ssp-lecture-notes solid state physics
ssp-lecture-notes solid state physics
Lecture notes
Alexander Tsirlin
Leipzig University
February 4, 2024
These lecture notes are released under the generic Creative Commons (CC-BY-SA) license. You
are free to disseminate and re-use the full document or any of its parts by providing attribution as
follows: Alexander Tsirlin, Leipzig University with a link to the homepage of this module.
Contents
1
8. Bonding in crystals: covalent, metallic, and van der Waals. . . . . . . . 36
8.1. Covalent crystals
8.2. Metallic crystals
8.3. Polymorphism
8.4. van der Waals radius and Lennard-Jones potential
8.5. Lattice energy
8.6. Bulk modulus
9. Mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . 39
9.1. Hydrostatic pressure, Equation of state
9.2. Uniaxial pressure
9.3. Shear deformation
9.4. Stress and strain tensors
9.5. Elastic constants
9.6. Plastic deformation
10. Dielectric properties . . . . . . . . . . . . . . . . . . . . . . . . . . 44
10.1. Permittivity
10.2. Dielectric loss
10.3. Induced dipoles and polarizability
10.4. Local electric field
10.5. Clausius-Mossotti relation
10.6. Permanent dipoles, Debye relaxation
10.7. Cole-Cole plots
11. Phonons and sound . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
11.1. Underlying approximations
11.2. Monoatomic chain
11.3. Elastic waves and sound
11.4. Longitudinal and transverse phonons
11.5. Diatomic chain
11.6. Acoustic and optical phonons
12. Phonons and light . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
12.1. Refractive index and reflectivity
12.2. Fluctuating dipoles
12.3. LO vs. TO
12.4. Interaction with light, polaritons
12.5. Lyddane-Sachs-Teller relation
12.6. Infra-red active phonon modes
13. Phonons and the reciprocal lattice . . . . . . . . . . . . . . . . . . . 61
13.1. Brillouin zone
13.2. Dynamical matrix
A. Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
A.1. Maxwell’s equations
A.2. Electromagnetic waves
A.3. Longitudinal waves
B. Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
C. List of experimental techniques . . . . . . . . . . . . . . . . . . . . . 64
2
Introduction
Condensed matter represents the largest sub-field of physics. It is not without a reason, because
condensed matter, or colloquially cond-mat, covers many aspects of physics that we encounter
in our daily life. While different branches of condensed-matter research have many concepts
and experimental techniques in common, they are also highly diverse, especially when it comes
to research objects.
One generally distinguishes research on soft matter (polymers, gels, membranes, cellular
structures) from hard condensed matter or simply solid-state physics that addresses properties
of solid materials. This distinction is in fact rather subtle. It goes back to the different dynamics
of “soft” and “hard” systems. Soft matter shows pronounced dynamics at ambient conditions,
it does not have a “fixed” structure, which is integral to solids. However, soft-matter objects
may become crystalline and turn into solids, whereas solids may show certain features of a
liquid.
Solid-state materials are variable too. They can be crystalline or amorphous. Crystalline
solids feature long-range-ordered structures (crystal structures) that are typically periodic.1 By
contrast, amorphous solids lack long-range order in any form, although they still show some
organization on the short-range scale.
Traditionally, solid-state physics has been developed for crystalline materials. Some of its
findings can be extended to amorphous materials too, but many of the key concepts, such as
band structure, are properly defined for crystalline materials only. In fact, they strongly rely
on periodicity and would require some re-thinking even for aperiodic crystals.
The content of these lecture notes splits into three parts:
• Structure of crystals (Chapters 1 to 6). Here, we discuss how crystals are organized
and use symmetry for classification of crystals and their properties.
• Atoms in crystals (Chapters...). This part concentrates on lattice degrees of freedom,
namely, bonding between the atoms as well as atomic vibrations and motions, and how
all of this contributes to different crystal properties.
• Electrons in crystals (Chapters...) This part will mostly address metals and analyze
how electronic degrees of freedom (free or, more precisely, itinerant electrons) affect dif-
ferent physical properties.
Many of the crystals also have spin degrees of freedom, but we disregard those for the sake of
simplicity. Magnetic properties of solids are addressed in a separate module.
Hyperlinks are highlighted in blue color. They will usually direct you to useful web resources
or an additional information on the topic.
1
A few examples of aperiodic crystalline solids will be introduced in Chapters 2 and 5.
3
1. Bravais lattice, or how to pack a crystal?
1.1. Bravais lattice and unit cell
Periodic pattern of the crystal can be described by a Bravais lattice defined as a set of points
R = n1 a + n2 b + n3 c (1.1)
where n1 , n2 , n3 are integers and a, b, c are three vectors that do not lie in the same plane.
They are known as lattice translations or lattice vectors. The parallelepiped spun by these three
vectors is the repetition unit of the lattice or its unit cell with the lattice parameters (a, b, c)
and lattice angles (α, β, γ). It is customary to define the angle between b and c as α, the angle
between a and c as β, and the angle between a and b as γ.
Not every lattice is a Bravais lattice. Consider for example hexagonal (honeycomb) lattice,
one of the popular geometries in modern solid-state physics. This lattice is clearly periodic, but
it is not a Bravais lattice per se, because its lattice sites can not be described with Eq. (1.1).
Bravais lattice of the honeycomb lattice is constructed by connecting second neighbors, as
shown in Fig. 1.1. The unit cell then contains two sites of the parent honeycomb lattice, one
at the corner and one in the interior. This situation is known as a Bravais lattice with basis.
Here, basis is the group of atoms contained by the unit cell. Of course, it is also possible to
choose hexagon as the repetition unit of the honeycomb lattice. However, such a choice would
introduce a disparity with other lattices, such as square lattice where the repetition unit is
obviously a square. The advantage of Eq. (1.1) lies in the unified description of all periodic
structures, so it is not surprising that Bravais lattice became the cornerstone of solid-state
physics.
unit
cell
Figure 1.1: Left: Bravais lattice, lattice vectors, and unit cell. Right: honeycomb lattice is not a
Bravais lattice per se; its Bravais lattice is indicated by the red lines, with two atoms per unit cell.
4
Li Be
Na Mg Al
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg Tl Pb
Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Figure 1.2: Structures of simple metals. dhcp is the double-hexagonal close-packing, ...ABACABAC...
cubic lattice. We can also introduce the coordination number, i.e., the number of neighboring
atoms, which is as low as 6 in this case. Simple cubic lattice does not allow dense structures.
Finding a dense packing of spheres (better known as close packing in solid-state physics) is
in fact a common mathematical problem. By solving it, or by using general intuition of putting
together spherical objects like billiard balls, we know that each sphere can be surrounded by
six other spheres, resulting in a close-packed honeycomb layer. The next layer should follow
the dips formed by the first one. Different possible stackings give rise to a series of close-packed
structures, including:
• ...ABABAB... (hcp or hexagonal close packing)
• ...ABCABC... (ccp or cubic close packing, or fcc = face-centered cubic structure)
These two structures are much denser than the simple cubic lattice, thanks to the coordination
number of 12 (six neighbors in each layer, plus three neighbors in each of the two adjacent
layers). They are in fact common for simple metals, which are well described by this packing
concept because only one type of atoms is present in the crystal and because denser packing
allows higher electron concentration, which is favorable for the metallic bonding. Fig. 1.2 shows
that more than half of simple metals adopt either hcp or fcc structures. Most of the remaining
metals form a bcc (body-centered cubic) structure, which is only slightly less dense than the
close-packed ones.
5
1.3. Electron microscopy
Back in time researchers could only guess about lattice periodicity and packing. Nowadays it
is possible to visualize the crystal lattice and even see individual atoms with the direct imaging
done by electron microscopy. Electron microscopes are somewhat similar to common optical
microscopes, but they use electrons as radiation with the much shorter wavelength, and of
course they require a much more sophisticated system of electromagnetic lenses to direct and
focus the electron beams.
Electron microscopes come in two main varieties:
• transmission electron microscopes (TEM) collect electrons on a detector behind the sam-
ple, just like an optical microscope does with visible light
• scanning electron microscopes (SEM) collect secondary electrons that are produced when
electrons are scattered on the sample
SEM’s are more compact, because one does not need a long separation between the sample
and detector. They can also map a larger area by scanning the sample with the electron beam
(hence the name). Concurrently, they lack in resolution. A typical instrument of this type has
the resolution of about 10 nm and can be used for imaging microstructures but not individual
atoms. By contrast, modern TEM’s have the resolution of 1 Å. They can operate in different
imaging modes and highlight heavy or light atoms, or map out chemical composition and even
electronic states on the sub-nm scale.
Two points of concern when using electron microscopy are:
• small size of the probe; a region of only 20 − 30 nm is probed in a single image taken with
the atomic resolution. This problem is partially mitigated by STEM (scanning TEM),
but even in this case only a tiniest fraction of the sample volume can be studied
• sample damage caused by the strong electron beam and high vacuum
Electron microscopy is the method of choice for studying defects and microstructures. When
it comes to long-range-ordered crystals, electron microscopy plays a somewhat auxiliary role,
while main information about the crystal structure comes from diffraction methods (Ch. 6).
6
2. Symmetry as the guiding principle
2.1. Symmetry operations
Long-range order of a crystal reflects its underlying symmetry. Most generally, symmetry of an
object is a set of transformations that leave this object invariant. We will consider geometrical
transformations that can be represented as
Rxx Rxy Rxz tx
symmetry operation = R ⊕ t = Ryx Ryy Ryz ⊕ ty (2.1)
Rzx Rzy Rzz tz
ap bp
cp
Figure 2.1: Conventional and primitive unit
cells of the body-centered cubic lattice. The
lattice angles of the primitive cell deviate
primitive from 90◦ , so it has a lower symmetry than
the conventional cell. Note that ap , bp , cp are
also shorter than a, b, c, hence the primitive
c conventional
cell has the twice smaller volume than the
conventional one.
b
7
• Conventional cell, which is the smallest repetition unit of the highest symmetry defined
by the lattice vectors a, b, c
In a primitive lattice, both unit cells coincide. In a centered lattice, the volume of the primitive
cell is twice (A, B, C, I), three times (R), or four times (F ) smaller than the volume of the
conventional unit cell.
From this point on, we will distinguish lattice vectors a, b, c that define the conventional
unit cell, from lattice translations t that include a, b, c along with other vectors allowed by the
lattice centering.
Lattice centering is an important part of describing crystal symmetry, but it may not have
any immediate implications. We do not expect crystals with face-centered lattices to be distinct
from the body-centered ones. However, in special cases, such as simple metals from Fig. 1.2,
lattice centering determines the packing fraction and directly affects density of the crystal.
8
2.4. Reflection and inversion symmetry
Eq. (2.2) covers only some of the unitary transformations. Two further important operations
are inversion center (labeled 1̄ or −1) and mirror plane (m),
−1 0 0 1 0 0
Rinversion = 0 −1 0 , Rreflection = 0 1 0 , (2.3)
0 0 −1 0 0 −1
where the latter describes a mirror plane perpendicular to z (m ⊥ c). These symmetry elements
can be combined with the rotation axes, namely, both can be present in the crystal at the same
time.
