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

Simultaneous description of β\beta decay and low-lying structure of neutron-rich even- and odd-mass Rh and Pd nuclei

K. Nomura [email protected]    L. Lotina Department of Physics, Faculty of Science, University of Zagreb, HR-10000 Zagreb, Croatia    R. Rodríguez-Guzmán Departamento de Física Aplicada I, Escuela Politécnica Superior, Universidad de Sevilla, Seville, E-41011, Spain    L. M. Robledo Departamento de Física Teórica and CIAFF, Universidad Autónoma de Madrid, E-28049 Madrid, Spain Center for Computational Simulation, Universidad Politécnica de Madrid, Campus de Montegancedo, Bohadilla del Monte, E-28660-Madrid, Spain
Abstract

The low-energy structure and β\beta decay properties of neutron-rich even- and odd-mass Pd and Rh nuclei are studied using a mapping framework based on the nuclear density functional theory and the particle-boson coupling scheme. Constrained Hartree-Fock-Bogoliubov calculations using the Gogny-D1M energy density functional are performed to obtain microscopic inputs to determine the interacting-boson Hamiltonian employed to describe the even-even core Pd nuclei. The mean-field calculations also provide single-particle energies for the odd systems, which are used to determine essential ingredients of the particle-boson interactions for the odd-nucleon systems, and of the Gamow-Teller and Fermi transition operators. The potential energy surfaces obtained for even-even Pd isotopes as well as the spectroscopic properties for the even- and odd-mass systems suggest a transition from prolate deformed to γ\gamma-unstable and to nearly-spherical shapes. The predicted β\beta decay logft\log{ft} values are shown to be sensitive to the details of the wave functions for the parent and daughter nuclei, and therefore serve as a stringent test of the employed theoretical approach.

I Introduction

Precise measurements and theoretical descriptions associated with the low-energy nuclear structure are crucial to the accurate modeling and better understanding of fundamental nuclear processes, such as, β\beta and double-β\beta (ββ\beta\beta) decays intimately connected to stellar nucleosynthesis. In this context, the low-energy excitations and decay properties of neutron-rich nuclei with mass A100A\approx 100 and neutron number N60N\approx 60 are of particular interest from both the nuclear structure and astrophysical points of view. Those nuclei exhibit a rich variety of phenomena such as shell evolution, onset of collectivity, quantum (shape) phase transitions and shape coexistence. They are also involved in the rapid neutron-capture (rr) process responsible for the nucleosynthesis of heavy chemical elements in explosive environments.

The β\beta decay half-lives of heavy neutron-rich nuclei have been extensively measured using radioactive-ion beams at major experimental facilities around the world. For example, the neutron-rich A110A\approx 110 nuclei from Kr to Tc [1], and from Rb to Sn [2] have been studied at the RIBF facility at RIKEN. The A90A\approx 90 region from Se to Zr isotopic chains has been studied at the NSCL at MSU [3]. Moreover, several A100110A\approx 100-110 nuclei are of special interest, including 96Zr, 96Mo, 100Mo, 100Ru, 110Pd, and 110Cd, since they correspond to the parent or daughter nuclei for the possible neutrinoless ββ\beta\beta decays [4].

From a theoretical point of view, the consistent description of both low-lying nuclear states and β\beta decay properties represents a major challenge. Theoretical studies of the β\beta decay process have been carried out within the interacting boson model (IBM) [5, 6, 7, 8, 9, 10, 11, 12, 13, 14], the quasiparticle random-phase approximation (QRPA) [15, 16, 17, 18, 19, 20, 21, 22, 23], and the large-scale shell model (LSSM) [24, 25, 26, 27, 28]. The calculation of β\beta decay properties serves as a stringent test of a given theoretical approach, since the decay rate of this process is very sensitive to the structure of the wave functions corresponding to the low-energy states of both the parent and daughter nuclei.

In this paper, we present a simultaneous description of the low-energy collective excitations and β\beta-decay properties of even- and odd-AA neutron-rich Pd and Rh isotopes in the mass range A100120A\approx 100-120. They represent a region of interest for future experiments and for astrophysical applications. Calculations are performed within a theoretical framework based on the nuclear density functional theory and the particle-core coupling scheme. In it even-even nuclei are described using the IBM [29]. The particle-core couplings for the odd-mass, and odd-odd nuclei are described using the interacting boson-fermion model (IBFM) [30, 31] and the interacting boson-fermion-fermion model (IBFFM) [31, 32], respectively. The bosonic-core Hamiltonian is built using microscopic input from self-consistent Hartree-Fock-Bogoliubov (HFB) [33] calculations based on the parametrization D1M [34] of the Gogny energy density functional (EDF) [35, 36]. Essential building blocks of the particle-boson interactions and of the Gamow-Teller (GT) and Fermi (F) transition operators for the β\beta decay are also determined with the aid of the same Gogny-EDF results. The method has already been applied to study the shape evolution and β\beta decay properties of the odd-AA [11] and even-AA [12] nuclei in the mass A130A\approx 130 region. It has also been employed to study even- and odd-AA As and Ge nuclei in the A70A\approx 70-80 region using microscopic input from relativistic Hartree-Bogoliubov calculations, based on the density-dependent point-coupling interaction [14].

The main goal of this work is to examine the performance of the method mentioned above in the case of neutron-rich nuclei, including those for which experimental information is scarce. The results to be discussed latter on in the paper also illustrate the predictive power of the EDF-based IBM to describe the low-lying structure and β\beta decay in this region of the nuclear chart where future experiments are expected. To identify the relevance of the low-lying structures of individual nuclei in the β\beta decay, we perform a detailed analysis of the wave functions obtained for both the parent and daughter nuclei of the decay. In addition, we perform conventional IBM calculations, with the parameters for the even-even boson core Hamiltonians taken from the earlier phenomenological calculation [37]. The corresponding results are compared with those from the EDF-based IBM calculations. Note that the present study is restricted to both types of allowed β\beta decays, i.e., the transition conserves parity and takes place between states that differ in the total angular momentum II by ΔI=0\Delta I=0 or 1.

To support our choice we note that, like other nonrelativistic [38] and relativistic [39, 40] EDFs, theoretical approaches based on the parametrizations D1M and D1S [41] of the Gogny-EDF both at the mean-field level and beyond have been extensively employed to study the low-energy nuclear structure and dynamics in various regions of the nuclear chart as well as fundamental nuclear processes (see Ref. [36] for a review, and references therein). In particular spectroscopic studies involving collective degrees of freedom have been carried out within the symmetry-projected generator coordinate method (GCM) [33] using the Gogny forces and involving different levels of sophistication [42, 36, 43, 44, 45, 46, 47, 48]. Furthermore, the mapping procedure leading to an IBM Hamiltonian from microscopic Gogny mean-field input has already shown its ability to describe spectroscopic properties associated with shape phase transitions, shape coexistence, and octupole deformations in nuclei [49, 50, 51, 52, 53, 54, 55, 56].

The paper is organized as follows. The theoretical framework is briefly outlined in Sec. II. The excitation spectra and electromagnetic transition properties obtained for even-even Pd (Sec. III), odd-AA Pd and Rh (Sec. IV), and odd-odd Rh nuclei (Sec. V) are discussed. The computed logft\log{ft} values for the β\beta decays of the odd- and even-AA Rh into Pd nuclei are discussed in detail in Sec. VI. Finally, Sec. VII is devoted to the concluding remarks.

II Theoretical framework

In this section, we describe the particle-core Hamiltonian (Sec. II.1), and the procedure to build it (Sec. II.2). Electromagnetic transition operators are discussed in Sec. II.3, and Gamow-Teller and Fermi operators are introduced in Sec. II.4.

II.1 Particle-core Hamiltonian

In this study, we use the neutron-proton IBM (IBM-2) [57, 58]. In this model both neutron and proton monopole (sνs_{\nu} and sπs_{\pi}), and quadrupole (dνd_{\nu} and dπd_{\pi}) bosons are considered as fundamental degrees of freedom. From a microscopic point of view [58, 57], the sνs_{\nu} (sπs_{\pi}) and dνd_{\nu} (dπd_{\pi}) bosons are associated with the collective SνS_{\nu} (SπS_{\pi}) and DνD_{\nu} (DπD_{\pi}) pairs of valence neutrons (protons) with angular momenta and parity 0+0^{+} and 2+2^{+}, respectively. In comparison with the simpler IBM-1, in which the neutrons and protons are not distinguished, the IBM-2 appears to be more suitable to treat β\beta decay, since in this process both proton and neutron degrees of freedom should be explicitly taken into account. For the model space the neutron N=N= 50-82 and proton Z=Z= 28-50 major shells are used. Hence for 104-124Pd, the number of neutron bosons, NνN_{\nu}, varies within the range 2Nν82\leqslant N_{\nu}\leqslant 8, while the number of the proton bosons is fixed, Nπ=2N_{\pi}=2.

To deal with even-even, odd-mass, and odd-odd nuclei on an equal footing, both collective and single-particle degrees of freedom are treated within the framework of the neutron-proton IBFFM (IBFFM-2). The IBFFM-2 Hamiltonian reads

H^=H^B+H^Fν+H^Fπ+V^BFν+V^BFπ+V^νπ,\displaystyle\hat{H}=\hat{H}_{\mathrm{B}}+\hat{H}_{\mathrm{F}}^{\nu}+\hat{H}_{\mathrm{F}}^{\pi}+\hat{V}_{\mathrm{BF}}^{\nu}+\hat{V}_{\mathrm{BF}}^{\pi}+\hat{V}_{\nu\pi}, (1)

where H^B\hat{H}_{\mathrm{B}} is the IBM-2 Hamiltonian representing the bosonic even-even core, H^Fν\hat{H}_{\mathrm{F}}^{\nu} (H^Fπ\hat{H}_{\mathrm{F}}^{\pi}) is the one-body, single-neutron (-proton) Hamiltonian, and V^BFν\hat{V}_{\mathrm{BF}}^{\nu} (V^BFπ\hat{V}_{\mathrm{BF}}^{\pi}) stands for the interaction between the odd neutron (proton) and the even-even IBM-2 core. The last term V^νπ\hat{V}_{\nu\pi} represents the residual interaction between the odd neutron and the odd proton.

The IBM-2 Hamiltonian takes the form

H^B=ϵd(n^dν+n^dπ)+κQ^νQ^π,\displaystyle\hat{H}_{\mathrm{B}}=\epsilon_{d}(\hat{n}_{d_{\nu}}+\hat{n}_{d_{\pi}})+\kappa\hat{Q}_{\nu}\cdot\hat{Q}_{\pi}, (2)

where in the first term, n^dρ=dρd~ρ\hat{n}_{d_{\rho}}=d^{\dagger}_{\rho}\cdot\tilde{d}_{\rho} (ρ=ν\rho=\nu or π\pi) is the dd-boson number operator, with ϵd\epsilon_{d} the single dd-boson energy relative to the ss-boson one, and d~ρμ=(1)μdρμ\tilde{d}_{\rho\mu}=(-1)^{\mu}d_{\rho-\mu}. The second term stands for the quadrupole-quadrupole interaction between neutron and proton boson systems with strength κ\kappa, and Q^ρ=dρsρ+sρd~ρ+χρ(dρ×d~ρ)(2)\hat{Q}_{\rho}=d_{\rho}^{\dagger}s_{\rho}+s_{\rho}^{\dagger}\tilde{d}_{\rho}+\chi_{\rho}(d^{\dagger}_{\rho}\times\tilde{d}_{\rho})^{(2)} represents the bosonic quadrupole operator, with the dimensionless parameter χρ\chi_{\rho}.

The single-nucleon Hamiltonian H^Fρ\hat{H}_{\mathrm{F}}^{\rho} takes the form

H^Fρ=jρϵjρ2jρ+1(ajρ×a~jρ)(0),\displaystyle\hat{H}_{\mathrm{F}}^{\rho}=-\sum_{j_{\rho}}\epsilon_{j_{\rho}}\sqrt{2j_{\rho}+1}(a_{j_{\rho}}^{\dagger}\times\tilde{a}_{j_{\rho}})^{(0)}, (3)

where ϵjρ\epsilon_{j_{\rho}} stands for the single-particle energy of the odd neutron (ρ=ν)(\rho=\nu) or proton (ρ=π\rho=\pi) orbital jρj_{\rho}. ajρa_{j_{\rho}} and ajρa_{j_{\rho}}^{\dagger} are annihilation and creation operators of the single particle, respectively. The operator a~jρ\tilde{a}_{j_{\rho}} is defined as a~jρmρ=(1)jρmρajρmρ\tilde{a}_{j_{\rho}m_{\rho}}=(-1)^{j_{\rho}-m_{\rho}}a_{j_{\rho}-m_{\rho}}.

