Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Theoretical nuclear structure

Josipa Diklić
jdiklic@irb.hr
Ruđer Bošković institute
July 31, 2020

1 Shell Model
1.1 Motivation for the Shell Model
It is known that nuclei with even numbers of protons and neutrons are more stable than those with odd.
In particular, there are "magic numbers" of neutrons and protons which seem to be particularly favored
in terms of nuclear stability: 2, 8, 20, 28, 50, 82, 126... "Doubly magic" nuclei, which have both neutron
number and proton number equal to one of the magic numbers, are especially stable. Such stability
is shown in very long half-life, higher binding energy (compared to the other nuclei), high first excited
state... The existence of these magic numbers suggests closed shell configurations, like the shells in atomic
structure, so they are one of the reasons for the development the shell model of atomic nuclei.
There is also example experimental evidence confirming the shell structure, and some of them will be
presented in this report.

Abundance of elements for which Z or N is a magic number


For highly stable nuclei, (doubly) magic nuclei, we expect the smallest cross-sections for the kind of scat-
tering which would disrupt them so we expect to have an large abundance of such nuclei.
In figure 1 we can observe peaks at (doubly) magic nuclei. Such occurrence is slightly less visible due
to the high abundance of iron, which doesn’t have magic number of Z or N, but it is most stable isotope
due to the details of how nucleosynthesis works. The most abundant 56 Fe is a more common endpoint of
fusion chains inside the extremely massive stars and is therefore more common in the universe, relative
to other metals.

"Magic Number" Nuclei at the End of Radioactive Series


The end points of all four of the natural radioactive series (232 Th... →208 Pb, 237 Np...→209 Bi, 238 U...→206 Pb,
235
U...→207 Pb) are stable heavy nuclei of Pb which have magic numbers of either N or Z. This nicely
illustrates magic number properties.

Neutron Absorption Cross-sections


In the following figure (2) we can notice much lower cross-section (probability) for the absorption of
neutrons by nuclei with magic number of neutrons. The reason for this is, again, the high stability of such
nuclei.

Binding Energy for the Last Neutron


The dependence of the binding energy (relative to the Weizsaeker formula) of the last neutron on the
neutron number is shown in the figure 3. We can observe high binding energies for nuclei with magic
numbers of neutrons, so those nuclei are less likely to loose nucleons by neutron bombardment.

1
Figure 1: Relative abundance vs. mass number of element in logarithmic scale.[4]

Electric Quadrupole Moments of Nuclei


The shell model for the closed shell predicts that the nuclear charge is spherically symmetric. If a nucleus
is not spherically symmetric, it will have a non-zero electric quadrupole moment, thus the measurement of
the quadrupole moment is one of the validation test for the shell theory. The reduced electric quadrupole
moment vs. odd number of nucleons is shown in figure 4. One can observe near-zero quadrupole moments
for magic nuclei.

1.2 Derivation of the shell model


The basic idea of the nuclear shell model theory is to assume independent motion of each nucleon in an
average nuclear potential.
A phenomenological shell model is based on the Schrodinger equation for the single-particle levels i
with eigenstates ψi (r) and eigenvalues εi :

~2 2
 
Ĥψi = − ∇ + V̂ (r) ψi (r) = εi ψi (r), (1)
2m

where ~ is a reduced Planck’s constant, m is a mass of the nucleon and V (r) is a averge nuclear potential.
The average nuclear potential V (r) is constructed with the several assumptions.
In the first approximation it is spherically symmetrical and with this model only doubly magical nuclei
with one nucleon more or one nucleon less can be described.
The potential averages nucleon-nucleon interaction for all pairs, as the name implies and the total force
at the center of the nuclei vanish ∂V

∂r r=0 = 0.
The strength of the force increases towards the center of the nuclei ∂V

∂r r<R0 > 0, where R0 is the
nuclei radius.
Naturally, the effective nuclear force has a finite range V (r) ≈ 0, r > R0 .
There are several functions, such as the square well, Wood-Saxon and the harmonic oscillator, which
are commonly used and approximate the potential reasonably well (figure 5).
Although the Wood-Saxon is the more realistic nuclear single-particle potential, I will derive the result
for a harmonic oscillator since it has the analytic solution.
If the potential energy and the boundary conditions are spherically symmetric, it is useful to transform
the Hamiltonian H into spherical coordinates and seek solutions to Schrodinger’s equation which can
be written as the product of a radial portion and an angular portion: ψ(r, θ, φ) = R(r)Y (θ, φ), or even
R(r)Θ(θ)Φ(φ).

