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

Three-Dimensional Approach Applied to Quasi-stationary States of Deformed α\alpha-Emitters

Ruijia Li School of Physics, Nanjing University, Nanjing 210093, China    Chang Xu [email protected] School of Physics, Nanjing University, Nanjing 210093, China Institute for Nonperturbative Physics, Nanjing University, Nanjing 210093, China
Abstract

We apply a three-dimensional (3D) approach to investigate the quasi-stationary states of well-deformed α\alpha-emitters. With a splitting of the anisotropic 3D potential into internal and external parts at a separation surface, the 3D α\alpha-cluster decay width is determined by the initial wave function of a true bound state of an anisotropic harmonic oscillator potential and a non-resonance scattering wave function of Coulomb potential. Substantial difference between the 1D and 3D decay width is found for typical α\alpha-emitters with large quadrupole and hexadecapole deformations.

pacs:
23.60. + e, 21.60.Gx

I Introduction

The quantum tunneling problem of α\alpha-cluster decay provides insight into not only the quasi-stationary states of unstable nuclei but also the nature of α\alpha-clustering in surrounding nuclear medium DLW ; Delion09 ; Buck ; cxu06 ; Mohr06 ; review1 ; review2 ; Ismail16 ; Perez19 ; Denisov09 ; Royer ; Poenaru . Much of our understanding of α\alpha-cluster decay comes from the well-established quantum picture where an α\alpha-cluster tunnels through the Coulomb barrier after its formation on the surface of nuclei. The α\alpha-clustering process is rather complex to handle, which in principle involves a quantum four-body problem embedded in nuclear medium THSR ; quartet2017 ; Po ; Yasu21 ; Eiji20 ; SY20 ; SY21 ; Xu2017 ; Xu2016 . In contrast, the tunneling probability can be quantitatively estimated from the WKB approximation by assuming an α\alpha-cluster interacting with a spherical daughter nucleus. This assumption is appropriate for α\alpha-emitters with small deformations Buck ; cxu06 ; Mohr06 . However, large deformations could be involved in the α\alpha-cluster decay of heavy nuclei and exotic nuclei far away from the β\beta-stability line Oganessian ; Batchelder ; Te2019 ; Zhang . The exact 3D decay width of well-deformed α\alpha-emitters is difficult to obtain by treating self-consistently both deformation and mixing of angular momenta. An empirical way to include effect of large deformations is to average the tunneling probabilities of all angles, which yields an enhanced α\alpha-cluster tunneling probability. For α\alpha-transitions with measured core excitation energies, the total wave function of the system can be expanded in terms of the ground and excited states channels and the coupled Schrödinger equations can be solved Barmore ; channel1 ; Delion06 .

The exact solution of quasi-stationary state with an anisotropic potential is quite interesting in many quantum systems QT ; Spins ; Electrons ; QD ; Superconductor ; TPA ; MTPA . In this work, we apply a 3D approach to investigate the quasi-stationary state of α\alpha-cluster decay problem, namely the 3D two-potential approach (3D-TPA) MDTPA . We emphasize that the quasi-stationary state of α\alpha-cluster decay highly resembles a bound state more than a scattering state. The anisotropic α\alpha-core potential is properly divided into inner and outer potentials on a separation surface inside the Coulomb barrier. The formed α\alpha-cluster is considered to stay initially in the bound state generated by the inner potential, and then transforms to a quasi-stationary state by switching on the outer potential. Using the time-dependent perturbation theory, both the decay width and the energy shift of α\alpha-cluster decay can be well defined by inner 3D bound state wave function and outer scattering state wave function on the separation surface. The multi-dimensional quantum tunneling problem for well-deformed α\alpha-emitters is reduced to a problem of true bound state plus non-resonance scattering state MDTPA . It is worth noting that the choice of separation surface does not affect the final results as long as it is inside the classical forbidden region. The challenge here is the numerical solution of the inner 3D bound state wave function on the separation surface, whose value is approximately of the order of 10-15. Although several numerical methods can be applied to solve the 3D Schrödinger equation to obtain the bound state wave function such as the grid-based approach, the imaginary time propagation method, and the basis expansion method, however, it is still an open question to obtain accurately the inner 3D bound state wave function at large distances due to the limitation of matrix size or the number of bases Imaginary ; Gaussian . In this sense, the anisotropic harmonic oscillator potential with exact solutions is a good choice to simulate the internal α\alpha-core 3D potential. The 1D and 3D decay widths are compared for several typical non-spherical α\alpha-emitters with large quadrupole and hexadecapole deformations. Note that the formation process of the α\alpha-cluster on nuclear surface is not touched here, which in principle does not affect the comparison between 1D and 3D decay widths.

The paper is organized as follows: In Sec.II, we give the formalism of 3D two-potential approach for the multi-dimensional quantum tunneling problem. The anisotropic α\alpha-core effective potential and the choice of separation surface are introduced in Sec.III. In Sec.IV, both the inner 3D bound state wave function and outer scattering state wave function are explicitly given. In Sec.V, the 3D approach is applied to several deformed α\alpha-emitters U236{}^{236}\textrm{U}, Pu244{}^{244}\textrm{Pu}, Cm246{}^{246}\textrm{Cm}, Cf250{}^{250}\textrm{Cf}, and Fm254{}^{254}\textrm{Fm}. The possible theoretical uncertainties are also discussed. Conclusions are presented in Sec.VI.

