Skip to main content

Lattice Vibrations and Thermal Properties

  • Chapter
  • First Online:
Solid-State Physics

Abstract

Chapter 1 was concerned with the binding forces in crystals and with the manner in which atoms were arranged. Chapter 1 defined, in effect, the universe with which we will be concerned. We now begin discussing the elements of this universe with which we interact. Perhaps the most interesting of these elements are the internal energy excitation modes of the crystals. The quanta of these modes are the “particles” of the solid. This chapter is primarily devoted to a particular type of internal mode—the lattice vibrations.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 119.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 159.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Had we chosen the sum to run over only the outer electrons associated with each atom, then we would have to replace the last term in (2.3) by an ion–ion interaction term. This term could have three and higher body interactions as well as two-body forces. Such a procedure would be appropriate [51, p. 3] for the practical discussion of lattice vibrations. However, we shall consider only two-body forces.

  2. 2.

    We have used the terms Born–Oppenheimer approximation and adiabatic approximation interchangeably. More exactly, Born–Oppenheimer corresponds to neglecting Cnn, whereas in the adiabatic approximation Cnn is retained.

  3. 3.

    For further details of the Born–Oppenheimer approximation,  [46, 82], [22, Vol. 1, pp. 611–613] and the references cited therein can be consulted.

  4. 4.

    See Dick and Overhauser [2.12].

  5. 5.

    See, for example, Cochran [2.9].

  6. 6.

    See Toya [2.34].

  7. 7.

    See Lehman et al. [2.23]. For a more general discussion, see Srivastava [2.32].

  8. 8.

    See Montroll [2.28].

  9. 9.

    The discussion of 1D (and 2D) lattices is perhaps mainly of interest because it sets up a formalism that is useful in 3D. One can show that the mean square displacement of atoms in 1D (and 2D) diverges in the phonon approximation. Such lattices are apparently inherently unstable. Fortunately, the mean energy does not diverge, and so the calculation of it in 1D (and 2D) perhaps makes some sense. However, in view of the divergence, things are not as simple as implied in the text. Also see a related comment on the Mermin–Wagner theorem in Chap. 7 (Sect. 7.2.5 under Two Dimensional Structures).

  10. 10.

    When we discuss lattice vibrations in three dimensions we give a more general technique for handling the case of two atoms per unit cell. Using the dynamical matrix defined in that section (or its one-dimensional analog), it is a worthwhile exercise to obtain (2.87a) and (2.87b).

  11. 11.

    See [2.39].

  12. 12.

    For the derivation of (2.108), see the article by Maradudin op cit (and references cited therein).

  13. 13.

    Elliott and Dawber [2.15].

  14. 14.

    See, for example, Jensen [2.19].

  15. 15.

    The way to do this is explained later when we discuss the classical calculation of the dispersion relation.

  16. 16.

    Maradudin et al. [2.26].

  17. 17.

    Note that this substitution assumes the results of Bloch’s theorem as discussed after (2.39).

  18. 18.

    We will later introduce more general ways of deducing the density of states from the dispersion relation, see (2.258).

  19. 19.

    [2.10, 1973, Chap. 8].

  20. 20.

    U0 is included for completeness, but we end up only using a vanishing temperature derivative so it could be left out.

  21. 21.

    See, e.g., Ghatak and Kothari [2.16, Chap. 4] or Brown [2.7, Chap. 5].

  22. 22.

    If one can assume central forces Cauchy proved that c12 = c44, however, this is not a good approximation in real materials.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to James D. Patterson .

Problems

Problems

  1. 2.1

    Find the normal modes and normal-mode frequencies for a three-atom “lattice” (assume the atoms are of equal mass). Use periodic boundary conditions.

  2. 2.2

    Show when m and m′ are restricted to a range consistent with the first Brillouin zone that

    $$ \frac{1}{N}\sum\limits_{n} {\exp \left( {\frac{{2\pi {\text{i}}}}{N}\left( {m - m^{{\prime }} } \right)n} \right) = \delta_{m}^{{m^{{\prime }} }} } , $$

    where \( \delta_{m}^{{m^{{\prime }} }} \) is the Kronecker delta.

  3. 2.3

    Evaluate the specific heat of the linear lattice [given by (2.80)] in the low temperature limit.

  4. 2.4

    Show that Gmn = Gnm, where G is given by (2.100).

