Chapter 4: Crystal Lattice Dynamics
Chapter 4: Crystal Lattice Dynamics
Debye
Contents
1
4 Theory of Neutron Scattering 26
4.1 Classical Theory of Neutron Scattering . . . . . . . . . . 27
4.2 Quantum Theory of Neutron Scattering . . . . . . . . . . 30
4.2.1 The Debye-Waller Factor . . . . . . . . . . . . . . 35
4.2.2 Zero-phonon Elastic Scattering . . . . . . . . . . 36
4.2.3 One-Phonon Inelastic Scattering . . . . . . . . . . 37
2
A crystal lattice is special due to its long range order. As you ex-
plored in the homework, this yields a sharp diffraction pattern, espe-
cially in 3-d. However, lattice vibrations are important. Among other
things, they contribute to
3
ionic mass:
m
¿1 (1)
M
which allows us to derive an accurate theory.
Due to Newton’s third law, the forces on the ions and electrons are
comparable F ∼ e2 /a2 , where a is the lattice constant. If we imagine
that, at least for small excursions, the forces binding the electrons and
the ions to the lattice may be modeled as harmonic oscillators, then
F ∼ e2 /a2 ∼ mωelectron
2 2
a ∼ M ωion a (2)
4
n’th unit cell
atom α
rα sα
rn
rnα Rnα = r n + rα + sα
O
Figure 1: Nomenclature for the lattice vibration problem. sn,α is the displacement of
the atom α within the n-th unit cell from its equilibrium position, given by rn,α =
rn + rα , where as usual, rn = n1 a1 + n2 a2 + n3 a3 .
We will define the zero potential such that when all Rn are at their
equilibrium positions, φ = 0. Then
X Pn2
H= + φ(R1 , · · · RN ) (5)
n 2M
Typical lattice vibrations involve small atomic excursions of the or-
der 0.1A or smaller, thus we may expand about the equilibrium position
of the ions.
∂φ 1 ∂ 2φ
φ({rnαi + snαi }) = φ({rnαi }) + snαi + snαi smβj (6)
∂rnαi 2 ∂rnαi ∂rmβj
The first two terms in the sum are zero; the first by definition, and
the second is zero since it is the first derivative of a potential being
5
evaluated at the equilibrium position. We will define the matrix
∂ 2φ
Φmβj
nαi = (7)
∂rnαi ∂rmβj
From the different conservation laws (related to symmetries) of the
system one may derive some simple relationships for Φ. We will discuss
these in detail later. However, one must be introduced now, that is,
rα sα
Figure 2: Since the coefficients of potential between the atoms linked by the blue lines
(m−n)βj
(or the red lines) must be identical, Φmβj
nαi = Φ0αi .
(m−n)βj ∂ 2φ
Φmβj
nαi = Φ0αi = (8)
∂r0αi ∂r(n−m)βj
ie, it can only depend upon the distance. This is important for the next
subsection.
6
1.1 The Equation of Motion
From the derivative of the potential, we can calculate the force on each
site
∂φ({rmβj + smβj })
Fnαi = − (9)
∂snαi
so that the equation of motion is
−Φmβj
nαi smβj = Mα s̈nαi (10)
If there are N unit cells, each with r atoms, then this gives 3N r equa-
tions of motion. We will take advantage of the periodicity of the lattice
by using Fourier transforms to achieve a significant decoupling of these
equations. Imagine that the coordinate s of each site is decomposed
into its Fourier components. Since the equations are linear, we may just
consider one of these components to derive our equations of motion in
Fourier space
1
snαi = √ uαi (q)ei(q·rn −ωt) (11)
Mα
where the first two terms on the rhs serve as the polarization vector
for the oscillation, uαi (q) is independent of n due to the translational
invariance of the system. In a real system the real s would be composed
of a sum over all q and polarizations. With this substitution, the
equations of motion become
1
ω 2 uαi (q) = q Φmβj
nαi e
iq·(rm −rn )
uβj (q) sum repeated indices .
Mα Mβ
(12)
7
t=0
t = ∆t
t = 2 ∆t
Figure 3: uαi (q) is independent of n so that a lattice vibration can propagate and
respect the translational invariant of the lattice.
