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

Microscopic collective inertial masses for nuclear reaction in the presence of nucleonic effective mass

Kai Wen wenkai@nucl.ph.tsukuba.ac.jp Center for Computational Sciences, University of Tsukuba, Tsukuba 305-8577, Japan    Takashi Nakatsukasa Center for Computational Sciences, University of Tsukuba, Tsukuba 305-8577, Japan Faculty of Pure and Applied Sciences, University of Tsukuba, Tsukuba 305-8571, Japan RIKEN Nishina Center, Wako 351-0198, Japan
Abstract

Collective inertial mass coefficients with respect to translational, relative, and rotational motions are microscopically calculated, along the collective reaction path self-consistently determined, based on the adiabatic self-consistent collective coordinate (ASCC) method. The impact of the time-odd component of the mean-field potential on the inertial masses are investigated. The results are compared with those calculated with the cranking formulae. The inertial masses based on the ASCC method reproduce the exact total nuclear mass for the translational motion as well as the exact reduced masses as the asymptotic values for the relative and rotational motions. In contrast, the cranking formulae fail to do so. This is due to the fact that the (local) Galilean invariance is properly restored in the ASCC method, but violated in the cranking formulae. A model Hamiltonian for low-energy nuclear reaction is constructed with the microscopically derived potentials and inertial masses. The astrophysical SS-factors are calculated, which indicates the importance of microscopic calculation of proper inertial masses.

pacs:
21.60.Ev, 21.10.Re, 21.60.Jz, 27.50.+e

I Introduction

Theoretical description of low-energy nuclear reaction with a solid microscopic foundation is still a challenging subject in nuclear physics. It may provide us with a deep insight into the reaction mechanism and the quantum dynamics of many-nucleon systems.

Starting from a microscopic many-body Hamiltonian (an energy density functional), the time-dependent Hartree-Fock (TDHF) method Negele (1982); Simenel (2012); Maruhn et al. (2014) has been established as a useful tool to study low-energy nuclear reactions. The TDHF real-time simulations have been intensively carried out with significant developments Nakatsukasa and Yabana (2005); Simenel (2012); Nakatsukasa et al. (2016); Sekizawa and Yabana (2016); Umar et al. (2016); Sekizawa (2017); Umar et al. (2017), including the time-dependent Hartree-Fock-Bogoliubov (TDHFB) calculations Ebata et al. (2010); Stetcu et al. (2011); Hashimoto (2012, 2013); Hashimoto and Scamps (2016); Nakatsukasa et al. (2016); Bulgac et al. (2016); Magierski et al. (2017); Scamps and Hashimoto (2019). In the TDHF(B) simulations, the reaction path is given by a time-dependent generalized Slater determinant which is determined by the TDHF(B) equations as an initial-value problem. It is successful in describing the microscopic dynamics in major reaction channels on average. However, it is known that important quantum effects are missing in the TDHF(B). For instance the sub-barrier fusion reaction and spontaneous fission can not be properly described within the real-time TDHF(B) simulations (Ring and Schuck, 1980; Negele, 1982; Nakatsukasa et al., 2016).

To recover the quantum fluctuations of nuclear collective dynamics that is missing in TDHF(B), we aim at the requantization of TDHF(B) on the collective subspace Nakatsukasa et al. (2016). First, we find a collective subspace spanned by a few selected collective canonical variables, which is well decoupled from the intrinsic excitations. The collective subspace and the collective variables can be extracted using the adiabatic self-consistent collective coordinate (ASCC) method Matsuo et al. (2000); Hinohara et al. (2007, 2009); Nakatsukasa (2012); Nakatsukasa et al. (2016). The ASCC is derived from the invariance principle of the TDHF(B) equation Marumori et al. (1980) under an assumption of adiabatic (slow) collective motion, namely, second-order expansion with respect to the collective momenta Matsuo et al. (2000). It also has a close connection with a theory of large amplitude collective motion developed in Refs. Klein et al. (1991); Dang et al. (2000). The collective subspace in the adiabatic regime is given by a series of time-even Slater determinants and generators of collective coordinates and momenta locally defined at each state. In the present paper, we apply the method to the low-energy nuclear reaction, to identify the optimal reaction path and the canonical variables. The requantization is performed on this self-consistently determined collective reaction path.

One of the key ingredients of the requantization procedure is the inertial mass parameters with respect to the collective coordinates. This is essential for the construction of the collective Hamiltonian. For the nuclear reaction, the relative distance RR between two colliding nuclei should be a proper choice of the collective coordinate in the asymptotic region (RR\rightarrow\infty). The corresponding inertial mass should be the reduced mass μred=APATm/(AP+AT)\mu_{\rm red}=A_{P}A_{T}m/(A_{P}+A_{T}), where APA_{P} (ATA_{T}) is the mass number of the projectile (target) nucleus and mm is the bare nucleon mass. Therefore, a theory can be tested by examining its asymptotic limit. On the other hand, the inertial mass in the interior region where two nuclei touch each other is highly nontrivial. Thus, a microscopic theory to calculate the mass on the entire reaction path is necessary.

A commonly used method to calculate the collective mass coefficient is the cranking formula (Baranger and Kumar, 1968a, b; Yuldashbaeva et al., 1999) which is derived based on the adiabatic perturbation theory. However, it is known that the cranking formula fails to reproduce the total mass of a whole nucleus for the translational motion, when the mean-field potential has velocity dependence Ring and Schuck (1980). This is due to violation of the Galilean invariance in the mean-field level which must be corrected by residual fields caused by the translational motion Bohr and Mottelson (1975). From this failure, we may expect that the cranking formula cannot reproduce the reduced mass for the relative motion of two nuclei at RR\rightarrow\infty. Since most of the realistic nuclear energy density functionals give effective mass, m/m<1m^{*}/m<1, the mean-field potential indeed has the velocity dependence. Therefore, it is highly desirable to find a proper method for the calculation of inertial masses. In this paper, we show that the ASCC method, which is capable of taking into account the residual effects missing in the cranking formula, provides a promising tool to microscopically calculate the proper inertial masses.

In our previous works Wen and Nakatsukasa (2016, 2017), we calculated the ASCC inertial masses for the relative motion between two nuclei, for the velocity-independent mean-field potential. We also examined those of the cranking formula, which turned out to be almost identical to the ASCC mass at RR\rightarrow\infty. However, this is not true in general. In this paper, we show that the asymptotic value in the cranking formula for a velocity-dependent potential does not agree with the reduced mass, while that in the ASCC method exactly reproduces it.

In general, the non-local mean-field potential produces the effective mass m/m1m^{*}/m\neq 1. For the energy density functionals of the Skyrme type, the effective mass comes from the t1t_{1} and t2t_{2} terms (momentum-dependent terms of the Skyrme interaction). In order to guarantee the Galilean invariance of the energy density functional, we need densities that are time-odd with respect to the time-reversal transformation. These time-odd densities vanish in the ground state of even-even nuclei. Therefore they do not contribute to the static mean-field potential. However, the time-odd mean fields may appear and play important roles in time-dependent dynamics, because the time-odd densities are non-zero in general for the time-dependent states. In this paper, we investigate the impact of the time-odd mean fields on the collective inertial masses in the context of low-energy reaction dynamics of light nuclei.

The paper is organized as follows: In Sec. II, we recapitulate the theoretical methods, mainly the ASCC method and cranking formulae, for the calculation of inertial masses. In Sec. III, in the presence of time-odd mean-field potential, we apply different methods to calculate the mass coefficients for the translational, relative and rotational motions, the applications are carried out for the reaction systems of α+α\alpha+\alpha\rightarrow 8Be, α+16\alpha+^{16}O \rightarrow 20Ne and 16O+16+^{16}O \rightarrow 32S. The impacts of inertial masses on the SS-factors of sub-barrier fusion are discussed. A summary and concluding remarks are given in Sec. IV.