II Formalism of 3D two-potential approach for multi-dimensional quantum tunneling problem

The time-dependent perturbation theory is applied to solve the multi-dimensional tunneling problems in 3D-TPA MDTPA . The main integrant is the separation of the 3D α\alpha-core potential V(𝒓)V({\mbox{\boldmath$r$}}) into the inner U(𝒓)U({\mbox{\boldmath$r$}}) and outer W(𝒓)W({\mbox{\boldmath$r$}}) potentials on the separation surface 𝒮\mathcal{S}

V(𝒓)=U(𝒓)+W(𝒓).V({\mbox{\boldmath$r$}})=U({\mbox{\boldmath$r$}})+W({\mbox{\boldmath$r$}}). (1)

The inner potential is

U(𝒓)={V(𝒓)inner regionU0outer region,U({\mbox{\boldmath$r$}})=\left\{\begin{aligned} &V({\mbox{\boldmath$r$}})&\text{inner region}\\ &U_{0}&\text{outer region},\end{aligned}\right. (2)

where U0U_{0} is the minimal value of V(𝒓)V({\mbox{\boldmath$r$}}) on the separation surface 𝒮\mathcal{S}. The outer potential is

W(𝒓)={0inner regionV(𝒓)U0outer region.W({\mbox{\boldmath$r$}})=\left\{\begin{aligned} &0&\text{inner region}\\ &V({\mbox{\boldmath$r$}})-U_{0}&\text{outer region}.\end{aligned}\right. (3)

The quasi-stationary state of α\alpha-cluster decay can be described by the time-dependent Schrödinger equation,

it|Ψ(t)\displaystyle i\hbar{\partial\over\partial t}|\Psi(t)\rangle =[22μ2+V(𝒓)]|Ψ(t)\displaystyle=\bigg{[}-\frac{\hbar^{2}}{2\mu}{\mbox{\boldmath$\nabla$}}^{2}+V({\mbox{\boldmath$r$}})\bigg{]}|\Psi(t)\rangle (4)
=[H0+W(𝒓)]|Ψ(t),\displaystyle=\bigg{[}H_{0}+W({\mbox{\boldmath$r$}})\bigg{]}|\Psi(t)\rangle,

where H0=22μ2+U(𝒓)H_{0}=-\frac{\hbar^{2}}{2\mu}{\mbox{\boldmath$\nabla$}}^{2}+U({\mbox{\boldmath$r$}}) is the Hamiltonian of bound state |Φi|\Phi_{i}\rangle confined in the inner potential U(𝒓)U({\mbox{\boldmath$r$}}), and the corresponding time-independent Schrödinger equation is [22μ2+U(𝒓)]|Φi=E0|Φi[-\frac{\hbar^{2}}{2\mu}{\mbox{\boldmath$\nabla$}}^{2}+U({\mbox{\boldmath$r$}})]|\Phi_{i}\rangle=E_{0}|\Phi_{i}\rangle. The separation surface 𝒮\mathcal{S} is not necessarily spherical, as long as it is taken between the equipotential surfaces S1S_{1} and S2S_{2}. At t>0t>0, the “unperturbed” bound state |Φi|\Phi_{i}\rangle is no longer an eigenstate of the total Hamiltonian H=H0+W(𝒓)H=H_{0}+W({\mbox{\boldmath$r$}}), but a wave packet spreading in time due to the perturbation W(𝒓)W({\mbox{\boldmath$r$}})

|Ψ(t)=b0(t)|ΦieiE0t/+bk(t)|φkeiEkt/dk(2π)3,|\Psi(t)\rangle=b_{0}(t)|\Phi_{i}\rangle e^{-iE_{0}t/\hbar}+\int b_{\textbf{k}}(t)|\varphi_{\textbf{k}}\rangle e^{-iE_{\textbf{k}}t/\hbar}\frac{d\textbf{k}}{(2\pi)^{3}}, (5)

where b0(t)b_{0}(t) and bk(t)b_{\textbf{k}}(t) are the probability amplitudes of finding the system in the eigenstates |Φi|\Phi_{i}\rangle and |φk|\varphi_{\textbf{k}}\rangle, respectively. The amplitudes b0(t)b_{0}(t) and bk(t)b_{\textbf{k}}(t) can be found from Eq.(4) with the initial condition: b0(t)=1,bk(t)=0b_{0}(t)=1,b_{\textbf{k}}(t)=0. The energy shift Re(ϵ0)Re(\epsilon_{0}) and the width Γ=2Im(ϵ0)\Gamma=-2Im(\epsilon_{0}) of the quasi-stationary state are directly related to the pole in the complex E-plane using the Green’s function technique TPA

ϵ0=EE0=Φi|W|Φi+Φi|WG~(E)W|Φi.\epsilon_{0}=E-E_{0}=\langle\Phi_{i}|W|\Phi_{i}\rangle+\langle\Phi_{i}|W\tilde{G}(E)W|\Phi_{i}\rangle. (6)

The Green’s function G~\tilde{G} is given by

G~(E)=G0(E)[1+W~G~(E)],\tilde{G}(E)=G_{0}(E)\Big{[}1+\tilde{W}\tilde{G}(E)\Big{]}, (7)

where W~=W+U0\tilde{W}=W+U_{0} is used instead of WW, to ensure the potential vanishes for rr\rightarrow\infty. G0(E)G_{0}(E) is given by

G0(E)=1ΛE+U0KU(𝒓),Λ=|ΦiΦi|.G_{0}(E)=\frac{1-\Lambda}{E+U_{0}-K-U({\mbox{\boldmath$r$}})},\qquad\Lambda=|\Phi_{i}\rangle\langle\Phi_{i}|. (8)

Note that the above derivations are general, but the numerical solution of G~\tilde{G} is rather difficult and converges very slowly. To make it feasible within the capacity of computer calculation, G~\tilde{G} is expanded in powers of GW~G_{\tilde{W}}, namely the Green function corresponding to W~(𝒓)\tilde{W}({\mbox{\boldmath$r$}})

