Prévia do material em texto
ar X iv :2 30 1. 00 20 2v 1 [ nu cl -t h] 3 1 D ec 2 02 2 A transport model description of Time-Dependent Generator Coordinate under Gaussian overlap approximation Fangyuan Wang,1 Yingxun Zhang,1, ∗ and Zhipan Li2 1Department of Nuclear Physics, China Institute of Atomic Energy, Beijing 102413, People’s Republic of China 2School of Physical Science and Technology, Southwest University, Chongqing 400715, People’s Republic of China (Dated: January 3, 2023) In this work, we derived a transport equation based on a generalized equation of time-dependent generator coordinate method (TDGCM) under the Gaussian overlap approximation (GOA). The transport equation is obtained by using quantum-mechanics phase space distributions under a “quasi-particle” picture and strategy of Bogoliubov-Born-Green-Kirkood-Yvon (BBGKY) hierar- chy. The theoretical advantage of this transport equation is that time evolution of s-body phase space density distribution is coupled with s + 1-body phase space density distributions, and thus, non-adiabatic effects and dynamical fluctuations could be involved by more collective degrees and entanglement of phase space trajectories. In future, we will perform the numerical calculations for fission nuclei after obtaining collective inertia and potential energy surface (PES). I. INTRODUCTION After the discovery of nuclear fission by Hahn and Strassmann [1] in 1939, nuclear fission has become one of the most challenging topics in physics since it is a key ingredient for modeling nucleosynthesis [2], energy pro- duction [3], medicine [4], and nuclear safeguard [5]. Even with recent progress in experimental techniques, mea- surements of nuclear fission are not possible for all fis- sion nuclei. Thus, theoretical simulations are mandatory for fully understanding the fission dynamics and comple- menting the missing data [6–12]. One kind of models is useful macroscopic model by taking into account shell effects, collective variables, and correlations between collective degree and single parti- cles motions, such as Brownian shape motion [13–15] and Langevin model [16–18]. With four or five collective degrees, it could depict fission dynamics process appro- priately and reproduce fission yields distribution well. Another kind of models is microscopic model, which based on nucleonic Hamiltonian and solve fission dy- namics with Schrödinger or Dirac equation in time domain. For example, time-dependent density func- tional theory, such as time-dependent superfluid lo- cal density approximation (TDSLDA) [19, 20], Con- strained and time-dependent Hartree-Fock calculations with dynamical Bardeen–Cooper–Schrieffer pairing cor- relations (CHF+BCS) [21, 22], time dependent Hartree- Fock-Bogliubov (TDHFB)/TDHF+BCS [23], the adia- batic time-dependent HFB (ATDHFB) [24] and time- dependent covariant density functional theory (TD- CDFT) [25] describe fission process with full quantum microscopic approaches. At tremendous costs of compu- tations, these models successfully describe the fission dy- namics and predict the most probable fission yields. The great understanding of fission dynamics from microscopic model are obtained [19, 22]. However, describing fission ∗ zhyx@ciae.ac.cn yields distribution with these models is still a theoreti- cal challenge due to the lacking of fluctuation in initial state and fission process. The efforts on this direction is to describe quantum fluctuation by a sampling of initial conditions followed by TDDFT[26] but without quantum interference. An alternative method to include the correlation is to represent a many-body wave function of the system with a mixture of states with different shapes. It stim- ulates the description of fission dynamics with time- dependent generator coordinate method under Gaussian overlap approximation (TDGCM+GOA) [27–29]. In the TDGCM+GOA, fission is assumed as adiabatic process since the typical time for the motion of individual nucle- ons inside the fission nucleus (roughly 10−22 s) is roughly ten times smaller than the time scale of the system’s col- lective deformation (10−21 s) [29]. Thus, the fission dy- namics are approximately described in terms of a few shape coordinates. Currently, most of the TDGCM+GOA calculations were performed by using only two degrees of freedom, usually (q20, q30) or (β2, β3), under adiabatic assump- tion [30–32]. While, the semiphenomenological and fully microscopic approaches illustrate that at least four or five collective variables play a role in the dynamics of fission [19, 33–37]. Regnier et al. [38] have started some trials on the rigorous three degrees of freedom calculation of the PES for 240Pu in the collective space (q20, q30, q40), and their work is still in progress. In Ref. [39], Zhao et al. did the calculations with the dynamical pairing de- gree of freedom as the third degree of freedom besides (β2, β3), and their results also demonstrate the impor- tance of including more degree of freedom. Furthermore, one should note that an ad hoc Gaussian smoothing have to be used