II Theoretical methods

It is customary to use a model Hamiltonian for studies of low-energy nuclear reactions. The relative distance RR between the projectile and the target nuclei is assumed to be a dynamical coordinate, then, the Schrödinger equation is given in a form,

{12μredd2dR2+(+1)2μredR2+V(R)E}u(R)=0,\left\{-\frac{1}{2\mu_{\rm red}}\frac{d^{2}}{dR^{2}}+\frac{\ell(\ell+1)}{2\mu_{\rm red}R^{2}}+V(R)-E\right\}u_{\ell}(R)=0, (1)

where μred\mu_{\rm red} is the reduced mass, μred=APATm/(AP+AT)\mu_{\rm red}=A_{P}A_{T}m/(A_{P}+A_{T}). We use the unit =1\hbar=1 throughout this paper. Here, in addition to the assumption of the coordinate RR, the two nuclei are approximated as structureless “point” particles, which leads to the trivial inertial masses, μred\mu_{\rm red} and μredR2\mu_{\rm red}R^{2}. This approximation becomes exact in the asymptotic limit (RR\rightarrow\infty). However, at finite distance RR, even if the two nuclei are well apart, it is not trivial to justify the point-particle approximation. In fact, as we show later, the rigid-body values of the moments of inertia are significantly different from those of the point-particle approximation, 𝒥rigidμredR2{\cal J}_{\rm rigid}\neq\mu_{\rm red}R^{2}, even at R>RP+RTR>R_{P}+R_{T} where RPR_{P} (RTR_{T}) indicates the radius of the projectile (target) nucleus. It should be noted that the moments of inertia for deformed nuclei are given by the rigid-body values in the harmonic-oscillator-potential model Bohr and Mottelson (1975). Therefore, it is of great interest to examine how far the point-particle approximation can be justified.

To reveal the coordinate (RR) dependence and dynamical properties of the inertial masses, we apply the ASCC method to an energy density functional. In this section we recapitulate the ASCC method for the calculations of inertial masses. For comparison, we also give a brief description of the cranking formulae.

II.1 ASCC inertial masses

The ASCC method provides a collective subspace which describes nuclear reaction. The generators of the collective variables (q,p)(q,p) are locally given by (Q^(q),P^(q))(\hat{Q}(q),\hat{P}(q)) and determined by the ASCC equations. In the asymptotic region on the reaction path (collective subspace) at p=0p=0, these should correspond to the relative distance and its conjugate momentum. Thus, we expect that

Q^(q)|Ψ(q)R^|Ψ(q),P^(q)|Ψ(q)R|Ψ(q),\hat{Q}(q)|{\Psi(q)}\rangle\propto\hat{R}|{\Psi(q)}\rangle,\quad\hat{P}(q)|{\Psi(q)}\rangle\propto\frac{\partial}{\partial R}|{\Psi(q)}\rangle, (2)

where the proportional constants are arbitrary as far as they satisfy the weak canonicity condition

Ψ(q)|[iP^(q),Q^(q)]|Ψ(q)=1.\displaystyle\langle\Psi(q)|[i\hat{P}(q),\hat{Q}(q)]|\Psi(q)\rangle=1. (3)

In this paper, we study systems that preserve the axial symmetry on the reaction path. The one-body operator R^\hat{R} for the relative distance between projectile and target nuclei can be defined in the following way. The symmetry axis is chosen as the zz axis, assuming the projectile on the right (z>0z>0) and target on the left (z<0z<0). We introduce a separation plane at z=zsz=z_{s}, so that

𝑑𝐫θ(zzs)ρ(𝐫)=AP,\displaystyle\int d\mathbf{r}\ \theta(z-z_{s})\rho(\mathbf{r})=A_{P}, (4)

where ρ(𝐫)\rho(\mathbf{r}) is the nucleon density and θ(x)\theta(x) is the step function. The operator R^\hat{R} is given by R^Z^PZ^T\hat{R}\equiv\hat{Z}_{P}-\hat{Z}_{T}, with

Z^P(T)=s=,q=n,p𝑑𝐫ψ^sq(𝐫)ψ^sq(𝐫)zθ(±(zzs))AP(T),\hat{Z}_{P(T)}=\sum_{s=\uparrow,\downarrow}\sum_{q=n,p}\int d\mathbf{r}\hat{\psi}_{sq}^{\dagger}(\mathbf{r})\hat{\psi}_{sq}(\mathbf{r})z\frac{\theta(\pm(z-z_{\rm s}))}{A_{P(T)}}, (5)

where the upper sign is adopted for Z^P\hat{Z}_{P} and the lower for Z^T\hat{Z}_{T}.

The operator R^\hat{R} is artificially defined as a convenient measure for the relative distance. Its meaning is clear and well defined in the asymptotic region, though it is no longer meaningful when the projectile and the target are merged into a single nucleus. Nevertheless, this choice of R^\hat{R} does not affect the final result as far as there is a one-to-one correspondence between qq and RR, because the reaction path |Ψ(q)|{\Psi(q)}\rangle and its canonical variables (q,p)(q,p) are self-consistently determined without any assumption. After determining all the quantities in the collective Hamiltonian, we perform the coordinate transformation from qq to RR, by using the relation R(q)=Ψ(q)|R^|Ψ(q)R(q)=\langle{\Psi(q)}|\hat{R}|{\Psi(q)}\rangle. This is merely a change in the scale (and the dimension) of the coordinate.

In the ASCC method, the reaction path |Ψ(q)|{\Psi(q)}\rangle at p=0p=0 and the local generators (Q^(q),P^(q))(\hat{Q}(q),\hat{P}(q)) are determined by solving the ASCC equations

δΨ(q)|H^mv|Ψ(q)=0,\displaystyle\delta\langle\Psi(q)|\hat{H}_{\rm mv}|\Psi(q)\rangle=0, (6)
δΨ(q)|[H^mv,1iP^(q)]2V(q)q2Q^(q)|Ψ(q)=0,\displaystyle\delta\langle\Psi(q)|[\hat{H}_{\rm mv},\frac{1}{i}\hat{P}(q)]-\frac{\partial^{2}V(q)}{\partial q^{2}}\hat{Q}(q)|\Psi(q)\rangle=0, (7)
δΨ(q)|[H^mv,iQ^(q)]1M(q)P^(q)|Ψ(q)=0,\displaystyle\delta\langle\Psi(q)|[\hat{H}_{\rm mv},i\hat{Q}(q)]-\frac{1}{M(q)}\hat{P}(q)|\Psi(q)\rangle=0, (8)

where curvature terms are neglected for simplicity Matsuo et al. (2000). The Hamiltonian in the moving frame H^mv\hat{H}_{\rm mv} and the potential V(q)V(q) are respectively defined as

H^mvH^V(q)qQ^(q),V(q)Ψ(q)|H^|Ψ(q).\displaystyle\hat{H}_{\rm mv}\equiv\hat{H}-\frac{\partial V(q)}{\partial q}\hat{Q}(q),\quad V(q)\equiv\langle\Psi(q)|\hat{H}|\Psi(q)\rangle. (9)

V(q)V(q) in Eq. (9) is used as V(R)V(R) in Eq. (1) at R=Ψ(q)|R^|Ψ(q)R=\langle{\Psi(q)}|\hat{R}|{\Psi(q)}\rangle.

Equation (6) which depends on Q^(q)\hat{Q}(q) determines the reaction path |Ψ(q)|{\Psi(q)}\rangle. The local generators, P^(q)\hat{P}(q) and Q^(q)\hat{Q}(q), are given by a solution of Eqs. (7), and (8). However, since solutions of Eqs. (7) and (8) are not unique, we need to select the one corresponding to the reaction path we study. Normally, we choose a solution with the lowest frequency, ω2=(2V/q2)M1(q)\omega^{2}=(\partial^{2}V/\partial q^{2})M^{-1}(q), except for the Nambu-Goldstone (NG) modes. The generators should satisfy the asymptotic condition of Eq. (2).

In Eq. (8) there appears the inertial mass parameter M(q)M(q) with respect to the coordinate qq. The magnitude of M(q)M(q) depends on the scale of the coordinate qq which is arbitrary. We can choose to set the mass M(q)M(q) in Eq. (8) to be a constant, for instance Mq=1M_{q}=1 MeV-1[q]2[q]^{-2} without losing generality Wen and Nakatsukasa (2016), where [q][q] represents the unit of qq.

