Model For Independent Particle Motion

Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

Model for independent particle motion

A. V. Afanasjev ∗
arXiv:2203.04460v1 [nucl-th] 9 Mar 2022

Abstract Independent particle model in nuclear physics assumes that the nucleon
in the nucleus moves in the average (mean field) potential generated by all other
nucleons. This chapter gives a short overview of basic features of the independent
particle motion in atomic nuclei and its theoretical realization in the framework of
shell models for spherical, deformed and rotating nuclei as well as in more sophis-
ticated approaches such as microscopic+macroscopic model and density functional
theories. Independent particle motion of nucleons leads to global and single-particle
consequences. The global ones manifest themselves in the shell structure and its
consequences for global structure of nuclear landscape, the existence of superheavy
nuclei and the superdeformation at high spin are briefly reviewed. The latter shows
itself in the single-particle properties such as energies, alignments and densities;
their manifestations are illustrated on specific examples.

Introduction

The basic idea of the independent particle model is that nucleon (proton or neutron)
moves in an average (mean field) nuclear potential and this motion is independent
of motion of other nucleons. This allows to replace the complicated solution of the
problem of A interacting particles (A is total number of nucleons in the nucleus)
with much simpler problem of A noninteracting particles in the mean field potential.
This mean field potential can be either designed in a phenomenological way or cal-
culated fully self-consistently from effective interaction. As discussed in this chap-
ter, many properties of the nuclei have been described with high accuracy within

A. V. Afanasjev
Department of Physics and Astronomy, Mississippi State University, MS 39762, USA, e-mail:
[email protected]
∗ corresponding author

1
2 A. V. Afanasjev

such frameworks and many new phenomena (for example, superheavy nuclei and
superdeformation at high spin) have been predicted theoretically and later observed
experimentally. Independent particle model in its different realizations provides also
the basis [mean field] for the models which take into account residual interactions
such as pairing, vibrations, particle-vibration coupling etc. These interactions are
typically included either by adding respective terms to the version of independent
particle model or by the modification of the formalism.
The basic features of independent particle model, its realization in the shell model
variants, microscopic+macroscopic model and covariant density functional theory
will be discussed in this chapter. In addition, some manifestations of the indepen-
dent particle motion emerging from the shell structure and single-particle properties
will be considered. In no way, this chapter should be considered as ”all-inclusive”:
it only scratches the surface of huge body of experimental and theoretical results.
More comprehensive and detailed reviews on specific phenomena/theoretical ap-
proaches are quoted in respective sections. In addition, some aspects of the inde-
pendent particle motion are discussed in the books [1, 2, 3, 4].

Independent particle model

The solution of the many-body nuclear problem for a nucleus consisting of A nu-
cleons (Z protons and N neutrons) requires the solution of the eigenvalue problem
[1, 2]

HΨα (~r, σ , τ) = Eα Ψα (~r, σ , τ). (1)

The Hamiltonian of the system could be either in non-relativistic (Schrödinger


equation) or relativistic (Dirac equation) forms. It contains the kinetic and potential
energy contributions from each nucleon
A
h̄2 2
H=∑ ∇i + ∑ Vi j , (2)
i=1 2mi i6= j

and includes two-body interaction between i- and j-th nucleons.


The total wave function Ψα (~r, σ , τ) of the nucleus in the state with quantum
numbers α has to be expressed in terms of the ones of the individual nucleons
ψi (~ri , σi , τi ). Here ~r, σ and τ are position, spin and isospin variables, respectively.
For simplicity, in further discussion ~ri will represent all independent variables of
the i-th nucleon. Total wave function of the system of the A particles has to be anti-
symmetric with respect of the exchange of the coordinates of two particles. Thus, a
many-body state is written in the form of the Slater determinant [2]
Model for independent particle motion 3

ψ1 (~r1 ) ψ1 (~r2 ) ... ψ1 (~rA )


1 ψ (~r ) ψ2 (~r2 ) ... ψ2 (~rA )
Ψα (~r1 ,~r2 , ...,~rA ) = √ det 2 1 . (3)
A! ... ... ... ...
ψA (~r1 ) ψA (~r2 ) ... ψA (~rA )

The factor √1A! is due to normalization. The choice of the single-particle wavefunc-
tions defines the type of the many-body state. It is important to remember that the
single-particle spectrum is infinite one. Thus, in practical applications respective
Hilbert space is truncated most frequently based on energy considerations.
In nuclear systems it is possible to recast Eq. (1) in the form
A A
H = ∑ h(~ri ) + ∑ Ṽ (~ri ,~r j ), (4)
i=1 i6= j=1

where Ṽ (~ri ,~r j ) is the residual two-body interaction and h(~ri ) is the single-particle
hamiltonian. This recast essentially means that the part of original two-body inter-
action Vi j is moved to the single-particle hamiltonian h(~ri ). The latter can be chosen
in such a way that the contribution of residual interaction becomes rather small or
even negligible.
Independent particle model corresponds to the situation when second term in Eq.
(4) is neglected. Thus, the nuclear Hamiltonian is a sum of single-particle terms
A
H = ∑ h(~ri ), (5)
i=1

and eigenvalue problem is reduced to

h(~ri )ψk (~ri ) = ek ψk (~ri ), (6)

where ek stand for the single-particle energy.


Independent particle model allows to reduce the nuclear many-body problem
from complicated two-body treatment to much simpler one-body one. It also relies
on the use of the single-particle potentials. Independent particle model corresponds
to the description of a nucleus in terms of noninteracting particles in the orbitals of
these single-particle potentials which itself are produced by all the nucleons. Thus,
nucleons move essentially free in these potentials. In this model, the ground state
is formed by filling all the single-particle states located below the Fermi level. The
excitation of the particle from occupied state below this level to an empty one above
it leads to an excited state. This process is called as a particle-hole excitation.
The justifications for an independent particle model have been discussed in a
number of publications [1, 3]. The mean free path between collisions of the con-
stituent nucleons is large compared to the average distance between them and in
some cases it could be larger than the dimensions of the nucleus. As a consequence,
the interactions between nucleons contribute mostly to the smoothly varying aver-
age potential in which the particles moves independently. One can arrive to the same
4 A. V. Afanasjev

conclusion by taking into account the fact that the nucleons occupy only approxi-
mately 1% of the volume of the nucleus. This estimate follows from the considera-
tion of nucleons as hard spheres of radius c ≈ 0.5 fm. This value of c corresponds
to the half of the closest distance which two nucleons can approach each other due
to infinite repulsion (see discussion in Sec. 2-5b of Ref. [1]). For such value of c,
nuclear matter behaves close to the free gas of nucleons with very rare head-on
collisions of nucleons. The Pauli principle and the weakness of strong nuclear in-
teraction when compared with characteristic kinetic energies of nucleons inside the
nucleus are other contributing factors for the justification of independent particle
model [1, 3].

Spherical shell model

The discussion of previous section clearly illustrates the need for the introduction
of the single-particle potential. This one-body potential can be introduced either in
phenomenological or in a fully self-consistent ways. In the former case, one deals
with phenomenological Nilsson, Woods-Saxon or folded Yukawa potentials. Self-
consistent single-particle potentials are formed as a result of the solution of the
many-body nuclear problem within non-relativistic and relativistic density func-
tional theories (DFTs); note that these potentials are not treated separately from
the rest of the nuclear many-body problem.
To illustrate the major physics aspects behind these potentials and for pedagogi-
cal reasons, the harmonic oscillator (HO) potential and its modifications relevant for
nuclear physics problems will be considered here. The single-particle hamiltonian
of the nucleon with mass m entering into Eq. (5) is given by

h̄2 2
h(~r) = − ∇ +V (~r), (7)
2m i
where the central potential
1
V (~r) = mω 0 r2 (8)
2
is pure HO potential representing one-body (mean) potential (field). This hamilto-
nian in spherical coordinates is given by

h̄2 1 ∂ 2 l 2 (θ , φ )
h(~r) = − r + +V (r), (9)
2m r ∂ r2 2mr2
and the wave function ψ representing the solution of the eigenvalue problem is sep-
arable in angular (θ , φ ) and radial (r) coordinates: ψ = R(r)Ylm (θ , φ ). Here, R(r)
and Ylm (θ , φ ) are radial and angular [given by spherical harmonics] wave functions,
respectively. Note that the latter is the eigenfunction of the angular momentum op-
erator l 2 :
Model for independent particle motion 5

2h
N=7 1k 2h11/2
µ′= 0.021 1j3/2
κ = 0.06 1k17/2
µ′ = 0.024 4s 184
4s
3d 3d3/2 1/2
2g
1j 3d5/2 7/2
2g
N=6 1j15/2
1i11/2
κ = 0.06 2g9/2
µ′= 0.024 1i 126
3p 3p1/2
2f5/2
3p3/2
2f 1i13/2
N=5 1h9/2
κ = 0.06 2f7/2
µ′ = 0.024 1h
82
1h11/2
3s
2d 3s1/2 2d3/2
N=4 1g
2d5/2 7/2
κ = 0.06 1g
µ′ = 0.024 50
1g9/2
2p1/2
2p 1f5/2
N=3 2p3/2
1f
κ = 0.075 28
µ′ = 0.0263 1f7/2
20
µ′ = 0 1d3/2
N = 2 κ = 0.08 2s+1d
2s1/2
_ 2 1d5/2
Hosc -µ′h ω0 [l -N(N+3)/2] _
-2κh ω0 l·s
Fig. 1 The sequential build-up of realistic nucleonic potential. The left column shows the single-
particle states of the pure harmonic oscillator. The employed parameters κ and µ 0 of the MO
potential are displayed: note that they are different for the different N-shells. The modifications
introduced by Eq. (11) are shown in the middle column. Finally, the right column shows the im-
pact of spin-orbit interaction on the energies of single-particle states and their ordering. Particle
numbers corresponding to spherical shell closures are encircled. Black and red colors are used for
positive and negative parity states, respectively. The figure is based on the results presented in Fig.
6.3 of Ref. [4].

l 2Ylm (θ , φ ) = h̄2 l(l + 1)Ylm (θ , φ ). (10)

The solutions of the hamiltonian with only central potential included are shown
in the left column of Fig. 1. The spectrum of single-particle states, characterized
6 A. V. Afanasjev

by principal quantum number N, is equidistant in energy with high degree of the


degeneracy of the single-particle states.
The realistic nuclear potential is located somewhat between pure HO and square
well potentials. To correct for that a centrifugal potential has to be added to V (r); it
is usually parametrized as [4, 5, 6]:

Vcorr = −µ 0 h̄ω 0 l 2 − l 2 N .

(11)

Within the N shell this term leads to a lowering in energy of high-l states relatively
to low-l ones and to a removal of the degeneracy between the l states (see middle
panel of Fig. 1).
In addition, there is a coupling between spin (s = 1/2) and orbital motion of the
single particle which, in general, is given by the following interaction term:

1 ∂VSO (r) ~
VLS = λ l ·~s. (12)
r ∂r
Here VSO indicates the spin-orbit potential which may be different from the central
one and λ is the coupling constant of spin-orbit interaction. In the case of harmonic
oscillator potential this term can be further simplified to VLS = −2κ h̄ω 0~l ·~s. The
presence of spin-orbit potential leads to the lowering and rising in energy of the j =
l + 1/2 and j = l − 1/2 states emerging from the state with a given orbital angular
momentum l (see right column in Fig. 1). Only with this interaction included, it is
possible to reproduce experimentally observed shell closures at particle numbers 2,
8, 20, 28, 50, 82 and 126 [1, 3, 6]. Note that the single-particle state is completely
defined by a set of quantum numbers [Nl j].
The combined potential

VMO = V (r) +Vcorr +VLS (13)

is usually called as modified oscillator (MO) or Nilsson potential. This potential is


defined by three parameters ω0i , κi and µi0 for each kind of nucleons (i = π or ν). ω0i
defines the radius of respective matter distribution, µi0 simulates the surface diffuse-
ness depth, and κi is the strength of spin-orbit interaction. The Coulomb potential is
not directly included into the MO potential but it is effectively accounted by the dif-
ferences of the above mentioned parameters in the proton and neutron subsystems.
Although the MO potential is still extensively used in nuclear structure studies,
more realistic treatment of the single-particle degrees of freedom is achieved by
means of the Woods-Saxon (WS) [7, 8, 9] and folded Yukawa (FY) [10, 11] poten-
tials. This is because they have more realistic shape of the potential (thus eliminating
the need for Vcorr ) and explicitly include the Coulomb interaction. Their structure is
illustrated below for the WS potential

VW S = VW S (r) +VLS +VC , (14)

where central potential is given by


Model for independent particle motion 7

V0
VW S (r) = (15)
1 + exp[(r − R)/a]

with V0 being the potential depth, a the surface thickness and R = r0 A1/3 the nuclear
radius (with typical value of r0 ≈ 1.2 fm). Here VC represents the Coulomb potential.
However, numerical realization of these potentials is more complicated as compared
with the MO potential. The detailed comparison of the WS and MO potentials is
presented in Ref. [9].