at the end of the TDGCM+GOA calculations, to account for the fluctuations in particle number of the fragments due to both pairing effects and the finite num- ber of particles in the neck region for points along the fission line. Thus, one would expect to develop a micro- scopic method that can include more collective degrees http://arxiv.org/abs/2301.00202v1 mailto:zhyx@ciae.ac.cn 2 and the correlations to account for dynamical fluctua- tion and non-adiabatic effects, and reasonably describing fission dynamics and distribution of fission yields. In this work, we derive a transport equation based on a generalized N-dimensional TDGCM+GOA equation to describe fission dynamics, in which non-adiabatic effects are introduced by more collective degrees, fluctuations are introduced by initial state fluctuation and entan- glement of phase space trajectories. The paper is or- ganized as follows: in Sec.II, the transport equation is obtained by using quantum-mechanics phase space dis- tributions under a “quasi-particle” picture and strategy of Bogoliubov-Born-Green-Kirkood-Yvon (BBGKY) hi- erarchy. One of the advantages of this hierarchy is that correlations from high-order degree can be involved in the evolution of the one-body phase-space density dis- tribution. In Sec.III, a numerical recipes for solving the transport equation is provided. Sec.IV is the summary and outlook. II. THEORY FRAMWORK A. Overview of TDGCM for fission For convenience, we briefly review the TDGCM +GOA theory which describes induced fission as a slow adiabatic process determined by a small number of collective de- grees of freedom. Under the Griffin-Hill-Wheeler ansatz, the many-body state of fissioning system at any time reads |Ψ(t)〉 = ∫ q∈E dq|φ(q)〉f(q, t). (1) The set {|φ(q)〉} is a family of the generator states which are the solutions of a constrained Hartree-Fock- Bogoliubov equation. f(q, t) is the complex-valued weights of the quantum mixture of states. The gener- ator coordinate q = {q1, ..., qN}, and each of these qi is a collective variable chosen based on the physics of fission. The time-dependent Schrödinger equation for the many-body state of fission system |Ψ(t)〉, (Ĥ − i~ d dt )|Ψ(t)〉 = 0, (2) can yield an equation of the unknown weight func- tion f(q, t), i.e., the Hill-Wheeler equation with time- dependent form, ∫ dq′〈φq| ( Ĥ − i~ d dt ) |φq′〉f(q ′, t) = 0. (3) Here, Ĥ is the Hamiltonian acting on the full many-body system. Principally, Eq. (3) can be solved numerically, but it needs a tremendous amount of computations. To overcome these difficulties, a popular approach named as Gaussian overlap approximation (GOA) is used. The simplest formulation of GOA assumes that the overlap between two generator states 〈φq|φq′〉 has a Gaussian shape, N (q,q′) = 〈φq|φq′〉 ≡ exp [ − 1 2 (q− q′)tG(q̄)(q − q′) ] . (4) N (q,q′) = 〈φq|φq′〉 is peaked functions for q = q′, and q̄ = (q+ q′)/2. By changinga new collective coordinate α by the relation α(q) = ∫ a∈Cq 0 G1/2(a)da, (5) in terms of which the overlap matrix becomes N (α,α′) = exp [ − 1 2 (α−α′)2 ] , (6) G (q) is the metric of new coordinates of α(q), and G (q) is the determinant of G. Within this approximation, the time-dependent Hill- Wheeler equation is reduced to a local, time-dependent Schrödinger-like equation as i~ ∂g(q, t) ∂t = Ĥcoll(q)g(q, t). (7) g(q, t) is related to the weight function f(q, t) as g = N 1/2f , and contains all the information about the fis- sion dynamics of system [29]. The collective Hamiltonian Ĥcoll(q) is a local operator acting on g(q, t), Ĥcoll(q) = (8) [ − ~ 2 2 ∑ kl 1 √ G (q) ∂ ∂qk √ G (q)Bkl (q) ∂ ∂ql + V (q) ] . The potential V (q), V (q) = 〈q|Ĥ |q〉 − ǫ0(q), (9) with the zero-point energy-correction ǫ0 = 1 2 Gij(q) ∂2h ∂qi∂q′j ∣ ∣ ∣ ∣ q=q′ . (10) The symmetric collective inertial tensor B(q) ≡ Bij(q), Bkl(q) = 1 2~2 Gkm(q) ( ∂2h(q,q′) ∂qm∂q′n − ∂2h(q,q′) ∂qm∂qn + { i mn } ∂h(q,q′) ∂qi )∣ ∣ ∣ ∣ q=q′ Gnl(q), (11) the expression in braces is the Christoffel symbol of the second kind. h(q,q′) is h(q,q′) = 〈 φq|Ĥ |φq′ 〉 〈φq|φq′〉 . (12) 3 They are usually calculated from the nuclear Hamilto- nian Ĥ and the generator states |φq〉 with HFB [30] or RMF+BCS [40]. The number of collective degree of freedom are usually selected as N = 2, and shape coordinates q are the mul- tipole moments Q20 and Q30 in Ref. [30–32], or β2 and β3 as in Refs. [39, 40], and G (q) = 1 Ref. [31]. In this case, the equation of TDGCM+GOA is i~ ∂ ∂t g (q1, q2; t) = (13) [ − ~ 2 2 ∑ kl ∂ ∂qk Bkl (q1, q2) ∂ ∂ql + V (q1, q2) ] g (q1, q2; t) . This equation has been solved by the software package FELIX-1.0 [31] or FELIX-2.0 [32] with finite element method. B. A transport equation for N-dimensional TDGCM+GOA The TDGCM has achieved great progress on describ- ing the fission dynamics [30–32, 39–41], but previous cal- culations in Refs. [33–36, 38] also showed it is necessary to include more degrees to describe nonadiabatic effects which may arise from the coupling between collective and intrinsic degrees of freedom, and invovle dynamical fluc- tuations to describe fission products