0% found this document useful (0 votes)
189 views213 pages

PHY307 Solid State

This document provides information on the course PHY307 Solid State Physics offered by the National Open University of Nigeria. The course is divided into 4 modules covering various topics in solid state physics across 15 study units. Assessment of the course consists of tutor-marked assignments accounting for 40% of the grade and a final exam accounting for 60%. The course aims to provide students with an understanding of solid state physics and its applications.

Uploaded by

Faderera Aikoye
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
189 views213 pages

PHY307 Solid State

This document provides information on the course PHY307 Solid State Physics offered by the National Open University of Nigeria. The course is divided into 4 modules covering various topics in solid state physics across 15 study units. Assessment of the course consists of tutor-marked assignments accounting for 40% of the grade and a final exam accounting for 60%. The course aims to provide students with an understanding of solid state physics and its applications.

Uploaded by

Faderera Aikoye
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 213

NATIONAL OPEN UNIVERSITY OF NIGERIA

SCHOOL OF SCIENCE AND TECHNOLOGY

COURSE CODE: PHY307

COURSE TITLE: Solid State Physics


Course Code: PHY307
Course Title Solid State Physics
Course Writer Dr Nasir Khalid Abdullahi
Department of Applied Sciences
College of Science and
Technology
Kaduna Polytechnic
P.M.B.2021, Kaduna
Nigeria

Course Editor

Programme Leader

Course Coordinator

2
CONTENTS
PAGE
Introduction - - - - - - - - - -
1
What you will Learn in this course - - - - - -
1
Course Aims - - - - - - - - -
1
Course Objectives - - - - - - - - -
1
The Course Materials - - - - - - - -
2
The study units - - - - - - - - -
2
Presentation Schedule - - - - - - - -
5
Assessment - - - - - - - - - -
6
Tutor Marked Assignment - - - - - - -
6
Final Examination and Grading - - - - - - -
6
Course Marking Scheme - - - - - - - -
6
Facilitators/Tutors and Tutorials - - - - - -
6

3
Introduction
Solid state physics is a very wide field, and it includes many branches. It is
concerned with the physical properties of solids, particularly the special properties
exhibited by atoms and molecules because of their association in the solid phase. The
existence of powerful theoretical methods and concepts applicable to a wide range of
problems has been an important unifying influence in the field

Learning solid-state physics requires a certain degree of maturity, since it involves tying
together diverse concepts from many areas of physics. The objective is to understand, in a
basic way, how solid materials behave. To do this one needs both a good physical and
mathematical background. One definition of solid-state physics is that it is the study of
the physical (e.g. the electrical, dielectric, magnetic, elastic, and thermal) properties of
solids in terms of basic physical laws. In one sense, solid-state physics is more like
chemistry than some other branches of physics because it focuses on common properties
of large classes of materials. It is typical that solid-state physics emphasizes how physical
properties link to the electronic structure. The rapid rise of interest in solid state physics
in recent years has suddenly presented universities with the problem of offering adequate
instruction in the subject. For this reason, there should be an introductory or survey
course followed by, as a minimum program for graduate students intending to do research
in the field, a course in x-ray crystallography and a course in the quantum theory of
solids. These two subjects are large, important, and well-developed; it is not possible to
deal with them adequately in an introductory course.

What You will Learn in this Course

The course consists of units and a course guide. This course guide tells you briefly what
the course is about, what course materials you will be using and how you can work your' with
these, materials. In addition, it advocates some general guidelines for the amount of time you are
likely to spend on each unit of the course in order to complete it successfully.

It gives you guidance in respect of your Tutor-Marked Assignment which will be


made available in the assignment file. There will be regular tutorial classes that are related
to the course. It is advisable for, you to attend these tutorial sessions. The course will prepare you
for the challenges you will meet in the field of solid state physics.