In order to obtain an intuitive picture of the collective dynamics, we map the collective coordinate qq to the relative distance RR between projectile and target nuclei. The inertial mass should be transformed as

M(R)=M(q)(dqdR)2=M(q)(dRdq)2,\displaystyle M(R)=M(q)\left(\frac{dq}{dR}\right)^{2}=M(q)\left(\frac{dR}{dq}\right)^{-2}, (10)

where the derivative dR/dqdR/dq can be obtained as

dRdq\displaystyle\frac{dR}{dq} =\displaystyle= ddqΨ(q)|R^|Ψ(q)=Ψ(q)|[R^,1iP^(q)]|Ψ(q)\displaystyle\frac{d}{dq}\langle\Psi(q)|\hat{R}|\Psi(q)\rangle=\langle\Psi(q)|[\hat{R},\frac{1}{i}\hat{P}(q)]|\Psi(q)\rangle (11)
=\displaystyle= 2np,jhRnj(q)Pnj(q),\displaystyle 2\sum_{n\in p,j\in h}R_{nj}(q)P_{nj}(q),

with the local generator P^(q)\hat{P}(q). Rnj(q)R_{nj}(q) and Pnj(q)P_{nj}(q) are the particle-hole (ph) matrix elements of R^(q)\hat{R}(q) and P^(q)\hat{P}(q), obtained by solving ASCC equations (7) and (8) Wen and Nakatsukasa (2016). The calculated mass M(R)M(R) will replace the constant mass μred\mu_{\rm red} in the first term of Eq. (1).

Apart from the the relative motion of our current interest, Eqs. (7) and (8) provide solutions for the NG modes, such as the translation and the rotation. These modes have natural and global generators, namely, the total linear momentum 𝐏^\hat{\mathbf{P}} for the translation and the total angular momentum 𝐉^\hat{\mathbf{J}} for the rotation. For example, the rotational motion around the xx axis can be generated by J^x\hat{J}_{x} without any cost of energy (ω2=0\omega^{2}=0), leading to the Thouless-Valatin equations Thouless and Valatin (1962); Ring and Schuck (1980),

δΨ(q)|[H^mv,J^x]|Ψ(q)=0,\displaystyle\delta\langle{\Psi(q)}|\left[\hat{H}_{\rm mv},\hat{J}_{x}\right]|{\Psi(q)}\rangle=0, (12)
δΨ(q)|[H^mv,iΘ^(q)]1𝒥(q)J^x|Ψ(q)=0,\displaystyle\delta\langle{\Psi(q)}|\left[\hat{H}_{\rm mv},i\hat{\Theta}(q)\right]-\frac{1}{{\cal J}(q)}\hat{J}_{x}|\Psi(q)\rangle=0, (13)
Ψ(q)|[iJ^x,Θ^(q)]|Ψ(q)=1,\displaystyle\langle{\Psi(q)}|[i\hat{J}_{x},\hat{\Theta}(q)]|{\Psi(q)}\rangle=1, (14)

where Θ^\hat{\Theta} is the angle variable conjugate to J^x\hat{J}_{x}, and 𝒥(q){\cal J}(q) is the moment of inertia around xx axis. Equations (12) and (13) indicate that (Q^(q),P^(q))=(Θ^(q),J^x)(\hat{Q}(q),\hat{P}(q))=(\hat{\Theta}(q),\hat{J}_{x}) correspond to a solution of the ASCC equations (7) and (8) with 2V/q2=0\partial^{2}V/\partial q^{2}=0 and M(q)=𝒥(q)M(q)={\cal J}(q). The calculated rotational moments of inertia can be used to replace μredR2\mu_{\rm red}R^{2} in the second term of Eq. (1).

To calculate the inertial masses for the NG modes, an efficient method was proposed in Ref. Hinohara (2015). The inertial mass of the NG modes M(q)M(q) are given by the zero-frequency limit of the strength function in the linear response for the momentum operator P^\hat{P}. We apply this method to the calculations of the translational and rotational inertial masses.

II.2 CHF+cranking method

A simple method based on the cranking formula to calculate the inertial mass is widely used for nuclear collective motion. The collective path (reaction path) is usually produced by the constrained-Hartree-Fock (CHF) calculation with a constraining operator C^\hat{C} given by hand. The CHF states are given by the variation

δΨ0(λ)|H^λ|Ψ0(λ)=0,H^λH^λC^,\delta\langle{\Psi_{0}(\lambda)}|\hat{H}_{\lambda}|{\Psi_{0}(\lambda)}\rangle=0,\quad\hat{H}_{\lambda}\equiv\hat{H}-\lambda\hat{C}, (15)

where λ\lambda is the Lagrange multiplier. Upon the CHF states |Ψ0(R)=|Ψ0(λ)|{\Psi_{0}(R)}\rangle=|{\Psi_{0}(\lambda)}\rangle with R=Ψ0(λ)|R^|Ψ0(λ)R=\langle{\Psi_{0}(\lambda)}|\hat{R}|{\Psi_{0}(\lambda)}\rangle, the inertial mass with respect to the coordinate RR is calculated using the cranking formula Baran et al. (2011):

Mcrnp(R)\displaystyle M_{\rm cr}^{\rm np}(R) =\displaystyle= 2m|Ψm(R)|/R|Ψ0(R)|2Em(R)E0(R)\displaystyle 2\sum_{m}\frac{|\langle{\Psi_{m}(R)}|\partial/\partial R|{\Psi_{0}(R)}\rangle|^{2}}{E_{m}(R)-E_{0}(R)} (16)
=\displaystyle= 2np,jh|φn(R)|/R|φj(R)|2en(R)ej(R),\displaystyle 2\sum_{n\in p,j\in h}\frac{|\langle{\varphi_{n}(R)}|\partial/\partial R|{\varphi_{j}(R)}\rangle|^{2}}{e_{n}(R)-e_{j}(R)}, (17)

where the single-particle states |φμ(R)|{\varphi_{\mu}(R)}\rangle and energies eμ(R)e_{\mu}(R) are defined with respect to the CHF single-particle Hamiltonian, h^λ=h^HF[ρ]λC^\hat{h}_{\lambda}=\hat{h}_{\rm HF}[\rho]-\lambda\hat{C}, as

h^λ|φμ(R)=eμ(R))|φμ(R),μp,h.\hat{h}_{\lambda}|{\varphi_{\mu}(R)}\rangle=e_{\mu}(R))|{\varphi_{\mu}(R)}\rangle,\quad\mu\in p,h. (18)

Note that the Lagrange multiplier λ\lambda is a function of RR given by the condition, R=Ψ0(λ)|R^|Ψ0(λ)R=\langle{\Psi_{0}(\lambda)}|\hat{R}|{\Psi_{0}(\lambda)}\rangle. The constraining operator C^\hat{C} can be in general different from the relative-distance operator R^\hat{R}. In the following calculations in Sec. III, we use the mass quadrupole and octupole operators as C^\hat{C}.

Another cranking formula, which is even more frequently used in many applications, can be derived by neglecting the rearranged fields induced by the change of λ\lambda, namely, δhHF/δρdρ/dλ=0\delta h_{\rm HF}/\delta\rho\cdot d\rho/d\lambda=0. Normally, one chooses the constraining operator as the collective coordinate, C^=R^\hat{C}=\hat{R}, for the CHF calculation. However, in the present calculation, since we need to find the separation plane z=zsz=z_{s} to determine the operator R^\hat{R}, it is convenient to adopt a different one-body operator C^\hat{C}. Let us derive the formula for this case. From Eq. (18) and the orthonormality condition φμ(R)|φν(R)=δμν\langle\varphi_{\mu}(R)|\varphi_{\nu}(R)\rangle=\delta_{\mu\nu}, it is easy to find

φn|R|φj=φn|hλ/R|φjenejdλdRφn|C^|φjenej.\langle{\varphi_{n}}|\frac{\partial}{\partial R}|{\varphi_{j}}\rangle=-\frac{\langle{\varphi_{n}}|\partial h_{\lambda}/\partial R|{\varphi_{j}}\rangle}{e_{n}-e_{j}}\approx\frac{d\lambda}{dR}\frac{\langle{\varphi_{n}}|\hat{C}|{\varphi_{j}}\rangle}{e_{n}-e_{j}}. (19)

Neglecting the rearrangement, h^λ+Δλh^λΔλC^\hat{h}_{\lambda+\Delta\lambda}\approx\hat{h}_{\lambda}-\Delta\lambda\hat{C}, the derivative dR/dλdR/d\lambda is estimated with the first-order perturbation as dR/dλ=2S1(R^,C^){dR}/{d\lambda}=2S_{1}(\hat{R},\hat{C}), where

Sk(Q^1,Q^2)\displaystyle S_{k}(\hat{Q}_{1},\hat{Q}_{2}) \displaystyle\equiv 12m>0Ψ0|Q^1|ΨmΨm|Q^2|Ψ0+c.c.(EmE0)k\displaystyle\frac{1}{2}\sum_{m>0}\frac{\langle{\Psi_{0}}|\hat{Q}_{1}|{\Psi_{m}}\rangle\langle{\Psi_{m}}|\hat{Q}_{2}|{\Psi_{0}}\rangle+\mbox{c.c.}}{\left(E_{m}-E_{0}\right)^{k}} (20)
=\displaystyle= 12np,jhφj|Q^1|φnφn|Q^2|φj+c.c.(enej)k\displaystyle\frac{1}{2}\sum_{n\in p,j\in h}\frac{\langle{\varphi_{j}}|\hat{Q}_{1}|{\varphi_{n}}\rangle\langle{\varphi_{n}}|\hat{Q}_{2}|{\varphi_{j}}\rangle+\mbox{c.c.}}{\left(e_{n}-e_{j}\right)^{k}}
=\displaystyle= Sk(Q^2,Q^1).\displaystyle S_{k}(\hat{Q}_{2},\hat{Q}_{1}).

