
Fundamentals of the Physics of Solids / 16-back-matter
.pdf
E.2 The Van Hove Formula for Cross Section |
659 |
the cross section is written as
|
d2σ |
= |
N |
|
|
|
|
|
AK S(K, ε/ ) . |
(E.2.19) |
|
|
dΩ dε |
2π |
Looking back at (E.2.5), the previous expression is factorized into three terms:
k mn 2, which is specific to the scattered particle alone, |UK |2 which de- k 2π 2
scribes the interaction, and S(K, ε/ ), the dynamical structure factor, which depends only on the internal dynamics of the scattering system. The above expression for the cross section is called the Van Hove formula. It is very generally valid, since only the applicability of the Born approximation has been assumed. Whether the correlation function of nuclear positions, electronic charge density, or spin density appears in the cross section is determined by the dominant interaction between the incoming particles and the sample.
In nonmagnetic scattering neutrons interact with nuclei at positions rm. Using the Fermi pseudopotential
U (r) = |
2π 2 |
amδ(r − rm) |
(E.2.20) |
mn |
for characterizing the interaction, the point-like character of the scattering center makes the Fourier transform independent of K:
UK = |
2π 2 |
(E.2.21) |
|
|
am . |
||
|
|||
|
mn |
|
For identical atoms the factor AK takes the K-independent form (k/k )|am|2, and the Fourier transform of the correlation function for the density of the nuclei appears in the cross section. From this formula information can be obtained about nuclear positions, about the equilibrium atomic positions in the crystal, and, through the motion of the nuclei, about phonons. This is discussed in more detail in Chapter 13.
Expression (E.2.10) for the cross section is valid also for the scattering of electrons by crystals. Through electronic position coordinates the spatial distribution of electrons bound to atoms appear in this case. If the electronic wavefunction is φi (r) in the initial and φf (r) in the final state, the transition matrix element is
k, i |Hint| f, k = drn drφi (r)φf (r)e−i(k−k )·rn U (r − rn) . (E.2.22)
When the relative coordinate is used, then, besides the Fourier transform of the potential, an additional factor
F (K) = drφi (r)φf (r)e−iK·r (E.2.23)

660 E Scattering of Particles by Solids
appears. If there is no change in the state of the electrons then F (K) is the Fourier transform of the electron density inside the atom. If the electron is around an atom at rm, that is, if the wavefunctions are in fact functions of r − rm, then
F (K)e−iK·rm |
(E.2.24) |
appears instead of F (K). The structure constant that is left behind after the separation of |F (K)| in the transition probability is now determined only by the structure and motion of the centers of mass of atomic electron clouds. Similarly to potential scattering, this leads to Bragg peaks and phonon peaks. From |F (G)| – the expression specifying the dependence of the Bragg peak intensities on the reciprocal-lattice vectors G – the spatial distribution of the electrons around the atom can be inferred.
As it was mentioned in Section 13.3, if the scattering amplitude is not the same for each scattering center then, in addition to coherent scattering, an additional incoherent contribution appears. This contribution accounts for interference that occurs not between rays scattered by di erent atoms but between rays scattered by the same atom at di erent times. Consequently the cross section of incoherent scattering is proportional to the Fourier transform of the self-correlation function of the atoms.
E.2.2 Magnetic Scattering
Neutrons can be scattered not only by nuclei but, on account of their magnetic moment, by electrons, too. The magnetic field of a neutron of magnetic moment μn at rn can be given by the vector potential A(r). Either of the three equivalent forms
A(r) = |
μ0 |
|
μn × (r − rn) |
|
|
||||
|
|
|
|
||||||
|
4π |
|r − rn|3 |
|||||||
= −4π |
μn |
× grad |r − rn| |
|||||||
|
|
μ0 |
|
1 |
|
||||
|
μ0 |
|
μn |
||||||
= |
|
curl |
|
|
|||||
4π |
|r − rn| |
can be used. The derived magnetic induction is
B(r) = curl A(r) .