Deformed shell model

In deformed nuclei, the shape of the nuclear surface is generally parametrized by


means of a multipole expansion of the radius in terms of the shape parameters [2];
a typical parametrization is
" #
R(θ , φ ) = R0 1 + ∑ αλ µ Yλ µ (θ , φ ) , (16)
λµ

where αλ µ are the deformation parameters and R0 is the radius of the sphere with the
same volume. For axially symmetric nuclear shapes µ = 0 and quadrupole deforma-
tion β2 = α20 is dominant [2, 12] . For simplicity, higher multipolarity deformations
such as octupole β3 , hexadecapole β4 , ... are not taken into account in the present
discussion. Note that in the literature different parametrizations of the deformations
of the single-particle potential exist (see discussion in Refs. [4, 12]).
There are two facts which affect the consideration of single-particle potentials in
deformed nuclei [5]. First, this potential should follow the nuclear density distribu-
tion. Second, in deformed nuclei the oscillator frequencies ω i (i = x, y and z) are
different along different principal axis of nuclear ellipsoid. As a result, the central
potential of Eq. (8) is modified in the following way:
m 2 2
ω x x + ω y2 y 2 + ω z2 z 2 .

V (~r) = (17)
2
Since nuclear matter is highly incompressible, the change of the shape of the nucleus
from spherical to ellipsoidal should not modify the volume of the nucleus. This is
accounted by the volume-conservation condition

ω 03 = ω x ω y ω z . (18)

Assuming axial symmetry around the z axis, i.e. ω x = ω y , the central potential can
be rewritten as
1
V (~r) = mω 02 r 2 − β2 mω 02 r 2Y20 (θ , φ ), (19)
2
8 A. V. Afanasjev

3/2 3d3/2
2f5/2 [6 3d5/2
42 4]

6]
-1 ] -4 13] 60

60
5/2[6 2[

Single-neutron energies ei [MeV]


9/

Single-proton energies ei [MeV]

/2[
114

11
2g7/2
-2 -5

1/2
1/2[6
2f7/2 725]
20]

[
9/2[624]

7]
164
613 ] 11/2[ 3/2[
622]
1/2 7/2[

1/2[
[52 ]
-3 106 1] -6 15

651]
3/2 104 2g9/2 [6 152
[52
1] 9/2
1j15/2 9/2[7
1i13/2 34]
-4 102 -7 5/2
1/ [622]

5]
7/2 2[

[50
[63 63
3]

9/2
1] 148
4] 7/2
1]
-5 [51 -8 138
5/2
[74
[50
92 7/25 96 [75 3]
1/2
/2[ 2]
64
1/2 2] 0]
[53 [40 NL3*
-6 254 0] 1/2 -9 4]
No 7/2[62
1i11/2
0 0.1 0.2 0.3 0 0.1 0.2 0.3
β2 deformation β2 deformation

Fig. 2 Single-particle energies, i.e., the diagonal elements of the single-particle Hamiltonian h in
the canonical basis [2], for the lowest in total energy solution in the nucleus 254 No calculated as
a function of the quadrupole equilibrium deformation β2 for covariant energy density functional
NL3*. Solid and dashed lines are used for positive- and negative-parity states, respectively. Rele-
vant spherical and deformed gaps are indicated. Figure taken from Ref. [14].

where β2 stands for the quadrupole deformation of the potential (the measure of
the deviation from spherical shape). Thus, the hamiltonian of the Nilsson model
becomes

h̄2 2 1
h(~r) = − ∇ + mω 02 r 2 − β2 mω 02 r 2Y20 (θ , φ )
2m i 2
− µ 0 h̄ω 0 l 2 − l 2 N − 2κ h̄ω 0~l ·~s.

(20)

Note that for simplicity only general outline of the Nilsson model is provided here
[5]. Technical details of the solution of the Nilsson potential are discussed in Refs.
[4, 5, 9]. Note also that there are generalizations of the Woods-Saxon and folded
Yukawa potentials to deformed shapes [7, 8, 10, 11, 12].
As a result, the focus here is on the consequences of the breaking of spherical
symmetry on the single-particle states. They are usually illustrated by means of the
Nilsson diagrams which show the evolution of the energies of deformed single-
particle states as a function of deformation (see, for example, Figs. 3 and 5 in Ref.
[5], Figs. 8.3 and 8.5 in Ref. [4] and Fig. 3 in Ref. [13] for such diagrams obtained
with the Nilsson potential). The general structure (and frequently fine details) of the
Nilsson diagrams obtained in phenomenological single-particle potentials and in
self-consistent models are very similar. This clearly indicates that the former have
deep microscopic roots. Thus, the results obtained in covariant density functional
theory presented in Fig. 2 are used here to illustrate the impact of deformation on
the single-particle states.
Several major features emerge on transition from spherical to deformed shapes.
First, it removes the 2 j + 1 degeneracy of the spherical subshells and deformed
single-particle states are only two-fold degenerate. Second, the deformed single-
Model for independent particle motion 9

particle states are defined by approximate Nilsson quantum numbers Ω [NnzΛ ] (note
they are frequently shown in inverted order of [NnzΛ ]Ω ). Here, Ω (Λ ) stands for
the projection of the total (orbital) single-particle angular momentum on the axis of
symmetry, N is the principal quantum number and nz is the number of nodes of the
wavefunction along the symmetry axis. Note that the parity π of the state is defined
by (−1)N . For prolate (β2 > 0) shapes, the deformed orbitals emerging from a given
spherical j-subshell are split in such a way that the state with Ω = 1/2 is always
lowest in energy while that with Ω = j is the highest one. The states with interme-
diate Ω 0 values are arranged in such a way that the energies E of two neighboring
deformed states satisfy the condition E(Ω 0 + 1) > E(Ω 0 ) (see Fig. 2).The order of
the states is inverted for the oblate (β2 < 0) shapes [4, 5] . Third, as a consequence
of these modifications of the single-particle states with deformation, spherical shell
gaps disappear and new deformed gaps, which are comparable in size with minor
shell gaps at spherical shape, appear. These gaps are encircled in Fig. 2).

Cranked shell model

Rotation is the phenomenon which appears in all branches of physics, from galaxies
down to atomic nuclei. In the latter, it is a collective phenomenon in which many
nucleons define a nuclear deformation and contribute to rotational motion. Note
that contrary to classical mechanics, a collective rotation along the symmetry axis
of nuclear density distribution is forbidden in quantum mechanics.
The simplest way to describe the properties of rotating nuclei in a microscopic
way is to use the cranking-model approximation suggested by Inglis [15, 16]. Over
the years this model has been successfully applied to the description of different
phenomena in rotating nuclei [4, 13, 17, 18, 19]. The consideration here is restricted
to one-dimensional cranking in which the nuclear field is rotated externally with a
constant angular velocity ω around a principal axis usually defined as the x-axis.
This is done in order to outline the basic features of this model.
The basic idea of the cranking model is that a nucleus with angular momentum
I 6= 0 can be described in terms of an intrinsic state Ψ ω at rest in a rotating frame. In
the one-dimensional cranking approximation for collective rotation, the total crank-
ing Hamiltonian (or Routhian) for a system of independent particles is given by

H ω = H − ωIx = ∑ hωi , (21)


i occ

where H is the total Hamiltonian in the laboratory system, Ix is the x-component of


the total angular momentum and hω is the single-particle Hamiltonian in the rotating
system

hω = h − ω jx (22)
10 A. V. Afanasjev

[642]5/2
-7 [521]1/2

Neutron Routhians [MeV]


[523]5/2
[523]5/2
[640]1/2 [523]5/2
[624]9/2
[633]7/2 [624]9/2
[402]5/2
[761]3/2 [402]5/2
-8 [402]5/2 [640]1/2
[514]9/2
[521]3/2
[514]9/2
[514]9/2 [770]1/2
[521]3/2
[521]3/2

-9 [642]3/2

86
[880]1/2
[770]1/2
85 [752]5/2

-10
[651]3/2
[642]5/2
[651]1/2 [411]1/2
[651]3/2
-11 [411]1/2 [411]1/2

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Rotational Frequency [MeV]
Fig. 3 Single-neutron energy levels obtained with the Woods-Saxon potential as a function of
rotational frequency for deformation parameters β2 = 0.61, β4 = 0.11 typical for the yrast su-
perdeformed band in 152 Dy and its closest neighbors. The orbitals are defined by the parity π and
signature quantum number α. Solid, dotted, short-dashed and long-dashed lines are used for the
orbitals with (π = −, α = −1/2) (π = −, α = +1/2), (π = +, α = +1/2) and π = +, α = −1/2),
respectively. The dominant Nilsson components of the wavefunctions of the levels are shown at the
lowest and highest calculated frequencies. Figure is reproduced from Ref. [20].

with h being the single-particle Hamiltonian in the laboratory system and jx the x-
component of the single-particle angular momentum. In Eq. (21), the sum extends
over the occupied proton and neutron orbitals. The term −ωIx in Eq. (21) is analo-
gous to the Coriolis and centrifugal forces in classical mechanics.
Then the total energy Etot in the laboratory system is given as

Etot = ∑ hψiω |h|ψiω i = ∑ eωi + ω ∑ hψiω | jx |ψiω i, (23)


i occ i occ i occ

and the total spin I by

I ≈ Ix = ∑ hψiω | jx |ψiω i, (24)


i occ

where ψiω are the single-particle eigenfunctions in the rotating system and eω i =
hψiω |hω |ψiω i the corresponding eigenvalues (single-particle routhians). The approx-
imation used in Eq. (24) is valid only in the limit of high spin (I  1).
Model for independent particle motion 11

The evolution of the single-particle states in rotating potential is shown in Fig. 3.


There are several important features. First, the time-reversal symmetry is broken in
rotating potential and two-fold degeneracy of deformed single-particle states, which
exist in non-rotating nuclei, is removed. As a consequence, each single-particle or-
bital has additionally to be characterized by signature quantum number (either ri or
αi ) [4, 13, 17]. This quantum number is a consequence of the fact that full cranking
Hamiltonian is invariant with respect to a rotation through an angle π around the
cranking axis (x-axis)

Rx = exp(−iπ jx ), Rx ψi = exp(−iπ jx )ψi . (25)

The eigenvalues of Rx are exp(−iπα), where α is the signature exponent quantum


number. Alternatively, one can define signature quantum number as r = exp(−iπα).
Signature quantum number αi (ri ) of a single-particle orbital could take the αi = + 21
(ri = −i) or αi = − 21 (ri = +i) values. Similar to parity, this classification of the
single-particle states is an important tool for identifying the nucleon orbitals in the
rotating nuclear potential.
Second, the coupling between the different single-particle orbitals increases with
increasing rotational frequency. As illustrated in Fig. 3 this leads to the change of the
dominant Nilsson components of the wavefunction. Third, the slope of the orbitals
in Fig. 3 corresponds to the single-particle alignment h jx ii [4, 17]

∂ eω
i
h jx ii = − (26)
∂ω
for the case of cranking at fixed deformation. Note that single-particle routhians
are always plotted at fixed deformation in phenomenological potentials. In contrast,
they are given along the equilibrium deformation path in the DFT calculations. Thus,
the condition (26) is only approximately satisfied in the latter case provided that
deformation changes with increasing rotational frequency are modest. Note that in
general h jx ii depends on signature and this dependence is especially pronounced for
the single-particle orbitals with low value of Ω .
The knowledge of single-particle alignments is extremely useful for an under-
standing of physical situation and interpretation of experimental data. For exam-
ple, high- j or high-N intruder orbitals (such as those with Nilsson labels 1/2[770]
and 1/2[880]) have large values of h jx ii and, as a consequence, they are strongly
downsloping with increasing rotational frequency (see Fig. 3). Large energy split-
ting between two signature partner orbitals of a given single-particle state appear
for the states with Ω = 1/2 but no such splitting exist for the states with large value
of Ω . For example, this is a case for the 1/2[770] state for which a large signature
splitting between the α = ±1/2 branches leads to a formation of large N = 85 gap
at ω ≈ 0.3 MeV which is absent at ω = 0 MeV (see Fig. 3).
The advantage of the cranking model is that it provides a microscopic descrip-
tion of rotating nuclei, where the total angular momentum is described as a sum of
single-particle angular momenta, and thus collective and ‘non-collective’ rotations
can be treated on the same footing. On the other hand, there are limitations. The
12 A. V. Afanasjev

cranking-model approximation is semiclassical, because the rotation is imposed ex-


ternally. The model also breaks the rotational invariance, since a fixed rotation axis
is used. These latter limitations are not very important for very fast rotation (I  1).
Another shortcoming of the cranking model is that the wave functions are not eigen-
states of the angular momentum operator, which can lead to difficulties; for example,
a proper calculation of electromagnetic transition probabilities requires the use of
projection techniques [2]. Furthermore, the cranking model is a poor approximation
in the crossing region of two weakly interacting bands, but this difficulty can be
overcome by the removal of such crossings [13].