Course Aim
The course aim to provide an understanding of Solid state physics
Course Objectives

To achieve the aim set out, the course has a set of objectives. Each unit has specific
objectives which are included at the beginning of the unit. You should read these
objectives before you study the unit. Below are the comprehensive objectives of the
course as a whole. By meeting these objectives, you should have achieved the aim
of the course as a whole. After going through the course, you should be able to:

 Explain crystal structure of solids


4
 Explain Crystal binding
 Explain X-ray diffraction in crystals
 Explain Thermal properties of the crystal lattice
 Explain Elastic properties of crystals
 Explain Lattice vibration
 Explain the concept of Free-electron theory of metals
 Understand Energy bands in crystals
 Understand Semiconductors
 Understand Superconductors

The course Materials

The main components of the course are:

1. The course Guide


2. Study units
3. References/Further Readings
4. Assignments
5. Presentation Schedule

Study Unit

The study units in this course are as follows:

Module 1 Property of Crystal

Unit 1 Crystal geometry

Unit 2 Crystal classification

Unit 3 Simple lattices

Unit 4 Crystal Diffraction (I)

Unit 5 Crystal diffraction (ii)

Unit 6 Experimental crystal structure determination

Module 2 Crystal Elastic Constants and Vibrations

Unit 1 Elastic Constants of Crystals (I)

Unit 2 Elastic Constants of Crystals (II)

Unit 3 Crystals Binding


5
Unit 4 Lattice Vibration

Unit 5 Thermal Properties


Module 3 Free Electron Fermi Gas
Unit 1 Free Electron Theory of Metals
Unit 2 Electronic Transfer
Unit 3 Energy Band Theory
Unit 4 Electron Dynamics
Unit 5 Fermi Surfaces
Module 4 Semiconductors and Superconductors
Unit 1 Structure and Bonding in semiconductors
Unit 2 Semiconductor Statistics
Unit 3 Electrical Conductivity and Real Semiconductors
Unit 4 Super Conductivity (I): The Basic Phenomenon
Unit 5 Superconductivity (II): Experiments and Theories
Module 1 which consists of six units deal with crystal structures and their determination.
Module 2 (five units) is devoted to fundamental determination of elastic constants of
crystal. The free electron which discusses the physical basis of the formation of bands,
the most important concept in the band – Fermi surfaces were treated in five units which
constitute model 3. Module 4 in five units provides discussions on the properties of
semiconductors as well as discussions on basic phenomenon of superconductors.
Each unit consists of either one or two weeks’ work and include an introduction,
objectives, definition, conclusion, summary, tutor marked assignments (TMA) and
references. The TMA will help you in achieving the stated learning objectives of the
individual units and the course as a whole.
Presentation Schedule

Students are encouraged to complete and submit on time their TMAs and to
guard against falling behind in attending tutorials.

Assessment

There are three aspects to the assessment of the course. These are the self-
assessment exercises, tutor marked assignments and the written examination/end
of course examination. The assignments must be deal with by applying the
knowledge and techniques gathered during the course and must be submitted to
your facilitator for formal assessment in accordance with the deadlines stated in
the presentation schedule. The assessment will account for 40% of the total
course work while the examination will count for the remaining 60%.

Tutor marked Assignment (TMA)

The TMA is a continuous assessment component of the course work. It accounts


for 40% of the total score. You will be given six (6) TMAs to answer out of
which four must be answered before a student is allowed to sit for the end of the
course examination. Students are not allowed to present other people’s work as

6
their own (including copying another student's work). Make sure that each
assignment reaches your facilitator on or before the deadline given. Extension
will not be granted after the due date unless there are exceptional cases

Final Examination and Grading

The end of course examination for solid state physics will be for three (3) hours
and it has a value of 60% of the total course work. All areas of the course will be
assessed.

Course Marking Scheme

Assignment Marks

Assignment 1-6 Six assignments, best four marks at


10% each totaling 40% of the
course marks

End of course examination 60% of overall course marks

Total 100% of course materials

Facilitators/Tutors and Tutorials


There will be tutorials provided in support of this course at the end of each unit.
Students will be notified of the dates, times and location of these tutorials as well as
the name and phone number of your facilitator. Your facilitator will mark and
comment on your assignments and returned to you as soon as possible.