In this study, we employ the following boson-fermion interaction V^BFρ\hat{V}_{\mathrm{BF}}^{\rho} [31]

V^BFρ=ΓρV^dynρ+ΛρV^excρ+AρV^monρ.\hat{V}_{\mathrm{BF}}^{\rho}=\Gamma_{\rho}\hat{V}_{\mathrm{dyn}}^{\rho}+\Lambda_{\rho}\hat{V}_{\mathrm{exc}}^{\rho}+A_{\rho}\hat{V}_{\mathrm{mon}}^{\rho}. (4)

The first, second, and third terms are dynamical quadrupole, exchange, and monopole interactions, respectively. Within the generalized seniority scheme [59, 31], the dynamical and exchange terms are assumed to be dominated by the interaction between unlike particles. On the other hand, the monopole term is assumed to be dominated by the interaction between like particles. The explicit form of the different terms in Eq. (4) then read

V^dynρ=jρjργjρjρ(ajρ×a~jρ)(2)Q^ρ,\displaystyle\hat{V}_{\mathrm{dyn}}^{\rho}=\sum_{j_{\rho}j_{\rho}^{\prime}}\gamma_{j_{\rho}j_{\rho}^{\prime}}(a^{\dagger}_{j_{\rho}}\times\tilde{a}_{j_{\rho}^{\prime}})^{(2)}\cdot\hat{Q}_{\rho^{\prime}}, (5)
V^excρ=(sρ×d~ρ)(2)jρjρjρ′′10Nρ(2jρ+1)βjρjρβjρ′′jρ\displaystyle\hat{V}^{\rho}_{\mathrm{exc}}=-\left(s_{\rho^{\prime}}^{\dagger}\times\tilde{d}_{\rho^{\prime}}\right)^{(2)}\cdot\sum_{j_{\rho}j_{\rho}^{\prime}j_{\rho}^{\prime\prime}}\sqrt{\frac{10}{N_{\rho}(2j_{\rho}+1)}}\beta_{j_{\rho}j_{\rho}^{\prime}}\beta_{j_{\rho}^{\prime\prime}j_{\rho}}
:((dρ×a~jρ′′)(jρ)×(ajρ×s~ρ)(jρ))(2):+(H.c.),\displaystyle{\quad}:\left((d_{\rho}^{\dagger}\times\tilde{a}_{j_{\rho}^{\prime\prime}})^{(j_{\rho})}\times(a_{j_{\rho}^{\prime}}^{\dagger}\times\tilde{s}_{\rho})^{(j_{\rho}^{\prime})}\right)^{(2)}:+(\text{H.c.}), (6)
V^monρ=n^dρjρ2jρ+1(ajρ×a~jρ)(0),\displaystyle\hat{V}_{\mathrm{mon}}^{\rho}=-\hat{n}_{d_{\rho}}\sum_{j_{\rho}}\sqrt{2j_{\rho}+1}(a_{j_{\rho}}^{\dagger}\times\tilde{a}_{j_{\rho}})^{(0)}, (7)

where the coefficients γjρjρ=(ujρujρvjρvjρ)Qjρjρ\gamma_{j_{\rho}j_{\rho}^{\prime}}=(u_{j_{\rho}}u_{j_{\rho}^{\prime}}-v_{j_{\rho}}v_{j_{\rho}^{\prime}})Q_{j_{\rho}j_{\rho}^{\prime}}, and βjρjρ=(ujρvjρ+vjρujρ)Qjρjρ\beta_{j_{\rho}j_{\rho}^{\prime}}=(u_{j_{\rho}}v_{j_{\rho}^{\prime}}+v_{j_{\rho}}u_{j_{\rho}^{\prime}})Q_{j_{\rho}j_{\rho}^{\prime}} are proportional to the matrix elements of the fermion quadrupole operator in the single-particle basis Qjρjρ=ρ12jρY(2)ρ12jρQ_{j_{\rho}j_{\rho}^{\prime}}=\braket{\ell_{\rho}\frac{1}{2}j_{\rho}}{Y^{(2)}}{\ell^{\prime}_{\rho}\frac{1}{2}j_{\rho}^{\prime}}. The operator Q^ρ\hat{Q}_{\rho^{\prime}} in Eq. (5) is the same boson quadrupole operator as in the boson Hamiltonian (2). In Eq. (II.1) the notation :()::(\cdots): stands for normal ordering. Within this formalism, the single-particle energy ϵjρ\epsilon_{j_{\rho}} in Eq. (3) is replaced with the quasiparticle energy ϵ~jρ\tilde{\epsilon}_{j_{\rho}}.

For the residual neutron-proton interaction V^νπ\hat{V}_{\nu\pi} in Eq. (1), we adopt the form [60]

V^νπ=4πvd\displaystyle\hat{V}_{\nu\pi}=4\pi{v_{\mathrm{d}}} δ(𝒓)δ(𝒓νr0)δ(𝒓πr0)\displaystyle\delta(\bm{r})\delta(\bm{r}_{\nu}-r_{0})\delta(\bm{r}_{\pi}-r_{0})
+vt[3(𝝈ν𝐫)(𝝈π𝐫)r2𝝈ν𝝈π],\displaystyle+v_{\mathrm{t}}\left[\frac{3({\bm{\sigma}}_{\nu}\cdot{\bf r})({\bm{\sigma}}_{\pi}\cdot{\bf r})}{r^{2}}-{\bm{\sigma}}_{\nu}\cdot{\bm{\sigma}}_{\pi}\right], (8)

where the first and second terms are surface-delta and tensor interactions with strength parameters vdv_{\mathrm{d}}, and vtv_{\mathrm{t}}, respectively. Note that 𝒓=𝒓ν𝒓π\bm{r}=\bm{r}_{\nu}-\bm{r}_{\pi} and r0=1.2A1/3r_{0}=1.2A^{1/3} fm.

Table 1 summarizes the even-even Pd core nuclei, neighboring odd-AA Pd and Rh, and odd-odd Rh nuclei considered in this study.

II.2 Procedure to build the Hamiltonian

Refer to caption
Figure 1: (a), (b) The single-particle energies ϵjρ\epsilon_{j_{\rho}}, obtained from the Gogny-D1M HFB calculations at the spherical configuration, (c), (d) the quasiparticle energies ϵ~jρ\tilde{\epsilon}_{j_{\rho}}, and (e), (f) the occupation probabilities vjρ2v^{2}_{j_{\rho}}, obtained from the BCS calculations. Results shown in the left column are for the odd neutron in the odd-AA Pd and even-AA Rh nuclei, and those in the right column are for the odd proton in the even- and odd-AA Rh nuclei.
Table 1: Even-even Pd core, and the neighboring odd-NN Pd, odd-ZZ Rh, and odd-odd Rh nuclei considered in this study.
even-even core odd-NN odd-ZZ odd-odd
46A{}^{A}_{46}PdN (58N6458\leqslant N\leqslant 64) 46A+1{}^{A+1}_{46}PdN+1 45A1{}^{A-1}_{45}RhN 45A{}^{A}_{45}RhN+1
46112{}^{112}_{46}Pd66 45111{}^{111}_{45}Rh66
46A{}^{A}_{46}PdN (68N7868\leqslant N\leqslant 78) 46A1{}^{A-1}_{46}PdN-1 45A1{}^{A-1}_{45}RhN 45A2{}^{A-2}_{45}RhN-1

In the initial step a set of constrained HFB calculations for even-even Pd isotopes based on the parametrization D1M of the Gogny-EDF is carried out to obtain the microscopic input to build the IBFFM-2 Hamiltonian. For each even-even Pd isotope, those calculations provide the corresponding energy surfaces, i.e., the total mean-field energies as functions of the triaxial quadrupole deformations β\beta and γ\gamma [61]. For each nucleus, the Gogny-D1M HFB energy surface is mapped onto the expectation value of the IBM-2 Hamiltonian H^B\hat{H}_{\mathrm{B}} (2) in the boson condensate state [62]. This procedure specifies the parameters of the boson Hamiltonian, i.e., ϵd\epsilon_{d}, κ\kappa, χν\chi_{\nu}, and χπ\chi_{\pi}. For more details about the mapping procedure, the reader is referred to Refs. [63, 64].

Next, the Hamiltonian H^F\hat{H}_{\mathrm{F}} of Eq. (3) and the boson-fermion interactions V^BF\hat{V}_{\mathrm{BF}} of Eq. (4) are determined using the procedure of Refs. [65, 66]. The single-particle energies ϵjρ\epsilon_{j_{\rho}} of the odd nucleon are obtained from HFB calculations constrained to zero quadrupole deformation. Once the single-particle energies are available, the quasiparticle energies ϵ~jρ\tilde{\epsilon}_{j_{\rho}} and occupation probabilities vjρ2v^{2}_{j_{\rho}} are computed within the BCS approximation, separately for neutron and proton single-particle spaces. The empirical pairing gap 12A1/212A^{-1/2} is used. We include in the BCS calculations the 2s1/22s_{1/2}, 1d3/21d_{3/2}, 1d5/21d_{5/2}, 0g7/20g_{7/2}, 0g9/20g_{9/2}, and 0h11/20h_{11/2} orbitals for the odd neutron, and the 1d5/21d_{5/2}, 0g7/20g_{7/2}, 0g9/20g_{9/2}, 2p1/22p_{1/2}, 2p3/22p_{3/2}, and 1f5/21f_{5/2} orbitals for the odd proton. The corresponding quasiparticle energies ϵ~jν\tilde{\epsilon}_{j_{\nu}} (ϵ~jπ\tilde{\epsilon}_{j_{\pi}}), and occupation probabilities vjν2v^{2}_{j_{\nu}} (vjπ2v^{2}_{j_{\pi}}) for the odd neutron (proton) 2s1/22s_{1/2}, 1d3/21d_{3/2}, 1d5/21d_{5/2}, and 0g7/20g_{7/2} (1d5/21d_{5/2}, 0g7/20g_{7/2}, and 0g9/20g_{9/2}) orbitals are taken as the inputs to H^Fν\hat{H}_{\mathrm{F}}^{\nu} (H^Fπ\hat{H}_{\mathrm{F}}^{\pi}) and V^BFν\hat{V}_{\mathrm{BF}}^{\nu} (V^BFπ\hat{V}_{\mathrm{BF}}^{\pi}), respectively. The strength parameters Γρ\Gamma_{\rho}, Λρ\Lambda_{\rho}, and AρA_{\rho} for V^BFρ\hat{V}_{\mathrm{BF}}^{\rho} are then fixed so that the observed low-energy positive-parity levels for the odd-AA Pd (ρ=ν\rho=\nu) or odd-ZZ Rh (ρ=π\rho=\pi) nuclei are reproduced reasonably well.

Finally, the parameters vdv_{\mathrm{d}} and vtv_{\mathrm{t}} for the residual neutron-proton interaction in Eq. (II.1) are determined [67] so that the observed low-lying positive-parity states for each odd-odd Rh nucleus are reasonably well reproduced. Note that the same strength parameters as those obtained in the previous step for the neighboring odd-AA nuclei are employed in the IBFFM-2 calculations for odd-odd nuclei. On the other hand, the quasiparticle energies and occupation probabilities of the odd particles are independently computed.

Figure 1 shows the neutron and proton spherical single-particle energies (ϵjν\epsilon_{j_{\nu}} and ϵjπ\epsilon_{j_{\pi}}), resulting from the Gogny-HFB calculations, and the quasiparticle energies (ϵ~jν\tilde{\epsilon}_{j_{\nu}} and ϵ~jπ\tilde{\epsilon}_{j_{\pi}}) and occupation probabilities (vjν2v^{2}_{j_{\nu}} and vjπ2v^{2}_{j_{\pi}}) used in the IBFM-2 and IBFFM-2 calculations.

II.3 Electromagnetic transition operators

Theories with effective degrees of freedom, like the IBFFM, require the definition of transition operators to be used in the evaluation of electromagnetic transition probabilities. For the electric E2E2 transition the operator T^(E2)\hat{T}^{(E2)} to be used in the IBFFM-2 takes the form [31]

T^(E2)=T^B(E2)+T^F(E2),\displaystyle\hat{T}^{(E2)}=\hat{T}^{(E2)}_{\text{B}}+\hat{T}^{(E2)}_{\text{F}}\;, (9)

where the first and second terms are the boson and fermion parts, respectively. They are given by

T^B(E2)=ρ=ν,πeρBQ^ρ,\displaystyle\hat{T}^{(E2)}_{\mathrm{B}}=\sum_{\rho=\nu,\pi}e_{\rho}^{\mathrm{B}}\hat{Q}_{\rho}\;, (10)

and

T^F(E2)=15\displaystyle\hat{T}^{(E2)}_{\mathrm{F}}=-\frac{1}{\sqrt{5}} ρ=ν,πjρjρ(ujρujρvjρvjρ)\displaystyle\sum_{\rho=\nu,\pi}\sum_{j_{\rho}j_{\rho}^{\prime}}(u_{j_{\rho}}u_{j_{\rho}^{\prime}}-v_{j_{\rho}}v_{j_{\rho}^{\prime}})
×ρ12jρeρFr2Y(2)ρ12jρ(ajρ×a~jρ)(2).\displaystyle\times\left\langle\ell_{\rho}\frac{1}{2}j_{\rho}\bigg{\|}e^{\mathrm{F}}_{\rho}r^{2}Y^{(2)}\bigg{\|}\ell_{\rho}^{\prime}\frac{1}{2}j_{\rho}^{\prime}\right\rangle(a_{j_{\rho}}^{\dagger}\times\tilde{a}_{j_{\rho}^{\prime}})^{(2)}\;. (11)

The fixed values eνB=eπB=0.1e^{\mathrm{B}}_{\nu}=e^{\mathrm{B}}_{\pi}=0.1 eeb for the boson effective charges are taken so that the experimental B(E2;21+01+)B(E2;2^{+}_{1}\rightarrow 0^{+}_{1}) transition probabilities are reproduced for even-even Pd isotopes. The standard neutron and proton effective charges eνF=0.5e^{\mathrm{F}}_{\nu}=0.5 eeb eπF=1.5e^{\mathrm{F}}_{\pi}=1.5 eeb are employed for all the studied odd-nucleon systems. The M1M1 transition operator T^(M1)\hat{T}^{(M1)} is defined as

T^(M1)=34π\displaystyle\hat{T}^{(M1)}=\sqrt{\frac{3}{4\pi}} ρ=ν,π[gρBL^ρ13jρjρ(ujρujρ+vjρvjρ)\displaystyle\sum_{\rho=\nu,\pi}\Biggl{[}g_{\rho}^{\mathrm{B}}\hat{L}_{\rho}-\frac{1}{\sqrt{3}}\sum_{j_{\rho}j_{\rho}^{\prime}}(u_{j_{\rho}}u_{j_{\rho}^{\prime}}+v_{j_{\rho}}v_{j_{\rho}^{\prime}})
×jρglρ𝐥+gsρ𝐬jρ(ajρ×a~jρ)(1)].\displaystyle\times\left\langle j_{\rho}\|g_{l}^{\rho}{\bf l}+g_{s}^{\rho}{\bf s}\|j_{\rho}^{\prime}\right\rangle(a_{j_{\rho}}^{\dagger}\times\tilde{a}_{j_{\rho}^{\prime}})^{(1)}\Biggr{]}. (12)

The empirical gg factors gνB=0μNg_{\nu}^{\mathrm{B}}=0\,\mu_{N} (nuclear magneton) and gπB=1.0μNg_{\pi}^{\mathrm{B}}=1.0\,\mu_{N}, are adopted for the neutron and proton bosons. For the neutron (proton) gg factors, the standard Schmidt values glν=0μNg_{l}^{\nu}=0\,\mu_{N} and gsν=3.82μNg_{s}^{\nu}=-3.82\,\mu_{N} (glπ=1.0μNg_{l}^{\pi}=1.0\,\mu_{N} and gsπ=5.58μNg_{s}^{\pi}=5.58\,\mu_{N}) are used, with gsρg_{s}^{\rho} quenched by 30% with respect to the free value.