2
Figure 2: Cross section for neutron absorption vs. neutron number.[4]

Figure 3: Binding energy (relative to the Weizsaeker formula) of the last neutron as a function of neutron
number.[4]

After transforming Laplacian in spherical coordinates and manipulations, the equations for the factors
become:
d2 Φ 2
dφ2 = −m Φ,

d 2
sin θ dΘ
 2
sin θ dθ dθ + l(l + 1) sin θΘ = m Θ,
, (2)

d 2mr 2
r2 dR

dr dr − ~2 [V (r) − E]R = l(l + 1)R
where m2 and l(l + 1) are constants of separation.
The product of solutions to the angular equations is:
 1/2
m (2l + 1) (l − |m|)!
Yl (θ, φ) =  eimφ Plm (cos θ). (3)
4π (l + |m|)!
where  = (−1)m for m ≥ h0 and  =i 1 for m ≤ 0, and the spherical harmonics are orthonormal:
Rπ R 2π ∗ 0

0
dθ sin θ 0 dφ [Ylm (θ, φ)] Ylm
0 (θ, φ) = δll0 δmm0 .
The angular part of the wave-function are spherical harmonics, Ylm (θ, φ) for all spherically symmetric
situations, so we are interested in the radial part.
By substituting u(r) = rR(r) in the radial equation (2) we obtain:
~2 d2 u(r) ~2 l(l + 1)
 
− + V (r) + u = Eu(r). (4)
2m dr2 2m r2

3
Figure 4: Reduced electric quadrupole moment vs. odd number of nucleons.[4]

Figure 5: Sketch of the functional form of three popular phenomenological shell-model potentials: Woods-
Saxon, harmonic oscillator, and the square well.

Given the boundary conditions for bound states (E < 0), u(0) = 0 and u(r) limr→∞ = 0 we obtain the
solutions for eigenstates to equation 4 for the Harmonic oscillator potential (V (r) = 21 mw2 r2 ):
2 l+1/2
ukl (r) = Nkl rl+1 e−νr Lk 2νr2 ,

(5)
l+1/2
where ν = mω/2~ and Lk are Laguerre polynomials.
The solutions for energies are:
   
3 3
E = ~ω 2k + l + = ~ω N + , (6)
2 2
where N , l, k are quantum numbers:

major oscillator quantum number N = 0, 1, 2 . . .


orbital quantum number l = N, N − 2, . . . , 1 or 0
(7)
radial quantum number k = (N − 1)/2
number of nodes of the radial wave function in the interval n = k + 1 = (N − l)/2
Due to possible ml = −l, −l + 1, ...l − 1, l projections, states of a certain l have 2l + 1 degeneracy. If
we want to calculate number of nucleons in the oscillatory shell N , we need to sum the degeneracy in all
possible l and multiple by spin factor 2 :
N
X
DN = 2 (2l + 1) = (N + 1)(N + 2). (8)
l=0or1

4
The total number of nucleons to Nmax shell is then:
N max
X 1
Dmax = DN = (Nmax + 1) (Nmax + 2) (Nmax + 3) . (9)
3
N =0

Nmax = 0, 1, 2, . . . ⇒ Dmax = 2, 8, 20, 40, 70, 112, 168, . . . (10)


Nuclei with Dmax number of neutrons and protons are expected to be (doubly) magical. The empirically
obtained magic numbers (2, 8, 20, 28, 50, 82, 126, ...) correspond to the first three calculated in this way,
but it was necessary to extend the theory to obtain a good match with empirical magic number for larger
number of protons and neutrons.
This was achieved by adding a spin-orbit term in the average single-nucleon potential,

h = h0 + ζ(r)~l · ~s. (11)


The spin-orbit part can be treated as a small perturbation, so the energy can then be written as a sum
of the non-perturbed energy and a small perturbation:
(0)
εnlj = εnlj + ∆εnlj , (12)

where:
(0)
εnlj = hnljm |h0 | nljmi
(13)
∆εnlj = hnljm|ζ~l · ~s|nljmi = D
j(j + 1) − l(l + 1) − 34 .
 