7
UNIT1: CRYSTAL GEOMETRY Page

1.0 Introduction - - - - - - - - 3
2.0 Objectives - - - - - - - - 3
3.0 Definition - - - - - - - - 3
3.1 Translational symmetry - - - - - 3
3.2 Lattice and Unit cell - - - - - - 4
3.3 Primitive and Non-primitive cells - - - - 5
3.4 Bravais Lattice - - - - - - 7
3.5 Basis and crystal structure - - - - - 7
4.0 Conclusion - - - - - - - - 8
5.0 Summary - - - - - - - - 8
6.0 Tutor Marked Assignment - - - - - - 8
7.0 Further Reading/References - - - - - - 9

8
1.0 Introduction
The physical definition of a solid has several ingredients. We start by defining a solid as
a large collection of atoms that attract one another so as to confine the atoms to a
definite volume of space. Additionally, in this unit, the term solid will mostly be
restricted to crystalline solids. A crystalline solid is a material whose atoms have a
regular arrangement that exhibits translational symmetry. When we say that the atoms
have a regular arrangement, what we mean is that the equilibrium positions of the atoms
have a regular arrangement. At any given temperature, the atoms may vibrate with small
amplitudes about fixed equilibrium positions. Elements form solids because for some
range of temperature and pressure, a solid has less free energy than other states of
matter. It is generally supposed that at low enough temperature and with suitable
external pressure everything becomes a solid. The study of crystal and electrons in
crystal is a division of physics known as solid state physics. The solid state physics is an
extension of atomic physics following the discovery of X-ray diffractions of crystalline
properties.

2.0 Objectives
The candidates should be able to:
• Define crystals
• Explain the crystal structure
• Classify crystals

3.0 Definition of crystal


Crystal may defined on the macroscopic scale as homogeneous solids, in which some of
the physical properties are function of direction. Microspically, a crystal may be defined
as a solid having an arrangement of atoms (or molecules) in which the atoms are
arranged in some repetitive pattern in three dimensions.

3.1 Translational Symmetry


A solid is said to be a crystal if atoms are arranged in such a way that their positions are
exactly periodic. This concept is illustrated in Fig.1.1 using a two-dimensional (2D)
structure. A perfect crystal maintains this periodicity in both the x and y directions from
- to + . As follows from this periodicity, the atoms A, B, C, etc. are equivalent. In
other words, for an observer located at any of these atomic sites, the crystal appears
exactly the same. The same idea can be expressed by saying that a crystal possesses a
translational symmetry. The translational symmetry means that if the crystal is translated
by any vector joining two atoms, say T in Fig.1.1, the crystal appears exactly the same
as it did before the translation. In other words the crystal remains invariant under any
such translation.

9
Fig.1.1: Periodicity and concept of symmetry.

3.2 Lattice and Unit cell


The structure of all crystals can be described in terms of a lattice. A lattice can be
defined as a regular periodic array of points in space (Fig.1.2).Every lattice point can be

    

located as;
(1.1)

     

Or in three dimensional case
(1.2)

 ,
, are called Lattice vectors and l, m and n are integers.
The network of lattice lines divide the space into identical parts called unit cells. Hence,
because of inherent periodicity of space lattice; it can thus be represented by a unit cell.
A unit cell is a conveniently chosen fundamental block by repeating the entire space
lattice which is generated. The unit cell may be in form of a parallelogram (2D) or a
parallelepiped (3D) with lattice points at their corners. The size and shape of the unit cell
are described by three lattice vectors a, b, c, originating from one corner of the unit cell.
The axial lengths a, b, c and the inter axial angles α, β and γ are lattice parameters of the
unit cell. Fig.1.3 shows the unit cell with the axes lengths and inter axial angles while
Fig.1.4 shows the lattice and unit cells in 2-dimension.

10
a

Fig.1.2: Lattice point and Lattice vectors

Fig.1.3: Unit cell showing axes lengths and inter axial angles.
The convention for drawing the lattice parameters is as follows:
a parallel to x-axis
b parallel to y-axis
c parallel to z-axis
α angle between y and z
β angle between z and x
γ angle between x and y

3.3 Primitive and Non-Primitive cells


The cell is said to be primitive if the lattice points are at the corners of the cell (Fig. 1.5)
and if there are lattice points in the cell other than the corners, the cell is said to be
nonprimitive (Fig.1.5)

11
Fig.1.4: Lattice and unit cells in 2-Dimension(After Kittel,1979)

Fig.1.5: Primitive and Non-primitive cells (After Sihv K Gupta,


www.4shared.com)

For a single atom, the single atom is placed on the lattice site and is known as Bravais
lattice. On the other hand, if there are several atoms per unit cell, we have a lattice with
a basis.