  5. 2.5

    This is an essay length problem. It should clarify many points about impurity modes. Solve the five-atom lattice problem shown in Fig. 2.14. Use periodic boundary conditions. To solve this problem define A = β/α and δ = m/M (α and β are the spring constants) and find the normal modes and eigenfrequencies. For each eigenfrequency, plot 2 versus δ for A = 1 and 2 versus A for δ = 1. For the first plot: (a) The degeneracy at δ = 1 is split by the presence of the impurity. (b) No frequency is changed by more than the distance to the next unperturbed frequency. This is a general property. (c) The frequencies that are unchanged by changing δ correspond to modes with a node at the impurity (M). (d) Identify the mode corresponding to a pure translation of the crystal. (e) Identify the impurity mode(s). (f) Note that as we reduce the mass of M, the frequency of the impurity mode increases. For the second plot: (a) The degeneracy at A = 1 is split by the presence of an impurity. (b) No frequency is changed more than the distance to the next unperturbed frequency. (c) Identify the pure translation mode. (d) Identify the impurity modes. (e) Note that the frequencies of the impurity mode(s) increase with β.

    Fig. 2.14
    figure 14

    The five-atom lattice

  6. 2.6

    Let aq and \( a_{q}^{\dag } \) be the phonon annihilation and creation operators. Show that

    $$ \left[ {a_{q} ,q_{{q^{1} }} } \right] = 0\quad {\text{and}}\quad \left[ {a_{q}^{\dag } ,a_{{q^{1} }}^{\dag } } \right] = 0. $$
  7. 2.7

    From the phonon annihilation and creation operator commutation relations derive that

    $$ a_{q}^{\dag } \left| {n_{q} } \right\rangle = \sqrt {n_{q} + 1} \left| {n_{q} + 1} \right\rangle , $$

    and

    $$ a_{q} \left| {n_{q} } \right\rangle = \sqrt {n_{q} } \left| {n_{q} - 1} \right\rangle . $$
  8. 2.8

    If a1, a2, and a3 are the primitive translation vectors and if Ωa = a1 · (a2 × a3), use the method of Jacobians to show that dx dy dz = Ωa dη1 dη2 dη3, where x, y, z are the Cartesian coordinates and η1, η2, and η3 are defined by r = η1a1+ η2a2 + η3a3.

  9. 2.9

    Show that the bi vectors defined by (2.172) satisfy

    $$ \varOmega_{a} \varvec{b}_{1} = \varvec{a}_{2} \times \varvec{a}_{3} ,\quad \varOmega_{a} \varvec{b}_{2} = \varvec{a}_{3} \times \varvec{a}_{1} ,\quad \varOmega_{a} \varvec{b}_{3} = \varvec{a}_{1} \times \varvec{a}_{2} , $$

    where Ωa = a1 ∙ (a2 × a3).

  10. 2.10

    If Ωb = b1 · (b2 × b3), Ωa = a1 · (a2 × a3), the bi are defined by (2.172), and the ai are the primitive translation vectors, show that Ωb = 1/Ωa.

  11. 2.11

    This is a long problem whose results are very important for crystal mathematics. [See (2.178)–(2.184)]. Show that

    1. (a)
      $$ \frac{1}{{N_{1} N_{2} N_{3} }}\sum\limits_{{\varvec{R}_{l} }} {\exp \left( {{\text{i}}\varvec{q} \cdot \varvec{R}_{l} } \right) = \sum_{{\varvec{G}_{n} }} \delta_{{\varvec{q},\varvec{G}_{n} }} } , $$

      where the sum over Rl is a sum over the lattice.

    2. (b)
      $$ \frac{1}{{N_{1} N_{2} N_{3} }}\sum\limits_{q} {\exp \left( {{\text{i}}\varvec{q} \cdot \varvec{R}_{l} } \right) = \delta_{{\varvec{R}_{l} ,0}} } , $$

      where the sum over q is a sum over one Brillouin zone.