(m−n)βj
Recall that Φmβj
nαi = Φ0αi so that if we identify
βj 1 1
Dαi = q Φmβj
nαi e
iq·(rm −rn )
= q Φpβj
0αi e
iq·(rp )
(13)
Mα Mβ Mα Mβ
where rp = rm − rn , then the equation of motion becomes
βj
ω 2 uαi (q) = Dαi uβj (q) (14)
or
µ ¶
βj βj
Dαi − ω 2 δαi uβj (q) = 0 (15)
³ ´
which only has nontrivial (u 6= 0) solutions if det D(q) − ω 2 I = 0.
For each q there are 3r different solutions (branches) with eigenvalues
ω (n) (q) (or rather ω (n) (q) are the root of the eigenvalues). The de-
pendence of these eigenvalues ω (n) (q) on q is known as the dispersion
relation.
8
basis
M1 M2 α=1 α=2 f
Φn,1 n,2
n,1 = Φn,2 = 2f (20)
9
Φn,2 n,1 n−1,2
n,1 = Φn,2 = Φn,1
n+1,1
= Φn,2 = −f . (21)
10
2.0
optical mode
1.5
acoustic mode ω+
1.0
ω
ω-
0.5 ω ~ ck
0.0
-4 -2 0 2 4
q
Figure 5: Dispersion of the linear chain of oscillators shown in Fig. 4 when M 1 = 1,
M2 = 2 and f = 1. The upper branch ω+ is called the optical and the lower branch is
the acoustic mode.
As a result, the + mode is quite flat; whereas the − mode varies from
zero at the Brillouin zone center q = 0 to a flat value at the edge of the
zone. This behavior is plotted in Fig. 5.
It is also instructive to look at the eigenvectors, since they will tell
us how the atoms vibrate. Let’s look at the optical mode at q = 0,
q
ω+ (0) = 2f /µ. Here,
√
2f /M1 −2f / M1 M2
D=
√
. (27)
−2f / M1 M2 2f /M2
11
Eigenvectors are non-trivial solutions to (ω 2 I − D)u = 0, or
√
2f /µ − 2f /M1 2f / M1 M2 u1
0=
√
. (28)
2f / M1 M2 2f /µ − 2f /M2 u2
q
with the solution u1 = − M2 /M1 u2 . In terms of the actual displace-
ments Eqs.11 v
u
sn1 u M 2 u1
=t (29)
sn2 M 1 u2
or sn1 /sn2 = −M2 /M1 so that the two atoms in the basis are moving
out of phase with amplitudes of motion inversely proportional to their
masses. These modes are described as optical modes since these atoms,
12
number of atoms in the basis). We also expect (and implicitly assumed
above) that the allowed frequencies are real and positive. However,
from simple mathematical identities, the point-group and translational
symmetries of the lattice, and its time-reversal invariance, we can learn
more about the dispersion without solving any particular problem.
The basic symmetries that we will employ are
• Time-reversal invariance.
∗βj 1
Dαi = q Φpβj
0αi e
−iq·(rp )
(30)
Mα Mβ
1 −p,β,j iq·(rp )
= q Φ0,α,i e (31)
Mα Mβ
Then, due to the symmetric properties of the second derivative
∗βj 1 0,α,i 1
Dαi = q Φ−p,β,j eiq·(rp ) = q Φp,α,i
0,β,j e
iq·(rp ) αi
= Dβj (32)
Mα Mβ Mα Mβ
13
Thus, DT ∗ = D† = D so D is hermitian and its eigenvalues ω 2 are
real. This means that either ω are real or they are pure imaginary. We
will assume the former. The latter yields pure exponential growth of
our Fourier solution, indicating an instability of the lattice to a second-
order structural phase transition.
Time-reversal invariance allows us to show related results. We as-
sume a solution of the form
1
snαi = √ uαi (q)ei(q·rn −ωt) (33)
Mα
which is a plane wave. Suppose that the plane wave is moving to the
right so that q = x̂qx , then the plane of stationary phase travels to the
right with
ω
x=t. (34)
qx
Clearly then changing the sign of qx is equivalent to taking t → −t.
