Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Fundamentals of the Physics of Solids / 12-The Quantum Theory of Lattice Vibrations

.pdf
Скачиваний:
47
Добавлен:
08.01.2014
Размер:
786.99 Кб
Скачать

12.2 Density of Phonon States

407

Nλ

(ω) =

V

 

4π

(ω − ω0)3/2

=

V

 

(ω − ω0)3/2

.

(12.2.40)

(2π)3 3 (α1α2α3)1/2

 

 

 

 

6π2 (α1α2α3)1/2

When the derivative of this quantity with respect to frequency is divided by the volume, the density of states (12.2.38) is recovered.

In P1-type points one of the αi is negative and the two others are positive, and so the spectrum has a saddle point at q0. Surfaces of constant energy are hyperboloids of two sheets for ω < ω0 and hyperboloids of one sheet for ω > ω0. These are illustrated in Fig. 12.6. In the ω = ω0 case the hyperboloid degenerates into a cone.

Fig. 12.6. Surfaces of constant energy around a P1-type saddle point for ω < ω0 and ω > ω0

Using the form

ωλ(q) = ω0 + α1ξ12 + α2ξ22 − α3ξ32

(12.2.41)

for describing the frequency spectrum, and changing to the variables xi = α1i /2ξi, the density of states can be expressed as

gλ(ω) = (2π)3

 

(α1

α2α3)1/2

dx1 dx2 dx3 δ(ω − ω0 − x12 − x22 + x32) .

1

 

 

1

 

 

(12.2.42) To evaluate the integral, we shall introduce new coordinates that are adapted to the hyperboloid. We choose

x1 = r sinh θ cos ϕ,

x2 = r sinh θ sin ϕ,

x3 = ±r cosh θ (12.2.43)

for ω < ω0, and

x1 = r cosh θ cos ϕ,

x2 = r cosh θ sin ϕ,

x3 = r sinh θ

(12.2.44)

408 12 The Quantum Theory of Lattice Vibrations

for ω > ω0. By evaluating the Jacobian for the new sets of variables, we have

gλ(ω) = (2π)3

 

(α1

α2α3)1/2

 

r2 dr sinh θ dθ dϕ δ(ω − ω0 + r2)

(12.2.45)

1

 

 

1

 

 

 

for ω < ω0, and

 

(α1

α2α3)1/2

 

r2 dr cosh θ dθ dϕ δ(ω − ω0 − r2)

(12.2.46)

gλ(ω) = (2π)3

 

1

 

 

1

 

 

 

for ω > ω0. In either case the integrand vanishes unless r = |ω − ω0|1/2. Assuming once again that the quadratic approximation for the dispersion curve is valid in a finite neighborhood of the saddle point, integration is performed in the region where

x12 + x22 + x32 ≤ R2 .

(12.2.47)

Positive and negative values of x3 contribute equally, therefore it is sufficient to calculate the integral over positive values. To include only points within the sphere of radius R, the following restriction must be imposed on the variable θ:

 

cosh θ ≤

 

 

 

R2

+ 1

1/2

 

 

 

 

1

21

 

 

,

if

ω < ω0 ,

 

 

r2

 

(12.2.48)

 

sinh θ ≤

 

 

R2

1

1/2

 

 

 

 

0

21

 

 

,

if

ω > ω0 .

 

r2

 

 

 

 

Integration with respect to θ factor of 2π, we have

2

1

 

gλ(ω) = (2π)2 (α1α2α3)1/2

for ω < ω0, and

is straightforward. Since the ϕ-integral gives a

r dr 2 r2 + 1

1

δ(ω − ω0 + r )

 

 

R2

1/2

 

2

1

 

 

 

2

 

 

 

 

 

(12.2.49)

 

2

 

 

1

r

 

 

2

 

R2

1/2

 

 

gλ(ω) =

 

 

 

dr

 

δ(ω − ω0

− r

)

