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

Pairing effects on pure rotational energy of nuclei

K. Abe Department of Physics, Graduate School of Science and Engineering, Chiba University, Yayoi-cho 1-33, Inage, Chiba 263-8522, Japan    H. Nakada [email protected] Department of Physics, Graduate School of Science, Chiba University, Yayoi-cho 1-33, Inage, Chiba 263-8522, Japan
Abstract

By applying the angular-momentum projection (AMP) to the self-consistent axial mean-field solutions with the semi-realistic effective Hamiltonian M3Y-P6, the pairing effects on the pure rotational energy of nuclei, i.e., the rotational energy at a fixed intrinsic state, have been investigated. While it was shown at the Hartree-Fock (HF) level that the individual terms of the Hamiltonian contribute to the rotational energy with ratios insensitive to nuclides except for light or weakly-deformed nuclei, the pair correlations significantly change the contributions, even for the well-deformed heavy nuclei. The contribution of the interaction to the rotational energy is found to correlate well with the degree of proximity between nucleons, which is measured via the expectation value that two nucleons exist at the same position. While the nucleons slightly spread as the angular momentum increases at the HF level, accounting for the positive (negative) contribution of the attractive (repulsive) components of the interaction, the pair correlations reduce or invert the effect.

I Introduction

It is known that the rotational band appears in a number of nuclei, giving the excitation energy Ex(J+)=J(J+1)/(2)E_{\mathrm{x}}(J^{+})=J(J+1)/(2\mathcal{I}) above the ground state (g.s.) [1, 2]. These energy spectra indicate that the intrinsic state of the nucleus is deformed and rotates with the moment of inertia (MoI) \mathcal{I}. From a microscopic standpoint, nuclei are well described in the self-consistent mean-field (MF) theories, such as the Hartree-Fock (HF) and the Hartree-Fock-Bogoliubov (HFB) approximations [3], and the deformed intrinsic states of nuclei are obtained as MF solutions in which the rotational symmetry is spontaneously broken. The Nambu-Goldstone (NG) mode is accompanied by symmetry breaking, and it restores the corresponding symmetry in energy eigenstates. The restoration of the rotational symmetry corresponds with a whole rotation of a deformed intrinsic state. The superposition of the degenerate intrinsic states along the NG mode derives the angular-momentum projection (AMP) [3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23]. The AMP accounts for the J(J+1)J(J+1) rule of the excitation energy under a reasonable approximation for well-deformed heavy nuclei [3, 4, 5, 7, 8, 10]. However, rotational spectra have also been observed in light nuclei including those far off the β\beta-stability. It will deserve reinvestigating from a general perspective how the rotational energy of nuclei arises, not restricting ourselves to well-deformed heavy nuclei.

In classical mechanics, the rotational energy arises from the kinetic energy, proportional to 𝐉2/(2)\mathbf{J}^{2}/(2\mathcal{I}) for a rigid rotor with the angular momentum 𝐉\mathbf{J}, in which the distance between constituent particles is invariant. On the other hand, the above-mentioned rotational energy of nuclei should be formed from the effective Hamiltonian, including the nucleonic interaction. In the Thouless-Valatin formula for the MoI [24], which is derived from the random-phase approximation (RPA), the g.s. correlations due to the interaction are taken into account. In Ref. [23], we investigated the composition of the rotational energy of various nuclei by applying the AMP to the HF solutions. When the rotational energy is decomposed into contributions of the individual terms of the Hamiltonian, those of the central forces are sizable, although the kinetic energies carry the major part and are close to the rigid-rotor values. The ratios of the individual terms of the Hamiltonian to the total rotational energy are insensitive to nuclides and deformation, as long as they are well-deformed and not very light. In contrast, the ratios significantly depend on nuclei and deformation for light or weakly-deformed nuclei.

It is known that the pair correlations in nuclei significantly influence the rotational energy. Within the cranking model [3], the Belyaev formula [25] accounts for the reduction of the MoI compared with the Inglis formula [26] due to the pair correlations. Conversely, the pair correlations could be influenced by the rotation; the pair correlation is suppressed by the cranking effect accompanied by the breaking of the time-reversal symmetry, known as the Mottelson-Valatin effect [27].

The effects of the pair correlations on the rotational energy of nuclei are investigated in this paper, extending the study in Ref. [23]. The AMP is applied to the self-consistent solutions of the Bardeen-Cooper-Schrieffer approximation on top of the axial HF orbitals (HF+BCS) as well as the axial HFB solutions. The Michigan-three-range-Yukawa (M3Y)-type nucleonic interactions [28, 29, 30] have been developed for the self-consistent MF calculations. With phenomenological modification of the original M3Y interaction obtained from the GG-matrix calculations [31, 32, 33, 34], they may be regarded as semi-realistic effective interactions. They have been pointed out to be free from most of the instabilities in the nuclear matter response functions, which sometimes occur in other interactions [35]. The M3Y-P6 parameter-set is employed [29, 30], which well describes the magic numbers of nuclei over a wide range of the nuclear chart [36] and the deformation of nuclei [37, 38].

In the present study, the AMP is applied to the MF wave functions obtained by self-consistent axial-MF calculations with the M3Y-P6 interaction, including the pair correlations. We restrict ourselves to the energies arising solely from the rotation of the MF solutions in the projection-after-variation (PAV) framework as in Ref. [23], which may be called pure rotational energy111 Since the state is restricted to the axial HF state as in Ref. [4], the corresponding energy was called Peierls-Yoccoz rotational energy in Ref. [23]., separating them from the counter effects of the rotation on the intrinsic states. It should be kept in mind that the pure rotational energy is not necessarily enough to describe the rotational spectra in actual nuclei [6], because the rotation may significantly affect the intrinsic state as handled in the cranking model [6, 10, 3] and the variation-after-projection (VAP) schemes [3].

The correlations of the spatial proximity between two neutrons (protons) have attracted interest as di-neutron (di-proton) correlations [39]. The spatial correlations between nucleons could be relevant to the contributions of the nucleonic interaction to the rotational energy, giving us an insight into why the interaction contributes.

II Theoretical framework and numerical method

II.1 AMP and cumulant expansion

The AMP is the method by which an intrinsic state is projected on angular-momentum eigenstates [3]. We consider an axially-symmetric intrinsic state, which is an eigenstate of J^z\hat{J}_{z} with the eigenvalue M=0M=0. The intrinsic state |Φ0\ket{\Phi_{0}} is expanded by angular-momentum eigenstates |J0\ket{J0}, |Φ0=J|J0J0|Φ0\ket*{\Phi_{0}}=\sum_{J}\ket*{J0}\innerproduct*{J0}{\Phi_{0}}, where we omit indices other than JJ and M(=0)M\,(=0) for simplicity. The Wigner (small) dd function [3, 40, 41] of the angle β\beta, dMK(J)(β):=JM|eiJ^yβ|JKd^{(J)}_{MK}(\beta):=\matrixelement*{JM}{e^{-i\hat{J}_{y}\beta}}{JK}, takes a real number under the standard phase convention. The expectation values of the scalar operator 𝒮^\hat{\mathcal{S}} on the angular-momentum eigenstates are obtained as follows [3, 23]:

J|𝒮^|J=0π/2𝑑βsinβd00(J)(β)Φ0|𝒮^eiJ^yβ|Φ00π/2𝑑βsinβd00(J)(β)Φ0|eiJ^yβ|Φ0.\expectationvalue*{\hat{\mathcal{S}}}{J}=\,\frac{\displaystyle\int_{0}^{\pi/2}d\beta\sin\beta\,d^{(J)}_{00}(\beta)\expectationvalue*{\hat{\mathcal{S}}\,e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}}{\displaystyle\int_{0}^{\pi/2}d\beta\sin\beta\,d^{(J)}_{00}(\beta)\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}}. (1)

Equation (1) is the basic formula of the AMP. We here omit the index M(=0)M\,(=0) on the left-hand side (LHS) of Eq. (1). The energy difference J|H^|J0|H^|0\expectationvalue*{\hat{H}}{J}-\expectationvalue*{\hat{H}}{0}, where H^\hat{H} is the Hamiltonian, is the pure rotational energy.

It has been pointed out in Ref. [23] that the cumulant expansion can straightforwardly be applied to the right-hand side (RHS) of Eq. (1). We expand d00(J)(β)d^{(J)}_{00}(\beta) by the power series of β\beta,

d00(J)(β)=n=0c2nβ2n;c2n=()n(2n)!J0|J^y 2n|J0.d^{(J)}_{00}(\beta)=\sum_{n=0}^{\infty}c_{2n}\beta^{2n};\,\quad c_{2n}=\,\frac{(-)^{n}}{(2n)!}\expectationvalue*{\hat{J}_{y}^{\,2n}}{J0}. (2)

Note c0=1c_{0}=1, c2=J(J+1)/(2! 2)c_{2}=-J(J+1)/(2!\,2) and c2n=𝒪(J2n)c_{2n}=\mathcal{O}(J^{2n}). The function 𝒮01(β)\mathcal{S}^{01}(\beta) is defined and also expanded by the power series of β\beta as follows:

𝒮01(β):=Φ0|𝒮^eiJ^yβ|Φ0Φ0|eiJ^yβ|Φ0=n=0s2nβ2n;s2n=()n(2n)!Φ0|𝒮^;J^y;;J^y2n|Φ0cum,\mathcal{S}^{01}(\beta):=\,\frac{\expectationvalue*{\hat{\mathcal{S}}\,e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}}{\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}}=\sum_{n=0}^{\infty}s_{2n}\beta^{2n};\quad s_{2n}=\,\frac{(-)^{n}}{(2n)!}\expectationvalue*{\hat{\mathcal{S}};\underbrace{\hat{J}_{y};\cdots;\hat{J}_{y}}_{2n}}{\Phi_{0}}_{\mathrm{cum}}, (3)

with the cumulant defined for a set of commutable operators {X^i;i=1,2,,n}\{\hat{X}_{i};i=1,2,\cdots,n\} [42],

