-
In the nuclear covariant energy density functional with a point-coupling interaction, the starting point is the following effective Lagrangian density [50],
$ \begin{aligned}[b] {\cal{L}} =& \bar\psi({\rm i}\gamma_\mu\partial^\mu-m)\psi -\frac{1}{2}\alpha_{\rm S}(\bar\psi\psi)(\bar\psi\psi) \\ & -\frac{1}{2}\alpha_{\rm V}(\bar\psi\gamma_\mu\psi)(\bar\psi\gamma^\mu\psi) -\frac{1}{2}\alpha_{\rm TV}(\bar\psi\vec\tau\gamma_\mu\psi)\cdot(\bar\psi\vec\tau\gamma^\mu\psi) \\ &-\frac{1}{3}\beta_{\rm S}(\bar\psi\psi)^3-\frac{1}{4}\gamma_{\rm S}(\bar\psi\psi)^4 -\frac{1}{4}\gamma_{\rm V}[(\bar\psi\gamma_\mu\psi)(\bar\psi\gamma^\mu\psi)]^2 \\ &-\frac{1}{2}\delta_{\rm S}\partial_\nu(\bar\psi\psi)\partial^\nu(\bar\psi\psi) -\frac{1}{2}\delta_{\rm V}\partial_\nu(\bar\psi\gamma_\mu\psi)\partial^\nu(\bar\psi\gamma^\mu\psi) \\ &-\frac{1}{2}\delta_{\rm TV}\partial_\nu(\bar\psi\vec\tau\gamma_\mu\psi)\cdot\partial^\nu(\bar\psi\vec\tau\gamma^\mu\psi)\\ &-\frac{1}{4}F^{\mu\nu}F_{\mu\nu} - e\bar\psi\gamma^\mu\frac{1-\tau_3}{2}\psi A_\mu, \end{aligned} $
(1) which is composed of a free nucleon term, four-fermion point-coupling terms, higher-order terms introduced for the effects of medium dependence, gradient terms to simulate the effects of a finite range, and electromagnetic interaction terms.
For the Lagrangian density
$ {\cal{L}} $ ,$ \psi $ is the Dirac spinor field of the nucleon with mass m,$ A_{\mu} $ and$ F_{\mu\nu} $ are respectively the four-vector potential and field strength tensor of the electromagnetic field, e is the charge unit for protons, and$ \vec{\tau} $ is an isospin vector with$ \tau_{3} $ being its third component. The subscripts S, V, and T in the coupling constants$ \alpha $ ,$ \beta $ ,$ \gamma $ , and$ \delta $ indicate the scalar, vector, and isovector couplings, respectively. The isovector-scalar (TS) channel is neglected owing to its small contributions to the description of nuclear ground state properties. In the full text, as a convention, we mark the isospin vectors with arrows and the space vectors in bold.In the framework of finite-temperature CDFT [48], the Dirac equation for a single nucleon reads
$ [\gamma_\mu({\rm i}\partial^\mu-V^\mu({{r}})) -(m+S({{r}}))]\psi_k({{r}}) = 0, $
(2) where
$ \psi_k $ is the Dirac spinor, and$\tag{3} S({{r}}) = \Sigma_{\rm S}, $
$\tag{4} V^\mu({{r}}) = \Sigma^\mu+\vec\tau\cdot\vec\Sigma^\mu_{\rm TV}, $
are respectively the scalar and vector potentials in terms of the isoscalar-scalar
$ \Sigma_{\rm S} $ , isoscalar-vector$ \Sigma^\mu $ , and isovector-vector$ \vec\Sigma^\mu_{\rm TV} $ self-energies,$\tag{5} \Sigma_{\rm S} = \alpha_{\rm S}\rho_{\rm S}+\beta_{\rm S}\rho^3_{\rm S}+\delta_{\rm S}\Delta\rho_{\rm S}, $
$\tag{6} \Sigma^\mu = \alpha_{\rm V}j^{\,\mu}_{\rm V}+\gamma_V(j^{\,\mu}_{\rm V})^3+\delta_{\rm V}\Delta j^{\,\mu}_V+eA^\mu, $
$ \tag{7} \vec\Sigma^\mu_{\rm TV} = \alpha_{\rm TV}\vec J^{\;\mu}_{\rm V}+\delta _{\rm TV}\Delta \vec J^{\;\mu}_{\rm V}. $
The isoscalar density
$ \rho_{\rm S} $ , isoscalar current$ j_{\rm V}^{\,\mu} $ , and isovector current$ \vec{j}_{\rm TV}^{\;\mu} $ are represented as follows:$\tag{8} \rho_{\rm S}({{r}}) = \sum_{k}\bar{\psi}_{k}({{r}})\psi_{k}({{r}})\left[v_{k}^{2}(1-2f_{k})+f_{k}\right], $
$\tag{9} j_{\rm V}^{\mu}({{r}}) = \sum_{k}\bar{\psi}_{k}({{r}})\gamma^{\mu}\psi_{k}({{r}})\left[v_{k}^{2}(1-2f_{k})+f_{k}\right], $