(2π)2 (α1

α2α3)1/2

2

r2 1

 

 

 

 

 

 

 

1

 

 

 

 

2

 

(12.2.50) for ω > ω0. Because of the delta functions these integrals are also elementary. For frequencies close to the saddle point ω0, i.e., for |ω − ω0| R2, the final result is

 

 

C

1

(ω0 − ω)1/2

+

O

(ω

0

ω) ,

if

ω < ω

 

,

g (ω) =

 

4π2 α1α2α3 1/2

 

 

 

 

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

|

|

 

 

 

 

 

 

 

 

 

λ

 

 

 

 

 

 

 

 

if

 

 

(12.2.51)

 

C + O(ω − ω0) ,

 

 

 

 

 

 

ω > ω0 .

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The density of states has a kink at ω0, with a square-root singularity on one side.

12.3 The Thermodynamics of Vibrating Lattices

409

In P2-type analytical critical points two of the αi are negative and the third is positive, so once again the spectrum has a saddle point at q0. The density of states exhibits a kink again, but the singularity is now on the other side:

g

(ω) =

C + O(ω0 − ω) ,

1/2

 

if

ω < ω0 ,

λ

 

 

 

1

 

 

(ω

 

ω0)

 

 

 

(12.2.52)

 

 

C

 

 

 

 

 

 

 

+ (ω ω0) , if

ω > ω0 .

 

 

 

 

2

 

 

 

 

1/2

 

 

 

4π

 

1

α2α3|

O −

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Finally, in P3-type points each coe cient αi is negative, and the spectrum has a minimum at q0. The density of states is

gλ(ω) =

1

 

(ω0 − ω)1/2

.

(12.2.53)

 

 

 

4π2 1α2α3|1/2

 

Figure 12.7 shows the behavior of the density of states around the four kinds of the analytical critical point. It should be mentioned again that the velocity of acoustic phonons is finite at q = 0, and so no Van Hove singularity appears at ω = 0 in the phonon density of states. In three-dimensional systems the density of states is initially proportional to ω2, in agreement with the Debye model.

g!) Saddle point P1

 

Saddle point P2

P0

 

P3

 

 

 

!

Fig. 12.7. Van Hove singularities in the phonon density of states

12.3 The Thermodynamics of Vibrating Lattices

From now on we consider the vibrating lattice as a gas of free phonons governed by the Bose–Einstein statistics, which can be characterized by the dispersion relation of the elementary excitations or the density of states. We shall examine how the thermodynamic properties of the crystal lattice may be interpreted in this picture.

410 12 The Quantum Theory of Lattice Vibrations

12.3.1 The Ground State of the Lattice and Melting

In the ground state of a classical crystal atoms are fixed rigidly at their equilibrium positions. Such a state is ruled out in quantum mechanics: this is why the Hamiltonian contains, in addition to the particle number, a term 1/2, which corresponds to the energy of zero-point vibrations. If it were not for these zero-point vibrations, the mean square displacement of the atoms would vanish in the ground state.

The latter can be determined using (12.1.39), the formula for atomic displacements. Only those terms contribute to the average in which phonons of the same energy are created and annihilated, either in this or in the reverse order:

 

2

 

 

 

 

 

 

 

1

(λ)

 

(λ)

 

 

 

uα

(m, μ)

=

 

 

 

 

eμ,α

(q)eμ,α

(q)

 

 

2N Mμ

q,λ

ωλ(q)

 

 

 

 

 

 

 

 

 

 

 

 

 

(12.3.1)

 

 

×

 

 

 

 

 

 

(q)aλ(q)

 

 

 

 

aλ(q)aλ(q) + aλ

.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

For simplicity, we shall consider a crystal with a monatomic basis, so the label μ will be absent. By summing over the three Cartesian coordinates, and by assuming that polarization vectors are normalized to unity, interchanging the order of the creation and annihilation operators yields

u2

(m) =

 

 

1

 

 

aλ

(q)aλ

(q) +

21

.