distributions. Now, the question is that can we effectively involve more collec- tive degrees of freedom into the equation with two degrees that we are currently using? Principally, nuclear shape can be described by an ex- pansion in spherical harmonics, i.e., R(θ, φ, t) = R0 1 + N ∑ λ=0 λ ∑ µ=−λ α∗λµ(t)Yλµ(θ, φ) . (14) The number of shape coordinates (or collective degree of freedom) N depends on the choice of collective co- ordinates or generator coordinates and stage of fission. In the stage of fissionning system from the quasista- tionary initial state to the outer fission barrier, evolu- tion is slow and fission process can be described by a small number of collective degree, i.e., a small N , with adiabatic approximation[42]. In the stage of saddle-to- scission, the nucleus quickly elongates toward scission and non-adiabatic effects have to be considered and N may vary with the stage of fission process. Thus, the col- lective wave function is presented with N degrees, i.e., g(q1, q2, ..., qN ), for fissioning system, and Eq.(7) will be- come a generalized N-dimensional TDGCM+GOA equa- tion since the degrees are not limited to a few. In this work, we interpret the wave function of g(q1, q2, ..., qN ) for fissioning system as a wave function for ‘N-body quasiparticles in 1-Dimension space’ sys- tem (NB1D), and a convention gN(q1, q2, ..., qN ) is used in following to represent the wave function with parti- cle from 1 to N. Then, we derive a transport equation which can effectively couple one more collective degree of freedom to the degrees currently used. Firstly, we per- form Wigner transformation [43] on NB1D wave function gN (q1, · · · , qN ) to get their quantum mechanically phase space density fN as, fN (q1, · · · , qN ; p1, · · · , pN ) (15) = 1 (π~)N ∫ · · · ∫ dy1 · · · dyNg ∗ N(q1 − y1, · · · , qN − yN ) gN (q1 + y1, · · · , qN + yN) × exp[−2i(p1 · y1 + · · ·+ pN · yN )/~]. Here pi is the conjugated momentum of qi for quasi- particle i. After a trivial deviation, the corresponding transport equation reads, ∂fN ∂t = − ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql (16) + ∑ λ=1,3,··· ( ~ 2i )λ1+···+λN−1 1 λ1! · · ·λN ! × ∂λ1+···+λNVN (q) ∂qλ11 · · · ∂q λN N ∂λ1+···+λN fN ∂pλ11 · · ·∂p λN N = − ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql + ∑ l ∂VN ∂ql ∂fN ∂pl + ∑ λ=3,··· ( ~ 2i )λ1+···+λN−1 1 λ1! · · ·λN ! × ∂λ1+···+λNVN (q) ∂qλ11 · · · ∂q λN N ∂λ1+···+λN fN ∂pλ11 · · ·∂p λN N . Here, λ = ∑N i=1 λi and B̄ (b) kl (q) is an effective collective inertia, which is defined as B̄ (b) kl (q) = ( Bkl(q− y ∗ (1b)) +Bkl(q− y ∗ (2b)) ) /2. (17) y∗(1b) and y ∗ (2b) are corrections on q, and their origins can be found in appendix A. VN (q) is N-body potential. Principally, VN (q) = V (q1, q2, · · · , qN ). If the N-body potential is calculated from the two-body interaction, i.e., V (q1, q2, · · · , qN ) = ∑ i≤j V (qi, qj), (18) the transport equation is simplified as, ∂fN ∂t = − ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql + 1 2 ∑ k,m 6=k ∂Vkm ∂qk ∂fN ∂pk .(19) Vkm is two-body potential between quasi-particle k and m, which can be obtained by HFB/RMF+BCS as in Refs.[30, 40]. When the N-body potential is obtained from multi-dimensional PES by HFB/RMF+BCS, trans- port equation of Eq.(16) should be used. 4 A standard procedure to solve the N -body transport equation is to use BBGKY hierarchy, in which one-body degrees of freedom (DOF) is coupled to two-body DOF that are themselves coupled to three-body DOFs and so forth. As an example, we present the time evolu- tion of fs under the condition of V (q1, q2, · · · , qN ) = ∑ i≤j V (qi, qj). The s-body phase space density distri- bution fs is defined as, fs(q1, · · · , qs, p1, · · · , ps) (20) = 1 ΩN−s ∫ fN (q1, · · · , qN , p1, · · · , pN )dΓs+1 · · · dΓN , dΓi = dqidpi, here, Ω is volume in phase space. Thus, ∂fs(q1, · · · , qs, p1, · · · , ps) ∂t (21) = 1 ΩN−s ∫ ∂fN (q1, · · · , qN , p1, · · · , pN ) ∂t dΓs+1 · · · dΓN = 1 ΩN−s ∫ [ − ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql + ∑ 1≤k<m≤N ∂Vkm ∂qk ∂fN ∂pk ] dΓs+1 · · · dΓN . One should note that the derivation in this case is dif- ferent than a system with fixed-mass many particles, be- cause the inertial B̄ (b) kl (q) depends on collective coordi- nates. To overcome this difficulty, we move the B̄ (b) kl (q) out from the integration by assuming the following rela- tionship, i.e., ∫ Bkl(q)O(q,p)dΓs+1 · · · dΓN = (22) Bkl(qs, q ∗ s+1, · · · , q ∗ N ) ∫ O(q,p)dΓs+1 · · · dΓN , The q∗s+1, · · · , q ∗ N depend on O(q,p), and its values will be fixed once the O(q,p) is determined. By using above relationship, we get ∂fs ∂t = − s ∑ k=1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )pk ∂fs ∂ql (23) − N ∑ k=s+1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )p ∗ k ∂fs ∂ql + ∑ 1≤k<m≤s ∂Vkm ∂qk ∂fs ∂pk + N − s Ω ∫ s ∑ k=1 ( ∂Vk,s+1 ∂qk ) × ( ∂fs+1 ∂pk ) dΓs+1. The collective inertial B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N ) only varies with qs, since the values of q ∗ s+1, . . . , q ∗ N will be fixed once the integrand was selected. But the values of B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N ) could be different from Bkl(qs). The details of the derivation are in appendix B. By expressing the fourth terms on the right side of Eq. (23) as δIcoll, the transport equation can be rewritten as, ∂fs ∂t + s ∑ k=1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )pk ∂fs ∂ql (24) + N ∑ k=s+1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )p ∗ k ∂fs ∂ql − ∑ 1≤k<m≤s ∂Vkm ∂qk ∂fs ∂pk = δIcoll. with δIcoll = N − s Ω ∫ s ∑ k=1 ( ∂Vk,s+1 ∂qk )( ∂fs+1 ∂pk ) dΓs+1.