If the system is to display proper time-reversal invariance, so that the
plane wave retraces its path under time-reversal, it must have the same
frequency when time, and hence q, is reversed, so
αi
Note that this is fully equivalent to the statement that Dβj (q) =
∗αi
Dβj (−q) which is clear from the definition of D.
Now, return to the secular equation, Eq. 15.
µ ¶
βj 2 βj
Dαi (q) −ω (q)δαi ²βj (q) = 0 (36)
14
Lets call the (normalized) eigenvectors of this equation ². They are the
elements of a unitary matrix which diagonalizes D. As a result, they
have orthogonality and completeness relations
X ∗(n) (m)
²α,i (q)²α,i (q) = δm,n orthogonality (37)
α,i
X ∗(n) ∗(n)
²α,i (q)²β,j (q) = δα,β δi,j (38)
n
If we now take the complex conjugate of the secular equation
µ ¶
βj βj
Dαi (−q) − ω 2 (−q)δαi ²∗βj (q) = 0 (39)
15
βj βj
it is easy to see that Dαi (q + G) = Dαi (q) (since G · rp = 2πn, where
n is an integer). I.e., D is periodic in k-space, and so its eigenvalues
(and eigenvectors) must also be periodic.
s 11
Figure 7: If each ion is shifted by s1,1,i , then the lattice energy is unchanged.
16
ions, the energy of the system should remain unchanged.
1 X m−n,β,j
δE = Φ0,α,i sn,α,i sm,β,j = 0 (45)
2 m,n,α,β,i,j
1 X m−n,β,j
= Φ0,α,i s1,1,i s1,1,j (46)
2 mnα,β,i,j
1X X m−n,β,j
= s1,1,i s1,1,j Φ0,α,i (47)
2 i,j mnα,β
After the stress has been applied, the atoms in the bulk of the sample
Figure 8: After a stress is applied to a lattice, the movement of each ion (strain) is
not only in the direction of the applied stress. The response of the lattice to an applied
stress is described by the strain matrix.
17
the central (n = 0) atom this means that
X
0 = F0,α,i = − Φm,β,j γ,k
0,α,i mβ,j sm,γ,k (50)
m,β,j,γ,k
X
Φm,β,j
0,α,i sm,γ,k = 0 (51)
m
whereas, sm,γ,k is odd in m. Thus the sum over all m yields zero.
Now let’s apply these constraints to D for an elemental lattice where
r = 1, and we may suppress the basis indices α.
1 X
Dij (q) = Φp,j
0,i e
iq·(rp )
(52)
M p
Thus, the leading order (small q) eigenvalues ω 2 (q) ∼ q 2 . I.e. they are
acoustic modes. We have shown that all elemental lattices must have
acoustic modes for small q.
18
In fact, one may show that all harmonic lattices in which the energy
is invariant under a rigid translation of the entire lattice must have at
least one acoustic mode. We will not prove this, but rather make a
simple argument. The rigid translation of the lattice corresponds to a
q = 0 translational mode, since no energy is gained by this translation,
it must be that ωs (q = 0) = 0 for the branch s which contains this
mode. The acoustic mode may be obtained by perturbing (in q) around
this point. Physically this mode corresponds to all of the elements of
the basis moving together so as to emulate the motion in the elemental
basis.
19
2.1 Periodicity and the Quantization of States
or
q(n + N )a = qna + 2πm where m is an integer (57)
20
G vector
Bisector
Figure 9: The First Brillouin Zone. The end points of all vector pairs that satisfy
the Bragg condition k − k0 = Ghkl lie on the perpendicular bisector of Ghkl . The
smallest polyhedron centered at the origin of the reciprocal lattice and enclosed by
perpendicular bisectors of the G’s is called the first Brillouin zone.
21
2.3 Point Group Symmetry and Density of States
Two other tricks to reduce the complexity of these sums are worth
mentioning here although they are discussed in detail elswhere.