(12.3.2)

N M q,λ

ωλ(q)

 

 

 

 

 

 

 

 

 

 

 

 

 

Even though no phonons are present in the ground state, the mean square displacement of the atoms is nevertheless finite because of the quantum fluctuations:

 

 

 

 

1

 

 

u2

(m) =

 

 

 

,

(12.3.3)

2N M

 

 

 

 

q,λ

ωλ(q)

 

 

 

 

 

 

 

which is the straightforward generalization of expression (12.1.22) for a single oscillator. Using the density of states, the left-hand side can be expressed as a frequency integral:

u2(m)

 

= 2N M

dω g ω .

(12.3.4)

 

 

 

V

 

(ω)

 

According to (12.2.20), in d-dimensional crystals the density of states g(ω) is proportional to ωd−1 in the low-frequency limit, therefore in d = 2 and d = 3 dimensions integration up to the finite Debye frequency gives a finite result for the mean square displacement of the atoms. However, in d = 1 dimension the integral diverges at the lower limit, indicating that one-dimensional ordered chains of atoms with discrete translational symmetry cannot exist. Quantum fluctuations are so strong that they destroy long-range order. This is in line

12.3 The Thermodynamics of Vibrating Lattices

411

with Coleman’s theorem presented in Chapter 6, which states that quantum fluctuations restore continuous symmetry in one dimension.

Obviously, the divergence has arisen because of the transformation of the discrete sum into an integral. Strictly speaking, this means that only in the thermodynamic limit, when the size of the sample approaches infinity, does the atomic displacement become infinitely large. In finite systems the wave vectors are finite and discrete, therefore the contribution of low-energy phonons is finite. This can be su ciently large to destabilize regular, periodic arrays of atoms, nevertheless atoms may be arranged in one-dimensional chain-like patterns. This is observed in the compound Hg3−δ AsF6, where mercury atoms occupy the tubes formed by AsF6 molecules. As the arrangement is one-dimensional, the positions of the mercury atoms do not show long-range order, consequently no sharp Bragg peaks are observed in the di raction pattern. However, there is a short-range order, since nearest-neighbor distances are roughly equal.

Expression (12.3.2) for the mean square atomic displacement is valid at finite temperatures as well, provided the average is taken as a thermal average. Using the Bose–Einstein statistics for the occupation number of the phonons,

 

u2(m) = N M q,λ

ωλ(q) e ωλ(q)/kBT

 

1 +

21

 

 

 

 

 

1

 

 

1

 

 

 

 

 

 

1

 

 

 

 

 

 

(12.3.5)

 

=

 

 

coth ωλ(q) .

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2N M

 

 

ω

(q)

 

 

2k T

 

 

 

 

 

 

 

 

q,λ

λ

 

 

 

 

B

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

In terms of the density of states this reads

u2(m)

 

= 2N M

dω g ω

coth 2kBT .

(12.3.6)

 

 

 

V

 

(ω)

 

ω

 

At finite temperatures, for small values of ω, coth( ω/2kBT ) 2kBT / ω, so when the small-ω asymptotic form (12.2.20) of the density of states is used, the integrand shows an ωd−3 dependence. Thus the above integral diverges at the lower limit even in d = 2 dimensions. This indicates that while stable two-dimensional crystal structures may exist at T = 0, they are destroyed at arbitrarily low nonzero temperatures.

This result can be considered as a consequence of the generalization of the Mermin–Wagner theorem,5 originally formulated for magnetic systems. According to the general theorem, in two dimensions no long-range ordered state may exist at any finite temperature that breaks a continuous symmetry of the Hamiltonian. In our case the Hamiltonian is invariant under continuous translations, therefore in two- (or lower-) dimensional systems crystal structures that are invariant only under discrete translations cannot be stable at any finite temperature.6 The melting point of the two-dimensional crystal is

5N. D. Mermin and H. Wagner, 1966.