2
(0)
Due to ~j = ~l + ~s → j = l ± 1/2 there is a splitting of energy levels εnlj .
D
∆εnlj=l+1/2 = 2l
(14)
∆εnlj=l−1/2 = − D
2 (l + 1)

The splitting of different shells is shown in figure 6. For states with a certain orbital quantum number
l, the spectroscopic notation is used (l = 0, 1, 2, 3, 4, 5, 6.. = s, p, d, f, g, h, i...). We can observe a great
overlap of magical number whit empirical ones (2, 8, 20, 28, 50, 82, 126, ...).
The shell model with a single-nucleon potential as in equation 11 describes spherically symmetric nuclei
with neutron and proton number around magical numbers very well. However, to describe non-spherical
nuclei some modifications need to be introduced. For example, Nillson’s model describes nuclei in the range
150 < A < 190 and A > 230 well. The Nilsson model is a nuclear shell model treating the atomic nucleus
as a deformed sphere. In this model there is an additional part in the single nucleon Hamiltonian ∼ l2 and
the average nuclear potential V (R, θ, φ) depends also on θ and φ coordinate (in spherical coordinate), so
there is a different frequency in different spatial directions. The l2 term is introduced phenomenologically
to lower the energy of the single-particle states closer to the nuclear surface in order to correct for the
steep rise in the harmonic-oscillator potential. This model can than be described with the following single-
particle Hamiltonian:

h = h0 + C~l · ~s − Dl2 ,
(15)
−~2 2
+ 21 m ωx2 + ωy2 + ωz2 .

h0 = 2m ∇

Therefore, there is a dependence of energy levels on the deformation of the nucleus (figure 7).

5
Figure 6: Energy levels with and without spin-orbit splitting.[4]

Figure 7: Nilsson diagram; the energies of valence orbitals as a function of the deformation parameter
2 .[5]

6
1.3 Shell model in experimental research
To substantiate the experimental results, we used the NATHAN code to calculate the theoretical shell struc-
ture. The code NATHAN [7] is written in 1999. at the Institute in Strasbourg. Since then, it evolved and
the sd-pf residual interaction [6] has been included and developed. The code itself solves the Schrödinger
equation as described in the previous section and uses the Lancez method for diagonalization in a multi-
particle base. This method has been developed for the diagonalization of large matrices where the dimen-
sion exceeds d > 200.
For my master thesis[8] we analyzed and interpreted data related to the electromagnetic transitions
excited by the nuclear reaction 40 Ar + 208 Pb measured with the PRISMA magnetic spectrometer coupled
to the CLARA γ-ray detector. As an example, I will elaborate on experimental and theoretical results for
the 41 K isotope (figure 8). On the left experimental results for energy levels with corresponding observed
gamma rays are shown. The arrow thickness represents the corresponding intensity. On the right the
theoretical results for energy levels, spins and parity are shown. The probability of certain components of
the wave function for proton and neutron valence space on certain energy level is given in table 1. We can
observe good agreement of experimental and theoretical results for levels with positive parity.
The proton valence space (which is used in my thesis) includes a sd shell with a range of elements
8 ≤ Z ≤ 20, and the neutron valence space includes the sd and pf shells with a neutron number 8 ≤ N ≤
40. In the case of potassium isotopes, which were studied, there are 19 protons and 21-27 neutrons.
The NATHAN code calculates the energies, spins, parities, and contribution of the components of the
wave functions for a particular isotope. The states of the same spin can occur at most tree times, and the
calculation is done for maximum excitation energies up to ≈ 3 MeV. Thus, for even-odd nuclei, only the
state of negative parity can be calculated, while for odd-odd nuclei only positive parity can be calculated.
An intruder state that would have opposite parity cannot be found within given valence spaces, because
their configuration goes beyond the scope of those spaces. Therefore, although we observe such states in
the experiment, they are not predicted by this model, so we do not have theoretical results for such states.