II.4 Gamow-Teller and Fermi transition operators

As in the electromagnetic case, the transition operators for allowed β\beta decay have to be redefined in terms of the relevant degrees of freedom of the model. The Gamow-Teller T^GT\hat{T}^{\mathrm{GT}} and Fermi T^F\hat{T}^{\mathrm{F}} transition operators take the form

T^GT=jνjπηjνjπGT(P^jν×P^jπ)(1),\displaystyle\hat{T}^{\rm GT}=\sum_{j_{\nu}j_{\pi}}\eta_{j_{\nu}j_{\pi}}^{\mathrm{GT}}\left(\hat{P}_{j_{\nu}}\times\hat{P}_{j_{\pi}}\right)^{(1)}, (13)
T^F=jνjπηjνjπF(P^jν×P^jπ)(0),\displaystyle\hat{T}^{\rm F}=\sum_{j_{\nu}j_{\pi}}\eta_{j_{\nu}j_{\pi}}^{\mathrm{F}}\left(\hat{P}_{j_{\nu}}\times\hat{P}_{j_{\pi}}\right)^{(0)}, (14)

with the coefficients

ηjνjπGT\displaystyle\eta_{j_{\nu}j_{\pi}}^{\mathrm{GT}} =13ν12jν𝝈π12jπδνπ,\displaystyle=-\frac{1}{\sqrt{3}}\left\langle\ell_{\nu}\frac{1}{2}j_{\nu}\bigg{\|}{\bm{\sigma}}\bigg{\|}\ell_{\pi}\frac{1}{2}j_{\pi}\right\rangle\delta_{\ell_{\nu}\ell_{\pi}}, (15)
ηjνjπF\displaystyle\eta_{j_{\nu}j_{\pi}}^{\mathrm{F}} =2jν+1δjνjπ.\displaystyle=-\sqrt{2j_{\nu}+1}\delta_{j_{\nu}j_{\pi}}. (16)

In Eqs. (13) and (14), P^jρ\hat{P}_{j_{\rho}} represents one of the one-particle creation operators

Ajρmρ=ζjρajρmρ+jρζjρjρsρ(d~ρ×ajρ)mρ(jρ)\displaystyle A^{\dagger}_{j_{\rho}m_{\rho}}=\zeta_{j_{\rho}}a_{{j_{\rho}}m_{\rho}}^{\dagger}+\sum_{j_{\rho}^{\prime}}\zeta_{j_{\rho}j_{\rho}^{\prime}}s^{\dagger}_{\rho}(\tilde{d}_{\rho}\times a_{j_{\rho}^{\prime}}^{\dagger})^{(j_{\rho})}_{m_{\rho}} (17a)
Bjρmρ=θjρsρa~jρmρ+jρθjρjρ(dρ×a~jρ)mρ(jρ),\displaystyle B^{\dagger}_{j_{\rho}m_{\rho}}=\theta_{j_{\rho}}s^{\dagger}_{\rho}\tilde{a}_{j_{\rho}m_{\rho}}+\sum_{j_{\rho}^{\prime}}\theta_{j_{\rho}j_{\rho}^{\prime}}(d^{\dagger}_{\rho}\times\tilde{a}_{j_{\rho}^{\prime}})^{(j_{\rho})}_{m_{\rho}}, (17b)
and the annihilation operators
A~jρmρ=(1)jρmρAjρmρ\displaystyle\tilde{A}_{j_{\rho}m_{\rho}}=(-1)^{j_{\rho}-m_{\rho}}A_{j_{\rho}-m_{\rho}} (17c)
B~jρmρ=(1)jρmρBjρmρ.\displaystyle\tilde{B}_{j_{\rho}m_{\rho}}=(-1)^{j_{\rho}-m_{\rho}}B_{j_{\rho}-m_{\rho}}. (17d)

The operators in Eqs. (17a) and (17c) conserve the boson number, whereas those in Eqs. (17b) and (17d) do not. The operators T^GT\hat{T}^{\rm GT} and T^F\hat{T}^{\rm F} are expressed as a combination of two of the operators in Eqs. (17a)-(17d), depending on the type of the β\beta decay studied (i.e., β+\beta^{+} or β\beta^{-}) and on the particle or hole nature of the valence nucleons. In the present case,

P^jν={B~jν,mν(N66)A~jν,mν(N68)\displaystyle\hat{P}_{j_{\nu}}=\left\{\begin{array}[]{cc}\tilde{B}_{j_{\nu},m_{\nu}}&(N\leqslant 66)\\ \tilde{A}^{\dagger}_{j_{\nu},m_{\nu}}&(N\geqslant 68)\\ \end{array}\right. (20)

for the β\beta^{-} decay of the odd-AA Rh, while

P^jν={A~jν,mν(N65)B~jν,mν(N67)\displaystyle\hat{P}_{j_{\nu}}=\left\{\begin{array}[]{cr}\tilde{A}_{j_{\nu},m_{\nu}}&(N\leqslant 65)\\ \tilde{B}^{\dagger}_{j_{\nu},m_{\nu}}&(N\geqslant 67)\\ \end{array}\right. (23)

for the β\beta^{-} decay of the even-AA Rh. On the other hand, P^jπ=A~jπ,mπ\hat{P}_{j_{\pi}}=\tilde{A}_{j_{\pi},m_{\pi}} for all the considered β\beta^{-} decays. Note, that Eqs. (17a)–(17d) are simplified forms of the most general one-particle transfer operators in the IBFM-2 [31].

By using the generalized seniority scheme, the coefficients ζj\zeta_{j}, ζjj\zeta_{jj^{\prime}}, θj\theta_{j}, and θjj\theta_{jj^{\prime}} in Eqs. (17a) and (17b) can be written as [68]

ζjρ\displaystyle\zeta_{j_{\rho}} =ujρ1Kjρ,\displaystyle=u_{j_{\rho}}\frac{1}{K_{j_{\rho}}^{\prime}}, (24a)
ζjρjρ\displaystyle\zeta_{j_{\rho}j_{\rho}^{\prime}} =vjρβjρjρ10Nρ(2jρ+1)1KKjρ,\displaystyle=-v_{j_{\rho}}\beta_{j_{\rho}^{\prime}j_{\rho}}\sqrt{\frac{10}{N_{\rho}(2j_{\rho}+1)}}\frac{1}{KK_{j_{\rho}}^{\prime}}, (24b)
θjρ\displaystyle\theta_{j_{\rho}} =vjρNρ1Kjρ′′,\displaystyle=\frac{v_{j_{\rho}}}{\sqrt{N_{\rho}}}\frac{1}{K_{j_{\rho}}^{\prime\prime}}, (24c)
θjρjρ\displaystyle\theta_{j_{\rho}j_{\rho}^{\prime}} =ujρβjρjρ102jρ+11KKjρ′′.\displaystyle=u_{j_{\rho}}\beta_{j_{\rho}^{\prime}j_{\rho}}\sqrt{\frac{10}{2j_{\rho}+1}}\frac{1}{KK_{j_{\rho}}^{\prime\prime}}. (24d)

The factors KK, KjρK_{j_{\rho}}^{\prime}, and Kjρ′′K_{j_{\rho}}^{\prime\prime} are defined as

K=(jρjρβjρjρ2)1/2,\displaystyle K=\left(\sum_{j_{\rho}j_{\rho}^{\prime}}\beta_{j_{\rho}j_{\rho}^{\prime}}^{2}\right)^{1/2}, (25a)
Kjρ=[1+2(vjρujρ)2(n^sρ+1)n^dρ01+Nρ(2jρ+1)jρβjρjρ2K2]1/2,\displaystyle K_{j_{\rho}}^{\prime}=\left[1+2\left(\frac{v_{j_{\rho}}}{u_{j_{\rho}}}\right)^{2}\frac{\braket{(\hat{n}_{s_{\rho}}+1)\hat{n}_{d_{\rho}}}_{0^{+}_{1}}}{N_{\rho}(2j_{\rho}+1)}\frac{\sum_{j_{\rho}^{\prime}}\beta_{j_{\rho}^{\prime}j_{\rho}}^{2}}{K^{2}}\right]^{1/2}, (25b)
Kjρ′′=[n^sρ01+Nρ+2(ujρvjρ)2n^dρ01+2jρ+1jρβjρjρ2K2]1/2,\displaystyle K_{j_{\rho}}^{\prime\prime}=\left[\frac{\braket{\hat{n}_{s_{\rho}}}_{0^{+}_{1}}}{N_{\rho}}+2\left(\frac{u_{j_{\rho}}}{v_{j_{\rho}}}\right)^{2}\frac{\braket{\hat{n}_{d_{\rho}}}_{0^{+}_{1}}}{2j_{\rho}+1}\frac{\sum_{j_{\rho}^{\prime}}\beta_{j_{\rho}^{\prime}j_{\rho}}^{2}}{K^{2}}\right]^{1/2}, (25c)

where n^sρ\hat{n}_{s_{\rho}} is the number operator for the sρs_{\rho} boson and 01+\braket{\cdots}_{0^{+}_{1}} stands for the expectation value of a given operator in the 01+0^{+}_{1} ground state of the even-even nucleus. The amplitudes vjρv_{j_{\rho}} and ujρu_{j_{\rho}} appearing in Eqs. (24a)-(24d) and (25a)-(25c) are the same as those used in the IBFM-2 (or IBFFM-2) calculations for the odd-mass (or odd-odd) nuclei. No additional parameter is introduced for the GT and Fermi operators. For a more detailed account on β\beta-decay operators within the IBFM-2 or IBFFM-2 framework, the reader is also referred to Refs. [68, 6, 31].

The β\beta-decay ftft values are given by

ft=K|M(F)|2+(gAgV)2|M(GT)|2,\displaystyle ft=\frac{K}{|M({\mathrm{F}})|^{2}+\left(\frac{g_{\mathrm{A}}}{g_{\mathrm{V}}}\right)^{2}|M({\mathrm{GT}})|^{2}}, (26)

where the numeric constant KK takes the value K=6163K=6163 s. The quantities M(F)M({\mathrm{F}}) and M(GT)M({\mathrm{GT}}) are the reduced matrix elements of the operators T^F\hat{T}^{\mathrm{F}} of Eq. (14) and T^GT\hat{T}^{\mathrm{GT}} of Eq. (13), respectively. Here gVg_{\mathrm{V}} and gAg_{\mathrm{A}} are the vector and axial-vector coupling constants, respectively. In this study, we use the free nucleon values, gV=1g_{\mathrm{V}}=1 and gA=1.27g_{\mathrm{A}}=1.27, for the β\beta decays of both even- and odd-AA Rh.

Refer to caption
Refer to caption
Figure 2: The Gogny-D1M HFB and mapped IBM-2 potential energy surfaces as functions of the (β,γ)(\beta,\gamma) deformation parameters for the even-even 104-124Pd nuclei. The energy difference between neighboring contours is 200 keV. The global minimum is identified by a solid circle.

III Even-even nuclei

III.1 Potential energy surfaces

The Gogny-D1M HFB and mapped IBM-2 potential energy surfaces are shown in Fig. 2 as functions of the (β,γ)(\beta,\gamma) deformation parameters for the even-even 104-124Pd nuclei. The variation of the HFB potential energy surfaces as functions of the neutron number suggests a transition from prolate (for N62N\lesssim 62) to γ\gamma-soft (64N7064\lesssim N\lesssim 70), and to nearly spherical (N72N\gtrsim 72) shapes. In particular, both 112,114Pd exhibit rather flat potential energy surfaces along the γ\gamma direction. This is what is expected in the γ\gamma-unstable O(6) limit of the IBM [29]. In the case of 116Pd, a flat-bottomed potential with a weak γ\gamma dependence, characteristic of the E(5) critical-point symmetry [69], is obtained.

For each of the considered nuclei, the Gogny-HFB and IBM-2 energy surfaces display a similar topology in the neighborhood of the global minimum (the location of the minimum, and the softness in the β\beta and γ\gamma directions are similar). However, the mapped IBM-2 surfaces generally become flat at large β\beta deformation (β0.4\beta\gtrsim 0.4). This difference is a consequence of the fact that in the HFB approach all nucleonic degrees of freedom are taken into account while the IBM-2 is built on the more limited model (valence) space of nucleon pairs. However, since the mean-field configurations most relevant to the low-energy collective excitations are those in the vicinity of the global minimum, the mapping is considered specifically in that region [63, 64].

Refer to caption
Figure 3: Excitation energies of the (a) 21+2^{+}_{1}, (b) 41+4^{+}_{1}, (c) 02+0^{+}_{2}, and (d) 22+2^{+}_{2} (d) states in the even-even 104-124Pd nuclei. Results are obtained within the mapped and phenomenological (phen.) IBM-2. Experimental data are taken from Ref. [70].
Refer to caption
Figure 4: Parameters for the even-even boson-core Hamiltonian (2) employed in the mapped and phenomenological (phen.) IBM-2 calculations for even-even Pd isotopes.
Refer to caption
Figure 5: The reduced transition probabilities B(E2)B(E2) for the transitions (a) 21+01+2^{+}_{1}\to 0^{+}_{1}, (b) 41+21+4^{+}_{1}\to 2^{+}_{1}, (c) 02+21+0^{+}_{2}\to 2^{+}_{1}, and (d) 22+21+2^{+}_{2}\to 2^{+}_{1} in even-even Pd isotopes in comparison with the experimental data [70]

III.2 Spectroscopic properties

The mapped IBM-2 excitation energies of the 21+2^{+}_{1}, 41+4^{+}_{1}, 02+0^{+}_{2}, and 22+2^{+}_{2} states in the even-even 104-124Pd nuclei are shown in Fig. 3 as functions of the neutron number NN. Results obtained using the conventional IBM-2 approach (hereinafter referred to as phenomenological IBM-2), with parameters adopted from the earlier phenomenological study [37], are also included in the plot. As can be seen from the figure, the excitation energies decrease toward the middle of the major shell, i.e., N=66N=66. For N64N\leqslant 64, the mapped IBM-2 21+2^{+}_{1} and 41+4^{+}_{1} excitation energies underestimate the experimental ones while the energies of the non-yrast 02+0^{+}_{2} and 22+2^{+}_{2} states are overestimated. In the mapped (phenomenological) IBM-2 approach the ratios R4/2R_{4/2} of the 41+4^{+}_{1} to 21+2^{+}_{1} excitation energies are 2.96 (2.43), 2.86 (2.39), and 2.69 (2.34) for 104Pd, 106Pd, and 108Pd, respectively. These values should be compared with the experimental ratios of 2.38, 2.40, and 2.41. Thus, the mapped IBM-2 provides excitation spectra which are more rotational in character than the phenomenological IBM-2 and experimental ones. Around the neutron midshell N=66N=66, both the predicted and experimental 22+2^{+}_{2} levels have the lowest energies, being even below the 41+4^{+}_{1} state. The 22+2^{+}_{2} state is the bandhead of the quasi-γ\gamma band, and the lowering of this state reflects an emergence of pronounced γ\gamma softness.