G~(E)=GW~(E)+GW~(E)(UU0)G~(E)G~W~(E)Λ[1+W~G~(E)].\tilde{G}(E)=G_{\tilde{W}}(E)+G_{\tilde{W}}(E)(U-U_{0})\tilde{G}(E)-\tilde{G}_{\tilde{W}}(E)\Lambda\Big{[}1+\tilde{W}\tilde{G}(E)\Big{]}. (9)

Substitute Eq.(9) into Eq.(6), then retain the first order and assume that the energy shift is small compared to E0E_{0}, one can replace GW~(E)G_{\tilde{W}}(E) by GW~(E0)G_{\tilde{W}}(E_{0})

E=E0+Φi|W|Φi+Φi|WGW~(E0)W|Φi.E=E_{0}+\langle\Phi_{i}|W|\Phi_{i}\rangle+\langle\Phi_{i}|WG_{\tilde{W}}(E_{0})W|\Phi_{i}\rangle\,. (10)

The Schrödinger equation of the Green’s function is

[E0KW~(𝒓)]GW~(E0;𝒓,𝒓)=δ(𝒓𝒓),\left[E_{0}-K-\tilde{W}({\mbox{\boldmath$r$}})\right]G_{\tilde{W}}(E_{0};{\mbox{\boldmath$r$}},{\mbox{\boldmath$r$}}^{\prime})=\delta({\mbox{\boldmath$r$}}-{\mbox{\boldmath$r$}}^{\prime}), (11)

and the spectral representation for the Green’s function is

GW~=|φkφk|E0Ekiηdk(2π)3,G_{\tilde{W}}=\int{|\varphi_{\textbf{k}}\rangle\langle\varphi_{\textbf{k}}|\over E_{0}-E_{\textbf{k}}-i\eta}{d\textbf{k}\over(2\pi)^{3}}, (12)

where Ek=2k2/2μE_{\textbf{k}}=\hbar^{2}\textbf{k}^{2}/2\mu. The |φk|\varphi_{\textbf{k}}\rangle is the eigenstate corresponding to non-resonance scattering states, which satisfies the following Schrödinger equation

[K+W~(𝒓)]|φk=Ek|φk.\Big{[}K+\tilde{W}({\mbox{\boldmath$r$}})\Big{]}|\varphi_{\textbf{k}}\rangle=E_{\textbf{k}}|\varphi_{\textbf{k}}\rangle. (13)

One can obtain the total width Γ\Gamma as an integral over the partial width Γk\Gamma_{\textbf{k}} MDTPA

Γ=Γkdk1dk2(2π)2||k|=k0,\Gamma=\left.\int\Gamma_{\textbf{k}}{dk_{1}dk_{2}\over(2\pi)^{2}}\right|_{|\textbf{k}|=k_{0}}, (14)

where k0=2mE0/2k_{0}=\sqrt{2mE_{0}/\hbar^{2}}. The partial width Γk\Gamma_{\textbf{k}} is

Γ𝒌=24μk3|𝒓Φi(𝒓)rφk(𝒓)dσ||k|=k02,\Gamma_{{\mbox{\boldmath$\scriptstyle k$}}}={\hbar^{2}\over 4\mu k_{3}}\left|\int_{{\mbox{\boldmath$r$}}\in\mathcal{B}}\Phi_{i}({\mbox{\boldmath$r$}})\stackrel{{\scriptstyle\leftrightarrow}}{{{\mbox{\boldmath$\nabla$}}}}_{\textbf{r}}\varphi_{\textbf{k}}({\mbox{\boldmath$r$}})d\sigma\right|^{2}_{|\textbf{k}|=k_{0}}, (15)

where the symbol r\stackrel{{\scriptstyle\leftrightarrow}}{{{\mbox{\boldmath$\nabla$}}}}_{\textbf{r}} means the gradient on the right minus the gradient on the left. 𝒌={k1,k2,k3}{\mbox{\boldmath$k$}}=\{k_{1},k_{2},k_{3}\} is the momentum vector. We note that Eq.(15) is similar to the Bardeen formula for the tunneling coupling between adjoining wells from a many-particle point of view, which is widely used in the solid state and atomic physics Bardeen ; Gurvitz . Finally, the decay half-life is given by

T1/2=ln2PαΓ,T_{1/2}=\frac{\hbar\ln{2}}{P_{\alpha}\Gamma}, (16)

where Pα is the formation probability of α\alpha-cluster on nuclear surface. Recent microscopic calculation of the α\alpha-cluster formation probability for the ideal Po isotopes has been performed by the quartetting wave function approach (QWFA), in which the intrinsic motion of the four nucleons forming the α\alpha-cluster and the center of mass motion between the α\alpha-cluster and the core are correctly treated SY20 ; SY21 ; Xu2016 ; Xu2017 . Empirically, the α\alpha-cluster formation probability PαP_{\alpha} is known to change abruptly across the major shell closures and differs for even-even, odd-A and odd-odd nuclei, as indicated by the experimental systematics. The decay width Γ\Gamma can be directly obtained by using Eq.(14) and Eq.(15), however, it is more convenient to re-formulate Γ\Gamma in the following way