Spatial densities of the single-particle states

The total density of the nucleus is defined as the sum of the single-particle densities
of occupied states

ρtot (~r) = ∑ ρk (~r), (27)


k

which are given by

ρk (~r) = ψk∗ (~r)ψk (~r) (28)

for the k−th single-particle state. The single-particle wave function ψk (~r) can be
expanded into basis states |µ >

ψk (~r) = ∑ ckµ |µ >, (29)


µ

where µ represents the set of quantum numbers defining the basis state and ckµ
the expansion coefficients. Of special interest are the cases when this expansion is
dominated by a single basis state n (ckn ≈ 1) since then the nodal structure of the
wave function of the state ψk (~r) and consequently its single-particle density will
be predominantly defined by this basis state. This takes place either at spherical
shape or at extremely elongated nuclear shapes typical for rod shape structures or
megadeformation (MD) [21].
In the latter case, the wave function defined by asymptotic Nilsson quantum num-
ber is expanded into the basis states characterized by µ = [NnzΛ ]Ω [4, 21]. For
extremely elongated shapes in light nuclei this expansion is dominated by a single
basis state [21]. This is illustrated in Fig. 4 which shows the single-particle densities
of indicated states at the deformation β2 ≈ 1.6 typical for the MD shapes. They are
characterized by axial or nearly axially symmetric spheroidal/ellipsoidal like den-
sity clusters formed for the single-particle states with the [NN0]1/2 Nilsson quan-
tum numbers, doughnut density distributions for the [N01]Ω states, multiply (two
for nz = 1 and three for nz = 2) ring shapes for the [N, N − 1, 1]Ω Nilsson states with
N = 2 and 3. Fig. 4 clearly indicates the importance of the deformation which has
Model for independent particle motion 13

g9/2
0
f5/2

Neutron single-particle energies ei [MeV]


f7/2
-10 [321]1/2
/2 [4
2]5 40
[20 ]1/2
d3/2
s1/2 [321]3/2
/2
-20 [211]1
[330]
d5/2 1/2
[211]3/2
/2
1]1 3/2
-30 [10 [101]
p1/2 [220]1/2
p3/2
-40
[110]1/2
40
-50
Ca [000]1/2
s1/2
0 0.5 1 1.5
b2 - deformation

Fig. 4 Left panel: The Nilsson diagram for the neutron single-particle states in 40 Ca. It is based
on the results obtained in axial relativistic Hartree-Bogoliubov calculations with the NL3* func-
tional. Right panel: Single-neutron density distributions due to the occupation of indicated Nilsson
states.The box in right bottom corner exemplifies physical dimensions of the nucleus as well as the
colormap used for single-particle densities. Other density plots are reduced down to the shape and
size of the nucleus which is indicated by black solid line corresponding to total neutron density line
of ρ = 0.001 fm−3 . The colormap shows the densities as multiplies of 0.001 fm−3 ; the plotting
of the densities starts with yellow color at ρ = 0.001 fm−3 . Figure taken from Ref. [22] and it is
based on the results presented in Ref. [21].

two critical effects. First, it leads to the formation of density clusters with specific
nodal structure and to the separation of the clusters in space. Second, it lowers the
energies of the Nilsson states of the[NN0]1/2 type which favors the α-clusterization
and leads to the configurations in which all occupied single-particle states have this
type of structure.
It is impossible to experimentally measure the single-particle density distribu-
tions in rotating nuclei. That is a reason why contrary to the single-particle energies
and alignments they have not been used as fingerprints of independent particle mo-
tion. However, they are important for an understanding of the α-clusterization and
the evolution of extremely elongated shapes in light nuclei such as ellipsoidal and
rod shapes and nuclear molecules [21, 22]. The examples of total neutron densities
of later two types of nuclear shapes are provided in Fig. 5. A rod-shaped nuclear
configuration in 12 C is built from a linear chain of three α-clusters and could be
well understood from the summation of the single-particle densities of occupied
14 A. V. Afanasjev

0.1 0.1
[31,31], I = 21
5 rod−shape, I = 8.8 5
0.08 0.08

0.06 0.06

Y (fm)

Y (fm)
0 0
0.04 0.04

0.02 0.02
−5 12
C
(a) −5 36
Ar
(d)
0 0
−5 0 5 −5 0 5
Symmetry axis Z (fm) Symmetry axis Z (fm)

Fig. 5 Total neutron densities [in fm−3 ] of indicated configurations in the 12 C and 36 Ar nuclei
at specified spin values. The plotting of the densities starts with yellow color at 0.001 fm−3 . The
results are based on the cranked relativistic mean field calculations with the NL3* functional.
Figure taken from Ref. [21].

0.1 40
38
Ca
Ar
proton density (fm )

36
-3

34
S
Si

0.05

0
2 4
r (fm)

Fig. 6 Left panel: Proton densities of the N = 20 isotones obtained in the spherical relativistic
Hartree-Bogoliubov calculations with the DD-ME2 functional. Right panels: Neutron and proton
densities of the spherical 292 120172 nucleus obtained in non-relativistic Skyrme and Gogny as well
as covariant DFT calculations. Employed functionals are indicated. Figures are taken from Refs.
[23, 24].

single-particle states [21] (see also Refs. [25, 26, 27] for additional discussion of
rod-shaped nuclei built on linear chains of the α particles). Rod-shaped configura-
tions appear also in heavier nuclei such as 42 Ca and 44 Ti [21]. However, their densi-
ties are more uniform since the α-clusterization is suppressed by the occupation of
the single-particle orbitals which have either doughnut or ring type single-particle
Model for independent particle motion 15

density distributions (see Fig. 4 and Refs. [21, 28]). Nuclear molecules are character-
ized by the formation of the neck between two nuclear fragments: the configuration
of 36 Ar is an example of such structure (see Fig. 5).
A knowledge of the nodal structure of the single-particle densities of the states
allows to understand the microscopic mechanism of the transition between differ-
ent nuclear shapes. For example, in order to build nuclear molecules from typical
ellipsoidal density distributions one has to move matter from the neck (equatorial
region) into the polar one. This can be achieved by means of specific particle-hole
excitations which move the particles from (preferentially) doughnut type orbitals
or from the orbitals which have a density ring in an equatorial plane into the or-
bitals (preferentially of the [NN0]1/2 type) building the density mostly in the polar
regions of the nucleus [21, 22].
Another example of important role of the single-particle density of the states is
seen in so-called “bubble” structures of the ground states of spherical nuclei. This
is due to a specific radial nodal structure of the wavefunctions of the single-particle
states (see Fig. 6.2 in Ref. [4]): only s (l = 0) states build the density in the center
of the nucleus. On the contrary, the l > 0 states do not participate in the building of
the densities in the center of the nucleus due to presence of the centrifugal barrier
(see also Fig. 2 in Ref. [23]). The impact of this feature is especially pronounced
in proton subsystem of 34 Si [23, 29]. Left panel of Fig. 6 shows the evolution of
the proton densities in the N = 20 isotones. The 40 Ca, 38 Ar and 36 S nuclei in which
the proton 2s1/2 orbital is occupied show similar proton density patterns with the
density peak at the center. In contrast, the emptying of the proton 2s1/2 orbital in
34 Si leads to a considerable depletion of the proton density in the central region of

nucleus. This “bubble” phenomenon has been indirectly confirmed in experiment


[30].
Similar depletion of the density in the central region of the nucleus exist also in
superheavy nuclei (see Refs. [24, 31, 33] and right panel of Fig. 6). However, the
mechanism of its creation is different as compared with the one active in 34 Si since
the emptying of the s states is not possible in such nuclei. It relies on the fact that the
filling of low- j/high- j orbitals builds the density in the central/surface region of the
nucleus [24, 31]). Thus, starting from the flat density distribution in 208 Pb (which
is experimentally verified, see Fig. 2.4 in Ref. [4]) one can build “bubble”-type
structure in the 292 120172 superheavy nucleus by occupying predominantly high- j
orbitals outside the 208 Pb core [24]. It was suggested in Ref. [33] that the depletion
of density in central region is mostly due to Coulomb interaction. This, however,
contradicts to the observation that spherical superheavy nuclei with Z = 126 have
significantly smaller depletion of the density in the central region as compared with
the 292 120172 one [24].
16 A. V. Afanasjev

Microscopic+macroscopic models

The ”shell model” concepts discussed above assume some fixed mean field proper-
ties which do not depend on nucleonic configuration. In addition, they require the
use of deformation parameters in the case of deformed and cranked shell models
which have to be defined either from experimental data or from higher level model.
To overcome these deficiencies the microscopic+macroscopic (mic+mac) model
has been suggested in 1960s [34, 35, 36] which can be considered as an approxima-
tion to the Hartree-Fock approach [36]. In this model the total energy of the nucleus
Etot is separated into two parts, a macroscopic part Emacro and a microscopic part
Emicro ,

Etot = Emacro + Emicro . (30)

The macroscopic energy Emacro is defined by some version of the liquid-drop model
while the microscopic energy Emicro is obtained from quantal shell corrections Esh
calculated from a phenomenological potential using the Strutinsky prescription [34].
The shape of the nuclear surface is parametrized by means of a multipole expansion
of the radius in terms of the shape parameters (see Eq. (16)). Then the equilibrium
deformation in a specific nucleonic configuration is determined by a minimization
of the total energy Etot with respect to the shape parameters.
For a simplicity, only the basic features of the mic-mac method for the rotat-
ing nuclei are outlined here since the case of no rotation is easy to obtain by
dropping respective terms. The total nuclear energy Etot at a specific deformation
β̄ = β2 , γ, β4 , . . . and spin I0 is given as a sum of the rotating liquid drop energy and
the shell energy
1
I02 + Esh β̄ , I0 .
  
Etot β̄ , I0 = ELD β̄ , I = 0 + (31)
2Jrig (β̄ )

The shell energy is defined as the difference between the discrete and smoothed
(indicated by e ) single-particle energy sums,

Esh (I0 ) = ∑ ei (ω, β̄ ) I=I0 − ∑ eg


i (ω,
e β̄ )
I=I
e 0
(32)

with both terms evaluated at the same spin value I0 . The smoothed sum is calculated

using the Strutinsky procedure [34]. Then the total nuclear energy Etot β̄ , I0 and
equilibrium deformation β̄eq of specific nucleonic configuration can be calculated
as a function of spin I (defined via Eq. (24)). By considering numerous configura-
tions, defined by the occupation of single-particle orbitals with the specific sets of
quantum numbers such as parity, signature etc, one can build comprehensive spectra
of multiply rotational bands and compare them with experimental data.
Note that in the mic+mac approach the moment of inertia, defined from smoothed
single-particle quantities, has to be renormalized to a rotating liquid behavior. The
renormalization of the moment of inertia is especially important in the Nilsson po-
Model for independent particle motion 17

tential because of the l 2 -term but it is also required in the Woods-Saxon potential
because the potential radius is often different from the nuclear matter radius (see
Ref. [13] and references therein).
The consideration above is restricted to the situation of no pairing correlations.
The experience and detailed studies show that this is quite accurate approximation at
high spins [13, 37]. However, the pairing correlations play an important role at low
and medium spins. In the mic+mac approaches they are usually taken into account
at the BCS level by adding pairing energy term E pairing [38, 39, 64]
pair
Etot = Etot + E pairing . (33)

The mic+mac model is simpler than self-consistent approaches and it is also sub-
stantially cheaper in numerical calculations. Despite approximations and simplifi-
cations employed by this method, it provides an accurate description of the ground
state energies and deformations as well as fission barrier heights [38, 64] and rota-
tional properties of multiply bands in different nuclei [13, 39].