13/2 + 2775 11/2 2762


13/2 + 2734.05
247 5/2 + 2621.06
11/2 + 2528 11/2 + 2526.3
9/2 + 2494.9

851
3/2 + 1978.01
1471
7/2 + 1677 5/2 + 1663.68
3/2 1586
7/2 + 1469.55
(7/2) 1291

1/2 + 981
1677 1/2 + 854.63
1586
981 1291

3/2 + 0 3/2 + 0

Figure 8: Energy level sheme for 41 K isotope.[8]

7
Figure 9: Occupancy scheme of the 41 K isotope levels in the proton (π) and neutron (ν) valence space.[8]

A Jiπ Ei (keV) wave function component contribution (≥10%)


41 3/2+ 0 π(d5/2 )6 (s1/2 )2 (d3/2 )3 ⊗ ν(f7/2 )2 89
+ 6 2 3 2
1/2 854.63 π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 73
π(d5/2 )6 (s1/2 )1 (d3/2 )4 ⊗ ν(f7/2 )2 24
+ 6 2 3 2
7/2 1469.55 π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 92
+ 6 2 3 2
5/2 1663.68 π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 79
6 2 3 1 1
π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) (p3/2 ) 12
3/2+
2 1978.01 6 2 3
π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 2
83
+ 6 2 3 2
9/2 2454.87 π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 83
6 2 3 1 1
π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) (p3/2 ) 12
+ 6 2 3 2
11/2 2526.3 π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 94
5/2+
2 2621.06 6 2 3
π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 2
66
6 1 4 2
π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 15
+ 6 2 3 2
13/2 2734.05 π(d5/2 ) (s1/2 ) (d3/2 ) ⊗ ν(f7/2 ) 91

Table 1: The results of the shell model calculations for the isotope 41 K. The spins and parities of the excited
energy levels (J ipi ), the nucleon configuration and the share of a certain component of the wave function
are listed.[8]

8
2 Pairing Correlations
2.1 Observation of the Pairing Correlations
The existence of pairing correlations in nuclei was suggested more than 50 years ago by Bohr, Mottelson
and Pines. Pairing correlations in nuclei are phenomena where two nucleons are tightly bound. So it is
assumed that pairs of nucleons are strongly bound, thus for example we expect that to energy required to
separate pair of neutron is smaller the twice separation energy of one neutron because we first need the
energy to break the pair. I will describe some of experimental results in which pairing correlations were
shown to exist.

"Odd-even staggering"
Pairing phenomena can be observed by comparing the binding energy in the isotope chain. Thus, if paring
correlations exist stronger binding is expected in nuclei with even number of protons or neutrons. There-
fore, for odd mass numbers A, we expect the binding energy to be Ebind (A) < 12 (Ebind (A − 1) + Ebind (A + 1)).
In figure 10, experimental evidence of the phenomena is shown[10]. The dependence of the neutron sep-
aration energy (Bn ) on the neutron number N for different isotopes suggests pairing correlations in nuclei
with even neutron numbers,

Bn (N, Z) = M (N − 1, Z) + mn − M (N, Z)
(16)
= Ebind (N, Z) − Ebind (N − 1, Z).

Figure 10: Dependence of the neutron separation energy (Bn ) on the neutron number N in Ca, Sn, and
Pb isotopic chain.[9][10]

Ground state of even-even nuclei: J P = 0+


The tendency of particles with spin s = 1/2 to form bound pairs with total momentum J = 0 is observed
in experiments where the ground states of more than 800 known even–even nuclei have total momentum
J P = 0+ .

Energy gap
In even-even nuclei there is an energy gap between the ground and first excited state around 1-2 MeV,
while for the double magic nuclei like 40 Ca is even larger than 3 MeV, which is much higher than for
even-odd nuclei. For example (figure 11) the first excited state in 42 Ca and 44 Ca is > 1 MeV while the first
excited state in 43 Ca and 45 Ca is <0.4 MeV. These phenomena can be explained if we take into account
that the bond between paired nucleons has to be broken up in order to form the first excited state.