$ \tag{10} \vec{j}_{\rm TV}^{\mu}({{r}}) = \sum_{k}\bar{\psi}_{k}({{r}})\vec{\tau}\gamma^{\mu}\psi_{k}({{r}})\left[v_{k}^{2}(1-2f_{k})+f_{k}\right], $
where
$ \nu^2_k $ ($ \mu^2_k = 1-\nu^2_k $ ) is the BCS occupancy probability,$ \tag{11} \nu^2_k = \frac{1}{2}\left(1-\frac{\varepsilon_k-\lambda}{E_k}\right), $
$\tag{12} \mu^2_k = \frac{1}{2}\left(1+\frac{\varepsilon_k-\lambda}{E_k}\right), $
with
$ \lambda $ being the Fermi surface and$ E_k $ being the quasiparticle energy.At finite temperature, the occupation probability
$ \nu^2_k $ will be altered by the thermal occupation probability of quasiparticle states$ f_{k} $ , which is determined by temperature T as follows:$ f_{k} = \frac{1}{1+{\rm e}^{E_{k}/k_{\rm B}T}}, $
(7) where
$k_{\rm B}$ is the Boltzmann constant.In the BCS approach, the quasiparticle energy
$ E_k $ can be calculated as$ E_k = \sqrt{(\varepsilon_k-\lambda)^2+\Delta_k}, $
(8) where
$ \varepsilon_k $ is the single-particle energy, and the Fermi surface (chemical potential)$ \lambda $ is determined by meeting the conservation condition for particle number$ N_q $ ,$ N_q = 2\sum\limits_{k>0}\left[v_{k}^{2}(1-2f_{k})+f_k\right], $
(9) and
$ \Delta_k $ is the pairing energy gap, which satisfies the gap equation,$ \Delta_k = -\frac{1}{2}\sum\limits_{k'>0}V_{k\bar{k}k'\bar{k}'}^{pp}\frac{\Delta_{k'}}{E_{k'}}(1-2f_{k'}). $
(10) At finite temperature, the Dirac equation, mean-field potential, densities and currents, as well as the BCS gap equation in the CDFT, are solved iteratively on a harmonic oscillator basis. After a convergence is achieved, a single-particle spectrum up to
$ 30 $ MeV is extracted as an input to the following shell correction method. -
The shell corrections to the energy of a nucleus within the mean-field approximation is defined as
$ \delta E_{\rm{shell}} = E_{\rm S}-\widetilde{E}, $
(11) where
$ E_{\rm S} $ is the sum of the single-particle energy$ \varepsilon_k $ of the occupied states calculated with the exact density of states$ g_{\rm S}(\varepsilon) $ in an axially deformed space,$\tag{18} E_{\rm S} = \sum\limits_{\rm{occ.}}2\varepsilon_k = \int_{-\infty}^{\lambda}\varepsilon g_{\rm S}(\varepsilon){\rm d}\varepsilon, $
$\tag{19} g_{\rm S}(\varepsilon) = \sum\limits_k 2\delta (\varepsilon-\varepsilon_k), $
and
$ \widetilde{E} $ is the average energy calculated with the averaged density of states$ \widetilde{g}(\varepsilon) $ ,$\tag{20} \widetilde{E} = \int_{-\infty}^{\widetilde{\lambda}}\varepsilon \widetilde{g}(\varepsilon){\rm d}\varepsilon, $
$\tag{21} \widetilde{g}(\varepsilon) = \frac{1}{\gamma}\int_{-\infty}^{+\infty}f\left(\frac{\varepsilon'-\varepsilon}{\gamma}\right)g_{\rm S}(\varepsilon'){\rm d}\varepsilon', $
where
$ \widetilde{\lambda} $ is a smoothed Fermi surface,$ \gamma $ is the smoothing parameter, and$ f(x) $ is the Strutinsky smoothing function,$ f(x) = \frac{1}{\sqrt{\pi}}{\rm e}^{-x^2}L^{1/2}_M(x^2), $
(14) with
$ L^{1/2}_M(x^2) $ being the M-order generalized Laguerre polynomial.At finite temperature T, Eqs. (17)-(21) for the shell corrections can be generalized in a straightforward manner, i.e., [12],
$ \delta E_{\rm{shell}}(T) = E(T)-\widetilde{E}(T), $
(15) where for the energy
$ E(T) $ of a system of independent particles at finite temperature,$\tag{24} E(T) = \sum\limits_{\varepsilon_k}^{\lambda}2\varepsilon_k n_k^T, $