The IBM-2 parameters obtained for the even-even Pd isotopes from the mapping procedure, and those determined phenomenologically are shown in Fig. 4. The phenomenological IBM-2 parameters are extracted from earlier fitting calculations for Pd and Ru isotopes [37]. In Ref. [37], in addition to the terms that appear in Eq. (2), the like-boson interactions, and the so-called Majorana terms were included in the model Hamiltonian. These terms were, however, shown to play a minor role [37], and are omitted in the present study. From Fig. 4, one sees that the single-dd boson energy ϵd\epsilon_{d} and the strength κ\kappa have similar nucleon-number dependence for both the mapped and phenomenological IBM-2 models. A notable quantitative difference is that the derived κ\kappa values for the former are \approx 1.4 larger in magnitude than for the latter. The behavior of the parameter χν\chi_{\nu} is different in the two approaches for N70N\geqslant 70. The sign and absolute value of the sum χν+χπ\chi_{\nu}+\chi_{\pi} reflect the extent of γ\gamma softness and whether the nucleus is prolate or oblate deformed. In both calculations, the sum is negative, χν+χπ<0\chi_{\nu}+\chi_{\pi}<0, for N64N\lesssim 64, indicating prolate deformation, and takes nearly vanishing values, χν+χπ0\chi_{\nu}+\chi_{\pi}\approx 0, around the neutron midshell N=66N=66, reflecting γ\gamma softness. However, for N70N\geqslant 70, the sum is negative (positive) in the mapped (phenomenological) calculations, implying prolate (oblate) deformation. Note that a fixed value χπ=0.2\chi_{\pi}=0.2 is employed in the phenomenological IBM-2 calculations, whereas in the mapped approach this parameter exhibits a strong nucleon number dependence.

The B(E2)B(E2) transition probabilities, computed within the mapped and phenomenological IBM-2 models, are plotted in Fig. 5 as functions of the neutron number NN. The same E2E2 effective boson charge is used for the quadrupole operators in the two sets of the IBM-2 calculations. The B(E2;21+01+)B(E2;2^{+}_{1}\to 0^{+}_{1}) and B(E2;41+21+)B(E2;4^{+}_{1}\to 2^{+}_{1}) values obtained in the mapped IBM-2 calculations agree reasonably well with the experiment, exception made of 112Pd. Both the mapped and phenomenological IBM-2 calculations predict B(E2;02+21+)B(E2;0^{+}_{2}\to 2^{+}_{1}) and B(E2;22+21+)B(E2;2^{+}_{2}\to 2^{+}_{1}) rates with similar trends as functions of NN. However, the mapped IBM-2 scheme provides smaller B(E2;02+21+)B(E2;0^{+}_{2}\to 2^{+}_{1}) values for Pd isotopes with 58N6258\leqslant N\leqslant 62. The enhancement of the predicted B(E2;22+21+)B(E2;2^{+}_{2}\to 2^{+}_{1}) transition rates around the midshell N=66N=66 [see Fig. 5(d)] can be considered as another signature of γ\gamma soft deformation.

Refer to caption
Figure 6: Excitation energies of the low-lying positive-parity states obtained for odd-AA Pd isotopes within the IBFM-2 with the boson-core Hamiltonian determined by (a) mapping the Gogny-D1M EDF, and (b) using the phenomenological fit. The experimental data included in panel (c) are taken from Ref. [71] for 117Pd, from Ref. [72] for 119Pd, and from the NNDC database [70] for the other nuclei.
Refer to caption
Figure 7: The same as Fig. 6, but for odd-ZZ Rh nuclei. The experimental data are taken from Ref. [70].
Refer to caption
Figure 8: Parameters of the mapped and phenomenological IBFM-2 Hamiltonian (2) for odd-NN Pd nuclei.
Refer to caption
Figure 9: Fractions (in percent %) of the neutron ν2s1/2\nu 2s_{1/2}, ν1d3/2\nu 1d_{3/2}, ν1d5/2\nu 1d_{5/2}, and ν0g7/2\nu 0g_{7/2} single-particle configurations in the wave functions for the (a) 1/21+{1/2}^{+}_{1}, (b) 3/21+{3/2}^{+}_{1}, (c) 5/21+{5/2}^{+}_{1}, and (d) 7/21+{7/2}^{+}_{1} states in odd-AA Pd nuclei. The wave functions are obtained within the mapped IBFM-2 scheme based on Gogny-D1M EDF calculations.

IV Odd-AA Pd and Rh nuclei

The excitation energies of the low-lying positive-parity states obtained for the odd-AA Pd isotopes 105-123Pd are depicted in Fig. 6. The results obtained within the IBFM-2 model with boson-core Hamiltonian determined by mapping the Gogny-D1M EDF [Fig. 6(a)] and the those obtained from phenomenological calculations of Ref. [37] [Fig. 6(b)] are compared with experimental data [71, 72, 70]. The two IBFM-2 calculations, using different boson-core Hamiltonian parameters, provide an overall consistent description of the experimental excitation energies. As can be seen from the figure, the experimental data display a change in the ground state spin from N=67N=67 to 69. The corresponding even-even core nuclei, 114Pd and 116Pd, are in the transitional region, for which the potential energy surfaces are suggested to be considerably γ\gamma soft (see Fig. 2). The sudden change in the ground-state spin of the odd-AA neighbor, therefore, reflects the transition that takes place in the even-even core systems from the γ\gamma unstable shape, which is associated with an O(6)-like potential, to the E(5)-like structure characterized by a flat-bottomed potential.

The excitation energies of the low-lying positive-parity states obtained for the odd-AA isotopes 103-123Rh are depicted in Fig. 7. Experimentally, the ground states of these isotopes have spin Iπ=7/2+I^{\pi}={7/2}^{+}. Exceptions are made of some of the heaviest isotopes, and similar results are predicted within both the mapped and phenomenological calculations. Both theoretically and experimentally, some of the energy levels exhibit an approximate parabolic behavior with a minimum around the middle of the major shell, N66N\approx 66. For 103-123Rh, the order of most of the energy levels remains unchanged in the whole isotopic chain within both the mapped and phenomenological IBFM-2 calculations. This situation is in a sharp contrast with the one in the odd-AA Pd (see Fig. 6), in which the structural change along the isotopic chain occurs more rapidly. Note that the low-lying states of the odd-AA Rh nuclei are accounted for almost purely by the proton π0g9/2\pi 0g_{9/2} single-particle configuration while more than one single-particle orbital is considered for the odd-AA Pd. The occupation number of the odd proton in the π0g9/2\pi 0g_{9/2} orbital is also nearly constant along the whole Rh isotopic chain [see Fig. 1(f)], whereas the occupation probabilities for the odd neutron in the odd-AA Pd vary significantly with NN [see Fig. 1(e)]. Furthermore, as shown below, the strength parameters for V^BF\hat{V}_{\mathrm{BF}} are fixed in the case of odd-AA Rh nuclei while they depend on the boson number for odd-AA Pd isotopes.

The strength parameters of the boson-fermion interaction (4) for odd-NN Pd nuclei are shown in Fig. 8. These parameters are chosen so that the ground-state spin, and energies of a few low-lying levels are reproduced reasonably well. The parameters for the two IBFM-2 calculations are rather similar, with an exception made of the monopole strength AνA_{\nu} for 59N6359\leqslant N\leqslant 63. Note that common quasiparticle energies ϵ~jρ\tilde{\epsilon}_{j_{\rho}} and occupation probabilities vjρ2v^{2}_{j_{\rho}} are used for both IBFM-2 calculations. The parameters for the 123Pd77 nucleus, where no experimental data are available, are taken to be the same as those for the adjacent nucleus 121Pd75. As can be seen from the figure, the IBFM-2 parameters turn out to have a strong NN-dependence that reflects the rapid structural change in the odd-AA Pd isotopes. On the other hand, constant strength parameters Γπ=0.6\Gamma_{\pi}=0.6 (0.0) MeV, Λπ=0.6\Lambda_{\pi}=0.6 (0.75) MeV, and Aπ=0.0A_{\pi}=0.0 (0.25-0.25) MeV reproduce reasonably well the experimental data for odd-AA Rh nuclei in the mapped (phenomenological) calculations.

Table 2: B(E2)B(E2) rates (in Weisskopf units, W.u.), quadrupole moment Q(I)Q(I) (in eeb), B(M1)B(M1) rates (in W.u. ×103\times 10^{-3}), and magnetic dipole moments μ(I)\mu(I) (in μN\mu_{N}) obtained for odd-AA Pd nuclei within the mapped and phenomenological IBFM-2 calculations. Experimental data are taken from Ref. [70, 73].
Calc.
mapped phen. Expt.
105Pd B(E2;1/21+3/21+)B(E2;1/2^{+}_{1}\to 3/2^{+}_{1}) 2525 1313 2.016+912.0^{+91}_{-16}
B(E2;1/21+5/21+)B(E2;1/2^{+}_{1}\to 5/2^{+}_{1}) 9090 4545 2.64(15)2.64(15)
B(E2;1/22+3/21+)B(E2;1/2^{+}_{2}\to 3/2^{+}_{1}) 0.60.6 3.13.1 0.97+120.9^{+12}_{-7}
B(E2;1/22+5/21+)B(E2;1/2^{+}_{2}\to 5/2^{+}_{1}) 0.040.04 0.90.9 8.4(9)8.4(9)
B(E2;3/21+5/21+)B(E2;3/2^{+}_{1}\to 5/2^{+}_{1}) 4444 4040 4.6(7)4.6(7)
B(E2;3/21+3/23+)B(E2;3/2^{+}_{1}\to 3/2^{+}_{3}) 0.050.05 5.15.1 >0.21>0.21
B(E2;3/23+5/22+)B(E2;3/2^{+}_{3}\to 5/2^{+}_{2}) 0.010.01 2.72.7 >2.2>2.2
B(E2;5/21+5/22+)B(E2;5/2^{+}_{1}\to 5/2^{+}_{2}) 1515 2929 1.8(4)1.8(4)
B(E2;7/21+5/21+)B(E2;7/2^{+}_{1}\to 5/2^{+}_{1}) 2424 3333 0.30(4)0.30(4)
B(E2;9/21+5/21+)B(E2;9/2^{+}_{1}\to 5/2^{+}_{1}) 5757 4040 14.3(13)14.3(13)
B(M1;1/21+3/21+)B(M1;1/2^{+}_{1}\to 3/2^{+}_{1}) 372372 280280 14.921+2014.9^{+20}_{-21}
B(M1;1/21+1/22+)B(M1;1/2^{+}_{1}\to 1/2^{+}_{2}) 0.930.93 1.51.5 7.8(8)7.8(8)
B(M1;1/22+3/21+)B(M1;1/2^{+}_{2}\to 3/2^{+}_{1}) 3737 77 455+645^{+6}_{-5}
B(M1;3/21+5/21+)B(M1;3/2^{+}_{1}\to 5/2^{+}_{1}) 3131 4.44.4 20.3(22)20.3(22)
B(M1;3/21+3/23+)B(M1;3/2^{+}_{1}\to 3/2^{+}_{3}) 0.0120.012 0.00040.0004 >5.9>5.9
B(M1;3/23+5/22+)B(M1;3/2^{+}_{3}\to 5/2^{+}_{2}) 0.00260.0026 2.92.9 >47>47
B(M1;5/21+5/22+)B(M1;5/2^{+}_{1}\to 5/2^{+}_{2}) 1313 1.61.6 19(3)19(3)
B(M1;5/23+3/21+)B(M1;5/2^{+}_{3}\to 3/2^{+}_{1}) 4.74.7 0.470.47 >0.40>0.40
B(M1;5/23+7/22+)B(M1;5/2^{+}_{3}\to 7/2^{+}_{2}) 5252 3232 >25>25
B(M1;7/21+5/21+)B(M1;7/2^{+}_{1}\to 5/2^{+}_{1}) 3131 3.73.7 10.6(12)10.6(12)
Q(5/21+)Q(5/2^{+}_{1}) 0.54-0.54 0.27-0.27 +0.660(11)+0.660(11)
μ(3/21+)\mu(3/2^{+}_{1}) 0.56-0.56 0.64-0.64 0.074(13)-0.074(13)
μ(5/21+)\mu(5/2^{+}_{1}) 1.19-1.19 1.32-1.32 0.642(3)-0.642(3)
μ(5/22+)\mu(5/2^{+}_{2}) 0.67-0.67 0.76-0.76 +0.95(20)+0.95(20)
107Pd B(E2;1/21+5/21+)B(E2;1/2^{+}_{1}\to 5/2^{+}_{1}) 112112 9090 0.58(7)0.58(7)
μ(5/21+)\mu(5/2^{+}_{1}) 1.06-1.06 1.05-1.05 0.735(7)0.735(7)
109Pd B(E2;1/21+5/21+)B(E2;1/2^{+}_{1}\to 5/2^{+}_{1}) 9797 7676 1.36(18)1.36(18)
B(E2;3/21+5/21+)B(E2;3/2^{+}_{1}\to 5/2^{+}_{1}) 5858 4848 8(8)8(8)
B(M1;3/21+5/21+)B(M1;3/2^{+}_{1}\to 5/2^{+}_{1}) 4.44.4 4.44.4 2.2(8)2.2(8)
B(M1;5/22+3/21+)B(M1;5/2^{+}_{2}\to 3/2^{+}_{1}) 159159 142142 11.7(19)11.7(19)
B(M1;7/22+5/21+)B(M1;7/2^{+}_{2}\to 5/2^{+}_{1}) 3.23.2 0.130.13 3.6(4)3.6(4)
Table 3: The same as in Table 2, but for odd-AA Rh nuclei.
Calc.
mapped phen. Expt.
103Rh B(E2;5/21+7/21+)B(E2;5/2^{+}_{1}\to 7/2^{+}_{1}) 3333 3131 2.0(6)2.0(6)
B(E2;5/21+9/21+)B(E2;5/2^{+}_{1}\to 9/2^{+}_{1}) 2828 1313 0.107(33)0.107(33)
B(M1;5/21+7/21+)B(M1;5/2^{+}_{1}\to 7/2^{+}_{1}) 471471 354354 40(12)40(12)
B(M1;9/21+7/21+)B(M1;9/2^{+}_{1}\to 7/2^{+}_{1}) 1.01.0 1.91.9 43(12)43(12)
μ(7/21+)\mu(7/2^{+}_{1}) 4.854.85 4.884.88 +4.540(11)+4.540(11)
μ(9/21+)\mu(9/2^{+}_{1}) 5.695.69 5.625.62 +4.9(8)+4.9(8)
107Rh B(E2;3/21+7/21+)B(E2;3/2^{+}_{1}\to 7/2^{+}_{1}) 4.734.73 1.621.62 0.16(2)0.16(2)
109Rh B(E2;3/21+3/22+)B(E2;3/2^{+}_{1}\to 3/2^{+}_{2}) 0.140.14 0.180.18 1.7×102(5)1.7\times{10}^{2}(5)
B(E2;3/21+7/21+)B(E2;3/2^{+}_{1}\to 7/2^{+}_{1}) 4.414.41 0.010.01 0.0174(5)0.0174(5)
B(E2;3/22+7/21+)B(E2;3/2^{+}_{2}\to 7/2^{+}_{1}) 5.35.3 5.95.9 26.1(19)26.1(19)
B(E2;5/21+9/21+)B(E2;5/2^{+}_{1}\to 9/2^{+}_{1}) 7.97.9 5.35.3 >23>23
B(E2;5/22+3/21+)B(E2;5/2^{+}_{2}\to 3/2^{+}_{1}) 1212 5.85.8 1.7(7)1.7(7)
B(E2;5/22+3/22+)B(E2;5/2^{+}_{2}\to 3/2^{+}_{2}) 2222 99 7.E+1(3)7.E+1(3)
B(E2;7/22+3/21+)B(E2;7/2^{+}_{2}\to 3/2^{+}_{1}) 88 1515 131(12)131(12)
B(M1;5/21+3/23+)B(M1;5/2^{+}_{1}\to 3/2^{+}_{3}) 5.25.2 8.68.6 >220>220
B(M1;5/21+3/21+)B(M1;5/2^{+}_{1}\to 3/2^{+}_{1}) 818818 414414 >0.40>0.40
B(M1;5/22+3/21+)B(M1;5/2^{+}_{2}\to 3/2^{+}_{1}) 3737 289289 2.4(3)2.4(3)
B(M1;5/22+3/22+)B(M1;5/2^{+}_{2}\to 3/2^{+}_{2}) 152152 207207 2.2(15)2.2(15)
B(M1;5/22+3/23+)B(M1;5/2^{+}_{2}\to 3/2^{+}_{3}) 1818 210210 2.5(4)2.5(4)
B(M1;5/22+7/21+)B(M1;5/2^{+}_{2}\to 7/2^{+}_{1}) 231231 112112 4.1×102(6)4.1{\times}10^{-2}(6)
B(M1;7/22+9/21+)B(M1;7/2^{+}_{2}\to 9/2^{+}_{1}) 318318 611611 0.25(6)0.25(6)
B(M1;7/21+7/22+)B(M1;7/2^{+}_{1}\to 7/2^{+}_{2}) 7.67.6 8.98.9 6.6×102(8)6.6{\times}10^{-2}(8)
B(M1;3/21+3/22+)B(M1;3/2^{+}_{1}\to 3/2^{+}_{2}) 2727 4848 0.58(12)0.58(12)
B(M1;3/21+3/23+)B(M1;3/2^{+}_{1}\to 3/2^{+}_{3}) 158158 256256 1.18(11)1.18(11)
B(M1;3/22+3/23+)B(M1;3/2^{+}_{2}\to 3/2^{+}_{3}) 276276 3232 0.32(10)0.32(10)
B(M1;5/21+7/21+)B(M1;5/2^{+}_{1}\to 7/2^{+}_{1}) 233233 172172 >3.2>3.2
B(M1;9/21+7/21+)B(M1;9/2^{+}_{1}\to 7/2^{+}_{1}) 4.34.3 1818 >58>58