6As mentioned above, in the one-dimensional case the crystalline state, which breaks the continuous symmetry, cannot be stable even as a ground state.

412 12 The Quantum Theory of Lattice Vibrations

therefore Tm = 0. The finite-temperature instability of the symmetry-breaking state is due to the excessively high number of soft Goldstone modes – acoustic phonons in our case – that are present even at very low temperatures and that disrupt the crystalline order.

Let us now calculate the mean square displacement in a three-dimensional crystal at finite temperature. Using the density of states (12.2.19) determined for the Debye model (with p = 1 for the monatomic basis),

u2(m)

 

= M

ωD

ωD3

eβ ω

 

1 + 21

dω .

(12.3.7)

 

 

 

 

 

9

 

 

ω

1

 

 

 

 

 

 

 

0

 

 

 

 

 

 

Instead of the Debye frequency ωD the Debye temperature ΘD defined via

kBΘD = ωD

(12.3.8)

is commonly used. Changing to the variable t = β ω,

u2

(m) =

9 2

 

 

 

T

kBΘDM

ΘD

 

 

 

 

 

 

 

 

= kBΘDM

&

4

+

 

 

9 2

 

1

 

 

2

ΘD /T

t

et 1 + 21 dt

 

 

0

 

 

 

 

 

 

 

 

 

1

 

 

 

 

 

 

 

 

 

 

ΘD/T

 

 

(12.3.9)

 

 

 

 

 

2

t

 

dt'.

 

 

 

 

 

0

 

 

 

 

ΘD

 

 

et

1

 

 

 

 

 

 

 

 

 

 

 

T

 

 

 

 

 

 

 

 

 

When only the first-order temperature correction to the zero-point vibrations is taken into account at low temperatures (T ΘD),

u2(m) = 4kBΘDM &1 +

3

ΘD

+ . . . '.

(12.3.10)

 

9 2

2π2

 

T

2

 

Above the Debye temperature the integral can be evaluated using the expansion valid for t < 1,

 

 

 

t

 

 

 

t

 

 

 

t2

 

 

 

 

 

 

 

 

 

 

= 1

 

+

 

 

 

+ . . . .

 

(12.3.11)

This gives

 

 

et 1

2

12

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

9 2

 

T

&1 +

1

 

Θ

 

2

'.

 

u2(m)

=

 

 

 

 

 

D

+ . . .

(12.3.12)

kBΘDM

ΘD

36

T

It seems reasonable to assume that the crystal melts when the root-mean- square displacement becomes comparable to the lattice constant a – that is, at the melting point Tm

7

 

 

9 2Tm

≈ αa ,

(12.3.13)

kBΘD2 M

12.3 The Thermodynamics of Vibrating Lattices

413

where α should be less than, but on the order of unity. Using the volume v of the primitive cell instead of the lattice constant,

Tm

α2k

 

9 2B ΘD2 M v2/3 .

(12.3.14)

Experimental data are in good agreement with the above relation if α ≈ 1/4. This is the Lindemann criterion for melting,7 even though it was formulated somewhat di erently before the advent of the Debye model.

12.3.2 The Specific Heat of the Phonon Gas

We shall now determine the contribution of atomic vibrations to the specific heat of solids. The previous expression for the thermal energy, (12.1.12), is expected to be modified by zero-point vibrations as

E = q,λ

ωλ(q)

eβ ωλ(q)

 

1

+ 21

= 21

q,λ ωλ(q) coth

21 β ωλ(q) .

 

 

1

 

 

 

 

 

 

 

 

 

 

 

(12.3.15) To demonstrate this, we shall make use of a statistical mechanical formula for the internal energy of a system with discrete energy levels Ei:

 

Eie−βEi

 

 

 

 

E =

i i e−βEi

=

 

ln

e−βEi .