Γ=24μ02π0πk0sinθkdθkdϕk(2π)2×|02π0πΦi(r,θ,ϕ)rφk(r,θ,ϕ)Rs(θ,ϕ)\displaystyle\Gamma=\frac{\hbar^{2}}{4\mu}\int_{0}^{2\pi}\int_{0}^{\pi}\frac{k_{0}\sin{\theta_{k}}d\theta_{k}d\phi_{k}}{(2\pi)^{2}}\times\Bigg{|}\int_{0}^{2\pi}\int_{0}^{\pi}\Phi_{i}(r,\theta,\phi)\stackrel{{\scriptstyle\leftrightarrow}}{{{\mbox{\boldmath$\nabla$}}}}_{r}\varphi_{k}(r,\theta,\phi)R_{s}(\theta,\phi) (17)
[dRs(θ,ϕ)dϕ]2+sinθ2[[dRs(θ,ϕ)dθ]2+Rs2(θ,ϕ)]dθdϕ|r=Rs(θ,ϕ)2,\displaystyle\sqrt{[\frac{dR_{s}(\theta,\phi)}{d\phi}]^{2}+\sin{\theta}^{2}\bigg{[}\big{[}\frac{dR_{s}(\theta,\phi)}{d\theta}\big{]}^{2}+R_{s}^{2}(\theta,\phi)\bigg{]}}d\theta d\phi\Bigg{|}_{r=R_{s}(\theta,\phi)}^{2},

where Rs(θ,ϕ)R_{s}(\theta,\phi) is the radius of the separation surface 𝒮\mathcal{S}. For spherical emitters, Eq.(LABEL:gam3) reduces exactly to the 1D decay width Γ=2μk0|ϕi(R)χk(R)χk(R)ϕi(R)|2\Gamma={\hbar^{2}\over\mu k_{0}}|\phi_{i}(R)\chi_{k}^{\prime}(R)-\chi_{k}(R)\phi_{i}^{\prime}(R)|^{2} TPA .

III α\alpha-core 3D effective potential and separation surface

Refer to caption
Figure 1: The α\alpha-core 3D potential V0(r,θ)V_{0}(r,\theta) in Cartesian coordinate system. The inner part of V0(r,θ)V_{0}(r,\theta) is an anisotropic harmonic oscillator potential, and the outer part of V0(r,θ)V_{0}(r,\theta) a deformed Coulomb potential. For demonstration, the α\alpha-core potential V0(r,θ)V_{0}(r,\theta) with the angle θ=0\theta=0 is shown in the small panel. r1(θ=0)r_{1}(\theta=0) and r2(θ=0)r_{2}(\theta=0) are the first and second turning points, respectively. 𝓈(θ=0)\mathcal{R_{s}}(\theta=0) is the matching point where VI(r,θ=0)κω0[L2L2N]=VC(r,θ=0)V_{I}(r,\theta=0)-\kappa\omega_{0}\hbar\big{[}L^{2}-\langle L^{2}\rangle_{N}\big{]}=V_{C}(r,\theta=0).

The 3D-TPA requires the information on 3D bound-state wave function deep inside the classical forbidden region where the wave function decreases exponentially and is extremely small (of the order of  10-15). This poses a big challenge for numerical calculations, even with high-performance parallel computing. Here we simulate the 3D α\alpha-core potential with an anisotropic harmonic oscillator potential that is of rotational symmetry. The relevant Hamiltonian is of Nilsson-form, which has a L2L^{2} correction term Nilsson ; SPM

H=22μ2+V0(r,θ),H=-\frac{\hbar^{2}}{2\mu}{\mbox{\boldmath$\nabla$}}^{2}+V_{0}(r,\theta),\ (18)

with

V0(r,θ)={VI(r,θ)κω0[L2L2N]r𝓈(θ)VC(r,θ)r>𝓈(θ),V_{0}(r,\theta)=\left\{\begin{aligned} &V_{I}(r,\theta)-\kappa\omega_{0}\hbar\bigg{[}L^{2}-\langle L^{2}\rangle_{N}\bigg{]}&&r\leq\mathcal{R_{s}}(\theta)\\ &V_{C}(r,\theta)&&r>\mathcal{R_{s}}(\theta),\end{aligned}\right. (19)

where 𝓈(θ)\mathcal{R_{s}}(\theta) describes the boundary condition VI(r,θ)κω0[L2L2N]=VC(r,θ)V_{I}(r,\theta)-\kappa\omega_{0}\hbar\big{[}L^{2}-\langle L^{2}\rangle_{N}\big{]}=V_{C}(r,\theta). The anisotropic harmonic oscillator potential with depth parameter DD is

VI(r,θ)=12mω2[(rcosθ)2+γ2(rsinθ)2]D,V_{I}(r,\theta)=\frac{1}{2}m\omega^{2}\bigg{[}(r\cos\theta)^{2}+\gamma^{2}(r\sin\theta)^{2}\bigg{]}-D, (20)

where ω\omega, ω0\omega_{0} and γ\gamma read

ω=2Dμ[R(0)/ar]2,ω0=2Dμ[R0/ar]2,γ=R(0)R(π2).\omega=\sqrt{\frac{2D}{\mu\Big{[}R(0)/a_{r}\Big{]}^{2}}},~{}~{}~{}~{}\omega_{0}=\sqrt{\frac{2D}{\mu\Big{[}R_{0}/a_{r}\Big{]}^{2}}},~{}~{}~{}~{}\gamma=\frac{R(0)}{R(\frac{\pi}{2})}. (21)

The deformed Coulomb potential is

VC(r,θ)\displaystyle V_{C}(r,\theta) =1.44ZcZdr[1+35(R0r)2β2mY2m(θ)]\displaystyle=\frac{1.44Z_{c}Z_{d}}{r}\bigg{[}1+\frac{3}{5}(\frac{R_{0}}{r})^{2}\beta_{2}\sum_{m}Y_{2m}(\theta)\bigg{]} (22)
+187π(R0r)2[β22Y20(θ)+345(R0r)2β22Y40(θ)],\displaystyle+\sqrt{\frac{18}{7\pi}}(\frac{R_{0}}{r})^{2}\bigg{[}-\beta_{2}^{2}Y_{20}(\theta)+\frac{3}{4}\sqrt{5}(\frac{R_{0}}{r})^{2}\beta_{2}^{2}Y_{40}(\theta)\bigg{]},

The half-density radius R(θ)R(\theta) is given by

R(θ)=R0[1+β2Y20(θ)+β4Y40(θ)],R(\theta)=R_{0}\bigg{[}1+\beta_{2}Y_{20}(\theta)+\beta_{4}Y_{40}(\theta)\bigg{]}, (23)

where the parameter R0=1.07Ad1/3R_{0}=1.07A_{d}^{1/3}rrr , and β2\beta_{2} and β4\beta_{4} are quadrupole and hexadecapole deformations, respectively. The deformed Coulomb potential VCV_{C} can be regarded as an isotropic potential at very large distances. The L2L^{2} correction term is introduced to eliminate the angular momentum degeneracy, resulting in different depths of potential for different angular momenta.