Thus, the formula reads

Mcrp(R)=12S3(C^,C^){S1(R^,C^)}2.M_{\rm cr}^{\rm p}(R)=\frac{1}{2}S_{3}(\hat{C},\hat{C})\left\{S_{1}(\hat{R},\hat{C})\right\}^{-2}. (21)

According to Ref. Baran et al. (2011), we call the former one in Eq. (17) “non-perturbative” cranking inertia and the latter in Eq. (21) “perturbative” one. In contrast to the ASCC mass, the cranking masses of Eqs. (17) and (21) have a drawback that they both neglect residual effect of the time-odd density fluctuation. As we shall see in Sec. III, when the nucleonic effective masses are present, the cranking formulae produce wrong masses for the translation and for the relative motion at RR\rightarrow\infty.

For the NG modes such as the translation and the rotation, we know the generator of the collective coordinate. For instance, in the translational case, the center-of-mass (COM) coordinates XkX_{k} (k=x,y,zk=x,y,z) are the trivial collective coordinates. Then, the non-perturbative cranking formula (17) can be transformed into

Mcrnp(Xk)\displaystyle M_{\rm cr}^{\rm np}(X_{k}) =\displaystyle= 2np,jh|φn(Xk)|/Xk|φj(Xk)|2en(Xk)ej(Xk)\displaystyle 2\sum_{n\in p,j\in h}\frac{|\langle{\varphi_{n}(X_{k})}|\partial/\partial X_{k}|{\varphi_{j}(X_{k})}\rangle|^{2}}{e_{n}(X_{k})-e_{j}(X_{k})} (22)
=\displaystyle= 2np,jh|φn|p^k|φj|2enej,\displaystyle 2\sum_{n\in p,j\in h}\frac{|\langle{\varphi_{n}}|\hat{p}_{k}|{\varphi_{j}}\rangle|^{2}}{e_{n}-e_{j}}, (23)

where p^k\hat{p}_{k} is the single-particle linear momentum. In the second equation, we take advantage of the fact that the single-particle energies and the wave functions relative to the COM are invariant with respect to its position XkX_{k}. For the rotation, we may replace XkX_{k} by an angle ϕk\phi_{k} and p^k\hat{p}_{k} by the angular momentum j^k\hat{j}_{k}. This leads to nothing but the cranking formula for the moment of inertia, originally proposed by Inglis Inglis (1954),

(𝒥crnp)k=2np,jh|φn|j^k|φj|2enej.\left({\cal J}_{\rm cr}^{\rm np}\right)_{k}=2\sum_{n\in p,j\in h}\frac{|\langle{\varphi_{n}}|\hat{j}_{k}|{\varphi_{j}}\rangle|^{2}}{e_{n}-e_{j}}. (24)

It is instructive to investigate properties of the cranking mass formula for the NG modes Ring and Schuck (1980); Bohr and Mottelson (1975); Thouless and Valatin (1962). Here, let us examine the formula of Eq. (23) for the translation, following the argument given in Ref. Bohr and Mottelson (1975). The summation with respect to the index nn in Eq. (23) is restricted to the particle states. This restriction can be removed, because the summation with respect to the hole-hole components, nhjh\sum_{n\in h}\sum_{j\in h}, identically vanishes. When the mean-field (Hartree-Fock) Hamiltonian h^HF[ρ]\hat{h}_{\rm HF}[\rho] conserves the Galilean invariance,

[h^HF,x^k]=imp^k,k=x,y,z,\left[\hat{h}_{\rm HF},\hat{x}_{k}\right]=\frac{-i}{m}\hat{p}_{k},\quad k=x,y,z, (25)

the translational mass becomes

Mcrnp\displaystyle M_{\rm cr}^{\rm np} =\displaystyle= 2μjhφj|p^k|φμφμ|p^k|φjeμej\displaystyle 2\sum_{\mu}\sum_{j\in h}\frac{\langle{\varphi_{j}}|\hat{p}_{k}|{\varphi_{\mu}}\rangle\langle{\varphi_{\mu}}|\hat{p}_{k}|{\varphi_{j}}\rangle}{e_{\mu}-e_{j}} (26)
=\displaystyle= μjh(φj|im[h^HF,x^k]|φμφμ|p^k|φj\displaystyle\sum_{\mu}\sum_{j\in h}\left(\langle{\varphi_{j}}|im\left[\hat{h}_{\rm HF},\hat{x}_{k}\right]|{\varphi_{\mu}}\rangle\langle{\varphi_{\mu}}|\hat{p}_{k}|{\varphi_{j}}\rangle\right.
+φj|p^k|φμφμ|im[h^HF,x^k]|φj)1eμej\displaystyle\left.+\langle{\varphi_{j}}|\hat{p}_{k}|{\varphi_{\mu}}\rangle\langle{\varphi_{\mu}}|im\left[\hat{h}_{\rm HF},\hat{x}_{k}\right]|{\varphi_{j}}\rangle\right)\frac{1}{e_{\mu}-e_{j}}
=\displaystyle= imjhφj|[p^k,x^k]|φj=Am.\displaystyle im\sum_{j\in h}\langle{\varphi_{j}}|\left[\hat{p}_{k},\hat{x}_{k}\right]|{\varphi_{j}}\rangle=Am.

Therefore, the Galilean invariance of the mean-field Hamiltonian guarantees that the cranking formula reproduces the total mass AmAm for the translation. However, the Galilean invariance is violated by velocity dependence of the one-body mean-field potential in most of nuclear energy density functionals. This results in a wrong mass McrAmM_{\rm cr}\neq Am in the cranking formula. This violation should be corrected by the residual fields that depend on the velocity of the translational motion Bohr and Mottelson (1975). Indeed, the ASCC mass reproduces the exact total mass even when the mean-field potential violates the Galilean invariance, which will be shown in Sec. III.

III Application

In the following numerical calculations, in order to express the orbital wave functions, the grid representation is employed, discretizing the rectangular box into three-dimensional (3D) Cartesian mesh Nakatsukasa and Yabana (2005). The model space is set to be 12×12×1812\times 12\times 18 fm3 for the systems α\alpha+α\alpha \rightarrow 8Be and 16O+α\alpha \rightarrow 20Ne. It is 12×12×2412\times 12\times 24 fm3 for 16O+16O \rightarrow 32S. The mesh size is set to be 1.0 fm for the system α\alpha+α\alpha \rightarrow 8Be and 1.1 fm for the other two systems.

For numerical calculations of the ASCC method, we use the the finite amplitude method (FAM) (Nakatsukasa et al., 2007; Inakura et al., 2009; Avogadro and Nakatsukasa, 2011, 2013). The FAM provides us with high numerical efficiency with simple computer programs, because only the calculations of the mean-field (single-particle) Hamiltonian constructed with independent bra and ket states are required Nakatsukasa et al. (2007). The matrix FAM (m-FAM) prescription Avogadro and Nakatsukasa (2013) is adopted to solve the moving RPA equations (7) and (8). On the other hand, the iterative FAM (i-FAM) (Nakatsukasa et al., 2007; Inakura et al., 2009; Avogadro and Nakatsukasa, 2011) is adopted to calculate the response functions for the NG modes. The moving mean-field equation (6) is solved by using the imaginary-time method (Davies et al., 1980).

III.1 Modified BKN energy density functional

In order to investigate the effect of this time-odd mean-field potential on the collective inertial masses, we adopt an energy density functional of the simplest choice, namely, the BKN energy density functional Bonche et al. (1976) with the minimum extension.