Experimental data for electromagnetic transitions and moments are available for odd-AA Pd and Rh nuclei with N65N\leqslant 65. The predicted B(E2)B(E2) and B(M1)B(M1) transition strengths as well as the electric quadrupole Q(Iπ)Q(I^{\pi}) and magnetic dipole μ(Iπ)\mu(I^{\pi}) moments for the low-lying positive-parity states in odd-AA Pd are given in Table 2. In most of the cases, the mapped and phenomenological calculations provide similar results. Large values are obtained for the B(E2;1/21+5/21+)B(E2;{1/2}^{+}_{1}\to{5/2}^{+}_{1}) (in 105Pd, 107Pd and 109Pd), B(E2;3/21+5/21+)B(E2;{3/2}^{+}_{1}\to{5/2}^{+}_{1}) (in 105Pd and 109Pd), and B(E2;9/21+5/21+)B(E2;{9/2}^{+}_{1}\to{5/2}^{+}_{1}) (in 105Pd) transitions. The experimental data, however, suggest that these E2E2 transitions are weaker. The B(E2)B(E2) and B(M1)B(M1) rates corresponding to some transitions in odd-AA Rh nuclei are given in Table 3. The large B(E2;5/21+7/21+)B(E2;{5/2}^{+}_{1}\to{7/2}^{+}_{1}) and B(E2;5/21+9/21+)B(E2;{5/2}^{+}_{1}\to{9/2}^{+}_{1}) rates obtained for 103Rh overestimate the experimental rates by several orders of magnitude.

The deviation of the predicted B(E2)B(E2) and B(M1)B(M1) transition rates for odd-AA systems with respect to the experiment could be interpreted in terms of the structure of the corresponding IBFM-2 wave functions. The components of the IBFM-2 wave functions for the low-lying states of odd-AA Pd isotopes are shown in Fig. 9. They are associated with the single(quasi)-particle orbitals ν2s1/2\nu 2s_{1/2}, ν1d3/2\nu 1d_{3/2}, ν1d5/2\nu 1d_{5/2}, and ν0g7/2\nu 0g_{7/2}. Only components obtained within the mapped framework are shown as illustrative examples, while qualitatively similar results are obtained using the phenomenological approach. The states considered for odd-AA Rh nuclei are almost purely made of the proton 0g9/20g_{9/2} configuration (with a weight of 99\approx 99 %). Therefore, the corresponding wave function contents are not shown in the plot. As can be seen from the figure, the neutron 1d5/21d_{5/2} configuration accounts for most of the IBFM-2 wave functions for the 1/21+{1/2}^{+}_{1}, 3/21+{3/2}^{+}_{1}, 5/21+{5/2}^{+}_{1}, and 7/21+{7/2}^{+}_{1} in odd-AA Pd nuclei with N67N\lesssim 67. However, the description of these wave functions in both the mapped and phenomenological IBFM-2 calculations in the present study may not be adequate, and this leads to some of the considerable disagreements between the calculated and experimental electromagnetic properties, including the B(E2;1/21+5/21+)B(E2;{1/2}^{+}_{1}\to{5/2}^{+}_{1}) values in 105Pd, 107Pd, and 109Pd (see Table 2). The deficiency of the IBFM-2 wave functions could arise from various deficiencies of the present model calculations, such as the choice of the single-particle space, the quasiparticle energies and occupation probabilities of the odd particle, and the effective charges involved in the transition operators, which are kept constant for all nuclei. On the other hand, earlier IBFM-2 fitting calculations in the same mass region [74, 7] obtained E2E2 and M1M1 properties consistent with experiment.

Refer to caption
Figure 10: The same as Fig. 6, but for the odd-odd Rh isotopes.
Refer to caption
Figure 11: Parameters for the residual neutron-proton interactions (II.1) employed for odd-odd Rh isotopes in the mapped and phenomenological IBM approaches.

V Odd-odd Rh nuclei

The excitation energies of the low-lying positive-parity states obtained for odd-odd Rh isotopes are depicted in Fig. 10. The available experimental data [70] suggest that for N71N\leqslant 71 the ground state has spin Iπ=1+I^{\pi}=1^{+}. Excited 1+1^{+} states are also observed at low energy. Both the mapped [Fig. 10(a)] and phenomenological [Fig. 10(b)] IBFFM-2 calculations account for the ground-state spin 1+1^{+}. The calculations also reproduce reasonably well the energies of the 12+1^{+}_{2} states. From N71N\approx 71 to 73, both types of calculations suggest a change in the ground-state spin to Iπ=5+I^{\pi}=5^{+}. There are no spectroscopic data to compare with for even-AA Rh isotopes with N73N\geqslant 73. Note, that a ground-state spin different from Iπ=1+I^{\pi}=1^{+} is experimentally found in the neighboring odd-odd Ag and In isotopes. For instance, for 120Ag, 122Ag, 124Ag and 126In the ground state has spin Iπ=3+I^{\pi}=3^{+}. A low-lying 5+5^{+} level is observed in 122In at an excitation energy around 40 keV above the 1+1^{+} ground state.

The strength parameters vdv_{\mathrm{d}} and vtv_{\mathrm{t}} of the neutron-proton residual interaction V^νπ\hat{V}_{\nu\pi} in Eq. (II.1) are shown in Fig. 11 for odd-odd Rh isotopes as functions of the neutron number. Those parameters are determined so that the correct ground-state spin Ig.s.π=11+I^{\pi}_{g.s.}=1^{+}_{1} as well as the energy of the 12+1^{+}_{2} state are reproduced reasonably well. For N73N\geqslant 73, where experimental data are not available, the same values of the parameters as for 116Rh71 are employed. As can be seen from Fig. 11(a), the parameter vdv_{\mathrm{d}} changes suddenly from N=63N=63 to 67. This sudden change accounts for the experimental [see Fig. 10(c)] lowering of the 12+1^{+}_{2} level toward the middle of the major shell, N67N\approx 67. On the other hand, the tensor interaction strength exhibits a smooth decrease with NN.

Refer to caption
Figure 12: Fraction (in percent %) of the neutron-proton pair components in the wave functions for the (a) 11+1^{+}_{1} and (b) 51+5^{+}_{1} states of the odd-odd 104-122Rh isotopes under study. The wave functions are obtained within the mapped IBFFM-2 formalism based on the Gogny-D1M EDF.

The nature of the low-lying states in odd-odd Rh isotopes can be analyzed in terms of various neutron-proton pair components in the IBFFM-2 wave functions. The corresponding results for the 11+1^{+}_{1} and 51+5^{+}_{1} states, obtained within the mapped IBFFM-2 formalism, are shown in Fig. 12. For nuclei with A118A\leqslant 118, the 11+1^{+}_{1} state is mostly based on the configuration associated with the [ν0g7/2π0g9/2](J)[\nu 0g_{7/2}\otimes\pi 0g_{9/2}]^{(J)} neutron-proton pairs coupled to the even-even boson core, with the total angular momentum of the fermion system J=1,2,J=1,2,\ldots, 8. For 120Rh and 122Rh, the contributions of the [ν1d3/2π0g9/2](J)[\nu 1d_{3/2}\otimes\pi 0g_{9/2}]^{(J)} (J=3,4,5,6J=3,4,5,6) pairs also play a prominent role. As one can see from Fig. 12(b), the dominant contribution to the 51+5^{+}_{1} wave function for Rh isotopes with mass A112A\leqslant 112 comes from the [ν1d5/2π0g9/2](J)[\nu 1d_{5/2}\otimes\pi 0g_{9/2}]^{(J)} pair components, while the [ν0g7/2π0g9/2](J)[\nu 0g_{7/2}\otimes\pi 0g_{9/2}]^{(J)} pair components play a negligible role. For heavier Rh isotopes, with A114A\geqslant 114, the other pair components that involve the π0g9/2\pi 0g_{9/2} state, i.e., those based on the [ν2s1/2π0g9/2](J)[\nu 2s_{1/2}\otimes\pi 0g_{9/2}]^{(J)}, [ν1d3/2π0g9/2](J)[\nu 1d_{3/2}\otimes\pi 0g_{9/2}]^{(J)}, and [ν0g7/2π0g9/2](J)[\nu 0g_{7/2}\otimes\pi 0g_{9/2}]^{(J)} pairs, are rather fragmented in the Iπ=51+I^{\pi}=5^{+}_{1} wave functions. Qualitatively similar results are obtained using phenomenological IBFFM-2 wave functions.

Table 4: B(E2)B(E2), B(M1)B(M1) (in W.u.), and magnetic dipole moment μ(11+)\mu(1^{+}_{1}) (in μN\mu_{N}) for odd-odd Rh isotopes, computed within the mapped IBFFM-2 based on the Gogny-D1M EDF and the phenomenological IBFFM-2. Experimental data are taken from Refs. [70, 73].
Calc.
mapped phen. Expt.
104Rh B(E2;13+21+)B(E2;1^{+}_{3}\to 2^{+}_{1}) 1.351.35 1313 >5.2>5.2
B(M1;21+11+)B(M1;2^{+}_{1}\to 1^{+}_{1}) 0.030.03 0.060.06 >0.029>0.029
B(M1;13+11+)B(M1;1^{+}_{3}\to 1^{+}_{1}) 0.030.03 0.050.05 >0.00098>0.00098
106Rh μ(11+)\mu(1^{+}_{1}) 2.132.13 2.202.20 2.575(7)2.575(7)

The experimental information on the electromagnetic properties of the considered odd-odd Rh nuclei is rather limited. Table 4 compares the predicted and experimental B(E2)B(E2), B(M1)B(M1), and magnetic dipole moment μ(11+)\mu(1^{+}_{1}) for 104Rh and 106Rh. Both the mapped and phenomenological IBFFM-2 calculations provide a reasonable description of the experimental data for these odd-odd nuclei. Nevertheless, a more detailed assessment of the quality of the IBFFM-2 wave functions is difficult in this case, due to the lack of data.