The first is the use of the point group symmetry of the system. It
is clear from their definition in chapter 3, the reciprocal lattice vectors
have the same point group symmetry as the lattice. As we discussed
in chapter 2, the knowledge of the group elements and corresponding
degeneracies may be used to reduce the sums over k to the irreducible
wedge within the the First Brillioun zone. For example, for a cubic
system, this wedge is only 1/23 3! or 1/48th of the the FBZ!
The second is to introduce a phonon density of states to reduce the
multidimensional sum over k to a one-dimensional integral over energy.
This will be discussed in chapter 5.
In this section we will derive the equations of motion for the lattice, de-
termine the canonically conjugate variables (the the sense of Lagrangian
mechanics), and use this information to both first and second quantize
the system.
Any lattice displacement may be expressed as a sum over the eigen-
22
vectors of the dynamical matrix D.
1 X
sn,α,i = √ Qs (q, t)²sα,i (q)eiq·rn (58)
Mα N q,s
Recall that ²sα,i (q) are distinguished from usα,i (q) only in that they are
normalized. Also since q + G is equivalent to q, we need sum only over
the first Brillouin zone. Finally we will assume that Qs (q, t) contains
the harmonic time dependence and since sn,α,i is real Q∗s (q) = Qs (−q).
We may rewrite both the kinetic and potential energy of the system
as sums over Q. For example, the kinetic energy of the lattice
1 X
T = Mα (ṡnα,i )2 (59)
2 n,α,i
1 X X
= Q̇r (q)²rα,i (q)eiq·rn Q̇s (k)²sα,i (k)eik·rn (60)
2N n,α,i q,k,r,s
Then as
1 X i(k+q)·r X r ∗s
e n
= δk,−q and ²α,i ²α,i = δrs (61)
N n α,i
the kinetic energy may be reduced to
1 X ¯¯ ¯2
¯
T = ¯Q̇r (q)¯ (62)
2 q,r
The potential energy may be rewritten in a similar fashion
1 X
V = Φm,β,j sn,α,i sm,β,j
2 n,m,α,β,i,j n,α,i
m−n,β,j
1 X Φ0,α,i
= q
2 n,m,α,β,i,j N Mα Mβ
X
Qs (q, t)²sα,i (q)eiq·rn Qr (k, t)²rβ,j (k)eik·rm (63)
q,k,s,r
23
Let rl = rm − rn
l,β,j
1 X Φ0,α,i
V = q
2 n,l,α,β,i,j N Mα Mβ
X
Qs (q, t)²sα,i (q)eiq·rn Qr (k, t)²rβ,j (k)eik·(rl +rn ) (64)
q,k,s,r
1 X r
V = ²α,i (k)²∗s 2 ∗
α,i (k)ωr (k)Qs (k)Qr (k) (67)
2 α,i,k,r,s
P r ∗s
Finally, since α,i ²α,i (k)²α,i (k) = δr,s
1X 2
V = ωs (k) |Qs (k)|2 (68)
2 k,s
Thus we may write the Lagrangian of the ionic system as
1 X µ¯¯ ¯2
¯ 2 2
¶
L=T −V = ¯Q̇s (k)¯ − ωs (k) |Qs (k)| , (69)
2 k,s
where the Qs (k) may be regarded as canonical coordinates, and
∂L
Pr∗ (k) = = Q̇∗s (k) (70)
∂Qr (k)
(no factor of 1/2 since Q∗s (k) = Qs (−k)) are the canonically conjugate
momenta.
24
The equations of motion are
d ∂L ∂L
− ∗
or Q̈s (k) + ωs2 (k)Qs (k) = 0 (71)
dt ∂ Q̇s (k)
∗ ∂Qs (k)
for each k, s. These are the equations of motion for 3rN independent
harmonic oscillators. Since going to the Q-coordinates accomplishes the
decoupling of these equations, the {Qs (k)} are referred to as normal
coordinates.
P.A.M. Dirac laid down the rules of quantization, from Classical Hamilton-
Jacobi classical mechanics to Hamiltonian-based quantum mechanics
following the path (Dirac p.84-89):
25
Thus, following Dirac, we may now quantize the normal coordinates
[Q∗r (k), Ps (q)] = ih̄δk,q δr,s where the other commutators vanish .