Self-consistent approaches: covariant density functional theory

There are several types of self-consistent density functional theories (DFTs) based
either on the non-relativistic Shrödinger equation with finite range Gogny or zero
range Skyrme forces or on the relativistic Dirac equation. The latter is called as co-
variant density functional theory (CDFT) [37] and its brief outline will be presented
here. It was very successful in the description of many physical phenomena (see
reviews in Refs. [37, 40, 41, 42]). The detailed reviews of the Skyrme and Gogny
DFTs are presented in Refs. [43, 44].
There are several classes of the CDFT models and, for simplicity, only meson-
exchange (ME) models will be considered here. In these models the nucleus is de-
scribed as a system of Dirac nucleons interacting via the exchange of mesons with
finite masses leading to finite-range interactions. The starting point of the ME mod-
els is a standard Lagrangian density [45]
 
1 − τ3 1 1
L = ψ̄ γ · (i∂ − gω ω − gρ~ρ ~τ − e A) − m − gσ σ ψ + (∂ σ )2 − m2σ σ 2
2 2 2
1 1 1 1 1
− Ω µν Ω µν + m2ω ω 2 − ~Rµν ~Rµν + m2ρ~ρ 2 − Fµν F µν , (34)
4 2 4 2 4
which contains nucleons described by the Dirac spinors ψ with the mass m and
several effective mesons characterized by the quantum numbers of spin, parity, and
isospin. The Lagrangian (34) contains as parameters the meson masses mσ , mω , and
mρ and the coupling constants gσ , gω , and gρ . e is the charge of the protons and it
vanishes for neutrons. The coupling constants are density dependent in the density
dependent meson exchange (DDME) class of covariant energy density functionals
(CEDFs) [46, 47]. In contrast, they are constant in the so-called non-linear (NL)
18 A. V. Afanasjev

CEDFs in which the density dependence is introduced via the powers of the σ -
meson [48]:
1 1
LNL = L − g2 σ 3 − g3 σ 4 . (35)
3 4
The solution of these Lagrangians leads to the relativistic Hartree-Bogoliubov
(RHB) equations [37]. They are illustrated below on the example of the cranked
RHB (CRHB) equations for the fermions in the rotating frame (in one-dimensional
cranking approximation) [49]

ĥD − λτ − Ωx Jˆx ∆ˆ
    
Uk (rr ) Uk (rr )
∗ = Ek , (36)
−∆ˆ ∗ −ĥD∗ + λτ + Ωx Jˆx Vk (rr ) Vk (rr )

where λτ (τ = p, n) are chemical potentials defined from the average particle num-
ber constraints for protons and neutrons; Uk (rr ) and Vk (rr ) are quasiparticle Dirac
spinors; Ek denotes the quasiparticle energies; and Jˆx is the angular momentum com-
ponent entering into the Coriolis term −Ωx Jx . Here, ĥD is the Dirac Hamiltonian for
the nucleon with mass m

ĥD = α(−i∇ −V
V (rr )) +V0 (rr ) + β (m + S(rr )), (37)

which contains an attractive scalar potential S(rr )

S(rr ) = gσ σ (rr ), (38)

a repulsive vector potential V0 (rr )

1 − τ3
V0 (rr ) = gω ω0 (rr ) + gρ τ3 ρ0 (rr ) + e A0 (rr ), (39)
2
and a magnetic potential V (rr )

1 − τ3
V (rr ) = gω ω (rr ) + gρ τ3 ρ (rr ) + e A (rr ). (40)
2
These equations are solved numerically in triaxial harmonic oscillator basis and
signature basis is used for the single-particle states. The CRHB framework is ap-
plicable to the description of both rotating nuclei in the paired regime and one-
/two-quasiparticle configurations in rotating and non-rotating nuclei. An approxi-
mate particle number projection by means of the Lipkin-Nogami (LN) method is
also used in it [49] but its discussion is omitted for simplicity. The CRHB+LN cal-
culations have been successful in the description of rotating nuclei in paired regime
[37, 49, 50] and one-quasiparticle configurations in odd-A nuclei (see discussion
of Fig. 12 below). Moreover, the CRHB framework can be either reduced to un-
paired regime, leading to so-called cranked relativistic mean field (CRMF) approach
[51, 52], or upgraded to the 2- and 3-dimensional cranking approximation [19]. By
putting Ωx = 0 it can also be applied to the description of non-rotating nuclei, but
Model for independent particle motion 19

more specialized versions of the RHB computer codes (without cranking and signa-
ture basis) designed for triaxial, axially symmetric and spherical shapes also exist
[53].
It turns out that without any assumptions on the form of the single-particle poten-
tial self-consistent calculations generate single-particle properties (energies, align-
ments) which in general are similar to those obtained in phenomenological poten-
tials. This is illustrated in Fig. 9 below which compares the single-particle spectra of
the superheavy 292 120172 nucleus obtained with phenomenological folded Yukawa
(FY) potential, non-relativistic Skyrme functionals SkP, SkM*, SLy6, SLy7, SkI1,
SkI4 and Sk3, and CEDFs NL3, NL-Z, NL-Z2 and NL-VT1. The Nilsson diagrams
obtained in the RHB calculations with the CEDF NL3* (see Fig. 2) show a lot of
similarities with those obtained in the Woods-Saxon potential (see, for example,
Figs. 3 and 4 in Ref. [12]). The behavior of the single-particle states in the rotat-
ing potential obtained in the Woods-Saxon potential (see Fig. 3) is very similar to
that obtained in the CRMF calculations (see Fig. 4 in Ref. [52]). However, there
is a principal difference in the physical mechanisms related to the single-particle
degrees of freedom between two types of approaches and it is related to time-odd
mean fields [54, 55]. They are absent in phenomenological potentials but are present
in self-consistent models. For example, in the CRHB framework they are related to
the terms which break time-reversal symmetry, namely, a magnetic potential V (rr )
(see Eq. (40)) and the the Coriolis term −Ωx Jx . In the DFT approaches, time-odd
mean fields modify single-particle energies [56, 57], single-particle alignments [55]
and have large impact on rotational properties of nuclei [54, 55].

Manifestation of independent particle motion in non-rotating


and rotating nuclei

When considering the manifestations of independent particle motion one should


separate the effects which emerge from the shell structure (as a coherent effect of
motion of many particles) and those from individual motion of the particles. The
configuration-mixing interactions [which are accumulated in residual interaction
term of the Hamiltitonian given by Eq. (4)] such as pairing and the coupling to the
low-lying collective vibrational degrees of freedom act destructively on the individ-
ual properties of the single-particle orbitals. In contrast, the global effects emerging
from the underlying shell structure are not that much affected by these residual in-
teractions. Thus, a global shell structure at spin zero, the superdeformation at high
spin and the existence of superheavy nuclei are considered as the examples of such
effects.
Then the physical properties which sensitively depend on the single-particle fea-
tures of individual orbitals are discussed. First, the energies of experimental and
calculated one-quasiparticle states in non-rotating deformed nuclei are compared.
Then, the analysis is extended to rotating nuclei at low and medium spins. The rota-
tion acts as a tool to significantly reduce the role of pairing interaction [58, 59, 60].
20 A. V. Afanasjev

0.5
200 0.4
DD-PC1
Proton number Z
0.3
160 0.2
0.1
0
120 -0.1
N=258 -0.2
80 Z=82 -0.3
N=184 -0.4
40
Z=50 -0.5
N=126 -0.6
50 N=82 -0.7
0 -0.8
0 40 80 120 160 200 240 280 320 360 400 440
Neutron number N
Fig. 7 Nuclear landscape at spin zero as obtained in the RHB calculations with the DD-PC1 func-
tional. For proton numbers Z ≤ 130, the ground states have ellipsoidal shapes. These nuclei are
shown by the squares the color of which indicates the equilibrium quadrupole deformation β2 (see
colormap). Toroidal shapes dominate nuclear landscape for higher proton numbers (white region
located between two-proton and two-neutron drip lines for toroidal nuclei shown by solid black
lines). Figure taken from Ref. [61].

Thus, nuclear systems at very high spins in which the impact of pairing is negligible
are considered and the single-particle and polarization effects due to the occupation
of specific orbitals are analyzed. In addition, the phenomenon of band termination
is discussed as an example of the competition of the collective and single-particle
degrees of freedom.

Global shell structure at spin zero

The state-of-the-art view on the nuclear landscape is shown in Fig. 7. It is based on


the results of the calculations presented in Refs. [61, 62, 63]. The bands (shown by
gray color) of spherical nuclei along proton and neutron numbers 8, 20, 28, 50 and
82 (as well as for neutron number N = 126) are seen in this figure. They are due to
large spherical shell gaps at these particle numbers which provide an extra stability
of the nuclei. Note that these bands are more pronounced in neutron subsystem.
Outside of these bands of spherical nuclei, the ground states of the Z ≤ 120 nuclei
are either oblate or prolate being in typical range of quadrupole deformations |β2 | <
0.4. Similar structures and features are also obtained in non-relativistic calculations
of Refs. [64, 65, 66] but they are restricted to the nuclei below Z ≈ 120. In general,
existing experimental data confirms these model predictions [67] but there may be
some differences between specific model and experiment.
With increasing proton number these classical features disappear and only toroidal
shapes are calculated as the lowest in energy. This region (shown in white color be-
tween two black lines in Fig. 7) is penetrated only by three islands (shown in gray
Model for independent particle motion 21

color) of potentially stable spherical hyperheavy nuclei. Their existence is due to


substantial proton Z = 154 and 186 and neutron N = 228, 308, and 406 spherical
shell gaps [61] and substantial fission barriers around spherical minima [61, 63].
However, these states are highly excited with respect of minima corresponding to
toroidal shapes. They could become the ground states if relevant toroidal minima
are unstable with respect of so-called sausage deformations [61, 68].
Thus, the richness of nuclear structure seen in experimentally known part of nu-
clear landscape is replaced by a more uniform structure of the nuclear landscape in
the region of hyperheavy (Z ≥ 126) nuclei dominated by toroidal nuclei [61, 63].
This transition from compact ellipsoidal-like shapes to non-compact toroidal shapes
is driven by the enhancement of the role of Coulomb interaction with increasing Z
[63]. Thus, the former shapes become either unstable against fission or energetically
unfavored in hyperheavy nuclei.
Single-particle degrees of freedom play also an important role in the definition of
the boundaries of nuclear landscape. For example, the position of two-neutron drip
line of ellipsoidal nuclei depends sensitively on the single-particle energies of high-
j states located in the vicinity of neutron continuum threshold [69]. The transition
from ellipsoidal to toroidal shapes drastically modifies the underlying single-particle
structure and as a result lowers the energy of the Fermi level for protons [61]. As
a consequence, a substantial shift of the two-proton drip line from its expected po-
sition for ellipsoildal shapes (shown by dashed orange line in Fig. 7) towards more
proton-rich nuclei for toroidal shapes (shown by solid line in Fig. 7) takes place
[61].