    3. (c)

      In the limit as Vf.p.p. → ∞ (Vf.p.p. means the volume of the parallelepiped representing the actual crystal), one can replace

      $$ \sum\limits_{q} {f\left( \varvec{q} \right)} \quad {\text{by}}\quad \frac{{V_{{{\text{f}} . {\text{p}} . {\text{p}} .}} }}{{\left( {2\pi } \right)^{3} }}\int {f\left( \varvec{q} \right){\text{d}}^{3} \varvec{q}} . $$
    4. (d)
      $$ \frac{{\varOmega_{a} }}{{\left( {2\pi } \right)^{3} }}\int\limits_{{{\text{B}} . {\text{Z}} .}} {\exp \left( {{\text{i}}\varvec{q} \cdot \varvec{R}_{l} } \right){\text{d}}^{3} q = \delta_{{\varvec{R}_{l} ,0}} } , $$

      where the integral is over one Brillouin zone.

    5. (e)
      $$ \frac{1}{{\varOmega_{a} }}\int {\exp \left[ {{\text{i}}\left( {\varvec{G}_{l ^ {\prime} } - \varvec{G}_{l} } \right) \cdot \varvec{r}} \right]d^{3} r = \delta_{{l^{{\prime }} ,l}} } , $$

      where the integral is over a unit cell.

    6. (f)
      $$ \frac{1}{{\left( {2\pi } \right)^{3} }}\int {\exp \left[ {{\text{i}}\varvec{q} \cdot \left( {\varvec{r} - \varvec{r}^{{\prime }} } \right)} \right]{\text{d}}^{3} q = \delta \left( {\varvec{r} - \varvec{r}^{{\prime }} } \right)} , $$

      where the integral is over all of reciprocal space and \( \delta \)(r − r) is the Dirac delta function.

    7. (g)
      $$ \frac{1}{{\left( {2\pi } \right)^{3} }}\int\limits_{{V_{{{\text{f}} . {\text{p}} . {\text{p}} .}} { \to }\infty }} {\exp \left[ {{\text{i}}\left( {\varvec{q} - \varvec{q}^{{\prime }} } \right) \cdot \varvec{r}} \right]{\text{d}}^{3} r = \delta \left( {\varvec{q} - \varvec{q}^{{\prime }} } \right)} . $$

      In this problem, the ai are the primitive translation vectors. N1a1, N2a2, and N3a3 are vectors along the edges of the fundamental parallelepiped. Rl defines lattice points in the direct lattice by (2.171). q are vectors in reciprocal space defined by (2.175). The Gl define the lattice points in the reciprocal lattice by (2.173). Ωa = a1 · (a2 × a3), and the r are vectors in direct space.

  12. 2.12

    This problem should clarify the discussion of diagonalizing Hq (defined by 2.198). Find the normal mode eigenvalues and eigenvectors associated with

    $$ \begin{array}{*{20}c} {m_{i} \ddot{x}_{i} = - \sum\limits_{j = 1}^{3} {\gamma_{ij} x_{j} } ,} \\ {m_{1} = m_{3} = m,\quad m_{2} = M,\quad {\text{and}}\quad \left( {\gamma_{ij} } \right) = \left( {\begin{array}{*{20}c} {k,} & { - k,} & 0 \\ { - k,} & {2k,} & { - k} \\ {0,} & { - k,} & k \\ \end{array} } \right).} \\ \end{array} $$

    A convenient substitution for this purpose is

    $$ x_{i} = u_{i} \frac{{{\text{e}}^{{{\text{i}}\omega t}} }}{{\sqrt {m_{i} } }}. $$
  13. 2.13

    By use of the Debye model, show that

    $$ c_{v} \, { \propto } \, T^{3} \quad {\text{for}}\quad T \ll \theta_{D} $$

    and

    $$ c_{v} { \propto 3}k\left( {NK} \right)\quad {\text{for}}\quad T \gg \theta_{D} . $$

    Here, k = the Boltzmann gas constant, N = the number of unit cells in the fundamental parallelepiped, and K = the number of atoms per unit cell. Show that this result is independent of the Debye model.

  14. 2.14

    The nearest-neighbor one-dimensional lattice vibration problem (compare Sect. 2.2.2) can be exactly solved. For this lattice: (a) Plot the average number (per atom) of phonons (with energies between ω and ω + dω) versus ω for several temperatures. (b) Plot the internal energy per atom versus temperature. (c) Plot the entropy per atom versus temperature. (d) Plot the specific heat per atom versus temperature. [Hint: Try to use convenient dimensionless quantities for both ordinates and abscissa in the plots.]