Refer to caption
Figure 13: logft\log{ft} values for the β\beta^{-} decays from the odd-AA Rh into Pd nuclei, (a) 7/21+5/21+7/2^{+}_{1}\to 5/2^{+}_{1}, (b) 7/21+7/21+7/2^{+}_{1}\to 7/2^{+}_{1}, (c) 7/21+5/22+7/2^{+}_{1}\to 5/2^{+}_{2}, and (d) 7/21+7/22+7/2^{+}_{1}\to 7/2^{+}_{2} computed using wave functions obtained within the mapped and phenomenological IBFM-2 models. The available experimental data [70] are also included in the plot.
Table 5: logft\log{ft} values for the β\beta^{-} decays from odd-AA Rh into Pd nuclei, computed using wave functions obtained within the mapped IBFM-2 scheme based on the Gogny-D1M EDF and within the phenomenological IBFM-2 model. The experimental data are taken from Ref. [70].
Calc.
Decay IiIfI_{i}\to I_{f} mapped phen. Expt.
105Rh105\to^{105}Pd 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} 7.45 6.88 5.710(7)
7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} 8.12 7.66 5.797(16)
7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} 7.19 7.41 5.152(20)
7/21+7/22+{7/2}^{+}_{1}\to{7/2}^{+}_{2} 9.01 10.08 6.91(3)
107Rh107\to^{107}Pd 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} 6.81 6.47 6.1(2)
7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} 7.81 7.39 5.0(1)
7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} 7.78 6.99 6.2(1)
7/21+7/22+{7/2}^{+}_{1}\to{7/2}^{+}_{2} 6.23 8.05 5.8(1)
7/21+5/23+{7/2}^{+}_{1}\to{5/2}^{+}_{3} 8.00 7.45 6.1(1)
7/21+5/24+{7/2}^{+}_{1}\to{5/2}^{+}_{4} 5.82 7.87 5.3(1)
109Rh109\to^{109}Pd 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} 6.05 5.86 5.8(3)
7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} 7.02 6.19 6.69(12)
7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} 6.92 6.58 4.86(5)
7/21+7/22+{7/2}^{+}_{1}\to{7/2}^{+}_{2} 5.68 5.57 5.69(6)
7/21+5/23+{7/2}^{+}_{1}\to{5/2}^{+}_{3} 6.83 7.39 5.53(5)
7/21+9/21+{7/2}^{+}_{1}\to{9/2}^{+}_{1} 7.32 6.84 7.26(19)
113Rh113\to^{113}Pd 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} 4.58 4.51 5.4(1)
7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} 6.46 8.07 5.90(5)
7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} 4.35 4.28 5.00(4)111(5/2+,7/2+)({5/2}^{+},{7/2}^{+}) at 349 keV [70]
7/21+7/22+{7/2}^{+}_{1}\to{7/2}^{+}_{2} 5.71 5.57 5.00(4)111(5/2+,7/2+)({5/2}^{+},{7/2}^{+}) at 349 keV [70]
7/21+5/23+{7/2}^{+}_{1}\to{5/2}^{+}_{3} 5.42 5.59 6.7(2)222(5/2+,7/2+)({5/2}^{+},{7/2}^{+}) at 373 keV based on the XUNDL datasets [70]
7/21+7/23+{7/2}^{+}_{1}\to{7/2}^{+}_{3} 5.07 4.81 6.7(2)222(5/2+,7/2+)({5/2}^{+},{7/2}^{+}) at 373 keV based on the XUNDL datasets [70]
117Rh117\to^{117}Pd 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} 5.61 5.09 6.0333Uncertainties are not given with the logft\log{ft}.
7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} 5.81 5.31 5.7333Uncertainties are not given with the logft\log{ft}.
7/21+5/23+{7/2}^{+}_{1}\to{5/2}^{+}_{3} 4.27 5.34 5.8333Uncertainties are not given with the logft\log{ft}.
7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} 5.22 5.27 6.3333Uncertainties are not given with the logft\log{ft}.
7/21+5/24+{7/2}^{+}_{1}\to{5/2}^{+}_{4} 7.64 4.56 6.3333Uncertainties are not given with the logft\log{ft}.
7/21+5/25+{7/2}^{+}_{1}\to{5/2}^{+}_{5} 5.82 5.50 6.0444(5/2+,7/2+)({5/2}^{+},{7/2}^{+}) level at 436 keV, based on the XUNDL datasets [70]. Uncertainties are not given.
7/21+7/22+{7/2}^{+}_{1}\to{7/2}^{+}_{2} 4.97 5.28 6.0444(5/2+,7/2+)({5/2}^{+},{7/2}^{+}) level at 436 keV, based on the XUNDL datasets [70]. Uncertainties are not given.

VI β\beta decay

VI.1 β\beta decays between odd-AA nuclei

Figure 13 shows the logft\log{ft} values for the β\beta^{-} decays of the 7/21+{7/2}^{+}_{1} state of the odd-AA Rh into several low-lying states of the odd-AA Pd nuclei. Results are obtained using mapped and phenomenological IBFM-2 wave functions. In both cases, the predicted trend of the logft\log{ft} values, as functions of the nucleon number, reflects the structural change in the parent and daughter odd-AA nuclei. An illustrative example is a kink emerging at the mass A113A\approx 113 or 115 in the predicted logft\log{ft} values for the 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} [Fig. 13(a)] and 7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} [Fig. 13(b)] decays. The mass number at which the kink emerges corresponds to the transitional region, where the ground-state spin changes, observed in the odd-AA Pd daughter (see Fig. 6). The mass dependence of the predicted logft\log{ft} values is similar in the mapped and phenomenological calculations, exception made of the results from AA=113 to 115 in the 7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} decay and from AA=117 to 119 in the 7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} decay.

Both within the mapped and phenomenological schemes, the present calculations overestimate the observed logft\log{ft} values for the decays 105,107Rh(7/21+)105,107({7/2}^{+}_{1})\to^{105,107}Pd(5/21+)({5/2}^{+}_{1}) [Fig. 13(a)]. At both A=A= 105 and 107, the 5/21+{5/2}^{+}_{1} final-state wave function has been shown to be almost purely made of the ν1d5/2\nu 1d_{5/2} configuration [see Fig. 9(c)], while the parent state 7/21+{7/2}^{+}_{1} is of almost pure π0g9/2\pi 0g_{9/2} nature.

The dominant contribution to the GT matrix element for the above 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} decays indeed comes from the term that corresponds to the coupling of the ν1d5/2\nu 1d_{5/2} with π0g9/2\pi 0g_{9/2} single-particle states, which is of the form

[[d~ν×aν1d5/2](7/2)×a~π0g9/2](J=1).\displaystyle[[\tilde{d}_{\nu}\times a_{\nu 1d_{5/2}}^{\dagger}]^{(7/2)}\times\tilde{a}_{\pi 0g_{9/2}}]^{(J=1)}\;. (27)

The matrix element of this term is, however, rather small: 0.041 and 0.091-0.091 (0.069 and 0.118-0.118), for the 105Rh and 107Rh decays in the mapped (phenomenological) approach. There are many other terms similar to the one in Eq. (27), but their matrix elements are small and cancel each other, leading to a small GT transition rate. The same is true for the 105,107Rh(7/21+)105,107({7/2}^{+}_{1})\to^{105,107}Pd(7/21+)({7/2}^{+}_{1}) decays [Fig. 13(b)]. In this case, the Fermi transition matrix is also negligibly small.

The calculations underestimate the logft\log{ft} values for the 113Rh(7/21+)113({7/2}^{+}_{1})\to^{113}Pd(5/21+)({5/2}^{+}_{1}) decay. For this decay, approximately 75 % and 25 % of the wave function of the 5/21+{5/2}^{+}_{1} final state are comprised of the ν1d5/2\nu 1d_{5/2} and ν0g7/2\nu 0g_{7/2} configurations, respectively [see Fig. 9(c)]. Due to the large admixture of the ν0g7/2\nu 0g_{7/2} components into the 5/21+{5/2}^{+}_{1} state of 113Pd, the term that is proportional to

[a0νg7/2×a~π0g9/2](1)\displaystyle[a^{\dagger}_{0\nu g_{7/2}}\times\tilde{a}_{\pi 0g_{9/2}}]^{(1)} (28)

makes a sizable contribution to the GT transition strength. The matrix element of this component, which amounts to 0.788-0.788 (0.850) in the mapped (phenomenological) calculation, is so large that the corresponding logft\log{ft} value is too small as compared with the experimental value.

As noted above, there are notable quantitative differences between the mapped and phenomenological predictions for the logft\log{ft} values in the case of the 113Rh(7/21+)113({7/2}^{+}_{1})\to^{113}Pd(7/21+)({7/2}^{+}_{1}) decay. The GT transition matrix element obtained in the phenomenological calculation is two orders of magnitude smaller than the one obtained within the mapped scheme. This difference stems from a subtle balance between matrix elements of different terms in the GT transition operator. The dominant contribution to the GT matrix element in the former calculation come from the term proportional to the expression in Eq. (28), and the one of the form

sν[[d~ν×aν0g7/2](7/2)×aπ0g9/2](1).\displaystyle s^{\dagger}_{\nu}[[\tilde{d}_{\nu}\times a^{\dagger}_{\nu 0g_{7/2}}]^{(7/2)}\times a_{\pi 0g_{9/2}}]^{(1)}\;. (29)

Their matrix elements are of the same order of magnitude, but have the opposite signs, hence cancellation occurs between these terms. The degree of the cancellation, however, is much smaller in the mapped calculation. The contribution of the Fermi matrix element is negligibly small in both the mapped and phenomenological cases.

The logft\log{ft} values for the Rh decays into the non-yrast states, 5/22+{5/2}^{+}_{2} and 7/22+{7/2}^{+}_{2}, of the odd-AA Pd are shown in Figs. 13(c) and 13(d), respectively. The predicted logft\log{ft} values for the ARh(7/21+)A({7/2}^{+}_{1})\to^{A}Pd(5/22+)({5/2}^{+}_{2}) decay in the two sets of calculations are generally large, logft7\log{ft}\gtrsim 7 for A111A\lesssim 111. In particular, they overestimate the experimental values for the 105Rh, 107Rh, and 109Rh decays by a factor of two. The discrepancy could be attributed to the nature of the IBFM-2 wave functions and the components of the GT operator. The computed logft\log{ft} values for the ARh(7/21+)A({7/2}^{+}_{1})\to^{A}Pd(7/22+)({7/2}^{+}_{2}) decay in the mapped scheme are close to the experimental values, with an exception made of the 105Rh decay.

Table 5 gives complementary results for the logft\log{ft} values of the β\beta^{-} decays ARh(7/21+)A({7/2}^{+}_{1})\to^{A}Pd(If+)(I^{+}_{f}), with final states other than those already discussed above. The predicted logft\log{ft} values are compared with the available experimental data [70].

Previous IBFM-2 calculations [7] provided logft\log{ft} values for the β\beta^{-} decays 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} and 7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} in 105,107,109Rh which are consistent with the experimental ones. However, for the same nuclei the values logft\log{ft}\approx 4 were obtained for the 7/21+5/22+{7/2}^{+}_{1}\to{5/2}^{+}_{2} β\beta^{-} decay. Such logft\log{ft} values are systematically smaller than the experimental values and those obtained in this work. A more recent IBFM-2 calculation for the 115,117Rh115,117\to^{115,117}Pd β\beta^{-} decay [13] obtained a value logft\log{ft} = 5.90 for the 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} decay of 115Rh. This logft\log{ft} value is close to the one obtained in this study. On the other hand, for the 7/21+5/21+{7/2}^{+}_{1}\to{5/2}^{+}_{1} and 7/21+7/21+{7/2}^{+}_{1}\to{7/2}^{+}_{1} decays of 117Rh, the values logft\log{ft} = 6.78 and 6.68 were reported in [13]. They are approximately 20 % larger than those obtained in the present work.

Refer to caption
Figure 14: The same as in Fig. 13, but for the β\beta^{-} decays (a) 11+01+1^{+}_{1}\to 0^{+}_{1}, (b) 11+21+1^{+}_{1}\to 2^{+}_{1}, (c) 11+02+1^{+}_{1}\to 0^{+}_{2}, and (d) 11+22+1^{+}_{1}\to 2^{+}_{2} from the even-AA Rh into Pd nuclei.
Refer to caption
Figure 15: Reduced matrix elements of the ν0g7/2π0g9/2\nu 0g_{7/2}-\pi 0g_{9/2} terms in the GT transition operators, and total GT matrix elements for the β\beta^{-} decays (a) 11+01+{1}^{+}_{1}\to 0^{+}_{1}, (b) 11+21+{1}^{+}_{1}\to 2^{+}_{1}, (c) 11+02+{1}^{+}_{1}\to 0^{+}_{2}, and (d) 11+22+{1}^{+}_{1}\to 2^{+}_{2} of the even-AA Rh, resulting from the mapped and phenomenological calculations.

VI.2 β\beta decays of even-AA nuclei

The logft\log{ft} values for the β\beta^{-} decays of the even-AA Rh into Pd nuclei are plotted in Fig. 14. One immediately sees from Fig. 14(a) that the mapped and phenomenological logft\log{ft} values for the ARh(11+)A({1}^{+}_{1})\to^{A}Pd(01+)(0^{+}_{1}) decays are, approximately, a factor two smaller than the experimental ones. The corresponding GT matrix elements are almost purely determined by the contributions of the terms associated with the ν0g7/2π0g9/2\nu 0g_{7/2}-\pi 0g_{9/2} coupling, i.e.,

[a~ν0g7/2×a~π0g9/2](1),\displaystyle[\tilde{a}_{\nu 0g_{7/2}}\times\tilde{a}_{\pi 0g_{9/2}}]^{(1)}\;, (30)

for N65N\leqslant 65 and

sν[a~ν0g7/2×a~π0g9/2](1),\displaystyle s_{\nu}^{\dagger}[\tilde{a}_{\nu 0g_{7/2}}\times\tilde{a}_{\pi 0g_{9/2}}]^{(1)}\;, (31)

for N67N\geqslant 67. As shown in Fig. 15(a), the matrix elements of these terms are particularly large for the mass A116A\leqslant 116. Note also, that the IBFFM-2 wave functions for the initial 11+1^{+}_{1} state mainly consist of the pair configuration [ν0g7/2π0g9/2](J)[\nu 0g_{7/2}\otimes\pi 0g_{9/2}]^{(J)} for the even-AA Rh with A118A\leqslant 118 [see Fig. 12(a)]. For the larger mass A120A\geqslant 120, this pair configuration becomes less important in the 11+1^{+}_{1} wave function of the final nucleus. As a consequence, the GT transition strength decreases with increasing AA [see Fig. 15(a)].

To reproduce the β\beta-decay logft\log{ft} data, effective values of the gAg_{\mathrm{A}} factor, gA,effg_{\mathrm{A,eff}}, are often employed. Here we compare the predicted logft\log{ft} value for the ARh(11+)A({1}^{+}_{1})\to^{A}Pd(01+)(0^{+}_{1}) decay with the corresponding experimental one, and extract the gA,effg_{\mathrm{A,eff}} values for those decays for which logft\log{ft} data are available. The resulting gA,effg_{\mathrm{A,eff}} values are, on average, gA,eff0.152g_{\mathrm{A,eff}}\approx 0.152 (0.205) in the mapped (phenomenological) scheme. This amounts to a reduction of the free value by approximately by 88 (84) %. In the previous IBM-2/IBFFM-2 study of the β\beta and ββ\beta\beta decays of the Te and Xe isotopes with A130A\approx 130 [9], the gA,effg_{\mathrm{A,eff}} values extracted from a comparison with the logft\log{ft} data for the single-β\beta decays are 0.313 for the β+\beta^{+} decay 128I(11+)128(1^{+}_{1})\to^{128}Te(01+)(0^{+}_{1}), and 0.255 for the β\beta^{-} decay 128I(11+)128(1^{+}_{1})\to^{128}Xe(01+)(0^{+}_{1}).