The reflection and inversion symmetries also have important implications. They determine
polarity and chirality of the crystal. We will get back to this in Ch. 3 after we learn how
individual symmetry elements build up a symmetry group.
jα = σαβ Eβ (2.5)
Permittivity is then related to the second derivative of G with respect to E, and mixed deriva-
tives must be equal,
∂ 2G ∂ 2G
= ⇒ εαβ = εβα . (2.8)
∂Eα ∂Eβ ∂Eβ ∂Eα
A similar statement for transport properties is known as Onsager reciprocal relations. Without
going into details we only mention here that these relations need to be amended in the presence
of a magnetic field where conductivity tensor becomes antisymmetric (more on this in Ch. ??).
9
Symmetry of the tensor reduces the number of independent components from 9 to 6. Addi-
tional constraints can be derived from crystal symmetry using Eq. (2.4). For example, consider
the 4-fold rotation axis parallel to z. According to Eq. (2.2), it leads to the transformation
(x, y, z) → (y, −x, z) that converts σ into an equivalent tensor σ ′ ,
σyy −σyx σyz σxx σxy σxz
σ ′ = −σxy σxx −σxz vs. σ = σyx σyy σyz . (2.9)
σzy −σzx σzz σzx σzy σzz
The tensor components should be pairwise equal, so σyy = σxx , whereas −σyx = σxy . For a
symmetric tensor it means that all off-diagonal components vanish, leading to the final form of
σxx 0 0
σ = 0 σxx 0 (2.10)
0 0 σzz
10
3. Systematics of crystals: symmetry groups
3.1. Point symmetry operations
Only a few local (point) symmetry operations R are compatible with periodicity. Every periodic
crystal can be then assigned to one or another symmetry group that comprises a subset of the
following symmetry elements:
• Rotation axes: 2, 3, 4, 6 (C2 , C3 , C4 , C6 )
• Inversion center: 1̄ (i)
• Mirror plane: m (σv , σh )
• Rotoinversion axes: 3̄, 4̄, 6̄
Here, we use labels in the international (Hermann-Mauguin) notation, which is common in
crystallography. The symbols in brackets are Schönflies notation favored by spectroscopists.
The last element, rotoinversion axes, has not been discussed yet and requires a further
comment. This symmetry element is a rotation followed by inversion,
− cos φ − sin φ 0
Rn̄ = sin φ − cos φ 0 (3.1)
0 0 −1
The inversion center 1̄ is in fact the end member of this series, because it combines inversion
with the 1-fold rotation. Using Eqs. (2.2) and (2.3), one can verify that 2̄ = m. Three other
rotoinversion axes are independent symmetry elements, which are needed to describe special
situations, such as the symmetry of a tetrahedron (4̄).
No Schönflies symbols for rotoinversion axes exist, because a different symmetry element,
rotation-reflection axes (Sn ), is considered in this case. They are quite similar to n̄, but rotation
is followed by a mirror-plane reflection. This operation is also known as an improper rotation,
cos φ sin φ 0
RSn = − sin φ cos φ 0 (3.2)
0 0 −1
11
2/m 4mm mmm
along c. Each pair of planes generates a rotation axis, but we don’t need to list them all, and
the point group is labeled simply as mmm. It features eight symmetry elements,
mmm = { 1, 2a , 2b , 2c , ma , mb , mc , 1̄ } . (3.4)
Since one symmetry element can be obtained from a few others, it is customary to label the
symmetry group by its main generators, similar to the examples above. One usually chooses
rotation axis of the highest order and then symmetry elements perpendicular to it. Consider
the following examples:
• 4/m is the four-fold rotation along c and the mirror plane perpendicular to c.
• 4mm is the four-fold rotation along c, the mirror planes perpendicular to a and b, and
the mirror planes perpendicular to a ± b. The absence of / means that no mirror plane
perpendicular to c occurs. Moreover, the a and b directions are equivalent (orange planes
in Fig. 3.1), so they are mentioned only once, while the second m stands for the mirror
planes, which are perpendicular to the ab-diagonals (green).
There are altogether 32 point groups formed by the symmetry operations from Ch. 3.1.
The full list can be found on Wikipedia along with an explanation of the Schönflies symbols of
point groups that we will not consider here. Every point group defines its own crystal class, a
family of crystals with the similar shape that reflects the underlying symmetry of the crystal
structure.
The relation between the crystal class and crystal shape is most useful in mineralogy.
Synthetic crystals can be grown in different shapes, and they are often cut to a custom shape
as required by the experiments. The point symmetry remains important, though, because
it determines whether or not the crystal is polar. Polar direction is defined as the direction
that stays invariant under all symmetry operations of the system. Crystal or molecule with at
least one polar direction are capable of having dipole moment and electric polarization, which
is important for ferroelectric and pyroelectric properties. Inversion centers and rotoinversion
axes obviously forbid polarity. Examples of polar symmetries are 4mm and 3m, examples of
non-polar symmetries are 4/m and 3 2.
12
Cubic a = b = c, α = β = γ = 90◦ m3̄m, 4̄3m, 432, m3̄, 23
Tetragonal a = b ̸= c, α = β = γ = 90◦ 4, 4mm, 4/m, 4/mmm, 4̄, 422, 4̄2m
Orthorhombic a ̸= b ̸= c, α = β = γ = 90◦ 222, mm2, mmm
Hexagonal a = b ̸= c, α = β = 90◦ , γ = 120◦ 6, 6/m, 6mm, 6/mmm, 6̄, 622, 6̄m2
Trigonal a = b ̸= c, α = β = 90◦ , γ = 120◦ 3, 3m, 3̄, 3̄m, 32
Monoclinic a ̸= b ̸= c, α ̸= γ ̸= 90◦ , β = 90◦ 2, m, 2/m
Triclinic a ̸= b ̸= c, α ̸= β ̸= γ ̸= 90◦ 1, 1̄
Crystal system defines the form of tensor properties according to Neumann’s principle, as
explained in Ch. 2.5. The knowledge of the crystal system is also important for constructing
reciprocal lattice and analyzing diffraction experiments, as we will see in Ch. 4 and 5.
Crystal system can be combined with the lattice centering from Ch. 2.2. Back then, we
mentioned that centering only makes sense when the conventional unit cell has a higher sym-
metry than the primitive cell. Otherwise, lattice vectors could be re-defined without loss of
symmetry. Indeed, a centered triclinic lattice is redundant, because its unit cell can always be
reduced. On the other hand, both face and body-centering occur for cubic and some other crys-
tal systems where symmetry elements would be lost otherwise. Combining crystal systems and
lattice centering gives rise to 14 Bravais lattices that can be encountered in periodic crystals.
13
4
3.5. Chirality
Screw axes impart their sense of rotation to the atomic arrangement and, therefore, to the
crystal as a whole. For example 31 and 32 produce similar helical chains with the counter-
clockwise and clockwise rotations, respectively (Fig. 3.2). It is an example of chirality.
More generally, chirality is defined as the property of an object not to match its mirror
image. Mirror-plane symmetry renders an object non-chiral because the object does not change
under this operation at all. On the other hand, any symmetry group that does not contain
mirror operations (nor it contains inversion centers and rotoinversion axes, which are equivalent
to the rotation-reflection axes) is chiral. Examples of chiral point groups are 2 2 2 and 3 2. Note
that some symmetry groups are polar and non-chiral or chiral and non-polar. Both polarity
and chirality are rooted in the crystal symmetry, but different aspects of the symmetry are
relevant in each case.
Chiral objects are interesting because they exist in (at least) two forms, known as left and
right enantiomers (for molecules) and enantiomorphs (for crystals). These forms are almost
indistinguishable, yet they may show drastically different properties. For example, many of
the aminoacids, carbohydrates, and other biologically relevant molecules are chiral. Only one
enantiomer is usually present in nature, and this choice of chirality has tremendous repercussions
for biological systems.
14
or absorption. Circular birefringence is the manifestation of an optical rotation. The polar-
ization direction turns when light goes through a chiral material. This has been the earliest
experimental tool for identifying chirality. It can be observed in very common systems, such as
crystalline sugar and sugar syrup.
15
group symbol. Crystalloghers have also done a great job in enforcing standard settings and
standard notations for the crystal symmetry. You can take a peek into their very systematic
and perfectly organized world by visiting the website of International Union of Crystallography
with its own dictionary, teaching pamphlets, and a lot more.
16
4. Unpacking the crystal structure
4.1. Atomic coordinates and Wyckoff positions
Atomic positions in a crystal are defined as fractions of lattice translations,
r = xa + yb + zc, (4.1)
m
0 Figure 4.1: Selected Wyckoff positions of the
1a
space group P mm2 (No. 25). 4i is the general
2f m position not lying on any symmetry element. 2f
4i and 1a are special positions with the lower mul-
tiplicity. In this figure, we used some standard
1/2
notation of the symmetry diagrams: solid lines
show the mirror planes, whereas gray pointed
ovals are the two-fold rotation axes. You can
find more of the symbols here.
1
x 2-fold rotation
These Wyckoff positions can be general or special. A general Wyckoff position does not lie
on any point symmetry element and shows the highest multiplicity allowed by the given space
group. A special Wyckoff position lies on at least one point symmetry element, which then
does not act on this atom, resulting in a lower multiplicity. For example, the general position
in the space group P mm2 has the multiplicity of 4; it is labeled 4i. By contrast, the position
(0, 0, z) on the two-fold rotation axis has the multiplicity of 1 (it is labeled 1a) because it stays
invariant under all symmetry operations (Fig. 4.1).
Wyckoff positions have two implications. First, they help us to make the description of
crystal structures more systematic. Lists of possible Wyckoff positions are available for every
space group (again, on the Bilbao server or in the International Tables for Crystallography).
Second, the distribution of atoms over different Wyckoff positions has ramifications for crystal
properties, such as the type of phonon modes and their response in infrared and Raman exper-
iments. We will briefly touch on this issue in Ch. 12. You can also learn more by trying the
SAM utility at the Bilbao server.
6
Crystallographers use bars to indicate minus sign in front of the coordinate.
17
4.2. Crystal structure
The structure of a given crystal is described by several elements:
• lattice parameters: a, b, c and α, β, γ
• symmetry (space group) symbol
• list of Wyckoff positions occupied by atoms, with the corresponding coordinates: xj , yj , zj
You will almost never see the full list of atoms in the unit cell. Only the Wyckoff positions (i.e.,
positions unrelated by symmetry) will be given. Other atoms can be generated using symmetry
elements.
Let’s address several beginners’ questions about crystal structures:
Where to find crystal structures? In databases that together contain over a million of
crystal structures that have been determined experimentally. You can use Crystallography
Open Database (free), Karlsruhe database or ICSD (commercial, inorganic compounds), and
Cambridge Structural Database or CSD (commercial, organic compounds). Each of these
databases contains the so-called cif-files that summarize structural information in a common
format.
How to read crystal structures? Experts can learn a lot by reading the cif-file as plain text.