$\tag{25} n_k^T = \frac{1}{1+{\rm e}^{(\varepsilon_k-\lambda)/T}}. $
For the average energy
$ \widetilde{E}(T) $ ,$\tag{26} \widetilde{E}(T) = \int_{-\infty}^{\widetilde{\lambda}}\varepsilon \widetilde{g}(\varepsilon)n_{\varepsilon}^T{\rm d}\varepsilon, $
$\tag{27} n_{\varepsilon}^T = \frac{1}{1+{\rm e}^{(\varepsilon_k-\widetilde{\lambda})/T}}. $
The chemical potentials
$ \lambda $ and$ \widetilde{\lambda} $ are conserved by the number of neutrons (protons),$ \sum\limits_{k}2n_{k}^T = \int_{-\infty}^{\widetilde{\lambda}}{\rm d}\varepsilon\widetilde{g}(\varepsilon)n_{\varepsilon}^T = N_{q}. $
(18) The shell corrections to entropy S and free energy F at finite temperature read
$\tag{29} \delta S_{\rm{shell}}(T) = S(T)-\widetilde{S}(T), $
$\tag{30} \delta F_{\rm{shell}}(T) = F(T)-\widetilde{F}(T), $
and are related to each other as
$ \delta F_{\rm{shell}}(T) = \delta E_{\rm{shell}}(T)-T\delta S_{\rm{shell}}(T). $
(20) For the entropy
$ S_{\rm{shell}}(T) $ , the standard definition for the system of independent particles is adopted,$ S(T) = -k_{\rm B}\sum\limits_{k}2[n_k^T\ln n_k^T+(1-n_k^T)\ln (1-n_k^T)]. $
(21) The average part of
$ S(T) $ is defined in an analogous manner by replaying the sum in Eq. (32) by the integral,$ \widetilde{S}(T) = -k_{\rm B}\int_{-\infty}^{+\infty}\widetilde{g}(\varepsilon)[n_\varepsilon^T\ln n_\varepsilon^T+(1-n_\varepsilon^T)\ln (1-n_\varepsilon^T)]{\rm d}\varepsilon. $
(22) -
Taking the nucleus
$ ^{144} $ Sm with neutron shell closure as an example, the single-particle spectrum is calculated using the density functional PC-PK1 [50]. For the pairing correlation, the$ \delta $ pairing force$ V({{r}}) = V_q\delta({{r}}) $ is adopted, where the pairing strengths$ V_q $ are taken as$ -349.5 $ and$ -330.0 $ MeV$ \cdot $ fm$ ^3 $ for neutrons and protons, respectively. A smooth energy-dependent cutoff weight is introduced to simulate the effect of the finite range in the evaluation of the local pair density. Further details can be found in Ref. [48].At the mean-field level, the internal binding energies E at different axial-symmetric shapes can be obtained by applying constraints with a quadrupole deformation
$ \beta_2 $ ,$ \langle H'\rangle = \langle H\rangle +\frac{1}{2}C(\langle \hat{Q}_2\rangle-\mu_2)^2, $
(23) where C is a spring constant,
$ \mu_2 = \dfrac {3AR^2} {4\pi} \beta_2 $ is the given quadrupole moment with nuclear mass number A and radius R, and$ \langle \hat{Q}_2\rangle $ is the expectation value of quadrupole moment operator$ \hat{Q}_2 = 2r^2P_2(\cos\theta) $ .The free energy is evaluated by
$ F = E-TS $ . For convenience, the temperature used is$ k_{\rm B}T $ in units of MeV, and the entropy applied is$ S/k_{\rm B} $ , which is unitless.First, a numerical check of the binding energy convergence based on size is conducted. In Fig. 1, the average binding energy as a function of the major shell number of the harmonic oscillator basis
$ N_f $ is plotted. The binding energy is stable against the major shell number beginning from$ N_f = 16 $ and is thus fixed as a proper number. Further checks at different temperatures$ T = 0.0-2.0 $ MeV show that the temperature has a slight effect on the convergence.Figure 1. (color online) Average binding energy