As can be seen from Fig. 14(b), the logft\log{ft} values obtained within the mapped and phenomenological approaches for the ARh(11+)A({1}^{+}_{1})\to^{A}Pd(21+)(2^{+}_{1}) decay differ considerably. The difference between the two calculations is especially large at A=110A=110 and 116. One sees from Fig. 15(b), that the GT matrix element M(GT;11+21+)M(\text{GT};{1}^{+}_{1}\to 2^{+}_{1}) for the 116Rh decay in the phenomenological calculations is much larger in magnitude than the one obtained within the mapped approach, with the largest contribution coming from the term associated with the ν0g7/2π0g9/2\nu 0g_{7/2}-\pi 0g_{9/2} coupling. Generally, the predicted logft\log{ft} values for the 11+21+1^{+}_{1}\to 2^{+}_{1} β\beta^{-} decay, both within the mapped and phenomenological schemes, increase with AA (or NN). This is due to the fact that the pair configuration [ν0g7/2π0g9/2](J)[\nu 0g_{7/2}\otimes\pi 0g_{9/2}]^{(J)} gradually becomes less important in the 11+1^{+}_{1} wave function of the even-AA Rh for larger AA [see Fig. 12(a)].

For the ARh(11+)A({1}^{+}_{1})\to^{A}Pd(02+)(0^{+}_{2}) decay, the logft\log{ft} values predicted within the mapped and phenomenological approaches are similar. The most notable difference occurs at A=116A=116, with the mapped logft\log{ft} value being nearly half the phenomenological one. This is a consequence of the fact that in the mapped GT matrix element M(GT;11+02+)M(\text{GT};{1}^{+}_{1}\to 0^{+}_{2}) associated with the 116Rh decay, the component of Eq. (31) is an order of magnitude larger than the one in the phenomenological calculations [see Fig. 15(c)]. In addition, the computed logft\log{ft} values for the 11+02+1^{+}_{1}\to 0^{+}_{2} decay are larger than those for the 11+01+1^{+}_{1}\to 0^{+}_{1} decay because the matrix elements of the components involving the coupling ν0g7/2π0g9/2\nu 0g_{7/2}-\pi 0g_{9/2} in the M(GT;11+02+)M(\text{GT};{1}^{+}_{1}\to 0^{+}_{2}) strength are smaller in magnitude than those in the M(GT;11+01+)M(\text{GT};{1}^{+}_{1}\to 0^{+}_{1}) one.

The logft\log{ft} values corresponding to the ARh(11+)A({1}^{+}_{1})\to^{A}Pd(22+)(2^{+}_{2}) decay are depicted in Fig. 14(d). Both the mapped and phenomenological calculations largely underestimate the measured value at A=104A=104. However, the results obtained with both schemes reproduce the experimental trend reasonably well for 108A116108\leqslant A\leqslant 116. As can be seen from Fig. 15(d), the difference between the mapped and phenomenological results for 104A108104\leqslant A\leqslant 108 is due to the difference between the matrix elements for the components ν0g7/2π0g9/2\nu 0g_{7/2}-\pi 0g_{9/2} in both schemes, with the mapped matrix elements being an order of magnitude smaller than the phenomenological ones.

Table 6: The same as in Table 5, but for the β\beta^{-} decays from even-AA Rh to Pd nuclei.
Calc.
Decay IiIfI_{i}\to I_{f} mapped phen. Expt.
104Rh104\to^{104}Pd 11+01+1^{+}_{1}\to 0^{+}_{1} 3.27 3.21 4.55(1)
11+21+1^{+}_{1}\to 2^{+}_{1} 3.54 5.41 5.80(1)
11+02+1^{+}_{1}\to 0^{+}_{2} 5.91 5.85 7.36(2)
11+22+1^{+}_{1}\to 2^{+}_{2} 6.03 4.45 8.7(1)
11+03+1^{+}_{1}\to 0^{+}_{3} 6.42 6.05 5.5(1)
11+23+1^{+}_{1}\to 2^{+}_{3} 5.24 4.72 6.3(1)
51+41+5^{+}_{1}\to 4^{+}_{1} 7.26 8.30 7.3(1)
51+42+5^{+}_{1}\to 4^{+}_{2} 8.45 7.59 6.1(1)
51+43+5^{+}_{1}\to 4^{+}_{3} 8.06 8.04 6.2(1)
51+44+5^{+}_{1}\to 4^{+}_{4} 8.59 8.57 5.8(1)
106Rh106\to^{106}Pd 11+01+1^{+}_{1}\to 0^{+}_{1} 3.31 3.43 5.168(7)
11+21+1^{+}_{1}\to 2^{+}_{1} 3.72 4.29 5.865(17)
11+22+1^{+}_{1}\to 2^{+}_{2} 6.78 4.72 6.55(7)
11+02+1^{+}_{1}\to 0^{+}_{2} 5.39 6.82 5.354(19)
11+23+1^{+}_{1}\to 2^{+}_{3} 5.15 4.58 5.757(17)
108Rh108\to^{108}Pd 11+01+1^{+}_{1}\to 0^{+}_{1} 3.31 3.45 5.5(3)
11+21+1^{+}_{1}\to 2^{+}_{1} 3.97 4.14 5.7(4)
11+22+1^{+}_{1}\to 2^{+}_{2} 7.06 5.00 6.0(4)
11+02+1^{+}_{1}\to 0^{+}_{2} 5.07 6.01 5.6(4)
51+61+5^{+}_{1}\to 6^{+}_{1} 7.72 7.44 6.8(3)
51+42+5^{+}_{1}\to 4^{+}_{2} 8.28 7.00 4.84(9)1114+,5+,6+{4}^{+},{5}^{+},6^{+} level at 2864 keV
51+51+5^{+}_{1}\to 5^{+}_{1} 9.59 8.35 4.84(9)1114+,5+,6+{4}^{+},{5}^{+},6^{+} level at 2864 keV
51+62+5^{+}_{1}\to 6^{+}_{2} 9.30 9.42 4.84(9)1114+,5+,6+{4}^{+},{5}^{+},6^{+} level at 2864 keV
110Rh110\to^{110}Pd 61+61+6^{+}_{1}\to 6^{+}_{1} 8.29 8.26 6.38(13)
61+62+6^{+}_{1}\to 6^{+}_{2} 9.57 8.95 7.1(4)
61+51+6^{+}_{1}\to 5^{+}_{1} 9.16 8.69 6.34(25)
112Rh112\to^{112}Pd 11+01+1^{+}_{1}\to 0^{+}_{1} 3.55 3.61 \approx5.5
11+21+1^{+}_{1}\to 2^{+}_{1} 4.88 4.35 6.2(3)
11+22+1^{+}_{1}\to 2^{+}_{2} 5.53 5.86 6.4(3)
11+02+1^{+}_{1}\to 0^{+}_{2} 6.20 5.01 6.52(6)
11+03+1^{+}_{1}\to 0^{+}_{3} 7.48 6.36 6.88(9)222(0,1,2)+(0,1,2)^{+} level at 1140 keV
11+11+1^{+}_{1}\to 1^{+}_{1} 7.74 5.66 6.88(9)222(0,1,2)+(0,1,2)^{+} level at 1140 keV
11+23+1^{+}_{1}\to 2^{+}_{3} 5.83 5.39 6.88(9)222(0,1,2)+(0,1,2)^{+} level at 1140 keV
11+23+1^{+}_{1}\to 2^{+}_{3} 5.83 5.39 6.97(22)
11+23+1^{+}_{1}\to 2^{+}_{3} 5.83 5.39 6.50(7)
61+61+6^{+}_{1}\to 6^{+}_{1} 8.75 8.80 6.52333logft\log{ft} values should be considered approximate [70].
61+51+6^{+}_{1}\to 5^{+}_{1} 8.96 10.34 6.54
61+62+6^{+}_{1}\to 6^{+}_{2} 9.15 8.82 6.88
114Rh114\to^{114}Pd 11+01+1^{+}_{1}\to 0^{+}_{1} 3.59 4.37 5.9(2)
11+21+1^{+}_{1}\to 2^{+}_{1} 5.19 3.89 6.0(4)
11+22+1^{+}_{1}\to 2^{+}_{2} 6.60 6.08 5.7(2)
11+02+1^{+}_{1}\to 0^{+}_{2} 4.59 5.10 6.1(2)
11+23+1^{+}_{1}\to 2^{+}_{3} 5.57 5.28 6.1(2)
116Rh116\to^{116}Pd 11+01+1^{+}_{1}\to 0^{+}_{1} 3.75 4.38 5.62(22)
11+21+1^{+}_{1}\to 2^{+}_{1} 6.36 4.04 5.84(18)
11+22+1^{+}_{1}\to 2^{+}_{2} 6.99 6.63 5.76(19)
11+02+1^{+}_{1}\to 0^{+}_{2} 4.45 8.03 6.47(20)
11+03+1^{+}_{1}\to 0^{+}_{3} 5.29 8.60 6.36(19)
11+23+1^{+}_{1}\to 2^{+}_{3} 5.05 5.00 6.81(21)

For the sake of completeness, Table 6 compares the predicted and experimental [70] logft\log{ft} values for the β\beta^{-} decays of the even-AA Rh isotopes. Cases other than hose already discussed above are considered in the table. As compared with the ground-state-to-ground-state decay 11+01+1^{+}_{1}\to 0^{+}_{1}, the ftft values for the decays of the 11+1^{+}_{1} state into non-yrast 1+1^{+} and 2+2^{+} states, and the logft\log{ft} values for the 51+If5^{+}_{1}\to I_{f} and 61+If6^{+}_{1}\to I_{f} decays are calculated to be large. Note that the predicted logft\log{ft} values for the decays 104Rh(51+)104(5^{+}_{1})\to^{104}Pd(41+)(4^{+}_{1}) and 108Rh(51+)108(5^{+}_{1})\to^{108}Pd(61+)(6^{+}_{1}) are rather close to the experimental ones.

VII Conclusions

In this paper, the low-energy collective states and β\beta decays for even and odd-mass neutron-rich Rh and Pd isotopes have been studied using a mapping framework based on the Gogny-EDF and the particle-boson coupling scheme. The constrained HFB has been employed to provide microscopic input to the mapping procedure. Such an input consists of potential energy surfaces as functions of the (β,γ)(\beta,\gamma) shape degrees of freedom for the even-even 104-124Pd isotopes. The IBM-2 Hamiltonian, used to describe even-even core nuclei, has been determined by mapping the Gogny-D1M HFB fermionic potential energy surfaces onto the corresponding bosonic surfaces. The microscopic mean-field calculations also provided single-particle energies for the odd systems. Those represent essential building blocks of the boson-fermion interactions for the neighboring odd-AA and odd-odd nuclei as well as for the GT and Fermi transition operators. The strength parameters of the boson-fermion and residual neutron-proton interactions were fitted to low-energy data for the odd-AA and odd-odd systems.

The Gogny-HFB (β,γ)(\beta,\gamma) potential energy surfaces obtained for even-even Pd isotopes point towards a transition from prolate deformed (104-108Pd) to γ\gamma-soft (110-116Pd), and to nearly spherical shapes (118-124Pd). The low-energy excitation spectra and B(E2)B(E2) transition strengths resulting from the diagonalization of the mapped IBM-2 Hamiltonian reproduced the experimental trends reasonably well and reflect, to a large extent, the structural evolution of the ground-state shapes predicted at the mean-field level. The excitation energies obtained for the low-lying positive-parity levels in the odd-AA Pd and Rh, and even-AA Rh nuclei also exhibit signatures of this structural evolution. Within this context, a notable example is the change in the ground state spin from 113Pd to 115Pd. The computed logft\log{ft} values for the β\beta^{-} decays of the odd- and even-AA Rh into Pd nuclei have been shown to be sensitive to the nature of the wave functions of the parent and daughter nuclei. They also reflect the rapid structural evolution along the considered isotopic chains. The logft\log{ft} values for the odd-AA Rh decay have been predicted to be larger than the experimental ones for A109A\lesssim 109. This could be traced back to the structure of the IBFM-2 wave functions for the odd-AA daughter (Pd) nuclei. Furthermore, it has been shown that for the even-AA Rh decay, the neutron-proton pair components [ν0g7/2π0g9/2](J)[\nu 0g_{7/2}\otimes\pi 0g_{9/2}]^{(J)} play a key role in the GT transition matrix elements and are responsible for the too small logft\log{ft} values for the ARh(11+)A(1^{+}_{1})\to^{A}Pd(01+)(0^{+}_{1}) decay with respect to the experimental data.

The results of the mapped calculations have been compared with conventional IBM-2 calculations in which the parameters for the boson Hamiltonian have been fit to the experiment. The mapped and phenomenological IBM-2 excitation spectra for even-even, odd-AA, and odd-odd systems are similar. However, the two sets of calculations differ in their predictions for electromagnetic and β\beta-decay properties of the odd-nucleon systems.

The results obtained in this study could be considered a plausible step towards a consistent simultaneous description of the low-lying states and β\beta-decay properties of atomic nuclei. However, the difference between the predicted and experimental β\beta-decay logft\log{ft} values might require additional refinements of the employed theoretical framework. In particular, the small logft\log{ft} values obtained suggest that the role of the effective axial-vector coupling constant gAg_{\mathrm{A}} should be further studied in future calculations. The gA,effg_{\mathrm{A,eff}} values extracted in this work from the comparison with the experimental data turned out to be by a factor 7-8 smaller than the free nucleon value. This large quenching indicates deficiencies in the model space of the calculations or of the theoretical procedure itself. Investigation along these lines is in progress and will be reported elsewhere.

Acknowledgements.
This work is financed within the Tenure Track Pilot Programme of the Croatian Science Foundation and the École Polytechnique Fédérale de Lausanne, and Project No. TTP-2018-07-3554 Exotic Nuclear Structure and Dynamics, with funds of the Croatian-Swiss Research Programme. The work of LMR is supported by the Spanish Ministry of Economy and Competitiveness (MINECO) Grant No. PGC2018-094583-B-I00.