X^1;;X^ncum:=t1tnlnexp(i=1ntiX^i)|t1==tn=0.\expectationvalue*{\hat{X}_{1};\cdots;\hat{X}_{n}}_{\mathrm{cum}}:=\frac{\partial}{\partial t_{1}}\cdots\frac{\partial}{\partial t_{n}}\left.\ln\expectationvalue{\exp\left(\sum_{i=1}^{n}t_{i}{\hat{X}}_{i}\right)}\right|_{t_{1}=\cdots=t_{n}=0}. (4)

The low-order cumulants are represented as

s0=Φ0|𝒮^|Φ0,s2=12!C[𝒮^,J^y 2],s4=14!(C[𝒮^,J^y 4]6C[𝒮^,J^y 2](σ[J^y])2).s_{0}=\expectationvalue*{\hat{\mathcal{S}}}{\Phi_{0}},\quad s_{2}=-\frac{1}{2!}\,C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}],\quad s_{4}=\,\frac{1}{4!}\left(C[\hat{\mathcal{S}},\hat{J}_{y}^{\,4}]-6\,C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}](\sigma[\hat{J}_{y}])^{2}\right). (5)

Here σ[A^]\sigma[\hat{A}] is the fluctuation of an operator A^\hat{A}, and C[A^,B^]C[\hat{A},\hat{B}] is the correlation function of operators A^\hat{A} and B^\hat{B},

C[A^,B^]:=Φ0|A^B^|Φ0Φ0|A^|Φ0Φ0|B^|Φ0,σ[A^]:=C[A^,A^].C[\hat{A},\hat{B}]:=\expectationvalue*{\hat{A}\hat{B}}{\Phi_{0}}-\expectationvalue*{\hat{A}}{\Phi_{0}}\expectationvalue*{\hat{B}}{\Phi_{0}},\quad\sigma[\hat{A}]:=\sqrt{C[\hat{A},\hat{A}]}. (6)

Then Eq. (1) is exactly expressed in terms of the cumulants,

J|𝒮^|J=m,n=0c2ms2nΛ2m+2n=0c2Λ2,\expectationvalue*{\hat{\mathcal{S}}}{J}=\,\frac{\displaystyle\sum_{m,n=0}^{\infty}c_{2m}s_{2n}\varLambda_{2m+2n}}{\displaystyle\sum_{\ell=0}^{\infty}c_{2\ell}\varLambda_{2\ell}}, (7)

where

N2n:=0π/2𝑑βsinββ2nΦ0|eiJ^yβ|Φ0,Λ2n:=N2nN0,(n=0,1,2,).N_{2n}:=\int_{0}^{\pi/2}d\beta\sin\beta\,\beta^{2n}\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}},\quad\varLambda_{2n}:=\frac{N_{2n}}{N_{0}},\quad(n=0,1,2,\cdots). (8)

Expansion of Eq. (7) with respect to c2nc_{2n} yields the g.s. expectation value and the J(J+1)J(J+1) rule for J|𝒮^|J\expectationvalue*{\hat{\mathcal{S}}}{J},

J|𝒮^|J=\displaystyle\expectationvalue*{\hat{\mathcal{S}}}{J}= 0|𝒮^|0+J(J+1)2[𝒮]+;\displaystyle\,\expectationvalue*{\hat{\mathcal{S}}}{0}+\frac{J(J+1)}{2\,\mathcal{I}[\mathcal{S}]}+\cdots\,; (9a)
0|𝒮^|0:=\displaystyle\expectationvalue*{\hat{\mathcal{S}}}{0}:= n=0s2nΛ2n,1[𝒮^]:=n=1s2n[12(Λ2n+2Λ2nΛ2)].\displaystyle\,\displaystyle\sum_{n=0}^{\infty}s_{2n}\varLambda_{2n},\quad\frac{1}{\mathcal{I}[\hat{\mathcal{S}}]}:=\,\sum_{n=1}^{\infty}s_{2n}\left[-\frac{1}{2}(\varLambda_{2n+2}-\varLambda_{2n}\varLambda_{2})\right]. (9b)

The MoI of Peierls and Yoccoz [4, 7] is obtained by neglecting the s2ns_{2n} terms with n2n\geq 2 for 𝒮^=H^\hat{\mathcal{S}}=\hat{H}. We shall call the approximation of Eq. (9) up to the s2s_{2} terms Peierls-Yoccoz (PY) formula. It has been shown that the higher-order terms are not negligible in light nuclei or weakly-deformed intrinsic states [23].

The cumulant of Eq. (4) can be generalized as

X^1;;X^ncum:=t1tnlni=1nexp(tiX^imissing)|t1==tn=0,\expectationvalue*{\hat{X}_{1};\cdots;\hat{X}_{n}}_{\mathrm{cum}}:=\frac{\partial}{\partial t_{1}}\cdots\frac{\partial}{\partial t_{n}}\left.\ln\expectationvalue{\prod_{i=1}^{n}\exp\Big(t_{i}\hat{X}_{i}\Big{missing})}\right|_{t_{1}=\cdots=t_{n}=0}, (10)

which distinguishes the ordering of the operators on the lhs and therefore is applicable even when the operators {X^i;i=1,2,,n}\{\hat{X}_{i};i=1,2,\cdots,n\} are not commutable one another. The expansion is then extended to triaxially-deformed intrinsic states.

II.2 Effective Hamiltonian

The nuclear effective Hamiltonian has translational, rotational, parity, and time-reversal symmetries, with the Galilean invariance and the number conservation. We assume that the individual terms of the Hamiltonian also have isospin symmetries except for the Coulomb force. The Hamiltonian is composed of the kinetic energy K^=i𝕡i2/(2M)\hat{K}=\sum_{i}\mathbb{p}_{i}^{2}/(2M), the effective nucleonic interaction V^nucl=i<jv^ij\hat{V}_{\mathrm{nucl}}=\sum_{i<j}\hat{v}_{ij}, the Coulomb interaction between protons V^Coul\hat{V}_{\mathrm{Coul}}, and the center-of-mass term H^c.m.=2/(2AM)\hat{H}_{\mathrm{c.m.}}=\mathbb{P}^{2}/(2AM) with the total momentum =i𝕡i\mathbb{P}=\sum_{i}\mathbb{p}_{i} and the mass number A=Z+NA=Z+N,

H^=K^+V^nucl+V^CoulH^c.m..\hat{H}=\hat{K}+\hat{V}_{\mathrm{nucl}}+\hat{V}_{\mathrm{Coul}}-\hat{H}_{\mathrm{c.m.}}. (11)

The effective nucleonic interaction for the self-consistent MF calculations consists of the following terms [28, 43]:

V^nucl=V^(C)+V^(LS)+V^(TN)+V^(Cρ);V^(X)=i<jv^ij(X),(X=C,LS,TN,Cρ),\hat{V}_{\mathrm{nucl}}=\hat{V}^{(\mathrm{C})}+\hat{V}^{(\mathrm{LS})}+\hat{V}^{(\mathrm{TN})}+\hat{V}^{(\mathrm{C\rho})}\,;\quad\hat{V}^{(\mathrm{X})}=\sum_{i<j}\hat{v}_{ij}^{(\mathrm{X})},\quad(\mathrm{X}=\mathrm{C,LS,TN,C\rho}), (12)

where V^(C)\hat{V}^{(\mathrm{C})}, V^(LS)\hat{V}^{(\mathrm{LS})} and V^(TN)\hat{V}^{(\mathrm{TN})} are the central, LS and tensor forces. For the individual terms of Eq. (12), we consider the following forms:

v^ij(C)=n(tn(SE)PSE+tn(TE)PTE+tn(SO)PSO+tn(TO)PTO)fn(C)(rij),v^ij(LS)=n(tn(LSE)PTE+tn(LSO)PTO)fn(LS)(rij)𝕃ij(𝕤i+𝕤j),v^ij(TN)=n(tn(TNE)PTE+tn(TNO)PTO)fn(TN)(rij)rij2Sij,v^ij(Cρ)=(tρ(SE)PSE[ρ(𝕣i)]α(SE)+tρ(TE)PTE[ρ(𝕣i)]α(TE))δ(𝕣ij),\begin{split}\hat{v}_{ij}^{(\mathrm{C})}=&\sum_{n}\left(t^{\mathrm{(SE)}}_{n}P_{\mathrm{SE}}+t^{\mathrm{(TE)}}_{n}P_{\mathrm{TE}}+t^{\mathrm{(SO)}}_{n}P_{\mathrm{SO}}+t^{\mathrm{(TO)}}_{n}P_{\mathrm{TO}}\right)f^{\mathrm{(C)}}_{n}(r_{ij}),\\ \hat{v}_{ij}^{(\mathrm{LS})}=&\sum_{n}\left(t^{\mathrm{(LSE)}}_{n}P_{\mathrm{TE}}+t^{\mathrm{(LSO)}}_{n}P_{\mathrm{TO}}\right)f^{\mathrm{(LS)}}_{n}(r_{ij})\,\mathbb{L}_{ij}\cdot(\mathbb{s}_{i}+\mathbb{s}_{j}),\\ \hat{v}_{ij}^{(\mathrm{TN})}=&\sum_{n}\left(t^{\mathrm{(TNE)}}_{n}P_{\mathrm{TE}}+t^{\mathrm{(TNO)}}_{n}P_{\mathrm{TO}}\right)f^{\mathrm{(TN)}}_{n}(r_{ij})\,r^{2}_{ij}\,S_{ij},\\ \hat{v}_{ij}^{(\mathrm{C\rho})}=&\left(t^{\mathrm{(SE)}}_{\rho}P_{\mathrm{SE}}\cdot\left[\rho(\mathbb{r}_{i})\right]^{\alpha^{(\mathrm{SE})}}+t^{\mathrm{(TE)}}_{\rho}P_{\mathrm{TE}}\cdot\left[\rho(\mathbb{r}_{i})\right]^{\alpha^{(\mathrm{TE})}}\right)\delta(\mathbb{r}_{ij}),\end{split} (13)