However, most people will find it much easier to draw the crystal structure using software like
VESTA. Then you immediately see all atoms, their positions relative to the unit cell and lattice
translations, etc.
How to understand crystal structures? The answer to this question depends on the in-
formation that you seek to obtain. There are generally three ways of thinking about crystal
structures, sometimes they are called structural models:
• close packing, works for relatively simple compounds, such as monoatomic crystals dis-
cussed in Ch. 1. We will see some further applications in the ionic crystals too (Ch. 7).
• ball-and-stick means that you connect atoms, which are sufficiently close to each other,
and indicate chemical bonds. In this way, one finds connectivity of the crystal, whether
it is built of molecules, chains, layers... This approach is most fruitful in covalent and van
der Waals crystals (Ch. 8).
• polyhedral means that you connect an atom to its neighbors and treat the resulting unit
as a rigid body. Then you analyze connectivity of such polyhedra. This approach is
especially useful in more complex ionic crystals that can not be described as closed-packed
structures.
18
Figure 4.2: Bragg’s law. Reflection of waves
θ θ from parallel atomic planes gives rise to a con-
structive interference when the path difference is
d θ path difference
equal to an integer number of wavelengths. Ex-
perimentally, the angle 2θ between the incident
and scattered waves is measured.
19
dhkl
c /l dhkl
a/h d100
b /k d200
Figure 4.3: Left: definition of Miller indices h, k, l. The leftmost plane goes through the origin. An
adjacent plane crosses the lattice vectors a, b, c at da = a/h, db = b/k, and dc = c/l. The distance
between the planes is dhkl . Middle and right: the (200) planes are similar to the (100) planes but
show a twice smaller spacing.
XRD is a much more versatile technique compared to electron microscopy (Ch. 1.3). It does not
destroy the sample and does not require high vacuum. It is compatible with different sample
environments and can be performed in a broad range of temperatures, pressures, electric and
magnetic fields.
The Miller index of zero means that the lattice plane never crosses the respective axis, i.e.,
it is parallel to this axis. Three faces of the unit cell will then have indices of (100), (010),
and (001). Increasing the Miller indices shortens the spacing between the planes. For example,
(200) are the same lattice planes as (100), but with the twice shorter spacing (Fig. 4.3).
The Miller indices must be integer to comply with periodicity of the lattice. Later we will
come to see that they also describe a position in the reciprocal space and can be arbitrary in
this sense. However, only integer values of h, k, l describe something that matches periodicity
of the crystal.
The Miller indices can be used to determine the d value in Bragg’s law. For a crystal with
α = β = γ = 90◦ ,
1 h2 k 2 l2
2
= + + . (4.4)
dhkl a2 b2 c2
We will prove this relation shortly (Ch. 5.3). For now let’s explain how the actual experiment
works. An XRD measurement returns a set of peak positions θi that can be re-calculated into
di ’s using Bragg’s law. Then one has to find the values of a, b, c, α, β, γ such that every di is
described by some integer values of h, k, l. This procedure is known as indexing of the Bragg
peaks. A successful indexing returns lattice parameters of the crystal.
20
4.6. Crystal faces and crystal directions
It should be clear from the above that crystal faces (and, consequently, surfaces) are labeled
with the same Miller indices h, k, l. Here, one considers a single plane only, so there will be no
difference between (100), (200), (300), and so on, hence the smallest indices are always used.
The notation of lattice planes and crystal faces should not be confused with the notation
of crystal directions, even though three integer numbers u, v, w are also involved in this case,
R[u v w] = u a + v b + w c. (4.5)
In simple cases, these vectors R are perpendicular to the lattice planes with the same indices.
For example, viewing a cubic crystal along the [100] direction means that you look at the (100)
face of the cube. However, this is not true in general. We will see very soon that the h, k, l
values define a vector in the reciprocal space, whereas Eq. (4.5) corresponds to a real-space
direction. Note also the different type of brackets used for the (lattice planes) vs. [crystal
directions].
21
5. Spaces of crystallography: Reciprocal lattice
5.1. Laue condition
Bragg’s law is so simple that it does not describe all features of the scattering process. It does
not account for the presence of different atoms, nor for the different reflection intensities, and it
treats atomic planes as simple mirrors that reflect the incident beam in one direction. A more
realistic description has been proposed by Laue who considered interference of waves scattered
on two different atoms inside the crystals.
k
n’
Figure 5.1: Laue condition of the construc-
n
tive interference. The waves scattered by two
atoms, which are separated by the lattice vec-
tor R, acquire a path difference that should
θ R be a multiple of the wavelength λ. The uni-
θ’ tary vectors n and n′ show the directions of
k and k′ , the propagation vectors of the in-
path cident and scattered waves, respectively.
difference
k’
Without loss of generality, we consider k and k′ , the propagation vectors of the incident
and scattered waves, respectively. If two atoms are separated by a vector R, the path difference
between the two scattered waves becomes (Fig. 5.1)
where n and n′ are unitary vectors along k and k′ , respectively. Using the definition of the
propagation vector, k = (2π/λ) n and k′ = (2π/λ) n′ , the interference condition can be written
as
λ
R(k′ − k) = mλ ⇒ R (k′ − k) = 2πm (5.2)
2π
with integer m. This equation is known as the Laue condition. It should hold for every lattice
vector R defined by Eq. (1.1), because a pair of atoms can be found for every such vector.
eiGR = 1 (5.3)
for every lattice vector R. It is common to say that R’s belong to the direct lattice of the
crystal in real space, whereas G’s occur in the reciprocal space.
Reciprocal lattice can be constructed explicitly by choosing the vectors
[b × c] [c × a] [a × b]
a∗ = 2π , b∗ = 2π , c∗ = 2π . (5.4)
a · [b × c] a · [b × c] a · [b × c]
a∗ a = 2π, a∗ b = 0, a∗ c = 0 (5.5)
22
c (100) b Figure 5.2: Direct-lattice and reciprocal-
c*
lattice vectors for the orthogonal (right)
b*
and non-orthogonal (left) lattice vectors.
β Note that the reciprocal-lattice vectors are
a a not always parallel to those of the direct
a* a*
lattice, yet they are perpendicular to the
d10
0
corresponding lattice planes: for example,
a∗ is perpendicular to the (100) planes.
Now, a∗ and c∗ are no longer parallel to a and c (Fig. 5.2). Instead, they form an angle of
β ∗ = 180◦ − β.
We should also note that a · [b × c] = V , the unit-cell volume in real space according to the
standard definition of the triple product. To obtain unit-cell volume in the reciprocal space,
one needs vector identities,
that return
2π 2π ∗ (2π)3
V ∗ = a∗ · [b∗ × c∗ ] = [b × c] · [b∗ × c∗ ] = b · [c∗ × [b × c]] = . (5.9)
V V V
This result will be needed in the future when we do summations over the reciprocal space.
At this point, reciprocal lattice may still look like a rather arbitrary mathematical con-
struction, but it is in fact undeniably real because it is observed in every scattering experiment:
intensity maxima appear at every or almost every (see Ch. 6.3) reciprocal-lattice site. We can
also say that each crystal lives two parallel lives. One is in real space and constitutes tangible
crystal properties. Another one is in the reciprocal space and incorporates intrinsic effects such
as atomic vibrations and electronic transitions. Only by grasping reciprocal-space phenomena
can one understand real-space properties of the crystal!
23
b
the unitary vector perpendicular to these planes. Now choose G = 2πn/dhkl and consider it as
a propagation vector of a wave with the wavelength of dhkl . This wave has the form eiGr . One
expects eiGr = 1 at r = 0. Since r = 0 corresponds to one of the lattice planes, the condition
eiGr = 1 should also hold on any other lattice plane because they are separated by an integer
number of wavelengths (Fig. 5.3). Each point R of the direct lattice belongs to one or another
lattice plane. Therefore, eiGR = 1 for every R, and the condition (5.3) is fulfilled. Then G
must be a reciprocal-lattice vector.
Now we have to prove that the reciprocal-lattice vector G = ha∗ + kb∗ + lc∗ corresponds to
the planes with the Miller indices h, k, l. Let this vector define some lattice planes, which are
perpendicular to it. We know that the corresponding interplane distance is given by |G| = 2π/d.
The distance between the origin and the nearest plane is a vector of the length d in the direction
of G (Fig. 5.3), namely, d = d G/|G| = d2 G/(2π). To determine the distance da that this
plane intersects on the a-axis, one has to compute
d d 2πa 2πa a
da = = = = ∗
= , (5.10)
cos φ d · a/(da) G·a ha · a h
which is equivalent to Eq. (4.3) that served as the definition of the Miller indices h, k, l.
We now realize that the lattice planes used in Bragg’s law are real-space manifestations of
the reciprocal lattice. Incident light sees crystal as an optical grating, with different gratings
identified by different families of the lattice planes. There need not be atoms on a given lattice
plane to produce a Bragg peak. The positions of these Bragg peaks define the reciprocal lattice
and directly convey lattice parameters of the crystal.
This statement also sheds light on Eq. (4.4) for dhkl that immediately follows from |G|2 =
h2 |a∗ |2 + k 2 |b∗ |2 + l2 |c∗ |2 as the vector length in the reciprocal space (assuming a, b, c are
mutually orthogonal). The calculation of dhkl for an arbitrary lattice also becomes straight-
forward, albeit tedious when non-90◦ angles have to be taken into account.
24
vectors,
ti = p1 a∗ + p2 b∗ + p3 c∗ . (5.11)
When p1 , p2 , p3 are simple fractions like 21 or 15 , the structure is called commensurately modulated.
It is nothing but a periodic crystal with the larger unit cell. For example, p1 = 12 would mean
that a∗ should be twice shorter, hence the lattice parameter a should be twice longer: some
element of the structure develops a twice longer periodicity than the rest of the crystal and
requires a two-fold expansion of the unit cell. By contrast, a structure with random values of
p1 , p2 , p3 is incommensurately modulated.
Special mathematical formalism has been developed for modulated structures. It is based
on a proper (periodic) lattice defined in a 3 + n-dimensional space known as superspace where
modulation vectors ti are accommodated along additional, artificial dimensions, in order to
restore periodicity of the system. For example, quasicrystals can be described as periodic
structures in the 5D or 6D reciprocal space. This space is, of course, unphysical, but its
3D projection forms the physical reciprocal space where diffraction pattern is observed. The
introduction of additional dimensions may seem bizarre at first glance, but it becomes more
palatable if one considers that crystallographic restriction theorem forbids 5-fold symmetry
only in 3D space. In higher dimensions, five-fold rotations may be compatible with periodicity.
Incommensurate modulations may have different origin depending on the chemical nature
of the crystal. Sometimes it is related to deformations or rotations of structural units that
occupy a certain position in space, but vary with a different periodicity compared to the rest
of the lattice. Other examples of aperiodic crystals are framework structures with channels
where atoms in channels have their own periodicity compared to the framework. Aperiodic
crystals are inconspicuous but abundant. They may occur in elemental solids, including Bi and
Te under pressure, or in such a common material as Na2 CO3 , better known as washing soda.
