09.03.2015 Views

Nuclear Models - Nuclear Physics

Nuclear Models - Nuclear Physics

Nuclear Models - Nuclear Physics

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Nuclear</strong> <strong>Models</strong><br />

Postgraduate Course following on<br />

from PHYS490<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 1


<strong>Nuclear</strong> <strong>Models</strong><br />

1. Introduction<br />

2. Spherical Shell Model<br />

3. <strong>Nuclear</strong> Deformation<br />

4. Vibrational Motion<br />

5. Collective <strong>Nuclear</strong> Rotation<br />

6. <strong>Nuclear</strong> Pairing<br />

7. Cranked Shell Model<br />

8. Strutinsky Shell Correction<br />

9. Broken Symmetries<br />

10. Band Termination<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 2


1. Introduction<br />

The Unique Nucleus<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 3


<strong>Nuclear</strong> <strong>Physics</strong><br />

• The nucleus is one of nature’s most interesting quantal<br />

few-body systems<br />

• It brings together many types of behaviour, almost all of<br />

which are found in other systems but which in nuclei<br />

interact with one another<br />

• The major elementary excitations in nuclei can be<br />

associated with single-particle or collective modes<br />

• While these modes can exist in isolation, it is the<br />

interaction between them that gives nuclear spectroscopy<br />

py<br />

its rich diversity<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 4


The Nucleus is Unique!<br />

• The nucleus is a unique ensemble of strongly interacting<br />

fermions (nucleons)<br />

• Its large, yet finite, number of constituents controls<br />

the physics of this mesoscopic system<br />

• Both single-particle (out-of-phase) and collective (in-<br />

phase) effects occur<br />

• There is an analogy to a herd of wild animals. Individual<br />

id animals may break out of the herd but are rapidly<br />

drawn back to the safety of the collective<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 5


Generation of Angular Momentum<br />

• There are two basic ways<br />

of generating high-spin<br />

states in a nucleus<br />

1. Collective (in-phase)<br />

motions of the nucleons:<br />

vibrations, rotations etc<br />

2. Single-particle effects:<br />

pair breaking, particlehole<br />

excitations. The<br />

individual spins of a few<br />

nucleons j i generate the<br />

total nuclear spin<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 6


Collective Level Scheme<br />

• This nucleus has 347 known levels l and 516 gamma rays !<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 7


Noncollective Level Scheme<br />

• 148 Gd is an example of a<br />

nucleus showing single-<br />

particle behaviour<br />

• Complicated set of<br />

energy levels<br />

• No regular features<br />

e.g. band structures<br />

• Some states are<br />

isomeric<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 8


Do Nuclei Really Rotate?<br />

• Should we talk about collective motion in nuclei?<br />

• We need to identify fast and slow degrees of freedom<br />

• For example, in molecules electronic motion is the<br />

fastest, vibrations are 10 2 times slower and rotations 10 6<br />

times slower. These motions have very different time<br />

scales so the wavefunction can be separated into a<br />

product of the terms<br />

• For nuclei the differences are much smaller. Collective<br />

and single-particle modes can perhaps be separated, but<br />

they will interact strongly !<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 9


Single-particle/Collective Modes<br />

• Collective and single-particle modes will interact strongly<br />

• Core polarisation<br />

• Coriolis Forces: ‘backbending’,<br />

modification of shell structure,<br />

quenching of pairing<br />

• Finite size effects: ‘band termination’,<br />

blocking of collective excitations<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 10


What is a Model?<br />

• Quantum mechanics governs basic nuclear behaviour<br />

• The forces are complicated and cannot be written down<br />

explicitly<br />

• It is a many-body problem of great complexity<br />

• In the absence of a comprehensive nuclear theory we<br />

turn to models<br />

• A model is simply a way of looking at the nucleus that<br />

gives a physical insight into a wide range of its properties<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 11


2. Spherical Shell Model<br />

• Single-Particle Shell Model<br />

• Square Well, Harmonic Oscillator, Woods-Saxon<br />

• Spin-Orbit Obit Coupling<br />

• Shell Structure<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 12


Experimental Shell Effects<br />

• The energies of the<br />

first excited 2 + states<br />

in nuclei peak at the<br />

magic numbers of<br />

protons or neutrons<br />

• ‘B(E2)’ values (∝ 1/τ<br />

where τ is the mean<br />

lifetime) of the 2<br />

+<br />

states reach a minimum<br />

at the magic numbers<br />

• ‘Magic’ nuclei are<br />

spherical and the least<br />

collective<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 13


First 2 + Energies<br />

Z<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 14<br />

N


Systematics Near Z(N) = 50<br />

N = 50<br />

Z = 50<br />

• 100 Sn (Z=N=50) and 132 Sn (N=82) are doubly magic nuclei<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 15


Neutron Separation Energies<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 16


Shell Model – Mean Field<br />

N nucleons in<br />

a nucleus<br />

A nucleon in the<br />

Mean Field of<br />

N-1 nucleons<br />

• Assumption – ignore detailed two-body interactions<br />

• Each particle moves in a state independent of other<br />

particles<br />

• The Mean Field is the average smoothed-out<br />

interaction with all the other particles<br />

• An individual nucleon only experiences a central force<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 17


Shell Model Hamiltonian<br />

• If the short range interaction potential between two<br />

nucleons i and j is v(r ij ), then the average potential<br />

acting on each particle is:<br />

V i (r i ) = 〈 ∑ j v(r ij ) 〉<br />

• The Hamiltonian, H = ∑ i T i + ∑ ij v(r ij ), can be rewritten:<br />

H’ = ∑ i [T i + V i (r i )] + λ [∑ ij v(r ij ) - ∑ i V i (r i )]<br />

mean field residual interaction<br />

• For λ = 1, H’ = H. The shell model assumption is that<br />

λ → 0, i.e. the central interaction is much larger than<br />

the residual interactions<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 18


Choice of Potential<br />

• A central potential V(r i ) only depends on the distance<br />

r i and is made up of a superposition of short-range<br />

internucleonic potentials:<br />

V(r i ) = ∫v|r i – r’| ρ(r’) dr’<br />

• ‘ρ(r’)’ is the density distribution of the nucleus<br />

• The internucleonic potential may be represented by a<br />

delta function: v(r ij ) = -V 0 δ(r ij )<br />

• Then: V(r i ) = V 0 ρ(r)<br />

• The Schrödinger equation is: [T + V] Ψ(r) = E Ψ(r)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 19


Some Potential Wells<br />

• Square Well: V(r) = -V 0 for r ≤ R 0<br />

= 0 for r > R 0<br />

• Gaussian Well: V(r) = -V 0 exp[-(r/a) 2 ]<br />

• Exponential Well: V(r) = -V 0 exp[-2r/a]<br />

• Yukawa Well: V(r) = -(V 0 /r) exp[-r/a]<br />

• Harmonic Oscillator: V(r) = -V (/R) 0 [1-(r/R 0 2 ]<br />

• Woods-Saxon: V(r) = -V 0 / {1 + exp[(r-R 0 )/a]}<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 20


Well Comparisons<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 21


Square Well Potential<br />

Infinite square<br />

well potential<br />

• Simplest form of potential<br />

• Since we have a spherically<br />

symmetric potential we can<br />

separate the solutions into<br />

angular and radial parts<br />

• Radial solutions are Bessel<br />

functions which satisfy the<br />

boundary condition R nl (R) = 0<br />

• The eigenvalues are:<br />

R nl = {A/√(κr)} J l+½ (κr)<br />

where A is a constant and κ is<br />

the wave number of the nucleon:<br />

κ 2 = (2M/ħ 2 )[E nl + V]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 22


Square Well Labels<br />

• The levels are labelled by n and l (‘s’ = 0, ‘p’ = 1, ‘d’ = 2,<br />

‘f’ = 3, ‘g’ = 4, ‘h’ = 5, ‘i’ = 6, ‘j’ = 7, ‘k’ = 8)<br />

• Each level has 2(2l + 1) substates<br />

• The first few levels (different from H atom):<br />

Level Occupation Total<br />

1s 2 2<br />

1p 6 8<br />

1d 10 18<br />

2s 2 20<br />

1f 14 34<br />

2p 6 40<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 23


Harmonic Oscillator Potential<br />

Simple harmonic<br />

oscillator potential<br />

• Easy to handle analytically<br />

• Form of potential:<br />

V HO (r) = -V + ½mr 2 ω 2<br />

• Solutions are Laguerre<br />

polynomials<br />

• Eigenenergies are labelled by<br />

the oscillator quantum number N:<br />

E N = (N + 3/2) ħω<br />

• For each N there are degenerate<br />

levels l with n and l that satisfy:<br />

2(n-1) + l = N, N ≥ 0, 0 ≤ l ≤ N<br />

• The parity of each shell is (-1) N<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 24


Harmonic Oscillator Degeneracies<br />

• For each N there are degenerate energy levels with n<br />

and l that satisfy:<br />

2(n-1) + l = N, N ≥ 0, 0 ≤ l ≤ N<br />

• Even N contains only l even states; odd N, odd l<br />

• The degeneracy condition i is:<br />

∆l = 2 and ∆n = 1<br />

(e.g. N = 4 3s, 2d, 1g orbits)<br />

• It is the fundamental reason for shell structure, i.e.<br />

clustering of levels<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 25


Harmonic Oscillator Labels<br />

• The number of degenerate levels for a given N is<br />

(N+1)(N+2)<br />

N allowed l E N Occupation Total<br />

0 0 3/2 2 2<br />

1 1 5/2 6 8<br />

2 2,0 7/2 12 20<br />

3 3,1 9/2 20 40<br />

4 4,2,0 20 11/2 30 70<br />

5 5,3,1 13/2 42 112<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 26


(Wrong) Magic Numbers<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 27


Spin-Orbit Potential<br />

• In Atomic <strong>Physics</strong> the spin-orbit interaction comes<br />

about due to the interaction of an electron’s magnetic<br />

moment with the magnetic field generated by its<br />

motion about the nucleus<br />

• A similar interaction was introduced for nuclei to<br />

empirically fit the observed magic numbers<br />

• A term is added to the potential:<br />

V(r) → V(r) + µ l.ss<br />

• The new term makes the force felt by a nucleon<br />

dependent on the direction of its spin<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 28


Spin Orbit Energy<br />

• The spin-orbit term does not violate<br />

spherical symmetry and leaves l, j<br />

and j z as good quantum numbers,<br />

although l z and s z are not<br />

• The spin-orbit energy is:<br />

E l.s = {[4j(j+1)-4l(l+1)-1]/8}ħ 2 µ<br />

The vectors L and<br />

S precess about J<br />

• By making µ < 0, the magic numbers<br />

can be reproduced<br />

• States t with j = l + ½ are lower in<br />

energy than those states with j = l -<br />

½ (opposite way round to spin-orbit<br />

interaction ti in atoms !)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 29


Modified Harmonic Oscillator<br />

• The harmonic oscillator<br />

shells are shown to the<br />

left in this diagram<br />

• In the middle, an l 2 term<br />

is added to make the<br />

potential more realistic<br />

• A spin orbit term l.s is<br />

added to the right with<br />

its strength adjusted to<br />

obtain the correct<br />

nuclear magic numbers<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 30


Woods-Saxon Potential<br />

V 0 ~ 50 MeV<br />

R 0 ~ 6-7 fm (A=125-190)<br />

a ~ 0.5 fm<br />

4a = ‘skin thickness’<br />

• The Woods-Saxon (WS) nuclear potential is ‘supposedly’<br />

the most realistic<br />

• The potential has the form:<br />

V(r) = -V 0 / { 1 + exp[(r - R 0 ) / a] }<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 31


WS vs. MHO Potentials<br />

• The Woods-Saxon (WS)<br />

potential is the most realistic<br />

• The l 2 term in the Modified<br />

Harmonic Oscillator (MHO)<br />

potential flattens the bottom<br />

of the potential making it look<br />

more like the Woods-Saxon<br />

shape<br />

• There are slight differences<br />

between the MHO and WS<br />

energy levels, e.g. the<br />

ordering of the 2d 5/2 and 1g 7/2<br />

levels is interchanged<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 32


3. <strong>Nuclear</strong> Deformation<br />

• Shape parameterisation<br />

• Quadrupole deformation, β and γ<br />

• Triaxiality<br />

• Anisotropic Harmonic Oscillator, Nilsson Model<br />

• Large deformation<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 33


Evidence for Deformation<br />

1. Large electric quadrupole moments Q 0<br />

2. Low-lying rotational bands (E ∝ I[I+1] )<br />

The origin of deformation lies in the long range<br />

component of the nucleon-nucleon residual interaction:<br />

a quadrupole-quadrupole interaction gives increased<br />

binding energy for nuclei which lie between closed<br />

shells if the nucleus is deformed. d In contrast, the<br />

short range (pairing) component favours sphericity<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 34