12
3.4 Bravais Lattice
There are many ways in which an actual crystal may be built, thus possible crystal
structures are unlimited. However, the possible schemes of space lattices are highly
restricted. Each space lattice has some convenient set of axes which need not be
necessarily orthogonal and chosen length along the three axes may not be equal. Bravais
in 1848 proved that there are only fourteen space lattices in total which are required to
describe all possible arrangement of points in space subject to the condition that each
lattice point has exactly identical environment. The fourteen space environments are
called Bravais Lattices. The Bravais lattices are the distinct lattice types which when
repeated can fill the whole space. The lattice can therefore be generated by three unit
vectors, a, b and c and a set of integers k, l and m so that each lattice point, identified by
a vector r, can be obtained from:

r=ka+lb+mc (1.3)

Bravais showed that in two dimensions there are five distinct Bravais lattices, while in
three dimensions there exist no more than fourteen space lattices.

3.5 Basis and Crystal structure.


The arrangement of atoms in a solid is termed crystal structure. In order to convert the
geometrical array of points in space (lattice) into a crystal structure, we must locate
atoms or molecules on the lattice points. The repeating unit assembly of atoms or
molecules that are located at each lattice point is called the basis. The basis must be
identical in composition, arrangement and orientation such that the crystal appears
exactly the same at one point as it does not at other equivalent points. No basis contains
fewer atoms than a primitive basis contains.
The crystal structure is thus given by two specifications:
I. the lattice, and
II. The assembly that repeat itself.
Hence, the logical relation is
Space lattice + basis = crystal structure (1.4)
Equation (1.4) is illustrated in Fig.1.6

O X O X
O X

a b O X O X O X
a2
a1
O X
O X
O X

Fig.1.6: Two-dimensional lattices. (a) Bravais lattice; a1 and a2 are basis vectors;
(b) Lattice with a basis of three atoms; , 0, X
(After Kachhava, 1992)

13
4.0 Conclusion
The fundamental feature of a crystal is the periodicity of the structure.

5.0 Summary
• The size and shape of the unit cell are described by three lattice vectors a, b, c,
originating from one corner of the unit cell. The axial lengths a, b, c and the inter
axial angles α, β and γ are lattice parameters of the unit cell.
• A cell is said to be primitive if the lattice points are at the corners of the cell and
if there are lattice point in the cell other than the corners, the cell is said to be
nonprimitive

 = n1a +n2b+n3c
• A lattice is any array of points related by the translational operator

• The Bravais lattices are the distinct lattice types which when repeated can fill the
whole space generated by three unit vectors, a, b and c and a set of integers k, l
and m

6.0 Tutor marked Assignment


Q1. A group is represented by three matrices

1 0    
        
0 1    

Where  = sin 30% and  =cos 30% .


(a) Determine the multiplication table for this group.
(b) Give an example of a 2-D crystal with these point group symmetries.

Q2. (a) Filled circles in the tetragonal crystal in the figure below represent copper
oxide atoms and the copper oxide layers are stacked with spacing c. assume that
there are no other atoms in the crystal, sketch the Bravais lattice and
indicate a possible set of primitive vectors for this crystal.

(b) Define the following terms


(i) Unit cell and
(ii) Basis

14
7.0 Further reading/References
Ashcroft, N.W., Mermin, D.N, Solid state physics, Saunders College Publishing,
1976
Gupta, S. K, (www.4shared.com)
Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992

15
UNIT2: CRYSTAL CLASSIFICATION page

1.0 Introduction - - - - - - - - 11
2.0 Objectives - - - - - - - - 11
3.0 Definition - - - - - - - -- 11
3.1 Fundamental types of lattice- - - - - - 11
3.2 Direction indices - - - - - - - 12
3.3 Miller indices -- - - - - - - 12
3.4 Determination of miller indices- - - - - 14
3.5 Some general principles of miller indices- - - - 15
4.0 Conclusion - - - - - - - - 16
5.0 Summary - - - - - - - -- 16
6.0 Tutor Marked Assignment - - - - - - 17
7.0 Further Reading/References - - - - - - 17