Superheavy nuclei

With increasing proton number beyond Z = 100 the fission barriers provided by
the liquid drop become rather small and then for higher Z they disappear (see Fig.
10.6 in Ref. [4]). It turns out that the existence of the heaviest nuclei with Z > 104
is primarily determined by shell effects due to the quantum-mechanical motion of
protons and neutrons inside the nucleus [70, 71, 72, 73, 74]. In these nuclei the
heights of fission barriers are entirely determined by the shell corrections and they
would not exist without shell effects. These effects also play a central role for the
production, stability and spectroscopy of superheavy nuclei.
Although state-of-the-art theoretical models provide a reasonable description of
many aspects of the physics of superheavy nuclei, they face substantial challenges in
the prediction of the location of next spherical shell closures [31, 32, 72, 73, 74, 75].
These challenges are illustrated in Fig. 8 which shows the map of calculated ground
state quadrupole deformations obtained with two covariant energy density function-
als. The predictions of these two functionals are drastically different in the vicinity
of the Z = 120 and N = 184 lines. The PC-PK1 functional predicts wide bands of
spherical nuclei in the (Z, N) chart along Z = 120 and N = 184. In contrast, in the
calculations with DD-PC1, the band along Z = 120 does not exist and narrower band
22 A. V. Afanasjev

Fig. 8 Charge quadrupole deformations β2 of even-even superheavy nuclei obtained in the RHB
calculations with indicated functionals. Experimentally known nuclei are indicated by white cir-
cles. Figure taken from Ref. [75].

along N = 184 is seen only for Z ≤ 114. These discrepancies are due to modest dif-
ferences in the single-particle properties of these functionals and related differences
in the competition of shell effects at spherical and deformed shapes [75]. Note that
the width of the band of spherical nuclei in the (Z, N) chart along a specific particle
number corresponding to a shell closure indicates the impact of this shell closure on
the structure of neighboring nuclei.
There is a wide variety of the predictions for proton and neutron spherical shell
closures in superheavy nuclei obtained in different models. These are proton num-
bers at Z = 114, 120 and 126 and neutron numbers at N = 172 and 184 (see Fig. 9
and Refs. [31, 32, 73, 75]. However, in the CDFT the N = 172 could be eliminated
as a potential candidate when the deformation effects are taken into account [75]
and the Z = 120 nuclei become deformed in a number of functionals when both de-
formation and correlations beyond mean field effects are considered [76]. The chal-
lenges the models face in the prediction of spherical shell closures are due to two
factors. At present, self-consistent models do not provide a spectroscopic quality of
the description of the single-particle spectra [14]. In contrast, the phenomenologi-
cal potentials better describe single-particle spectra in known nuclei (see Ref. [32]
and references therein) but they fail to incorporate self-consistency effects (such
as a depletion of the density in central region of nucleus) which are important in
superheavy nuclei (see discussion of Fig. 6).
Fig. 8 illustrates that only for a few experimentally known Z = 116 and 118 nu-
clei the CDFT predictions differ. In general, with possible exception of these high-
Z isotopic chains, existing experimental data on superheavy nuclei are described
with comparable level of accuracy by all existing models [73, 74, 75]. This is un-
deniable success of independent particle model the global consequences of which
(shell structure) led to the predictions of superheavy nuclei in 1960s within rel-
atively simple models [70, 71]. Experimental data collected over these years fully
confirmed these predictions and some differences in the predictions of next spherical
shell closures are secondary to this success. This field or research (both experimen-
tal and theoretical) remains extremely active which is illustrated by recent reviews
[72, 73, 74]. However, at present available experimental data does not allow to give
a preference to the predictions of one or another model and to decide where next (if
Model for independent particle motion 23

Fig. 9 Single-particle energies of proton (top) and neutron (bottom) states in the 292 120172 nucleus
obtained at spherical shape with indicated non-relativistic and relativistic functionals. Figure taken
from Ref. [31].

any) proton and neutron spherical shell closures are located or whether there is an
island of spherical superheavy nuclei.

Superdeformation at high spin

Fig. 7 shows that the ground states of the nuclei in experimentally known part of
the nuclear chart are characterized by quadrupole deformation β2 which is located
typically in the range −0.35 ≤ β2 ≤ +0.35. Strutinsky has predicted excited min-
imum with β2 ∼ 0.6 in the second (superdeformed, SD) potential well of potential
24 A. V. Afanasjev

energy curves of deformed nuclei in Ref. [34]. At low spin such minima are located
at high excitation energies, but they can be brought down to the yrast line by fast
rotation since it favors extremely deformed shapes. This was confirmed later in the
mic+mac calculations of Ref. [77] which predicted the doubly magic SD band in
152 Dy with 2:1 semi-axis ratio. The existence of the SD bands is due to the shell

effects associated with large proton and neutron SD shell gaps. For example, in the
A ∼ 150 region of superdeformation these shell effects are produced by large proton
Z = 66 and neutron N = 86 SD shell gaps which exist both in phenomenological
potentials (see Fig. 3 in this paper and Refs. [59, 79]) and in the DFT calculations
[52].
The phenomenon of superdeformation at high spin has been confirmed experi-
mentally by the observation of the first SD band in 152 Dy in 1986 [78]. From that
time, a significant amount of experimental data on the SD bands in different mass
regions has been collected [80]. The mic+mac and DFT-based cranked approaches
rather well describe experimental observables such as dynamic J (2) and kinematic
J (1) moments of inertia and transitional quadrupole moments Qt of these bands (see
Refs. [49, 52, 59, 79] and references quoted therein). The calculations also reveal
a substantial dependence of these physical observables on the occupation of high-
N intruder orbitals; here N stands for the principal quantum number of dominant
component of the wave function. For example, they strongly depend on the number
of occupied N = 6 protons and N = 7 neutrons in the A ∼ 150 region of superde-
formation. This region is also of special interest since pairing is negligible in the
majority of the SD bands. As a consequence, it provides one of the best examples
of independent particle motion in nuclear physics (see detailed discussion in the last
section).

The phenomenon of band termination

One of clear manifestations of the independent particle motion is the phenomenon


of band termination [13, 81, 82]. It definitely reveals the fact that each single-
particle orbital occupied in nucleonic configuration possesses a limited and state-
dependent angular momentum. Of special interest are so-called smooth terminating
bands which show a continuous transition from high collectivity at low and medium
spin values to a pure particle-hole (terminating) state at the maximum spin which
can be built within the configuration [13, 83]. Note that this feature of finite multi-
fermion systems has so far only been observed and studied in atomic nuclei. In the
terminating state, the symmetry axis coincides with the rotation axis. Since collec-
tive rotation along the symmetry axis is forbidden in quantum mechanics, no fur-
ther angular momentum can be brought into the system with the same occupation of
single-particle orbitals and thus this state represents the termination of the rotational
band.
The phenomenon of band termination has been observed in several regions of
nuclear chart (see review in Ref. [13]). However, the best examples of smooth ter-
Model for independent particle motion 25

0 2
L
1
3

−1
E− 0.013 I(I+1) [MeV]

109
Sb expt.
−2

0 [21,3]
+/−

+
87/2

83/2

[21,2] +
−1 89/2

109
Sb theory
−2
10 15 20 25 30 35 40 45 50

I [h ]

Fig. 10 The comparison of experimental and calculated E − ERLD curves for the bands 1 − 3 in
109 Sb. The energies have been normalized in such a way that calculated and experimental curves

for the band 1 have the same value at the minimum of the E − ERLD curve. The energies of the
unlinked experimental bands 2 and 3 are chosen so that their E − ERLD minima have the same
energies as theoretical counterparts. Terminating states are indicated by large open circles and
their spins are displayed. Solid and dashed lines are used for the configurations with π = + and
π = −, respectively. The states with signature αtot = +1/2 and αtot = −1/2 are shown by solid an
open symbols, respectively. Figure taken from Ref. [13].

minating bands are seen in the A ≈ 110 region in which the nuclei have several
valence particles and holes outside the 100 Sn core. A classical example of termi-
nating bands in 109 Sb [13, 83] is considered here in order to illustrate the major
features of smooth band termination. A critical feature of these bands is the fact that
their dynamic moments of inertia J (2) gradually decrease with increasing rotational
frequency to unusually low values (near 1/3 of rigid-body value) at the highest ob-
served frequencies. This is definite indication of significant suppression of pairing
correlations [13] making this type of rotational structures as one of the best exam-
ples of independent particle motion.
Fig. 10 compares the experimental and calculated energies E of rotational bands
with respect of rigid rotor reference ERLD = AI(I + 1), where A is the moment
of inertia parameter. The calculations are performed in configuration-dependent
cranked Nilsson-Strutinsky approach (CNS) approach [13, 84]. The configurations
relatively to the 100 Sn core are labelled using the shorthand notation [p1 p2 , n]αtot ≡
26 A. V. Afanasjev

4.0
109
Sb − band 1
3.5 −
[21,2]
0.2 60° 50° 40°
3.0
109
Sb

Qt (eb)
30° 2.5

2.0
[21,3]−
39.5+

37.5 [21,2] − 20° 1.5
0.1 40.5+ [21,3]+
1.0
γ 16 20 24 28

I (h )
32 36 40

10°
21.5+
19.5−

22.5+

0.0 0.1 0.2 0.3
ε2
Fig. 11 (left panel) Calculated deformation paths in the (ε2 , γ) plane of the configurations assigned
to smooth terminating bands 1-3 in 109 Sb. The spins of some states are shown. (right panel) The
comparison of experimental and calculated transition quadrupole moments Qt of the band 1 and
assigned theoretical configuration. Figure taken from Ref. [13].

[π(g9/2 )−p1 (h11/2 ) p2 (g7/2 d5/2 )Z−50+p1 −p2 ⊗ν(h11/2 )n (g7/2 d5/2 )N−50−n ]αtot , where
αtot is the total signature of the configuration (only sign is shown) and p1 , p2 and n
are the numbers of proton holes in the g9/2 orbitals, of protons in the h11/2 orbitals
and of neutrons in the h11/2 orbitals, respectively. One can see that the CNS calcu-
lations without pairing very well reproduce experimental data. Some discrepancies
seen at low I ≤ 20h̄ spin are due to the neglect of pairing.
The CNS calculations suggest that rotational bands of interest are near prolate
at low spin and thus they involve collective rotation about an axis perpendicular
to the symmetry axis. With increasing spin the valence nucleons gradually align
their spin vectors with the axis of rotation via the Coriolis interaction. This causes
the nuclear shape to gradually trace a path through the triaxial (γ) plane, toward
the non-collective oblate shape at γ = +60◦ (see left panel of Fig. 11). After the
available spin is exhausted, consistent with the Pauli principle, the band termi-
nates. This gradual change from collective near-prolate (γ ∼ 0◦ ) to non-collective
oblate (γ = +60◦ ) shape leads to a gradual decrease of transition quadrupole mo-
ment Qt which agrees with available experimental data (see right panel in Fig.
11). The termination spins, which depend on the configuration, are well defined
property for terminating bands and they confirm the termination process. For ex-

ample, the detailed structure of the [21, 2]− (band 1) terminating I = 83 2 state
in 109 Sb is π(g9/2 )−2
8 (g d )2 (h )1 ⊗ ν(g
7/2 5/2 6 11/2 5.5 d )6 (h
7/2 5/2 12 11/2 10)2 . Note that fully

self-consistent cranked relativistic mean field calculations confirm this physical pic-
ture [37]. However, due to technical reasons it is difficult to follow smooth termi-
nating bands up to their terminating states in such calculations.
Model for independent particle motion 27

UNEDF2

UNEDF2

NL3*
NL3*
SLy4

SLy4

D1M
D1M

NL1
NL1
D1S

D1S
exp.

exp.
2
5/2[613] 5/2[613] 9/2[604]
249 1/2[651] 251
One-quasiparticle energies [MeV] Bk Cf 11/2[606]

1 9/2[604]
7/2[514] 9/2[615]
1/2[521] 1/2[7]
11/2[725]
3/2[622]
7/2[633] 7/2[613]
0 3/2[521] 1/2[620]

1/2[400] 9/2[734]
5/2[642] 5/2[622]
1/2[550] 3/2[651]

-1 7/2[624]
5/2[523] 7/2[743]
7/2[743]
11/2[505] 1/2[6]
1/2[6]
1/2[550] 1/2[631] 1/2[631]
-2

Fig. 12 Experimental and calculated quasiparticle spectra in 249 Bk and 251 Cf. Solid and dashed
lines are used for positive- and negative-parity states, respectively. The states are labeled by the
Nilsson label of the dominant component of the wave function when the squared amplitude of
this component exceeds 50%. Otherwise, they are labeled by Ω [N] where N is principal quantum
number of the components of wave function whose cumulative contribution into the wave function
is dominant. Figure taken from Ref. [14].

Single-particle states in deformed nuclei

Non-rotating nuclei

The detailed comparison of experimental and theoretical information on the prop-


erties of specific individual states can shed additional light on the validity in inde-
pendent particle motion in atomic nuclei. In non-rotating nuclei, such information
is provided by the energies of the single-particle states, their densities and by the
transitions between the single-particle states. However, such densities are not acces-
sible experimentally and calculated transition probabilities are prone to substantial
theoretical errors. One could imagine that spherical nuclei would provide the clean-
est and simplest set of data on the single-particle properties. It turns out that this is
not a case because of substantial residual interaction due to (quasi)particle-vibration
(quasiparticle-phonon) coupling. As a consequence, the wavefunctions of the states
in the spherical nuclei are not of pure single-particle nature since they are substan-
tially polluted by the vibrational admixtures [85, 86, 87].
The impact of the quasiparticle-vibration coupling is reduced in deformed nu-
clei since the part of such correlations is accounted at the level of deformed mean
field. Indeed, the admixtures of the phonons to the structure of the ground and low-
lying states in deformed rare-earth and actinide odd-mass nuclei is relatively small
28 A. V. Afanasjev

Fig. 13 Experimental (left panel) and calculated (right panel) Routhians relative to the g-band
reference Routhian Ere f as a function of rotational frequency h̄ω1 . Note that the data are restricted
to the simplest configurations since two-quasiproton and several three-quasineutron configurations
are left away. The convention for (π, α) is the following: open and solid symbols are used for
π = + and π = −, respectively, and the α = +1/2 and α = −1/2 states are shown by triangles up
and triangles down, respectively. Figures taken from Ref. [90].

