-
We consider a system comprising a nucleus with a specific number of muons. The total Hamiltonian of the nucleus-muons system can be expressed as follows:
$ \begin{equation} H_{N\mu s}= H_{N} + H_{\mu s} - \int \frac{ {\rm e}^2\rho_p({\bf{r}})\rho_\mu({\bf{r'}})}{|{\bf{r}} - {\bf{r'}}|} {\rm d} {\bf{r}}{\rm d} {\bf{r'}}, \end{equation} $
(1) where
$ H_{N} $ represents the nuclear Hamiltonian and$ H_{\mu s} $ denotes the Hamiltonian of the muons alone;$ \rho_p({\bf{r}}) $ and$ \rho_\mu({\bf{r}}) $ correspond to the proton density and muon density, respectively; and e represents the charge of a proton.$ H_{\mu s} $ can be written as follows:$ \begin{equation} H_{\mu s}= -\sum\limits_{i=1}^{N_\mu} \frac{\hbar^2}{2m_\mu} \nabla^2_i + \sum\limits_{i<j}^{N_\mu} \frac{{\rm e}^2}{|{\bf{r}}_i - {\bf{r}}_j|}, \end{equation} $
(2) where
$ \hbar $ is the Plank constant,$ N_\mu $ corresponds the total number of muons, and$ m_\mu $ denotes the mass of an muon.The nuclear interaction is modeled as a Skyrme's density-dependent interaction presented in Ref. [11]. Here, we only describe the general framework. The Skyrme interaction can be written as a potential:
$ \begin{equation} V=\sum\limits_{i<j} v_{i j}^{(2)}+\sum\limits_{i<j<k} v_{i j k}^{(3)}, \end{equation} $
(3) with a two-body part
$ v_{i j} $ and three-body part$ v_{i j k} $ . To simplify calculations, Skyrme used a short-range expansion for the two-body interaction and a zero-range force for the three-body force.Concerning the Skyrme interaction, there exists a straightforward manner to obtain the Hartree-Fock equations. Consider a nucleus whose ground state is represented by a Slater determinant ϕ of single-particle states
$ \phi_i $ :$ \begin{equation} \phi\left(x_1, x_2, \ldots, x_A\right)=\frac{1}{\sqrt{A !}} \operatorname{det}\left|\phi_i\left(x_j\right)\right|, \end{equation} $
(4) where x denotes the set
$ {\bf{r}},~ \sigma,~ q $ of space, spin, and isospin coordinates$ \left(q=+\dfrac{1}{2}\right. $ for a proton,$ -\dfrac{1}{2} $ for a neutron). The expectation value of the total energy is expressed as follows:$ \begin{equation} \begin{aligned} E= & \langle\phi|(T+V)| \phi\rangle \\ = & \sum\limits_i\langle {i}\left|\frac{p^2}{2 m}\right| {i}\rangle+\frac{1}{2} \sum\limits_{{i} j}\langle i j\left|\tilde{v}_{12}\right| ij\rangle +\frac{1}{6} \sum\limits_{{i} j k}\langle {i} j k\left|\tilde{v}_{123}\right| {i} j k\rangle \\ = & \int H({\bf{r}}) {\rm d} {\bf{r}}, \end{aligned} \end{equation} $
(5) where
$ \tilde{v} $ denotes an antisymmetrized matrix element. For the Skyrme interaction, the energy density$ H({\bf{r}}) $ is an algebraic function of the nucleon densities$ \rho_n\left(\rho_p\right) $ , kinetic energy$ \tau_n\left(\tau_p\right) $ , and spin densities$ {\bf{J}}_n\left({\bf{J}}_p\right) $ . These quantities depend in turn on the single-particle states$ \phi_i $ defining the Slater-determinant wave function ϕ as follows:$ \begin{aligned}[b] & \rho_q({\bf{r}})=\sum\limits_{i, \sigma}\left|\phi_i({\bf{r}}, \sigma, q)\right|^2, \\ & \tau_q({\bf{r}})= \sum\limits_{i, \sigma}\left|\mathit{\boldsymbol{\nabla}} \phi_i({\bf{r}}, \sigma, q)\right|^2, \\ & {\bf{J}}_q({\bf{r}})=(-{\rm i}) \sum\limits_{i, \sigma, \sigma^{\prime}} \phi_i^*({\bf{r}}, \sigma, q)\left[\mathit{\boldsymbol{\nabla}} \phi_i\left({\bf{r}}, \sigma^{\prime}, q\right) \times\langle\sigma | \mathit{\boldsymbol{\sigma}}| \sigma^{\prime}\rangle\right] . \end{aligned} $