Simple <strong>Nuclear</strong> Shapes<br />

• In the description of a ‘drop’<br />

of nuclear matter with a sharp<br />

surface, the equipotential<br />

surface R(θ,φ) can be<br />

expressed as a sum over<br />

spherical harmonics Y λµ (θ,φ):<br />

R(θ,φ) = R 0 [1 + ∑ λ ∑ µ α λµ Y λµ (θ,φ)]<br />

• Here R 0 is the radius of a<br />

sphere and the α λµ coefficients<br />

represent distortions from the<br />

equilibrium i spherical shape<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 35


Volume Conservation<br />

• By integrating over the shape of the nucleus, the volume<br />

for small deformation is:<br />

V ≈ (4π/3) [1 + 3α 00 /√(4π)] R 0<br />

3<br />

• To account for the incompressibility of nuclear matter<br />

we demand volume conservation under distortions and<br />

hence set α 00 = 0<br />

• A factor C(α λµ ) may be introduced to satisfy the<br />

conservation of volume more precisely:<br />

R(θ,φ) = C(α λµ ) R 0 [1 + ∑ λ ∑ µ α λµ Y λµ (θ,φ)]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 36


Most Important Multipoles<br />

• The λ = 1 term describes the displacement of the centre<br />

of mass and therefore cannot give rise to intrinsic<br />

excitation of the nucleus – ignore !<br />

• The λ = 2 term is the most important term and describes<br />

quadrupole deformation<br />

• The λ = 3 term describes octupole shapes which can look<br />

like pears (µ = 0), bananas (µ = 1) and peanuts (µ =2,3)<br />

• The λ = 4 term describes hexadecapole shapes<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 37


Variety of Shapes<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 38


Theoretical Deformations<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 39


Quadrupole Deformation<br />

The Euler angles<br />

relate the intrinsic<br />

(nucleus) and lab<br />

frame axes<br />

• The description of the nuclear<br />

shape simplifies if we make the<br />

principal axes of our coordinate<br />

system, i.e. (x, y, z), coincide with<br />

the nuclear axes (1, 2, 3)<br />

• Then α 22 = α 2-2 , and α 21 = α 2-1 = 0<br />

• The two independent coefficients<br />

α 20 and α 22 , together with the<br />

three Euler angles, then<br />

completely define the system<br />

• The shape then simplifies to:<br />

R = C R 0 [1 + α 20 Y 20 + α 22 (Y 22 +Y 2-2 )]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 40


β 2 and γ Parameters<br />

• An alternative parameterisation in the system of principal<br />

axes introduces the polar coordinates (β 2 , γ) through the<br />

relations:<br />

α 20 = β 2 cos γ and α 22 = -1/√2 β 2 sin γ<br />

• The parameter β 2 measures the total deformation:<br />

β<br />

2<br />

2 =∑ µ |α 2µ | 2<br />

• The parameter γ measures the lengths along the principal<br />

axes. For γ = 0°, the shape is prolate with the z-axis as<br />

the (long) symmetry axis<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 41


Quadrupole β 2 and γ Parameters<br />

prolate<br />

x > y = z<br />

60°<br />

oblate<br />

x = z > y<br />

Axially symmetric shapes<br />

γ = n 60°<br />

0°<br />

prolate<br />

x = y < z<br />

oblate<br />

x = y > z -60°<br />

prolate<br />

x = z < y<br />

oblate<br />

x < y = z<br />

Triaxial shapes : x ≠ y ≠ z<br />

γ ≠ n 60°<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 42


Lund Convention<br />

• In order to specify the<br />

triaxiality of a deformed<br />

quadrupole intrinsic<br />

shape, the range of γ<br />

values,<br />

0° ≤ γ≤60° is sufficient<br />

i • However, in order to<br />

specify a cranked system,<br />

we need three times this<br />

range, corresponding to<br />

the three principal axes<br />

about which the system<br />

can be cranked<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 43


Deformation Systematics<br />

Theory<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 44


Approximate Value of β 2<br />

• From an empirical fit to E2 transition rates (not valid<br />

near closed shells) it is found:<br />

T 4 γ (E2; 2 + → 0 + ) = (4 ± 2) x 10 10 Z 2 E γ4 A -1<br />

• This can be used for a ‘Grodzins’ estimate of the<br />

quadrupole deformation parameter:<br />

β 2 ≈ { 1225 / [ A 7/3 E(2 + ) ] } 1/2<br />

with the 2 + energy expressed in MeV<br />

• The energy of the 2 + state of an even-even nucleus hence<br />

gives an insight into the nuclear deformation. The lower<br />

the 2 + energy, the larger is β 2 and also the nuclear<br />

moment of inertia<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 45


First 2 + Energies<br />

Z<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 46<br />

N


Triaxiality<br />

• All three principal axes have different lengths:<br />

R x ≠ R y ≠ R z<br />

i.e. ‘short’, ‘long’ and ‘intermediate’ t axes<br />

• There is no symmetry axis (however, there is reflection<br />

symmetry) so K is not a good quantum number<br />

• The low-spin energy levels l in even-even nuclei move<br />

around as a function of γ<br />

• For γ≠0° an effective quadrupole deformation<br />

parameter may be defined:<br />

β 2 2 1/2<br />

eff = β { 4 sin (3γ) / [9 - √(81 -72 sin (3γ) )] } 2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 47


Asymmetric Rotor Model<br />

• The Asymmetric Rotor Model (ARM) investigates rigid<br />

triaxial shapes<br />

• The energies of the first two 2 + states are:<br />

E(2 + ) = (6ħ 2 /2I) {9[1 ± √(1 - 8/9 sin 2 (3γ))] /4 sin 2 (3γ)}<br />

• Hence from the experimental energies of the first 2 +<br />

states, a value of |γ| can be deduced<br />

• The higher spin states of the ground-state and γ-bands<br />

move around in energy as γ changes<br />

• Increasing γ tends to lower the energy levels of the<br />

γ-band relative to the ground-state band<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 48


ARM Energy Levels<br />

• These are the<br />

lowest energy<br />

levels of an<br />

asymmetric rigid-<br />

rotor predicted by<br />

the ARM<br />

• Note that the<br />

second 2 + state<br />

falls below the<br />

first 4 + state for<br />

|γ| ≥ 15°<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 49


More ARM Relations<br />

• For the odd-spin members of the γ band:<br />

E(3 + ) = E 1 (2 + ) + E 2 (2 + ) and E(5 + ) = 4 E 1 (2 + ) + E 2 (2 + )<br />

• Percentage differences for N = 76 isotones:<br />

Nucleus |γ| R 3 (%) R 5 (%)<br />

128<br />

Te 26.6° -1.21<br />

130<br />

Xe 27.6° +1.55<br />

132<br />

Ba 26.3° -0.98<br />

134 Ce 25.3° -179 -1.79<br />

136<br />

Nd 25.7° +0.38 +13.2<br />

138<br />

Sm 27.0° +0.79 +18.7<br />

140<br />

Gd 26.8° -2.53 +16.5<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 50


γ-rigidity or γ-softness?<br />

• The ARM considers<br />

the rotation of a rigid<br />

triaxial shape<br />

• The other extreme is<br />

a completely flat<br />

potential with respect<br />

to γ, with γ oscillating<br />

uniformly between<br />

γ = 0° (prolate) and<br />

γ = 60° (oblate)<br />

• Since the average is<br />

γ = 30°, we compare<br />

the two models at this<br />

value<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 51


Gamma-Band Staggering<br />

• As γ increases, a staggering arises between the odd-spin<br />

and even-spin members of the band<br />

• One way to measure this is to form the ratio:<br />

S(4,3,2) = { [ E 2 (4 + ) – E 1 (3 + ) ] - [ E 1 (3 + ) – E 2 (2 + ) ] } / E 1 (2 + )<br />

• The energies, in units of E 1 (2 + ), are:<br />

Model E + + + 2 (2 ) E 1 (3 ) E 2 (4 ) S(432) S(4,3,2)<br />

γ-rigid (30°) 2.0 3.0 5.67 +1.67<br />

γ-unstable 2.5 4.5 4.5 -2.0<br />

• Spherical Harmonic Vibrator: S(4,3,2) = -1.0<br />

• Symmetric Rotor (γ = 0°): S(4,3,2) = +0.33<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 52


S(4,3,2) Ratios vs. γ<br />

• The S(432) S(4,3,2)<br />

values for various<br />

types of motion<br />

are shown here as<br />

a function of the<br />

γ deformation<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 53


Do Triaxial Nuclei Really Exist?<br />

• Considerable effort has been made over the last<br />

twenty years to obtain conclusive evidence of (static)<br />

triaxial nuclear shapes<br />

• These efforts have recently been intensified by the<br />

experimental evidence of chirality (handedness) and<br />

the wobbling (precession) mode in nuclei (discussed<br />

later)<br />

• Triaxiality is an essential prerequisite for the<br />

manifestation of both of these effects in the<br />

atomic nucleus !<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 54


Anisotropic Harmonic Oscillator<br />

• The Anisotropic Harmonic Oscillator (AHO) potential<br />

for a spheroidal nucleus deformed along the z-axis may<br />

be written:<br />

V 2 osc = ½M[ω<br />

2<br />

⊥ (x 2 + y 2 ) + ω z2 z 2 ]<br />

• Here ω ⊥ and ω z represent the frequencies of the simple<br />

harmonic motion perpendicular and parallel to the<br />

nuclear symmetry axis, respectively, and are functions<br />

of the nuclear deformation:<br />

ω z ≈ ω 0 [1 – 2/3 δ], ω ⊥ ≈ ω 0 [1 + 1/3 δ]<br />

and ω 3 2 0 = ω ⊥2 ω z for volume conservation<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 55


Harmonic Oscillator Quantum<br />

• The Harmonic Oscillator quantum ω 0 is usually taken to<br />

have an isospin dependence:<br />

ħω 0 = 41 A -1/3 [1 ± (N-Z)/3A] MeV<br />

where the minus sign is used for protons and the plus<br />

sign for neutrons<br />

• In the ‘stretched’ coordinate system, the potential may<br />

then be written simply as:<br />

V osc = ½ħω 0 (ε 2 ) ρ 2 [1 – 2/3 ε 2 P 2 (cos θ t )]<br />

where ε 2 is (yet) another deformation parameter<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 56


δ, β and ε Parameters<br />

• Three deformation parameters are often used:<br />

1. Delta: δ = ∆R/R<br />

2. Epsilon: ε 2 defines a rotational ellipsoid<br />

3. Beta: β 2 defines a rotational quadrupoloid<br />

• If the deformation is not so large, then the following<br />

approximations hold:<br />

ε 2 ≈ 0.946 β 2 (1 - 0.1126β 2 )<br />

δ ≈ 0.946 β 2 (1 - 0.2700β 2 )<br />

• Also the hexadecapole β 4 parameter has opposite sign<br />

to the ε 4 parameter: ε 4 ≈ -0.85 β 4<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 57


Solutions of the AHO<br />

• The eigenvalues of the AHO potential are:<br />

E(n z ,n ⊥ ) = [n z + ½] ħω z + [n ⊥ + 1] ħω ⊥<br />

or<br />

E(N,n z ,n ⊥ ) ≈ [N +3/2] ħω 0 –1/3δ[2n z -n ⊥ ] ħω 0<br />

with N = n z + n ⊥<br />

• The latter expression is simply the energies of a<br />

Spherical Harmonic Oscillator minus a correction term,<br />

proportional to the deformation<br />

• The energy levels are labelled by the asymptotic<br />

quantum numbers:<br />

Ω π [N n z Λ]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 58


AHO Labels<br />

• The energy levels l are labelled ll by the asymptotic<br />

quantum numbers: Ω π [N n z Λ]<br />

• ‘N’: N = n x + n y +n z (= n z + n ⊥ ) is the oscillator quantum<br />

number<br />

• ‘n z ’: n z describes the z-axis component of N<br />

• ‘Λ’: Λ = l z is the projection of l onto the z-axis<br />

• ‘Ω’: Ω = Λ + Σ is the projection of j = l + s onto the z-<br />

axis<br />

• ‘π’: π = (-1) l is the parity of the state<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 59