(25) As one can see that the δIcoll is related to the phase space density fs+1, and the potential betweenk and s+1, i.e., Vk,s+1. III. NUMERICAL RECIPE FOR SOLVING TRANSPORT EQUATION In this section, we discuss practical numerical recipe of the time evolution of f2. In the following discussions, we take q1 = β2 and q2 = β3. Given s = 2, the time evolution of f2 becomes, ∂f2(q1, q2; p1, p2) ∂t (26) + 2 ∑ k=1 2 ∑ l=1 pkB̄ (b) kl (q1, q2, q ∗ 3 , · · · , q ∗ N ) ∂f2 ∂ql + N ∑ k=3 2 ∑ l=1 p∗kB̄ (b) kl (q1, q2, q ∗ 3 , · · · , q ∗ N ) ∂f2 ∂ql − 2 ∑ k=1 2 ∑ l 6=k 1 2 ∂Vkl ∂qk ∂f2 ∂pk = δIcoll. The time evolution of f2 is not only related to the col- lective inertia Bkl and potential Vkl, but also related to f3 and potential between qi=1,2 and q3 which actually reflect the high order correlation between different shape coordinates. A. Initialization For the TDGCM+GOA equation, the starting point is a collective wave packet at initial time, which represents the compound nucleus after excitation by absorption of a low-energy neutron or photon. One choice is to use the quasibound state [30], i.e., collective ground state g0(q, t = 0), and q = {q1, q2}. Its modulus is roughly a Gaussian centered on the minimum of the potential 5 Vmin(q), which is achieved by extrapolating inner poten- tial barrier with a quadratic form. The width of this Gaussian is characterized by a width close to the dimen- sion of the first potential well. To describe fission, g0 has to boost in q1(β2) direction for simulating the fission events, i.e., g(q, t = 0) = g0(q) exp(ikq1), (27) since its average energy is below the fission barrier. The amplitude k of the boost is determined so that the av- erage energy of the initial state lies few MeV above the inner fission barrier. For the transport equation described in this work, we need to do the initialization in {q1, q2; p1, p2} space ac- cording the phase space density f2. The initial f2, which represents the compound nucleus after the absorption of a low-energy neutron, can be realized by doing Wigner transformation on g0(q), i.e., f2(q1, q2; p1, p2, t = 0) = (28) 1 (π~)2 ∫ dy1dy2g ∗ 0(q+ y)g0(q− y)e 2ip·y/~. Numerically, a test particle method is used to describe f2(q1, q2; p1, p2, t = 0), which means each quasiparticle is replaced by a large number of test particles and the method was first proposed by Wong in nuclear Vlasov model [44]. f2(q1, q2; p1, p2, t = 0) = (29) 1 Ntest 2 ∑ k=1 Ntest ∑ i=1 δ(qki − q̄ki(t))δ(pki − p̄ki(t)). qki and pki are the time-dependent coordinates and mo- menta of the test-particle i for particle k=1 or 2. q̄ki(t = 0) and p̄ki(t = 0) are sampled according to the f2. Ntest is the number of test particles. Once p̄ki(t = 0)’s are obtained, p̄1i(t = 0) is boosted as p̄1i(t = 0) + k. k can be obtained as same as in TDGCM initialization. B. Time evolution To solve the transport equations with test particle method, we separately treat the left and right hand of Eq.(26) as same as in the transport model used for sim- ulating the heavy ion collisions[45]. The equation of motion of test particles under the mean field can be obtained by comparing, df2 dt = ∂f2(q1, q2; p1, p2) ∂t + 2 ∑ k=1 [ ∂f2 ∂qk ∂qk ∂t + ∂f2 ∂pk ∂pk ∂t ] = 0, (30) to Eq.(26) without considering the collision term, i.e., ∂f2 ∂t = − 2 ∑ k=1 2 ∑ l=1 B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N )pk ∂f2 ∂ql (31) − N ∑ k=3 2 ∑ l=1 B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N )p ∗ k ∂f2 ∂ql + ∑ 1≤k<m≤2 ∂Vkm ∂qk ∂f2 ∂pk . The equation of motion of test particle ki becomes, q̇ki = 2 ∑ l=1 B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N )pli (32) + N ∑ l=3 B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N )p ∗ li ṗki = − 1 2 ∑ 1≤l≤2 ∂Vkl ∂qki , (33) The abbreviation of ki in the lower index means par- ticle k and its ith test-particle, i.e., k = 1, 2 and i = 1, · · · , Ntest. As shown in Eq.(32), the evolution of position of test particle ki not only depend on the collective inertia B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N ) and momentum of pli with l = 1 and 2, but also on collective inertia B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N ) and effective momentum of parti- cle of p∗ki with l ≥ 3. In the calculations, one can approx- imate B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N) = η(q ∗ 3 , . . . , q ∗ N )Bkl(q1, q2) for l ≤ 2, in which the parameter η depend on the selection of qk≥3. Alternatively, one can use η as a phenomenological parameter to fit the data of fission yield. The value of B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N ) = η(q∗3 , . . . , q ∗ N )Bkl(q1, q2) for l ≥ 3 can be learned if the three-dimensional collective inertia can be provided. Within the framework of this equation, the correlations beyond q1 and q2 are involved via B̄ (b) kl (q1, q2, q ∗ 3 , . . . , q ∗ N ). In the second term of Eq.