9
Figure 11: Spectra of the excited states of 42−45 Ca isotopes.[9]

Moment of inertia
Moments of inertia for for even-even nuclei are shown in figure 5. We can observe that these moments
calculated for even–even nuclei, applying the independent particle model, are 2– 3 times higher than the
experimental values. The calculation with pairing contribution included comes into close agreement with
the experimental result. If the pairing was strong the system would be a super fluid having irrotational
flow and corresponding inertial dynamics.

Figure 12: Comparison of experimental results and theoretical calculations of the moment of inertia for
different even-even nuclei.[11]

Low-lying 2+ states
If 2+ can be described with vibrational or rotational degrees of freedom, a characteristic spectrum for these
two cases would be expected. For rotational degrees of freedom the excitation energy is proportional to
∼ J(J +1), so the ratio of the first two excited energies is expected to be: E(4+ )/E(2+ ) = (4×5)/(2×3) =
3.33. For the vibrational spectrum the expected ratio is E(4+ )/E(2+ ) ≈ 2.
Figure 13 shows a chart of the ratio of energies of the first two excited states E(4+)/E(2+) as a function
of Z and N. As we can observe there are dark regions (ratio< 2) around nuclei with magical numbers of
neutrons and protons which suggests that the 2+ state can’t be described by vibrational or rotational
motion.
This phenomenon can also be described with the pairing of identical nucleons in the outer shell.

10
Figure 13: A chart of the ratio of energies of the first two excited states E(4+)/E(2+) as a function of Z
and N.[9]

2.2 Theory of pairing correlation (BCS)


Theoretically, nuclear paring can be described with the BCS theory, which was first introduced to describe
electron pairing in metals. BCS theory or Bardeen–Cooper–Schrieffer theory (named after John Bardeen,
Leon Cooper, and John Robert Schrieffer (1957)) is the first microscopic theory of superconductivity.
The theory describes superconductivity as a microscopic effect caused by a condensation of Cooper pairs
(electrons). Leter it was adopted for nuclei by Bohr, Mottelson, Pines (1958), Belyaev(1959),Nilsson, Prior
(1960), ... [13]
Neutrons and protons just as electrons have spin 1/2, so they are fermions, but the total spin of a Cooper
pair is integer (0 or 1) making it a composite boson. Consequently the wave functions are symmetric under
particle interchange. Therefore, like electrons, neutrons and protons multiple Cooper pairs are allowed to
be in the same quantum state, which is responsible for the phenomena of superconductivity.
In BSC theory[12], the wave function for even-even nuclei is given by:
Y
uk + vk a+ +

|BCSi = k ak |−i. (17)
k>0

For each state k > 0 in the configuration space there exist a "conjugate" state k̄ < 0, and (k, k̄) generate
the whole single-particle space. Parameters vk2 and u2k represent the probability of occupation of certain
pairs (k, k̄).
The product can be written as:
X vk 1 X vk c k 0 + + + +
[BCSi ∝ |−i + a+ +
k ak̄ |−i + a a a 0 a |−i + · · · (18)
uk 2 0 uk uk0 k k̄ k k̄0
k>0 kk >0

Normalization requires:
Y 2 2
hBCS|BCSi = (u2k + vk2 ) ⇒ |uk | + |vk | = 1. (19)
k>0

The parameters u and v in the wave function are determined by variation of energy, but there is another
condition that expected value of the particle number must be N .
X
hBCS|N̂ |BCSi = BCS a+ +
k ak + ak̄ ak̄ BCS = 2 vk2 = N (20)
k>0

11
D E
Particle number uncertainty (∆N )2 = BCS N̂ 2 BCS − N 2 = 4 k>0 u2k vk2 arises from those
P

single-particle states that are fractionally occupied (u2k 6= 0, 1, vk2 6= 0, 1).