(E.2.25)
(E.2.26)
If the magnetic moment of the jth electron, located at rj , is μj , then the interaction of such moments with the magnetic field can be given by the Hamiltonian
|
|
Hint = − μj · B(rj ) . |
(E.2.27) |
j
Considering a single term of this sum, and introducing the relative coordinate r = rj − rn, some simple algebra leads to the following formula for the interaction with the magnetic moment of the jth electron:

|
|
E.2 The Van Hove Formula for Cross Section 661 |
||||||||||
|
μ0 |
μj · curl curl |
μ |
|
|
|
|
|
|
|
||
Hint = − |
|
n |
|
|
|
|
|
|
|
|
|
|
4π |
|r| |
|
|
|
|
1 |
(E.2.28) |
|||||
|
μ0 |
1 |
+ |
μ0 |
|
|
|
|||||
= − |
|
μj · (μn · ) |
|
|
μj |
· μn |
2 |
|
. |
|||
4π |
r |
4π |
r |
For r = 0 the first term yields the well-known dipole–dipole interaction:
|
μ0 |
|
μ |
μ |
n |
|
3(μ |
j |
· |
r)(μ |
n · |
r) |
|
|
Hdipole = |
|
|
|
j · |
− |
|
|
|
. |
(E.2.29) |
||||
4π |
|
r3 |
|
|
|
|
r5 |
|
|
An extra contribution arises at r = 0; this is clearly seen when the relation
1 |
= −4πδ(r) |
|
2 r |
(E.2.30) |
is used in the second term. However, a similar singular contribution arises from the first term, too, as it was shown in Section 3.3.1. By repeating the derivation presented there, the interaction Hamiltonian can be given as the sum of two terms:
Hint = Hdipole + Hcontact , |
(E.2.31) |
where, in addition to the usual dipole–dipole interaction term, the Fermi contact interaction
Hcontact = − |
2μ0 |
μj · μn δ(r) |
(E.2.32) |
3 |
appears. This term is related to the overlap of the two moments, and plays a fundamental role in hyperfine interactions.
In magnetic scattering the spin direction can change for neutrons as well as for electrons bound to atoms. Therefore, besides the wave vector of the neutron and the spatial part of the electron wavefunction, the spin states of the neutron and electron must also be explicitly included in the wavefunction of the initial and final states. Denoting the spatial part of the electron wavefunction before
and after scattering by φi and φf, the initial and final states are |
|
||||||||||
|
|
|i, k = eik·rn |sn φi(rj )|σj , |
|
(E.2.33) |
|||||||
|
|
|
|
|
= eik ·rn |
s |
φ (r |
) σ |
. |
|
|
|
|
|f, k |
|
| |
n |
f |
j | j |
|
|
||
The matrix element between these initial and final states is |
|
||||||||||
k, i |Hint| f, k |
= |
σj , sn |
|
|
drn drj φi (rj )φf (rj )e−i(k−k )·rn Hint |
sn, σj . |
|||||
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(E.2.34) |
||
|
|
|
|
|
|
|
|
|
To evaluate this expression, the form (E.2.28) of the interaction is best used. Integration by parts of the first term gives
|
|
1 |
|
|
|
|
1 |
|
|||||
dre−iK·r |
|
μj · μn · |
|
|
= iK · μj |
dre−iK·r μn · |
|
|
|
||||
r |
r |
|
|||||||||||
|
|
|
|
= − K · μj |
|
|
1 |
e−iK·r |
|||||
|
|
|
|
|
K · μn |
dr |
|
||||||
|
|
|
|
|
r |
||||||||
|
|
|
|
|
|
|
4π |
|
|
|
|
|
|
|
|
|
|
|
= − K · μj |
K · μn |
|
. |
(E.2.35) |
||||
|
|
|
|
|
K2 |
662 |
E Scattering of Particles by Solids |
|
|
|
|
|
|
|
|
|
|
|
|
|
||||||||||
Making use of (E.2.30) in the second term, the matrix element becomes |
||||||||||||||||||||||||
k, i |
|Hint| |
f, k |
|
= μ |
, s |
|
(K |
· |
μ |
)(K |
· |
μ |
|
)K−2 |
− |
μ |
j · |
μ |
n |
s |
, σ |
|
F (K) , |
|
|
|
0 σj |
|
n |
|
|
j |
|
|