(32), the contribution of p∗ki can also be thought as friction effects. The momentum of test particle update according to Eq.(33), and it will depend on the potential Vkm. Inside the scission line (hypersurface), potential Vkm can be ob- tained by RMF+BCS/HFB model. Out of the scission hypersurface, the fissioning trajectories will not go back and one can set the potential ∂Vkm/∂qk = 0. For high order correlation term in Eq. (25), i.e., the collision term in our approach, it comes from, δIcoll = N−2 Ω ∫ ∑2 k=1 ( ∂Vk,3 ∂qk )( ∂f3 ∂pk ) dΓ3. (34) Suppose f3 can be expressed as, f3(q1, q2, q3; p1, p2, p3) = f2(q1, q2; p1, p2)f1(q3; p3), (35) 6 Thus, δIcoll = 2 ∑ k=1 ( ∂Φ̄k ∂qk )( ∂f2 ∂pk ) , (36) and potential Φ̄k means, Φ̄k = N − 2 Ω ∫ Vk,3(qk, q3)f1(q3, p3)dΓ3, (37) which reflect the potential of k particle felt by surround- ing particles. However, exact calculations of Eq.(36) and Eq.(37) are impossible since they always beyond one more degree we have. One effective way to handle the collision term is to use a random collision among test particles. For example, one first select three particles among 2Ntest test parti- cles according to “collision section”, and then, perform a random collision as follows, pk1 + pk2 + pk3 = p ′ k1 + p ′ k2 + p ′ k3. (38) p′k1, p ′ k2 and p ′ k3 will be determined by using the Monte- Carlo sampling under the momentum conservation, p′k1 = (pk1 + pk2 + pk3) ∗ ξ1, (39) (p′k2 + p ′ k3) = (pk1 + pk2 + pk3) ∗ (1− ξ1), (40) p′k2 = (p ′ k2 + p ′ k3) ∗ ξ2, (41) p′k3 = (p ′ k2 + p ′ k3) ∗ (1− ξ2). (42) ξ1 and ξ2 are the random number satisfy a certain dis- tribution. The collision among test particles during the evolution describe entanglement among different trajec- tories of test particles. In practical calculations, the col- lision probability can be adjusted by introducing a ‘cross section’ of three-body collision. After random collision, the values of momentum of test particle will be randomly modified. As a result, the fluctuation on q will be au- tomatically involved and the mass distribution of fission fragment can be expected. C. Fission fragments distributions In this work, we only focused on fission fragment mass/charge distribution. First, one need to search scis- sion line (or hypersurface) on potential energy surface, which is composed from many scission points qsci. In- side the scission line(hypersurface), the nucleus is whole. Out of scission line (hypersurface), the system becomes two well-separated fragments which are connected by a thin neck. Thus, each scission points qsci is associated with a given fragmentation (AL, AR). AL and AR means the mass of fission fragment in the left and right of neck, respectively. The fission fragment mass can be obtained from the integration of single-body density over the do- main of left or right of neck, AL = ∫ r∈L drρ(r), (43) AR = ∫ r∈R drρ(r). Here, R and Lmeans the region of right and left of neck of fissioning system. ρ(r) is constructed from the collective coordinates. In transport model approach, the probability of mea- sured the fission fragment AL and AR, i.e., Y (AL) and Y (AR) can also be obtained from the time integrated flux through the hypersurfaceelement S. By using the test particle method, it will be obtained by counting the num- ber of test particles across the hypersurface S at scission point, i.e., with t → +∞, Y (A, S) = ∫ ∫ qs>qsci fs(qs,ps, t)dpsdqs (44) = 1 2Ntest 2 ∑ k=1 Ntest ∑ i=1 Θ(qki − q̄ki,sci(t)). Thus, the yield of mass of fragment Y (A) = ∑ S Y (A, S). (45) The sum on S runs over the whole scission hypersurface. IV. SUMMARY AND OUTLOOK In this work, we derive a transport equation based on a generalized N-dimensional TDGCM+GOA equation to describe fission dynamics. The transport equation is obtained by using quantum-mechanics phase space dis- tributions under a “quasi-particle” picture and strategy of BBGKY hierarchy. The advantages of this transport equation is that time evolution of s-body phase space density distribution is coupled with s + 1-body phase space density distributions. Thus, one can expect that non-adiabatic effects and dynamical fluctuations could be introduced by involving more collective degrees and entanglement of phase space trajectories. Different than directly solving the TDGCM+GOA equation with finite elements method, our approach is realized by using the test particle method. The coordi- nates and momentum of test particles in initialization of fissioning system are sampled according to initial phase space density distribution of system. The time evolu- tion of test particles are governed by a Hamiltonian-like equation coupled with a random scattering between test particles, which will naturally provide a fluctuation on the mass of fission