The Λ, Σ, Ω Quantum Numbers<br />

• Spin projections: Ω = Λ + Σ = Λ ± ½<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 60


AHO Degeneracies<br />

• Some of the degeneracies of the SHO are lifted<br />

• Consider the N = 4 shell spherical oscillator shell which<br />

has degeneracy (N + 1)(N + 2) = 30 with l = 4, 2, 0.<br />

• The onset of deformation causes these levels to split<br />

into (N + 1) levels, each of degeneracy 2(n ⊥ + 1):<br />

n z n ⊥ Occupation<br />

4 0 2<br />

3 1 4<br />

2 2 6<br />

1 3 8<br />

0 4 10<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 61


Levels of the AHO<br />

• The splitting of the<br />

N = 4 oscillator shell<br />

is shown here when<br />

deformation is<br />

introduced<br />

• Note that levels with<br />

large n z (and hence<br />

small n ⊥ ) are favoured<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 62


Nilsson Model<br />

• Nilsson added terms proportional to l 2 and l.ss similar to<br />

the spherical case<br />

• The resulting Modified d Harmonic Oscillator (MHO) or<br />

Nilsson potential may be written as:<br />

V MHO = V osc – κħω 0 [2l t .s + µ(l t<br />

2<br />

- 〈l t2 〉 N ]<br />

where κ and µ are adjustable parameters. They are<br />

different for each major oscillator shell<br />

• The l t .ss term imitates tes the nuclear spin-orbit interaction<br />

n<br />

in the stretched coordinate system<br />

• The l 2 t2 term deepens the effective potential ti for<br />

particles near the nuclear surface<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 63


Remaining Degeneracies<br />

Nilsson Diagram<br />

• The l t .s and l t2 terms<br />

lift the 2(n ⊥ + 1)<br />

degeneracy of the<br />

N = n z + n ⊥ states<br />

• States with different<br />

Ω now have different<br />

energy<br />

• Each Ω π [N n z Λ] state<br />

is only twofold<br />

degenerate,<br />

corresponding to<br />

particles with ±Ω<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 64


Nilsson Single-Particle Diagrams<br />

N<br />

Z<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 65


Splitting of Ω States<br />

• Low Ω states favour<br />

prolate shapes<br />

• High Ω states<br />

favour oblate<br />

shapes<br />

• Note that each Ω<br />

state is now only<br />

twofold degenerate<br />

(±Ω)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 66


Splitting of Ω States<br />

David Campbell<br />

Florida State<br />

University<br />

it<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 67


Asymptotic Quantum Numbers<br />

• Because of the additional l.s and l 2 terms the<br />

physical quantities labelled by n z and Λ are not<br />

constants of the motion, but only approximately so<br />

• These quantum numbers are called asymptotic as<br />

they only come good as ε 2 → ∞<br />

• However, the quantum numbers N, Ω and π are always<br />

good labels provided that:<br />

1. the nucleus is not rotating and<br />

2. there are no residual interactions<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 68


Proton Nilsson Diagram<br />

• A ‘Nilsson Diagram’<br />

shows nuclear energy<br />

levels as a function of a<br />

quadrupole deformation<br />

parameter (β 2 , ε 2 or δ)<br />

• In this diagram, the<br />

large spherical shell<br />

gap at Z = 50 is rapidly<br />

diminished by the onset<br />

of deformation for both<br />

prolate (β 2 > 0) and<br />

oblate (β 2 < 0) shapes<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 69


Intruder Orbitals<br />

• The slope of Nilsson<br />

levels is related to the<br />

single-particle matrix<br />

element of the<br />

quadrupole operator:<br />

dE/dβ = - 〈k|r 2 Y 20 |k〉<br />

• Unnatural-parity low<br />

Ω prolate orbitals may<br />

‘intrude’ down into a<br />

lower shell at large<br />

deformation<br />

• This is the origin of<br />

superdeformation<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 70


Large Deformations<br />

• Deformed shell gaps<br />

(new ‘magic numbers’)<br />

emerge when the ratio<br />

of the major and minor<br />

nuclear axes are equal<br />

to the ratio of small<br />

integers<br />

• A superdeformed<br />

d<br />

shape has a major to<br />

minor axis ratio of 2:1<br />

• A hyperdeformed<br />

shape has a major to<br />

minor axis ratio of 3:1<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 71


Superdeformed 152 Dy<br />

• The SD band in<br />

152<br />

Dy is a very<br />

regular structure<br />

t<br />

with equally<br />

spaced gamma-ray<br />

transitions<br />

Original SD γ-ray spectrum<br />

from 1986 (Daresbury)<br />

• The spacing is<br />

relatively small,<br />

i.e. the band has a<br />

large moment of<br />

inertia (close to<br />

the rigid body<br />

value)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 72


Superdeformed Axis Ratios<br />

• The moment of inertia of a rigid sphere is:<br />

I rig = (A 5/3 /72) ħ 2 MeV -1<br />

• The moment of inertia of a prolate ellipsoid undergoing<br />

rigid rotation is:<br />

I rig = (A 5/3 /72) (1 + x 2 )/ 2x 2/3 ħ 2 MeV -1<br />

where x is the ratio of major to minor axes<br />

• The moment of inertia is not always a good indicator of<br />

nuclear deformation (e.g. pairing)<br />

• The quadrupole moment (charge distribution) is a<br />

better indicator:<br />

Q 0 = (2/5) Z R 2 (x 2 –1)/ x 2/3 eb<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 73


SD Systematics<br />

Nucleus Q 0 (eb) Axis Ratio<br />

36<br />

Ar 1.18 1.55<br />

60<br />

Zn 2.75 1.54<br />

82<br />

Sr 3.54 1.47<br />

91<br />

Tc 8.1 1.85<br />

108 Cd >9.5 >1.8<br />

132<br />

Ce 7.4 1.45<br />

152 Dy 17.5 185 1.85<br />

192<br />

Hg 17.7 1.61<br />

236<br />

U 32 1.84<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 74


SD Regions<br />

Z<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 75<br />

A


4. Vibrational Motion<br />

• Spherical Harmonic Vibrator<br />

• Particle-vibration coupling<br />

• Rotation-Vibration ti ti Model<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 76


Spherical Harmonic Vibrator<br />

• A dynamic deformation<br />

• We assume the nucleus<br />

is spherical in its ground<br />

state and the excited<br />

states are due to<br />

harmonic oscillations of<br />

the nuclear surface<br />

• For a quadrupole vibration, raton, the potential nta may be written: wrtt V vib = ∑ µ {½C 2 |α 2µ | 2 +½B 2 |dα 2µ /dt| 2 }<br />

where C 2 is a parameter representing the restoring<br />

potential and B 2 is associated with the mass carried by<br />

the vibration. This mode is possible since C 2 , determined<br />

by the surface tension, is low<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 77


Vibrator Eigenvalues<br />

• The eigenvalues of the spherical harmonic vibrator<br />

are:<br />

E n = E 0 + n ħω 2 with ω 2 = √(C 2 /B 2 )<br />

• E 0 represents the intrinsic and zero point motion of<br />

the oscillations<br />

• The energy levels for different n are equally spaced<br />

• Each phonon carries angular momentum 2 (Y 2 ) and has<br />

positive parity<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 78


• For an E2 transition:<br />

Allowed Transitions<br />

• For quadrupole vibrations,<br />

electromagnetic<br />

transitions are only<br />

allowed between states<br />

with:<br />

∆n = ±1<br />

〈I=2,n=1||E2||I=0,n=0〉 I 2 0 0 = √(5) Q vib e<br />

where Q vib is calculated from the Liquid Drop Model:<br />

Q vib = (3ZR 2 /4π) √(ħ/2B 2 ω 2 )<br />

• The magnetic moment is constant for λ = 2 states and<br />

therefore M1 transitions are not allowed<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 79


Multiphonon Vibrational States<br />

N = 3 (3 phonon)<br />

N = 2 (2 phonon)<br />

124<br />

Sn,<br />

spherical<br />

N = 1 (1 phonon)<br />

N = 0<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 80


Octupole Vibrations<br />

• For octupole vibrations,<br />

each phonon n carries<br />

angular momentum 3<br />

(Y 3 ) and negative parity<br />

• The energy of the first excited state (3 - ) is roughly<br />

twice the energy of the quadrupole case<br />

• For real nuclei, an anharmonic oscillator is needed. This<br />

removes the degeneracy of the n = 2 states (0 + , 2 + , 4 + )<br />

of the quadrupole vibrator. It also displaces the λ = 2<br />

and λ = 3 states relative to each other<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 81


Vibrational Movies…<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 82


Beta (Y 20 ) Vibration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 83


Gamma (Y 22 ) Vibration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 84


Octupole (Y 30 ) Vibration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 85


Octupole (Y 31 ) Vibration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 86


Octupole (Y 32 ) Vibration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 87


Octupole (Y 33 ) Vibration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 88


Particle-Vibration Coupling<br />

• For an odd-A nucleus<br />

near a closed shell with<br />

small deformation, the<br />

odd particle may couple<br />

to the surface vibrations<br />

of the core<br />

• The Hamiltonian is: H = [H int + H vib ] + H coup = H 0 + H coup<br />

• If we assume the interaction H coup → 0, the motions are<br />

decoupled from one another and the eigenfunctions will<br />

take a product form: H 0 Ψ = EΨ with Ψ = Ψ int Ψ vib<br />

• Consider coupling an h 9/2 proton to the 3 - state in 208 Pb,<br />

forming states in 209 Bi. Seven ‘degenerate’ states are<br />

formed by coupling spin vectors 3 and 9/2<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 89


Vibration or Rotation?<br />

• The simple ratio of<br />

the 4 + and 2 +<br />

energy levels of an<br />

even-even nucleus<br />

gives an indication<br />

of the types of<br />

excitation<br />

• For a vibrator:<br />

E ∝ n<br />

E(4 + )/E(2 + ) = 2.0<br />

• For a rotor:<br />

E ∝ I(I+1)<br />

E(4 + )/E(2 + ) = 3.33<br />

Te (Z = 52) systematics show that they are vibrational<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 90


Development of Collectivity<br />

• Another limiting value of the E(4 + )/E(2 + ) ratio is 2.5,<br />

corresponding to a γ-soft rotor, or γ-unstable oscillator<br />

(O(6) limit of the interacting tin boson model: IBM)<br />

• Adding protons to tin:<br />

Nucleus E(4 + )/E(2 + ) Behaviour<br />

116<br />

Sn (Z=50) 1.65 spherical<br />

118<br />

Te (Z=52) 1.99 vibrational<br />

120<br />

Xe (Z=54) 2.47 γ-soft<br />

122<br />

Ba (Z=56) 2.89 transitional<br />

124<br />

Ce (Z=58) 3.15 rotational<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 91


E(4 + )/E(2 + ) Values<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 92


Evolution of Structure<br />

• This diagram shows<br />

the evolution of level<br />

structure from<br />

closed shell (doubly<br />

magic, spherical) to<br />

midshell (rotational,<br />

deformed) nuclei<br />

• The corresponding<br />

E(4 + )/E(2 + ) ratios<br />

are also shown<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 93


Rotation-Vibration Model<br />

• The RVM model considers a well deformed (static),<br />

axially symmetric even-even nucleus and allows small<br />

fluctuations (dynamic) about the equilibrium shape β 0<br />

• After a ‘few’ approximations the energy spectrum may<br />

be written as:<br />

E = ½є [I(I+1) –K 2 R ] + є βn β + є γn γ<br />

where the є parameters are energies associated with<br />

rotations and vibrations<br />

• є R is related to β 0 and the nuclear moment of inertia I<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 94


RVM Quantum Numbers<br />

• The quantum numbers are constrained such that:<br />

K = 0, 2, 4,…<br />

I = 0, 2, 4,… for K = 0<br />

= K, K+1, K+2,… for K ≠ 0<br />

n β = 0, 1, 2,…<br />

n γ = K/2, K/2 + 2, K/2 + 4,…<br />

• So what are the possible low-lying l energy levels l ?<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 95


RVM Band Structure<br />

• For K = n β = n γ = 0, we expect a set of levels:<br />

l<br />

E = ½є R I(I+1)<br />

with I = 0, 2, 4,… ‘ground-state band’ represents pure<br />

rotation<br />

• A rotational ti band can be built on a β vibration by setting<br />

n β = 1. The energy levels are:<br />

E = є β + ½є R I(I+1)<br />

again with I = 0, 2, 4,… ‘β band’ (α 20 varies)<br />

• To include a γ vibration requires a nonzero K. So<br />

beginning with K = 2 and n γ = 1, the levels are:<br />

E = є γ + ½є R [I(I+1) – 4]<br />

this time with I = 2, 3, 4,… ‘γ band’ (α 22 varies)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 96