(75)
Furthermore, since we have a system of 3rN uncoupled harmonic os-
cillators we may immediately second quantize by introducing
1 q i
as (k) = √ ωs (k)Qs (k) + q Ps (k) (76)
2h̄ ωs (k)
1 q i
a†s (k) = √ ωs (k)Q∗s (k) − q Ps∗ (k) , (77)
2h̄ ωs (k)
or
v
u
u h̄ ³ ´
Qs (k) = t as (k) + a†s (−k) (78)
2ωs (k)
v
u
u h̄ωs (k) ³ ´
t
Ps (k) = −i as (k) − a†s (−k) (79)
2
Where
[as (k), a†r (q)] = δr,s δq,k [as (k), ar (q)] = [a†s (k), a†r (q)] = 0 (80)
26
and a†s (k) and as (k) create and destroy phonons respectively, in the
state (k, s)
q
a†s (k) |ns (k)i = ns (k) + 1 |ns (k) + 1i (83)
q
as (k) |ns (k)i = ns (k) |ns (k) − 1i (84)
If |0i is the normalized state with no phonons present, then the state
with {ns (k)} phonons in each state (k, s) is
1
Y 1 2 Y ³ † ´ns (k)
|{ns (k)}i = as (k) |0i (85)
k,s ns (k)! k,s
27
Source of thermal
n λ ≈ | a 1 | or |a 2 |
neutrons a1
a2
Figure 10: Neutron Scattering. The De Broglie wavelength of the neutrons must
be roughly the same size as the lattice constants in order to learn about the lattice
structure and its vibrational modes from the experiment. This dictates the use of
thermal neutrons.
Thus the neutron cannot ”see” the detailed structure of the nucleus,
and so we may approximate the neutron-ion interaction potential as a
contact interaction
X
V (r) = Vn δ(r − rn ) (89)
rn
28
theory of diffraction. Nevertheless it is useful to compare the classical
result to what we will develop for the quantum problems.
For the classical problem we will assume that the lattice is elemental
(r = 1) and start with a generalization of the formalism developed in
the last chapter
I ∝ |ρ(K, Ω)|2 (90)
where
1
rn (t) = rn + sn (t) and sn (t) = √ u(q)ei(q·rn −ω(q)t) (92)
M
describes the harmonic motion of the s-mode with wave-vector q.
XZ
ρ(K, Ω) ∝ dtei[K·(rn +sn (t))−Ωt] . (93)
n
K = k0 − k = G and Ω = ω0 − ω = 0 (95)
29
When multiplied by h̄ , these can be interpreted as conditions for
the conservation of (crystal) momentum and energy when the scat-
tering event involves the creation (destruction) of a lattice excitation
(phonon). These processes are called Stokes and antistokes processes,
respectively, and are illustrated in Fig. 11.
k = k 0- q , ω = ω 0 - ω q k = k 0+ q , ω = ω 0 + ω q
q ,ω q q ,ω q
n n
k 0, ω0 k 0, ω0
Stokes Process Anti-Stokes Process
(phonon creation) (phonon absorbtion)
Figure 11: Stokes and antistokes processes in inelastic neutron scattering involving
the creation or absorption of a lattice phonon.
30
4.2 Quantum Theory of Neutron Scattering
• Callaway, p. 36–.
Initial Final
ψ 0 = 1− e i( k 0•⋅ r - ω 0 t ) ψ f = 1− i( k f •⋅ r - ω f t )
V
√ V e
√
n
k 0 ⋅ω 0
n k ⋅ω
φ0 E0 φf Ef
Figure 12: The initial (left) and final (right) states of the neutron and lattice during
a scattering event. The initial system state is given by Ψ0 = φ0 ψ0 , with energy
²0 = E0 + h̄ω0 where ω0 = k02 /2M . The final system state is given by Ψf = φf ψf ,
with energy ²f = Ef + h̄ωf where ωf = kf2 /2M .