n |
|
|
|
n |
j |
(E.2.36) |
||||||||
where F (K) is the atomic form |
factor introduced in (8.1.27), |
|
|
|
|
|||||||||||||||||||
|
|
|
|
F (K) = |
drj φi (rj )φf (rj )e−iK·rj . |
|
|
|
|
|
(E.2.37) |
Expressed in terms of the unit vector e along the direction of K, the matrix element can be rewritten as
k, i |
f, k |
= |
|
μ |
|
σ , s |
μ |
|
e μ e s |
, σ F (K) |
||||||||
|Hint| |
|
= |
− |
μ0 |
|
sn |
μn sn |
|
σj |
|
|
j × |
j |
|
n |
j |
(E.2.38) |
|
|
0 |
j |
n |
|
n · |
× |
|
|
× |
j |
||||||||
|
|
|
− |
|
|
| | |
· | |
|
× |
|
|
| |
F (K) . |
|||||
|
|
|
|
|
|
|
|
|
|
|
|
e |
μ |
|
|
e |
σ |
If zm electrons of the mth atom participate in the magnetic scattering process, and their common wavefunction is Ψm(r1, r2, . . . rzm ), then the cross section contains the magnetic form factor
Fm(K) = |
|
drν Ψm(r1, r2, . . . rzm ) |
|
|
|
|
? |
|
|
|
(E.2.39) |
|
zm |
e−iK·rν (sν · Sm) Ψ (r , r , . . . r ) , |
|||
|
|
|
|
|
|
|
× |
|
m 1 2 |
zm |
|
|
S(S + 1) |
|
|||
|
ν=1 |
|
|
|
|
which characterizes the spatial distribution of “magnetic” electrons. Here Sm is the total spin of the mth atom; the associated magnetic moment μm is given by
μm = gμBSm . |
(E.2.40) |
This total atomic moment appears in the matrix element, too,
k, i |Hint|f, k = −μ0 sn|μn|sn · σj | (e × [μm × e]) |σj Fm(K) . (E.2.41)
In expressions (E.2.37) and (E.2.39) for the atomic form factor rj and rν denote the absolute position of the electron. Changing to the relative coordinate
with respect to the position rm of the atom, the substitution |
|
Fm(K) → Fm(K)e−iK·rm |
(E.2.42) |
allows for the separation of a phase factor due to the actual atomic position, and the obtained Fm(K) is the true atomic magnetic form factor.
The cross section is obtained by summing over all atoms. The calculations below are for the simple case when the incident neutron beam is unpolarized and the polarization of the scattered beam is not measured, either. By characterizing the probabilities of the possible spin states of the neutron by the density matrix ρn, summing over final-state spins, and using the density matrix ρ for specifying the initial spin state of the atoms, the cross section is given by
E.2 The Van Hove Formula for Cross Section |
663 |
d2σ |
= |
k |
|
mnμ0 |
2 |
mn sn,sn |
σl },{σl } sn|ρn|sn {σl}|ρ|{σl} |
|
||
dΩ dε |
k |
2π 2 |
|
|
|
|||||
|
|
|
|
|
|
|
|
{ |
|
|
|
|
|
|
F |
(K)e−iK·rm F (K)eiK·rn |
|
||||
|
|
|
× m |
|
|
|
|
n |
|
|
|
|
|
× sn|μn|sn · {σl}| (e × [Sm × e]) |{σl} |
(E.2.43) |
||||||
|
|
|
× sn|μn|sn · {σl }| (e × [Sn × e]) |{σl} δ(ε − Ef + Ei) . |
Separating the factors that contain the neutron magnetic moment, and expressing the moment with the neutron g-factor, the nuclear magneton and the spin as
|
μn = gnμNsn , |
|
(E.2.44) |
|
we have |
|
|
|
|
sn,sn |
sn|ρn|sn sn|μnα|sn sn|μnβ |sn = sn|ρn|sn sn|μnαμnβ |sn |
|||
sn |
|
|
||
|
|
|
|
|
|
= 41 |
gnμN 2δαβ |
sn|ρn|sn |
|
|
|
|
sn |
|
|
|
|
|
|
|
= 41 |
gnμN |
2δαβ . |
(E.2.45) |
Now the integral representation can be used for the delta function of energy conservation, and the time dependence can be absorbed in one of the operators, as was done for potential scattering. Then, by making use of the completeness relation to sum over the intermediate spin states of the electron