The original BKN functional assumes the spin-isospin symmetry without the spin-orbit interaction, which is a functional of the isoscalar kinetic and local densities, τ(𝐫)\tau(\mathbf{r}) and ρ(𝐫)\rho(\mathbf{r}), only. The mean-field potential is local and has no velocity dependence. Thus, the nucleon’s effective mass is identical to the bare nucleon mass. However, in most of realistic energy density functionals, the effective mass is smaller than the bare mass, typically m/m0.7m^{*}/m\sim 0.7. In order to introduce the effective mass, we extend the energy density by adding terms ρτ𝐣2\rho\tau-\mathbf{j}^{2} where 𝐣(𝐫)\mathbf{j}(\mathbf{r}) is the isoscalar current density. The appearance of the current density is necessary to preserve the Galilean invariance.

The modified BKN energy density functional reads,

E[ρ]\displaystyle E[\rho] =\displaystyle= 12mτ(𝐫)𝑑𝐫+𝑑𝐫{38t0ρ2(𝐫)+116t3ρ3(𝐫)}\displaystyle\int\frac{1}{2m}\tau(\mathbf{r})d\mathbf{r}+\int d\mathbf{r}\left\{\frac{3}{8}t_{0}\rho^{2}(\mathbf{r})+\frac{1}{16}t_{3}\rho^{3}(\mathbf{r})\right\} (27)
+𝑑𝐫𝑑𝐫ρ(𝐫)v(𝐫𝐫)ρ(𝐫)\displaystyle+\int\int d\mathbf{r}d\mathbf{r}^{\prime}\rho(\mathbf{r})v(\mathbf{r}-\mathbf{r}^{\prime})\rho(\mathbf{r}^{\prime})
+B3𝑑𝐫{ρ(𝐫)τ(𝐫)𝐣2(𝐫)},\displaystyle+B_{3}\int d\mathbf{r}\left\{\rho(\mathbf{r})\tau(\mathbf{r})-\mathbf{j}^{2}(\mathbf{r})\right\},

where ρ(𝐫)\rho(\mathbf{r}), τ(𝐫)\tau(\mathbf{r}), and 𝐣(𝐫)\mathbf{j}(\mathbf{r}) are calculated as

ρ(𝐫)\displaystyle\rho(\mathbf{r}) =\displaystyle= 4j=1A/4|ψj(𝐫)|2,τ(𝐫)=4j=1A/4|ψj(𝐫)|2,\displaystyle 4\sum_{j=1}^{A/4}|\psi_{j}(\mathbf{r})|^{2},\quad\tau(\mathbf{r})=4\sum_{j=1}^{A/4}|\nabla\psi_{j}(\mathbf{r})|^{2},
j(𝐫)\displaystyle\textbf{j}(\mathbf{r}) =\displaystyle= 42ij=1A/4{ψj(𝐫)ψj(𝐫)ψj(𝐫)ψj(𝐫)}.\displaystyle\frac{4}{2i}\sum_{j=1}^{A/4}\left\{\psi_{j}^{*}(\mathbf{r})\nabla\psi_{j}(\mathbf{r})-\psi_{j}(\mathbf{r})\nabla\psi_{j}^{*}(\mathbf{r})\right\}.

In Eq. (27), v(𝐫)v(\mathbf{r}) is the sum of the Yukawa and the Coulomb potentials (Bonche et al., 1976),

v(𝐫)V0aer/ar+(e/2)2r.\displaystyle v(\mathbf{r})\equiv V_{0}a\frac{e^{-r/a}}{r}+\frac{(e/2)^{2}}{r}.

The new parameter B3B_{3} controls the effective mass and the velocity dependence of the mean-field potential.

The variation of the total energy with respect to the density (or equivalently single-particle wave functions) defines the mean-field (Hartree-Fock) Hamiltonian,

h^HF(𝐫)\displaystyle\hat{h}_{\rm HF}(\mathbf{r}) =\displaystyle= 12m(𝐫)+34t0ρ(𝐫)+316t3ρ2(𝐫)\displaystyle-\nabla\frac{1}{2m^{*}(\mathbf{r})}\nabla+\frac{3}{4}t_{0}\rho(\mathbf{r})+\frac{3}{16}t_{3}\rho^{2}(\mathbf{r}) (28)
+𝑑𝐫v(𝐫𝐫)ρ(𝐫)+B3(τ(𝐫)+i𝐣(𝐫))\displaystyle+\int d\mathbf{r^{\prime}}v(\mathbf{r}-\mathbf{r^{\prime}})\rho(\mathbf{r^{\prime}})+B_{3}(\tau(\mathbf{r})+i\nabla\cdot\mathbf{j}(\mathbf{r}))
+2iB3𝐣(𝐫),\displaystyle+2iB_{3}\mathbf{j}(\mathbf{r})\cdot\nabla,

where the effective mass is now deviated from bare nucleon mass,

12m(𝐫)=12m+B3ρ(𝐫).\displaystyle\frac{1}{2m^{*}(\mathbf{r})}=\frac{1}{2m}+B_{3}\rho(\mathbf{r}). (29)

For B30B_{3}\neq 0, Eq. (28) indicates the velocity (momentum) dependence of the mean-field potential and the presence of the time-odd mean fields, iB3(𝐣+2𝐣)iB_{3}(\nabla\cdot\mathbf{j}+2\mathbf{j}\cdot\nabla). For the time-even states, such as the ground state of even-even nuclei, the current density disappears, 𝐣(𝐫)=0\mathbf{j}(\mathbf{r})=0. Nevertheless, as will be shown later, the terms associated with 𝐣(𝐫)\mathbf{j}(\mathbf{r}) play an important role in the collective inertial mass.

The parameters, t0t_{0}, t3t_{3}, V0V_{0}, and aa are adopted from Ref. (Bonche et al., 1976), and we vary B3B_{3} to change the effective mass and the time-odd mean fields.

III.2 Inertial massses for translational motion: Alpha particle

First, we demonstrate the importance of the time-odd mean fields, taking the translational total mass as a trivial example. We adopt the simplest case, namely, the single alpha particle.

Refer to caption
Figure 1: (Color online) Translational mass of the α\alpha particle in unit of the nucleon mass as a function of the parameter B3B_{3}. The solid (red), dashed (green), and dash-dotted (blue) curves show the results of the ASCC, the non-perturbative cranking formula (23) and AA times the average effective mass, respectively. See the text for details.

In Fig. 1, we show the translational mass of a single alpha particle as a function of B3B_{3}. Increasing B3B_{3}, the effective mass mm^{*} decreases, as we see in Eq. (29). In the present case, we use the center-of-mass coordinate 𝐑cm\mathbf{R}_{\rm cm} as the collective coordinate RR in Sec. II.1 and II.2. Since the system is isotropic, we use its zz component ZcmZ_{\rm cm} in the numerical calculation. In the present model, neutrons and protons have the identical mass mm. Therefore, the total inertial mass of an alpha particle should be equal to 4m4m.

It is clearly demonstrated that the ASCC always reproduces the correct total mass AmAm, irrespective of values of the parameter B3B_{3}. The non-perturbative cranking mass reproduces the total mass, Mcrnp=4mM^{\rm np}_{\rm cr}=4m, at B3=0B_{3}=0, which agree with Eq. (26). This is due to the velocity-independent mean-field potential, for which the mean-field Hamiltonian h^HF\hat{h}_{\rm HF} preserves the Galilean invariance. However, at B3>0B_{3}>0, the effective mass becomes smaller than the bare nucleon mass, m<mm^{*}<m, inside the nucleus with ρ(𝐫)>0\rho(\mathbf{r})>0. This leads to the velocity-dependent potential, and the violation of the Galilean invariance in the mean-field level. Figure 1 indicates Mcrnp<4mM_{\rm cr}^{\rm np}<4m at B3>0B_{3}>0, and shows a monotonic decrease with respect to B3B_{3}. We calculate the average effective mass defined by

(mm)av12mAρ(𝐫)2m(𝐫)𝑑𝐫.\left(\frac{m^{*}}{m}\right)_{\rm av}^{-1}\equiv\frac{2m}{A}\int\frac{\rho(\mathbf{r})}{2m^{*}(\mathbf{r})}d\mathbf{r}. (30)

It turns out that the cranking mass for the translation approximately coincides with Am×(m/m)avAm\times(m^{*}/m)_{\rm av} as shown in Fig. 1.