A short memorial article can serve as a good introduction into aperiodic crystals.
25
6. Structure factor: All shades of diffraction
6.1. Structure factor
Reciprocal lattice serves as a connection between positions of Bragg peaks and lattice parame-
ters of the crystal. Similarly, structure factor connects intensities of Bragg peaks to the atomic
coordinates.
To calculate intensity of a Bragg peak, we go back to Laue’s representation of the scattering
process and take the scattering mechanism into account. Atoms are not point-like objects, and
atoms are not scatterers per se. Instead, one should consider periodic scattering density ρ(r)
that satisfies ρ(r + R) = ρ(r) for every lattice vector R.
Every small volume element dr scatters the incident wave. The phase of the scattered wave
depends on the position of dr inside the crystal. Indeed, from Laue’s argument (Ch. 5.1) the
path difference for a given position r relative to the origin is
λ λ
r(k − k′ ) = Gr (6.1)
2π 2π
and the corresponding phase shift is ∆φ = G r where G is a reciprocal-lattice vector. The
scattered wave is obtained by integrating such phase shifts over the unit cell, taking into
account ρ(r) as the concentration of scatterers in a given position r,
Z
i(k′ r̃−ωt) ′
e · ρ(r) eiGr dr = ei(k r̃−ωt) F (G) (6.2)
where the newly introduced structure factor F (G) is the amplitude of the scattered wave that
travels away from the crystal into some arbitrary position r̃. Intensity is proportional8 to the
squared amplitude, Z
I(G) ∼ |F (G)|2 , F (G) = ρ(r) eiGr dr. (6.3)
UC
Considering the periodicity of ρ(r), it is sufficient to integrate over the unit cell (UC) when
calculating F (G), because other unit cells will simply lead to an additional pre-factor, which
should be the same for every reflection.9
It is instructive to compare the structure factor, Eq. (6.3), with the Fourier decomposition
of the scattering density,
Z
X
iGr 1
ρ(r) = ρG e , ρG = ρ(r) eiGr dr, (6.4)
G
V
where we restricted the summation to the reciprocal-lattice vectors G because ρ(r) is periodic
in direct space. In fact, structure factor is merely the Fourier component of the scattering
density, up to a pre-factor.
26
5
1 r’ Figure 6.1: Scattering density of the crys-
r₁ 3 tal, ρ(r), is represented as a superposi-
r tion of the scattering densities of individ-
ual atoms, ρj (r′ ), with the vector r de-
2 composed into r1 + r′ .
4 6
of scattering densities ρj (r′ ) from individual atoms. Consider j = 1, . . . N atoms located at the
positions r1 , . . . rN within the unit cell. By introducing the volume element dr′ within an atom,
we represent r = rj + r′ (Fig. 6.1) and re-write the structure factor as a sum over individual
atoms,
Z N Z N
iGr′
X X
iGr iGrj ′ ′
F (G) = ρ(r) e dr = e ρj (r ) e dr = fj (G) eiGrj . (6.5)
j=1 atom j=1
Here, we introduced the atomic form factor fj (G) that describes the scattering of the wave by
a given atom j in the direction of G.
To a first approximation, any given atom scatters waves in the same way regardless of the
chemical environment of this atom. Then the same set of atomic form factors can be used for
all crystals, and the variation of the Bragg peak intensity with G arises from two ingredients:
• material-specific: positions of atoms rj within the unit cell
• generic: angular dependence of the scattering by a given atom, as defined by fj (G),
regardless of the exact nature of the crystal
(note an immediate advantage of fractional atomic coordinates in conjunction with the recip-
rocal lattice!) We can now write the structure factor as
N N/2 h i
1 1 1
fj (G) e2πi(hxj +kyj +lzj ) + e2πi[h(xj + 2 )+k(yj + 2 )+l(zj + 2 )]
X X
iGrj
F (G) = fj (G) e =
j=1 j=1
N/2
X
fj (G) e2πi(hxj +kyj +lzj ) 1 + eπi(h+k+l)
= (6.7)
j=1
27
intensities. Only reflections with h + k + l = 2n (even) are observed. This is the reflection
condition for a body-centered crystal. The systematic absence of reflections is known as an
extinction, so we could also introduce an extinction condition h + k + l = 2n + 1 (odd).
ii) Open symmetry elements. Glide planes and screw axes cause their own extinctions.
Consider the glide plane a ⊥ c. It leads to the transformation (x, y, z) → (x + 21 , y, z̄). Then
the structure factor becomes
N/2 h i
2πi[h(xj + 21 )+kyj −lzj ]
X
2πi(hxj +kyj +lzj )
F (G) = fj (G) e +e . (6.8)
j=1
This expression does not look very spectacular, but it becomes simpler in the special case of
l = 0. Then,
N/2
X
fj (hk0) e2πi(hxj +kyj ) 1 + eπih
Fhk0 = (6.9)
j=1
and we expect the reflection condition hk0, h = 2n for the a-glide plane perpendicular to c.
Reflection conditions for a given space group can be obtained from the HKLCOND utility
at the Bilbao Server. By analyzing extinctions, crystal symmetry can be determined or, more
precisely, narrowed down to a list of possible space groups.
More generally, one can see the reflection (extinction) conditions as a manifestation of the
periodicity. The reciprocal lattice defined using three lattice vectors, Eq. (5.4), corresponds to
the situation when no additional translations are allowed. Crystals with lattice centering have
a shorter periodicity indicated by the primitive cell, so their reciprocal lattices should be more
sparse. This effect is achieved by removing some of the reciprocal-lattice points via extinctions.
For example, reciprocal lattice of a bcc lattice is an fcc lattice with the periodicity of 4π/a.
10
More precisely, lattice centering can be determined, and a subset of possible space groups defined. Space
groups that contain neither lattice centering nor open symmetry elements do not have any reflection conditions
and can not be distinguished in this way.
28
ρ(r) f(q)
neutron
(nuclear)
electron density
x-ray
Figure 6.2: Scattering densities (left) and the resulting atomic form factors (right) for the x-ray and
neutron scattering. Different densities are not drawn to the scale. Spin density is typically associated
with the d- and f -atomic orbitals that have nodes at r = 0.
where q is a vector in the reciprocal space (the vectors q are no longer restricted to the
reciprocal-lattice vectors G, because atom is not periodic). Spherical symmetry of an atom
implies that f depends only on |q| and not on the direction of q. It is then customary to
represent |q| as 2π/d similar to Ch. 5.3 and consider 1/d ∼ sin θ/λ from Bragg’s law. Therefore,
atomic form factors are tabulated as a function of sin θ/λ, measured in Å−1 . They can be found
in the International Tables for Crystallography and on the web.
The scattering density in Eq. (6.10) depends on the nature of the wave being scattered.
Several cases are of special importance (Fig. 6.2):
• x-rays are scattered by electrons, so ρ(r) is electron density that gradually decreases
away from the atom. Then f (q) is also a slowly decreasing function. Larger values of
|q| correspond to smaller values of dhkl that, in turn, correspond to the higher angles θ.
Therefore, one sees lower intensities in XRD patterns at high angles
• neutrons are scattered by the atomic nuclei, which are very small. Then, ρ(r) looks
more like a δ-function, while its Fourier transform is a constant. In a neutron diffraction
experiment, reflection intensities only weakly depend on the angle θ.
• neutrons are also scattered by the spin density, namely, by valence electrons with an
unpaired spin. This is the primary experimental tool for studying magnetic order in
crystals. The corresponding magnetic form factor decreases even faster than the x-ray
one because only valence electrons are involved, so magnetic Bragg peaks are observed at
low angles only
• electrons are scattered by other electrons, so their f (q) is similar to that of x-rays. The
intensities decrease with the angle. However, electrons are impatient and will usually
scatter several times as they travel through the crystal. Intensities of Bragg peaks in
electron diffraction are, therefore, strongly affected by multiple scattering.
29
of neutrons to magnetic order, the feature that x-rays lack.11
Neutron diffraction experiments have major downsides too. Neutrons can not be produced
in the lab, they are obtained either from a nuclear reactor or from a so-called spallation source.
Both are huge installations, and only a handful of neutron research facilities exist around the
world. Neutron beams are hard to focus, so large instrumentation is needed, and sample volume
is usually much larger than in the case of x-rays. Tiny objects like thin films are hard to study
using neutrons, and spatial resolution is much lower.
11
Unless experiments are performed at the x-ray energy that matches the absorption edge, see resonant x-ray
scattering.
30
7. Bonding in crystals: ionic
7.1. Types of chemical bonding
Many of the crystal properties can be understood from the perspective of bonding between the
atoms. Four main types of chemical bonding are usually distinguished:
• ionic, due to electrostatic forces between charged units (ions)
• covalent, due to an overlap of atomic orbitals
• metallic, due to shared (itinerant) electrons
• van der Waals, due to dipole-dipole and other weak interactions, such as London disper-
sion force (between induced dipoles)
One often identifies a separate group of hydrogen bonds, which occur between atoms with
small, fractional charges, such as H and O in water and organic molecules. These bonds are
somewhat stronger than the typical van der Waals bonds, yet for our purpose they fall into
the same group of weak bonding. We will also discuss molecular crystals where finite units
(molecules) formed by covalent bonds are held together by van der Waals bonds.
where the subscripts (s) and (g) denote the solid and gas states, respectively.
Lattice energies are obtained from calorimetry experiments where the amount of heat re-
leased or absorbed in a given process is measured. Sublimation enthalpy of a metallic, covalent,
or molecular crystal is a direct measure of its lattice energy.13 More often, though, the quan-
tity measured in the experiment is the formation enthalpy ∆f H, which is neither cohesive nor
lattice energy of the crystal:
This formation enthalpy can be related to the lattice energy by considering all intermediate
processes (Fig. ??) and adding up relevant energies: sublimation energy of the Na metal,
ionization potential of the Na atom, bond dissociation energy of the Cl2 molecule, and electron
12
Older textbooks, including the one by Ashcroft and Mermin define cohesion energy simply as lattice energy
and describe the formation of crystal from its constituents as cohesion.
13
In thermochemistry, enthalpies are commonly used instead of energies because measurements are performed
at a constant pressure rather than constant volume. Then, strictly speaking, lattice enthalpy is obtained. This
difference is usually unimportant, since solids show only a minor difference between energy and enthalpy owing
to the small thermal expansion, see also Ch. ??.
31
E tot
affinity of the Cl atom. All these quantities can be measured in separate experiments, and
eventually lattice energy of NaCl can be obtained from ∆f H of NaCl. This method is known
as the Born-Haber cycle. Thermochemistry data can be found in multiple sources, such as
NIST Chemistry WebBook and CRC Handbook of Chemistry and Physics.