(especially, when compared with spherical open shell nuclei [86]) according to the
quasiparticle-phonon model [88, 89]. Thus, deformed nuclei can provide a better
and cleaner examples of independent particle motion.
Fig. 12 compares experimental and calculated energies of one-quasiparticle states
Eiqp in odd-A 249 Bk and 251 Cf nuclei. The results of the calculations with non-
relativistic Skyrme (SLy4, UNEDF2), Gogny (D1S, D1M) and covariant (NL1,
NL3*) energy density functionals are shown in this figure. These are fully self-
consistent calculations which means that the total energies Ei of the nucleonic con-
figurations with blocked i-th single-particle state of interest are obtained within
Hartree-Fock-Bogoliubov or RHB frameworks [2]. Then the ground state is asso-
ciated with nucleonic configuration which has the lowest energy Elowest and the
energies of the excited states Eiqp are defined as Eiqp = Ei − Elowest . Note that in
Fig. 12 the particle and hole states are plotted at Eiqp and −Eiqp , respectively. This
is done in order to facilitate the comparison of these self-consistent results with the
Nilsson diagrams.
One can see that the energies of some deformed one-quasiparticle states are rather
well described in specific functionals, but others deviate appreciably from the exper-
iment. The detailed comparison of experimental data with calculations is presented
in Ref. [14]. Considering an average level of the accuracy of the description of ex-
perimental data, it is difficult to give a clear preference to one or another functional.
At present, a systematic analysis of the accuracy of the reproduction of the single-
particle spectra in deformed nuclei is available only in the RHB framework [91]
and smaller in scope (only for the Rb, Yb and Nb isotopic chains) analysis is car-
ried out in the HFB calculations with Gogny forces (see Ref. [92] and references
therein). These investigations reveal two sources of inaccuracies in the description
of the energies of the single-particle states, namely, low effective mass leading to
a stretching of the energy scale of the calculated results as compared with experi-
Model for independent particle motion 29

mental ones and incorrect relative positions of some single-particle states [14, 91].
On the absolute scale these deficiencies are not large especially considering the
fact that no (or very limited) information on single-particle degrees of freedom has
been taken into account in the fitting protocols of covariant (non-relativistic) en-
ergy density functionals. The accounting of quasiparticle-vibration coupling (QVC)
improves the agreement with experiment by both improving the description of the
energies of individual states and increasing the density of the single-particle states
in the vicinity of the Fermi level. This was illustrated for the 251 Cf nucleus in the
RHB+QVC framework in Ref. [93].
In general, phenomenological potentials provide better description of the ener-
gies of the single-particle states in deformed odd-A nuclei because they are fitted
to this type of experimental data [8, 9, 84]. However, such accuracy is obtained at
the cost of neglect of self-consistency effects [91] and effective incorporation of the
effects of particle-vibration coupling [93].

Rotating nuclei in the pairing regime

As discussed earlier the rotation of the nuclei can generate substantial modifications
of the single-particle energies and provides single-particle alignments h jx ii as a new
measure of the single-particle properties. This leads to a new and very robust probes
of the single-particle structure since some pairs of the orbitals emerging from a
given non-rotating state show significant signature splitting (see, for example, the
[770]1/2(α = ±1/2) and [761]3/2(α = ±1/2) pairs of the orbitals in Fig. 3) while
other pairs (such as [532]5/2(α = ±1/2) and [633]7/2(α = ±1/2) in Fig. 3) show
no signature splitting.
These features can be used for a reliable interpretation of experimental data. This
is demonstrated in Fig. 13 which compares experimental Routhians with calculated
ones for selected set of rotational bands in odd-neutron 163 Er nucleus. The excitation
energies are taken relative to a reference configuration which is the ground state (g-)
configuration in even-even nucleus. Experimental Routhian is well approximated by
the Harris expression

ω2 ω4 h̄2
Ere f = Θ0 + Θ1 + (41)
2 4 8Θ0

with Θ0 = 32h̄2 MeV−1 and Θ1 = 32h̄4 MeV−3 extracted from the ground state band
in 164 Er. The calculations are performed in the cranked shell model (CSM) [94]
which employs fixed mean-field parameters for the Nilsson potential and pairing.
The parameters of its Ere f reference are adjusted to the calculated Routhian of the
g-configuration. The quasiparticle orbitals, on which rotational bands are build, are
labelled by the letters of the alphabet A, B, C , D (for high- j intruder states) and E, F,
G, ... (for normal parity states) [see Ref. [90] for details of this labelling convention].
One can see that this relatively simple model can describe rather well experimentally
observed features such as (i) large signature splitting in the pair of bands A/B, (ii)
30 A. V. Afanasjev

the lack or small amount of signature splitting in the pairs of bands E/F, X/Y and
G/H, (iii) the presence of paired band crossing at h̄ω1 ≈ 0.25 MeV in the E/F, G/H
and C bands which reflects itself in the change of the slope of the Routhians as a
function of rotational frequency and (iv) the absence of paired band crossing in the
A/B and X/Y pairs. The latter is due to the fact that the occupation of either of these
orbitals blocks paired band crossing in neutron subsystem.
Over the years huge amount of experimental data on rotating nuclei has been
accumulated and more sophisticated theoretical tools have been developed and suc-
cessfully applied for the description of rotating nuclei both in paired and unpaired
regimes. They go beyond the basic assumption of the CSM on the constancy of
the mean and pairing fields and take into account their configuration dependence
either on the level of mic+mac model or on a fully self-consistent level. These in-
clude cranked Nilsson-Strutinsky approach [13, 39], cranked approaches based on
the Skyrme and Gogny DFTs [95, 96, 97] and CDFT [50, 51, 98] and others.

Single-particle and polarization effects due to the occupation of


single-particle orbitals

The addition or removal of particle(s) to the nucleonic configuration modifies the


total physical observables. But it also creates the polarization effects on the physical
properties (both in time-even and time-odd channels) of initial configuration. The
comparison of relative properties of two configurations can shed important light
both on the impact of the added/removed particle(s) in specific orbital(s) on phys-
ical observable of interest and on the related polarization effects. In addition, the
comparison with experimental data on such properties can provide a measure of the
accuracy of the description of single-particle properties in model calculations.
Let us consider an example of rotational bands in the regime of weak pairing.
The relative properties of different physical observables such as relative charge
quadrupole moments

∆ Q(ω) = QB (ω) − QA (ω), (42)

and relative effective alignments [79]

ieB,A
f f (ω) = IB (ω) − IA (ω), (43)

are defined as the differences between either charge quadrupole moments Q or the
spins I of the bands A and B at the same rotational frequency ω with the band A
being a reference band.
They provide sensitive probes of the single-particle motion and polarization ef-
fects induced by the occupation of specific single-particle orbitals. In addition, the
experimental values of relative charge quadrupole moments are to a large degree
free from the uncertainties due to stopping powers which impact the absolute val-
Model for independent particle motion 31

Table 1 Experimental and calculated relative charge quadrupole moments ∆ Q = Q(Band) −


Q(152 Dy(1)) of the 149 Gd(1), 151 Tb(1) and 151 Dy(1) superdeformed bands. The detailed struc-
ture of the configurations of these bands relative to the doubly magic 152 Dy superdeformed core is
given in column 2. Column 5 shows the sum ∑i ∆ Qi of the ’independent’ contributions ∆ Qi of i-th
particle to the charge quadrupole moment calculated at rotational frequency ω = 0.50 MeV. Note
that the values in columns 3 and 4 are averaged over the observed spin range. Based on the results
of the CRMF calculations with the NL1 functionals and non-relativistic Skyrme DFT calculations
with SkP and SkM* functionals presented in Refs. [52, 99].

Band Configuration ∆ Qexp (eb) ∆ Qth (eb) ∑i ∆ Qi


1 2 3 4 5
149 Gd(1) ν[770]1/2(r = −i)−1 (π[651]3/2)−2 −2.5(0.3) −2.41 [NL1] −2.44 [NL1]

−2.42 [SkP]
−2.32 [SkM*]
151 Tb(1) π[651]3/2(r = +i)−1 −0.7(0.7) −1.01 [NL1] −0.99 [NL1]
−0.96 [SkP]
−0.96 [SkM*]
151 Dy(1) ν[770]1/2(r = −i)−1 −0.6(0.4) −0.53 [NL1] −0.55 [NL1]
−0.57 [SkP]
−0.48 [SkM*]

ues of these moments. The ∆ Q(ω) probes the differences in time-even mean fields
generated by the addition or removal of particle(s) from the configuration of the
reference band [52, 99].
The effective alignment ie f f depends both on the alignment properties of single-
particle orbital(s) by which the two bands differ and on the polarization effects (both
in time-even and time-odd channels of the DFTs) induced by the particles in these
orbitals [52]. It can also be used to investigate experimentally the structure of the
underlying single-particle orbitals in the configurations of interest [52, 79]. Note
that additivity principle for the single-particle observables has been tested for the
first time on the example of effective alignments of superdeformed bands in the
A ∼ 150 mass region [79].
The comparison of experimental and calculated ∆ Q(ω) values for a selected set
of superdeformed bands in the nuclei near 152 Dy is presented in Table 1 (for a more
extensive set of data see Table 2 in Ref. [99]). One can see that these quantities (col-
umn 3) are well described in model calculations (columns 4 or 5). In addition, the
additivity rule of quadrupole moments [99] which states that relative quadrupole
moments of two configurations ∆ Q(ω) can be well approximated by the sum of
individual contributions ∆ Qi of i-th particles to the charge quadrupole moment de-
fined with respect of core SD configuration in 152 Dy

∆ Q(ω) ≈ ∑ ∆ Qi (44)
i

is rather well fulfilled both in relativistic and non-relativistic DFT calculations (com-
pare columns 4 and 5 in Table 1 and see Ref. [99]). This means that the polarization
effects due to individual particles or holes are largely independent.
32 A. V. Afanasjev

CRMF - solid lines Exp. data - unlinked circles


CNS - dashed lines
3.0
(a) 1.0 (b)
149
2.0 Gd(1)/150Tb(1) π[651]3/2(r = + ι)
0.0
151
1.0 Tb(1)/152Dy(1)
-1.0
0.0
π[651]3/2(r = – ι) -2.0
-1.0
2.0
3.0 150
152
Dy(1)/153Dy(1)
Tb(1)/151Tb(1)
1.0
Effective alignment ιeff

(d)
2.0
ν[761]3/2(r = + ι)
(c) 0.0
1.0

0.0
0.4 π[301]1/2 151
ν[770]1/2(r = – ι) Tb(4)/152Dy(1)
-1.0 0.2 r= – ι
0.0 r= – ι
0.4 -0.2 (f)
(e) r=+ι
ν[402]5/2(r = ± ι)
0.2 -0.4
0.0 -0.6
-0.2 152 -0.8
Dy(1)/153Dy(2,3) 151 r=+ι
Tb(2)/152Dy(1)
-0.4 -1.0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Rotational frequency Ω (MeV) Rotational frequency Ω (MeV)

Fig. 14 Effective alignments, ie f f (in units h̄), extracted from experimental data are compared with
those obtained in the CRMF and CNS calculations of Refs. [52, 79]. The experimental effective
alignment between bands A and B is indicated as ‘A/B’. The band A in the lighter nucleus is
taken as a reference, so the effective alignment measures the effect of the additional particle. The
calculated configurations differ in the occupation of the orbitals indicated by the quantum numbers
in the panels. Figure taken from Ref. [102].

Experimental and calculated ie f f values for single-particle orbitals active in the


vicinity of the SD shell closures in the A ≈ 150 region of superdeformation are
compared in Fig. 14. The calculations are carried out within the cranked relativistic
mean field [52] and the cranked Nilsson-Strutinsky [79] approaches. The CRMF
calculations reproduce in average the experimental ie f f values better than the CNS
approach. This indicates that alignment properties of single-particle orbitals and
their polarization effects are correctly accounted for in the CRMF approach. The
discrepancy between the CRMF calculations and experiment seen in Fig. 14a at
Ω ≤ 0.5 MeV is most likely due to the fact that pairing correlations play some role
at low rotational frequency in the 149 Gd(1) band.
It is necessary to point on principal differences in the description of these
relative observables in the DFT and mic+mac based approaches. The relative
quadrupole moments are self-consistently described in the DFT based approaches
[52, 99, 100]. In contrast, the effective single-particle quadrupole moments in the
Model for independent particle motion 33

mic+mac method are not uniquely defined because of the lack of self-consistency
between the microscopic and macroscopic contributions [101]. In the mic+mac
models (for example, CNS) polarization effects caused by time-odd fields are ne-
glected and therefore ie f f is predominantly defined by the alignment properties of
the active single-particle orbitals [79]. On the contrary, they play an important role
in the DFT models [54, 55].
Similar studies of relative properties of rotational bands have been performed in
different regions of nuclear chart [60, 100, 101]). For example, a statistical analysis
of significant number of rotational configurations in the A ∼ 130 region of superde-
formation confirms the additivity principle for quadrupole moments and effective
alignments [100]. This justifies the use of an extreme single-particle model in an
unpaired regime typical of high angular momentum. Note that the basic idea behind
the additivity principle for one-body operators is rooted in the independent particle
model.