For the translational motion, the ASCC inertial mass becomes identical to the Thouless-Valatin moments of inertia Thouless and Valatin (1962) in Eqs. (12) and (13), with (J^x,Θ^)(\hat{J}_{x},\hat{\Theta}) being replaced by (P^z,Z^cm)(\hat{P}_{z},\hat{Z}_{\rm cm}). The fact that the ASCC inertial mass agrees with the total mass AmAm means that the Galilean invariance, violated in the mean-field level, is properly restored by the residual induced fields in the leading order. This shows a clear advantageous feature of the ASCC mass over the cranking formulae.

III.3 Inertial massses for relative motions

III.3.1 α\alpha+α\alpha 8\rightarrow^{8}Be

In this section we study how the time-odd mean-field changes the inertial mass for the relative motion between two alpha particles. The reaction path was obtained and shown in our former paper Wen and Nakatsukasa (2016) for B3=0B_{3}=0. With B30B_{3}\neq 0, basic features of the reaction path is the same, although the ground state of 8Be becomes more elongated, developing a prominent 2α2\alpha cluster structure. The curves in Fig. 2 show the inertial mass and total energy for the relative motion between two alpha particles as a function of the relative distance RR. The ASCC mass is calculated based on the self-consistent ASCC collective path. For the cranking masses, the mass quadrupole operator Q^20\hat{Q}_{20} is used as a constraint to construct the reaction path with the CHF calculation.

Refer to caption
Figure 2: (Color online) Inertial mass M(R)M(R) for the reaction α+α\alpha+\alpha as a function of the relative distance RR. The top panel shows the results of the ASCC method, while the middle panel shows those of the perturbative and non-perturbative cranking formulae McrpM_{\rm cr}^{\rm p} (thin lines) and McrnpM_{\rm cr}^{\rm np} (thick lines). The bottom panel shows the potential energies along the ASCC collective paths. Solid (red) and dashed (blue) lines indicate those of B3=0B_{3}=0 and 75 MeV fm5, respectively. In the upper panel, the dots correspond to the positions of the minimum energy in the bottom panel.

At large RR where the two alpha particles can be approximated as point particles, we expect that the inertial mass with respect to RR is identical to the reduced mass, 2m2m. For the non-perturbative cranking mass, this is true at B3=0B_{3}=0, while it monotonically decreases as B3B_{3} increases (Fig. 3). On the other hand, the ASCC mass reproduces the correct reduced mass at large RR, irrespective of the value of B3B_{3}. The main difference between the ASCC and the cranking masses is the inclusion of the effect of time-odd residual fields. Therefore, we may conclude that the time-odd residual effect is essential to reproduce the reduced mass in the asymptotic region of RR\rightarrow\infty.

The perturbative cranking mass seems to be unable to reproduce the correct reduced mass. It is larger than μred=2m\mu_{\rm red}=2m even for B3=0B_{3}=0. As is seen in the first equation of (19), the non-perturbative cranking mass comes from the RR-dependence of the Lagrange multiplier λ\lambda and from that of the HF Hamiltonian (rearrangement) h^HF\hat{h}_{\rm HF}. When we take the asymptotic limit of RR\rightarrow\infty, since the interaction between two α\alpha particles vanishes, we naturally expect λ0\lambda\rightarrow 0. Then, we have dλ/dR0d\lambda/dR\rightarrow 0, and McrnpM^{\rm np}_{\rm cr} solely comes from the rearrangement effect. In the perturbative treatment, we neglect the rearrangement in the last equality of Eq. (19). Instead, we replace dλ/dRd\lambda/dR by the perturbative value (2S1(R^,C^))1(2S_{1}(\hat{R},\hat{C}))^{-1}. It is thus difficult to justify the perturbative approximation in the asymptotic limit.

Refer to caption
Figure 3: (Color online) Non-perturbative cranking inertial masses calculated at R=8R=8 fm for the relative motion between two α\alpha particles divided by the reduced mass μred=2m\mu_{\rm red}=2m, as a function of B3B_{3}. The blue dash-dotted curve indicates (m/m)av(m^{*}/m)_{\rm av} of Eq. (30).

Figure 3 clearly demonstrates that the cranking inertial mass at B30B_{3}\neq 0 underestimates the exact reduced mass for the relative motion (μred=2m\mu_{\rm red}=2m). Again, it well agrees with the average effective mass, 2(m)av2(m^{*})_{\rm av}. Thus, the failure of the cranking mass is due to the violation of the Galilean invariance.

III.3.2 α\alpha+16O 20\rightarrow^{20}Ne and 16O+16O 32\rightarrow^{32}S

Next we show the inertial masses M(R)M(R) of the relative motion for two reaction systems α\alpha+16O 20\rightarrow^{20}Ne and 16O+16O 32\rightarrow^{32}S. The self-consistent reaction paths for these were presented in our former paper Wen and Nakatsukasa (2017), but only for B3=0B_{3}=0. Similar to the α+α8\alpha+\alpha\rightarrow^{8}Be case, the increase of B3B_{3} favors larger deformation. The upper panels in Figs. 4 and 5 show M(R)M(R) for α\alpha+16O \rightarrow 20Ne and 16O+16O \rightarrow 32S, respectively. The smaller effective mass enlarges the size of the nucleus, thus, the touching points between the two nuclei take place at larger RR. The minimum energy position is also shifted to larger RR for larger B3B_{3}. The drastic increase in M(R)M(R) in the interior region is due to the increase of dq/dRdq/dR in Eq. (10). In contrast to variations of M(R)M(R) in the interior region, M(R)M(R) is very stable and independent from the value of B3B_{3} when two nuclei are separated. In the asymptotic region (RR\rightarrow\infty), we always observe the correct limit, M(R)μredM(R)\rightarrow\mu_{\rm red}. The reduced mass for α\alpha+16O 20\rightarrow^{20}Ne is μred=3.2m\mu_{\rm red}=3.2m, and μred=8m\mu_{\rm red}=8m for 16O+16O 32\rightarrow^{32}S.

The cranking masses are shown in the middle panels in Figs. 4 and 5. The perturbative and non-perturbative cranking masses are calculated based on the CHF reaction paths that are constructed with the constraining operator C^=Q^30\hat{C}=\hat{Q}_{30} (mass octupole) for α\alpha+16O 20\rightarrow^{20}Ne, and with C^=Q^20\hat{C}=\hat{Q}_{20} (mass quadrupole) for 16O+16+^{16}O 32\rightarrow^{32}S. If we adopt C^=Q^20\hat{C}=\hat{Q}_{20} for the former reaction system, we cannot obtain a continuous reaction path Wen and Nakatsukasa (2017). Again, for B30B_{3}\neq 0, the cranking mass does not reproduce the correct reduced mass at RR\rightarrow\infty, neither with the perturbative cranking formula nor with the non-perturbative one. Furthermore, the perturbative cranking mass is larger than the reduced mass, even for B3=0B_{3}=0.

Refer to caption
Figure 4: (Color online) Inertial mass M(R)M(R) for the reaction path α\alpha+16O as a function of RR, in units of the nucleon mass. The top panel shows the results of the ASCC method, the middle panel shows those of the perturbative and non-perturbative cranking formulae McrpM_{\rm cr}^{\rm p} (thin lines) and McrnpM_{\rm cr}^{\rm np} (thick lines), and the bottom panel shows the potential energies along the three ASCC collective paths. Solid (red) and dashed (blue) lines indicate those of B3=0B_{3}=0 and 75 MeV fm5. In the bottom panel, in order to show it in the given scale, the potential for B3=75B_{3}=75 MeV fm5 (dashed line) is shifted downwards by 60 MeV.
Refer to caption
Figure 5: (Color online) Same as Fig. 4, but for 16O+16O. Solid (red) and dashed (blue) lines indicate results of B3=0B_{3}=0 and 30 MeV fm5. In the bottom panel, the potential for B3=30B_{3}=30 MeV fm5 (dashed line) is shifted downwards by 45 MeV.

III.4 Rotational moments of inertia

Next let us present results for the moment of inertia which is another important inertial mass for the collective Hamiltonian in Eq. (1). The rotational motion is a NG mode whose ASCC inertial mass can be calculated with the strength function for the angular momentum operator at zero frequency Hinohara (2015), instead of solving the Thouless-Valatin equation (13). For this purpose, we perform the i-FAM calculation Nakatsukasa et al. (2007); Inakura et al. (2009); Avogadro and Nakatsukasa (2011) based on the reaction path self-consistently constructed by the ASCC method. For comparison, we also apply the cranking formula to the moments of inertia, Eq. (24), based on the reaction path calculated with the CHF method.