The details of the α\alpha-core 3D potential V0(r,θ)V_{0}(r,\theta) are shown in Fig.1, in which the inner potential joins with the outer Coulomb potential at 𝓈(θ)\mathcal{R_{s}}(\theta). The total α\alpha-core potential is divided into two parts by the separation surface 𝒮\mathcal{S}, which can be chosen between the equipotential surfaces S1S_{1} and S2S_{2} with the same energy E0E_{0}. Here, we take the separation surface 𝒮\mathcal{S} as coinciding with 𝓈(θ)\mathcal{R_{s}}(\theta) (see Fig.2). In this way, one can minimize the numerical errors associated with Φi(r,θ,ϕ)\Phi_{i}(r,\theta,\phi) and φk(r,θ,ϕ)\varphi_{\textbf{k}}(r,\theta,\phi). Note that the boundary 𝓈(θ)\mathcal{R_{s}}(\theta) is azimuth-independent (ϕ\phi-independent). Thus, the separation surface 𝒮\mathcal{S} is also azimuth-independent in our calculations.

Refer to caption
Figure 2: The separation of the α\alpha-core potential V0(r,θ)V_{0}(r,\theta) into inner and outer regions. S1S_{1} and S2S_{2} denote the equipotential surfaces with the same energy E0E_{0}. The separation surface 𝒮\mathcal{S}, denoted by the radius Rs(θ)R_{s}(\theta), is chosen as coinciding with 𝓈(θ)\mathcal{R_{s}}(\theta).

IV Inner 3D bound state wave function and outer scattering state wave function

We use the analytical 3D wave function of anisotropic harmonic oscillator potential as an approximation of Φi(r,θ,ϕ)\Phi_{i}(r,\theta,\phi), which is the eigenstate of E0E_{0} corresponding to the inner potential. As a matter of fact, we only need the information of Φi(r,θ,ϕ)\Phi_{i}(r,\theta,\phi) on the surface 𝒮\mathcal{S} where the 3D wave function of anisotropic harmonic oscillator is considered to be a good approximation for well-deformed α\alpha-emitters

Φi(r,θ,ϕ)=Ψnρm(r,θ)Ψnz(r,θ)eimϕ2π,\Phi_{i}(r,\theta,\phi)=\Psi_{n_{\rho}}^{m}(r,\theta)\Psi_{n_{z}}(r,\theta)\frac{e^{im\phi}}{\sqrt{2\pi}}, (24)

where

Ψnρm(r,θ)\displaystyle\Psi_{n_{\rho}}^{m}(r,\theta) =Nnρmβ|m|+12(rsinθ)|m|e(r2sinθ2β2)/2Lnρ|m|(rsinθ),\displaystyle=N_{n_{\rho}}^{m}\beta_{\perp}^{|m|+1}\sqrt{2}(r\sin{\theta})^{|m|}e^{-(r^{2}\sin{\theta}^{2}\beta_{\perp}^{2})/2}L_{n_{\rho}}^{|m|}(r\sin{\theta}), (25)

and

Ψnz(r,θ)=Nnzβz1/2er2cosθ2/2Hnz(rcosθ),\Psi_{n_{z}}(r,\theta)=N_{n_{z}}\beta_{z}^{1/2}e^{-r^{2}\cos{\theta}^{2}/2}H_{n_{z}}(r\cos\theta), (26)

with parameters β\beta_{\perp}, βz\beta_{z}, NnρmN_{n_{\rho}}^{m} and NnzN_{n_{z}}

β=(mωγ)1/2\displaystyle\beta_{\perp}=(\frac{m\omega\gamma}{\hbar})^{1/2} ,βz=(mω)1/2,\displaystyle,\quad\beta_{z}=(\frac{m\omega}{\hbar})^{1/2}, (27a)
Nnρm=(nρ!(nρ+|m|)!)\displaystyle N_{n_{\rho}}^{m}=(\frac{n_{\rho}!}{(n_{\rho}+|m|)!}) ,Nnz=(1π2nznz!)1/2.\displaystyle,\quad N_{n_{z}}=(\frac{1}{\sqrt{\pi}2^{n_{z}}n_{z}!})^{1/2}.\quad (27b)
Refer to caption
Figure 3: (a) The 3D inner wave function Φi(r,θ,ϕ=0)\Phi_{i}(r,\theta,\phi=0) of the spherical case with β2=0\beta_{2}=0 and β4=0\beta_{4}=0. (b) Φi(r,θ,ϕ=0)\Phi_{i}(r,\theta,\phi=0) of the well-deformed case with β2=0.226\beta_{2}=0.226, β4=0.108\beta_{4}=0.108. For any ϕ\phi, the Φi(r,θ,ϕ)\Phi_{i}(r,\theta,\phi) has the same shape.
Refer to caption
Figure 4: The wave function χ(r,θ)\chi(r,\theta) in the outer potential W~(r,θ)\tilde{W}(r,\theta). The black, blue, and red curves refer to χ(r,θ)\chi(r,\theta) with θ=0\theta=0, θ=π/4\theta=\pi/4 and θ=π/2\theta=\pi/2, respectively, which almost coincide with each other.