31
interaction to be local
X 1 X Z d3 q X iq·(r−rn )
V (r) = V (r − rn ) = V (q)eiq·(r−rn ) = V0 e (97)
rn N q,n V n
If we now substitute in the ion-neutron potential Eq. 98, then the inte-
gral over r will yield a delta function V δ(q + k − k0 ) which allows the
q integral to be evaluated
¯ ¯
(2πh̄)3 X
2
¯X D
¯
¯ ¯
¯ −iK·rn ¯
E ¯2
¯
P =a 2
δ(E0 − Ef + h̄Ω) ¯¯ φ 0 ¯e ¯ φf ¯
¯
(101)
(M V ) f rn
32
Now, before proceeding to a calculation of the differential cross sec-
dσ
tion dΩdE we must be able to convert this probability (rate) for eigen-
states into a flux of neutrons of energy E and momentum p. A differ-
ential volume element of momentum space d3 p contains V d3 p/(2πh̄)3
neutron states. While this is a natural consequence of the uncertainty
principle, it is useful to show this in a more quantitative sense: Imagine
a cubic volume V = L3 with periodic boundary conditions so that for
any state Ψ in V ,
1 P iq·r
If we write Ψ(r) = N qe Ψ(q), then it must be that
with similar relations for the y and z components. So for each volume
³ ´
2π 3
element of q-space L there is one such state. In terms of states p =
h̄q, the volume of a state is (2πh̄/L)3 . Thus d3 p contains V d3 p/(2πh̄)3
states.
The incident neutron flux of states (velocity times density) is
¯ ¯
h̄k0 2 h̄k0 ¯¯ 1 ik0 ·r ¯¯2 h̄k0
j= |Ψ0 | = ¯√ e ¯ = (104)
M M ¯ V ¯ MV
Then since the number of neutrons is conserved
dσ h̄k0 dσ d3 p p2 dpdΩ
j dEdΩ = dEdΩ = P V = PV (105)
dEdΩ M V dEdΩ (2πh̄)3 (2πh̄)3
33
And for thermal (non-relativistic) neutrons E = p2 /2M , so dE =
pdp/M , and
h̄k0 dσ h̄kM dEdΩ
dEdΩ = P V (106)
M V dEdΩ (2πh̄)3
or
dσ k (M V )2
=P . (107)
dEdΩ k0 (2πh̄)3
Substituting in the previous result for P
¯ ¯
dσ k (M V )2 2 (2πh̄)3 X ¯X D
¯
¯ ¯
¯ −iK·rn ¯
E ¯2
¯
= 3
a 2
δ(E0 − Ef + h̄Ω) ¯¯ φ 0 ¯e ¯ φf ¯
¯
dEdΩ k0 (2πh̄) (M V ) f rn
(108)
or
dσ k N a2
= S(K, Ω) (109)
dEdΩ k0 h̄
where
¯ ¯
1 X ¯X D ¯ ¯ E ¯2
¯ ¯ −iK·rn ¯ ¯
S(K, Ω) = δ(E0 − Ef + h̄Ω) ¯¯ φ 0 ¯e ¯ φf ¯
¯
. (110)
N f rn
so that
1 XZ ∞ dt i((E0 −Ef )/h̄+Ω)t
S(K, Ω) = e
N f
−∞ 2π
X ¯¯D ¯ ¯
¯ iK·rn ¯
E¯ ¯D
¯¯
¯ ¯
¯ −iK·rm ¯
E¯
¯
¯ φ 0 ¯e ¯ φ f ¯ ¯ φ f ¯e ¯ φ0 ¯ . (112)
rn ,rm
then as
e−iHt/h̄ |φl i = e−iEl t/h̄ |φl i (113)
34
where H is the lattice Hamiltonian, we can write this as
1 XZ ∞ dt iΩt X D ¯¯ iHt/h̄ iK·rn −iHt/h̄ ¯¯ E
S(K, Ω) = e φ 0 ¯e e e ¯ φf
N f
−∞ 2π rn ,rm
D ¯ ¯ E
φf ¯¯e−iK·rm ¯¯ φ0 , (114)
Thus,
1 XZ ∞ dt iΩt X D ¯¯ iK·rn (t) ¯¯ E D ¯¯ −iK·rm ¯¯ E
S(K, Ω) = e φ 0 ¯e ¯ φf φ f ¯e ¯ φ0
N f
−∞ 2π rn ,rm
1 Z ∞ dt X D ¯ ¯ E
= eiΩt φ0 ¯¯eiK·rn (t) e−iK·rm ¯¯ φ0 . (116)
N −∞ 2π rn ,rm
35
4.2.1 The Debye-Waller Factor
This relation is in fact true to all orders, as long as A and B are linear
functions of a† and a . (c.f. Ashcroft and Mermin, p. 792, Callaway pp.