In Eq. (1), the point-particle approximation was adopted, leading to the moment of inertia, μredR2\mu_{\rm red}R^{2}. There is another classical limit, namely, the rigid-body moment of inertia. In the present case, the system preserves the axial symmetry on the reaction path. Choosing the axis of the symmetry as the zz axis, we have 𝒥x=𝒥y{\cal J}_{x}={\cal J}_{y} whose rigid-body value is given by

𝒥rig=m(y2+z2)ρ(𝐫)𝑑𝐫=m(z2+x2)ρ(𝐫)𝑑𝐫.\displaystyle{\cal J}_{\rm rig}=m\int(y^{2}+z^{2})\rho(\mathbf{r})d\mathbf{r}=m\int(z^{2}+x^{2})\rho(\mathbf{r})d\mathbf{r}. (31)

The rigid-body moment of inertia about the zz axis is non-zero, 𝒥z0{\cal J}_{z}\neq 0, which contradicts with the trivial quantum mechanical requirement that there exists no rotation around the symmetry axis. On the other hand, it is known that the cranking moments of inertia, 𝒥x{\cal J}_{x} and 𝒥y{\cal J}_{y}, for axially deformed nuclei in the harmonic-oscillator-potential model are given by the rigid-body value at the equilibrium Bohr and Mottelson (1975). The cranking formula also satisfies the quantum mechanical condition, producing 𝒥z=0{\cal J}_{z}=0 in the axially symmetric case.

When two nuclei are far away, we expect the point-particle approximation is good. Therefore, we expect that the moments of inertia, 𝒥x{\cal J}_{x} and 𝒥y{\cal J}_{y}, on the reaction path changes from the point-particle value 𝒥(R)=μredR2{\cal J}(R)=\mu_{\rm red}R^{2} at large RR to the rigid-body value 𝒥(R)=𝒥rig{\cal J}(R)={\cal J}_{\rm rig} near the equilibrium (ground) state. It is of significant interest and of importance to examine where and how this transition takes place during the reaction.

In Ref. Wen and Nakatsukasa (2020), we have published the result for α+α8\alpha+\alpha\rightarrow^{8}Be. Thus, in this paper, we present the results for α\alpha+16O 20\rightarrow^{20}Ne and 16O+16O 32\rightarrow^{32}S. First, let us show results for the velocity-independent potential with B3=0B_{3}=0. Figure 6 shows the moments of inertia calculated for the two reaction systems with B3=0B_{3}=0. The ground state of 20Ne corresponds to R=3.8R=3.8 fm in the upper panel of Fig. 6, and the superdeformed minimum of 32S is located at R=5R=5 fm in the lower panel. We can see that the ASCC and the (non-perturbative) cranking formulae give results very similar to each other. This is the consequence of the local Galilean invariance of the mean-field potential under the momentum transformation 𝐩𝐩m(𝝎×𝐫)\mathbf{p}\rightarrow\mathbf{p}-m(\bm{\omega}\times\mathbf{r}) Bohr and Mottelson (1975). The calculated moments of inertia are close to their rigid-body values near these equilibrium states, 𝒥(Re)𝒥rig(Re){\cal J}(R_{e})\approx{\cal J}_{\rm rig}(R_{e}), where ReR_{e} represents the value of RR at the potential equilibrium (dV/dR=0dV/dR=0). However, when the nucleus is more elongated along the reaction path, 𝒥(R){\cal J}(R) decreases as RR increases. Since the rigid-body value is a monotonically increasing function of RR, they become smaller than the rigid-body values, 𝒥(R)<𝒥rig(R){\cal J}(R)<{\cal J}_{\rm rig}(R), then, quickly approach the point-particle values, μredR2\mu_{\rm red}R^{2}. Beyond the scission point R=RsR=R_{s} where the system splits into two separated nuclei, we have 𝒥(R)μredR2{\cal J}(R)\approx\mu_{\rm red}R^{2} (RRsR\geq R_{s}). The calculated moments of inertia give parabolic convex lines as functions of RR, showing their minima located around the midpoint between ReR_{e} and RsR_{s}.

It is surprising that the moments of inertia decrease as the deformation develops near the equilibrium. This contradicts with our naive expectation based on the classical model. For instance, the rigid-body moments of inertia linearly increases with the deformation δ\delta, 𝒥rig𝒥0(1+δ/3){\cal J}_{\rm rig}\approx{\cal J}_{0}(1+\delta/3). Another classical model, irrotational fluid model suggests that it increases quadratically with the deformation as 𝒥irrot𝒥rigδ2{\cal J}_{\rm irrot}\approx{\cal J}_{\rm rig}\delta^{2} Bohr and Mottelson (1975). In any case, a single classical model cannot explain the reduction in 𝒥(R){\cal J}(R) as a function of RR. A hint to understand this behavior may come from the harmonic oscillator model in the quantum mechanics which reproduces both the rigid-body value and the irrotational-fluid value depending on the configuration of nucleus Bohr and Mottelson (1975). During the evolution of deformation as a function of RR, different configurations appear, which may lead to moments of inertia corresponding to different classical models. This suggests the importance of the quantum mechanical calculation.

Another striking feature in Fig. 6 is 𝒥(R)μredR2{\cal J}(R)\approx\mu_{\rm red}R^{2} at RRsR\geq R_{s}, which is significantly smaller than 𝒥rig(R){\cal J}_{\rm rig}(R). When the two nuclei are separated, the structure of the projectile and the target nuclei are almost invariant with respect to RR. Let us denote the rigid-body moments of inertia of the projectile and the target with respect to their own center-of-mass coordinates as 𝒥P{\cal J}_{P} and 𝒥T{\cal J}_{T}, respectively. The total rigid-body moments of inertia can be written as

𝒥(R)=𝒥P+𝒥T+μredR2.{\cal J}(R)={\cal J}_{P}+{\cal J}_{T}+\mu_{\rm red}R^{2}. (32)

This equation is easy to prove in the rigid-body case. In the quantum mechanical treatment of the present cases, there is another trivial result, namely, 𝒥P=𝒥T=0{\cal J}_{P}={\cal J}_{T}=0. This is because both the projectile and the target nuclei (α\alpha and 16O) are spherical in the ground state, thus, the quantum mechanics requires the vanishing moments of inertia (See also Eq. (24)). Assuming that Eq. (32) is valid for the quantum mechanical systems, we find 𝒥(R)μredR2{\cal J}(R)\approx\mu_{\rm red}R^{2} in the region where the projectile and the target become spherical. This is also a consequence of the quantum mechanics.

Next, let us show how the B3B_{3} term influences the moments of inertia. Figure 7 shows the results. For B30B_{3}\neq 0, the velocity dependence in the mean-field potential violates the local Galilean invariance. The calculated cranking moments of inertia are significantly smaller than the rigid-body value 𝒥rig{\cal J}_{\rm rig} at RReR\approx R_{e}. They are also smaller than the point-particle value μredR2\mu_{\rm red}R^{2} at R>RsR>R_{s}. On the other hand, the ASCC calculation includes the residual effects of the time-odd mean fields which restore the local Galilean invariance, then, nicely reproduces both rigid-body and point-particle values at RReR\approx R_{e} and at R>RsR>R_{s}, respectively. In fact, for these reaction systems, the ASCC moments of inertia are insensitive to the B3B_{3} value over the entire region.

Refer to caption
Figure 6: (Color online) Calculated rotational moments of inertias for B3=0B_{3}=0 as a function of relative distance RR. The upper panel shows the results for the system α\alpha+16O, the lower panel shows the results for the system 16O+16O. The solid (red), dashed (pink), dotted (blue) and dash-dotted (green) curves indicate the results of ASCC, non-perturbative cranking formula, rigid body approximation, and point-mass approximation, respectively.
Refer to caption
Figure 7: (Color online) Same as Fig. 6, but for B3=25B_{3}=25 MeV fm5.