In order to keep the expectation value of the particle number to the desired particle number, the
parameter λ is defined and −λN̂ is added to varational Hamiltonian,
0
H = H − λN̂

where the first term is, (21)


+ 1
v̄k1 ,k2 k3 ,k4 a+ +
P P
H= k1 k2 ≤0 tk1 k2 ak1 ak2 + 4 k1 k2 k3 k4 k1 ak2 ak4 ak3 .
The Langrange multiplier λ is also called the chemical potential or the Fermi energy because it relates
the energy and particle number,

dE dhBCS|H|BCSi
= λ= (22)
dN dN
0
From eq. 17 and eq. 21 we can find the expectation value of H to be:
 
X 1 X  X
hBCS |H 0 | BCSi = (tkk − λ) vk2 + v̄kk0 kk · vk2 vk20 + v̄kk̄k0 k̄0 uk vk uk0 vk0 (23)
 2 0  0
k k >0 k,k >0

The parameter vk determines the BCS wave function completely, so we can express uk in terms of vk
by the normalization condition (eq. 19),

δhBCS|Ĥ|BCSi = 0,
  (24)
∂ vk ∂ 0
∂vk − uk ∂uk hBCS |H | BCSi = 0.
After differentiating, we obtain the set of BCS equations:

k uk vk + ∆k vk2 − u2k = 0,
2˜ k>0

where,
(25)
1

(v̄k̄k0 k̄k0 + v̄kk0 kk0 ) vk20 − λ,
P
˜k = 2 tkk + tk̄k̄ + k0
P
and the gap parameter is ∆k = − k0 >0 v̄kk̄k0 k̄0 uk0 vk0 .
These equations can be solved for some special cases such as when ˜k and ∆k have fixed values, but in
general these nonlinear equations need to be solved iteratively.
If we assume a constant residual interaction, which acts only on pairs of nucleons, v̄k1 ,k2 k3 ,k4 in eq. 21
we obtain the following Hamiltonian:
X X
H= εk a+
k ak − G a+ +
k ak̄ ak̄0 ak .
0 (26)
k>0 kk0 >0
0
The expectation value of H is:

0
(εk − λ) a+ a+ +
P P
hBCS|H |BCSi = hBCS|H − λN̂ |BCSi = BCS k k ak − G kk0 k ak̄ ak̄0 ak BCS =
0

 2
X
vk2 vk4
P P 
2 k>0 (εk − λ) − k>0 G −G uk vk  =
|{z} |{z}  | {z }  ,
k>0
hBCS |a+
k ak |BCS i hBCS |a+
k a−k a−k ak |BCS i
+
hBCS |ak a−k a−k ak |BCS i
+ +

| {z }

∆2
˜k vk2 + 12 Gvk4 −
P 
2 k>0 G
(27)

12
where,
X
∆=G uk vk and ˜k = k − λ − Gvk2 . (28)
k>0

The terms in eq. 27 and 28 Gvk4 and Gvk2 are often neglected, so we have:
 
2

uk 1 k − λ
= 1 ± q (29)
vk2

2 2
(k − λ) + ∆2
From eq. 29 we can observe that if there is no pairing (G → 0), energy gap vanish (∆ = 0) and
probability for occupied levels vk2 = 1 and for unoccupied ones is u2k = 0, thus vk is a step function. If
pairing exist (G 6= 0) energy gap don’t vanish (∆ 6= 0) and vk have distributional form (figure ??).
The gap equation takes following form:
2 X 1
= p (30)
G (ek − λ)2 + ∆2
k>0

Figure 14: The occupation probability for interacting and non-interacting case.[12]

In figure 15 distributions of nucleons (2 ≤ n ≥26) over the five orbitals are shown. We can observe
that with increasing number of nucleons distribution function becomes less step-function like.

Figure 15: The occupation probability for different number of nucleon.[11]

In special case where there is a single j-shell all k are the same thus all vk are the same as well. From
particle-number condition (eq. 20) we got:

13
r r
X N N
hBCS|N̂ |BCSi = 2 vk2 = N ⇒ vk = , uk = 1−
2Ω 2Ω
k>0
s   (31)
N N
∆=G· Ω− ,
2 2
where Ω is the maximum number of of pairs in one shell.
The binding energy of the pairs (≈ 2∆) is maximum when the orbital is half filled (N = Ω). In this
case,
2∆ = GΩ. (32)
In more complicated case where are n-particles with seniority ν, in the shell j the general expression
for the spectrum of the pairing Hamiltonian is given:

G
Ev (n) − E0 (n) = ν(2Ω − ν + 2) (33)
4
Where, the seniority quantum number ν represent number of unpaired nucleons. We can observe that
energy difference is independent of n. In figure 16 spectrum for a pure pairing force (G set to 0.25)
within the h11/2 orbital with seniority ν = 0, ν = 2, ν = 4, and ν = 6 as a function of the particle number
n is shown.