41-48). Thus
D E 2 2
eiK·sn (t) e−iK·sm = e− 2 h(K·sn (t)) i e− 2 h(K·sm ) i ehK·sn (t)K·sm i .
1 1
(122)
Since the Hamiltonian has no time dependence, and the lattice is in-
variant under translations rn
D E 2
eiK·sn (t) e−iK·sm = e−h(K·sn ) i ehK·sn−m (t)K·s0 i , (123)
36
where the first term is called the Debye-Waller factor e−2W .
2
e−2W = e−h(K·sn ) i . (124)
Thus letting l = n − m
Z dt i(K·rl +Ωt) hK·sl (t)K·s0 i
−2W X ∞
S(K, Ω) = e e e . (125)
l
−∞ 2π
Here the Debye-Waller factor contains much of the crucial quantum
physics. It is finite, even at T = 0 due to zero-point fluctuations, and
since hK · sn i2 will increase with temperature, the total strength of the
Bragg peaks will diminish with increasing T . However, as long as a
crystal has long-ranged order, it will remain finite.
37
dσ0 k N a2 −2W X
= e δ(Ω)N δK,G (129)
dEdΩ k0 h̄ G
However, now the scattering intensity is reduced by the Debye-
Waller factor e−2W , which accounts for zero-point motion and thermal
fluctuations.
= a(q)e−iω(q)t (132)
where we have used the fact that (a† a)n a = (a† a)n−1 a† aa = (a† a)n−1 (aa† −
1)a = (a† a)n−1 a(a† a−1) = a(a† a−1)n . Similarly a† (q, t) = a† (q)eiω(q)t .
Thus,
√ iq·r
1 X h̄e n s ³ ´
sn,α (t) = √ q ²α (q) as (q)e−iωs (q)t + a†s (−q)eiωs (q)t
Mα N q,s 2ωs (q)
(133)
38
and
v
u
1 Xu h̄ ³ ´
s0,α (0) = √ t ²rα (p) ar (p) + a†r (−p) (134)
Mα N p,r 2ωr (p)
−2W
Z ∞dt iΩt X h̄ |K · ²(K)|2
S1 (K, Ω) = e e δK+q,G (136)
−∞ 2π q,G,s 2ωs (q)M
h D E D Ei
e−iωs (q)t as (−K)a†s (−K) + eiωs (q)t a†s (−K)as (−K)
39
this can happen at any temperature, since (1 + ns (K)) 6= 0 at any T .
The second term is only finite when Ω + ωs (K) = ω0 − ωf + ωs (K) = 0;
ie., the final energy of the neutron is larger than the initial energy. The
additional energy comes from the absorption of a phonon. Thus phonon
absorption is only allowed at finite temperatures, and in fact, the factor
ns (K) = 0 at zero temperature. These terms correspond to the Stokes
and anti-Stokes processes, respectively, illustrated in Fig. 13.
k = k 0- q , ω = ω 0 - ω q k = k 0+ q , ω = ω 0 + ω q
1 + n (K) n (K)
s s
q ,ω q q ,ω q
n n
k 0, ω0 k 0, ω0
Stokes Process Anti-Stokes Process
(phonon creation) (phonon absorbtion)
Figure 13: Stokes and antistokes processes in inelastic neutron scattering involving
the creation or absorption of a lattice phonon. The antistokes process can only occur
at finite-T, when ns (K) 6= 0.
40