It is also helpful to represent crystal energy as a function of an effective interatomic distance,
E (r), as shown in Fig. 7.1. Then the position of the energy minimum yields the equilibrium
distance r0 and, consequently, the lattice parameter of the crystal. The depth of the minimum
is the lattice energy E lat .
with the pre-factor 21 to avoid double-counting. This summation will include several terms
according to the number of different atomic positions in the crystal. In a simple AB crystal,
symmetry requires that α+ = α− = α. Using z+ = −z− = z, one finds
αz 2 e2
E Coul = − (7.4)
4πε0 r
14
Some instructive examples of the conditional convergence of electrostatic energy in ionic crystals can be
found here and here.
32
This energy does not have a minimum, and indeed an ionic crystal with only Coulomb
forces should collapse. An energy minimum in the vein of Fig. 7.1 appears when Coulomb
energy is augmented by a repulsive energy,
αz 2 e2 Crep
E tot (r) = − + m (7.5)
4πε0 r r
where Crep = const and m is also a constant that takes values between 6 and 10 depending on
the ions. This repulsive potential is empirical in nature and mimics the intuitive understanding
that different atoms cannot penetrate into each other. Beyond this common intuition, the
repulsion goes back to the Pauli exclusion principle, as explained here.
It shows the change in the crystal volume under pressure (more on this in Ch. 9.1). The
bulk modulus can be obtained from E tot if one considers thermodynamic definition of pressure,
p = −(∂E /∂V )T . The bulk modulus is essentially the second derivative of E tot with respect
to V . In contrast to lattice energy, Eq. (7.7), the exact expression for the bulk modulus depends
on the structure type, which defines the relation between V and r in the crystal.
Let’s choose rocksalt-type structure (Ch. 7.6) as an example. The distance r is the separa-
tion between the cation and anion, r = a/2. Then, V = a3 /4 = 2r3 (volume per formula unit)
and
d 1 d d dE tot r d 1 dE tot
= 2 ⇒ B=V = . (7.9)
dV 6r dr dV dV 18 dr r2 dr
33
This value should be taken at r = r0 where dE tot /dr = 0. Then, the equation simplifies to
1 d2E tot m − 1 αz 2 e2
B= = (7.10)
18 r0 dr2 r=r0 18 4πε0 r04
34
Zn
Cl Cs
S
Na Cl
Figure 7.2: Common structure types of ionic crystals: zinc blende (fcc lattice), rocksalt (fcc lattice),
and CsCl-type (primitive cubic lattice).
35
8. Bonding in crystals: covalent, metallic, and van der Waals
8.1. Covalent crystals
The formation of covalent bonds can be probed using the covalent radius.16 When rAB ≤
rA +rB , a covalent bond occurs between the atoms A and B. Shorter distances indicate stronger
covalent bonds. This is well known from carbon in organic molecules where single (1.54 Å),
double (1.34 Å), and triple (1.20 Å) C–C bonds can be distinguished by their typical length and
chemical environment. Covalent bonds in solids are not always identifiable as single, double,
or triple, but the same trend holds.
Complex nature of the covalent bonds does not allow a generic description in the same vein
as Eq. (7.5). We can only say that shorter covalent bonds should lead to the increased lattice
energy and higher bulk modulus. The typical lattice energies of covalent crystals are on the
order of 10 eV/atom and the bulk moduli are in the range of 50 − 200 GPa.
3 ℏ2 2
E lat = (3π 2 ne ) 3 , (8.1)
10 me
so it depends on the electron concentration ne and decreases with increasing the lattice param-
eter.
Metals can have lattice energies as low as 1 eV/atom and feature bulk moduli of several GPa
only. Some of the elemental metals have melting points close to room temperature (cesium,
gallium), whereas mercury is even a liquid. Nevertheless, other metals like tungsten are char-
acterized by very high melting points and low compressibilities, exceeding those of the typical
ionic crystals. It all depends on the electron concentration.
8.3. Polymorphism
In elemental solids, the type of the crystal structure will be usually determined by chemical
bonding. Metals tend to adopt simple and dense structures (see Ch. 1.2), non-metals will
often develop similar structures but with the packing of weakly bonded atoms (noble gases)
or diatomic molecules (hydrogen, oxygen). Elements spanning the boundary between metals
and non-metals show the most diverse behavior, as they can form different types of covalently-
bonded structures and simultaneously appear in the metallic form. Tin is a celebrated example.
It forms metallic crystals of white tin (body-centered tetragonal structure) that abruptly trans-
form into non-metallic gray tin (diamond structure) on cooling. These two forms of tin are
called allotropes. Allotropes are also known for carbon, sulphur, phosphorous.
Binary and more complex compounds will usually stick to one type of bonding and differ
only in their structural details, although these details are by no means unimportant. They can
have major implications for observed properties. One usually distinguishes:
• polymorphs as different structural forms of the same chemical compounds. They are
labeled with Greek letters: α-Sn, β-Sn, etc. Strictly speaking, allotropes can be also
considered as polymorphs, as we just did for tin, and so all researchers do, especially
16
The term atomic radius is somewhat more loosely defined, but most often it implies the covalent radius.
36
when they deal with high-pressure structures of elemental solids. The somewhat old-
fashioned word “allotrope” is basically reserved for different forms of elemental solids
observed at ambient pressure.
• enantiomorphs as left and right forms of a chiral structure (see Ch. 3.5)
• polytypes as different structures arising from different stacking sequences. Polytypes are
common in layered structures like graphite, but they could also occur in structures derived
from close-packed layers, as in Ch. 7.7. Labels of polytypes include the number of layers
per unit cell and the indication of symmetry (cubic, hexagonal, rhombohedral). For
example, zinc blende and wurtzite are the 3C- and 2H- polytypes of ZnS. Such polytypes
are often observed in binary semiconductors.
37
where r is the nearest-neighbor interatomic distance and we introduced lattice sums A6 and
A12 . They are similar in their meaning to the Madelung constant (Ch. 7.3), yet much easier to
calculate because the series converge very fast. For an fcc crystal, A6 = 14.45 and A12 = 12.13.
Setting the derivative of E tot to zero yields the equilibrium distance,
" 6 12 # 1
dE tot 6A6 σ 12A12 σ 2A12 6
= 0 ⇒ 2ϵ − = 0 ⇒ r0 = σ (8.6)
dr r0 r0 r0 r0 A6
so lattice energy solely depends on ϵ. This is certainly more elegant than the ionic case.
An interesting feature of van der Waals crystals is that they become more stable on in-
creasing the lattice parameter. This is because A increases with the atomic radius, as the
atoms become more polarizable. The B value should increase too to ensure the increase in
1
σ = (B /A ) 6 and the lattice parameter, but ϵ = A 2 /4B increases concurrently, because it con-
tains A 2 . Larger atoms and molecules develop stronger van der Waals bonds. For example,
melting points increase across the family of noble gases from neon to xenon, and likewise in-
crease across halogens from F2 to I2 . On the absolute scale, lattice energies of van der Waals
crystal remain quite low, though, typically in the range of 10 − 200 meV/atom.
There is now the ratio of ϵ and σ 3 , so the trend may be less intuitive than in the case of
energy, but nevertheless bulk modulus of van der Waals crystals will typically increase with
increasing the lattice parameter. Iodine is less compressible than solid bromine. The absolute
values of the bulk modulus are typically below 10 GPa and even below 1 GPa for some of the
lighter atoms and molecules.
38
9. Mechanical properties
Starting from this chapter, we will go through different crystal properties and discuss the main
parameters that describe them. We will also see how these properties are related to microscopic
aspects of the crystals, especially the atoms and their chemical bonding that constitute lattice
degrees of freedom.
V
Murnaghan
3!"-order Birch-Murnaghan
39
Fx
ΔLx /2 ΔLy Fy
Lx ΔLy /2 Lx
Ax
Figure 9.2: Hydrostatic compression (left), uniaxial pressure (middle), and shear deformation (right).
Uniaxial strain is defined as ϵx = ∆Lx /Lx , whereas shear strain is defined as ϵxy = tan φ ≃ φ.
This equation contains the same three parameters as Eq. (9.3), but entails a different underlying
approximation because no linear pressure dependence of B is assumed.
All equations of state show that volume decreases with pressure and develops a positive
curvature, as shown in Fig. 9.1, because crystals harden on compression. Direct measurement of
this V (p) curve using high-pressure XRD (i.e., x-ray diffraction on a crystal placed into a pres-
sure cell) is the main experimental tool for studying compressibility of solids and determining
their bulk moduli.
σx = Y ϵx (9.5)
40
9.3. Shear deformation
Force can be applied not only perpendicular to the sample surface, but also parallel to it, leading
to a so-called shear deformation that changes sample shape without changing its volume. Such
shear stress τxy = Fy /Ax is defined in the same way as the uniaxial stress, whereas shear strain
ϵxy = ∆Ly /Lx = tanφ is the relative displacement of the sample surface caused by Fy . For low
displacements, tanφ ≃ φ, and shear strain is measured simply as angle (Fig. 9.2, right).
The same linear relation between stress and strain holds in this case,
with the shear modulus G that complements B and EY in describing elastic properties of
materials.
Up to this point we never referred to periodicity of the crystals. In fact, all of the above
applies to any solid, be it crystalline or amorphous. In engineering, one often considers an
isotropic medium, namely, a medium that shows the same behavior along all the directions.
Such an isotropic medium is characterized by
Its elastic properties are then fully described by only two parameters. This is the case for a
typical amorphous material like glass. Note however that Eq. (9.7) does not hold for crystals,
because even cubic crystals are not isotropic. Their [100], [110], and [111] directions are distinct,
as they are not related by any symmetry.
They have the same units of pressure (Pa) as all the elastic moduli. Elastic constants describe
mechanical response of anisotropic media, including crystals. Technically, Cijkl is a fourth-
rank tensor (elasticity tensor) with 81 components, but symmetries of the stress and strain
tensors allow several simplifications. Indeed, with only six independent components in σ and
ϵ, respectively, it becomes convenient to introduce the aliases,
xx yy zz xy xz yz
1 2 3 4 5 6
and consider 36 elastic constants from C11 to C66 .
Elastic constants can be also defined in a thermodynamic fashion as second derivatives of
the elastic energy (energy acquired by the crystal due to its deformation),
1X
E = E0 + Cijkl ϵij ϵkl . (9.9)
2 ij,kl
Mixed second derivatives are equal, so Cijkl = Cklij (see also Ch. 2.5), thus reducing the number
of independent elastic constants to 21. Symmetry constrains them further. For example,
41
elasticity tensor of a cubic crystal takes the form
C11 C12 C12 0 0 0
C12 C11 C12 0 0 0
C12 C12 C11 0 0 0
C= 0
(9.10)
0 0 C44 0 0
0 0 0 0 C44 0
0 0 0 0 0 C44
In Ch. 7 and 8, we saw that bulk modulus depends on the chemical bonding in crystals.
Stronger bonds give rise to less compressible crystals. The same would be true for all the elastic
constants. They are determined by the type of the crystal structure and the nature of chemical
bonds.
Experimentally, elastic constants are determined from resonant ultrasound spectroscopy.