The expectation value E0E_{0} is given by SPM

E0=nznρ\displaystyle E_{0}=\langle n_{z}n_{\rho} m|H|nznρm=(2nρ+m+1)ωγ+(nz+12)ω\displaystyle m|H|n_{z}n_{\rho}m\rangle=(2n_{\rho}+m+1)\hbar\omega\gamma+(n_{z}+\frac{1}{2})\hbar\omega (28)
+ω0κ[12(2nzG12)2m218]D,\displaystyle+\hbar\omega_{0}\kappa\big{[}\frac{1}{2}(2n_{z}-G-\frac{1}{2})^{2}-m^{2}-\frac{1}{8}\big{]}-D,

where the depth DD can be obtained by matching the value of experimental decay energy QαQ_{\alpha} with E0E_{0}. The choice of nρn_{\rho} and nzn_{z} should fulfill the so-called Wildermuth-Tang rule WT

G=2nρ+nz=i=14gi,G=2n_{\rho}+n_{z}=\sum_{i=1}^{4}g_{i}, (29)

where GG is the global quantum number and gig_{i} are the oscillator quantum numbers of the nucleons forming the α\alpha-cluster. The GG value is restricted by the Pauli principle. For instance, GG is usually taken as G=22G=22 for heavy nuclei with neutron numbers N>126N>126. In principle, nzn_{z} can be taken as an even number from 0 to GG. Here, nzn_{z} is taken as nz=10n_{z}=10 as an example.

The scattering wave function φk(r,θ,ϕ)\varphi_{k}(r,\theta,\phi) is solved from the scattering Schrödinger equation Eq.(13) with the outer potential (see Fig.3(b)), which has the form scattering ; QM

φk(r,θ,ϕ)=2πl,m2πile±iσlYlm(θ,ϕ)Ylm(θk,ϕk)χ(r,θ)kr,\varphi_{k}(r,\theta,\phi)=2\pi\sum_{l,m}\sqrt{\frac{2}{\pi}}i^{l}e^{\pm i\sigma_{l}}Y_{lm}(\theta,\phi)Y_{lm}^{*}(\theta_{k},\phi_{k})\frac{\chi(r,\theta)}{kr}, (30)

where χ(r,θ)\chi(r,\theta) can be represented by the linear combination of the regular Fl(r)F_{l}(r) and irregular Gl(r)G_{l}(r) solutions of Coulomb potential. The corresponding coefficients in the linear combination of Fl(r)F_{l}(r) and Gl(r)G_{l}(r) are directly related to the non-resonant scattering phase shift for the outer potential W~(r)\tilde{W}(r) MTPA

χ(r,θ)=cosδl(θ)Fl(r)+sinδl(θ)Gl(r),\chi(r,\theta)=\cos{\delta_{l}(\theta)}F_{l}(r)+\sin{\delta_{l}(\theta)}G_{l}(r), (31)

where the relative phase δl\delta_{l} obeys

tanδl(θ)=FlRs(θ)+αFlRs(θ)GlRs(θ)αGlRs(θ),\tan{\delta_{l}}(\theta)=\frac{-F_{l}R_{s}^{\prime}(\theta)+\alpha F_{l}R_{s}(\theta)}{G_{l}R_{s}^{\prime}(\theta)-\alpha G_{l}R_{s}(\theta)}, (32)

and

α=2μ(U0Qα).\alpha=\frac{\sqrt{2\mu(U_{0}-Q_{\alpha})}}{\hbar}. (33)

The inner wave function Φi(r,θ,ϕ)\Phi_{i}(r,\theta,\phi) and the outer Coulomb function χ(r,θ)\chi(r,\theta) are shown in Fig.3 and Fig.4, respectively. For spherical emitters, the inner wave function shown in Fig.3(a) is the spherical harmonic oscillator wave function. For well-deformed emitters, the inner wave function is approximated by the anisotropic harmonic oscillator wave function as shown in Fig.3(b). As for the scattering state, the wave function χ(r,θ)\chi(r,\theta) becomes almost isotropic at large distances (see Fig.4).

V Comparison between 1D and 3D cases and uncertainty analysis

As shown in Table 1, all the α\alpha-cluster emitters selected in calculations are well-deformed with β2>0.22\beta_{2}>0.22 and far away from the major shell closures Z=82Z=82 and N=126N=126. Moreover, only the α\alpha-transitions between ground-states (0+0+0^{+}\longrightarrow 0^{+}) are considered in order to check the validity of 3D-TPA. There are three adjustable parameters in 3D-TPA that should be determined, namely ara_{r}, κ\kappa, and DD. ara_{r} is associated with the angular frequency ω\omega of harmonic oscillator. κ\kappa is the strength coefficient of orbit correction term. DD is the depth of harmonic oscillator potential, which can be obtained by matching the experimental decay energy QαQ_{\alpha} with the energy E0E_{0} once the quantum numbers G,nz,mG,n_{z},m are determined. It is noted that different quantum numbers G,nz,mG,n_{z},m can be chosen. Here the same parameters and quantum numbers are used for all emitters, namely, ar=0.7883,κ=0.0054,G=22,nz=10a_{r}=0.7883,\kappa=0.0054,G=22,n_{z}=10, and m=0m=0.

Table 1: Comparison between 1D and 3D α\alpha-decay half-lives (in log base 10 and in seconds) for 254Fm, 250Cf, 246Cm, 244Pu, and 236U. The information on both parent and daughter nuclei are listed in columns 1–4. Column 5 gives the experimental α\alpha-decay energy QαQ_{\alpha}. The theoretical quadrupole and the hexadecapole deformations are listed in columns 6 and 7, respectively. The experimental α\alpha-decay half-lives are given in column 8. In the last two columns, the 1D and 3D α\alpha-cluster decay half-lives from TPA are given.
ApA_{p} ZpZ_{p} AdA_{d} ZdZ_{d} Qα(MeV)Q_{\alpha}(MeV) β2\beta_{2} β4\beta_{4} TExp.T_{Exp.} TCal.1DT_{Cal.}^{1D} TCal.3DT_{Cal.}^{3D}
236 U 232 Th 4.572 0.226 0.108 15.003 16.2416 14.9246
244 Pu 240 U 4.6655 0.237 0.061 15.5015 16.5278 15.5721
246 Cm 242 Pu 5.4748 0.249 0.051 11.2615 12.3441 11.2006
250 Cf 246 Cm 6.1284 0.250 0.027 8.6984 9.32804 8.63641
254 Fm 250 Cf 7.307 0.251 0.015 4.1415 4.73273 4.14148
Refer to caption
Figure 5: The comparison of α\alpha-decay half-lives (in logarithm with a base 10) in spherical and deformed cases. The red hexagrams refer to the experimental half-lives and the blue circles the calculated results from 3D-TPA.