β and γ Bands<br />

• β-vibrational and γ-<br />

vibrational bands<br />

coexist with the<br />

rotational groundstate<br />

band in<br />

deformed nuclei<br />

• Such bands are found<br />

predominantly in the<br />

regions:<br />

150 ≤ A ≤ 190 and<br />

A ≥ 230<br />

which are far from<br />

shell closures<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 97


E0 Transitions<br />

• β vibrations give rise to enhanced E0 transitions due to<br />

radial shape oscillations in β<br />

• A measure of E0 strength is ‘rho squared’ :<br />

ρ 2 (E0) = |M(E0)/eR 2 | 2<br />

where M(E0) is the monopole moment operator:<br />

M(E0) = ∫ρ(r)r 2 dτ<br />

• The ρ 2 values are usually yquoted in units of ρ 2 (E0) x 10 3<br />

• In the RVM model we expect values around 200, but<br />

these are somewhat larger than experiment<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 98


More Vibrational Bands<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 99


Octupole Vibrational Bands<br />

238<br />

U<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 100


Nonadiabatic Vibration<br />

• For the surface modes of<br />

vibration, the frequency<br />

(velocity) of the<br />

oscillations is much smaller<br />

than that of the individual<br />

nucleonic motion<br />

• The motion is ‘adiabatic’<br />

(as is nuclear rotation) and<br />

individual quantum levels<br />

are evident<br />

• However, ‘nonadiabatic’<br />

collective motion can<br />

occur: ‘giant resonances’<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 101


Giant Resonances<br />

Monopole<br />

L = 0<br />

Isoscalar<br />

Isovector<br />

Dipole<br />

L = 1<br />

Quadrupole<br />

L = 2<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 102


5. Collective <strong>Nuclear</strong><br />

Rotation<br />

• <strong>Nuclear</strong> moments of inertia<br />

• Rotational band properties, signature<br />

• Particle-rotor t coupling<br />

• High-K bands, wobbling motion<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 103


Gammasphere<br />

• The Hulk<br />

• News Item<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 104


Moment of Inertia<br />

• Deformation provides an element of anisotropy allowing<br />

the definition of a nuclear orientation and the possibility<br />

of observing rotation<br />

• Classically the energy associated with rotation is:<br />

E 2 2 rot = ½ I ω = I / 2 I ; ω = I / I<br />

• Collective rotation involves the coherent contributions<br />

from many nucleons and gives rise to a smooth relation<br />

between energy and spin:<br />

E = (ħ 2 /2I) I[I + 1]<br />

which h defines the ‘static’ ti moment of inertia, sometimes<br />

denoted I (0)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 105


Energy Levels of a Rotor<br />

• The energy levels of a rotor<br />

are proportional p rti to I(I+1)<br />

• The ratios of energy levels<br />

for a rotor are:<br />

E(4 + )/E(2 + ) = 3.333<br />

E(6 + )/E(2 + ) = 7.0<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 106


Rotational Frequency<br />

• The intensive variable ω<br />

(rotation about the x axis)<br />

is related to the extensive<br />

variable I by the relation:<br />

ħω = dE/dI x<br />

≈ ½[E(I+1) – E(I-1)<br />

• Here I x is the projection of I<br />

onto the rotation axis (x):<br />

I x = √[I(I+1)-K 2 ] ħ<br />

The rotational frequency ω is distinct from the oscillator<br />

quantum ω 0 . In practice ω « ω 0 and the collective<br />

rotation can be considered as an adiabatic motion<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 107


Rigid Body Moment of Inertia<br />

• The rigid-body moment of inertia for a spherical nucleus<br />

is:<br />

I rig = (2/5) MR 2 = (2/5) A 5/3 m N r<br />

2<br />

0<br />

where m N is the mass of a nucleon (M = A m N ) and<br />

R = r 0 A 1/3 with r 0 = 12 1.2 fm<br />

• For a deformed nucleus:<br />

I rig = (2/5) A 5/3 m N r 02 [1 + 1/3 δ]<br />

where δ = ∆R / R 0<br />

• Typically nuclear moments of inertia are less than 50%<br />

of the rigid-body value at low spin<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 108


<strong>Nuclear</strong> Moments of Inertia<br />

• <strong>Nuclear</strong><br />

moments of<br />

inertia are<br />

lower than the<br />

rigid-body<br />

value – a<br />

consequence<br />

of nuclear<br />

pairing<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 109


<strong>Nuclear</strong> Rotation<br />

• The assumption of the ideal flow<br />

of an incompressible nonviscous<br />

fluid (Liquid Drop Model) leads<br />

to a hydrodynamic moment of<br />

inertia (surface waves):<br />

I hydro = I rig δ 2<br />

• This estimate is much too low !<br />

• We require short-range pairing ii<br />

correlations to account for the<br />

experimental values<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 110


Kinematic and Dynamic MoI’s<br />

• Assuming maximum alignment on the<br />

x-axis (I x ~ I), the kinematic moment<br />

of inertia is defined:<br />

I (1) = (ħ 2 I) [dE(I)/dI] -1 = ħ I/ω<br />

• The dynamic moment of inertia<br />

(response of system to a force) is:<br />

• Note that I (2) = I (1) + ω dI (1) /dω<br />

I (2) = (ħ 2 ) [d 2 E(I)/dI 2 ]<br />

-1 = ħ dI/dω<br />

• Rigid body: I (1) = I (2) Nucleus at high spin: I (1) ≈ I (2)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 111


General Rotation<br />

• A deformed rotor has a Hamiltonian of the form:<br />

H 2 2 rot = Σ k A k R k , A k = ħ /2I k<br />

where I k is the moment of inertia about the k th axis<br />

• For triaxial shapes the moments of inertia are:<br />

I k = (4/3) I 0 sin 2 [γ + k 2π/3 ]<br />

• For an axial nucleus deformed along the z-axis,<br />

I 1 = I 2 = I 0 and I 3 = 0, and the Hamiltonian is:<br />

H 2 2 22 2 2<br />

rot = (ħ /2I 0 ) [R 1 + R ] = (ħ /2I 0 ) R<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 112


Irrotational Moments of Inertia<br />

• This diagram shows<br />

the variation of the<br />

moments of inertia I k<br />

as a function of the<br />

triaxiality parameter γ<br />

• For a prolate nuclear<br />

shape (γ = 0°), I 1 = I 2<br />

and I 3 = 0<br />

• For γ = 30° , I 2 reaches<br />

a maximum and this<br />

represents the ‘most<br />

collective’ shape<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 113


Angular Momentum Coupling<br />

• Provided that the collective rotation is slow relative to<br />

the single-particle motion (adiabatic condition), the<br />

nuclear Hamiltonian can be separated into intrinsic and<br />

rotational parts:<br />

H = H int + H rot<br />

with eigenvalues Ψ = Ψ int Ψ rot<br />

• The intrinsic motion has angular momentum J, which is<br />

not a conserved quantity. It couples to the collective<br />

rotation R to give total spin:<br />

I = R + J<br />

• The total spin I is a constant of the motion together<br />

with its projection M<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 114


Various Spin Projections<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 115


Rotation Matrices<br />

• The intrinsic wavefunction can be characterised by the<br />

K projection. The three variables I 2 , M and K completely<br />

specify the state of motion The eigenfunctions are given<br />

by:<br />

Ψ rot = |IMK 〉 = √[(2I +1)/8π 2 ] D IMK (θ,φ,ψ)<br />

where the functions D IMK are ‘rotation matrices’<br />

• Note: Î 2 D IMK = I(I+1)ħ 2 D IMK<br />

; Î Z D IMK = Kħ D IMK<br />

Î ± D IMK = √[I(I+1) – K(KK1)]ħ D IMKK1<br />

• The rotational energy is:<br />

(1/2I 2 x )(Î 2 – Î z2 ) Ψ rot i.e. E rot = (ħ 2 /2I x )[I(I+1) – K 2 ]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 116


Signature Quantum Number ‘r’<br />

• For K = 0, the D IMK functions reduce to spherical<br />

harmonics Y IM and the nuclear wavefunction is:<br />

•<br />

Ψ r,IMK=0 = (1/√2) Ψ r,K=0 Y IM<br />

• The quantum number r is the ‘signature’ , related to the<br />

invariance of the system when rotated 180° about an axis<br />

perpendicular to the symmetry axis (z): operator R(π)<br />

• A second rotation by 180° brings the system back to its<br />

original orientation. Hence:<br />

R 2 (π) Ψ r,IMK = r 2 Ψ r,IMK = Ψ r,IMK<br />

• The allowed values of r are: (-1) I<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 117


Bands of Good Signature<br />

• For K = 0, we may classify rotational bands in terms of<br />

the signature quantum number<br />

• For r = +1, the allowed spins are:<br />

I = 0, 2, 4,…<br />

• For r = -1, the allowed spins are:<br />

I = 1, 3, 5,…<br />

• Hence for each signature we obtain a rotational band<br />

with the energy levels separated by 2ħ<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 118


Rotational Bands with K ≠ 0<br />

• For K ≠ 0, the total nuclear wavefunction takes the<br />

antisymmetrised form in order to satisfy the rotation<br />

(reflection) symmetry:<br />

Ψ IMK = √[(2I+1)/16π 2 ] {Ψ K D IMK + (-1) I+K Ψ -K D IM-K }<br />

where Ψ -K corresponds to a projection of the spin –K and<br />

is obtained by the operation R(π) Ψ K<br />

• The consequence of R(π) invariance for K ≠ 0 is that the<br />

intrinsic states Ψ K and Ψ -K , with eigenvalues ±K of J z , are<br />

degenerate and constitute only a single sequence of<br />

states with spins:<br />

I = K, K+1, K+2,…<br />

i.e. states with alternating signature<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 119


Particle-Rotor Coupling<br />

• For an axially symmetric deformed d rotor:<br />

H rot = (ħ 2 /2I 0 0) R 2 = (ħ 2 /2I 0 0) [I –J] 2<br />

= (ħ 2 /2I 0 ) [I.I + J.J -2I.J]<br />

where the I.J couples the degrees of freedom of the<br />

valence particles to the rotational motion and is<br />

analogous to the classical Coriolis and centrifugal forces<br />

• Now consider J to consist of a single particle<br />

(J →j) coupled to an even-even core<br />

H rot = (ħ 2 /2I 0 ) [(I 2 –I z2 ) + (j 2 –j z2 )–(I + j - + I - j + )]<br />

The final term couples intrinsic and rotational states<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 120


Particle-Rotor Coupling Schemes<br />

• (a) shows the strong-coupling limit or deformation-<br />

aligned (DAL) coupling scheme<br />

• (b) shows the weak-coupling limit or rotation-aligned<br />

(RAL) coupling scheme<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 121


Strong Coupling (DAL)<br />

• This limit is recognised when the level splitting of the<br />

deformed shell-model single-particle energies for<br />

different Ω values is large compared with the Coriolis<br />

perturbation, i.e. large deformation or small Coriolis<br />

matrix elements (low j, high Ω)<br />

• The angular momentum vector j precesses around the<br />

deformation axis and K is approximately a good quantum<br />

number<br />

• The energy spectrum is given by the set of levels:<br />

E rot = (ħ 2 /2I 0 0) ) [I(I+1) –K 2 ]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 122


Decoupling Limit (RAL)<br />

• For weakly deformed nuclei, or fast enough rotation, the<br />

Coriolis force may be so strong that the coupling of the<br />

valence nucleon to the deformed core is negligible<br />

• The Coriolis force tends to align the nucleonic angular<br />

momentum j with that of the rotational angular<br />

momentum R<br />

• In this limit, the rotation band has spins:<br />

I = j, j+2, j+4,…<br />

• The energies are: E rot = (ħ 2 /2I 0 0) (I - j x )(I – j x +1)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 123


