Download as pdf or txt
Download as pdf or txt
You are on page 1of 3

The Cranking Model for Rotations

Giacomo Accorto
August 25, 2023

Almost every fully microscopic theory of nuclear rotations are based in some way on the crank-
ing model. The model was first introduced by Inglis in a semiclassical way, but it can be quantized,
at least in the limit of large deformations and not too strong K-admixtures (K ≪ 1). The crank-
ing model does not introduce redundant variables, and offers an insight on the dynamics of the
rotational motion. It describes the collective angular momentum as the sum of single-nucleon an-
gular momenta: The collective and the single-particle rotation are treated on an equal footing. The
model holds for very large angular momenta, where classical arguments can be made. The theory
is nonlinear, and only in the limit of small angular momenta one can linearize it via perturbation
theory. The obtained wave functions are not eigenstates of the angular momentum operators, and
one should use projection techniques to get the wave functions in the laboratory frame.

1 Semiclassical derivation
The main assumption of the model is that if one introduces a frame of reference that rotates with
constant angular velocity ω around a fixed axis, the motion of the nucleons in such system is
supposed to be rather simple. There, nucleons can be treated as independent particles moving in
an average, rotating potential well.
If one completely neglects residual interactions and considers a single-particle potential V with
fixed shape, that rotates in space, then the single-particle Hamiltonian reads

p2
h(t) = + V (⃗r, t) (1)
2m
with the corresponding the Schrödinger equation


h(t)ψ(t) = ih̄ ψ(t). (2)
∂t
Introducing spherical coordinates with respect to the rotation axis allows one to represent the time
dependence of the potential well as

V (⃗r, t) = V (r, θ, φ − ωt, 0) (3)

where V (⃗r, 0) is the potential at t = 0. Thus, V is time dependent if it depends on the coordinate
φ, namely V should not have axial symmetry around the rotational axis. In fact, for a quantum

1
mechanical system no collective rotation is possible around an axis of symmetry. One can introduce
a unitary transformation to get rid of the explicit time dependence of the potential,

U = eiωlt (4)
with ω · ⃗l = −ih̄ω · ∂/∂φ called Coriolis term. The single-particle Hamiltonian evolves in time
accordingly
U h(t)U = h(0). (5)
With the definition ψ̃ = U ψ,
˙ ⃗ ψ̃.
ih̄ψ̃ = ih̄U ψ̇ + ih̄U̇ ψ = (h(0) − ω I) (6)

The effective Hamiltonian hω ≡ h(0) − ω I⃗ is time-independent and one can solve the standard
eigenvalue problem
⃗ ψ̃ = ϵ′ ψ̃
hω ψ̃ = (h(0) − ω I) (7)
ω
and obtain the energies of the original Hamiltonian by calculating
ϵω = ⟨ψ|h(t)|ψ⟩ = ⟨ψ̃|h(0)|ψ̃⟩ = ϵ′ω + ω ⟨ψ̃|⃗l|ψ̃⟩ (8)
In the specific case of rotations, hω is identical to the Hamiltonian as seen from the rotating frame
of reference [1].
One normally chooses the orientation of the rotation axis to be parallel to the x-axis, that is
perpendicular to the axis of symmetry, which for ω = 0 is the z-axis. The many-body Hamiltonian
of the cranking model is given by
A
X
Hω = H − ωJx = h(i)
ω (9)
i=1
PA (i)
with Jx − i=1 jx and ⃗j = ⃗l + ⃗l. A diagonalization provides the ground state wave function |Φω ⟩,
that is a Slater determinant. In order to minimize the total energy, single-particle levels are filled in
the usual way to obtain the yrast levels.
In the laboratory system, from Eq. (8)
E(ω) = ⟨Φω |H|Φω ⟩ = ⟨Φω |Hω |Φω ⟩ + ω ⟨Φω |Jx |Φω ⟩ . (10)
Since the energy must be independent of the sign of the velocity,
1 2
E(ω) = E(0) + ω + ... (11)
2J1
and since ⟨Φ0 |Jx |Φ0 ⟩ = 0,
J(ω) = ⟨Φω |Jx |Φω ⟩ = J2 ω + . . . (12)
The constants J1 and J2 are equal due to the fact that E(ω) is the lowest eigenvalue of Hω ,
δ ⟨Φω |Hω |Φω ⟩ = 0 (13)
d d

⟨Φω′ |H|Φω′ ⟩ − ω ′ ⟨Φω′ |Jx |Φω′ ⟩ = 0 (14)
dω dω
1 d ′ d
J1 = E(ω ) = J(ω ′ ) = J2 . (15)
ω dω ′ ′
ω =0 dω ′
ω ′ =0

2
Moreover, one also gets that ω = dEdJ . The value of the angular velocity can be derived by including
p
the zero-point oscillations semiclassically by requiring J = I(I + 1). In the first order
p
I(I + 1)
ω= (16)
J1
and the yrast bands are
1
E(I) = E(0) + I(I + 1). (17)
2J1
Deviations from the I(I + 1)-rule are expected at high velocities. The moment of inertia is defined
by J = ωJ .

2 The Cranking Formula


For a pure I(I + 1) spectrum, one only needs to calculate the moment of inertia, which can be
determined simply from the 2+ state. It is then possible to apply perturbation theory for small I-
values. In the unperturbed system |Φ0 ⟩ of a deformed potential, that is filled up to the Fermi level,
one calls holes (i, i′ ) the levels below the FermiPenergy, and particles (m, m′ ) the levels above it.
The shell model basis consists of |Φ⟩ = |Φ0 ⟩ + im a†m ai |Φ0 ⟩ + 2p-2h states + . . . and since the
perturbation ωJx is a one-body operator, it only excites one ph-pair at a time. Up to linear order in
ω, the perturbed wave function reads
X ⟨mi|Jx |Φ0 ⟩
|Φ⟩ = |Φ0 ⟩ + ω a†m ai |Φ0 ⟩ (18)
im
ϵ m − ϵi

and the expectation value of Jx is


X | ⟨mi|Jx |Φ0 ⟩ |2
J = ⟨Φ|Jx |Φ⟩ = 2ω . (19)
im
ϵm − ϵi
Using Eq. (12) one derives the Inglis formula for the moment of inertia,
X | ⟨m|Jx |i⟩ |2
JInglis = 2 (20)
im
ϵm − ϵi
which usually produces values very close to the rigid body solution.
Experimental moments of inertia are generally a factor 2 to 3 smaller then their rigid body
values. It is known that this is due to a superfluid slippage of some nucleons as the nucleus rotates.
Including residual two-body interactions, especially pairing correlations, would lower these values.
Finally, at higher angular momenta one may use perturbation theory in higher order and include the
effects of a residual interaction. The most important effect that arises is the Coriolis anti-pairing,
that is the disruption of pp-pairs due to the centrifugal forces. In any case, it is only at very high
spin states (I > 30) that one can expect the pairing correlations to vanish.

References
[1] J. G. Valatin. Nucleon Motion in a Rotating Potential. Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences, 238(1212):132–141, 1956.

You might also like