$E_b/A$ as a function of the major shell number of the harmonic oscillator basis$N_f$ obtained by the finite temperature CDFT+BCS calculations using the PC-PK1 density functional at zero temperature.Second, the mandatory plateau condition for the shell correction method is checked. The shell correction energy should be insensitive to the smoothing parameter
$ \gamma $ and the order of the generalized Laguerre polynomial M, i.e.,$ \frac{\partial \delta E_{\rm{shell}}(T)}{\partial \gamma} = 0,\qquad\frac{\partial \delta E_{\rm{shell}}(T)}{\partial M} = 0. $
(24) In Fig. 2, the shell correction energy as a function of the above parameters
$ \gamma $ and M for$ ^{144} $ Sm is plotted. The unit of the smoothing range$ \gamma $ is$ \hbar \omega_0 = 41 A^{-1/3} (1\pm $ $\dfrac{1}{3}\dfrac{N-Z}{A}) $ MeV, where the plus (minus) sign holds for neutrons (protons). It can be seen from Fig. 2 that the optimal values are$ \gamma = 1.3 $ $ \hbar \omega_0 $ and$ M = 3 $ , which are consistent with previous relativistic calculations [35, 36].Figure 2. (color online) Neutron shell correction energy
$\delta E_{\rm{shell}}$ as a function of the smoothing parameter$\gamma$ and the order of the generalized Laguerre polynomial M for$^{144}$ Sm obtained by the finite temperature CDFT+BCS calculations using PC-PK1 density functional at zero temperature. The four different curves correspond to the orders M = 1, 2, 3, and 4, respectively. -
The free energy curves at temperatures 0, 0.4, 0.8, 1.2, 1.6 and 2.0 MeV for
$ ^{144} $ Sm are plotted in Fig. 3. The nucleus$ ^{144} $ Sm has spherical minima for all temperatures, which are consistent with the shell closure at neutron number$ N = 82 $ . The energy curve is hard against the deformation near the spherical region. In addition, at low temperatures, a local minimum occurs at approximately$ \beta_2 = 0.7 $ and a flat minimum occurs at approximately$ \beta_2 = -0.4 $ . However, it is shown that the fine details on the potential energy curves are washed out with increases in temperatures above T = 1.2 MeV, whereas the relative structures are maintained well at low temperatures.Figure 3. (color online) The relative free energy curves for
$^{144}$ Sm at different temperatures in the range of$0$ to$2$ MeV with a step of$0.4$ MeV obtained by the constrained CDFT+BCS calculations using the PC-PK1 energy density functional. The ground state free energy at zero temperature is set to zero and is shifted up by$4$ MeV for every$0.4$ MeV temperature rise.Furthermore, the shell corrections to the energy, entropy, and free energy as functions of quadrupole deformation
$ \beta_2 $ at various temperatures T are shown in Fig. 4. The shell correction to the energy$ \delta E_{\rm{shell}} $ shows a deep valley at the spherical region demonstrating a strong shell effect. In addition, the valley becomes deeper for$ T\leqslant 0.8 $ MeV and then shallower with increasing temperature, whereas the two peaks decrease dramatically after T = 0.4 MeV. The peaks and valleys on the$ \delta E_{\rm{shell}} $ curve are basically consistent with details of the free energy curve in Fig. 3. In Fig. 4(b), the entropy shell correction curve$ T \delta S_{\rm{shell}} $ changes slightly. The corresponding amplitudes are generally much smaller compared with those of$ \delta E_{\rm{shell}} $ . As the difference between$ \delta E_{\rm{shell}} $ and$ T \delta S_{\rm{shell}} $ , the curves of shell correction to the free energy$ \delta F_{\rm{shell}} $ in Fig. 4(c) have similar shapes as$ \delta E_{\rm{shell}} $ . By contrast, with increasing temperature, both the peaks and valleys of$ \delta F_{\rm{shell}} $ diminish gradually. Similar to the shell correction at zero temperature, applying a shell correction at finite temperature is a good way to quantify the shell effects, which provide rich information.Figure 4. (color online) Neutron shell corrections to the energy