K = ½ Bands in Odd-A Nuclei<br />

• The rotational energy of a K = ½ band is:<br />

E(I) = (ħ 2 /2I a(-1) I+½ 0 ) [I(I+1) + (I+½)]<br />

where a is the decoupling parameter<br />

• Bands can mix if ∆K = ±1<br />

• For K = ½ bands there is a<br />

diagonal matrix element of<br />

the form: 〈K=½|j + |K=-½〉<br />

where j + = j x + ij y which<br />

perturbs the energy<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 124


High K (I z ) Bands<br />

• If we have many paired<br />

nucleons outside the closed<br />

shell in the ground state t<br />

then alignment with the x-<br />

axis becomes difficult<br />

because the valence<br />

nucleons lie closer to the<br />

z-axis axis, i.e. they have high<br />

Ω values<br />

K = I z = ∑j z = ∑Ω<br />

• The sum K of these<br />

projections onto the<br />

deformation (z) axis is now<br />

a good quantum number<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 125


K Forbidden Transitions<br />

• It is difficult for rotational bands with high K values to<br />

decay to bands with smaller K since the nucleus has to<br />

change the orientation on of its angular momentum.<br />

• For example, the K π = 8 - band head in 178 Hf is isomeric<br />

with a lifetime of 4 s. This is much longer than the<br />

lifetimes of the rotational states built on it.<br />

• The K π = 8 - band head is formed by breaking a pair of<br />

protons and placing them in the ‘Nilsson configurations’:<br />

Ω [N n 3 Λ] = 7/2 [4 0 4] and 9/2 [5 1 4]<br />

• In this case: K = 7/2 + 9/2 = 8 and π = (-1) N(1) .(-1) N(2) = -1<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 126


K Isomers in 178 Hf<br />

• A low lying state with<br />

spin I = 16 and K = 16 in<br />

178 Hf is isomeric with a<br />

half life of 31 years !<br />

• It is yrast (lowest state<br />

for a given spin) and is<br />

‘trapped’ since it must<br />

change K by 4 units in its<br />

decay<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 127


K Forbiddenness<br />

• Strictly, in the decay of a high-K h band-head, d K can only<br />

change by an amount up to the multipolarity λ of the<br />

transition<br />

• The ‘degree of K forbiddenness’ is:<br />

ν = |∆K| - λ<br />

• The ‘hindrance factor’ is:<br />

f = F W = T 1/2γ / T<br />

W<br />

1/2<br />

where T 1/2γ is the partial γ-ray half-life and T<br />

W<br />

1/2 is the<br />

theoretical Weisskopf estimate<br />

• The ‘reduced hindrance factor’ is:<br />

f ν = f 1/ν = [ T 1/2γ / T<br />

W<br />

1/2 ] 1/ν<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 128


Hindrance Factors<br />

• The solid line shows<br />

the dependence of F W<br />

on ∆K for some E1<br />

transitions according<br />

to an empirical rule:<br />

log F W = 2{|∆K| - λ}<br />

= 2ν<br />

• i.e. F W values increase<br />

approximately by a<br />

factor of 100 per<br />

degree of K<br />

forbiddenness<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 129


Wobbling Motion and Triaxiality<br />

• Wobbling is a fundamental mode due to triaxiality which<br />

occurs when the axis of collective rotation does not<br />

coincide with one of the principal axes<br />

• For a deformed rotor the Hamiltonian is:<br />

H rot = (ħ 2 /2I x ) I x<br />

2<br />

+ (ħ 2 /2I y ) I y2 + (ħ 2 /2I z ) I z<br />

2<br />

• For a well-deformed but triaxial nucleus with I x » I y ≠ I z<br />

the energy of the wobbling rotor is:<br />

E(I,n W ) = (ħ 2 /2I x ) I(I+1) + ħω W (I) (n W + ½)<br />

where n W is the number of wobbling phonons and ω W is<br />

the wobbling frequency<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 130


Wobbling Frequency<br />

• The wobbling frequency is related to the rotational<br />

frequency as:<br />

with<br />

ω W = ω rot √[ (I x - I y)(I x - I z ) / (I y I z ) ]<br />

ω rot = ħ I / I x<br />

• Note for an axially symmetric prolate nucleus, I z goes to<br />

zero and ω W → ∞, i.e. there is no wobbling motion<br />

• A family of wobbling bands is expected for n W = 0, 1, 2,…<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 131


Wobbling Motion<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 132


Wobbling Bands in 165 Lu<br />

• A family of wobbling<br />

bands is expected to<br />

show very similar<br />

internal structure<br />

• TSD (Triaxial<br />

SuperDeformed) bands<br />

1, 2 and 3 in 165 Lu<br />

represent bands with 0,<br />

1 and 2 wobbling<br />

phonons, respectively<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 133


Electromagnetic Properties<br />

• A characteristic signature<br />

of wobbling motion is the<br />

occurrence of ∆I = ±1<br />

interband transitions with<br />

unusually large B(E2) out<br />

values that compete with<br />

the strong ∆I = 2 inband<br />

transitions, B(E2) in<br />

• ∆n W = 2 transitions are<br />

forbidden<br />

• Measured multipole mixing ratios for the interband<br />

∆I = 1 transitions in 165 Lu show them to be ~90% E2<br />

and only ~10% M1 !<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 134


6. <strong>Nuclear</strong> Pairing<br />

• Pairing and superfluidity<br />

• Odd-even mass difference<br />

• Quasiparticles<br />

• Coriolis antipairing and backbending<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 135


Experimental Evidence<br />

• The ground states of all even-even nuclei have I π = 0 +<br />

• The binding energies of odd-even nuclei are less than the<br />

mean value of the two neighbouring even-even nuclei<br />

• Doubly odd nuclei are even less bound<br />

• <strong>Nuclear</strong> moments of inertia are only 30-50% of the rigid-<br />

body value at low spin<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 136


Time Reversed Orbits<br />

• The greatest overlap<br />

would occur if two<br />

particles could orbit<br />

in the same level<br />

• Not allowed (PEP) !<br />

• The next greatest<br />

overlap occurs for<br />

particles in ‘time<br />

reversed’ orbits<br />

• The spins cancel to<br />

give I π = 0 +<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 137


Coupling Two Particles<br />

• The short-range (pairing) residual interaction yields<br />

an energetically favoured 0 + state<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 138


Scattering Between Orbits<br />

• Pairs of particles scatter from one<br />

orbit to another, induced by the<br />

pairing interaction<br />

• The particles change orbits in pairs<br />

so I π = 0 +<br />

• Since the orbits have<br />

different energies, the<br />

Fermi surface is smeared<br />

out over a region ±∆<br />

(±1.5 MeV)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 139


Odd-Even Mass Difference<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 140


Pairing Energies<br />

• The neutron separation energy is:<br />

S n = B(A,Z) –B(A-1,Z) = M(A-1,Z) –M(A,Z) + M n<br />

where B(A,Z) in the nuclear binding energy<br />

• The proton separation energy is:<br />

S p = B(A,Z) – B(A-1 1, Z-1) = M(A-1 1,Z-1) – M(A,Z) + M H<br />

• The pairing energies are:<br />

P n (A,Z) = S n (A,Z) – S n (A-1,Z) (neutron)<br />

P p (A,Z) = S p (A,Z) – S p (A-1,Z-1) 1Z 1) (proton)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 141


Pairing Hamiltonian<br />

• The Hamiltonian including a two-body monopole (i.e. I = 0)<br />

pairing interaction is:<br />

H = H sp + H pair = ∑є u [a u<br />

†<br />

a u + a ū<br />

†<br />

a ū ] - G∑a u1<br />

†<br />

a ū1<br />

†<br />

a ū2 a u2<br />

• Here a † and a are particle creation and annihilation<br />

operators<br />

• The first term is the sum of single-particle energies<br />

• The second term contains the pairing interaction that<br />

annihilates a pair of particles in time reversed orbits<br />

|u 2 〉 and |ū 2 〉 and simultaneously creates a pair in time<br />

reversed orbits |u 1 〉 and |ū 1 〉<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 142


Chemical Potential λ<br />

• The energy increase of the condensate per particle<br />

added defines the chemical potential λ<br />

• The Hamiltonian is: H’ = H – λÑ = H sp + H pair – λÑ<br />

where Ñ is the particle number operator<br />

• The two-body monopole pairing interaction is:<br />

H pair = -¼G P † P<br />

where the pair creation and annihilation operators are:<br />

P † = ∑ a † u a † ū<br />

and P = ∑ a ū a u<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 143


Pairing Strength G<br />

• The strength of the pairing term G is a positive constant<br />

• It is larger for high-j orbitals and depends on the spatial<br />

overlap of the two nucleons<br />

• The strength decreases with mass since in heavier nuclei<br />

the outer nucleons are further apart<br />

• The strength is also lower for protons than neutrons<br />

because of Coulomb repulsion<br />

• Approximately:<br />

G p = 17/A MeV and G n = 23/A MeV<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 144


Pairing Gap ∆<br />

• The pairing term contains the product of two creation<br />

and two annihilation operators<br />

• In order to simplify the calculations, the term P † P<br />

(product) is replaced by P † + P (sum) and:<br />

H pair = -½∆ [P † + P]<br />

which h introduces the pairing ii gap parameter ∆<br />

• Particle number is now not conserved ! The chemical<br />

potential λ is now treated as a Lagrange multiplier l and is<br />

varied to produce the correct particle number:<br />

〈Ψ|Ñ|Ψ〉 Ñ = N λ = - ∂E/∂N<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 145


Single Particle Levels with Pairing<br />

• An energy gap between<br />

the ground state and<br />

first excited state t of<br />

~ ∆ opens up<br />

• The excited states<br />

become bunched<br />

togetherth<br />

• A rough estimate of<br />

the energy required to<br />

create a particle-hole<br />

excitation ti is 2∆<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 146


<strong>Nuclear</strong> Ground State<br />

• Nuclei in their ground states are in specific<br />

configurations: some pairs of nucleons are above the<br />

Fermi surface (λ) and some states below the Fermi<br />

surface are empty<br />

• With pairing, states are not always full or always empty<br />

but filled for part of the time or empty for part of the<br />

time<br />

• The probability of a given level є u being occupied by a<br />

particle is:<br />

P u (є u ) = ½{ 1 + (є u – λ) / √[ (є u – λ) 2 + ∆ 2 ] }<br />

• Now P u (є u ) ≠ 0 or 1 around the Fermi surface !<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 147


Quasiparticles<br />

• A further simplification is to replace the pairwise<br />

interacting particles by a gas of noninteracting<br />

‘quasiparticles’, whose energies are then simply additive<br />

• A quasiparticle may be considered as a mixture of a<br />

particle and hole states<br />

• The Bogoliubov-Valatin l transformation ti changes the<br />

particle basis (a † ,a) into the quasiparticle basis (α † ,α):<br />

α u<br />

†<br />

= U u a u† + V u a ū<br />

; a u† = U u α u† -V u α ū<br />

α<br />

† ū = U u a ū† - V u a u ; a ū† = U u α ū† + V u α u<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 148


The Quasiparticle Vacuum<br />

• The transformation coefficients U u and V u can be<br />

obtained following a BCS treatment (superconductivity)<br />

• The BCS wavefunction is of the form:<br />

|Ψ † † BCS 〉 = Π u [U u + V u a u† a ū† ] |0〉<br />

where |0〉 〉 denotes the vacuum state of the particles<br />

and |Ψ BCS 〉 represents the quasiparticle vacuum<br />

• U u and V u represent occupation amplitudes (‘empty’ and<br />

‘filled’, respectively) and hence:<br />

|U 2 2 u | + |V u | = 1<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 149


Quasiparticle Energies<br />

• Expressions for U u and V u are:<br />

U u = (1/√2) ){ 1 + (є + λ) / E 1/2<br />

u ) u }<br />

V u = (1/√2) { 1 + (є u – λ) / E u } 1/2<br />

• The quasiparticle energy of a state |u〉 relative to the<br />

ground state is:<br />

E u = √[ (є u – λ) 2 + ∆ 2 ]<br />

where є u is the single-particle energy. Note that the<br />

lowest excited state of a paired nucleus is at a higher<br />

energy than for the unpaired case<br />

• The pair gap parameter may be expressed: ∆ = G∑U u V u<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 150


Destruction of Pairing<br />

• Strong external<br />

influences may destroy<br />

the superfluid nature<br />

of the nucleus<br />

• In the case of a superconductor, a strong magnetic field<br />

can destroy the superconductivity: i the ‘Meissner Effect’<br />

• For the nucleus, the analogous role of the magnetic field<br />

is played by the Coriolis force, which at high spin, tends<br />

to decouple pairs from spin zero and thus destroy the<br />

superfluid pairing correlations<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 151