system, by exploiting the identity |
|
(e × [a × e]) · (e × [b × e]) = a · b − (a · e) (b · e) , |
(E.2.46) |
and finally by eliminating the atomic moments in favor of the spin variables, one obtains
d2σ |
= |
1 |
|
|
|
2 |
|
2 k |
|
mnμ0 |
|
2 |
mn,αβ(δαβ − eαeβ ) |
|
|||
dΩ dε |
4 |
gnμN |
|
gμB k |
2π 2 |
|
|
||||||||||
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
× |
F |
(K)F (K)e−iK·rm eiK·rn |
(E.2.47) |
||||||||||||
|
|
|
m |
|
|
|
n |
|
|
|
|
|
|||||
|
|
|
|
|
|
|
∞ |
Smα (t)Snβ (0) . |
|
||||||||
|
|
× 2π |
dt eiεt/ |
|
|||||||||||||
|
|
|
|
1 |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
−∞ |
|
|
|
|
Atomic motions were ignored in the previous derivation. When due care is taken of them, atomic positions also appear in the thermal averages, leading to
664 E |
Scattering of Particles by Solids |
|
|
|
|
|||||||||||||||
|
d2σ |
= |
1 |
gnμN |
2 |
|
|
2 k |
|
mnμ0 |
|
2 |
mn,αβ(δαβ − eαeβ ) |
|
||||||
|
dΩ dε |
4 |
|
gμB k |
2π 2 |
|
|
|||||||||||||
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
× |
F |
m |
(K)F (K)e−iK·Rm eiK·Rn |
(E.2.48) |
|||||||||||||
|
|
|
|
|
|
∞ |
n |
|
|
|
|
|
||||||||
|
|
|
|
|
|
|
|
|
dt eiεt/ |
|
e−iK·um (t)Smα (t)eiK·un Snβ (0) . |
|
||||||||
|
|
|
× 2π |
|
|
|
||||||||||||||
|
|
|
|
|
|
|
1 |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
−∞ |
|
|
|
|
|
|
|
If spin–phonon interactions are neglected, the averaging procedures can be performed separately for spins and atomic positions. When calculating the contribution of magnetic scattering, atoms are assumed to occupy their equilibrium positions, and so
d2σ |
= |
1 |
gnμN |
2 |
gμB |
2 k |
|
mnμ0 |
2 |
mn Fm(K)Fn (K)e− |
iK |
(R R ) |
||||||
dΩ dε |
4 |
|
|
k |
2π 2 |
|
· |
m − n |
||||||||||
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
× αβ |
|
|
|
|
|
|
|
∞ |
dt eiεt/ |
|
Smα (t)Snβ (0) . |
(E.2.49) |
|||
|
|
|
(δαβ − eαeβ ) 2π |
|
||||||||||||||
|
|
|
|
|
|
|
|
|
|
1 |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
−∞ |
|
|
|
The spatial distribution of the moment is characterized by the atomic magnetic form factor, while the spin correlation function describes spin dynamics.
References
1.E. Balcar and S. W. Lovesey, Theory of Magnetic Neutron and Photon Scattering, Oxford University Press, Oxford (1989).
2.S. W. Lovesey, Theory of Neutron Scattering from Condensed Matter, Vol. I: Nuclear Scattering, Clarendon Press, Oxford (1986).
3.S. W. Lovesey, Theory of Neutron Scattering from Condensed Matter, Vol. II: Polarization E ects and Magnetic Scattering, Clarendon Press, Oxford (1986).
4.W. Marshall and S. W. Lovesey, Theory of Thermal Neutron Scattering, Clarendon Press, Oxford (1971).

F
The Algebra of Angular-Momentum and Spin
Operators
In this Appendix we shall first list some basic relations for quantum mechanical angular-momentum operators, and then turn to the special properties of spin operators.
F.1 Angular Momentum
It was mentioned in Appendix D that the dimensionless orbital angular momentum operator defined via L = r × p is the generator of the transformations of the continuous rotation group. We shall discuss this relation in more detail below, and then present the irreducible representations of the rotation group as well as the addition theorem for angular momenta.