fragments. Finally, the numerical re- sults on this transport equations are still on the way, since we need to select reasonable collective coordinates to obtain the results of PES and collective inertia, and 7 the high order effects of degree of freedom on collective inertia should also be investigated in this approach before presenting numerical results. ACKNOWLEDGEMENTS The authors thank Zhuxia Li, Zaochun Gao and Siyu Zhuo for reading the manuscript and providing useful feedback. This work was partly supported by the National Natural Science Foundation of China Nos. 11875323, 12275359, 11875225, 11705163, 11790320, 11790323, and 11961141003, the National Key R&D Pro- gram of China under Grant No. 2018 YFA0404404, the Continuous Basic Scientific Research Project (No. WDJC-2019-13, BJ20002501), Key Laboratory of Nu- clear Data foundation (No.JCKY2022201C158) and the funding of China Institute of Atomic Energy YZ222407001301. The Leading Innovation Project of the CNNC under Grant No. LC192209000701, No. LC202309000201. Appendix A: Derivation of transport equation The equation of motion of gN is determined by the Schrödinger-like equation, i.e., i~ ∂ ∂t gN (q, t) = (A1) [ − ~ 2 2 ∑ kl ∂ ∂qk Bkl (q) ∂ ∂ql + VN (q) ] gN (q, t) . By using the Wigner transformation [43], fN (q,p) is ob- tained. The equation of motion of fN (q,p) reads as, ∂fN(q,p, t) ∂t = 1 (π~)N ∫ dy1 · · · dyNe −2ip·y/~ (A2) { ∂g∗N(q− y) ∂t gN (q+ y) + g ∗ N (q− y) ∂gN(q + y) ∂t } . where p is the conjugate momentum of q. In the derivations of Eq. (A2), the ∂g∗N/∂t and ∂gN/∂t are replaced with the right hand side of Eq. (A1) and we have, ∂fN(q,p, t) ∂t = 1 (π~) N ∫ dy1 · · · dyNe −2ip·y/~ (A3) { ∑ kl [ − i~ 2 ∂ ∂(qk − yk) ( Bkl (q− y) ∂g∗N (q− y) ∂(ql − yl) ) gN (q+ y) + i~ 2 g∗N (q− y) ∂ ∂(qk + yk) ( Bkl (q+ y) ∂gN (q+ y) ∂(ql + yl) )] + i ~ (V (q− y)− V (q+ y)) g∗N (q− y)gN(q + y) } . One should note that the kinetic energy term contains the collective coordinate q dependence of inertia Bkl, which is different than the N-body system with fixed particle mass. In coordinate space, the kinetic energy terms in Eq. (A3) are as follows, Mkl = − i~ 2 1 (π~)N ∫ dy1 · · ·dyNe −2ip·y/~ (A4) [ ∂ ∂yk ( Bkl(q− y) ∂g∗N (q− y) ∂yl ) gN(q+ y) −g∗N(q− y) ∂ ∂yk ( Bkl(q+ y) ∂gN (q+ y) ∂yl )] = Mkl,1 +Mkl,2. Since the gN and g ∗ N depend on the integration vari- able yk, one can replace the differentiations with respect to qk+yk by differentiations with respect to yk in deriva- tions. By doing one partial integration with respect to yk, one has Mkl,1 = (A5) − i~ 2 1 (π~)N Bkl(q− y) ∂g∗N (q− y) ∂yl gN (q+ y)e −2ip·y/~ ∣ ∣ ∣ ∣ +∞ −∞ + i~ 2 1 (π~)N ∫ dy1 · · · dyNBkl(q− y) ∂g∗N (q− y) ∂yl × ∂ ∂yk ( gN (q+ y)e −2ip·y/~ ) The first term in Eq. (A5) vanishes due to the boundary condition at infinity, and second term becomes Mkl,1 = i~ 2 1 (π~)N ∫ dy1 · · ·dyN (A6) ×Bkl(q− y) ∂g∗N (q− y) ∂yl ∂ ∂yk ( gN (q+ y)e −2ip·y/~ ) = i~ 2 1 (π~)N ∫ dy1 · · ·dyNs ×Bkl(q− y) ∂g∗N (q− y) ∂yl [ ∂gN(q + y) ∂yk e−2ip·y/~ +gN(q+ y)e −2ip·y/~(− 2ipk ~ ) ] = i~ 2 Bkl(q− y ∗ (1a)) 1 (π~)N ∫ dy1 · · · dyN (A7) ∂g∗N(q − y) ∂yl ∂gN (q+ y) ∂yk e−2ip·y/~ + i~ 2 Bkl(q− y ∗ (1b)) 1 (π~)N ∫ dy1 · · ·dyN ∂g∗N(q − y) ∂yl gN(q+ y)e −2ip·y/~(− 2ipk ~ ). In above derivations, the Bkl(q− y) is moved out by assuming the following relationship, i.e., ∫ Bkl(q − y)O(q, ∂/∂q)dy1 · · · dyN = (A8) Bkl(q− y ∗ (I)) ∫ O(q, ∂/∂q)dy1 · · · dyN . 8 Similarly, the Mkl,2 becomes Mkl,2 = − i~ 2 Bkl(q+ y ∗ (2a)) 1 (π~)N ∫ dy1 · · · dyN(A9) ∂g∗N(q− y) ∂yk ∂gN(q + y) ∂yl e−2ip·y/~ − i~ 2 Bkl(q+ y ∗ (2b)) 1 (π~)N ∫ dy1 · · · dyN g∗N(q− y) ∂gN (q+ y) ∂yl e−2ip·y/~(− 2ipk ~ ) By defining B̄ (a) kl , B̄ (b) kl , δB (a) kl and δB (b) kl as, B̄ (a) kl = ( Bkl(q− y ∗ (1a)) +Bkl(q+ y ∗ (2a)) ) /2, B̄ (b) kl = ( Bkl(q− y ∗ (1b)) +Bkl(q+ y ∗ (2b)) ) /2, δB (a) kl = ( Bkl(q− y ∗ (1a))−Bkl(q+ y ∗ (2a)) ) , δB (b) kl = ( Bkl(q− y ∗ (2a))−Bkl(q+ y ∗ (2b)) ) , and assuming δB (a) kl ≈ 0 and δB (b) kl ≈ 0, Mkl becomes Mkl = Mkl,1 +Mkl,2 (A10) = − i~ 2 B̄ (a) kl 1 (π~)N ∫ dy1 · · · dyNe −2ip·y/~ [ ∂g∗N (q− y) ∂ql ∂gN(q+ y) ∂qk − ∂g∗N (q− y) ∂qk ∂gN(q+ y) ∂ql ] − i~ 2 B̄ (b) kl 1 (π~)N ∫ dy1 · · · dyNe −2ip·y/~ [ ∂g∗N (q− y) ∂ql g(q+ y) − g∗N(q − y) ∂gN (q+ y) ∂ql ] (− 2ipk ~ ). In Eq.(A10), an identical relationship between differen- tial with respect to yk(yl) and with respect to qk(ql) is used. Due to the symmetric property of B̄ (a) kl and summation of ∑ kl in Eq.(A3), the contributions from first term in Eq.