Coriolis Antipairing Effects<br />

• Classically the Coriolis force is given by:<br />

F Cor = -2m [ω х v]<br />

• Coriolis Antipairing (CAP): the magnitude of ∆ gradually<br />

and smoothly decreases and the nuclear moment of<br />

inertia (∝ ω 2 ) increases<br />

• Rotational Alignment: At spin ~ 12ħ, the Coriolis force is<br />

strong enough to break a specific pair of valence<br />

nucleons and align their individual id angular momenta along<br />

the rotation axis<br />

• High-j low-Ω particles are the most susceptible<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 152


Demise Of Pairing<br />

• CAP and rotational<br />

alignments diminish the<br />

magnitude of the<br />

nuclear pairing<br />

• Eventually the nucleus<br />

may enter an unpaired<br />

phase at high spin<br />

• In addition to static<br />

pairing, dynamic pairing<br />

correlations occur<br />

resulting from<br />

fluctuations in ∆<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 153


Backbending<br />

• The breaking of a specific nucleonic pair and the<br />

rotational alignment of the angular momenta leads to<br />

a characteristic ‘S’ shape of the nuclear Spin vs<br />

Frequency<br />

• A contribution of the nuclear spin now comes from single<br />

particles:<br />

I = R + J<br />

with ih J ≈ (j x,max + j x,max – 1) in accordance with the PEP<br />

• The nucleus ‘changes gear’, i.e. slows down while<br />

maintaining the angular momentum<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 154


Backbending<br />

• The moment of inertia<br />

increases with increasing<br />

rotational frequency<br />

• Around spin 10ħ a<br />

dramatic rise occurs<br />

(rotational frequency) 2<br />

• The characteristic ti ‘S’<br />

shape is called a backbend<br />

( 158 Er)<br />

• A more gradual increase is<br />

called an upbend ( 174 Hf)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 155


Backbending Movie<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 156


Backbending Demonstration<br />

This movie shows Mark<br />

Riley’s “backbending<br />

machine” built here in<br />

Liverpool<br />

This movie shows<br />

backbending in the 1960s<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 157


Band Crossings<br />

• Backbending can be<br />

interpreted as the<br />

crossing of two<br />

bands<br />

• The ‘G’ band (Ground<br />

state) is a fully<br />

paired configuration<br />

• The ‘S’ band (Super<br />

or Stockholm)<br />

contains one broken<br />

pair<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 158


Quadrupole Pairing<br />

• Higher order pairing correlations may occur leading to<br />

configuration-dependent pairing which depends on the<br />

relative orientation of nuclei orbits in a deformed<br />

potential<br />

• The Y 21 quadrupole component has the largest effect<br />

on the moment of inertia. The nuclear shape still has<br />

Y 20 symmetry and hence the quadrupole pairing is of a<br />

dynamical nature<br />

• The generalised pair creation operator is:<br />

P † λ † 뵆 = ∑ 〈u 1 |r Y λµ |u 2 〉 a u1† a<br />

†<br />

ū2<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 159


Neutron-Proton Pairing<br />

• The concept of<br />

superconductivity,<br />

related to like nucleon<br />

pairs coupled to spin I = 0<br />

and isospin T = 1, can be<br />

extended to neutron-<br />

proton pairs with T = 0<br />

• The greatest overlap<br />

occurs if the particles<br />

are in the same orbitals<br />

• Strong neutron-proton<br />

pairing can occur for<br />

nuclei with N = Z<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 160


Nucleon Pairing<br />

• The isovector (T=1)<br />

n-p pairing (c) is<br />

similar to the n-n (a)<br />

and p-p p( (b)pairing<br />

• The isoscalar (T=0)<br />

n-p pairing (d) is<br />

clearly different<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 161


7. Cranked Shell Model<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 162


Cranking<br />

• A rotation is externally imposed on a nucleus about the<br />

x axis. The Schrödinger equation is:<br />

iħ ∂Ψ lab /∂t = H lab Ψ lab<br />

• Ui Using the rotation operator<br />

R x = exp [-iI x ωt]<br />

with Ψ lab = R x Ψ int and H lab = R x H int R<br />

-1<br />

x , the SE within the<br />

intrinsic i i frame becomes:<br />

iħ ∂Ψ int /∂t = [H int – ħωI x ] Ψ int<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 163


The Routhian<br />

• The cranked Hamiltonian or Routhian (just the energy in<br />

the rotating frame) is then:<br />

H ω = H int – ħωI x<br />

where ωI x is analogous to the classical Coriolis and<br />

centrifugal forces<br />

• In terms of single-particle states, the cranking<br />

Hamiltonian is:<br />

H ω = ∑ ω i h (i) = ∑ i [h int (i) – ħωj x (i)]<br />

where j x (i) are the components of the nucleonic angular<br />

momenta on the rotation axis (x)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 164


Single-Particle Routhians<br />

• The single-particle Routhian can be evaluated by solving<br />

the eigenvalue envalue equation:<br />

h ω |ν ω 〉 = e ω ω ν |ν 〉<br />

where |ν ω 〉 are the single-particle eigenfunctions in the<br />

rotating frame<br />

• The Routhian is simply the energy in the rotating frame<br />

of reference: e<br />

ω<br />

ν<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 165


The Alignment<br />

• The alignment is just the expectation value of j x : 〈j x 〉<br />

and is equal to the (negative) differential of the<br />

Routhian with respect to rotational frequency, i.e.<br />

de νω /dω = -ħ 〈ν ω |j x |ν ω 〉<br />

• Those orbits with large j x values, and hence low Ω<br />

values, are most affected by the rotation, i.e. the<br />

Coriolis Corols and centrifugal centrfugal forces<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 166


Symmetries of Rotating Nuclei<br />

• The time-reversal (twofold) degeneracy of the ±Ω<br />

states is lifted by the ωj x term<br />

j x<br />

• Axially symmetric potentials exhibit invariance with<br />

respect to rotations ti by 180° (π) about the three<br />

principal axes (reflection symmetry)<br />

• However, the cranking Hamiltonian is only invariant for<br />

rotation of π about the x axis<br />

R x (π) = exp(-iπI x )<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 167


Signature Quantum Number<br />

• A rotation of 2π leaves the wavefunction unchanged,<br />

except for a possible phase factor (±1), i.e.<br />

R x2 (π) Ψ = r 2 Ψ = (-1) A Ψ with r 2 = ±1<br />

• The eigenvalues r of the rotation operator R x (π), called<br />

signature, are good quantum numbers, i.e. . constants of<br />

the motion<br />

• The signature exponent quantum number α is defined:<br />

r = exp(-iπα)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 168


Signature Quantum Numbers<br />

• The signature is:<br />

r = ±1 (even A) α = 0, 1<br />

r = ±i (odd A) α = ±½<br />

• The spins are restricted<br />

according to<br />

α = I mod 2<br />

Only good quantum<br />

numbers: π and α<br />

• Total signature: α tot = ∑α<br />

• Total parity: π tot = Ππ<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 169


Signature Partners<br />

• Signature is a good<br />

quantum number at high<br />

spin<br />

• A splitting between the<br />

α = 0/1 (even A) or α =<br />

±½ (odd A) states gives<br />

rise to two distinct<br />

‘signature partner’ bands<br />

• For an orbital with angular momentum j the ‘favoured’<br />

band has signature: α f = j mod 2<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 170


CSM Calculations<br />

• Including pairing the CSM Hamiltonian is:<br />

H νω = H sp – ∆(P † + P) – λÑ – ωI x<br />

with H<br />

ω = ∑h ω , H sp = ∑h sp , I x = ∑j x and the total energy<br />

E = ∑e νω , i.e. a sum of the single-particle Routhians,<br />

e νω = 〈h νω 〉<br />

• The parameters λ (-∂E/∂N) and ω (-∂E/∂I) can be<br />

considered as Lagrange multipliers needed to constrain<br />

the particle number and angular momentum, respectively:<br />

〈Ñ〉 = N and 〈I = √[ I(I+1) –K 2 x 〉 [ ]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 171


Quasiparticle Diagrams<br />

• Here is an<br />

example of a<br />

Woods-Saxon<br />

quasiparticle<br />

diagram<br />

• Label (π, α)<br />

A (C) (+,+½)<br />

B (D) (+,-½)<br />

E (G) (-,-½)<br />

F (H) (-,+½)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 172


Comparison to Experiment<br />

• The results of cranking<br />

calculations provide<br />

Routhians rather than<br />

energies<br />

• The experimental data<br />

must be transformed<br />

into the intrinsic frame<br />

to afford detailed<br />

d<br />

comparisons<br />

• Quantities are<br />

approximated as<br />

quotients of finite<br />

differences<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 173


Moments of Inertia<br />

• The frequency for γ ray 1 is:<br />

ω(I) = (1/ħ) dE(I)/dI x ≈ E γ1 /2ħ<br />

• Expressions for the moments of inertia are:<br />

I (1) (I) = ħ 2 { (2I-1) / (E(I+1) – E(I-1) } = ħ 2 (2I-1) /E γ1<br />

I (2) (I-1) = ħ 2 { 4 / (E γ1 – E γ2 ) = 4ħ 2 / ∆E γ<br />

• The dynamic moment of inertia corresponds to spin I-1<br />

and the associated frequency should be calculated l at I-1.<br />

In practice, the average frequency is used:<br />

ω(I-1) ( ≈ (E γ1 + E γ2 ) / 4ħ<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 174


Experimental Routhian and<br />

Alignment<br />

• The experimental Routhian may be expressed as:<br />

E ω expt(I) = ½{ E(I+1) + E(I-1) } – ω(I) I x (I)<br />

but we are interested in obtaining a quasiparticle<br />

Routhian and so must subtract the collective rotational<br />

energy:<br />

e’(I) = E ω expt(I) - E ω ref(I)<br />

• Similarly, il l we remove the collective spin to produce the<br />

quasiparticle alignment:<br />

i x (I) = I x (I) – I x,ref (I)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 175


Choice of Reference<br />

• The reference removes the collective effects of rotation<br />

and leaves the energies and spins solely from the valence<br />

quasiparticles, in the rotating frame<br />

• The reference can be obtained from the ground state<br />

band (zero quasiparticle, vacuum) of a (neighbouring)<br />

even-even nucleus<br />

• At low spin, it is found that I ∝ ω 2 . Hence a ‘variable<br />

moment of inertia’ (VMI) reference can be fitted,<br />

introducing ‘Harris Parameters’ I 0 , I 1 :<br />

I (1) ref = I 0 + I 1 ω 2<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 176


Harris Parameters<br />

• The Harris Parameters can be obtained by fitting:<br />

I x,ref (ω) = ω{I 0 + I 1 ω 2 } + i x<br />

to the reference band. Note that i x = 0 for the ground-<br />

state band of an even-even nucleus<br />

• The energy reference is then given by:<br />

E ω fdω 2 4 2 ref = -ħ∫I x,ref = -½ω I 0 - ¼ ω I 1 + ⅛ħ /I 0<br />

where the final term, an integration constant, ensures<br />

that the ground-state energy, E ω ref(I=0), is zero<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 177


Expt. Alignments and Routhians<br />

• Experimental results<br />

for N = 74 isotones are<br />

shown here<br />

• A clear ‘backbend’ is<br />

seen for 132 Ce while the<br />

heavier nuclei show<br />

‘upbends’<br />

• E and F correspond to<br />

proton h 11/2 orbitals,<br />

and these are the<br />

quasiparticles that align<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 178


Band Crossings<br />

• Band crossings can be classified by the rotational<br />

frequency at which they occur ω c and the gain in<br />

alignment ∆i x at the crossing<br />

• Experimentally: These quantities can be obtained by<br />

plotting e’ vs. ω and i x vs. ω<br />

• Theoretically: Crossing frequencies can be obtained<br />

from CSM quasiparticle diagrams.<br />

• The gain is alignment is given by the slopes of the<br />

interacting levels, E and F:<br />

∆i x = -(1/ħ) { de’ E /dω + de’ F /dω }<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 179


Signature Splitting<br />

• The signature splitting between the components of an<br />

orbital is the difference in excitation energy (or<br />

Routhian) at a fixed frequency, e.g.<br />

∆e’ FE (ω) = e’ F (ω) – e’ E (ω)<br />

• The magnitude of the signature splitting is related to<br />

the admixture of the Ω = ½ component in the<br />

wavefunction and is larger for low Ω values<br />

• The Coriolis interaction connects states with Ω = ±1<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 180


Staggering Parameter<br />

• One way to enhance<br />

signature effects is to<br />

plot the<br />

staggering parameter S(I)<br />

S(I) = E(I) – E(I-1) - ½[ E(I+1) – E(I) + E(I-1) – E(I-2)]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 181