where 𝕣ij:=𝕣i𝕣j\mathbb{r}_{ij}:=\mathbb{r}_{i}-\mathbb{r}_{j}, rij:=|𝕣ij|r_{ij}:=|\mathbb{r}_{ij}|, 𝕣^ij:=𝕣ij/rij\mathbb{\hat{r}}_{ij}:=\mathbb{r}_{ij}/r_{ij}, 𝕡ij:=(𝕡i𝕡j)/2\mathbb{p}_{ij}:=(\mathbb{p}_{i}-\mathbb{p}_{j})/2, 𝕃ij:=𝕣ij×𝕡ij\mathbb{L}_{ij}:=\mathbb{r}_{ij}\times\mathbb{p}_{ij}, Sij:=4[3(𝕤i𝕣^ij)(𝕤j𝕣^ij)𝕤i𝕤j]S_{ij}:=4\big{[}3(\mathbb{s}_{i}\cdot\mathbb{\hat{r}}_{ij})(\mathbb{s}_{j}\cdot\mathbb{\hat{r}}_{ij})-\mathbb{s}_{i}\cdot\mathbb{s}_{j}\big{]}, and ρ(𝐫)\rho(\mathbf{r}) is the nucleon density. The central density-dependent term is distinguished from V^(C)\hat{V}^{(\mathrm{C})} and represented by V^(Cρ)\hat{V}^{(\mathrm{C\rho})}. The projection operators on the singlet-even (SE), triplet-even (TE), singlet-odd (SO) and triplet-odd (TO) two-nucleon states are denoted by PYP_{\mathrm{Y}} (Y=SE,TE,SO,TO\mathrm{Y}=\mathrm{SE},\mathrm{TE},\mathrm{SO},\mathrm{TO}).

We adopt the semi-realistic interaction M3Y-P6 [29, 30, 37, 38], in which the Yukawa function fn(X)(r)=eμn(X)r/(μn(X)r)f_{n}^{(\mathrm{X})}(r)=e^{-\mu_{n}^{(\mathrm{X})}r}/(\mu_{n}^{(\mathrm{X})}r) is used for the radial functions, except for v^ij(Cρ)\hat{v}_{ij}^{(\mathrm{C\rho})}. The longest-range term in v^ij(C)\hat{v}^{\mathrm{(C)}}_{ij} is fixed to be that of the one-pion exchange potential (OPEP). This central OPEP, denoted by V^(OPEP)\hat{V}^{\mathrm{(OPEP)}}, is an example of spin-dependent interactions. The values of the parameters for M3Y-P6 are given in Ref. [29].

II.3 MF calculations

The self-consistent MF calculations have been implemented via the Gaussian expansion method (GEM) [29, 30, 37, 38, 44]. The generalized Bogoliubov transformation in the HFB theory is given as [3, 45]

αi:=k(ck𝖴ki+ck𝖵ki).\alpha^{\dagger}_{i}:=\sum_{k}\left(c^{\dagger}_{k}\mathsf{U}_{ki}+c_{k}\mathsf{V}_{ki}\right). (14)

We assume the axial, time-reversal, and parity symmetry on the MF state |Φ0\ket{\Phi_{0}}, and then the variational parameters 𝖴ki\mathsf{U}_{ki} and 𝖵ki\mathsf{V}_{ki} are taken to be real numbers. Adding the terms constraining the particle numbers, we modify the Hamiltonian as

H^:=H^μp(N^pZ)μn(N^nN),\hat{H}^{\prime}:=\hat{H}-\mu_{p}(\hat{N}_{p}-Z)-\mu_{n}(\hat{N}_{n}-N), (15)

where μp\mu_{p} (μn\mu_{n}) is the chemical potential for protons (neutrons), and N^p\hat{N}_{p} (N^n\hat{N}_{n}) is the proton (neutron) number operator. The energy Φ0|H^|Φ0\expectationvalue*{\hat{H}^{\prime}}{\Phi_{0}} is minimized in the HFB calculations. We do not consider the proton-neutron pairing in this paper.

The HFB handles the pairing effects self-consistently by taking into account the influences of the pairing on the particle-hole channel. However, in the HFB, the pairing may influence the particle-hole channel and alter the HF configuration. For clarifying the effects of the pairing, the HF+BCS (Bardeen-Cooper-Schrieffer) method is useful as well, in which the HF configuration is fixed. For analyzing the pairing effects on the pure rotational energy, we introduce a parameter gg as

H^g=H^dns+gH^pair,\expectationvalue*{\hat{H}_{g}}=\expectationvalue*{\hat{H}_{\mathrm{dns}}}+g\expectationvalue*{\hat{H}_{\mathrm{pair}}}, (16)

where H^dns\expectationvalue*{\hat{H}_{\mathrm{dns}}} consists of the terms including only the density matrix, while H^pair\expectationvalue*{\hat{H}_{\mathrm{pair}}} is the pair energy containing the pairing tensor [3]. In the HF+BCS calculations, the axial-HF solution has been solved self-consistently, and the BCS equation is solved for H^g\expectationvalue*{\hat{H}_{g}} on top of the HF single-particle (s.p.) states. We denote its solution by |Φ0g\ket{\Phi_{0}}_{g}. The state |Φ0g=0\ket{\Phi_{0}}_{g=0} corresponds to the HF state, and the state |Φ0g=1\ket{\Phi_{0}}_{g=1} does to the HF+BCS state with the original Hamiltonian H^\hat{H}^{\prime}. Throughout this paper, the parameter gg is employed only in the HF+BCS scheme. We always apply H^\hat{H}^{\prime} of Eq. (15) to the HFB calculation, not using H^g\expectationvalue*{\hat{H}_{g}}.

II.4 Implementation of AMP

In this work, the PAV has been applied for the AMP calculations of Eq. (1). The number projection is not applied. In reality, the intrinsic state could gradually change with increasing JJ, often accompanied by a breakdown of the axial and the time-reversal symmetry. While these effects can be handled in the cranking model [3, 6, 10] and in the VAP approaches [3], they are ignored in the present study, and we focus on the pure rotational energy, i.e., the rotational energy arising from a fixed intrinsic state, as stated in Introduction.

The overlap function Φ0|eiJ^yβ|Φ0\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}} has been calculated from the Onishi formula [3, 9, 11, 23]. Concerning the sign problem of the Onishi formula, solutions have been proposed [46, 47] and the non-negativity of the overlap function has been proven for the HF states under the time-reversal symmetry in Appendix C of Ref. [23]. We have here confirmed via the continuity with respect to β\beta that the sign of the overlap function is positive in all the cases under consideration.

There is a problem in the density-dependent coefficients in v^ij(Cρ)\hat{v}_{ij}^{\mathrm{(C\rho)}} in the AMP calculations [19, 21, 22, 23]. In the present calculations, the standard treatment in Refs. [16, 23] has been adopted, replacing the density ρ(𝐫)\rho(\mathbf{r}) in Eq. (13) with the “generalized density ” ρ¯(𝐫;β)\bar{\rho}(\mathbf{r};\beta),

ρ¯(𝐫;β):=τσΦ0|ρ^(𝐫στ)eiJ^yβ|Φ0Φ0|eiJ^yβ|Φ0;ρ^(𝐫στ):=ψ(𝐫στ)ψ(𝐫στ),\bar{\rho}(\mathbf{r};\beta):=\sum_{\tau}\sum_{\sigma}\frac{\expectationvalue*{\hat{\rho}(\mathbf{r}\sigma\tau)\,e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}}{\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}};\quad\hat{\rho}(\mathbf{r}\sigma\tau):=\psi^{\dagger}(\mathbf{r}\sigma\tau)\psi(\mathbf{r}\sigma\tau), (17)

where ψ(𝐫στ)\psi^{\dagger}(\mathbf{r}\sigma\tau) and ψ(𝐫στ)\psi(\mathbf{r}\sigma\tau) stand for the creation and annihilation operators for the spinor field [48], which satisfy the fermionic anticommutation relations:

{ψ(x1),ψ(x2)}=δ(x1x2),{ψ(x1),ψ(x2)}={ψ(x1),ψ(x2)}=0.\{\psi(x_{1}),\psi^{\dagger}(x_{2})\}=\delta(x_{1}-x_{2}),\quad\{\psi(x_{1}),\psi(x_{2})\}=\{\psi^{\dagger}(x_{1}),\psi^{\dagger}(x_{2})\}=0. (18)

with the shorthand notation xi=(𝐫iσiτi)x_{i}=(\mathbf{r}_{i}\sigma_{i}\tau_{i}) (i=1,2i=1,2) and δ(x1x2):=δ(𝐫1𝐫2)δσ1σ2δτ1τ2\delta(x_{1}-x_{2}):=\delta(\mathbf{r}_{1}-\mathbf{r}_{2})\delta_{\sigma_{1}\sigma_{2}}\delta_{\tau_{1}\tau_{2}}. Although ρ¯(𝐫;β)\bar{\rho}(\mathbf{r};\beta) in Eq. (17) is a real number, owing to the time-reversal symmetry, ρ¯α(𝐫;β)\bar{\rho}^{\,\alpha}(\mathbf{r};\beta) is multivalued unless the power α\alpha is an integer. In the M3Y-P6 interaction, α(SE)=1\alpha^{\mathrm{(SE)}}=1 and α(TE)=1/3\alpha^{\mathrm{(TE)}}=1/3 [29]. The phase of ρ¯α(TE)(𝐫;β)\bar{\rho}^{\,\alpha^{\mathrm{(TE)}}}(\mathbf{r};\beta) has been chosen to be negative when ρ¯(𝐫;β)\bar{\rho}(\mathbf{r};\beta) is negative [23].

II.5 Degree of proximity between nucleons

To investigate the relevance of the spatial correlations between nucleons to the rotational energy, we consider an operator comprised of the two-body delta function,

D^:=i<jδ(𝐫ij).\hat{D}:=\sum_{i<j}\delta(\mathbf{r}_{ij}). (19)

The expectation value D^\expectationvalue*{\hat{D}} measures the degree how frequently two constituent nucleons sit at an equal position, representing the degree that two nucleons get spatially close to each other. We call D^\expectationvalue*{\hat{D}} degree of proximity (DoP) in this paper.