F.1.1 Angular Momentum and the Rotation Group
The set of all possible rotations around any axis through the origin form a continuous group, the full rotation group SO(3). The infinitesimal rotation through angle δα around the axis characterized by the unit vector n is denoted by Cn (δα); in this operation the end point of the vector r is displaced by
δr = δα n × r . |
(F.1.1) |
In the space of the functions ψ(r) an operator O Cn(δα) is associated with this transformation; it takes function ψ(r) into
O Cn(δα) ψ(r) = ψ Cn−1(δα)r = ψ(r − δr) = [1 − δα n · (r × )]ψ(r) ,
that is, the rotation operator can be written as
O Cn(δα) = 1 − δα n · (r × ) = 1 − iδα(n · L) .
(F.1.2)
(F.1.3)
666 F The Algebra of Angular-Momentum and Spin Operators
It then follows that for a rotation through an arbitrary angle α
O Cn (α) = e−iα(n·L) .
For conciseness we shall use the commoner notation
Rn(α) ≡ O Cn (α) .
(F.1.4)
(F.1.5)
The components of the operator L satisfy the commutation relations
[Lx, Ly] = iLz , [Ly , Lz ] = iLx , [Lz , Lx] = iLy , |
(F.1.6) |
which can be concisely written as |
|
|
|
[Lα, Lβ ] = i εαβγ Lγ , |
(F.1.7) |
γ |
|
or as |
|
L × L = iL , |
(F.1.8) |
where εαβγ is the Levi-Civita tensor. We shall call any operator J an angularmomentum operator if it appears as a generator for rotations and its components satisfy the commutation relation
[Jα, Jβ ] = i |
|
(F.1.9) |
εαβγ Jγ , |
γ
even if the transformation does not concern a complex function ψ(r) but a spinor or a vector function.
It is not always convenient to give the rotation operator in this form. Namely, if the rotation axis n is specified by its polar angles θ and ϕ, the term
−iα(sin θ cos ϕJx + sin θ sin ϕJy + cos θJz ) |
(F.1.10) |
appears in the exponent – and since the components of the angular-momentum operator do not commute, this expression is more di cult to evaluate. Therefore rotations are customarily characterized using the Euler angles. Several widely used conventions exist; we shall adopt the one in which a rotation through φ around the z-axis is followed by a rotation through ϑ around the new y -axis, and finally by a rotation through ψ around the even newer z -axis:
Rn(α) = R(φ, ϑ, ψ) = Rz (ψ)Ry (ϑ)Rz (φ) . |
(F.1.11) |
In terms of the generators of rotations,
R(φ, ϑ, ψ) = e−iψJz e−iϑJy e−iφJz . |
(F.1.12) |
An even simpler form is obtained when it is recognized that the y -axis is obtained from the y-axis via a rotation around the z-axis through angle φ, that is
|
|
|
|
F.1 Angular Momentum |
667 |
|
J |
= R |
|
(φ)J R−1 |
(φ) , |
(F.1.13) |
|
y |
|
z |
y z |
|
||
and therefore |
|
|
|
|
|
|
e−iϑJy = e−iφJz e−iϑJy eiφJz . |
(F.1.14) |
Using the corresponding expression for the operator of the rotation around the z -axis,
R(φ, ϑ, ψ) = e−iφJz e−iϑJy e−iψJz |
(F.1.15) |
is finally obtained.
A possible representation of the rotation matrices can be easily obtained from determining their action on the function ψ(r) = r. For a rotation around the z-axis through angle φ,
|
cos φ |
sin φ 0 |
|
|
||
− |
sin φ |
cos φ 0 |
(F.1.16) |
|||
Rz (φ) = |
|
. |
||||
|
|
0 |
0 |
1 |
|
|
Likewise, for a rotation around the y-axis through angle ϑ,
Ry(ϑ) = |
|
0 1 |
− |
0 |
|
, |
(F.1.17) |
|
|
cos ϑ 0 |
|
sin ϑ |
|
|
|
|
sin ϑ 0 |
cos ϑ |
|
|
while for a rotation around the z-axis through angle ψ,
|
cos ψ |
sin ψ 0 |
|
|
||
− |
sin ψ cos ψ 0 |
(F.1.18) |
||||
Rz(ψ) = |
|
. |
||||
|
|
0 |
0 |
1 |
|
|
Owing to the transformations of the axes, rotations appear in the reverse order in (F.1.15):
R (φ, ϑ, ψ) = |
− sin φ cos φ 0 |
|
0 1 |
− |
0 |
|
− sin ψ cos ψ 0 |
. |
||
|
cos φ |
sin φ 0 |
|
cos ϑ 0 |
|
sin ϑ |
|
cos ψ |
sin ψ 0 |
|
|
|
|
|
|
|
|
|
|||
|
0 |
0 1 |
|
sin ϑ 0 |
cos ϑ |
|
0 |
0 1 |
|
(F.1.19) Thus we have obtained the representation of the rotation group by 3 × 3 orthogonal matrices of unit determinant, as R T = R −1. Spatial rotations in three dimensions are indeed isomorphic to the group SO(3).