References

  • Nishimura et al. [2011] S. Nishimura, Z. Li, H. Watanabe, K. Yoshinaga, T. Sumikama, T. Tachibana, K. Yamaguchi, M. Kurata-Nishimura, G. Lorusso, Y. Miyashita, A. Odahara, H. Baba, J. S. Berryman, N. Blasi, A. Bracco, F. Camera, J. Chiba, P. Doornenbal, S. Go, T. Hashimoto, S. Hayakawa, C. Hinke, E. Ideguchi, T. Isobe, Y. Ito, D. G. Jenkins, Y. Kawada, N. Kobayashi, Y. Kondo, R. Krücken, S. Kubono, T. Nakano, H. J. Ong, S. Ota, Z. Podolyák, H. Sakurai, H. Scheit, K. Steiger, D. Steppenbeck, K. Sugimoto, S. Takano, A. Takashima, K. Tajiri, T. Teranishi, Y. Wakabayashi, P. M. Walker, O. Wieland, and H. Yamaguchi, Phys. Rev. Lett. 106, 052502 (2011).
  • Lorusso et al. [2015] G. Lorusso, S. Nishimura, Z. Y. Xu, A. Jungclaus, Y. Shimizu, G. S. Simpson, P.-A. Söderström, H. Watanabe, F. Browne, P. Doornenbal, G. Gey, H. S. Jung, B. Meyer, T. Sumikama, J. Taprogge, Z. Vajta, J. Wu, H. Baba, G. Benzoni, K. Y. Chae, F. C. L. Crespi, N. Fukuda, R. Gernhäuser, N. Inabe, T. Isobe, T. Kajino, D. Kameda, G. D. Kim, Y.-K. Kim, I. Kojouharov, F. G. Kondev, T. Kubo, N. Kurz, Y. K. Kwon, G. J. Lane, Z. Li, A. Montaner-Pizá, K. Moschner, F. Naqvi, M. Niikura, H. Nishibata, A. Odahara, R. Orlandi, Z. Patel, Z. Podolyák, H. Sakurai, H. Schaffner, P. Schury, S. Shibagaki, K. Steiger, H. Suzuki, H. Takeda, A. Wendt, A. Yagi, and K. Yoshinaga, Phys. Rev. Lett. 114, 192501 (2015).
  • Quinn et al. [2012] M. Quinn, A. Aprahamian, J. Pereira, R. Surman, O. Arndt, T. Baumann, A. Becerril, T. Elliot, A. Estrade, D. Galaviz, T. Ginter, M. Hausmann, S. Hennrich, R. Kessler, K.-L. Kratz, G. Lorusso, P. F. Mantica, M. Matos, F. Montes, B. Pfeiffer, M. Portillo, H. Schatz, F. Schertz, L. Schnorrenberger, E. Smith, A. Stolz, W. B. Walters, and A. Wöhr, Phys. Rev. C 85, 035807 (2012).
  • Avignone et al. [2008] F. T. Avignone, S. R. Elliott, and J. Engel, Rev. Mod. Phys. 80, 481 (2008).
  • Navrátil and Dobe [1988] P. Navrátil and J. Dobe, Phys. Rev. C 37, 2126 (1988).
  • Dellagiacoma and Iachello [1989] F. Dellagiacoma and F. Iachello, Phys. Lett. B 218, 399 (1989).
  • Yoshida et al. [2002] N. Yoshida, L. Zuffi, and S. Brant, Phys. Rev. C 66, 014306 (2002).
  • Brant et al. [2004] S. Brant, N. Yoshida, and L. Zuffi, Phys. Rev. C 70, 054301 (2004).
  • Yoshida and Iachello [2013] N. Yoshida and F. Iachello, Prog. Theor. Exp. Phys. 2013, 043D01 (2013).
  • Mardones et al. [2016] E. Mardones, J. Barea, C. E. Alonso, and J. M. Arias, Phys. Rev. C 93, 034332 (2016).
  • Nomura et al. [2020a] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 101, 024311 (2020a).
  • Nomura et al. [2020b] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 101, 044318 (2020b).
  • Ferretti et al. [2020] J. Ferretti, J. Kotila, R. I. M. n. Vsevolodovna, and E. Santopinto, Phys. Rev. C 102, 054329 (2020).
  • Nomura [2022] K. Nomura, Phys. Rev. C 105, 044306 (2022).
  • Álvarez-Rodríguez et al. [2004] R. Álvarez-Rodríguez, P. Sarriguren, E. M. de Guerra, L. Pacearescu, A. Faessler, and F. Šimkovic, Phys. Rev. C 70, 064309 (2004).
  • Sarriguren [2015] P. Sarriguren, Phys. Rev. C 91, 044304 (2015).
  • Boillos and Sarriguren [2015] J. M. Boillos and P. Sarriguren, Phys. Rev. C 91, 034311 (2015).
  • Pirinen and Suhonen [2015] P. Pirinen and J. Suhonen, Phys. Rev. C 91, 054309 (2015).
  • Šimkovic et al. [2013] F. Šimkovic, V. Rodin, A. Faessler, and P. Vogel, Phys. Rev. C 87, 045501 (2013).
  • Mustonen and Engel [2016] M. T. Mustonen and J. Engel, Phys. Rev. C 93, 014304 (2016).
  • Suhonen [2017] J. T. Suhonen, Frontiers Phys. 5, 55 (2017).
  • Ravlić et al. [2021] A. Ravlić, E. Yüksel, Y. F. Niu, and N. Paar, Phys. Rev. C 104, 054318 (2021).
  • Minato et al. [2021] F. Minato, T. Marketin, and N. Paar, Phys. Rev. C 104, 044321 (2021).
  • Langanke and Martínez-Pinedo [2003] K. Langanke and G. Martínez-Pinedo, Rev. Mod. Phys. 75, 819 (2003).
  • Caurier et al. [2005] E. Caurier, G. Martínez-Pinedo, F. Nowacki, A. Poves, and A. P. Zuker, Rev. Mod. Phys. 77, 427 (2005).
  • Yoshida et al. [2018] S. Yoshida, Y. Utsuno, N. Shimizu, and T. Otsuka, Phys. Rev. C 97, 054321 (2018).
  • Suzuki et al. [2018] T. Suzuki, S. Shibagaki, T. Yoshida, T. Kajino, and T. Otsuka, Astrophys. J. 859, 133 (2018).
  • Kumar et al. [2020] A. Kumar, P. C. Srivastava, J. Kostensalo, and J. Suhonen, Phys. Rev. C 101, 064304 (2020).
  • Iachello and Arima [1987] F. Iachello and A. Arima, The interacting boson model (Cambridge University Press, Cambridge, 1987).
  • Iachello and Scholten [1979] F. Iachello and O. Scholten, Phys. Rev. Lett. 43, 679 (1979).
  • Iachello and Van Isacker [1991] F. Iachello and P. Van Isacker, The interacting boson-fermion model (Cambridge University Press, Cambridge, 1991).
  • Brant et al. [1984] S. Brant, V. Paar, and D. Vretenar, Z. Phys. A 319, 355 (1984).
  • Ring and Schuck [1980] P. Ring and P. Schuck, The nuclear many-body problem (Springer, Berlin, 1980).
  • Goriely et al. [2009] S. Goriely, S. Hilaire, M. Girod, and S. Péru, Phys. Rev. Lett. 102, 242501 (2009).
  • J. Decharge and M. Girod and D. Gogny [1975] J. Decharge and M. Girod and D. Gogny, Phys. Lett. B 55, 361 (1975).
  • Robledo et al. [2019] L. M. Robledo, T. R. Rodríguez, and R. R. Rodríguez-Guzmán, J. Phys. G: Nucl. Part. Phys. 46, 013001 (2019).
  • Van Isacker and Puddu [1980] P. Van Isacker and G. Puddu, Nucl. Phys. A 348, 125 (1980).
  • Bender et al. [2003] M. Bender, P.-H. Heenen, and P.-G. Reinhard, Rev. Mod. Phys. 75, 121 (2003).
  • Vretenar et al. [2005] D. Vretenar, A. V. Afanasjev, G. A. Lalazissis, and P. Ring, Phys. Rep. 409, 101 (2005).
  • Nikšić et al. [2011] T. Nikšić, D. Vretenar, and P. Ring, Prog. Part. Nucl. Phys. 66, 519 (2011).
  • Berger et al. [1984] J. F. Berger, M. Girod, and D. Gogny, Nucl. Phys. A 428, 23 (1984).
  • Borrajo and Egido [2016] M. Borrajo and J. L. Egido, The European Physical Journal A 52, 277 (2016).
  • Garrett et al. [2020] P. E. Garrett, T. R. Rodríguez, A. Diaz Varela, K. L. Green, J. Bangay, A. Finlay, R. A. E. Austin, G. C. Ball, D. S. Bandyopadhyay, V. Bildstein, S. Colosimo, D. S. Cross, G. A. Demand, P. Finlay, A. B. Garnsworthy, G. F. Grinyer, G. Hackman, B. Jigmeddorj, J. Jolie, W. D. Kulp, K. G. Leach, A. C. Morton, J. N. Orce, C. J. Pearson, A. A. Phillips, A. J. Radich, E. T. Rand, M. A. Schumaker, C. E. Svensson, C. Sumithrarachchi, S. Triambak, N. Warr, J. Wong, J. L. Wood, and S. W. Yates, Phys. Rev. C 101, 044302 (2020).
  • Siciliano et al. [2020] M. Siciliano, I. Zanon, A. Goasduff, P. R. John, T. R. Rodríguez, S. Péru, I. Deloncle, J. Libert, M. Zielińska, D. Ashad, D. Bazzacco, G. Benzoni, B. Birkenbach, A. Boso, T. Braunroth, M. Cicerchia, N. Cieplicka-Oryńczak, G. Colucci, F. Davide, G. de Angelis, B. de Canditiis, A. Gadea, L. P. Gaffney, F. Galtarossa, A. Gozzelino, K. Hadyńska-Klȩk, G. Jaworski, P. Koseoglou, S. M. Lenzi, B. Melon, R. Menegazzo, D. Mengoni, A. Nannini, D. R. Napoli, J. Pakarinen, D. Quero, P. Rath, F. Recchia, M. Rocchini, D. Testov, J. J. Valiente-Dobón, A. Vogt, J. Wiederhold, and W. Witt, Phys. Rev. C 102, 014318 (2020).
  • Siciliano et al. [2021] M. Siciliano, J. J. Valiente-Dobón, A. Goasduff, T. R. Rodríguez, D. Bazzacco, G. Benzoni, T. Braunroth, N. Cieplicka-Oryńczak, E. Clément, F. C. L. Crespi, G. de France, M. Doncel, S. Ertürk, C. Fransen, A. Gadea, G. Georgiev, A. Goldkuhle, U. Jakobsson, G. Jaworski, P. R. John, I. Kuti, A. Lemasson, H. Li, A. Lopez-Martens, T. Marchi, D. Mengoni, C. Michelagnoli, T. Mijatović, C. Müller-Gatermann, D. R. Napoli, J. Nyberg, M. Palacz, R. M. Pérez-Vidal, B. Sayği, D. Sohler, S. Szilner, and D. Testov, Phys. Rev. C 104, 034320 (2021).
  • Rodríguez-Guzmán et al. [2020] R. Rodríguez-Guzmán, Y. M. Humadi, and L. M. Robledo, J. Phys. G: Nucl. Part. Phys. 48, 015103 (2020).
  • Rodríguez-Guzmán and Robledo [2021] R. Rodríguez-Guzmán and L. M. Robledo, Phys. Rev. C 103, 044301 (2021).
  • Rodríguez-Guzmán et al. [2021] R. Rodríguez-Guzmán, L. M. Robledo, K. Nomura, and N. C. Hernandez, J. Phys. G: Nucl. Part. Phys. 49, 015101 (2021).
  • Nomura et al. [2012] K. Nomura, R. Rodríguez-Guzmán, L. M. Robledo, and N. Shimizu, Phys. Rev. C 86, 034322 (2012).
  • Nomura et al. [2013] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 87, 064313 (2013).
  • Nomura et al. [2016a] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 94, 044314 (2016a).
  • Nomura et al. [2015] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 92, 014312 (2015).
  • Nomura et al. [2020c] K. Nomura, R. Rodríguez-Guzmán, Y. M. Humadi, L. M. Robledo, and J. E. García-Ramos, Phys. Rev. C 102, 064326 (2020c).
  • Nomura et al. [2021a] K. Nomura, R. Rodríguez-Guzmán, L. Robledo, and J. García-Ramos, Phys. Rev. C 103, 044311 (2021a).
  • Nomura et al. [2021b] K. Nomura, R. Rodríguez-Guzmán, L. M. Robledo, J. E. García-Ramos, and N. C. Hernández, Phys. Rev. C 104, 044324 (2021b).
  • Nomura et al. [2021c] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 104, 054320 (2021c).
  • Otsuka et al. [1978a] T. Otsuka, A. Arima, and F. Iachello, Nucl. Phys. A 309, 1 (1978a).
  • Otsuka et al. [1978b] T. Otsuka, A. Arima, F. Iachello, and I. Talmi, Phys. Lett. B 76, 139 (1978b).
  • Scholten [1985] O. Scholten, Prog. Part. Nucl. Phys. 14, 189 (1985).
  • Brant and Paar [1988] S. Brant and V. Paar, Z. Phys. A 329, 151 (1988).
  • Bohr and Mottelson [1975] A. Bohr and B. R. Mottelson, Nuclear Structure (Benjamin, New York, 1975).
  • Ginocchio and Kirson [1980] J. N. Ginocchio and M. W. Kirson, Nucl. Phys. A 350, 31 (1980).
  • Nomura et al. [2008] K. Nomura, N. Shimizu, and T. Otsuka, Phys. Rev. Lett. 101, 142501 (2008).
  • Nomura et al. [2010] K. Nomura, N. Shimizu, and T. Otsuka, Phys. Rev. C 81, 044307 (2010).
  • Nomura et al. [2016b] K. Nomura, T. Nikšić, and D. Vretenar, Phys. Rev. C 93, 054305 (2016b).
  • Nomura et al. [2017] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 96, 064316 (2017).
  • Nomura et al. [2019] K. Nomura, R. Rodríguez-Guzmán, and L. M. Robledo, Phys. Rev. C 99, 034308 (2019).
  • Dellagiacoma [1988] F. Dellagiacoma, Beta decay of odd mass nuclei in the interacting boson-fermion model, Ph.D. thesis, Yale University (1988).
  • Iachello [2000] F. Iachello, Phys. Rev. Lett. 85, 3580 (2000).
  • [70] Brookhaven National Nuclear Data Center, http://www.nndc.bnl.gov.
  • Kurpeta et al. [2018] J. Kurpeta et al.Phys. Rev. C 98, 024318 (2018).
  • Kurpeta et al. [2022] J. Kurpeta et al.Phys. Rev. C 105, 034316 (2022).
  • Stone [2005] N. Stone, At. Data Nucl. Data Tables 90, 75 (2005).
  • Arias et al. [1987] J. Arias, C. Alonso, and M. Lozano, Nucl. Phys. A 466, 295 (1987).