Signature Inversion<br />

• In some odd-odd nuclei<br />

at low spin the<br />

signatures are the<br />

‘wrong way round’ i.e.<br />

the ‘favoured’ signature<br />

is energetically<br />

unfavoured !<br />

• At hih higher spin the<br />

signatures revert to<br />

their ‘expected’<br />

ordering<br />

Doubly odd La systematics<br />

• This is still not fully<br />

understood<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 182


8. Strutinsky Shell Correction<br />

• Shell correction energy<br />

• Cranked Nilsson Strutinsky Model<br />

• Total Routhian Surfaces<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 183


Shell Effects<br />

• A nuclear ‘property’ (e.g.<br />

binding energy) usually<br />

shows an irregular<br />

behaviour with mass A<br />

(or Z or N). It is made<br />

up of an oscillatory part<br />

∆E on top of a smooth<br />

part E smooth<br />

• Strutinsky’s s idea was to use the shell model to obtain<br />

∆E as the local variation from the average smoothed<br />

(shell model) value, but then to use the Liquid Drop<br />

Model to calculate l the real ‘smooth’ behaviour E smooth<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 184


Strutinsky Shell Correction<br />

• To obtain both the global<br />

(liquid drop) and local<br />

(shell model) variations<br />

with δ, Z and A, Strutinsky<br />

developed a method to<br />

combine the best<br />

properties of both models<br />

(a) Liquid drop: g F (e) = g AV (e)<br />

(b) and (c) show shell effects.<br />

A change in nuclear binding<br />

arises from: g AV (e) – g F (e)<br />

• He considered the<br />

behaviour of the level<br />

density g(e) in the two<br />

models and calculated the<br />

‘fluctuation’ energy<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 185


Level Density<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 186


Strutinsky Procedure<br />

• The nuclear (binding) energy is considered to have to<br />

have an oscillatory ypart ∆E shell, caused by yquantal<br />

effects (shell model), superposed upon a smoothly<br />

varying liquid drop part E LD :<br />

E = E LD + ∆E shell<br />

• Strutinsky proposed that only ∆E shell (‘microscopic’)<br />

should be calculated within the framework of the shell<br />

model, while the smoothly varying part E LD<br />

(‘macroscopic’) should be taken from the Liquid Drop<br />

Model<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 187


Strutinsky Procedure (cont)<br />

• Similarly, the total shell-model energy E SH does not<br />

vary smoothly and is composed of oscillatory and<br />

smooth parts:<br />

E SH =∑ 1A є i = Ĕ SH + ∆E SH<br />

• The real and smoothed level densities can be defined<br />

by g(є) and ğ(є), respectively<br />

• The number of levels between є and є + dє is given by<br />

g(є) dє and the level density is:<br />

g(є) = ∑ i δ(є – є i )<br />

where δ(є – є i ) is the Dirac delta dl function<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 188


Strutinsky Procedure (cont)<br />

• The particle number can be evaluated as:<br />

A = ∫ λ g(є) dє<br />

• The total shell-model energy and the smoothed part are<br />

then given, respectively, as:<br />

E SH = ∫ λ є g(є) dє<br />

and Ĕ SH = ∫ λ’ єğ(є) dє<br />

• Note that λ≠ λ’ ’ because of a smearing of the Fermi<br />

surface when calculating Ĕ SH<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 189


Strutinsky Procedure (cont)<br />

• The total energy of the nucleus may finally be written<br />

as:<br />

E = E LD + ∆E SH = E LD + [E SH – Ĕ SH ]<br />

where E LD is the macroscopic contribution and [E SH – Ĕ SH ]<br />

is the microscopic shell correction<br />

• Note that the shell correction can be positive or negative<br />

• Negative values give increased binding and stability<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 190


Shell Correction Energies<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 191


Superheavy Island<br />

• Shell effects,<br />

particularly<br />

hexadecapole<br />

deformation, can<br />

stabilise very heavy<br />

nuclei<br />

• Such superheavy nuclei<br />

only exist because of<br />

subtle quantum<br />

mechanical effects<br />

leading to a localised<br />

region (‘island’) of<br />

increased stability<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 192


Strutinsky Method for Spin<br />

• The Strutinsky technique can be extended to include<br />

rotation<br />

• We introduce another ‘level density’:<br />

g 2 (є) = ∑ i 〈j x 〉 i δ(є – є i )<br />

• The total t single-particle l energy is obtained from the<br />

cranking Hamiltonian as:<br />

where I = ∫ λ g 2 (є) dє<br />

E SP (I) = ∫ λ є g(є) dє + ħω I<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 193


Strutinsky Method for Spin<br />

• The smoothed energy is:<br />

Ĕ(є) = ∫ λ’ єğ(є) dє + ħω ğ 2 (є) dє<br />

• The cranked Nilsson Strutinsky method includes<br />

deformation (ε 2 , ε 4 , γ) and spin<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 194


Cranked Nilsson Strutinsky<br />

• The total energy is:<br />

E (ε 2 ,ε 4 ,γ, I) = E LD (ε 2 ,ε 4 ,γ, I) + ∆E SH (ε 2 ,ε 4 ,γ, I)<br />

• The macroscopic energy contribution is can be calculated<br />

from:<br />

E LD = E surf + E Coul + (9ħI) 2 /2I rig<br />

• This method usually ignores pairing correlations and is<br />

hence only valid for high-spin states (I > 20 ħ)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 195


Total Routhian Surfaces<br />

• This method is based on the Woods-Saxon potential and<br />

includes pairing. The total energy of a nucleus (Z, N) as a<br />

function of deformation β’ = (β 2 , β 4 , γ) is:<br />

E(ω,Z,N,β’) = E macro (ω,Z,N,β’) + ∆E shell (ω,Z,N,β’)<br />

• The total Routhian is:<br />

+ ∆E pair (ω,Z,N,β’)<br />

E(ω,Z,N,β’) = E(ω=0,Z,N,β’)<br />

+ [〈Ψ ω |H ω (Z,N,β’)|Ψ ω 〉 - 〈Ψ|H ω=0 (Z,N,β’)|Ψ〉]<br />

-½ω 2 [I macro (A,β’) - I Strut (N,Z,β’)]<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 196


Total Routhian Surfaces (cont)<br />

• The term E(ω=0,Z,N,β’) corresponds to the liquid-drop<br />

energy, the single-particle shell correction energy, and<br />

the pairing in energy at zero rotational ti frequency<br />

• ‘[〈Ψ ω |H ω (Z,N,β β’)|Ψ ω 〉 - 〈Ψ|H ω=0 (Z,N,β β’)|Ψ〉]’ is the<br />

change in energy due to rotation<br />

• The term ½ω 2 [I macro (A,β’) - I Strut (N,Z,β’)] represents a<br />

renormalisation of the LDM energy which is required<br />

due to unrealistically large proton and neutron radii<br />

used in some parameterisations of the Woods-Saxon<br />

potential<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 197


TRS Maps<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 198


9. Broken Symmetries<br />

• Reflection Asymmetry: Octupole Bands<br />

• Handedness: Chiral Bands<br />

• Magnetic Rotation: ti Shears Bands<br />

• Transitional Nuclei: Critical Points<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 199


Reflection Asymmetry<br />

• If a nucleus is ‘reflection asymmetric’ (i.e. the odd<br />

multipole deformation parameters are non-zero, e.g.<br />

β 3 ≠ 0 is the most important) then the nuclear<br />

wavefunction in its intrinsic frame is not an eigenvalue<br />

of the parity operator:<br />

Ψ 2 (x ,y ,z) ≠ Ψ 2 (-x, -y, -z)<br />

• If β 3 ≠ 0 for a nucleus it is said to possess octupole<br />

deformation<br />

• The deformation can however be static, 〈β 3 〉 ≠ 0, or<br />

dynamic, 〈β 3 〉 = 0 (oscillating octopule shape)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 200


Octupole Band Structures<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 201


Octupole Vibrations in 238 U<br />

• This nucleus<br />

shows three<br />

octupole<br />

vibrational<br />

bands with<br />

dff different K<br />

values<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 202


Parity Splitting<br />

• For a static octupole shape, the negative parity states<br />

are interleaved (midway between) with the positive<br />

parity states<br />

• A measure of such a feature is the ‘parity splitting’,<br />

defined as:<br />

δE = E(I) - -½[ E(I+1) + + E(I-1) + ]<br />

• This quantity generally decreases towards zero with<br />

increasing spin and suggests that rotation may stabilise<br />

the octupole shape<br />

• A similar quantity is the difference in alignment:<br />

∆i x = i x- - i<br />

+<br />

x<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 203


Octupole Vibration or Deformed?<br />

• For an octupole<br />

vibrational<br />

phonon coupled<br />

to the positiveparity<br />

states:<br />

∆i x = 3 ħ<br />

• For a static<br />

octupole<br />

deformation:<br />

∆i x = 0<br />

x<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 204


Reflection (A)symmetry<br />

1 band 2 bands 2 bands 4 bands<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 205


Electric Dipole Moment<br />

• In a nucleus with octupole<br />

deformation, the centre of<br />

mass and centre of charge<br />

tend to separate, creating a<br />

non-zero electric dipole<br />

moment<br />

• Bands of opposite parity<br />

connected by strong E1<br />

transitions occur<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 206


Enhanced E1 Transitions<br />

• In heavy nuclei, E1 strengths typically lie between<br />

10 -4 and 10 -7 Wu<br />

• In nuclei with octupole deformation, the E1 strengths<br />

can be much higher: 10 -3 –10 -2 Wu<br />

• The intrinsic dipole moment of an octupole deformed<br />

nucleus is:<br />

D 0 = C LD A Z e β 2 β 3<br />

with the liquid drop constant C LD = 0.0007 fm<br />

• In a Strutinsky type approach, macroscopic and<br />

microscopic effects can be considered and:<br />

D = D macro + D shell<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 207


Experimental Dipole Moments<br />

• Experimental values of D 0 can be obtained by measuring<br />

B(E1)/B(E2) ratios, related simply to γ-ray energies and<br />

intensities<br />

• The B(E1) reduced transition rate is:<br />

B(E1;I→I-1) = (3/4π) e 2 D 0<br />

2<br />

|〈 I i K i 1 0 | I f K f 〉| 2<br />

• The B(E2) reduced d transition rate is:<br />

B(E2;I→I-2) = (5/16π) e 2 Q<br />

2<br />

0 |〈 I i K i 2 0 | I f K f 〉| 2<br />

• Hence if Q 0 is known (e.g. from the quadrupole<br />

deformation β 2 ) then a value for D 0 can be extracted, i.e:<br />

D 0 = √[5B(E1)/16B(E2)] Q 0<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 208


Simplex Quantum Number<br />

• The only symmetries for a rotating reflection<br />

symmetric nucleus are parity p and signature r<br />

• For a reflection asymmetric shape (e.g. octupole) these<br />

are no longer good quantum numbers but the nucleus is<br />

invariant with respect to a combination of rotation of<br />

180° about the x axis (R(π)) and change of parity (P)<br />

• The ‘simplex’ operator is defined as:<br />

S = P R(π) -1<br />

with eigenvalues: s = -pr = ±i , ±1<br />

(p = s exp[iπI])<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 209


Parity Doublets<br />

• For K ≠ 0, four ∆I = 2 (E2) bands are formed based on<br />

states with K ± and (K+1) ±<br />

• The simplex quantum number can be used to classify<br />

these structures<br />

• For an even-even nucleus:<br />

s = +1 describes states (0 + ), 1 - , 2 + , 3 - , 4 + …<br />

s = -1 describes states (0 - ), 1 + , 2 - , 3 + , 4 - …<br />

• For an odd-A nucleus:<br />

s = +i describes states 1/2 + , 3/2 - , 5/2 + , 7/2 - ,…<br />

s = -i describes states 1/2 - , 3/2 + , 5/2 - , 7/2 + ,…<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 210


Parity Doublets in 223 Th<br />

• The nucleus 223 Th shows<br />

parity doublets<br />

• The two ∆I = 2 bands, shown<br />

to the left, are connected by<br />

strong E1 transitions and have<br />

simplex s = -i<br />

• The two ∆I = 2 bands, to the<br />

right, have simplex s = +i<br />

s = -i s = +i<br />

• M1 transitions also connect<br />

some of the bands<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 211


Octupole Magic Numbers<br />

• Octupole correlations<br />

occur between orbitals<br />

which differ in both<br />

orbital (l) and total (j)<br />

angular momenta by 3<br />

• Magic numbers occur at<br />

34, 56, 88 and 134<br />

• Nuclei with both proton<br />

and neutron numbers<br />

close to these are the<br />

best candidates to show<br />

octupole effects<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 212


Rotational Invariance<br />

• From Kris Starosta (Michigan State University)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 213