III.5 Impact on astrophysical SS factors

Refer to caption
Figure 8: (Color online) Astrophysical SS factor calculated with B3=0B_{3}=0 for α+16\alpha+^{16}O (upper panel) and for 16O+16O (lower panel), as a function of incident energy Ec.m.E_{\rm c.m.}. (1) The point-particle approximation both for the relative and the rotational degrees of freedom, μred\mu^{\rm red} and μredR2\mu_{\rm red}R^{2}, respectively. (2) The ASCC mass M(R)M(R) for the relative motion together with μredR2\mu_{\rm red}R^{2} for the rotation. (3) ASCC M(R)M(R) and rigid-body 𝒥rig(R){\cal J}_{\rm rig}(R). (4) ASCC M(R)M(R) and ASCC 𝒥(R){\cal J}(R). See text for details.
Refer to caption
Figure 9: (Color online) Same as Fig. 8, but calculated with the inertial masses for B3=25B_{3}=25 MeV fm5. The same potentials as those of B3=0B_{3}=0 are used. See text for details.

In this section, we present the calculation of the astrophysical SS factor for fusion reactions, using the potentials and the inertial masses obtained in Sec. III.3 and III.4. The ASCC calculation provides us with the collective Hamiltonian with the optimal reaction paths for 16O+α\alpha \rightarrow 20Ne and 16O+16O32\rightarrow^{32}S. The Hamiltonian for the reaction is given as that of Eq. (1) in which the moment of inertia μredR2\mu_{\rm red}R^{2} is replaced by 𝒥(R){\cal J}(R), and the reduced mass μred\mu_{\rm red} by M(R)M(R). We investigate the dependence of the sub-barrier fusion cross sections on these inertial masses.

The sub-barrier fusion cross section is calculated with the WKB approximation, following the procedure in Ref. Reinhard et al. (1984). The total sub-barrier fusion cross section is given by the sum of all the partial waves with angular moment \ell. Under the WKB approximation, the transmission coefficient for the partial wave \ell at incident energy Ec.m.E_{\rm c.m.} is given by

T(Ec.m.)=[1+exp(2I)]1,\displaystyle T_{\ell}(E_{\rm c.m.})=[1+\exp(2I_{\ell})]^{-1}, (33)

with

I(Ec.m.)\displaystyle I_{\ell}(E_{\rm c.m.}) =\displaystyle= abdR{2M(R)\displaystyle\int_{a}^{b}dR\Big{\{}2M(R) (34)
×\displaystyle\times (V(R)+(+1)2𝒥(R)Ec.m.)}1/2,\displaystyle\left(V(R)+\frac{\ell(\ell+1)}{2{\cal J}(R)}-E_{\rm c.m.}\right)\Big{\}}^{1/2},

where aa and bb are the classical turning points on the inner and outer sides of the barrier respectively. In addition to the potential V(R)V(R), the coordinate-dependent inertial mass for the relative motion M(R)M(R) and for the rotational moment of inertia 𝒥(R){\cal J}(R) are necessary for the calculation. The term (+1)/2𝒥(R)\ell(\ell+1)/2{\cal J}(R) represents the centrifugal potential.

The fusion cross section is given by

σ(Ec.m.)=π2μredEc.m.(2+1)T(Ec.m.).\displaystyle\sigma(E_{\rm c.m.})=\frac{\pi}{2\mu_{\rm red}E_{\rm c.m.}}\sum_{\ell}(2\ell+1)T_{\ell}(E_{\rm c.m.}). (35)

For identical incident nuclei, 16O+16O, Eq. (35) must be modified according to the proper symmetrization. Only the partial wave with even LL contribute to the cross section as

σ(Ec.m.)=π2μredEc.m.[1+()](2+1)T(Ec.m.).\displaystyle\sigma(E_{\rm c.m.})=\frac{\pi}{2\mu_{\rm red}E_{\rm c.m.}}\sum_{\ell}[1+(-)^{\ell}](2\ell+1)T_{\ell}(E_{\rm c.m.}).

Instead of σ(Ec.m.)\sigma(E_{\rm c.m.}), we show the astrophysical SS factor defined by

S(Ec.m.)=Ec.m.σ(Ec.m.)exp[2πZ1Z2e2/v],\displaystyle S(E_{\rm c.m.})=E_{\rm c.m.}\sigma(E_{\rm c.m.})\exp[2\pi Z_{1}Z_{2}e^{2}/\hbar v], (37)

where vv is the relative velocity at RR\rightarrow\infty. The astrophysical SS factor is preferred for sub-barrier fusion because it removes the change by tens of orders of magnitude present in the cross section due to the penetration through the Coulomb barrier. The SS factor can reveal in a more transparent way the influence of the nuclear structure and dynamics.

Figure 8 shows the calculated SS factor with B3=0B_{3}=0 for the scattering of α+16\alpha+^{16}O (upper panel) and 16O+16O (lower panel), respectively. In order to clarify the effect of the inertial masses, we use the same potential V(R)V(R) for all the curves in each panel of Fig. 8, which is the ASCC potential obtained for B3=0B_{3}=0 (solid lines in the bottom panels of Fig. 4 and 5). Different curves show the calculations with different inertia masses. For the case of B3=0B_{3}=0, the ASCC and the cranking inertial masses are similar to each other. Thus, the calculation with the cranking inertial masses produces the astrophysical SS factor similar to the ASCC result. Generally speaking, the larger the inertial mass M(R)M(R) for the relative motion is, the smaller the astrophysical SS factor is. The opposite effect can be seen for the rotational moments of inertia 𝒥(R){\cal J}(R). The larger moments of inertia give the larger SS factor. The replacement of the constant mass μred\mu_{\rm red} by ASCC M(R)M(R) gives a significant effect for the fusion cross section. For the reaction system 16O+16O, we find a strong suppression of the fusion cross section at Ec.m.<4E_{\rm c.m.}<4 MeV compared to the calculation with μred\mu_{\rm red}, which is an order of magnitude or even larger.

Figure 9 shows the same SS factors with inertial masses calculated with B3=25B_{3}=25 MeV fm5 for the scattering of α+16\alpha+^{16}O (upper panel) and 16O+16O (lower panel), respectively. In order to see the effect of inertial masses, we use the potential V(R)V(R) obtained for B3=0B_{3}=0. Thus, the differences between Figs. 8 and 9 come from the change in inertial masses. Again, the change from the reduced mass μred\mu_{\rm red} into M(R)M(R) gives the largest impact on the SS factor. The suppression effect of M(R)M(R) is even more significant than the B3=0B_{3}=0 case, because the M(R)M(R) is larger for finite values of B3B_{3}, as we can find in Fig. 4 and 5.

Although the BKN density functional provides us only with qualitative results, these calculations reveal the important roles of the inertial masses, and suggest significant influence of the inertial masses on the sub-barrier fusion cross sections.

IV Summary

We calculated the ASCC inertial mass coefficients with respect to the translational, relative and rotational motions. The numerical calculations are performed using the FAM in the 3D coordinate space representation for the reaction systems α\alpha+α\alpha \rightarrow 8Be, α+16\alpha+^{16}O \rightarrow 20Ne and 16O+16+^{16}O \rightarrow 32S. We investigated the impact of time-odd mean-field potentials on the collective inertial masses. In the presence of nucleonic effective mass, the cranking mass can neither reproduce the total mass AmAm of translational motion nor the correct reduced mass μred\mu_{\rm red} for the relative motion. In addition, the cranking formula for the rotation produces neither the point-particle value μredR2\mu_{\rm red}R^{2} in the asymptotic region (RR\rightarrow\infty), nor the rigid-body value 𝒥rid{\cal J}_{\rm rid} near the equilibrium states. In contrast, the ASCC masses is able to properly take into account the residual time-odd effects, producing the total mass AmAm for translation, the reduced mass μred\mu_{\rm red} for relative motion, and the point-particle value μredR2\mu_{\rm red}R^{2} (the rigid-body value 𝒥rid{\cal J}_{\rm rid}) after the scission (near the equilibrium). We also calculated the astrophysical SS factors with these microscopic inputs as the inertial masses of a reaction model. It turns out that replacement of the reduced mass μred\mu_{\rm red} by M(R)M(R) obtained with ASCC gives the largest impact on the SS factors. It could lead to a strong suppression of the fusion cross section at low energy.

In the present study, a schematic BKN interaction plus time-odd terms are adopted for a qualitative investigation. It is highly desired to use realistic modern nuclear energy functionals. The paring correlation is also expected to play a critical role in low-energy collective dynamics. These issues are currently under investigation.

Acknowledgements.
This work is supported in part by JSPS KAKENHI Grant No. 18H01209, No. 19H05142, and No. 20K14458. This research in part used computational resources provided through the HPCI System Research Project (Project ID: hp200069), and by Multidisciplinary Cooperative Research Program in Center for Computational Sciences, University of Tsukuba.

References