The pair-distribution function has been employed to investigate the spatial correlation between two particles [49], such as the di-neutron correlation [39]. The pair-distribution function is defined as

𝒢(x1,x2):=ψ(x1)ψ(x2)ψ(x2)ψ(x1)ρ^(x1)ρ^(x2),\mathcal{G}(x_{1},x_{2}):=\,\frac{\expectationvalue*{\psi^{\dagger}(x_{1})\psi^{\dagger}(x_{2})\psi(x_{2})\psi(x_{1})}}{\expectationvalue*{\hat{\rho}(x_{1})}\expectationvalue*{\hat{\rho}(x_{2})}}, (20)

with xi=(𝐫iσiτi)x_{i}=(\mathbf{r}_{i}\sigma_{i}\tau_{i}). Notice 𝒢(x1,x2)=0\mathcal{G}(x_{1},x_{2})=0 for x1=x2x_{1}=x_{2}, and 𝒢(x1,x2)0\mathcal{G}(x_{1},x_{2})\geq 0. The DoP is regarded as a summation of the pair-distribution function at the same position, i.e.,

D^=12τ1τ2σ1σ2d 3rψ(𝐫σ1τ1)ψ(𝐫σ2τ2)ψ(𝐫σ2τ2)ψ(𝐫σ1τ1)=12τ1τ2σ1σ2d 3rρ^(𝐫σ1τ1)ρ^(𝐫σ2τ2)𝒢(𝐫σ1τ1,𝐫σ2τ2).\begin{split}\expectationvalue*{\hat{D}}&=\frac{1}{2}\sum_{\tau_{1}\tau_{2}}\sum_{\sigma_{1}\sigma_{2}}\int d^{\,3}r\expectationvalue*{\psi^{\dagger}(\mathbf{r}\sigma_{1}\tau_{1})\psi^{\dagger}(\mathbf{r}\sigma_{2}\tau_{2})\psi(\mathbf{r}\sigma_{2}\tau_{2})\psi(\mathbf{r}\sigma_{1}\tau_{1})}\\ &=\frac{1}{2}\sum_{\tau_{1}\tau_{2}}\sum_{\sigma_{1}\sigma_{2}}\int d^{\,3}r\expectationvalue*{\hat{\rho}(\mathbf{r}\sigma_{1}\tau_{1})}\expectationvalue*{\hat{\rho}(\mathbf{r}\sigma_{2}\tau_{2})}\mathcal{G}(\mathbf{r}\sigma_{1}\tau_{1},\mathbf{r}\sigma_{2}\tau_{2}).\end{split} (21)

Since D^\hat{D} is a rotational scalar, the DoP for angular-momentum eigenstates J|D^|J\expectationvalue*{\hat{D}}{J} is calculable via Eq. (1). By applying D^PY\hat{D}P_{\mathrm{Y}} (Y=SE,TE\mathrm{Y}=\mathrm{SE},\mathrm{TE}) instead of D^\hat{D} itself, the DoP can be separated between the SE and TE channels. Restricting ii and jj on the RHS of Eq. (19), we can calculate the DoP for individual isospin components.

III Results

In the present work, the AMP of Eq. (1) is implemented on top of the axial MF solutions for deformed 12Mg [30, 37] and 40Zr [38] nuclei, including stable and unstable ones. It has been established, e.g., from the ratios of excitation energies Ex(4+)/Ex(2+)E_{\mathrm{x}}(4^{+})/E_{\mathrm{x}}(2^{+}) [2, 50, 51, 52], that 1224{}^{24}_{12}Mg, 123438{}^{34-38}_{~{}~{}~{}~{}12}Mg  [50], 4080{}^{80}_{40}Zr [53] and  40100110{}^{100-110}_{~{}~{}~{}~{}~{}~{}\,40}Zr [2, 51, 52] are well-deformed. 1240{}^{40}_{12}Mg lies near the neutron dripline [54], and a deformed halo structure has been predicted [44, 55]. In the following, we shall show the AMP results on top of the HFB solutions in   1224,34,40{}^{24,34,40}_{~{}~{}~{}~{}~{}\,\,12}Mg and 4080,100,104{}^{80,100,104}_{~{}~{}~{}~{}~{}~{}~{}~{}40}Zr, whereas we limit those on top of the HF+BCS results with varying gg to 1224{}^{24}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr, which are suffcient for the present discussion.

III.1 Influence of pairing on deformation

We define the quadrupole deformation parameter a20a_{20} as follows [1],

a20:=q01.09A5/3fm2,a_{20}:=\frac{q_{0}}{1.09A^{5/3}\,\mathrm{fm}^{2}}, (22)

where q0q_{0} is the mass quadrupole moment of the MF state [37]. In Fig.  1, the dependence of the deformation parameter a20a_{20} on the pairing strength gg in the HF+BCS solutions [see Eq. (16)] is depicted for the 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr nuclei. The state |Φ0g=0\ket{\Phi_{0}}_{g=0} corresponds with the HF state. As gg grows, the pair correlations arise at the critical values, corresponding with the normal-to-superfluid phase transition. The critical gg values for proton and neutron pairs are close, though not equal. While a20a_{20} is insensitive to gg for the 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr nuclei, a20a_{20} slightly decreases as gg grows for the 1234{}^{34}_{12}Mg nucleus. The values of a20a_{20} for the HFB minima are also shown. The a20a_{20} values for the HF+BCS solutions are close to that of the HFB solution for the 4080{}^{80}_{40}Zr nucleus, while they do not match well for the 1234{}^{34}_{12}Mg and 40100{}^{100}_{~{}40}Zr nuclei, indicating influence of the pairing on the HF configurations. Whereas the a20a_{20} values for the HFB solutions are smaller than those for the HF solutions in 1234{}^{34}_{12}Mg and 40100{}^{100}_{~{}40}Zr, the opposite is realized at 4080{}^{80}_{40}Zr. This enhancement of deformation at 4080{}^{80}_{40}Zr takes place owing to the s.p. levels near the Fermi energy. The pairing moves a portion of neutrons at the highest occupied level with Ωπ=5/2+\Omega^{\pi}=5/2^{+}, which originates from the 0g9/20g_{9/2} spherical orbit, to the lowest unoccupied level with Ωπ=1/2+\Omega^{\pi}=1/2^{+}, by which the prolate deformation is slightly enhanced. Table 1 presents the a20a_{20} values for the HF and HFB solutions at their lowest minima for 1224{}^{24}_{12}Mg, 1240{}^{40}_{12}Mg and 40104{}^{104}_{~{}40}Zr.

Refer to caption
Figure 1: The gg dependence of the deformation parameter a20a_{20} in the HF+BCS results with H^g\expectationvalue*{\hat{H}_{g}} for 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The red crosses represent the a20a_{20} values in the HFB solutions.
Table 1: a20a_{20} for the HF and HFB solutions at their lowest minima for 1224{}^{24}_{12}Mg, 1240{}^{40}_{12}Mg and 40104{}^{104}_{~{}40}Zr.
nuclide HF HFB
1224{}^{24}_{12}Mg 0.540.54 0.540.54
1240{}^{40}_{12}Mg 0.470.47 0.430.43
40104{}^{104}_{~{}40}Zr 0.460.46 0.430.43

III.2 Comparison of Ex(2+)E_{\mathrm{x}}(2^{+}) with rigid-rotor model and experiment

Let us denote the expectation value of Eq. (1) measured from that of the g.s. by

𝒮x(J+):=J|𝒮^|J0|𝒮^|0.\mathcal{S}_{\mathrm{x}}(J^{+}):=\expectationvalue*{\hat{\mathcal{S}}}{J}-\expectationvalue*{\hat{\mathcal{S}}}{0}. (23)

We explicitly attach the parity quantum number (++) on the LHS. For 𝒮^=H^\mathcal{\hat{S}}=\hat{H}^{\prime}, 𝒮x(J+)\mathcal{S}_{\mathrm{x}}(J^{+}) corresponds with the excitation energy (i.e., the pure rotational energy),

Ex(J+)=J|H^|J0|H^|0.E_{\mathrm{x}}(J^{+})=\expectationvalue*{\hat{H}^{\prime}}{J}-\expectationvalue*{\hat{H}^{\prime}}{0}. (24)
Refer to caption
Figure 2: The unprojected and the projected energies for 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The black crosses represent the unprojected HF+BCS energies depending on gg, and the red circles do the corresponding projected energies E(0+)E(0^{+}). The red crosses (circles) represent the unprojected (projected) HFB energies.

In Fig. 2, the gg dependence of the unprojected and the projected (E(0+)=0|H^|0E(0^{+})=\expectationvalue*{\hat{H}^{\prime}}{0}) g.s. energies is shown for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. Up to the critical gg value where the pair correlations arise, the energies are equal to the HF case. As gg increases from the critical values, both the unprojected and projected energies decrease for the HF+BCS solutions. The unprojected and projected energies for the HFB minima are also shown. Irrespective of the unprojected or the projected energies, the energies for the HFB solutions are close to those for the HF+BCS ones in the region between g=1.1g=1.1 and 1.21.2.

Refer to caption
Figure 3: Ex(J+)/Ex(2+)E_{\mathrm{x}}(J^{+})/E_{\mathrm{x}}(2^{+}) for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The lines display the rigid-rotor values J(J+1)/6J(J+1)/6.