Single crystal of the material is placed between two plates. Mechanical vibration is induced in
one of these plates and transmitted to the opposite plate that shows vibrations at those frequen-
cies where external modulation resonates with own frequencies of the crystal. These frequencies
depend on the elastic constants of the material, but also on the shape and dimensions of the
crystal. With the sufficient number of the resonant frequencies, all the 21 elastic constants can
be determined and further measured as a function of temperature or other external parameters.
42
Materials characterized by a broad region of the plastic deformation are called ductile, in
contrast to brittle materials that break close to the yield point. Ductility is usually defined
with respect to tensile strain. The robustness of a material toward plastic deformation upon
compressive or shear strain is called malleability. It is of crucial importance for metals and
shows whether they can be rolled and bent. Finally, hardness is defined as the robustness of a
material against indentation.
Microscopically, plastic deformations are related to the propagation of dislocations, linear
defects that hinder periodicity of the crystal. The motion of dislocations shifts parts of the
crystal relative to each other and increases the crystal length. Dislocations are especially
abundant in metallic crystals where itinerant electrons are responsible for the chemical bonding,
and relative positions of the atoms are less important. Therefore, metals, in contrast to the
covalent and ionic crystals, are amenable to plastic deformations. On the other hand, alloys like
steel where iron atoms are interspersed with carbon atoms, are less ductile because impurity
atoms block the motion of dislocations.
Importantly, only elastic properties are intrinsic properties of the crystal, i.e., they are fully
determined by the structure of the crystal, its constituent atoms and chemical bonds. On the
other hand, ductility and other properties related to the plastic deformation are rooted in the
microstructure, a combination of impurities and defects that affect the motion of dislocations.
43
10. Dielectric properties
10.1. Permittivity
Response to the external electric field E is described by permittivity (ε) of the crystal. This
parameter shows the onset of the electric polarization P and of the corresponding displacement
field D,
D = εε0 E = ε0 E + P, (10.1)
as local dipoles build up in the crystal and/or become aligned with the applied field.
Polarization is defined as the dipole moment per unit of volume, P = nd µd where nd is the
dipole concentration and µd is the moment of a single dipole. P has the units of C/m2 and
also serves as a measure of charge per area, but one should remember that this charge is always
compensated by the same charge of the opposite sign sitting somewhere else in the crystal.
It is also customary to define dielectric susceptibility,
P
χe = ⇒ ε = 1 + χe . (10.2)
ε0 E
Both ε and χe are 3 × 3 tensors19 that reduce to scalars in cubic crystals.
Both dielectric susceptibility and permittivity are frequency-dependent complex numbers.
One typically separates them into the real and imaginary parts,
that show, respectively, the polarization in-phase and out-of-phase with the oscillating electric
field, E = E0 e−iωt . Indeed,
P = χ′e E0 e−iωt + χ′′e E0 ie−iωt = (ε′ − 1)E0 e−iωt + ε′′ E0 ie−iωt . (10.4)
At low frequencies, ε′ approaches static permittivity εst , whereas ε′′ vanishes. At high
frequencies, ε′ approaches 1 (vacuum permittivity), because electric field oscillates too fast to
induce any polarization, and ε′′ again vanishes.
44
in power applications to avoid heating by the ac-field. On the other hand, commercial microwave
ovens are tuned to the frequency that renders the large dielectric loss in water. Therefore, food
can be microwaved and heated, whereas its container remains cold.
µd = α Elocal (10.7)
where α is polarizability (again, tensor property that in lucky cases can be reduced to a scalar).
This definition is quite tricky because it involves local electric field near the atom/molecule,
which is generally different from the applied electric field. We will deal with this problem
shortly. For now let’s notice that polarizability has rather non-intuitive SI units of C m2 /V or
F m2 . It becomes much more palatable in electrostatic CGS units [ε0 = 1/(4π)] where α is
measured in cm3 and corresponds to an effective volume of an atom/molecule.
One is the external field Eext , the second one is the depolarizing field Edepol arising from charges
that accumulate on the surface, and the third one is the Lorentz field EL due to charges in the
interior of the dielectric. The first two terms, E = Eex + Edepol , build up the electric field used
in Maxwell’s equations.
One good thing about Edepol and EL is that they are proportional to polarization, because
both fields arise from charges inside the dielectric and on its surface. The depolarizing field
depends on the sample shape,
P
Edepol = −f , (10.9)
ε0
where f is the depolarization factor, basically a form factor. Its calculation is a rather tedious
exercise in electrostatics that, reportedly, Landau gave to his students as one problem of the
“theory minimum”. Less ambitious students should be alright with the simple knowledge that
flat sample features f = 1 because a lot of charge is accumulated on its surfaces, and it acts as
a capacitor of some sort. By contrast, needle-like sample features f = 0. Spherical sample is
somewhat in between with its f = 13 , as we will see shortly.
Lorentz field arises from charges in the interior of the dielectric. Envisage a small spherical
hole that has been cut out inside the sample. The charge accumulated on this new surface
produces electric field in the center of the sphere and acts opposite to the depolarizing field.
We introduce spherical coordinates with the polar angle θ that gauges the angle with respect
to E. We then slice the sphere perpendicular to the direction of E, such that each slice is a ring
with the radius R sin θ and thickness R dθ (Fig. 10.1, right). The polarization is maximum in
45
++++++++++++++ ++++++++++++++ +++++
dq
- - - - - -- - - - - - - - - -
---- Rsinθ -
--- -
- - - - - - - - - - ----- - -- --
R dθ
Eext - -
- -
-
-
- -
- θ R
δEL
EL
Edepol + +
+ +
++ +
++
-------------- ----------
Figure 10.1: Left: local electric field is obtained as a superposition of the external field Eext , depolar-
izing field Edepol produced by charges on the surface, and Lorentz field EL produced by charges in the
interior of the dielectric. Right: calculation of the Lorentz field by slicing the sphere and evaluating
the electric field δEL created by each slice in its center.
the zenith and minimum at θ = π/2, hence for a given slice P (θ) = −P cos θ, and the overall
charge on the slice is polarization times area, namely,
δq = −P cos θ × 2πR sin θ × Rdθ = −2πR2 P sin θ cos θ dθ. (10.10)
46
ε
ε’
E(t) εst
8
t ω
1/τ
Figure 10.2: Debye model of relaxation. Left: an abrupt change in the electric field E gives rise to an
exponential change of the polarization with the single relaxation time τ . Right: real and imaginary
parts of the permittivity as a function of frequency.
the Clausius-Mossotti relation, also known as the Lorentz-Lorenz equation. Simple at first
glance, it has been very significant historically, because it offered the very first experimental
tool for probing molecular property (polarizability) using a macroscopic measurement (permit-
tivity).
ε0 χ0 E0 P χ0
P0 = ⇒ χe = = , (10.17)
1 − iωτ ε0 E 1 − iωτ
the Debye relaxation. It is a standard form of a response function for a system with the
single relaxation time. We will see it again when we come to discuss ac-conductivity of metals
(Ch. ??). For now, let’s get more familiar with the expressions. Using χ = χ′ + iχ′′ , one finds
χ0 χ0 ωτ
χ′e (ω) = , χ′′e (ω) = , (10.18)
1 + ω2τ 2 1 + ω2τ 2
where χ0 is static susceptibility.
We can also use Eq. (10.2) to calculate permittivity,
εst − 1 (εst − 1) ωτ
ε′ (ω) = 1 + , ε′′ (ω) = , (10.19)
1 + ω2τ 2 1 + ω2τ 2
with the static permittivity εst = 1 + χ0 . The imaginary component ε′′ (ω) has a peak-like
structure with the maximum at ωτ = 1, i.e., at a characteristic relaxation frequency determined
by the relaxation time τ (Fig. 10.2). It is the target frequency if we want to maximize dielectric
47
ε’’
Davidson-Cole
Debye Figure 10.3: Dielectric relaxation rep-
resented by the Cole-Cole plot. De-
bye relaxation (single value of τ ) man-
ifests itself by a semi-circle. The Cole-
Cole and Davidson-Cole models de-
Cole-Cole scribe dielectrics with multiple relax-
ation times.
ε’
ε εst
8
loss and heat the sample by microwave radiation.20 The real part ε′ equals static permittivity
at ω = 0 and decays to 1 at ω → ∞.
Debye relaxation occurs at frequencies determined by the value of τ . Liquids feature re-
laxation times on the order of picoseconds and the ε′′ (ω) peak in the GHz range. In solids,
dipoles are more constrained, so their relaxation time increases to milliseconds, whereas the
peak shifts into the kHz range. All of these frequencies are anyway much lower than typical
phonon frequencies, which lie in the THz range. In fact, Debye relaxation is never the only pro-
cess induced by the oscillating electric field, and ε′ (ω) never reaches 1 at high frequencies (see
also Ch. 12). This aspect is taken into account by the modified expression where we introduced
the high-frequency limit ε∞ ,
εst − ε∞
ε(ω) = ε∞ + . (10.20)
1 + iωτ
Here, “high frequency” means the frequency well above the frequency ω = 1/τ of the Debye
relaxation.
Needless to say, α and β can be used at the same time, thus leaving even more flexibility. The
Cole-Cole equation renders the ε′′ vs. ε′ graph non-circular, while keeping the symmetry with
respect to the mid-point at (εst − ε∞ )/2. In contrast, the Davidson-Cole relaxation leads to a
semi-circle stretched on one side and suppressed on the other (Fig. 10.3). The Cole-Cole plots
20
Commercial microwave ovens operate at 2.4 GHz, which is somewhat away from the peak of ε′′ (ω) for water,
in order to avoid overheating.
48
are widely used for analyzing relaxation behavior based on frequency-dependent measurements,
not only in dielectrics but also for example in spin glass.
49
11. Phonons and sound
11.1. Underlying approximations
Atomic motion determines many, if not all, properties of a crystal. Molecules show characteristic
vibrations (normal modes) that occur at special frequencies. Crystals feature collective atomic
vibrations too, but their nature is somewhat more involved than in the case of molecules. To
understand these vibrations in the simplest possible way, we will take advantage of several
approximations:
• classical, namely, we will treat atomic motion using classical mechanics and obtain dis-
placement waves that describe collective atomic motion in the crystal. Quantum mechan-
ics requires this motion to be quantized. The corresponding quantum is called phonon,
but we usually do not need to solve the respective problem on the quantum level. It is
sufficient to quantize the waves obtained classically, as we will consciously do in Ch. ??.
• harmonic, namely, elastic energy is quadratic in the displacement, or in other words
any displacement creates a restoring force, which is proportional to the displacement.
Harmonic approximation works well most of the time, although in Ch. ?? we will see the
need to go beyond it.