(12.3.16)

∂β

i

 

 

 

 

 

 

 

 

It is a simple matter to evaluate the sum on the right hand side, the partition function

 

(12.3.17)

Z = e−βEi .

i

Writing the energy of the phonon system as a sum of individual modes,

Ei = Ei(q, λ) =

ωλ(q)[ni(q, λ) + 1

] ,

(12.3.18)

 

2

 

 

q,λ

q,λ

 

 

where ni(q, λ) can be any nonnegative integer. Then

Z = ? e−βEi (q,λ)

q,λ i

 

 

?

=eβ ωλ(q)/2 + e3β ωλ (q)/2 + e5β ωλ(q)/2 + . . .

q,λ

 

 

 

 

 

?

e−β ωλ(q)/2

 

 

 

 

 

=

1

 

e−β ωλ(q)

.

(12.3.19)

q,λ

 

 

 

 

 

7 F. A. Lindemann, 1910.

414 12 The Quantum Theory of Lattice Vibrations

From this formula the above expression for the internal energy is indeed recovered. Alternatively, the internal energy of the phonon gas can be expressed in terms of its Helmholtz free energy

F = kBT ln Z = q,λ

ωλ(q)

 

 

 

 

 

ωλ (q)/kB T

 

2

 

+ kBT

q,λ ln 1 e

 

 

 

 

 

 

 

 

 

 

= kBT q,λ ln 2 sinh

 

 

2kBT

,

 

(12.3.20)

 

 

 

 

 

 

ωλ(q)

 

 

 

using the thermodynamic identity

 

 

 

 

 

 

 

 

 

 

 

 

∂F

 

 

ln Z

 

 

 

E = F − T

 

= kBT 2

 

 

,

 

(12.3.21)

∂T

∂T

 

which is equivalent to (12.3.16). In terms of the density of states, the Helmholtz free energy and the internal energy read

F = kBT V

 

ln 2 sinh

21 β ω ! g(ω) dω ,

(12.3.22-a)

E = 21 V

ω coth( 21 β ω) g(ω) dω .

(12.3.22-b)

Let us return to form (12.3.15) of the internal energy. At high temperatures the condition β ωλ(q) = ωλ(q)/kBT 1 holds for all frequencies of the spectrum, and so it is su cient to keep the first two terms in the series expansion of coth x given in (3.2.83):

E = q,λ

kBT

1 + 12

 

kBT

 

2

.

(12.3.23)

+ . . .

 

 

1

 

ωλ(q)

 

 

 

 

 

 

 

 

 

 

 

 

 

When the specific heat is calculated from this expression, the leading term gives the classical Dulong–Petit value. Its first correction would be of order 1/T 2, however this is suppressed by other contributions that are neglected in the harmonic approximation.

The frequency of acoustic phonons usually ranges from zero to order 1012 Hz. As the phonon density of states is proportional to ω2, acoustic phonons of frequency 1012 Hz can be considered as typical. Even for optical phonons the frequency is at most 1013 Hz – so typical phonon energies are between 10 and 100 meV. This also means that the Dulong–Petit law is expected to be valid only at temperatures well above 100 K. At lower temperatures energy and specific heat can be determined only numerically from the above expressions, and only when the spectrum is known. The energy can be

split into a temperature-independent term

 

E0 =

21 ωλ(q)

(12.3.24)

q,λ

12.3 The Thermodynamics of Vibrating Lattices

415

and a temperature-dependent part

 

 

 

 

 

 

 

ωλ(q)

 

 

 

 

 

 

 

 

ET

=

eβ ωλ (q)

 

 

.

(12.3.25)

 

q,λ

 

1

 

 

 

 

 

 

 

Expressing the sum over the phonon quantum numbers by the density of states,

ET = V

eβ ωω1 g(ω) dω .