(6) The sums in above equations are taken over all occupied single-particle states. The exact expression for
$ H({\bf{r}}) $ is as follows [11]:$\begin{aligned}\\[-10pt] H({\bf{r}})= \frac{\hbar}{2m}\tau({\bf{r}})+ \frac{1}{2}t_0[ (1+\frac{1}{2}x_0) \rho^2- (x_0 +\frac{1}{2})(\rho_n^2 + \rho_p^2)] + \frac{1}{4} (t_1+t_2)\rho\tau \end{aligned} $ $ \begin{aligned}[b] &+\frac{1}{8}(t_2-t_1)(\rho_n \tau_n + \rho_p \tau_p) + \frac{1}{16}(t_2- 3t_1)\rho \nabla^2 \rho + \frac{1}{32}(3t_1 + t_2)(\rho_n \nabla^2 \rho_n + \rho_p \nabla^2 \rho_p) \\&+ \frac{1}{16}(t_1- t_2)({\bf{J}}_n^2 + {\bf{J}}_p^2)+ \frac{1}{4}t_3 \rho_n \rho_p \rho + H_C({\bf{r}})- \frac{1}{2}w_0(\rho \mathit{\boldsymbol{\nabla}} \cdot{\bf{J}} + \rho_n \mathit{\boldsymbol{\nabla}}\cdot{\bf{J}}_n + \rho_p \mathit{\boldsymbol{\nabla}} \cdot{\bf{J}}_p), \end{aligned} $
(7) where
$ \rho= \rho_n +\rho_p $ ,$ \tau= \tau_n+ \tau_p $ and$ {\bf{J}}={\bf{J}}_p+{\bf{J}}_p $ ; x0, t0,$t_1,\,t_2,\,t_3,\,w_0$ describe the parameterization of the nuclear force. The direct part of Coulomb interaction in$ H_C({\bf{r}}) $ is$ \dfrac{1}{2} V_C({\bf{r}}) \rho_p({\bf{r}}) $ , where$ \begin{equation} V_C({\bf{r}})=\int \rho_p({\bf{r}}^{\prime}) \frac{{\rm e}^2}{\left|{\bf{r}}-{\bf{r}}^{\prime}\right|} \rm d{\bf{r}}^{\prime}. \end{equation} $
(8) We refer to
$ V_C({\bf{r}}) $ as the Coulomb potential generated by protons, and one obtains the Coulomb potential of muons by replacing$ \rho_{p} $ with$ \rho_{\mu} $ . The Hartree-Fock equations for the Skyrme interaction are obtained by assuming that the total energy E is stationary with respect to individual variations of the single-particle states$ \phi_i $ , with the subsidiary condition that$ \phi_i $ are normalized:$ \begin{equation} \frac{\delta}{\delta \phi_i}\left(E-\sum\limits_i e_i \int\left|\phi_i({\bf{r}})\right|^2 \rm d^3 r\right)=0 . \end{equation} $
(9) It can be shown
$ \phi_i $ satisfy the following set of equations,$ \begin{equation} \left[ -\mathit{\boldsymbol{\nabla}} \cdot \frac{\hbar^2}{2 m^*_q({\bf{r}})}\mathit{\boldsymbol{\nabla}} + U_q({\bf{r}}) + {\bf{W}}_q({\bf{r}})\cdot (-{\rm i})(\mathit{\boldsymbol{\nabla}} \times \mathit{\boldsymbol{\sigma}})\right] \phi_i = e_i \phi_i. \end{equation} $
(10) Equation (10) involves an effective mass
$ m^*_q ({\bf{r}}) $ which depends on the density as follows:$ \begin{equation} \frac{\hbar^2}{2 m^*_q ({\bf{r}})}= \frac{\hbar^2}{2 m_q}+ \frac{1}{4}(t_1 +t_2)\rho + \frac{1}{8} (t_2 - t_1) \rho_q. \end{equation} $
(11) The potential
$ U_q({\bf{r}}) $ is expressed as follows:$ \begin{aligned}[b] U_q({\bf{r}})= & t_0[ (1 +\frac{1}{2}x_0)\rho- (x_0 +\frac{1}{2})\rho_q] + \frac{1}{4}t_3 (\rho^2 - \rho_q^2) \\&- \frac{1}{8} (3t_1 - t_2) \nabla^2 \rho + \frac{1}{16} (3t_1 + t_2) \nabla^2 \rho_q \\ & +\frac{1}{4} (t_1 + t_2) \tau + \frac{1}{8} (t_2 - t_1) \tau_q \\&- \frac{1}{2} W_0( \mathit{\boldsymbol{\nabla}}\cdot {\bf{J}} + \mathit{\boldsymbol{\nabla}}\cdot {\bf{J}}_q) + \delta_{q,+\frac{1}{2}}V_C({\bf{r}}). \end{aligned} $
(12) The form factor