$\delta E_{\rm{shell}}$ , entropy$T\delta S_{\rm{shell}}$ , and free energy$\delta F_{\rm{shell}}$ as functions of quadrupole deformation$\beta_2$ for$^{144}$ Sm at different temperatures from$0$ to$2$ MeV with steps of$0.4$ MeV obtained by the constrained CDFT+BCS calculations using PC-PK1 energy density functional.For the minimum states of
$ ^{144} $ Sm corresponding to increases in temperatures up to 4 MeV, the shell corrections to the energy$ \delta E_{\rm{shell}} $ , entropy$ T \delta S_{\rm{shell}} $ , and free energy$ \delta E_{\rm{shell}} $ are shown in Fig. 5. The non-monotonous behavior of$ \delta E_{\rm{shell}} $ with respect to temperature is significantly different from the exponential fading. The$ \delta E_{\rm{shell}} $ first decreases and then increases, monotonously approaching zero at high temperatures. This is consistent with the Woods-Saxon potential calculations carried out in Ref. [12]. In Ref. [8], a piecewise temperature-dependent factor is multiplied by the shell correction$ \delta E_{\rm{shell}} $ . The factor remains one for low temperatures below 1.65 MeV and then decreases exponentially. Here, the absolute amplitude first enlarges to approximately 120% at a temperature of 0.8 MeV and then bounces back to approximately 90% above 1.65 MeV. For this low temperature range, such behavior is roughly consistent with that in Ref. [8]. The exponential fading holds true for high temperatures for the current case and in Refs. [8] and [12].Figure 5. (color online) The temperature dependence of the shell corrections to the energy
$\delta E_{\rm{shell}}$ (black line), entropy$T\delta S_{\rm{shell}}$ (red line), and free energy$\delta F_{\rm{shell}}$ (blue line) with corresponding fitted empirical Bohr-Mottelson forms [7] (dashed lines) for the states with minimum free energy in$^{144}$ Sm shown in Fig. 3 obtained using the constrained CDFT+BCS calculations applying the PC-PK1 energy density functional.Because
$ \delta E_{\rm{shell}} $ is related to single-particle energy$ \varepsilon_k $ , Fermi surface$ \lambda $ , smoothed Fermi surface$ \widetilde{\lambda} $ , and temperature T according to Eqs. (23)-(27), the single particle levels near the neutron Fermi surface against the temperature for$ ^{144} $ Sm is plotted in Fig. 6. It is shown that the spectrum is almost constant within the region of$ T<0.8 $ MeV and only changes slightly at high temperature. Meanwhile, both the original Fermi surface$ \lambda $ and the smoothed surface$ \widetilde{\lambda} $ decrease synchronically with increasing temperatures. Thus, excluding$ \varepsilon_k $ ,$ \lambda $ , and$ \widetilde{\lambda} $ , the contribution directly from the temperature may play an important role in the behavior of the obtained shell correction to energy$ \delta E_{\rm{shell}} $ , as plotted in Fig. 5.Figure 6. (color online) Neutron single-particle levels as a function of temperature for