Conclusions

The concept of independent particle motion is the foundation of the absolute ma-
jority of the nuclear structure models. It leads to verifiable consequences such as
shell structure and individual single-particle properties of nucleons. Both mic+mac
and DFT models are used nowadays for the description of numerous aspects of low
energy nuclear phenomena. The former allows significant flexibility (such as the
calculation of large number of the nucleonic configurations in a single nucleus as it
is done in the CNS approach [13]) but lacks self-consistency. This neglect of self-
consistency limits the applicability of the mic+mac models to ellipsoidal nuclei with
similar density distributions in proton and neutron subsystems. In contrast, the DFT
models are fully self-consistent which makes them applicable to very exotic nuclear
shapes including clustered, ”bubble” and toroidal ones. However, the calculation of
many nucleonic configurations in a given nucleus involving one or several blocked
single-particle states still remains a challenge especially in the formalism which in-
cludes pairing. At present, the use of these two theoretical frameworks should be
considered as complimentary.
The validity of independent particle model has been confirmed by experimental
observation of superheavy nuclei and the phenomenon of superdeformation at high
spin; these observations were motivated by model predictions. They together with
global structure of the nuclear landscape are the consequences of underlying shell
structure, which emerges from independent particle motion of the nucleons in the
nucleus. The energies and alignments of the single-particle states serve as a clear
fingerprint of individual properties of the single-particle orbitals. Signature splitting
properties are of particular value since either their large values or no splittings help
in the identification of the single-particle states involved in the structure of observed
experimental bands. The experimental features seen at extremely high spins where
the pairing correlations are negligible provide the best [at the level of single-particle
34 A. V. Afanasjev

states] examples of independent particle motion; these are found in smooth termi-
nating bands of the A ∼ 110 mass region and in superdeformed bands of the A ∼ 150
mass region.

Acknowledgments

This material is based upon work supported by the U.S. Department of Energy,
Office of Science, Office of Nuclear Physics under Award No. DE-SC0013037. I
would like to thank I. Debes and J. Dudek for the creation of improved quality Fig.
3. Useful discussions with I. Ragnarsson are greatly appreciated.

References

1. A. Bohr and B. R. Mottelson, Nuclear Structure, vol. 1, (W. A. Benjamin, Inc., 1969).
2. P. Ring and P. Schuck, The Nuclear Many-Body Problem (Springer-Verlag, Berlin, 1980)
3. R. F. Casten, Nuclear Structure from a Simple Perspective, (Oxford University Press, 1990).
4. S. G. Nilsson and I. Ragnarsson, Shapes and shells in nuclear structure, (Cambridge Univer-
sity Press, 1995).
5. S. G. Nilsson, Binding states of individual nucleons in strongly deformed nuclei, Mat.-Fys.
Medd.-K.Dan. Vidensk. Selsk. 29, no. 16 (1955).
6. I. Ragnarsson, S. G. Nilsson and R. K. Sheline, Shell structure in nuclei, Phys. Rep. 45, 1
(1978).
7. F. A. Gareev, S. P. Ivanova, V. G. Soloviev and S. I. Fedotov, Single-particle energies and
wave functions of the Saxon-Woods potentials and nonrotational states of odd-mass nuclei in
the region 150 < A < 190, Phys. Elem. Part. and Atom. Nuclei, v. 4., part 2 (1973).
8. S. Cwiok, J. Dudek, W. Nazarewicz, J. Skalski, and T. Werner, Single-particle energies, wave
functions, quadrupole moments and g-factors in an axially deformed Woods-Saxon poten-
tial with applications to the two-centre-type nuclear problems, Comp. Phys. Comm. 46, 379
(1987).
9. R. Bengtsson, J. Dudek, W. Nazarewicz and P.Olanders, A systematic comparison between
the Nilsson and Woods-Saxon deformed shell model potentials, Phys. Scripta 39, 196 (1989).
10. P. Möller and J. R. Nix, Nuclear mass formula with a Yukawa-plus-exponential macroscopic
model and a folded-Yukawa single-particle potential, Nucl. Phys. A 361, 117 (1981).
11. A. Dobrowolski, K. Pomorski and J. Bartel, Solving the eigenvalue problem of the nuclear
Yukawa-folded mean-field Hamiltonian, Comp. Phys. Comm. 199, 118 (2016).
12. R.R. Chasman, I. Ahmad, A.M. Friedman and J.R. Erskine, Survey of single-particle states in
the mass region A > 228, Rev. Mod. Phys. 49, 833 (1977).
13. A.V. Afanasjev, D.B. Fossan, G.J. Lane and I. Ragnarsson, Termination of Rotational Bands:
Disappearance of Quantum Many-body Collectivity, Phys. Rep. 322, 1 (1999).
14. J. Dobaczewski, A.V. Afanasjev, M.Bender, L.M.Robledo, and Yue Shi, Properties of nuclei
in the nobelium region studied within the covariant, Skyrme, and Gogny energy density func-
tionals, Nucl. Phys. A 944, 388 (2015).
15. D. R. Inglis, Particle derivation of nuclear rotation properties associated with a surface wave,
Phys. Rev. 96, 1059 (1954).
16. D.R. Inglis, Nuclear moments of inertia due to nucleon motion in a rotating well, Phys. Rev.
103, 1786 (1956).
17. M.J.A. de Voigt, J. Dudek, Z. Szymański, High-spin phenomena in atomic nuclei, Rev. Mod.
Phys. 55, 949 (1983).
Model for independent particle motion 35

18. Z. Szymański, Fast Nuclear Rotation, (Claredon Press, Oxford, 1983).


19. J. Meng, J. Peng S.-Q.- Zhang, P.-W. Zhao, Progress on tilted axis cranking covariant density
functional theory for nuclear magnetic and antimagnetic rotation, Front. of Phys. 8, 55 (2013).
20. J. Robin, Th. Byrski, G. Duchêne, F. A. Beck, D. Curien, N. Dubray, J. Dudek, A. Gó ź dź,
A. Odahara, N. Schunck, N. Adimi, D. E. Appelbe, P. Bednarczyk, A. Bracco, B. Cederwall,
S. Courtin, D. M. Cullen, O. Dorvaux,S. Ertück, G. de France, B. Gall, P. Joshi, S. L. King,
A. Korichi, K. Lagergren, G. Lo Bianco, S. Leoni, A. Lopez-Martens, S. Lunardi, B. Million,
A. Nourredine, E. Pachoud, E. S. Paul, C. Petrache, I. Piqueras, N. Redon, A. Saltarelli, J.
Simpson, O. Stezowski, R. Venturelli, J. P. Vivien, and K. Zuber, Extended investigation of
superdeformed bands in 151,152 Tb nuclei, Phys. Rev. C 77, 014308 (2008).
21. A.V. Afanasjev and H. Abusara, From cluster structures to nuclear molecules: The role of
nodal structure of the single-particle wave functions, Phys. Rev. C, 97, 024329 (2018).
22. A.V. Afanasjev, Cluster structures, ellipsoidal shapes and nuclear molecules in light A = 12 −
50 nuclei, EPJ Web of Conf. 194, 06001 (2018).
23. K. Karakatsanis, G.A. Lalazissis, P. Ring, and E. Litvinova, Spin-orbit splittings of neutron
states in N = 20 isotones from covariant density functionals and their extensions, Phys. Rev.
C 95, 034318 (2017).
24. A.V. Afanasjev and S. Frauendorf, Central depression in nuclear density and its consequences
for the shell structure of superheavy nuclei, Phys. Rev. C 71, 024308 (2005).
25. T. Ichikawa, J.A.Maruhn, N. Itagaki, and S.Ohkubo, Linear Chain Structure of Four-α Clus-
ters in 16 O, Phys. Rev. Lett. 107, 112501 (2011).
26. J. M. Yao, N. Itagaki, and J. Meng, Searching for a 4α linear-chain structure in excited states
of 16 O with covariant density functional theory, Phys. Rev. C 90, 054307 (2014).
27. P.W. Zhao, N. Itagaki, and J. Meng, Rod-shaped Nuclei at Extreme Spin and Isospin, Phys.
Rev. Lett. 115, 022501 (2015).
28. J.-P. Ebran, E. Khan, R.-D. Lasseri, and D. Vretenar, Single-particle spatial dispersion and
clusters in nuclei, Phys. Rev. C 97, 061301(R) (2018).
29. M. Grasso, L. Gaudefroy, E. Khan, and T. Nikšić, J. Piekarewicz, O. Sorlin, N. Van Giai, and
D. Vretenar, Nuclear “bubble” structure in 34 Si, Phys. Rev. C 79, 034318 (2009).
30. A. Mutschler, A. Lemasson, O. Sorlin, D. Bazin, C. Borcea, R. Borcea, Z. Dombrádi, J.-P.
Ebran, A. Gade, H. Iwasaki, E. Khan, A. Lepailleur, F. Recchia, T. Roger, F. Rotaru, D. Sohler,
M. Stanoiu, S.R. Stroberg, J. A. Tostevin, M. Vandebrouck, D. Weisshaar and K. Wimmer, A
proton density bubble in the doubly magic 34 Si nucleus, Nature Phys. 13, 152 (2017).
31. M. Bender, K.Rutz, P.-G. Reinhard, J.A. Maruhn and W. Greiner, Shell structure of superheavy
nuclei in self-consistent mean-field models, Phys. Rev. C 60, 034304 (1999).
32. A. Sobiczewski and K. Pomorski, Description of structure and properties of superheavy
nuclei, Prog. Part. Nucl. Phys. 58, 292 (2007).
33. B. Schuetrumpf, W. Nazarewicz, W. and P.-G. Reinhard, Central depression in nucleonic den-
sities: Trend analysis in the nuclear density functional theory approach, Phys. Rev. C 96,
024306 (2017).
34. V.M. Strutinsky, Shell effects in nuclear masses and deformation energies, Nucl. Phys. A 95,
420 (1967).
35. S.G. Nilsson, C.F. Tsang, A. Sobiczewski, Z. Szymański, S. Wycech, C. Gustafsson, I.-L.
Lamm, P. Möller, On the nuclear structure and stability of heavy and superheavy elements, B.
Nilsson, Nucl. Phys. A 131, 1 (1969).
36. M. Brack, J. Damgaard, A.S. Jensen, H.C. Pauli, V.M. Strutinsky, C.Y. Wong, Funny Hills:
The shell-correction approach to nuclear shell effects and its application to the fission process,
Rev. Mod. Phys. 44 (1972) 320.
37. D. Vretenar, A.V. Afanasjev, G.A. Lalazissis and P. Ring, Relativistic Hartree-Bogoliubov
Theory: Static and Dynamic Aspects of Exotic Nuclear Structure, Phys. Rep. 409, 101 (2005).
38. K. Pomorski and J. Dudek, Nuclear liquid-drop model and surface-curvature effects, Phys.
Rev. C 67, 443161 (2003).
39. B. G. Carlsson, I. Ragnarsson, R. Bengtsson, E. O. Lieder, R. M. Lieder, and A. A. Pasternak,
Triaxial shape with rotation around the longest principal axis in 142 Gd, Phys. Rev. C78, 034316
(2008).
36 A. V. Afanasjev