(A10) vanishes. Thus, the kinetic part can be written as, Mkl = −pkB̄ (b) kl ∂fN ∂ql (A11) For the potential part in Eq. (A3), a Taylor series with respect to q is performed with potential fields VN (q+ y) and VN (q− y). N = 1 (π~) N ∫ dy1 · · · dyNe −2ip·y/~ (A12) × i ~ [V (q− y)− V (q+ y)] g∗N (q− y)gN(q + y) = 1 (π~) N ∫ dy1 · · · dyNe −2ip·y × 2i ~ ∑ λ 1 λ1! · · ·λN ! ∂λ1+···λNVN (q) ∂qλ11 · · · ∂q λN N yλ11 · · · y λN N g ∗ N (q− y) gN (q+ y) = 2i ~ ∑ λ ( ~ 2i )λ1+···+λN 1 λ1! · · ·λN ! × ∂λ1+···+λNVN (q) ∂qλ11 · · · ∂q λN N ∂λ1+···+λN fN ∂pλ11 · · ·∂p λN N Finally, Eq. (A3) could be written as ∂fN(q,p, t) ∂t = − ∑ kl pkB̄ (b) kl (q) ∂fN ∂ql (A13) + ∑ λ ( ~ 2i )λ1+···+λN−1 1 λ1! · · ·λN ! × ∂λ1+···+λNVN (q) ∂qλ11 · · ·∂q λN N ∂λ1+···+λN fN ∂pλ11 · · · ∂p λN N = − ∑ kl pkB̄ (b) kl (q) ∂fN ∂ql + N ∑ l ∂V ∂ql · ∂fN ∂pl + ∑ λ ( ~ 2i )λ1+···+λN−1 1 λ1! · · ·λN ! × ∂λ1+···+λNVN (q) ∂qλ11 · · ·∂q λN N ∂λ1+···+λN fN ∂pλ11 · · · ∂p λN N .. where all λ1, · · · , λN are non-negative integer values and λ1 + · · · + λN is an odd number. This is a general form of transport equation for TDGCM+GOA. When the potential is calculated from the two-body interaction, i.e., V (q1, q2, · · · , qN ) = ∑ i≤j V (qi, qj), (A14) the equation is further simplified as, ∂fN ∂t = − ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql + ∑ k≤m ∂Vkm ∂qk ∂fN ∂pk . (A15) 9 Appendix B: Time evolution of s-body phase space density distribution The s-body phase space density is defined as, fs(q1, · · · , qs, p1, · · · , ps) (B1) = 1 ΩN−s ∫ fN (q1, · · · , qN , p1, · · · , pN )dΓs+1 · · · dΓN , dΓi = dqidpi. When the potential is calculated from the two-bodyin- teraction, the time-dependent probability distribution fs is obtained by the similar strategy of the derivation of the Bogoliubov-Born-Green-Kirkood-Yvon (BBGKY) hierarchy, i.e., ∂fs(q1, · · · , qs, p1, · · · , ps) ∂t = 1 ΩN−s ∫ dΓs+1 · · · dΓN ∂fN(q1, · · · , qN , p1, · · · , pN ) ∂t = 1 ΩN−s ∫ dΓs+1 · · · dΓN [ − ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql + ∑ k≤m ∂Vkm ∂qk ∂fN ∂pk ] . (B2) Different than the transport equation for fixed-mass many-particle system, the inertia Bkl(q) depends on the coordinate and it causes the equation much more com- plexity. To avoid the difficulty caused by B̄kl(q), we move the B̄kl out from the integration by using the equivalent of the integration, i.e., ∫ B̄kl(q)O(q,p)dΓs+1 · · · dΓN = (B3) B̄kl(qs, q ∗ s+1, · · · , q ∗ N ) ∫ O(q,p)dΓs+1 · · · dΓN . Consequently, the derivation will be simplified. For the first term in r.h.s of Eq. (B2), we perform one partial integration with respect to ql and the result is I1 =− 1 ΩN−s ∫ [ ∑ kl B̄ (b) kl (q)pk ∂fN ∂ql ] dΓs+1 · · · dΓN =− ∑ kl B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N ) × 1 ΩN−s ∫ [ pk ∂fN ∂ql ] dΓs+1 · · · dΓN =− s ∑ k=1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )pk ∂fs ∂ql − N ∑ k=s+1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N ) × 1 ΩN−s ∫ pk ∂fN ∂ql dΓs+1 · · · dΓN =− s ∑ k=1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )pk ∂fs ∂ql − δI1 (B4) In above derivation, we use the term of ∑N l=s+1 B̄ (b) kl (q)pkfN | ql→∞ ql→−∞ = 0. This is because the finite values of fN , which means fN (q) = 0 at q → ±∞. The δI1 is defined as, δI1 = N ∑ k=s+1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N ) (B5) · 1 ΩN−s ∫ pk ∂fN ∂ql dΓs+1 · · · dΓN = N ∑ k=s+1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )p ∗ k ∂fs ∂ql which is connected to the N -body density distribution fN , and reflects how the momentum field evolve with the coordinates. The second term in Eq. (B2) reads, I3 = 1 ΩN−s ∫ ∑ k≤m ∂Vkm ∂qk ∂fN ∂pk dΓs+1 · · · dΓN (B6) = 1 ΩN−s ∫ ∑ 1≤k≤m≤s ∂Vkm ∂qk ∂fN ∂pk dΓs+1 · · · dΓN + (N − s) 1 ΩN−s ∫ l ∑ k=1 ( ∂Vk,s+1 ∂qk )( ∂fN ∂pk ) dΓs+1 · · · dΓN = ∑ 1≤k≤m≤s ∂Vkm ∂qk ∂fs ∂pk + N − s Ω ∫ s ∑ k=1 ( ∂Vk,s+1 ∂qk ) × ( ∂fs+1 ∂pk ) dΓs+1 10 Finally, the time evolution of fs is, ∂fs ∂t = − s ∑ k=1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )pk ∂fs ∂ql (B7) − N ∑ k=s+1 s ∑ l=1 B̄ (b) kl (qs, q ∗ s+1, . . . , q ∗ N )p ∗ k ∂fs ∂ql − ∑ 1≤k≤m≤s ∂Vkm ∂qk ∂fs ∂pk + N − s Ω ∫ s ∑ k=1 ( ∂Vk,s+1 ∂qk ) × ( ∂fs+1 ∂pk ) dΓs+1 [1] O. Hahn and F. Strassmann, Naturwissenschaften 27, 11 (1939). [2] A. Baran, M. Kowal, P.-G. Reinhard, L. Robledo, A. Staszczak, and M. Warda, Nuclear Physics A 944, 442 (2015). [3] E. Fermi, E. Amaldi, B. Pontecorvo, F. Rasetti, and E. Segrä, Ricerca. Scientifica 5, 282 (1934). [4] W. A. Weber, J. Czernin, C. J. Anderson, R. D. Badawi, H. Barthel, F. Bengel, L. Bodei, I. Buvat, M. DiCarli, M. M. Graham, et al., Journal of Nuclear Medicine 61, 263S (2020). [5] A. Nichols, M. Verpelli, and D. Aldama, Handbook of nuclear data for safeguards: database extensions, August 2008 (IAEA, 2008). [6] N. Bohr and J. A. Wheeler, Physical Review 56, 426 (1939). [7] N. Schunck and L. M. Robledo, Reports on Progress in Physics 79, 116301 (2016). [8] A. N. Andreyev, K. Nishio, and K. H. Schmidt, Reports on Progress in Physics 81, 016301 (2017). [9] K.-H. Schmidt and B. Jurado, Reports on Progress in Physics 81, 106301 (2018). [10] M. Bender, R. Bernard, G. Bertsch, et al., Journal of Physics G: Nuclear and Particle Physics 47, 113002 (2020). [11] A. Bulgac, S. Jin, and I. Stetcu, Frontiers in Physics 8, 63 (2020). [12] N. Schunck and D. Regnier, Progress in Particle and Nuclear Physics 125, 103963 (2022). [13] J. Randrup and P. Möller, Physical Review Letters 106, 132503 (2011). [14] J. Randrup, P. Möller, and A. J. Sierk, Physical Review C 84, 034613 (2011). [15] M. R. Mumpower, P. Jaffke, M. Verriere, and J. Randrup, Physical Review C 101, 054607 (2020). [16] C. Ishizuka, M. D. Usang, F. A. Ivanyuk, J. A. Maruhn, K. Nishio, and S. Chiba, Physical Review C 96, 064616 (2017). [17] L.