Figure 16: Spectrum for a pure pairing force (G set to 0.25) within the h11/2 orbital with seniority ν = 0,
ν = 2, ν = 4, and ν = 6 as a function of the particle number n.

14
2.3 Pairing in experimental research of heavy ion collisions
How pairing correlations can be probed in heavy ion collisions, is still an open question. Several experi-
ments have been performed in the past, searching for signatures mainly via the extraction of enhancement
coefficients [15], defined as the ratio of the actual transfer cross-section to the prediction of models using
uncorrelated states. Unfortunately the experimental evidence of these factors are marred by the fact that
almost all existing studies involve inclusive cross-sections at energies higher than the Coulomb barrier and
at angles forward of the grazing [16] where the reaction mechanism is complicated by the interplay be-
tween nuclear and Coulomb trajectories. Most data collected on transfer reactions measured above the
barrier is represented via the transfer probability
dσtr
Ptr = . (34)
dσel
(defined as ratio of the transfer yield over the quasi-elastic one) are expressed as a function of the distance
of closest approach (d),
Zt Zp e2 1
d= (1 + ), (35)
2Ec.m. sin(θc.m. /2)
(where Zt and Zp are atomic numbers of target and projectile, while Ecm and sin(θcm ) are the beam
energy and angle of detection), however quite contradictory conclusions have been deduced.
In the 116 Sn+60 Ni system[17] (figure 17), for the first time in a heavy ion collision, we have been able to
provide a consistent description of one and two neutron transfer reactions by incorporating, in the reaction
mechanism, all known structure information of entrance and exit channels nuclei, without needing to
introduce any enhancement factor for the description of two neutron transfers. The achievement was
possible only because the chosen system is very well Q-value matched so that the reaction was dominated
by the ground-ground state transition. An interesting and almost unexplored issue is whether and to what
extent the effect of nucleon-nucleon correlations in the evolution of the reaction is modified in the presence
of high Coulomb fields. In fact, in the collision amongst very heavy ions, the population of final states with
high excitation and angular momenta and/or multi-step processes may significantly change the transfer
strength of the ground to ground state transitions.

Figure 17: The (+1n) and (+2n) channels in 116 Sn+60 Ni. Where the
pfitting function for fit is: fun(d) =
factor × exp(−2 · α·d). The expected value of parameter α is: α = 2mB/~2 , where B is defined with
form factor B = -Sn , and m = 939.565 MeV/c2 , ~c = 197.327 MeV fm, e2 = 1.44 are constants.[17]

After analysing the 116 Sn+60 Ni system,a new experiment involving 209 Pb+118 Sn, on which I am cur-
rently working on, was proposed. Measurement of Ptr for neutron transfer channels in a wide range of d,

15
from ∼ 15 fm up to ∼ 19-20 fm were performed. In this way one can quantitatively probe the effect of
neutron-neutron correlations by comparing the experimental Ptr for the odd and even neutron transfers
to microscopic calculations, using as a reference the previously studied system.
To substantiate the experimental results, we use a semi-classical GRAZING model which was developed
by A.Winther and Copenhagen group. [18][19] Interacting forces of two ions, Coulomb and nuclear forces
are described by two-point charge expression and Wood-Saxon parametrization respectively. The exchange
of many nucleons is obtained by a multi-step mechanism for single-particle nucleons. The two nuclei are
treated as an ensemble of independent nucleons that can vibrate around their spherical equilibrium shapes,
where the basic degrees of freedom are surface vibrations and single-particle transfer where the lowest
2+ and 3− states and associated giant resonances are included. For the excitation of the surface mods,
the model uses the macroscopic approximation whose form factors strength is given by the experimental
B(Eλ). The exchange of nucleons is governed by microscopic form factors which take into account the
single-particle properties of the projectile and target and an average single-particle level density. Therefore,
the approximation in the treatment of particle transfer, is the use of representative form factors for the
transfer and the substitution of the actual distribution of single-particle states with a density function.
As an example, in figure 18 the GRAZING [20] code calculations for the differential cross sections at
Elab =1200 MeV for the +1n channel (for experiment 209 Pb+118 Sn) as function of both the distance of
closest approach and laboratory angle are shown. The plot on the left side represents the corresponding
transfer probability. At lower bombarding energies the angular distributions will flatten out with maximal
moving at θlab =0◦ (i.e. θcm =180◦ ).
Once I complete the analysis, I will compare the results with the theoretically obtained one, shown
here.