As it is equal to J(J+1)/6J(J+1)/6 in the rigid-rotor model, the ratio Ex(J+)/Ex(2+)E_{\mathrm{x}}(J^{+})/E_{\mathrm{x}}(2^{+}) can be a measure of how well the rotational band develops. The gg dependence of Ex(J+)/Ex(2+)E_{\mathrm{x}}(J^{+})/E_{\mathrm{x}}(2^{+}) is shown for the 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr nuclei in Fig. 3. The ratios Ex(J+)/Ex(2+)E_{\mathrm{x}}(J^{+})/E_{\mathrm{x}}(2^{+}) are insensitive to gg and close to the J(J+1)/6J(J+1)/6 lines except for high J(10)J\,(\gtrsim 10) at 1234{}^{34}_{12}Mg. For the 1234{}^{34}_{12}Mg nucleus, the ratios Ex(J+)/Ex(2+)E_{\mathrm{x}}(J^{+})/E_{\mathrm{x}}(2^{+}) decrease for high JJ as gg increases. The deviation from the J(J+1)/6J(J+1)/6 lines indicates that the higher-c2nc_{2n} terms are not negligible in Eq. (7).

Refer to caption
Figure 4: Ex(2+)E_{\mathrm{x}}(2^{+}) for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr, which is represented by the red circles. The orange crosses represent the values for the HFB solutions. The black dashed lines display the experimental values [2], and the green dashed lines are the rigid-rotor values in Eq. (25[1].

Figure 4 shows the gg dependence of the excitation energies Ex(2+)E_{\mathrm{x}}(2^{+}) for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. Those for the HFB solutions are also shown. As gg increases, Ex(2+)E_{\mathrm{x}}(2^{+}) gets higher; the pair correlations reduce the MoI. The excitation energies Ex(2+)E_{\mathrm{x}}(2^{+}) of the HF+BCS solutions between g=1.0g=1.0 and 1.21.2 are close to those of the HFB solutions. The Ex(2+)E_{\mathrm{x}}(2^{+}) values for the HFB solutions in these nuclei are about 1.5 – 2 times higher than the rigid-rotor value [1],

Ex(RR)(J+)=J(J+1)2(RR);(RR) 0.0138A5/3[MeV1].E_{\mathrm{x}}^{(\mathrm{RR})}(J^{+})=\,\frac{J(J+1)}{2\,\mathcal{I}^{\mathrm{(RR)}}};\quad\mathcal{I}^{\mathrm{(RR)}}\approx\,0.0138\,A^{5/3}[\mathrm{MeV}^{-1}]. (25)

Compared to the experimental values, the Ex(2+)E_{\mathrm{x}}(2^{+}) value is high in the HFB solution of the 1234{}^{34}_{12}Mg nucleus, while slightly low in 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr.

III.3 Influence of higher-order terms in cumulant expansion

We next investigate pairing effects on higher-order terms in Eq. (7). In Fig. 5, the gg dependence of Λ2n\varLambda_{2n} and Λ2n+2/Λ2n\varLambda_{2n+2}/\varLambda_{2n} in Eq. (8) is shown for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. If Λ2n+2/Λ2n\varLambda_{2n+2}/\varLambda_{2n} stays small, both the J(J+1)J(J+1) rule and the PY approximation are validated. For growing gg, Λ2n\varLambda_{2n} and Λ2n+2/Λ2n\varLambda_{2n+2}/\varLambda_{2n} slightly increase for fixed nn.

Refer to caption
Figure 5: Λ2n\varLambda_{2n} and Λ2n+2/Λ2n\varLambda_{2n+2}/\varLambda_{2n} in Eq. (8) for the HF+BCS results of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The symbols correspond with the nn values indicated in the inset.

The g.s. correlation is defined and expanded by s2ns_{2n} for 𝒮^=H^\hat{\mathcal{S}}=\hat{H}^{\prime} as

ΔEg.s.c.:=Φ0|H^|Φ00|H^|0=n=1s2nΛ2n.\varDelta E_{\mathrm{g.s.c.}}:=\expectationvalue*{\hat{H}^{\prime}}{\Phi_{0}}-\expectationvalue*{\hat{H}^{\prime}}{0}=\displaystyle-\sum_{n=1}^{\infty}s_{2n}\varLambda_{2n}. (26)

To view the influence of the higher-order terms, we have calculated the following quantities:

εg.s.c.(k):=n=1ks2nΛ2nΔEg.s.c.ΔEg.s.c.,\displaystyle\varepsilon^{(k)}_{\mathrm{g.s.c.}}:=\,\frac{\displaystyle-\sum_{n=1}^{k}s_{2n}\varLambda_{2n}-\varDelta E_{\mathrm{g.s.c.}}}{\varDelta E_{\mathrm{g.s.c.}}}, (27a)
εx(k):=3n=1ks2n[12(Λ2n+2Λ2nΛ2)]Ex(2+)Ex(2+).\displaystyle\varepsilon^{(k)}_{\mathrm{x}}:=\,\frac{\displaystyle 3\sum_{n=1}^{k}s_{2n}\left[-\frac{1}{2}(\varLambda_{2n+2}-\varLambda_{2n}\varLambda_{2})\right]-E_{\mathrm{x}}(2^{+})}{E_{\mathrm{x}}(2^{+})}. (27b)

We calculate s2s_{2} and s4s_{4} from Eq. (3) via numerical differentiation. The values of εg.s.c.(k)\varepsilon^{(k)}_{\mathrm{g.s.c.}} and εx(k)\varepsilon^{(k)}_{\mathrm{x}} (k=1,2k=1,2) are shown for 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr in Fig. 6. Insensitive to gg, εg.s.c.(1)\varepsilon^{(1)}_{\mathrm{g.s.c.}} and εx(1)\varepsilon^{(1)}_{\mathrm{x}} almost vanish for 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. Namely, the contributions of the higher-s2ns_{2n} terms to ΔEg.s.c.\varDelta E_{\mathrm{g.s.c.}} and Ex(2+)E_{\mathrm{x}}(2^{+}) are negligible, and the PY formula (Eq. (9b) truncated at n=1n=1) is good. On the other hand, the εg.s.c.(1)\varepsilon^{(1)}_{\mathrm{g.s.c.}} and εx(1)\varepsilon^{(1)}_{\mathrm{x}} values become larger as gg increases in the 1234{}^{34}_{12}Mg nucleus. The pair correlations enhance the higher-s2ns_{2n} terms of the cumulant expansion, the terms including s4s_{4} in practice, in the g.s. correlations and the MoI of this light nucleus. The εg.s.c.(2)\varepsilon^{(2)}_{\mathrm{g.s.c.}} and εx(2)\varepsilon^{(2)}_{\mathrm{x}} values are vanishing.

Refer to caption
Figure 6: εg.s.c.(k)\varepsilon^{(k)}_{\mathrm{g.s.c.}} and εx(k)\varepsilon^{(k)}_{\mathrm{x}} (k=1,2k=1,2) in the HF+BCS and HFB solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr.

In Fig. 6, the values of εg.s.c.(k)\varepsilon^{(k)}_{\mathrm{g.s.c.}} and εx(k)\varepsilon^{(k)}_{\mathrm{x}} for k=1,2k=1,2 are also shown for the HFB solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The results are similar to the HF+BCS cases. While the εg.s.c.(1)\varepsilon^{(1)}_{\mathrm{g.s.c.}} and εx(1)\varepsilon^{(1)}_{\mathrm{x}} values are less than 10% for 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr, the s4s_{4} terms are sizable for the 1234{}^{34}_{12}Mg nucleus. This consequence is qualitatively similar also to the HF case in Ref. [23].

III.4 Contribution of constituent terms of effective Hamiltonian

In Ref. [23], we analyzed the composition of the pure rotational energy of the axial-HF solutions. In this subsection, we present influences of the pairing on the composition of the pure rotational energy. By taking 𝒮^\hat{\mathcal{S}} to be constituent terms of H^\hat{H}^{\prime}, the 𝒮x(J+)\mathcal{S}_{\mathrm{x}}(J^{+}) values of Eq. (23) yield their contributions to the rotational energy. In the following, 𝒮^\hat{\mathcal{S}} is an element of the following set,

𝒮^{H^,K^,V^(C),V^(LS),V^(TN),V^(Cρ),V^(OPEP),H^pair}.\hat{\mathcal{S}}\in\{\hat{H}^{\prime},\hat{K},\hat{V}^{\mathrm{(C)}},\hat{V}^{\mathrm{(LS)}},\hat{V}^{\mathrm{(TN)}},\hat{V}^{\mathrm{(C\rho)}},\hat{V}^{\mathrm{(OPEP)}},\hat{H}_{\mathrm{pair}}\}. (28)

Each element has been defined in Sec. II. Since J|N^p|JZ\expectationvalue*{\hat{N}_{p}}{J}\neq Z and J|N^n|JN\expectationvalue*{\hat{N}_{n}}{J}\neq N, the chemical-potential terms in Eq. (15) have contributions to the rotational energy, which should be attributed to other terms of the Hamiltonian if the particle-number projection is simultaneously implemented. However, they are insignificant, staying within 0.517%-0.5-17\,\% for the MF states under investigation.

Refer to caption
Figure 7: 𝒮x(4+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(4^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) for 𝒮^=K^\hat{\mathcal{S}}=\hat{K} (red circles), V^(C)\hat{V}^{\mathrm{(C)}} (blue squares), V^(LS)\hat{V}^{\mathrm{(LS)}} (yellow stars), V^(TN)\hat{V}^{\mathrm{(TN)}} (green triangles), V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} (pink diamonds) and V^(OPEP)\hat{V}^{\mathrm{(OPEP)}} (sky-blue pluses) in the HF+BCS results of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. 𝒮x(4+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(4^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) for H^pair\hat{H}_{\mathrm{pair}} are represented by the orange crosses. The rigid-rotor value 10/310/3 is displayed by the horizontal lines.

