoratska, P. von Neumann-Cosel, A. Tamii, T. Adachi, C. A. Bertulani, J. Carter, M. Dozono, H. Fujita, K. Fujita, Y. Fujita, K. Hatanaka, M. Itoh, T. Kawabata, Y. Kalmykov, A. M. Krumbholz, E. Litvinova, H. Matsubara, K. Nakanishi, R. Neveling, H. Okamura, H. J. Ong, B. Özel-Tashenov, V. Y. Ponomarev, A. Richter, B. Rubio, H. Sakaguchi, Y. Sakemi, Y. Sasamoto, Y. Shimbara, Y. Shimizu, F. D. Smit, T. Suzuki, Y. Tameshige, J. Wambach, M. Yosoi,  and J. Zenihiro, Phys. Rev. C 85, 041304(R) (2012).
  • Birkhan et al. (2016) J. Birkhan, H. Matsubara, P. von Neumann-Cosel, N. Pietralla, V. Y. Ponomarev, A. Richter, A. Tamii,  and J. Wambach, Phys. Rev. C 93, 041302(R) (2016).
  • Tselyaev (2007) V. I. Tselyaev, Phys. Rev. C 75, 024306 (2007).
  • Soloviev (1992) V. G. Soloviev, Theory of Atomic Nuclei: Quasiparticles and Phonons (Institute of Physics, Bristol and Philadelphia, 1992).
  • Martin (2007) M. Martin, Nuclear Data Sheets 108, 1583 (2007).