Figure 18: Right : GRAZING code calculations of transfer probability. Left: GRAZING code calculations
for the differential cross sections for the +1n channel (as function of the distance of closest approach (top)
and corresponding laboratory angle θlab (bottom)) at Elab =1200 MeV.

16
References
[1] Samuel S. M. Wong, Introductory Nuclear Physics, Wiley-Interscience, 1999.

[2] Krane, Kenneth S., Introductory Nuclear Physics, Wiley, 1987.


[3] Booth, C. N. and Combley, F. H., Nuclear Physics 303, http://www.shef.ac.uk/uni/academic/N-Q/
phys/teaching/phy303/phy303-3.html.
[4] HyperPhysics web-page http://hyperphysics.phy-astr.gsu.edu/hbase/Nuclear/shell.html.

[5] Thummerer, Severin & Oertzen, W. & Gebauer, B. & Lenzi, S. & Gadea, A. & Napoli, D. & Beck, C. &
Rousseau, Michèle. (2001). The population of deformed bands in 48 Cr by emission of 8 Be from the
32
S + 24 Mg reaction.
[6] Nowacki, F., and A. Poves. "New effective interaction for 0 ~ω shell-model calculations in the sd-pf
valence space." Physical Review C 79.1 (2009): 014310.

[7] Caurier, Etienne, and Frederic Nowacki. "Present status of shell model techniques." Acta Physica
Polonica B 30 (1999): 705.
[8] Diklić, J. (2019). Neutronski bogate jezgre kalija (Master thesis) https://urn.nsk.hr/urn:nbn:hr:
217:177748.
[9] Ishkhanov, B. & Stepanov, M. & Tretyakova, Tatiana. (2013). Nucleon pairing in atomic nuclei. Moscow
University Physics Bulletin. 69. 10.3103/S0027134914010068.
[10] G. Audi, M. Wang, A.H. Wapstra, F.G. Kondev,M. MacCormick, X. Xu, and B. Pfeiffer, Chinese.Phys.
C 36, p. 1287; M. Wang, G. Audi, A. H. Wapstra,F. G. Kondev, M. MacCormick, X. Xu, and B. Pfeif-
fer,Chinese Phys. C 36, 1603 (2012).

[11] http://www.phys.ens.fr/cours/Sandro/04-04-05/richter.pdf.
[12] Ring, P. and Schuck, P. - The Nuclear Many-Body Problem-Springer (2004).
[13] A. Bohr, B. R. Mottelson, and D. Pines. Possible analogy between the excitation spectra of nuclei and
those of the superconducting metallic state. Physical Review, 110:936, May 1958.

[14] Corradi, L., Pollarolo, G. & Szilner, S. (2009) Multinucleon transfer processes in heavy-ion reactions.
Journal of Physics G - Nuclear and Particle Physics, 36 (11), 113101-1.
[15] W. von Oertzen et al., Z. Phys. A 326, 463 (1987).
[16] W. von Oertzen and A. Vitturi, Rep. Prog. Phys. 64, 1247 (2001).

[17] D. Montanari et al., Phys. Rev. Lett. 113, 052501 (2014).


[18] A. Winther. Grazing reactions in collisions between heavy nuclei. Nuclear Physics A, 572:191–235,
May 1994.
[19] A. Winther. Dissipation, polarization and fluctuation in grazing heavy-ion collisions and the boundary
to the chaotic regime. Nuclear Physics A, 594:203–245, November 1995.
[20] GRAZING code http://personalpages.to.infn.it/~nanni/grazing/.

17

You might also like