$ {\bf{W}}_q({\bf{r}}) $ of the spin-orbit potential is expressed as follows:$ \begin{equation} {\bf{W}}_q({\bf{r}})= \frac{1}{2}W_0 (\mathit{\boldsymbol{\nabla}}\rho + \mathit{\boldsymbol{\nabla}}\rho_q)+ \frac{1}{8}(t_1- t_2){\bf{J}}_q({\bf{r}}). \end{equation} $
(13) We employed the force II parameterization from Ref. [11] for the Skyrme force in the numerical code developed. Specifically, we used the following parameter values:
$ x_0= 0.34 $ ,$ t_0 $ = –1169.9 MeV fm$ ^3 $ ,$ t_1 $ = 585.6 MeV fm$ ^5 $ ,$ t_2 $ = –27.1 MeV fm$ ^5 $ ,$ t_3 $ = 9331.1 MeV fm$ ^6 $ , and$ W_0 $ = 105 MeV fm$ ^5 $ . By successfully reproducing the outcomes reported in the aforementioned reference for various nuclei, we validated the reliability of the code.When binding a specific number of muons to the nuclei, it becomes imperative to incorporate the Coulomb potential contributed by the muons into the self-consistent mean field calculation for determining the single-particle orbit of protons. The mean field that governs the single-particle orbits of muons comprises the Coulomb potential generated by the protons and the Coulomb potential generated by the muons themselves; specifically, the single-particle states denoted as
$ \varphi_i({\bf{r}},\sigma) $ satisfy the following equations,$ \begin{equation} \left( -\frac{\hbar^2}{2 m_\mu} {\nabla}^2 - \int \frac{{\rm e}^2(\rho_p({\bf{r'}})-\rho_\mu({\bf{r'}}))}{|{\bf{r}} - {\bf{r'}}|} {\rm d} {\bf{r'}}\right) \varphi_i({\bf{r}},\sigma) = \varepsilon_i \varphi_i({\bf{r}},\sigma). \end{equation} $
(14) To obtain self-consistent results, we performed a series of numerical iterations until convergence was achieved. During each iteration, the updated potentials derived from the previous iteraction were employed to compute the single-particle orbits. From these orbits, the new single-particle densities were calculated and utilized to construct the updated potentials for the next iteration. This iterative process continued until the desired convergence was attained.
In the numerical implementation, we assumed spherical symmetry of the system so that the computation was reduced essentially to integrating the system along the radial direction. We used a lattice system to model the radial dimension; the lattice constant can be smaller than 0.08 fm while the radial size of the system can reach up to 60 fm.
Given our objective of numerically estimating the shift in proton chemical potentials resulting from the presence of muons, the simplifications made in our model can be justified. These simplifications include the assumption of spherical symmetry, preclusion of nuclear pairing interactions, and non-relativistic treament of muons.
Expanding proton dripline by employing a number of muons
- Received Date: 2023-07-05
- Available Online: 2023-11-15
Abstract: Through mean-field calculations, we demonstrate that, in a large Z nucleus binding multiple muons, these heavy leptons localize within a few dozen femtometers of the nucleus. The mutual Coulomb interactions between the muons and protons can lead to a substantial decrease in proton chemical potential, surpassing 1 MeV. These findings imply that, in principle, the proton-dripline can be expanded on the nuclear chart, suggesting the possible production of nuclei with Z around 120.