$^{144}$ Sm obtained by the constrained CDFT+BCS calculations using PC-PK1 energy density functional. The blue dashed line and red dash-dotted line represent the original and smoothed Fermi surfaces, respectively.The shell correction to the free energy
$ \delta F_{\rm{shell}} $ increases monotonously and approaches zero at high temperatures. The shell correction to the entropy$ T\delta S_{\rm{shell}} $ behaves similar to$ \delta E_{\rm{shell}} $ . For comparison, the fitted shell corrections to free energy$ \delta F_{\rm{shell}} $ and entropy$ T\delta S_{\rm{shell}} $ in the Bohr-Mottelson form [7] are also plotted as dashed lines in Fig. 5. The Bohr-Mottelson [7] form for the shell correction to the free energy$ \delta F_{\rm{shell}} $ is expressed as$ \delta F_{\rm{shell}}(T)/\delta F_{\rm{shell}}(0) = \Psi_{\rm BM}(T) = \frac{\tau}{\sinh(\tau)}, $
(25) where
$ \tau = c_0 \cdot 2\pi^2T/\hbar\omega_0 $ with the fitting parameter$ c_0 = 2.08 $ . Similar to$ \delta F_{\rm{shell}} $ ,$ T\delta S_{\rm{shell}} $ can also be approximated as$ T\delta S_{\rm{shell}}(T) / \delta F_{\rm{shell}}(0) = \frac{T \delta S_0 [\tau \coth(\tau)-1]}{\sinh(\tau)}, $
(26) when introducing the additional parameter
$ \delta S_0 = 2.15 $ . With these two empirical formula, the shell corrections to the energy$ \delta E_{\rm{shell}} $ as the sum of$ \delta F_{\rm{shell}} $ and$ T\delta S_{\rm{shell}} $ take the following form,$ \delta E_{\rm{shell}}(T) = \delta E_{\rm{shell}}(0) \frac{\tau +T \delta S_0 [\tau \coth(\tau)-1]}{\sinh(\tau)}, $
(27) noting that
$ \delta E_{\rm{shell}}(0) $ equals$ \delta F_{\rm{shell}}(0) $ . From Fig. 5, it can be clearly seen that both the shell corrections to the free energy$ \delta F_{\rm{shell}} $ and the entropy$ T \delta S_{\rm{shell}} $ can be approximated well using the Bohr-Mottelson forms.For more evidence, the same temperature dependence of the shell correction, for both neutrons and protons, is explored in other closed-shell nuclei. In Fig. 7, the shell corrections to the energy
$ \delta E_{\rm{shell}} $ , entropy$ T\delta S_{\rm{shell}} $ , and free energy$ \delta F_{\rm{shell}} $ in$ ^{100} $ Sn and$ ^{208} $ Pb with the corresponding fitted empirical Bohr-Mottelson forms are plotted. In general, the curve shapes for all quantities are extremely similar to those of$ ^{144} $ Sm in Fig. 5, proving the same temperature dependence. In addition, the fitting parameters$ c_0 $ for the neutron and proton shell corrections to the free energy$ \delta F_{\rm{shell}} $ of$ ^{100} $ Sn and$ ^{208} $ Pb are 1.90, 2.08, 2.24, and 2.28, respectively, which are close to those of$ ^{144} $ Sm 2.08. For the neutron and proton shell corrections to the entropy$ T\delta S_{\rm{shell}} $ , the values of parameter$ \delta S_0 $ are 1.78, 2.00, 2.23, and 2.16, respectively, which are close to those of$ ^{144} $ Sm 2.15. It was demonstrated that the Bohr-Mottelson forms describe well the shell corrections for closed-shell nuclei.Figure 7. (color online) Same as Fig. 5, but for neutrons and protons in
$^{100}$ Sn and$^{208}$ Pb.
Shell corrections with finite temperature covariant density functional theory
- Received Date: 2020-09-03
- Available Online: 2021-02-15
Abstract: The temperature dependence of the shell corrections to the energy