We show in Table 1 the comparison between 1D and 3D α\alpha-decay half-lives for typical α\alpha-emitter U236{}^{236}\textrm{U}, Pu244{}^{244}\textrm{Pu}, Cm246{}^{246}\textrm{Cm}, Cf250{}^{250}\textrm{Cf}, and Fm254{}^{254}\textrm{Fm}. Note that the quadrupole and hexadecapole deformations are taken from Ref.bb24 . QαQ_{\alpha} and experimental results of α\alpha-decay half-lives are taken from Refs.AME1 ; AME2 . One can see from Table 1 that the experimental α\alpha-decay half-lives vary in a quite large range from 10410^{4} to 101510^{15} seconds. This is also helpful for testing the validity of present 3D model for not only short-lived but also long-lived α\alpha-emitters. The formation problem of α\alpha-cluster on the surface of parent nucleus is not considered here and its value is assumed to be unity (Pα=1). This assumption is reasonable because the α\alpha-cluster formation probability is known to change dramatically in the vicinity of shell closures but smoothly in the open-shell region here. One can see from Table 1 that deviations of the calculated α\alpha-decay half-lives exist between the 1D case and 3D case. For all the α\alpha-emitters considered in this work, the largest deviation occurs for the decay of 236U, which is possibly due to its small decay energy. In general, the 1D results are all reduced by taking nuclear deformation into account (shown also in Fig.5). Similar conclusions can also be found in the empirical approaches cxu06 . More importantly, the 3D-TPA reproduces nicely the systematics of experimental data with only one set of parameters. This is quite different from the 1D-TPA without the consideration of deformation, in which one may need to adjust the parameters such as the depth of potential for each emitter.

Finally, we discuss the possible theoretical uncertainties of present 3D-TPA calculations, which mainly come from the following aspects: a) the error associated with the approximations such as the replacement of GW~(E)G_{\tilde{W}}(E) by GW~(E0)G_{\tilde{W}}(E_{0}) in 3D-TPA. The correction due to this approximation is expected to be negligible because of very small energy shift from E0E_{0} to EE; b) the inner 3D wave function is approximated by the exact solutions of an anisotropic harmonic oscillator potential. We solved numerically the bound-state wave function corresponding to the inner potential in the spherical case, and found its value only deviates from the exact result by several percentages on the separation surface; c) the scattering wave function χ(r,θ)\chi(r,\theta) is considered to be isotropic on the separation surface. We have checked this approximation by taking the decay of U236{}^{236}U as an example, and found that the final results are almost not affected.

VI Summary

Large deformation is relevant to the α\alpha-cluster decay of heavy nuclei and nuclides far away from the β\beta-stability line. In this work, we apply a three-dimensional approach with the Nilsson-form Hamiltonian to calculate the decay widths of typical α\alpha-emitters U236{}^{236}\textrm{U}, Pu244{}^{244}\textrm{Pu}, Cm246{}^{246}\textrm{Cm}, Cf250{}^{250}\textrm{Cf}, and Fm254{}^{254}\textrm{Fm} by dividing the 3D effective potential into a bound-state inner region and a scattering outer region. The inner wave function can be well approximated by the exact solution of the anisotropic harmonic oscillator potential, and the scattering wave function can be safely considered as isotropic on the separation surface. Substantial difference is found between 1D and 3D decay width for favored transitions of these α\alpha-emitters. The systematics of experimental α\alpha-cluster decay half-lives is nicely reproduced. In the future, state-of-art numerical approaches can be applied to evaluate accurately the inner 3D wave function at large distances for arbitrary 3D potentials. Moreover, the combination of the present 3D-TPA and approaches of clustering such as quartetting wave function approach should be performed in order to predict reliably the α\alpha-decay half-lives of unknown nuclei far away from the major shell closures.

𝐀𝐜𝐤𝐧𝐨𝐰𝐥𝐞𝐝𝐠𝐦𝐞𝐧𝐭𝐬\bf{Acknowledgments}

The work is supported by the National Natural Science Foundation of China (Grant No. 12275129) and the Fundamental Research Funds for the Central Universities (Grant No. 020414380209).