-L. Liu, X.-Z. Wu, Y.-J. Chen, C.-W. Shen, Z.-X. Li, and Z.-G. Ge, Physical Review C 99, 044614 (2019). [18] K. Shimada, C. Ishizuka, F. A. Ivanyuk, and S. Chiba, Physical Review C 104, 054609 (2021). [19] A. Bulgac, P. Magierski, K. J. Roche, and I. Stetcu, Physical Review Letters 116, 122504 (2016). [20] A. Bulgac, S. Jin, K. J. Roche, N. Schunck, and I. Stetcu, Physical Review C 100, 034615 (2019). [21] G. Scamps, C. Simenel, and D. Lacroix, Physical Review C 92, 011602 (2015). [22] G. Scamps and C. Simenel, Nature 564, 382 (2018). [23] G. Scamps, D. Lacroix, G. F. Bertsch, and K. Washiyama, Physical Review C 85, 034328 (2012). [24] S. A. Giuliani and L. M. Robledo, Physics Letters B 787, 134 (2018). [25] Z. X. Ren, D. Vretenar, T. Nikšić, P. W. Zhao, J. Zhao, and J. Meng, Physical Review Letters 128, 172501 (2022). [26] Y. Tanimura, D. Lacroix, and S. Ayik, Phys. Rev. Lett. 118, 152501 (2017). [27] P. Ring and P. Schuck, The nuclear many-body problem (Springer Science & Business Media, 2004). [28] H. J. Krappe and K. Pomorski, Theory of nuclear fission: a textbook, Vol. 838 (Springer Science & Business Media, 2012). [29] M. Verriere and D. Regnier, Frontiers in Physics 8, 233 (2020). [30] D. Regnier, N. Dubray, N. Schunck, and M. Verrière, Physical Review C 93, 054611 (2016). [31] D. Regnier, M. Verrière, N. Dubray, and N. Schunck, Computer Physics Communications 200, 350 (2016). [32] D. Regnier, N. Dubray, M. Verrière, and N. Schunck, Computer Physics Communications 225, 180 (2018). [33] P. Möller, D. Madland, A. Sierk, and A. Iwamoto, Nature 409, 785 (2001). [34] W. Younes and D. Gogny, Physical Review C 80, 054313 (2009). [35] N. Dubray and D. Regnier, Computer Physics Communications 183, 2035 (2012). [36] N. Schunck, D. Duke, H. Carr, and A. Knoll, Physical Review C 90, 054305 (2014). [37] H. Eslamizadeh and H. Raanaei, Physics Letters B 783, 163 (2018). [38] D. Regnier, N. Dubray, N. Schunck, and M. Verrière, in EPJ Web of Conferences , Vol. 146 (2017) p. 04043. [39] J. Zhao, T. Nikšić, and D. Vretenar, Physical Review C 104, 044612 (2021). [40] H. Tao, J. Zhao, Z. P. Li, T. Nikšić, and D. Vretenar, Physical Review C 96, 024319 (2017). [41] J. Zhao, T. Nikšić, D. Vretenar, and S.-G. Zhou, Physical Review C 101, 064605 (2020). [42] J. W. Negele, S. E. Koonin, P. Möller, J. R. Nix, and A. J. Sierk, Phys. Rev. C 17, 1098 (1978). https://doi.org/10.1007/BF01488241 https://doi.org/https://doi.org/10.1016/j.nuclphysa.2015.06.002 https://ui.adsabs.harvard.edu/abs/1934RicSc...5..282F https://doi.org/10.2967/jnumed.120.254532 https://doi.org/10.1103/PhysRev.56.426 https://doi.org/10.1088/0034-4885/79/11/116301 https://doi.org/10.1088/1361-6633/aa82eb https://doi.org/10.1088/1361-6633/aacfa7 https://doi.org/10.1103/PhysRevC.96.024319 https://doi.org/10.3389/fphy.2020.00063 https://doi.org/https://doi.org/10.1016/j.ppnp.2022.103963 https://doi.org/10.1103/PhysRevLett.106.132503 https://doi.org/10.1103/PhysRevC.84.034613 https://doi.org/10.1103/PhysRevC.101.054607 https://doi.org/10.1103/PhysRevC.96.064616 https://doi.org/10.1103/PhysRevC.99.044614 https://doi.org/10.1103/PhysRevC.104.054609 https://doi.org/10.1103/PhysRevLett.116.122504 https://doi.org/10.1103/PhysRevC.100.034615 https://doi.org/10.1103/PhysRevC.92.011602 https://doi.org/10.1038/s41586-018-0780-0 https://doi.org/10.1103/PhysRevC.85.034328 https://doi.org/https://doi.org/10.1016/j.physletb.2018.10.045 https://doi.org/10.1103/PhysRevLett.128.172501 https://doi.org/10.1103/PhysRevLett.118.152501 https://doi.org/10.3389/fphy.2020.00233 https://doi.org/10.1103/PhysRevC.93.054611 https://doi.org/10.1016/j.cpc.2015.11.013 https://doi.org/10.1016/j.cpc.2017.12.007 https://doi.org/10.1038/35057204 https://doi.org/10.1103/PhysRevC.80.054313 https://doi.org/https://doi.org/10.1016/j.cpc.2012.05.001https://doi.org/10.1103/PhysRevC.90.054305 https://doi.org/https://doi.org/10.1016/j.physletb.2018.06.050 https://doi.org/10.1051/epjconf/201714604043 https://doi.org/10.1103/PhysRevC.104.044612 https://doi.org/10.1103/PhysRevC.96.024319 https://doi.org/10.1103/PhysRevC.101.064605 https://doi.org/10.1103/PhysRevC.17.1098 11 [43] E. Wigner, Physical Review 40, 749 (1932). [44] C.-Y. Wong, Physical Review C 25, 1460 (1982). [45] H. Wolter, M. Colonna, D. Cozma, P. Danielewicz, C. M. Ko, R. Kumar, A. Ono, M. B. Tsang, J. Xu, Y.-X. Zhang, E. Bratkovskaya, Z.-Q. Feng, T. Gaitanos, A. Le Fèvre, N. Ikeno, Y. Kim, S. Mallik, P. Napolitani, D. Oli- inychenko, T. Ogawa, M. Papa, J. Su, R. Wang, Y.-J. Wang, J. Weil, F.-S. Zhang, G.-Q. Zhang, Z. Zhang, J. Aichelin, W. Cassing, L.-W. Chen, H.-G. Cheng, H. Elfner, K. Gallmeister, C. Hartnack, S. Hashimoto, S. Jeon, K. Kim, M. Kim, B.-A. Li, C.-H. Lee, Q.-F. Li, Z.-X. Li, U. Mosel, Y. Nara, K. Niita, A. Ohnishi, T. Sato, T. Song, A. Sorensen, N. Wang, and W.-J. Xie, Progress in Particle and Nuclear Physics 125, 103962 (2022). https://doi.org/10.1103/PhysRev.40.749 https://doi.org/10.1103/PhysRevC.25.1460 https://doi.org/https://doi.org/10.1016/j.ppnp.2022.103962