  15. 2.15

    Find the reciprocal lattice of the two-dimensional square lattice shown above.

  16. 2.16

    Find the reciprocal lattice of the three-dimensional body-centered cubic lattice. Use for primitive lattice vectors

    $$ \varvec{a}_{1} = \frac{a}{2}\left( {\hat{\varvec{x}} + \hat{\varvec{y}} - \hat{\varvec{z}}} \right),\quad \varvec{a}_{2} = \frac{a}{2}\left( { - \hat{\varvec{x}} + \varvec{\hat{y} + \hat{z}}} \right),\quad \varvec{a}_{3} = \frac{a}{2}\left( {\hat{\varvec{x}} - \varvec{\hat{y} + \hat{z}}} \right). $$
  17. 2.17

    Find the reciprocal lattice of the three-dimensional face-centered cubic lattice. Use as primitive lattice vectors

    $$ \varvec{a}_{1} = \frac{a}{2}\left( {\hat{\varvec{x}} + \hat{\varvec{y}}} \right),\quad \varvec{a}_{2} = \frac{a}{2}\left( {\varvec{\hat{y} + \hat{z}}} \right),\quad \varvec{a}_{3} = \frac{a}{2}\left( {\varvec{\hat{y} + \hat{x}}} \right). $$
  18. 2.18

    Sketch the first Brillouin zone in the reciprocal lattice of the fcc lattice. The easiest way to do this is to draw planes that perpendicularly bisect vectors (in reciprocal space) from the origin to other reciprocal lattice points. The volume contained by all planes is the first Brillouin zone. This definition is equivalent to the definition just after (2.176).

  19. 2.19

    Sketch the first Brillouin zone in the reciprocal lattice of the bcc lattice. Problem 2.18 gives a definition of the first Brillouin zone.

  20. 2.20

    Find the dispersion relation for the two-dimensional monatomic square lattice in the harmonic approximation. Assume nearest-neighbor interactions.

  21. 2.21

    Write an exact expression for the heat capacity (at constant area) of the two-dimensional square lattice in the nearest-neighbor harmonic approximation. Evaluate this expression in an approximation that is analogous to the Debye approximation, which is used in three dimensions. Find the exact high- and low-temperature limits of the specific heat.

  22. 2.22

    Use (2.200) and (2.203), the fact that the polarization vectors satisfy

    $$ \sum\limits_{p} {e_{{\varvec{q}p\varvec{b}}}^{ * \alpha } e_{{\varvec{q}p\varvec{b}^{{\prime }} }}^{ * \beta } = \delta_{\alpha }^{\beta } \delta_{\varvec{b}}^{{\varvec{b}^{{\prime }} }} } $$

    (the α and β refer to Cartesian components), and

    $$ X_{{ - \varvec{q},p}}^{11\dag } = X_{{\varvec{q},p}}^{11\dag } ,P_{{ - \varvec{q},p}}^{11\dag } = P_{{\varvec{q},p}}^{11} . $$

    (you should convince yourself that these last two relations are valid) to establish that

    $$ {\mathbf{X}}_{{\varvec{q,b}}}^{1} = - {\text{i}}\sum\limits_{p} {\sqrt {\frac{\hbar }{{2m_{\varvec{b}} \omega_{{\varvec{q},p}} }}} \varvec{e}_{{\varvec{q},p,\varvec{b}}}^{ * } \left( {a_{{\varvec{q},p}}^{\dag } - a_{{ - \varvec{q},p}} } \right)} . $$
  23. 2.23

    Show that the specific heat of a lattice at low temperatures goes as the temperature to the power of the dimension of the lattice as in Table 2.5.

  24. 2.24

    Discuss the Einstein theory of specific heat of a crystal in which only one lattice vibrational frequency is considered. Show that this leads to a vanishing of the specific heat at absolute zero, but not as T cubed.

  25. 2.25

    In (2.270) show vl is longitudinal and v2, v3 are transverse.

  26. 2.26

    Derive wave velocities and physically describe the waves that propagate along the [110] directions in a cubic crystal. Use (2.269).

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG, part of Springer Nature

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Patterson, J.D., Bailey, B.C. (2018). Lattice Vibrations and Thermal Properties. In: Solid-State Physics. Springer, Cham. https://doi.org/10.1007/978-3-319-75322-5_2

Download citation

Publish with us

Policies and ethics