• adiabatic, namely, atomic (nuclear) and electronic degrees of freedom are separated from
each other. This is possible because nuclei are much heavier than electrons, so electrons
move a lot faster. One can then think of electrons as creating a potential energy landscape
for the nuclear motion. The typical time scale of electronic and nuclear motion is 10−15 s
and 10−12 s, respectively. This approximation is also known as the Born-Oppenheimer
approximation, especially in the context of molecules. It goes back to the very general
adiabatic theorem of quantum mechanics. The adiabatic approximation fails when two
different electronic states have similar energies, and nuclei no longer know in which po-
tential energy landscape to move. Such cases are fairly rare in the ground state,21 yet
they become more common when excited electronic states are considered, for example,
when crystal is hit by a laser that drives an electronic excitation of some sort.
d2 u p ∂E elastic
m 2
=− = −k (2up − up+1 − up−1 ). (11.2)
dt ∂up
where m is the atomic mass. Instead of trying a brute force solution of these coupled differential
equations, we will choose an ansatz,
50
ω
ave
p -1 p p +1
w
s!c
k
ela
a up-1 up
q = π/a
q
-π/a π/a
Figure 11.1: Atomic displacements up in the monoatomic chain. The right panel shows the dispersion
relations for the displacement wave (orange) and elastic wave (blue). The light-blue arrows indicate
the atomic displacements at q = π/a.
that essentially describes oscillations with the frequency ω that propagate along the chain with
the propagation vector q (we consider the 1D situation, so the scalar suffices). The first part,
eiqpa , is basically a phase shift between the oscillations of different atoms.
Using Eq. (11.3) in Eq. (11.2), one finds −mω 2 = −k (2 − eiqa − e−iqa ) and
r
k k qa
ω 2 (q) = 2 [1 − cos(qa)] ⇒ ω(q) = 2 × sin (11.4)
m m 2
where we used the absolute value because frequency must be positive. This is a dispersion
relation for the displacement wave (phonon) in a monoatomic chain. One immediately notices
that ω(q) is symmetric with respect to q = 0 and shows periodicity of 2π/a (Fig. 11.1) that
intriguingly matches periodicity of the reciprocal lattice (Ch. 5.2). Full implications of this
fact will become clear in Ch. 13. For now we will stay away from it and analyze the dispersion
relation. By definition, the value of q shows the phase shift between the displacements of
neighboring atoms. When this phase shift is small, springs stretch very little, and the wave
energy (hence frequency) is low. It increases with increasing q. At q = π/a, the phase shift
becomes π, so adjacent atoms move opposite to each other, thus forming a standing wave
(Fig. 11.1).
At low q, one can use sin(qa/2) ≃ qa/2 and write the linear dispersion relation
r
k
ω(q) = a ×q (11.5)
m
that corresponds to the group velocity of
r
dω k
vph = =a (11.6)
dq m
This is the speed of sound, as we will see shortly.
∂ 2 ux X
m = F = F (x + dx) − F (x) (11.7)
∂t2
51
ux
x x + dx shear
where forces act on both sides of the volume element. By introducing volume density ρ = m/V ,
one can write
∂ 2 ux ∂ 2 ux 1 ∂F ∂ςxx
(ρA dx) 2
= F (x + dx) − F (x) ⇒ ρ 2
= = (11.8)
∂t ∂t A ∂x ∂x
Hooke’s law relates stress and strain, ςxx = Y ϵxx via the Young’s modulus Y , whereas strain
can be represented as ϵxx = ∂ux /∂x, so overall
∂ 2 ux Y ∂ 2 ux
= × . (11.9)
∂t2 ρ ∂x2
This is the standard wave equation solved by our earlier ansatz, Eq. (11.3), but with the
dispersion relation s
Y
ω 2 = vs2 q 2 where vs = . (11.10)
ρ
The speed vs resembles our earlier result, Eq. (11.6) obtained in the q → 0 limit, and serves as
the speed of sound.
An important difference between the compressional elastic wave in a continuous medium, as
we considered here, and the displacement wave in a crystal (Ch. 11.2) is that linear dispersion
relation, ω = vs q, breaks down in the latter at higher q’s (Fig. 11.1). This happens because
crystal is discrete, so it can’t be treated as a continuous medium when period of the displacement
wave becomes comparable to the interatomic distance or, in other words, q approaches π/a.
However, the continuum approximation is perfectly justified at low q’s where period of the
displacement wave is large compared to the interatomic distance. Continuum approximations
are widely used in solid-state physics for describing mesoscopic phenomena, for example long-
period spin textures and magnetic skyrmions.
We can also envisage a wave of displacements uy , which are perpendicular to the bar
(Fig. 11.2). The solution is similar to the previous case and involves the shear stress, τxy = F/A,
as well as the shear strain, ϵxy = ∂uy /∂x, related by τxy = G ϵxy with the shear modulus G.
The resulting wave equation,
∂ 2 uy G ∂ 2 uy
= × , (11.11)
∂t2 ρ ∂x2
p
describes a shear wave propagating with the velocity of G/ρ.
Compressional wave propagates in any elastic medium, whereas shear waves only propagate
in solids because gases and liquids feature G = 0 (their shape can be changed at no energy
cost). This is the main reason why sound is a compressional wave. Shear wave can be thought
as sound too, but it’s not audible because our hearing mechanism involves sound transmission
through air.
52
11.4. Longitudinal and transverse phonons
The phonon described in Ch. 11.2 is called acoustic because q → 0 part of its dispersion
relation describes sound propagation in crystals. A similar acoustic phonon with displacements
perpendicular to the propagation direction exists too. Different names can be used to describe
these two types of motion depending on the context,
acoustic phonon longitudinal transverse
elastic wave compressional shear
seismic wave p-wave s-wave
Speeds of the corresponding waves are given by
s s
B + 43 G G
vLA = , vTA = (11.12)
ρ ρ
which are, for example, velocities of the primary (p) and secondary (s) seismic waves. Since
vLA > vTA , p-wave always arrives earlier than s-wave. This time delay measured by a seismome-
ter gauges the distance to the epicentre of an earthquake. Likewise, the data on the propagation
of p-waves and s-waves provides information on the internal structure of Earth. The existence
of the liquid outer core and inner solid core is inferred from the refraction of p-waves and from
the fact that s-waves do not reach points on the opposite side of Earth, because shear waves
do not propagate through liquid.
In Eq. (11.12), vTA is equal to the velocity of the shear wave as determined in Ch. 11.3.
However, vLA is different from the velocity of the compressional wave in a thin bar, because
Poisson’s ratio should be taken into account in a 3D solid.
In liquid and gas, G = 0 and speed of sound is described by the Newton-Laplace equation,
s
B
v= (11.13)
ρ
d2 up
m1 = −k (up − vp ) − k (up − vp−1 ),
dt2
d2 vp
m2 2 = −k (vp − up ) − k (vp − up+1 ).
dt
Our ansatz for up and vp returns a system of two linear equations for the amplitudes A1 and
A2 , (
−m1 ω 2 A1 = −k A1 + k A2 − k A1 + k A2 e−iqa
−m2 ω 2 A2 = −k A2 + k A1 − k A2 + k A1 eiqa
53
ω
k k k k
ω+
m₁ m₂ op!cal
up vp
acous!c
ω-
a
q
-π/a π/a
Figure 11.3: Left: atomic displacements in the diatomic chain comprising atoms with the masses m1
and m2 connected by the springs with the same stiffness k . Right: dispersion relation for the two
phonon modes, acoustic and optical.
or (
(m1 ω 2 − 2k )A1 + k (1 + e−iqa )A2 = 0
(11.15)
k (1 + eiqa )A1 + (m2 ω 2 − 2k )A2 = 0
The solution for A1 and A2 exists when the determinant is zero, namely,
m1 ω 2 − 2k k (1 + e−iqa )
=0 (11.16)
k (1 + eiqa ) m2 ω 2 − 2k
This condition returns a quadratic equation for ω 2 ,
m1 m2 ω 4 − 2k (m1 + m2 ) ω 2 + 2k 2 [1 − cos(qa)] = 0, (11.17)
and r
k 1 4 qa
ω 2 (q) = ± k − sin2 (11.18)
m m 2 m1 m2 2
where we introduced the reduced mass,
1 1 1
= + (11.19)
m m1 m2
54
12. Phonons and light
Whereas acoustic phonons are responsible for propagation of sound, optical phonons interact
with light. To understand this interaction, we will first review the meaning of optical constants
and then derive optical reflectivity of an ionic crystal due to phonons.
ε ω 2 = c2 k 2 with ε = n2 (12.1)
where ε is permittivity, k is the length of the propagation vector, and c = 3 × 108 m/s is
speed of light in vacuum. Since ε is generally a complex number, n is a complex number too,
n = n′ + in′′ .22 Using k = (n/c) ω = n k0 , one writes the oscillating electric field as
′′ k ′
E = E0 ei(kr−ωt) = E0 ei(nk0 r−ωt) = E0 e−n 0r
ei(n k0 r−ωt) (12.2)
where k0 is the propagation vector in vacuum (ε = 1). It is then clear that n′ , real part of the
refractive index, describes refraction, namely, the change in speed of light inside the material.
In contrast, the imaginary part n′′ describes absorption, the attenuation of the wave amplitude
inside the material.
x incident (Ei )
Figure 12.1: Propagation of light at
the crystal surface. The incident beam
splits into the reflected and transmit-
ted beams. Their intensities are deter-
transmitted (Et ) mined by the complex refractive index
of the crystal, n = n′ + in′′ .
reflected (Er )
Absorption measurements require a suitably chosen sample thickness, such that the fraction
of the absorbed intensity is neither too large nor too small. It is often more convenient to
measure reflectivity, which is defined as the intensity ratio of the incident and reflected light.
In the case of normal incidence, all of the incident, reflected, and transmitted waves travel along
the z direction perpendicular to the surface (Fig. 12.1),
At the interface, which we choose as z = 0, electric field should change continuously, so the
amplitudes fulfill the condition Et0 = Ei0 + Er0 . A similar condition should hold for the
magnetic field too, ωB = k × E according to Eq. (A.8). Therefore,
55
It is then easy to eliminate Et0 and calculate reflectivity,23
|Er0 |2 |1 − n|2
R= = . (12.4)
|Ei0 |2 |1 + n|2
In Ch. 12.4, we will see that R values close to 1 are obtained when n′′ is large, whereas
′
n vanishes. Then any light transmitted through the interface is immediately absorbed, so the
sample is non-transparent: no light can go through. In a more general situation, part of the
transmitted amplitude, Et0 , is then lost due to absorption (A), whereas transmittance of the
whole sample is T = 1 − R − A.
d2 vp
m2 dt2
= −2k (vp − up ) − qe Elocal
The phonon is considered at q → 0, so all atoms of a given type undergo the same displacement,
and we end up with only one differential equation for l = up − vp , the dipole length. Indeed,
dividing the first equation by m1 and the second equation by m2 returns the single equation
for l,
d2 l d2 l
1 1 1 1
= − k + l + + q e Elocal ⇒ m + m ω02 l = qe Elocal . (12.5)
dt2 m1 m2 m1 m2 dt2
p
This is a standard equation for a driven harmonic oscillator with ω0 = 2k /m , the frequency
of the optical phonon at q = 0 in the absence of any electric charges (as in Ch. 12).