In Fig. 7, the gg dependence of 𝒮x(4+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(4^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}), the ratios given by the constituent terms of the effective Hamiltonian, is shown for the 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr nuclei. Contributions of the pairing tensors are excluded except for H^pair\hat{H}_{\mathrm{pair}}. The ratios are close to 10/310/3 for 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr, almost independent of gg and 𝒮^\hat{\mathcal{S}}. In 1234{}^{34}_{12}Mg, the ratios 𝒮x(4+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(4^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) for 𝒮^=K^\hat{\mathcal{S}}=\hat{K}, V^(TN)\hat{V}^{\mathrm{(TN)}} and H^pair\hat{H}_{\mathrm{pair}} are also almost independent of gg, while those for V^(C)\hat{V}^{\mathrm{(C)}}, V^(LS)\hat{V}^{\mathrm{(LS)}}, V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} and V^(OPEP)\hat{V}^{\mathrm{(OPEP)}} become deviating from 10/310/3 as gg increases. The irregular behavior for V^(C)\hat{V}^{\mathrm{(C)}} and V^(OPEP)\hat{V}^{\mathrm{(OPEP)}} observed at g1.1g\approx 1.1 in 1234{}^{34}_{12}Mg happens because 𝒮x(2+)0\mathcal{S}_{\mathrm{x}}(2^{+})\approx 0 in this region, with sign inversion (see Fig. 8).

Refer to caption
Figure 8: The ratios 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}). See Fig. 7 for conventions.

As long as any 𝒮x(J+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(J^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) is close to J(J+1)/6J(J+1)/6, 𝒮x(J+)\mathcal{S}_{\mathrm{x}}(J^{+}) is well described by [𝒮^]\mathcal{I}[\hat{\mathcal{S}}] in Eq. (9a), and it is sufficient to inspect 𝒮x(2+)\mathcal{S}_{\mathrm{x}}(2^{+}) in analyzing the rotational energy. The contributions of the constituent terms of the effective Hamiltonian to the total rotational energy are represented by 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}). In Fig. 8, the gg dependence of 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}) is shown for 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. As gg increases from the critical points, the 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}) values for 𝒮^=K^\hat{\mathcal{S}}=\hat{K}, V^(C)\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} vary; decrease for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}}, while increase for V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}. Even their signs are inverted near g=1.1g=1.1. As gg increases, the contribution of gH^pairg\hat{H}_{\mathrm{pair}} to the rotational energy is enhanced, which is dominated by the central force. The contributions of V^Coul\hat{V}_{\mathrm{Coul}}, H^c.m.\hat{H}_{\mathrm{c.m.}} and V^(OPEP)\hat{V}^{\mathrm{(OPEP)}} to Ex(2+)E_{\mathrm{x}}(2^{+}) are ±10%\pm 10\% at most.

Refer to caption
Figure 9: 𝒮x(J+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(J^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) for the HFB solutions of the   1224,34,40{}^{24,34,40}_{~{}~{}~{}~{}~{}\,\,12}Mg and 4080,100,104{}^{80,100,104}_{~{}~{}~{}~{}~{}~{}~{}~{}40}Zr nuclei. See Fig. 7 for conventions. The experimental values of Ex(J+)/Ex(2+)E_{\mathrm{x}}(J^{+})/E_{\mathrm{x}}(2^{+}) are taken from Refs. [2, 50, 56].

Figure 9 shows the ratios 𝒮x(J+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(J^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) for the HFB solutions of the   1224,34,40{}^{24,34,40}_{~{}~{}~{}~{}~{}\,\,12}Mg and 4080,100,104{}^{80,100,104}_{~{}~{}~{}~{}~{}~{}~{}~{}40}Zr nuclei, all of which have prolate shapes. The ratios Ex(4+)/Ex(2+)E_{\mathrm{x}}(4^{+})/E_{\mathrm{x}}(2^{+}) obtained in the present work are close to those of the experiments and 10/310/3 except at 1240{}^{40}_{12}Mg. For the deformed 40Zr nuclei, the ratios of the constituent terms 𝒮x(J+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(J^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) are also close to J(J+1)/6J(J+1)/6, although less close than in the HF case. On the other hand, some ratios 𝒮x(J+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(J^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) do not obey the J(J+1)J(J+1) rule for the 12Mg nuclei. We find several cases in which even J|𝒮^|J\expectationvalue*{\hat{\mathcal{S}}}{J} is not monotonic for JJ, e.g., with 𝒮x(4+)/𝒮x(2+)<1\mathcal{S}_{\mathrm{x}}(4^{+})/\mathcal{S}_{\mathrm{x}}(2^{+})<1.

Refer to caption
Figure 10: 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}) for the HFB solutions (lower panel) of the   1224,34,40{}^{24,34,40}_{~{}~{}~{}~{}~{}\,\,12}Mg and 4080,100,104{}^{80,100,104}_{~{}~{}~{}~{}~{}~{}~{}~{}40}Zr nuclei, in comparison with those for the HF solutions (upper panel) [23].

In Fig. 10, 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}) are shown for the HFB solutions of the   1224,34,40{}^{24,34,40}_{~{}~{}~{}~{}~{}\,\,12}Mg and 4080,100,104{}^{80,100,104}_{~{}~{}~{}~{}~{}~{}~{}~{}40}Zr nuclei. The results significantly depend on the nuclei. The contribution of the kinetic energy 𝒮^=K^\hat{\mathcal{S}}=\hat{K} is large. It is remarked that those of the interactions V^(C)\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} are scattered, in sharp contrast to the results for the deformed HF solutions in Fig. 3 of Ref. [23]. For the HF solutions, the composition of the pure rotational energy for the well-deformed heavy nuclei is insensitive to nuclides and deformation. However, as elucidated in Fig. 8, 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}) is sensitive to the pairing. Since the degree of the pair correlations depends on nuclei, the components of the pure rotational energy of nuclei also do, even for the well-deformed heavy nuclei.

III.5 Angle dependence of integrands of Eq. (1)

In Ref. [23], we showed that the overlap functions Φ0|eiJ^yβ|Φ0\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}} and 𝒮01(β)\mathcal{S}^{01}(\beta) in Eq. (3) are relevant to the J(J+1)J(J+1) rule and the composition of the rotational energy. The dependence of the overlap functions on the angle β\beta will be instructive in the HF+BCS and HFB cases, as well.

Refer to caption
Figure 11: The overlap functions Φ0|eiJ^yβ|Φ0g\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}_{g} in the HF+BCS and HFB solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The individual lines correspond to the gg values shown in the inset.

Figure 11 shows the gg dependence of the overlap functions Φ0|eiJ^yβ|Φ0g\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}_{g} for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. For 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr, Φ0|eiJ^yβ|Φ0g\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}_{g} have sharp peaks near β=0\beta=0, and become slightly broader as gg increases. In contrast, Φ0|eiJ^yβ|Φ0g\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}_{g} have broad peaks near β=0\beta=0 and getting much broader for increasing gg for the 1234{}^{34}_{12}Mg nucleus.

Figure 11 also depicts Φ0|eiJ^yβ|Φ0\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}} for the HFB solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The overlap functions Φ0|eiJ^yβ|Φ0\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}} for 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr have sharper peaks near β=0\beta=0 than those for 1234{}^{34}_{12}Mg. The broad peak near β=0\beta=0 makes the ratios 𝒮x(J+)/𝒮x(2+)\mathcal{S}_{\mathrm{x}}(J^{+})/\mathcal{S}_{\mathrm{x}}(2^{+}) deviate from J(J+1)/6J(J+1)/6 in Fig. 9, as analogous arguments given in Ref. [23]. The curvature of Φ0|eiJ^yβ|Φ0\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}} at β=0\beta=0 is equal to the variance (σ[J^y])2(\sigma[\hat{J}_{y}])^{2} apart from the sign (see Eq. (6)),

d 2dβ 2Φ0|eiJ^yβ|Φ0|β=0=(σ[J^y])2.-\left.\frac{d^{\,2}}{d\beta^{\,2}}\,\expectationvalue*{e^{-i\hat{J}_{y}\beta}}{\Phi_{0}}\,\right|_{\beta=0}=(\sigma[\hat{J}_{y}])^{2}. (29)

In Fig. 12, we show (σ[J^y])2(\sigma[\hat{J}_{y}])^{2} for 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. Regardless of nuclei, the (σ[J^y])2(\sigma[\hat{J}_{y}])^{2} values decrease as gg increases, and (σ[J^y])2(\sigma[\hat{J}_{y}])^{2} of the HFB solutions are smaller than that of the HF solutions. The 40100{}^{100}_{~{}40}Zr nucleus exemplifies that a20a_{20} does not well correlate to σ[J^y]\sigma[\hat{J}_{y}] straightforwardly; (σ[J^y])2(\sigma[\hat{J}_{y}])^{2} substantially decreases as gg grows, while a20a_{20} shown in Fig. 1 is insensitive to gg.

Refer to caption
Figure 12: (σ[J^y])2(\sigma[\hat{J}_{y}])^{2} in the HF+BCS and HFB solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr.
Refer to caption
Figure 13: Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} in the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The individual lines correspond to the gg values shown in the inset.

We next investigate the inversion near g=1.1g=1.1 for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} observed in Fig. 8, at the level of the overlap functions. By using Eq. (3), we define Δ𝒮01(β)\varDelta\mathcal{S}^{01}(\beta) as

Δ𝒮01(β):=𝒮01(β)𝒮01(β=0).\varDelta\mathcal{S}^{01}(\beta):=\mathcal{S}^{01}(\beta)-\mathcal{S}^{01}(\beta=0). (30)

The curvature of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) is related to the correlation function C[𝒮^,J^y 2]C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}],

d 2dβ 2Δ𝒮01(β)|β=0=C[𝒮^,J^y 2].\left.-\,\frac{d^{\,2}}{d\beta^{\,2}}\,\varDelta\mathcal{S}^{01}(\beta)\right|_{\beta=0}=C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}]. (31)

Although this relation is not exact for v^ij(Cρ)[ρ¯(𝐫i;β)]\hat{v}^{\mathrm{(C\rho)}}_{ij}[\bar{\rho}(\mathbf{r}_{i};\beta)] because of the β\beta-dependence of ρ¯\bar{\rho}, Eq. (31) holds approximately. Figure 13 shows the gg dependence of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) in Eq. (30) for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} in 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. As gg increases, the behavior of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) drastically changes. The signs of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) near β=0\beta=0, which is related to C[𝒮^,J^y 2]C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}] via Eq. (31), changes from positive (negative) to negative (positive) for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} (V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}). The results in Fig. 13 correspond to those in Fig. 8 via the MoI of Eq. (9). As a function of β\beta, Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) for 1234{}^{34}_{12}Mg changes more slowly than those for 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. This almost flat structure of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) for 1234{}^{34}_{12}Mg gives rise to the deviation from the J(J+1)J(J+1) rule and the irregular JJ ordering in Fig. 9. The results in Fig. 13 correspond well to those in Figs. 7 and 8.