References

  • (1) D. S. Delion, R. J. Liotta, and R. Wyss, Phys. Rev. C 92, 051301(R) (2015).
  • (2) D. S. Delion, Phys. Rev. C 80, 024310 (2009).
  • (3) B. Buck, A. C. Merchant, and S. M. Perez, Phys. Rev. C 45, 2247 (1992).
  • (4) C. Xu and Z. Ren, Phys. Rev. C 74, 014304 (2006).
  • (5) P. Mohr, Phys. Rev. C 73, 031301(R) (2006).
  • (6) R. G. Lovas, R. J. Liotta, A. Insolia, K. Varga, and D. S. Delion, Phys. Rep. 294, 265 (1998).
  • (7) C. Qi, R. J. Liotta, and R. Wyss, Prog. Part. Nucl. Phys. 105, 214 (2019).
  • (8) M. Ismail, W. M. Seif, and A. Abdurrahman, Phys. Rev. C 94, 024316(2016).
  • (9) J. E. Perez Velasquez, N. G. Kelkar, and N. J. Upadhyay, Phys. Rev. C 99, 024308(2019).
  • (10) V. Yu. Denisov and A. A. Khudenko, Phys. Rev. C 79, 054614 (2009).
  • (11) G. Royer and R. A. Gherghescu, Nucl. Phys. A 699, 479 (2002).
  • (12) D. N. Poenaru and R. A. Gherghescu, Phys. Rev. C 97, 044621 (2018) .
  • (13) A. Tohsaki, H. Horiuchi, P. Schuck, and G. Röpke, Phys. Rev. Lett. 87, 192501 (2001).
  • (14) G. Röpke, P. Schuck, C. Xu et al., J. Low Temp. Phys. 189, 383 (2017).
  • (15) G. Röpke, P. Schuck, Y. Funaki, H. Horiuchi, Z. Ren, A. Tohsaki, C. Xu, T. Yamada, and B. Zhou, Phys. Rev. C 90, 034304 (2014).
  • (16) Y. Taniguchi, K. Yoshida, Y. Chiba, Y. K. En’yo, M. Kimura, and K. Ogata, Phys. Rev. C 103, L031305 (2021).
  • (17) E. Nakano, K. Iida, and W. Horiuchi, Phys. Rev. C 102, 055802 (2020).
  • (18) S. Yang, C. Xu, and G. Röpke, Phys. Rev. C 104, 034302 (2021).
  • (19) S. Yang, C. Xu, G. Röpke, P. Schuck, Z. Ren, Y. Funaki, H. Horiuchi, A. Tohsaki, T. Yamada, and B. Zhou, Phys. Rev. C 101, 024316 (2020).
  • (20) C. Xu, G. Röpke, P. Schuck, Z. Ren, Y. Funaki, H. Horiuchi, A. Tohsaki, T. Yamada, and Bo Zhou, Phys. Rev. C 95, 061306 (2017).
  • (21) C. Xu, G. Röpke, P. Schuck, Z. Ren, Y. Funaki, H. Horiuchi, A. Tohsaki, T. Yamada, and Bo Zhou, Phys. Rev. C 93, 011306(R) (2016).
  • (22) Yu. Ts. Oganessian et al., Phys. Rev. C 106, L031301 (2022).
  • (23) J. C. Batchelder, K. S. Toth, C. R. Bingham, L. T. Brown, L. F. Conticchio, C. N. Davids, D. Seweryniak, J. Wauters, J. L. Wood, and E. F. Zganjar, Phys. Rev. C 55, R2142 (1997).
  • (24) Y. Xiao, S. Go, R. Grzywacz et al., Phys. Rev. C 100, 034315 (2019).
  • (25) Z. Y. Zhang et al., Phys. Rev. Lett. 126, 152502 (2021).
  • (26) B. Barmore, A. T. Kruppa, W. Nazarewicz, and T. Vertse, Phys. Rev. C 62, 054315 (2000).
  • (27) D. Ni and Z. Ren, Phys. Rev. C 80, 051303(R) (2009).
  • (28) D. S. Delion, S. Peltonen, and J. Suhonen, Phys. Rev. C 73, 014315 (2006).
  • (29) M. Razavy, Quantum theory of tunneling, World Scientific Publishing (2013).
  • (30) J. L. Van Hemmen and A. Sütö, Physica B + C 141, 37 (1986).
  • (31) D. V. Averin and A. A. Odintsov, Phys. Lett. A 140, 251 (1989).
  • (32) C. Timm, Phys. Rev. B 77, 195416 (2008).
  • (33) N. Giordano and E. R. Schuler, Phys. Rev. Lett. 63, 2417 (1989).
  • (34) S. A. Gurvitz, Phys. Rev. A 38, 1747 (1988).
  • (35) S. A. Gurvitz, P. B. Semmes, W. Nazarewicz, and T. Vertse, Phys. Rev. A 69, 042705 (2004).
  • (36) S. A. Gurvitz, Multiple facets of quantization and supersymmetry: Michael Marinov memorial volume, World Scientific Publishing (2002).
  • (37) L. Lehtovaara, J. Toivanen, and J. Eloranta, J. Comp. Phys. 221, 148 (2007).
  • (38) E. Hiyama, Y. Kino, and M. Kamimura, Prog. Part. Nucl. Phys. 51, 223 (2003).
  • (39) J. Bardeen, Phys. Rev. Lett. 6, 57 (1961).
  • (40) S. A. Gurvitz, D. Mozyrsky and G. P. Berman, Phys. Rev. B 72, 205341 (2005).
  • (41) S. G. Nilsson, Dan. Mat. Fys. Medd. 29, 1 (1955).
  • (42) A. Bohr, B. R. Mottelson, Nuclear Structure, World Scientific Publishing (1998).
  • (43) B. Hahn, D. G. Ravenhall, and R. Hofstadter, Phys. Rev. 101, 1131 (1956).
  • (44) K. Wildermuth and Y. C. Tang, A unified theory of the nucleus, Springer-Verlag Publishing (2013).
  • (45) C. Cohen-Tannoudji, B. Diu and F. Laloë, Quantum mechanics, EDP Sciences Publishing (1986).
  • (46) T. Berggren, Nucl. Phys. A 109, 265 (1968).
  • (47) P. Möller, A. J. Sierk, T. Ichikawa, and H. Sagawa , Nucl. Phys. A 109, 1–204 (2016).
  • (48) W. J. Huang, M. Wang, F. G. Kondev, G. Audi, and S. Naimi, Chin. Phys. C 45, 030002 (2021).
  • (49) M. Wang, W. J. Huang, F. G. Kondev, G. Audi, and S. Naimi, Chin. Phys. C 45, 030003 (2021).