(12.3.26)

 

 

 

From the previous expressions for the energy, the specific heat is

CV

=

 

∂ ωλ(q)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

∂T eβ ωλ(q)

 

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

q,λ

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(12.3.27)

 

= kB q,λ

 

2kBT

 

2

 

 

 

 

2kBT

 

 

 

 

sinh2

,

 

 

 

 

 

 

ωλ(q)

 

 

 

 

 

 

ωλ(q)

 

 

or alternatively

 

 

∂T eβ ωω1 g(ω) dω

 

 

 

 

 

CV = V

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

= V kB

 

 

 

2

sinh2

 

 

 

 

 

(12.3.28)

2kBT

2kBT g(ω) dω .

 

 

 

 

 

 

 

 

ω

 

 

 

 

 

 

 

 

ω

 

 

At low temperatures, where practically only low-energy acoustic phonons (which belong to the linear regime of the dispersion curve) are excited, the Debye model provides a good approximation. Using the Debye model density of states in (12.3.26),

ET = 9pN

With the new variable t = β ω,

ET = 9pN kBT

kBT

ωD

ωD

ω

 

 

ω2

 

 

 

0

 

 

 

 

 

 

 

 

 

dω .

(12.3.29)

eβ ω 1

ωD3

3

ωD/kB T

t3

 

dt .

(12.3.30)

et

1

 

0

 

 

 

 

 

Besides the factor p, which specifies the number of atoms in the basis, all information about the material characteristics of the crystal is contained in the parameter ωD. Using the Debye temperature ΘD instead,

ET = 9pN kBT

 

ΘD

3

ΘD/T

3

 

(12.3.31)

 

 

et t1 dt .

 

 

T

 

 

 

 

 

0

416 12 The Quantum Theory of Lattice Vibrations

Introducing the Debye function through the definition

 

 

 

3

x

 

 

t3

 

 

 

 

 

 

 

 

D3(x) =

0

 

 

 

 

 

dt ,

(12.3.32)

 

x3

 

 

et 1

we have

 

 

 

 

 

 

 

 

Θ

.

 

 

 

 

 

 

 

 

 

 

 

 

 

ET = 3pN kBT D3

D

 

(12.3.33)

 

T

The specific heat is then

 

 

TD D3

 

 

TD .

 

CV = 3pN kB

D3 TD

 

 

(12.3.34)

 

 

Θ

 

 

Θ

 

 

 

 

 

Θ

 

A simple integration by parts shows that this is equivalent to the expression

CV = 9pN kB

 

 

3

ΘD

 

 

T

 

ΘD/T

4

t

 

0

t e

 

dt .

(12.3.35)

(et

1)2

When the condition T ΘD is met, the energy and the specific heat contain the asymptotic form of the Debye function valid for large values of x. According to (C.2.14),

D3(x)

3 π4

 

(12.3.36)

x3

 

15

 

 

in this limit, and so

 

 

 

 

 

 

 

3

 

 

12π4

 

 

 

 

T

 

CV = pN

kB

.

(12.3.37)

5

ΘD

In the low-temperature region the temperature dependence of the specific heat is usually well approximated by the cubic term. When CV /T is plotted against T 2, as in Fig. 12.8, a straight line can be fit fairly well to the measured data.

While a pure T 3 temperature dependence is observed in ionic crystals that are electrical insulators, the finite intercept for metals indicates that their specific heat contains an additional term which is linear in T . We shall see in Chapter 16 (Volume 2) that this is due to the electrons that are not bound to the ion cores and move almost freely.

On the other hand, it follows from (C.2.13), the expansion of the Debye function valid for small values of x, that at high temperatures the Dulong– Petit value is recovered from (12.3.35). Since (12.3.34) and (12.3.35) give good approximations for the specific heat both in the lowand high-temperature regions, the numerical evaluation of the integral in the Debye function may give an interpolation formula that is valid in the intermediate region as well. As indicated in Fig. 12.9, experimental data for the temperature dependence of specific heat show remarkable agreement with this interpolation formula.