Refer to caption
Figure 14: Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) for the constituent terms of the Hamiltonian for the HFB solutions of 1234,40{}^{34,40}_{~{}~{}~{}12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr.

In Fig. 14, we show Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) for the HFB solutions of 1234,40{}^{34,40}_{~{}~{}~{}12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. These results are relevant to those in Fig. 10. The |C[𝒮^,J^y 2]||C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}]| values (see Eq. (31)) significantly increase for 𝒮^=K^\hat{\mathcal{S}}=\hat{K}, V^(C)\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} as the mass number increases. The signs of C[𝒮^,J^y 2]C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}] for H^pair\hat{H}_{\mathrm{pair}} are positive without exceptions. The signs of C[𝒮^,J^y 2]C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}] for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} are negative, and those for V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} are positive for 1234{}^{34}_{12}Mg and 4080{}^{80}_{40}Zr, which are opposite to the results for the HF solutions in Fig. 13 of Ref. [23]. The pair correlations could change the signs of C[𝒮^,J^y 2]C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}] for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}, leading to the results in Fig. 10. Even though the HFB solution of 40100{}^{100}_{~{}40}Zr has the pair correlations, C[𝒮^,J^y 2]C[\hat{\mathcal{S}},\hat{J}_{y}^{\,2}] is positive (negative) for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} (V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}). The flat structure of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta) for 𝒮^=V^(LS)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(LS)}} in 1234{}^{34}_{12}Mg and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}} in 1240{}^{40}_{12}Mg corresponds to the deviation from J(J+1)/6J(J+1)/6 and the irregular JJ ordering in Fig. 9. The results in Fig. 14 well account for those in Figs. 9 and 10.

III.6 Degree of proximity for nucleons associated with nucleonic interaction

We have found that the pairing greatly influences the contribution of V^(C)\hat{V}^{(\mathrm{C})} and V^(Cρ)\hat{V}^{(\mathrm{C}\rho)} to the rotational energy. In the preceding subsection, their contributions have been analyzed in terms of Δ𝒮01(β)-\varDelta\mathcal{S}^{01}(\beta). To investigate what governs the contributions of V^(C)\hat{V}^{(\mathrm{C})} and V^(Cρ)\hat{V}^{(\mathrm{C}\rho)} further, we calculate the DoP D^\expectationvalue*{\hat{D}} by using the AMP, which has been defined in Sec. II.5.

Refer to caption
Figure 15: ΔD\varDelta D for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The black pentagons in the lower panels are the total values of ΔD\varDelta D. The separated values of ΔD\varDelta D into the individual isospin components are also shown; the pppp (red circles), nnnn (blue squares), and pnpn correlations (yellow stars) of the SE channel in D^dns\expectationvalue*{\hat{D}_{\mathrm{dns}}}, where pp (nn) stands for protons (neutrons). The arithmetic average of the spin components is plotted for the ΔD\varDelta D values of the TE channel (green triangles). The pppp and nnnn correlations of the pairing channel in D^pair\expectationvalue*{\hat{D}_{\mathrm{pair}}} in Eq. (33) are represented as pink diamonds and sky-blue pluses, respectively.

Denoting the increment of D^\expectationvalue*{\hat{D}} by ΔD\varDelta D as

ΔD:=Φ0|D^|Φ0gΦ0|D^|Φ0g=0,\varDelta D:=\expectationvalue*{\hat{D}}{\Phi_{0}}_{g}-\expectationvalue*{\hat{D}}{\Phi_{0}}_{g=0}, (32)

the gg dependence of ΔD\varDelta D is shown for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr in Fig. 15. The Φ0|D^|Φ0g=0\expectationvalue*{\hat{D}}{\Phi_{0}}_{g=0} values for 1234{}^{34}_{12}Mg, 4080{}^{80}_{40}Zr and 40100{}^{100}_{~{}40}Zr are 1.221.22, 3.433.43 and 4.284.28, respectively. As expected, ΔD\varDelta D increases for increasing gg. Analogously to Eq. (16), we separate the DoP D^\expectationvalue*{\hat{D}} as

D^=D^dns+D^pair.\expectationvalue*{\hat{D}}=\expectationvalue*{\hat{D}_{\mathrm{dns}}}+\expectationvalue*{\hat{D}_{\mathrm{pair}}}. (33)

Owing to the locality of the operator D^\hat{D}, D^dns\expectationvalue*{\hat{D}_{\mathrm{dns}}} consists of the SE (viz. T=1T=1) and the TE (viz. T=0T=0) channels, while D^pair\expectationvalue*{\hat{D}_{\mathrm{pair}}} has only the SE channel. The ΔD\varDelta D values for the individual isospin components, pppp, nnnn, pnpn with T=1T=1 and 0, are also shown. For the TE channel, we take the arithmetic average of the three components of the spin-triplet. The contributions of Φ0|D^pair|Φ0g\expectationvalue*{\hat{D}_{\mathrm{pair}}}{\Phi_{0}}_{g} to ΔD\varDelta D tend to be larger than those of Φ0|D^dns|Φ0g\expectationvalue*{\hat{D}_{\mathrm{dns}}}{\Phi_{0}}_{g}.

Refer to caption
Figure 16: Dx(2+)D_{\mathrm{x}}(2^{+}) in the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. See Fig. 15 for conventions. The orange crosses are the total values of Dx(2+)D_{\mathrm{x}}(2^{+}) for the HFB solutions.

We have calculated J|D^|J\expectationvalue*{\hat{D}}{J}, the DoP at angular-momentum eigenstates, for the HF+BCS and HFB solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr. The gg dependence of J|D^|JgJ|D^|Jg=0\expectationvalue*{\hat{D}}{J}_{g}-\expectationvalue*{\hat{D}}{J}_{g=0} behaves similarly to that of ΔD\varDelta D in Fig. 15. We then define Dx(J+)D_{\mathrm{x}}(J^{+}) as

Dx(J+):=J|D^|J0|D^|0,D_{\mathrm{x}}(J^{+}):=\expectationvalue*{\hat{D}}{J}-\expectationvalue*{\hat{D}}{0}, (34)

which corresponds with Eq. (23) for 𝒮^=D^\hat{\mathcal{S}}=\hat{D}. As the intrinsic state |Φ0\ket{\Phi_{0}} is identical among |J\ket{J} (J=0,2,4,J=0,2,4,\cdots), Dx(J+)D_{\mathrm{x}}(J^{+}) are small, but do not vanish. The gg dependence of Dx(2+)D_{\mathrm{x}}(2^{+}) for the HF+BCS solutions of 1234{}^{34}_{12}Mg and 4080,100{}^{80,100}_{~{}~{}~{}~{}40}Zr are exhibited in Fig. 16. The values of Dx(2+)D_{\mathrm{x}}(2^{+}) for the HFB solutions (with g=1g=1) are also shown for reference. In 4080{}^{80}_{40}Zr, Dx(2+)D_{\mathrm{x}}(2^{+}) for the HFB solution is close to those for the HF+BCS solutions at g1.2g\approx 1.2. No gg values match the HFB results in 1234{}^{34}_{12}Mg and 40100{}^{100}_{~{}40}Zr, suggesting sizable influence of the pairing on the HF configurations as discussed in Subsec. III.1.

For the HF (viz. g=0g=0) solutions, the Dx(2+)D_{\mathrm{x}}(2^{+}) values are negative. This means that the DoP J|D^|J\expectationvalue*{\hat{D}}{J} decreases and the nucleons slightly spread as JJ goes up. As gg increases from the critical points, Dx(2+)D_{\mathrm{x}}(2^{+}) increases. The signs of Dx(2+)D_{\mathrm{x}}(2^{+}) change between g=1.1g=1.1 and 1.21.2. Thus, although the nucleons spread with increasing JJ at the HF level, the pair correlations reduce and finally invert the effect. The Dx(2+)D_{\mathrm{x}}(2^{+}) values are decomposed into the individual isospin components. As gg increases, contributions of D^dns\expectationvalue*{\hat{D}_{\mathrm{dns}}} to Dx(2+)D_{\mathrm{x}}(2^{+}) also increase irrespective of the four isospin components, while those of D^pair\expectationvalue*{\hat{D}_{\mathrm{pair}}} decrease from zero. We should note that there will be counter effects in actual nuclei that are not included in the pure rotational energy. The intrinsic state may gradually stretch with increasing JJ, as handled in the cranking model [6, 10, 3] and the VAP schemes [3].

These results correlate well with 𝒮x(2+)/Ex(2+)\mathcal{S}_{\mathrm{x}}(2^{+})/E_{\mathrm{x}}(2^{+}) in Fig. 8, particularly for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}. This correlation seems to reflect the short-range nature of the nucleonic interaction. It is interpreted that the contributions of the individual components of the interaction to the rotational energy are governed by the spatial proximity among constituent nucleons measured by the DoP. Although the proximity changes with JJ only by a small fraction, its effects on the pure rotational energy are significant. This argument is supported by comparing the fraction with those for V^(C)\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}, i.e., the percentage of the contribution to the rotational energy in the whole V^(C)\hat{V}^{\mathrm{(C)}} and V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}. If we estimate it through 𝒮x(2+)/0|𝒮^|0\mathcal{S}_{\mathrm{x}}(2^{+})/\expectationvalue*{\hat{\mathcal{S}}}{0} and take 4080{}^{80}_{40}Zr as an example, the value is 3.1×1053.1\times 10^{-5} (1.8×1051.8\times 10^{-5}) for 𝒮^=V^(C)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C)}} (𝒮^=V^(Cρ)\hat{\mathcal{S}}=\hat{V}^{\mathrm{(C\rho)}}). The corresponding value for the DoP (𝒮^=D^\hat{\mathcal{S}}=\hat{D}) is 1.5×1051.5\times 10^{-5}. The ratios with the same order of magnitude are compatible with the interpretation that the DoP plays an essential role in the contribution of the nucleonic interaction to the rotational energy. The consequence also applies to the HF results reported in Ref. [23], and to the effects of the pairing. When the DoP diminishes (i.e., the constituent nucleons tend to spread) for increasing JJ, the attractive (repulsive) forces like V^(C)\hat{V}^{\mathrm{(C)}} (V^(Cρ)\hat{V}^{\mathrm{(C\rho)}}) increase (decrease) the rotational energy, and vice versa, mediated by the overlap functions.