40. J. Meng, H. Toki, S. G. Zhou, S. Q. Zhang, W. H. Long and L. S. Geng, Relativistic


continuum Hartree-Bogoliubov theory for ground state properties of exotic nuclei, Prog. Part.
Nucl. Phys. 470 (2006).
41. T. Nikšić, D. Vretenar and P. Ring, Relativistic nuclear energy density functionals: mean-field
and beyond, Prog. Part. Nucl. Phys. 66, 519 (2011).
42. Relativistic Density Functional for Nuclear Structure, (World Scientific Publishing Co), Edited
by Jie Meng, Int. Rev. Nucl. Phys. 10 (2016).
43. M. Bender, P.-H. Heenen and P.-G. Reinhard, Self-consistent mean-field models for nuclear
structure, Rev. Mod. Phys. 75, 121 (2003).
44. S. Peru and M. Martini, Mean field based calculations with the Gogny force: Some theoretical
tools to explore the nuclear structure, Eur. Phys. J A50, 88 (2014).
45. Y. K. Gambhir, P. Ring, and A. Thimet, Relativistic mean field theory for finite nuclei, Ann.
Phys. (NY) 198, 132 (1990).
46. S. Typel and H. H. Wolter, Relativistic mean field calculations with density dependent meson-
nucleon coupling, Nucl. Phys. A656, 331 (1999).
47. G.A. Lalazissis, T.Nikšić, D. Vretenar and P. Ring, New relativistic mean-field interaction
with density-dependent meson-nucleon couplings, Phys. Rev. C 71, 024312 (2005).
48. J. Boguta and R. Bodmer, Relativistic calculation of nuclear matter and the nuclear surface,
Nucl. Phys. A292, 413 (1977).
49. A.V. Afanasjev, P. Ring and J. König, Cranked relativistic Hartree-Bogoliubov theory: formal-
ism and application to the superdeformed bands in the A ∼ 190 region, Nucl. Phys. A676, 196
(2000).
50. A. V. Afanasjev and O. Abdurazakov, Pairing and rotational properties of actinides and
superheavy nuclei in covariant density functional theory, Phys. Rev. C 88, 014320 (2013).
51. W. Koepf and P. Ring, A relativistic description of rotating nuclei: the yrast line of 20 Ne,
Nucl. Phys. A 493, 61 (1989).
52. A.V. Afanasjev, G. Lalazissis and P. Ring, Relativistic mean field theory in rotating frame:
Single-particle properties at superdeformation, Nucl. Phys. A 634, 395 (1998).
53. T. Nikšić and N. Paar and D. Vretenar and P. Ring, DIRHB - a relativistic self-consistent
mean-field framework for atomic nuclei, Comp. Phys. Comm. 185, 1808 (2014).
54. J. Dobaczewski and J. Dudek, Time-odd components in the mean field of rotating superde-
formed nuclei, Phys. Rev. C 52, 1827 (1995).
55. A.V. Afanasjev and P. Ring, Time-odd mean fields in the rotating frame: Microscopic nature
of nuclear magnetism, Phys. Rev. C 62, 031302(R), (2000).
56. A. V. Afanasjev and H. Abusara, Time-odd mean fields in covariant density functional
theory: nonrotating systems, Phys. Rev. C, 81, 014309 (2010).
57. N. Schunck, J. Dobaczewski, J. McDonnell, J. Moré, W. Nazarewicz, J. Sarich and
M. V. Stoitsov, One-quasiparticle states in the nuclear energy density functional theory, Phys.
Rev. C 81, 024316 (2010).
58. Y.R. Shimizu, J.D. Garrett, R. A. Broglia, M. Gallardo and E. Vigezzi, Pairing fluctuations
in rapidly rotating nuclei, Rev. Mod. Phys. 61, 131 (1989).
59. W. Nazarewicz, R. Wyss and A. Johnson, Structure of superdeformed bands in the A ≈ 150
mass region, Nucl. Phys. A 503, 285 (1989).
60. A.V. Afanasjev and S. Frauendorf, Description of rotating N = Z nuclei in terms of isovector
pairing, Phys. Rev. C 71, 064318 (2005).
61. S.E. Agbemava and A.V. Afanasjev, Hyperheavy spherical and toroidal nuclei: The role of
shell structure, Phys. Rev. C 103 034323 (2021).
62. S.E. Agbemava, A.V. Afanasjev, D.Ray and P.Ring, Global performance of covariant energy
density functionals: Ground state observables of even-even nuclei and the estimate of theoret-
ical uncertainties, Phys. Rev. C 89, 054320 (2014)
63. S.E. Agbemava, A.V. Afanasjev, A. Taninah, and A. Gyawali, Extension of the nuclear land-
scape to hyperheavy nuclei, Phys. Rev. C 99, 034316 (2019).
64. P. Möller, A.J. Sierk, T. Ichikawa and H. Sagawa, Nuclear ground-state masses and deforma-
tions: FRDM(2012), Atomic Data and Nuclear Data Tables, 109-110, 1 (2016).
Model for independent particle motion 37

65. J.-P. Delaroche, M. Girod, J. Libert, H. Goutte, S. Hilaire, S. Peru, N. Pillet and G.F. Bertsch,
Structure of even-even nuclei using a mapped collective Hamiltonian and the D1S Gogny
interaction, Phys. Rev. C 81, 014303 (2010).
66. J. Erler, N. Birge, M. Kortelainen, W. Nazarewicz, E. Olsen, A.M. Perhac and M. Stoitsov,
The limits of the nuclear landscape, Nature, 486, 509 (2012).
67. O. Sorlin and M.-G. Porquet, Nuclear magic numbers: New features far from stability, Prog.
Part. Nucl. Phys. 61, 602 (2008).
68. C. Y. Wong, Toroidal and spherical bubble nuclei, Ann. Phys. 77, 279 (1973).
69. A.V. Afanasjev, S.E. Agbemava, D. Ray and P. Ring, Neutron drip line: Single-particle degrees
of freedom and pairing properties as sources of theoretical uncertainties, Phys. Rev. C 91,
014324 (2015).
70. A. Sobiczewski, F.A. Gareev, and B.N. Kalinkin, Closed shells for Z ≥ 82 and N ≥ 126 in a
diffuse potential well, Phys. Lett. 22, 500 (1966).
71. H. Meldner, Predictions of new magic regions and masses for super-heavy nuclei from calcu-
lations with realistic shell model single particle Hamiltonians, Ark. Fys. 36, 593 (1967).
72. Yu.Ts. Oganessian and V.K. Utyonkov, Super-heavy element research, Rep. Prog. Phys. 78,
036301 (2015)
73. S.A. Giuliani, Z. Matheson, W. Nazarewicz, E. Olsen, P.-G. Reinhard, J. Sadhukhan, B.
Schuetrumpf, N. Schunck, and P. Schwerdtfeger, Colloquium: Superheavy elements: Oganes-
son and beyond, Rev. Mod. Phys. 91, 011001 (2019).
74. G.G. Adamian, N.V. Antonenko, H. Lenske, L. A. Malov and Shan-Gui Zhou, Self-consistent
methods for structure and production of heavy and superheavy nuclei, Eur. Phys. J A57, 89
(2021).
75. S.E. Agbemava, A,V. Afanasjev, T. Nakatsukasa, and P. Ring, Covariant density functional
theory: Reexamining the structure of superheavy nuclei, Phys. Rev. C 92, 054310 (2015).
76. Z. Shi, A.V. Afanasjev, Z.P. Li, and J. Meng, Superheavy nuclei in a microscopic collective
Hamiltonian approach: The impact of beyond-mean-field correlations on ground state and
fission properties, Phys. Rev. C 99, 064316 (2019).
77. G. Andersson, S.E. Larsson, G.Leander, P. Möller, S.G. Nilsson, I. Ragnarsson, S. Åberg,
R. Bengtsson, J. Dudek, B. Nerlo-Pomorska, K.Pomorski and Z.Szymański, Nuclear shell
structure at very high angular momentum, Nucl. Phys. A 268, 205 (1976).
78. P.J. Twin, B.M. Nyakó, A.H. Nelson, J. Simpson, M.A. Bentley, H.W. Cranmer-Gordon, P.D.
Forsyth, D. Howe, A.R. Mokhtar, J.D. Morrison, J.F. Sharpey-Schafer and G. Sletten, Obser-
vation of a Discrete-Line Superdeformed Band up to 60h̄ in 152 Dy, Phys. Rev. Lett. 57, 811
(1986).
79. I. Ragnarsson, Orbital and spin assignment of SD bands in the Dy/Gd region — identical
bands, Nucl. Phys. A 557 (1993) 167c.
80. B. Singh, R. Zywina and R.B. Firestone, Table of superdeformed nuclear bands and fission
isomers, Nucl. Data Sheets 97, 241 (2002).
81. O. Häusser, A.J. Ferguson, A.B. Mcdonald, I.M.Szöghy, T.K. Alexander and D.L.Disdier,
High spin states in 20 Ne from 16 O(α, α) scattering, Nucl. Phys. A 179, 465 (1972).
82. A. Watt, D. Kelvin and R. R. Whitehead, Shell-model calculations in the sd shell: X. Termi-
nation of rotational bands, J. Phys. G 6, 35 (1980)
83. I. Ragnarsson, V.P. Janzen, D.B. Fossan, N.C. Schmeing, R. Wadsworth, Smooth termination
of collective rotational bands, Phys. Rev. Lett. 74 (1995) 3935
84. T. Bengtsson and I. Ragnarsson, Rotational bands and particle-hole excitations at very high
spin, Nucl. Phys. A 436, 14 (1985).
85. C. Mahaux, P. F. Bortignon, R. A. Broglia and C. H. Dasso, Dynamics of the shell model,
Phys. Rep. 120, 1 (1985).
86. A.V. Afanasjev and E. Litvinova, Impact of collective vibrations on quasiparticle states of
open-shell odd-mass nuclei and possible interference with the tensor force, Phys. Rev. C 92,
044317 (2015).
87. L.-G. Cao, G. Colò, H. Sagawa and P. F. Bortignon, Properties of single-particle states in a
fully self-consistent particle-vibration coupling approach, Phys. Rev. C 89, 044314 (2014).
38 A. V. Afanasjev

88. B. A. Alikov, K. N. Badalov, V. O. Nesterenko, A. V. Sushkov, and J. Wawryszczuk, On the


role of the Coriolis and quasiparticle-phonon interactions in describing E1 transition proba-
bilitues in odd Eu and Tb isotopes, Z. Phys. A 331, 265 (1988).
89. N. Y. Shirikova, A. V. Sushkov, L. A. Malov, and R. V. Jolos, Structure of the low-lying states
of the odd-neutron nuclei with Z ≈ 100, Eur. Phys. J. A 51, 21 (2015).
90. S. Frauendorf, Beyond the Unified Model, Phys. Scr. 93, 043003 (2018).
91. A.V. Afanasjev and S. Shawaqfeh, Deformed one-quasiparticle states in covariant density
functional theory, Phys. Lett. B 706, 177 (2011).
92. R. Rodriguez-Guzman, P. Sarriguren, and L. M. Robledo, Shape evolution in yttrium and
niobium neutron-rich isotopes, Phys. Rev. C 83, 044307 (2011).
93. Y. Zhang, A. Bjelc̆ić, T. Niks̆ić, E. Litvinova, P. Ring, and P. Schuck, A many-body approach
to superfluid nuclei in axial geometry, submitted to Phys. Rev. C., see also nuclear theory
arXiv:2104.14513.
94. R. Bengtsson and S. Frauendorf, Quasiparticle spectra near the yrast line, Nucl. Phys. A 327,
139 (1979).
95. H. Molique, J. Dobaczewski and J. Dudek, Superdeformed bands in 32 S and neighboring nu-
clei predicted within the Hartree-Fock method, Phys. Rev. C 61, 044304 (2000).
96. J.L. Egido and L.M. Robledo, High spin properties of the Er and Yb isotopes with the Gogny
force, Nucl. Phys. A 570, 69 (1994).
97. T. Duguet, P. Bonche and P.-H. Heenen, Rotational properties of 252,253,254 No: influence of
pairing correlations, Nucl. Phys. A 679, 427 (2001).
98. Z.-H. Zhang, M. Huang and A.V. Afanasjev, Rotational excitations in rare-earth nuclei: A
comparative study within three cranking models with different mean fields and treatments of
pairing correlations. Phys. Rev. C 101, 054303 (2020).
99. W. Satuła, J. Dobaczewski, J. Dudek, and W. Nazarewicz, Additivity of quadrupole moments
in superdeformed bands: single-particle motion at extreme conditions. Phys. Rev. Lett. 77
(1996) 5182.
100. M. Matev, A. V. Afanasjev, J. Dobaczewski, G. A. Lalazissis, and W. Nazarewicz, Additivity
of effective quadrupole moments and angular momentum alignments in A ∼ 130 nuclei, Phys.
Rev. C 76, 034304 (2007).
101. L.B. Karlsson, I. Ragnarsson and S. Åberg, Polarization effects in superdeformed nuclei,
Nucl. Phys. 639, 654 (1998).
102. A.V. Afanasjev and P. Ring, Superdeformations in relativistic and non-relativistic mean field
theories, Phys. Scripta. T 88, 10 (2000).

You might also like