F.1.2 The Irreducible Representations of the Rotation Group
To determine further representations of the rotation group, we shall exploit the fact that J 2 and Jz commute, and therefore have a common set of eigenfunctions. Denoting the eigenvalues of J 2 by j(j + 1) and those of Jz by m,

668 F The Algebra of Angular-Momentum and Spin Operators
it follows from the properties of the operators that j can take only positive, while m any integer or half-integer1 value,
j = 0, 1/2, 1, 3/2, 2, . . .,
m = 0, ±1/2, ±1, ±3/2, ±2, . . ., |
(F.1.20) |
|
|
but for a given j the eigenvalues of m can only be |
|
m = −j, −j + 1, . . . , j . |
(F.1.21) |
The angular-momentum operator can be given in the space of states with quantum numbers j, m – denoted as |j, m – by (2j + 1) ×(2j + 1) matrices. To determine the matrix element, the raising and lowering operators J± = Jx±iJy are introduced; they increase and decrease the component Jz by unity:
|
J+|j, m = [j(j + 1) − m(m + 1)]1/2|j, m + 1 |
|
|
||||||||||||
|
|
|
|
= (j − m)1/2(j + m + 1)1/2|j, m + 1 , |
|
(F.1.22) |
|||||||||
|
J−|j, m = [j(j + 1) − m(m − 1)]1/2|j, m − 1 |
|
|
||||||||||||
|
|
|
|
= (j + m)1/2(j − m + 1)1/2|j, m + 1 , |
|
|
|||||||||
in line with the relations |
|
|
|
|
|
|
|
|
|
|
|||||
|
|
|
|
|
J+J− = J 2 − Jz (Jz − 1) , |
|
|
|
|
||||||
|
|
|
|
|
J−J+ = J 2 − Jz (Jz + 1) , |
|
|
|
(F.1.23) |
||||||
|
|
|
|
|
J 2 = J |
2 + |
1 |
(J J + J J ) . |
|
|
|
||||
|
|
|
|
|
|
z |
2 |
|
+ − |
− + |
|
|
|
|
|
For j = 1/2 the generators are 2 × 2 matrices of the form |
|
|
|||||||||||||
2 |
1 0 |
|
2 |
i 0 |
|
2 |
0 −1 |
|
|
||||||
Jx = 1 |
0 1 |
, |
Jy = 1 |
0 |
−i |
, |
Jz = 1 |
|
1 0 |
, |
(F.1.24) |
||||
while for j = 1 |
|
|
, Jy = √ |
|
i |
0 −i , |
|
|
|
. |
|
||||
Jx = √ |
|
1 0 1 |
|
Jz = |
0 0 0 |
(F.1.25) |
|||||||||
1 |
|
0 1 0 |
|
1 |
|
|
0 |
−i 0 |
|
|
1 0 0 |
|
|
||
|
|
0 1 0 |
|
|
0 i 0 |
|
0 0 −1 |
|
|||||||
|
2 |
2 |
|
|
It is straightforward to show that a suitable unitary transformation can take these matrices into (D.1.72), that is, the two forms are equivalent.
Expressing the rotation matrices in terms of these generators, as in
(F.1.15), the matrix elements in the space of states |j, m |
are |
|
||
Rmmj (φ, ϑ, ψ) = |
j, m e−iφJz e−iϑJy e−iψJz j, m |
. |
(F.1.26) |
|
|
|
|
|
|
1By half-integer or half-odd-integer numbers we mean numbers of the form (2n + 1)/2, where n is integer.