IV Conclusion

The pure rotational energy of nuclei, i.e., the rotational energy for a fixed intrinsic state, has extensively been analyzed by applying the AMP to the self-consistent axial-MF solutions with the semi-realistic effective Hamiltonian M3Y-P6. The contributions of the constituent terms of the Hamiltonian to the total rotational energies have been inspected, focusing on effects of the pairing.

Due to the pair correlations, the compositions of the pure rotational energy drastically change, sometimes inverting their signs, and depend strongly on nuclides even for the well-deformed nuclei. When the pairing becomes stronger, the contributions of the kinetic energies increase. Those of the attractive (repulsive) forces decrease (increase), and their signs could invert.

The degree of proximity (DoP) between nucleons slightly depends on the angular momentum JJ, and could account for the effects of nucleonic interactions on the rotational energy. The nucleons slightly spread as JJ increases at the HF level, while the pair correlations can reduce or invert the effect. It is concluded that the rotational energy of nuclei is carried by the kinetic energy in its major part, but is contributed by the nucleonic interaction as well, sensitively reflecting the DoP, i.e., the degree how frequently two constituent nucleons come close. This spatial correlation accounts for the stable composition of the pure rotational energy in the HF states of well-deformed medium-to-heavy nuclei found in Ref. [23], the deviation in light nuclei and weakly-deformed states, and the disarrangement due to the pairing. For actual nuclei, the intrinsic state varies as JJ increases, and this effect influences the DoP and the rotational energy, which is ignored in the present study and left for future works. Still, even when the intrinsic state depends on JJ, the role of the DoP discovered here will give us an insight into the effects of the interaction on the rotational energy of nuclei.

Acknowledgments

The authors are grateful to H. Kurasawa, K. Yoshida and K. Washiyama for discussions. A part of this work was performed under the long-term international workshop on “Mean-field and Cluster Dynamics in Nuclear Systems (MCD2022)”, sponsored by the Yukawa International Program for Quark-Hadron Sciences and held at Yukawa Institute for Theoretical Physics (YITP), Kyoto University, Japan. This research was partly supported by the research assistant program at Chiba University. Numerical calculations were carried out on Yukawa-21 at YITP, Oakforest PACS at Center for Computational Sciences, University of Tsukuba under the Multidisciplinary Cooperative Research Program, and HITACHI SR24000 at the Institute of Management and Information Technologies, Chiba University.

References

  • [1] A. Bohr and B. R. Mottelson, Nuclear Structure, Vols. I, II (World Scientific, Singapore, 1998).
  • [2] National Nuclear Data Center, https://www.nndc.bnl.gov/.
  • [3] P. Ring and P. Schuck, The Nuclear Many-Body Problem (Springer-Verlag, Berlin, 1980).
  • [4] R. E. Peierls and J. Yoccoz, Proc. Phys. Soc. London, Sect. A 70, 381 (1957).
  • [5] J. Yoccoz, Proc. Phys. Soc. London, Sect. A 70, 388 (1957).
  • [6] R. E. Peierls and D. J. Thouless, Nucl. Phys. 38, 154 (1962).
  • [7] B. J. Verhaar, Nucl. Phys. 45, 129 (1963).
  • [8] B. J. Verhaar, Nucl. Phys. 54, 641 (1964).
  • [9] N. Onishi and S. Yoshida, Nucl. Phys. 80, 367 (1966).
  • [10] A. Kamlah, Z. Phys. 216, 52 (1968).
  • [11] R. Balian and E. Brezin, Nuov Cim. B 64, 37 (1969).
  • [12] H. J. Mang, Phys. Pep. 18, No. 6, 325 (1975).
  • [13] K. Hara and S. Iwasaki, Nucl. Phys. A332, 61 (1979).
  • [14] K. Hara, S. Iwasaki and K. Tanabe, Nucl. Phys. A332, 69 (1979).
  • [15] K. Hara and S. Iwasaki, Nucl. Phys. A348, 200 (1980).
  • [16] P. R. Rodríguez-Guzmán, J. L. Egido and L. M. Robledo, Nucl. Phys. A 709, 201 (2002).
  • [17] M. Bender, P. H. Heenen and P. G. Reinhard, Rev. Mod. Phys. 75, 121 (2003).
  • [18] K. W. Schmid, Prog. Part. Nucl. Phys. 52, 565 (2004).
  • [19] M. Bender and P. H. Heenen, Phys. Rev. C 78, 024309 (2008).
  • [20] L. M. Robledo, T. R. Rodríguez and P. R. Rodríguez-Guzmán, J. Phys. G: Nucl. Part. Phys. 46, 013001 (2019).
  • [21] B. Bally and M. Bender, Phys. Rev. C 103, 024315 (2021).
  • [22] J. A. Sheikh, J. Dobaczewski, P. Ring, L. M. Robledo and C. Yannouleas, J. Phys. G: Nucl. Part. Phys. 48, 123001 (2021).
  • [23] K. Abe and H. Nakada, Phys. Rev. C 106, 054317 (2022).
  • [24] D. J. Thouless and J. G. Valatin, Nucl. Phys. 31, 211 (1962).
  • [25] S. T. Belyaev, Mat. Fys. Medd. Dan. Vid. Selsk. 31, No. 11 (1959).
  • [26] D. R. Inglis, Phys. Rev. 96, 1059 (1954); 97, 701 (1955).
  • [27] B. R. Mottelson and J. G. Valatin, Phys. Rev. Lett. 5, 511 (1960).
  • [28] H. Nakada, Phys. Rev. C 68, 014316 (2003).
  • [29] H. Nakada, Phys. Rev. C 87, 014336 (2013).
  • [30] H. Nakada, Int. J. Mod. Phys. E 29, 1930008 (2020).
  • [31] G. Bertsch, J. Borysowicz, H. McManus and W. G. Love, Nucl. Phys. A 284, 399 (1977).
  • [32] W. N. Cottingham, M. Lacombe, B. Loiseau, J. M. Richard and R. Vinh Mau, Phys. Rev. D 8, 800 (1973).
  • [33] M. Lacombe, B. Loiseau, J. M. Richard, R. Vinh Mau, J. Côté, P. Pirés and R. de Tourreil, Phys. Rev. C 21, 861 (1980).
  • [34] N. Anantaraman, H. Toki and G. F. Bertsch, Nucl. Phys. A 398, 269 (1983).
  • [35] D. Davesne, A. Pastore and J. Navarro, Prog. Part. Nucl. Phys. 120, 103870 (2021).
  • [36] H. Nakada and K. Sugiura, Prog. Theor. Exp. Phys. 2014, 033D02 (2014); 2016, 099201 (2016).
  • [37] Y. Suzuki, H. Nakada and S. Miyahara, Phys. Rev. C 94, 024343 (2016).
  • [38] S. Miyahara and H. Nakada, Phys. Rev. C 98, 064318 (2018).
  • [39] M. Matsuo, K. Mizuyama and Y. Serizawa, Phys. Rev. C 71, 064326 (2005).
  • [40] D. A. Varshalovich, A. N. Moskalev and V. K. Khersonskii, Quantum Theory of Angular Momentum (World Scientific, Singapore, 1988).
  • [41] J. J. Sakurai, Modern Quantum Mechanics Revised Edition (Addison Wesley Longman, 1994).
  • [42] R. Kubo, J. Phys. Soc. Jpn. 17, No. 7, 1100 (1962).
  • [43] H. Nakada and M. Sato, Nucl. Phys. A 699, 511 (2002); 714, 696 (2003).
  • [44] H. Nakada, Nucl. Phys. A 808, 47 (2008).
  • [45] H. Nakada, Nucl. Phys. A 764, 117 (2006); 801, 169 (2008).
  • [46] K. Neergård and E. Wüst, Nucl. Phys. A 402, 311 (1983).
  • [47] L. M. Robledo, Phys. Rev. C 79, 021302(R) (2009).
  • [48] J. W. Negele and H. Orland, Quantum Many-Particle Systems (Addison-Wesley, 1995).
  • [49] G. F. Giuliani and G. Vignale, Quantum Theory of the Electron Liquid (Cambridge Univ. Press, Cambridge, 2005).
  • [50] P. Doornenbal, H. Scheit, S. Takeuchi, N. Aoi, K. Li, M. Matsushita, et al., Phys. Rev. Lett. 111, 212502 (2013).
  • [51] T. Sumikama, K. Yoshinaga, H. Watanabe, S. Nishimura, Y. Miyashita, K. Yamaguchi, et al., Phys. Rev. Lett. 106, 202501 (2011).
  • [52] N. Paul, A. Corsi, A. Obertelli, P. Doornenbal, G. Authelet, H. Baba, et al., Phys. Rev. Lett. 118, 032501 (2017).
  • [53] C. J. Lister, M. Campbell, A. A. Chishti, W. Gelletly, L. Goettig, R. Moscrop, et al., Phys. Rev. Lett. 59, 1270 (1987).
  • [54] T. Baumann, A. M. Amthor, D. Bazin, B. A. Brown, C. M. Folden III\rm{III}, A. Gade, et al., Nature 449, 1022 (2007).
  • [55] H. Nakada and K. Takayama, Phys. Rev. C 98, 011301(R) (2018).
  • [56] H. L. Crawford, P. Fallon, A. O. Macchiavelli, P. Doornenbal, N. Aoi, F. Browne, et al., Phys. Rev. Lett. 122, 052501 (2019).