Electric field is local in the sense of Ch. 10.4. It includes the internal field due to light, as well
as Lorenz and depolarization fields, which are proportional to polarization. The polarization
oscillates too, following oscillations of the dipole moment µd = qe l. Therefore, it makes sense
to choose the oscillating form of Elocal = E0 e−iωt and search for the oscillating solution, l(t) =
A e−iωt . The result is,
qe /m qe2 /m
A(ω) = E0 ⇒ µ d (t) = q e l(t) = Elocal (t), (12.6)
ω02 − ω 2 ω02 − ω 2
and polarizability of the crystal due to an optical phonon becomes
qe2 /m
α(ω) = 2 . (12.7)
ω0 − ω 2
23
This expression is derived for the case of normal incidence. More complex Fresnel equations are required
for an arbitrary incidence angle. In particular, polarization of light changes upon the non-90◦ reflection, which
is the cornerstone of the optical method called ellipsometry.
56
Electric field drives oscillations and creates resonance when its frequency matches the frequency
of the system, which is the frequency of the optical phonon at q → 0.
12.3. LO vs. TO
We should now use this polarizability α to calculate optical parameters from Ch. 12.1. Before
doing that, let’s compare electric polarizations created by the TO and LO phonons. These
two types of phonons differ by the direction of their displacements relative to the propagation
direction q. The displacements take the form u = u0 ei(qr−ωt) where q is the chain direction,
so we expect P = P0 ei(qr−ωt) with P ∥ q (LO) and P ⊥ q (TO).
In the absence of an external electric field, Eq. (10.1) becomes
D = ε0 Edepol + P, (12.8)
and we expect the same wave-like form of D = D0 ei(qr−ωt) and Edelop = E0 ei(qr−ωt) because
both of them arise from the same atomic displacements. Additionally, one of Maxwell’s equa-
tions requires divD = 0 ⇒ D · q = 0. For the LO phonon, this condition holds with D = 0
only, hence Edepol = −P/ε0 corresponding to the depolarization factor f = 1. On the other
hand, the TO phonon entails D ⊥ q for an arbitrary D and satisfies the above condition, but an-
other Maxwell’s equation, rot E = 0, requires q × E = 0, which is possible with Edepol = 0 only
because Edepol ⊥ q in this case. Therefore, the TO phonon corresponds to the depolarization
factor f = 0. Altogether,
P P 2P P
LO : Elocal = − =− , TO : Elocal = (12.9)
3 ε0 ε0 3 ε0 3 ε0
where P (t) is electric polarization created by the fluctuating dipoles µd (t).
Using these expressions in the definition of the electric polarization, P = nd µd = nd αElocal ,
one finds simple relations for the LO and TO phonon frequencies
nd qe2 −2P 2 nd qe2
P = 2 2
⇒ ωTO = ω02 − (12.10)
3m ε0 ω0 − ωLO 3m ε0
nd qe2 P 2 2nd qe2
P = 2 2
⇒ ωLO = ω02 + (12.11)
3m ε0 ω0 − ωTO 3m ε0
We thus conclude that ωLO ≥ ωTO . The difference between these frequencies gauges ionicity
p of
the crystal. We basically see that atomic charges modify the phonon frequency ω0 = 2k /m
that has been obtained for neutral atoms. The LO and TO phonons shift charges in different
ways and create different local polarizations. The LO phonon separates the charges and creates
internal electric fields that counteract the atomic displacements. Therefore, higher energy is
required for the LO-type atomic motion.
The oversimplified nature of this model is not to be overlooked, though. We assumed the
same stiffness k for the longitudinal and transverse atomic motion, which is of course far from
realistic in the light of the differences between the compressional and shear waves (Ch. 11.4).
Nevertheless, we will see that even this primitive model does lead to a qualitatively correct
physical picture, which justifies the exaggerated approximations involved.
57
ε
ω R
1
εst
ε ωLO
8
ω ωTO
ωTO ωLO
k 0 ω
ωTO ωLO
Figure 12.2: Interaction of optical phonon with light. Left: frequency dependence of the permittivity.
Middle: polariton dispersion caused by the interaction of light with the phonon. Right: frequency
dependence of the reflectivity with the Reststrahlen band at ωTO < ω < ωLO .
The conjectured frequency ω0 disappears, and only the tangible frequencies, ωTO and ωLO ,
remain. Permittivity diverges at ω = ωTO and shows a zero crossing at ω = ωLO (Fig. 12.2,
left).
This form of the permittivity modifies an electromagnetic wave at frequencies near ωTO
and ωLO . Indeed, by using Eq. (12.12) in the dispersion relation, Eq. (12.1), one finds that the
standard linear dispersion ω ∼ k does not hold in the vicinity of the phonon frequency, and no
real solution exists for ω(k) at ωTO ≤ ω < ωLO . It means that photons can not propagate in the
crystal within this frequency range. The characteristic dispersion shown in Fig. 12.2 (middle)
is often ascribed to a polariton, an electromagnetic wave strongly coupled to an intrinsic and
dipole-carrying excitation of the crystal.
The polariton formation causes the crystal to reflect electromagnetic waves. Indeed, at
ω < ωTO and at ω > ωLO , one finds ε > 0, such that n = n′ + in′′ is a real number with
n′ ̸= 0 and n′′ = 0. Crystal refracts light, as we know from NaCl, quartz, and many other
ionic crystals. On the other hand, at ωTO < ω < ωLO , negative ε renders n a purely imaginary
number with n′ = 0 and n′′ ̸= 0.24 The resulting reflectivity is
|1 − in′′ |2
R= = 1, (12.13)
|1 + in′′ |2
so ionic crystals reflect light in the frequency range between ωTO and ωLO (Fig. 12.2, right).
These so-called Reststrahlen bands are observed, for example, in ionic crystals in the infrared
range (ω = 0.3 − 400 THz) typical for optical phonons. A similar mechanism underlies the
response of polar molecules that absorb infrared radiation at frequencies of their vibrational
modes.
58
write permittivity as25
2
ωLO − ω2
ε(ω) = ε∞ 2
, (12.14)
ωTO − ω2
resulting in
εst ω2
= LO
2
, (12.15)
ε∞ ωTO
the Lyddane-Sachs-Teller relation.
The former condition requires D = 0 or k ⊥ D (hence k ⊥ E), while the latter requires E = 0
or k ∥ E (hence k ∥ D).
Consider the TO phonon. In this case, D ⊥ k, so E must be zero for the second condition
to be fulfilled. Then, ε → ∞ because D = εε0 E, so ε(ω) diverges at ωTO . On the other hand,
the LO phonon implies E∥k, so D must be zero, which is only possible at ε = 0. Therefore,
a zero crossing of ε(ω) should occur at ωLO . The pre-condition for these arguments is our
assumption that D ∥ E or, basically, the diagonal form of the permittivity tensor expected in
crystals with the sufficiently high symmetry (Ch. 2.5).
25
This equation can be obtained rigorously by augmenting the ionic polarizability in Eq. (12.7) with the
atomic polarizability α0 as the origin of ε∞ ̸= 1. Unfortunately, this small amendment leads to a far more
tedious algebra than what we did here, so this exercise is recommended for nerds only.
59
The situation becomes more complex in crystals with lower symmetry. Indeed, even dis-
tinguishing between longitudinal and transverse phonons is not possible in this case, because
atomic displacements are not constrained to be parallel or perpendicular to the propagation
vector of the phonon (see also Ch. 13). However, what matters for our analysis is not q but
k, the propagation direction of light. One can think of splitting atomic displacements of an
arbitrary phonon mode into two components. The component perpendicular to k leads to a
divergence of ε(ω) at ω = ωTO , whereas the component parallel to k results in the zero crossing
at some effective frequency ω = ωLO .
It may seem that almost every phonon mode of a non-cubic crystal should show up in ε(ω)
and in the optical (infrared) spectrum. Not quite. No coupling to light occurs when phonon
mode does not generate any dipole moment. This may happen in a covalent crystal or when the
phonon preserves inversion symmetry.26 What ultimately matters is the formation of a dipole
moment (i.e., atomic displacements) parallel to E of light. The coupling between this dipole
moment and oscillating electric field renders phonon-mode IR-active. Polarized light is often
used to single out modes with a given direction of the atomic displacements.
26
Such modes are called even (gerade), in contrast to odd (ungerade) modes that break the inversion symmetry.
For a given crystal, the distribution of modes into odd and even can be analyzed on the basis of crystal symmetry
and occupied Wyckoff positions, for example using the SAM utility at the Bilbao server.
60
13. Phonons and the reciprocal lattice
13.1. Brillouin zone
13.2. Dynamical matrix
...
61
A. Electrodynamics
A.1. Maxwell’s equations
For a non-magnetic solid one writes the four Maxwell’s equations in the form
div D = ρe , (A.1)
div B = 0, (A.2)
∂B
rot E + = 0, (A.3)
∂t
∂D
rot H − = j, (A.4)
∂t
where ρe is the charge density and j is electric current density. The displacement field D = εε0 E
is defined via the permittivity ε, while the two magnetic fields are related by B = µ0 H.
One can also derive the continuity equation by taking divergence of Eq. (A.4) and using the
vector identity div (rot H) = 0,
∂ ∂ρe
div (rot H) − div D = div j ⇒ div j + = 0. (A.5)
∂t ∂t
It shows that any change in the charge density is caused by the flux of the electric current, and
vice versa.
The above equations are written in the SI units where ε0 = 8.85 × 1012 C/(V m) and
µ0 = 4π × 10−7 V s/(A m) are the vacuum permittivity and permeability, respectively. It is not
advisable to use CGS units unless you really have to, because Maxwell’s equations will contain
additional pre-factors in this case.
where k is the propagation vector and amplitudes E0 , H0 can be complex to allow for a phase
shift between E and B. By replacing Eqs. (A.6) into the Maxwell’s equations, one finds
εε0 (k · E) = 0 µ0 (k · H) = 0 (A.7)
k × E = ωB k × H + εε0 ω E = j/i. (A.8)
The current j can be eliminated using Ohm’s law, j = σE, which transforms Eq. (A.8) into
The most natural solution to Eqs. (A.7)–(A.8) is the transverse wave with k ⊥ E and k ⊥ B,
thus satisfying Eqs. (A.7). From the remaining two equations one concludes that E ⊥ B. This
way, E and B oscillate along two different (orthogonal) directions, whereas the wave propagates
along the third one. It is the standard electromagnetic wave (light) as we know it.
To explore the behavior of such transverse waves, we use the vector relation
62
(k · E = 0 since the two vectors are orthogonal) and apply it to Eqs. (A.8). Then we obtain
c2 k 2 = ε̂(ω) ω 2 , (A.11)
√
with c = 1/ ε0 µ0 = 3 × 108 m/s, speed of light in vacuum, and complex permittivity
iσ
ε(ω) = ε + . (A.12)
ε0 ω
It is what we call permittivity throughout these lecture notes. The practical advantage of
the complex permittivity is the simple and convenient linear dispersion relation, k ∼ ω. The
conceptual physical reason for merging permittivity and conductivity into a single parameter
is that both of them describe system’s response to the electric field, but charge displacements
(ε) have a phase shift of π/2 with respect to charge velocities (σ). Ch. ?? offers some further
insights into this issue.
63
B. Thermodynamics
64