Space Inversion Invariance<br />

• From Kris Starosta (Michigan State University)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 214


Chirality (Handedness)<br />

• ‘I call any geometric figure, or<br />

group of points, chiral, and say<br />

it has chirality, if its image in a<br />

plane mirror, ideally realised,<br />

cannot be brought to coincide<br />

with itself’ Lord Kelvin 1904<br />

• Examples of chiral systems are<br />

found throughout h nature and in<br />

several disciplines of science<br />

• Axial vectors of angular momenta<br />

systems of opposite chirality are<br />

related by time reversal<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 215


Chiral Geometry<br />

• Spontaneous n chiral symmetry breaking can occur in<br />

triaxial doubly odd nuclei when there are three mutually<br />

perpendicular p spin vectors of differing lengths that can<br />

form a left-handed or right-handed configuration<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 216


Odd-Odd Mass 130 Nuclei<br />

• Region of triaxial shapes (x ≠ y ≠ z)<br />

• Consider the πhh 11/2 νhh 11/2 configuration<br />

n<br />

1. The proton Fermi surface lies at the bottom of the<br />

h 11/2 subshell: the proton single-particle j aligns along<br />

the short axis<br />

2. The neutron Fermi surface lies at the top of the h 11/2<br />

subshell: the neutron single-particle j aligns along the<br />

long axis<br />

3. The irrotational moment of inertia is largest for γ =<br />

30°: the core angular momentum aligns along the<br />

intermediate t axis<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 217


Irrotational Moments of Inertia<br />

• This diagram shows<br />

the variation of the<br />

moments of inertia I k<br />

as a function of the<br />

triaxiality parameter γ<br />

• For a prolate nuclear<br />

shape (γ = 0°), I 1 = I 2<br />

and I 3 = 0<br />

• For γ = 30° , I 2 reaches<br />

a maximum and this<br />

represents the ‘most<br />

collective’ shape<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 218


Chiral Operator<br />

• The chiral operator is a combination of time reversal and<br />

rotation by 180°: Ô = TR y (π)<br />

• The left-handed and right-handed systems are related<br />

to each other by this operator:<br />

|L〉 = Ô|R〉 and |R〉 = Ô|L〉<br />

• For a prolate nucleus, chiral symmetry is good: |R〉 = |L〉<br />

• However, for the triaxial odd-odd case: |R〉 ≠ |L〉<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 219


Restoration of Chiral Symmetry<br />

• Note that |R〉 and |L〉 are not solutions of the nuclear<br />

Hamiltonian in the lab frame and chiral symmetry must<br />

be restored by forming wavefunctions ns of the form<br />

(similar to the octupole case):<br />

|+〉 = (1/√2) [|R〉 + |L〉]<br />

|-〉 = (i/√2) [|R〉 -|L〉]<br />

• This leads to the doubling of the states and the<br />

occurrence of two (near) degenerate ∆I = 1 bands of<br />

the same parity<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 220


Chiral Twin Bands<br />

Two near degenerate ∆I = 1 bands of the same parity arise<br />

(cf octupole bands: two ∆I = 1 bands of opposite parity)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 221


Cranking Symmetries<br />

• If the nuclear spin I lies<br />

along one of the principal<br />

axes, one ∆I = 2 band arises<br />

• If the spin lies in the plane<br />

defined by two principal<br />

axes, one ∆I = 1 band arises<br />

• If the spin moves out of<br />

these planes, two degenerate<br />

∆I = 1 bands occur (chiral<br />

twins)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 222


Magnetic Rotation<br />

• In spherical lead nuclei,<br />

regular bands of intense M1<br />

transitions have been found<br />

• The valence proton and<br />

neutron orbitals lie<br />

perpendicular to each other<br />

and produce a magnetic<br />

moment vector that breaks<br />

the spherical symmetry of<br />

the system and allows<br />

‘magnetic’ rotation<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 223


Shears Mechanism<br />

• In magnetic rotation, higher<br />

angular momentum is generated<br />

by the reorientation of the<br />

neutron and proton spin<br />

vectors<br />

• Originally perpendicular, the<br />

vectors close like the blades<br />

of a pair of shears to generate<br />

the higher angular momentum<br />

states<br />

• The B(M1) strength decreases<br />

with increasing spin as µ ⊥<br />

decreases<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 224


Shears Systematics<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 225


Antimagnetic Rotation<br />

• Expected in weakly deformed nuclei<br />

• In 106 Cd the spin is generated by<br />

closing the πg<br />

-1<br />

9/2 vectors ( j<br />

-1<br />

π bottom diagram )<br />

• Each πgg 9/2 hole combines with one<br />

νh 11/2 particle forming a pair of backto-back<br />

shears<br />

• Note that the magnetic moment for<br />

this situation is zero, i.e. µ ⊥ = 0<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 226


Antimagnetic Rotation in 106 Cd<br />

• The yrast band appears to stop at 26 + with a measured<br />

drop in B(E2) values, or collectivity it (cf band termination)<br />

ti 2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 227


Transitional Nuclei<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 228


Interacting Boson Model<br />

• Bosons are constructed from fermion pairs<br />

• <strong>Nuclear</strong> collective excitations are described in terms<br />

of N interacting s (l = 0) and d (l = 2) bosons<br />

• Algebraic model based on U(6) group<br />

• Limits:<br />

• SU(3) rotational<br />

• U(5) vibrational<br />

• O(6) gamma-unstable<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 229


Critical Point Symmetries<br />

gamma soft<br />

vibrator<br />

rotor<br />

• The Casten Triangle<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 230


10. Band Termination<br />

• Favoured and unfavoured termination<br />

• Abrupt and smooth termination<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 231


Introduction: Band Termination<br />

• A deformed prolate nucleus can increase its angular<br />

momentum by collective rotation about an axis<br />

perpendicular to its symmetry axis<br />

• However, the nucleus is a many-body quantal system and<br />

such collective behaviour must have an underlying<br />

microscopic basis<br />

• There is a limiting angular momentum that a given<br />

configuration can generate<br />

• Successive alignments occur until all the valence particles<br />

are aligned and move in equatorial orbits giving the<br />

nucleus an oblate appearence<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 232


Band Termination: 158 Er<br />

neutron<br />

backbend<br />

proton<br />

backbend<br />

No more ( rays<br />

Gamma Ray Energy<br />

• A band ‘terminates’ when all valence particles outside a<br />

doubly magic (spherical) core are aligned<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 233


Band Termination in 158 Er<br />

• When the valence np protons and nn neutrons align, the<br />

total spin is: I = ∑<br />

np<br />

i j i (p) + ∑<br />

nn<br />

i j i (n)<br />

and the rotational band is said to ‘terminate’<br />

• At termination ti 158 Er can be considered d as a spherical<br />

146<br />

Gd core plus 4 protons and 8 neutrons, generating<br />

a maximum spin 46ħ<br />

• The configuration is: π(h 11/2 ) 4 ⊗ ν(i 13/2 ) 2 (h 9/2 ) 3 (f 7/2 ) 3<br />

• The terminating spin value of I max = 46 is generated as:<br />

(11/2+9/2+7/2+5/2) / / / + (13/2+11/2) / + (9/2+7/2+5/2) / / + (7/2+5/2+3/2)<br />

/ / 2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 234


Favoured Oblate States<br />

• Full termination represents the maximum alignment of<br />

all the valence particles outside a doubly magic core,<br />

consistent with the Pauli Exclusion Principle<br />

• Certain noncollective states representing maximal<br />

m<br />

alignment of a subset of the valence particles may be<br />

yrast leading to the observation of (energetically)<br />

favoured noncollective oblate states at a certain spin<br />

• Example: In<br />

157,158 Er energetically favoured states are<br />

seen at (I max –6),corresponding to two f 7/2 neutrons<br />

still being paired, i.e. they contribute 0 spin rather than<br />

7/2 + 5/2 = 6<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 235


Noncollective Oblate States in 121 I<br />

• Low-lying states are seen at<br />

I = 39/2 and 55/2<br />

• 121 I can be considered as a<br />

core ( 114 Sn) plus 3 valence<br />

protons and 4 valence<br />

neutrons<br />

• The configuration is<br />

π {h 11/2 g 7/22 } ν {h 11/24 }<br />

with maximum spin 55/2<br />

• If two of the protons remain<br />

paired we get the 39/2 state<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 236


Rigid Rotor Plot<br />

• The favoured nature of<br />

noncollective oblate<br />

states can be seen by<br />

plotting energy levels<br />

against spin<br />

• A rotating liquid-drop<br />

energy reference is<br />

subtracted:<br />

E LD = (ħ 2 /2I rig ) I(I+1)<br />

I ie 007 rig scaled to 158 Er, i.e. (ħ 2 /2I rig ) = 0.007 {158/A} 5/3 MeV<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 237


Shape Coexistence<br />

• Noncollective oblate<br />

states may coexist<br />

with collective<br />

rotational structures<br />

• In 119 I, collective<br />

structures are seen<br />

to the left and oblate<br />

states to the right<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 238


Smooth Band Termination<br />

• A novel type of ‘smooth’ band termination has been<br />

observed in several nuclei of the mass A = 110 region<br />

• Bands extend to very high energy (frequency) and the<br />

spacings between the γ rays increase (moment of<br />

inertia decreases)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 239


Drift Through the γ Plane<br />

• Smooth termination has<br />

been interpreted in the<br />

framework of the<br />

Cranked Nilsson<br />

Strutinsky method<br />

• It represents a gradual<br />

shape change from<br />

collective prolate (γ = 0°)<br />

to noncollective oblate<br />

(γ = 60°) over a wide spin<br />

range<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 240


Abrupt and Smooth Termination<br />

• The contrasting high-<br />

spin behaviour of<br />

117 Xe<br />

and 122 Xe is shown here<br />

• The rotational nature of<br />

122<br />

Xe abruptly breaks<br />

down above I = 22 ħ<br />

• The behaviour of 117 Xe<br />

appears more smooth<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 241


Termination Modes<br />

• Abrupt or<br />

favoured<br />

termination is<br />

shown at the<br />

top<br />

• Smooth or<br />

unfavoured<br />

termination is<br />

shown at the<br />

bottom<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 242


Termination Systematics<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 243


Band Termination in 152 Gd<br />

David Campbell<br />

Florida State<br />

University<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 244


Beyond Termination<br />

• At termination, several valence particles are aligned with<br />

the ‘rotation’ axis outside an ‘inert’ closed core (‘doubly<br />

magic’ spherical core)<br />

• How do we generate higher spin states?<br />

• We must break the core and form energetically<br />

expensive particle-hole excitations across the magic<br />

shell gaps of the core<br />

• A classic (state-of-the-art) case is the nucleus 157 Er<br />

which was studied with the Gammasphere spectrometer<br />

in Berkeley<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 245


157<br />

Er at High Spin<br />

• New high-energy h (15 (1.5-2.5 25 MeV), high-spin h transitions<br />

have been identified above I π = 87/2 - , 89/2 - and 93/2 +<br />

(new) terminating states<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 246


157<br />

Er Spectrum<br />

Several (weak) transitions are seen in the energy range<br />

1.0 – 2.5 MeV. Measured ∆I=2 transitions are labelled as<br />

Q (quadrupole) and ∆I=1 transitions as D (dipole)<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 247


157<br />

Er Rigid-Rotor Plot<br />

• Favoured oblate<br />

states are shown in<br />

gold<br />

• It costs a lot of<br />

energy to generate<br />

states of higher<br />

spin<br />

• The Z=64 proton<br />

core has to be<br />

broken to generate<br />

the highest spins<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 248


157<br />

Er Termination States<br />

• Relative to the 146 Gd<br />

core (Z=64, N=82), the<br />

[π{h 11/2 } 4 ]<br />

+<br />

16<br />

proton configuration is<br />

coupled to the neutron<br />

configurations (left) to<br />

produce energetically<br />

favoured noncollective<br />

oblate states<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 249


New Bands in 157,158 Er<br />

γ 5 spectra<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 250


Return of Collectivity in 158 Er<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 251


Triaxial SD Band in 158 Er<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 252


The End<br />

2/16/2009 Postgraduate <strong>Nuclear</strong> <strong>Models</strong> Course, Liverpool : E.S. Paul 253

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!