16
1.0 Introduction
Crystal lattices are classified according to their symmetry properties, such as inversion,
reflection and rotation. Also, it is sometimes more convenient to deal with non-primitive
or conventional cells, which have additional lattice sites either inside the cell or on its
surface. In three dimensions there are 14 different Bravais crystal lattices which belong
to 7 crystal systems. These systems are triclinic, monoclinic, orthorhombic, tetragonal,
cubic, hexagonal and trigonal.
2.0 Objectives
• To revise the classification of crystal lattices
• To understand direction indices
• To understand miller indices

3.0 Definition of Crystal Lattice


Crystal lattice classification is the regular geometric arrangement of points in the atom
of a crystal

3.1 Fundamental types of lattices


The most obvious feature of a crystal is its regularity or symmetry. The basis of
classification of crystal is the symmetry exhibited by them. In a well defined crystal, the
various symmetry elements (rotation, reflection, inversion etc.) intersect at a point. Each
set of symmetry elements intersecting at a point (the centre of unit cell) is called a point-
group. Since there are 32 point groups, there are equal numbers of crystal classes, which
can be grouped together into seven groups known as crystal systems. Table 1.1, consists
of the list describing the various systems. Fig 2.1 shows how seven crystal systems can
be obtained by successive distortion of a cube.
Table 1.1: Seven Crystal Systems
Crystal System Axial Lengths and Unit cell Number of
angles Lattices

Cubic a = b = c, α = β = γ = a cube 3
900

Tetragonal a = b ≠c , α = β = γ = a squared-based right 2


900 prism
Orthorhombic a ≠ b ≠ c, α = β = γ = a rectangular-based 4
900 right prism
Rhombohedra a = b = c, α = β = γ ≠ a rhombohedron 1
900

Hexagonal a = b ≠c , α = β = 900, a rhombus-based right 1


γ=1200 angles
Monoclinic a ≠ b ≠ c, α = γ = 900 ≠ A parallelepiped-based 2
β right prism

Triclinic a ≠ b ≠ c, α ≠ β ≠ γ ≠ a parallelepiped 1
900

17
3.2 Direction indices
To find the direction indices, the following rules are used:
I. Find any vector in the desired direction.
II. Express this vector in terms of the basis (a, b, c).
III. Divide the coefficient of (a, b, c) by their greatest common divisor.
The resultant set of three integers u, v, w usually included in parentheses [uvw] defines a
direction. uvw means that all vectors are equivalent to [uvw]. Negative sign in any of
the numbers are indicated by placing a bar over the number ( u ). Let a = 2, b = 3, c = 4
units and the vector be
r = 6 i +12 j +10 k
Then r = 3(2) i +4(3) j + 2.5(4) k
Thus, the coefficients of (a, b, c) are 3, 4, 2.5. The relevant greatest common divisor is
0.5. Thus, the three numbers 6, 8, 5 are found. Hence, for the example considered the
indices of direction are [685].
In the cubic system, u, v, w are proportional to the direction cosines of the chosen
vector. The cube edge a would be denoted by [100], that of direction b by [010], and c

mean a plane parallel to [100] but cutting a axis at 1)2 *. Fig 2.1 shows the indices of
by [001].The negative direction of a would be [ 100]. When we speak of [200] plane, we

some important planes and directions in crystals. Note that:


I. All parallel rows of atoms have the same [uvw].
II. The angle θ between two crystallographic direction[u1v1w1] and [u2v2w2] in a
cubic system is given by

cos + 
,- ,. /0- 0. /1- 1.
-⁄ . -⁄.
2,-. /0-. /1-. 3 2,.. /0.. /1.. 3
(2.1)

3.3 Miller indices


Miller indices are the most commonly used notation for specifying points, directions,
and planes in crystal lattice systems. Not only do they simplify the description of
locations and directions within the lattice, but they also allow vector operations like dot
and cross products. Miller Indices are a symbolic vector representation for the
orientation of an atomic plane in a crystal lattice and are defined as the reciprocals of
the fractional intercepts which the plane makes with the crystallographic axes. Before
Miller indices can be used, a coordinate system for the crystal structure must first be
selected. The right-hand Cartesian coordinate system is the usual choice for this
(Fig.2.2). Points within the coordinate system are specified by Miller indices as h, k, l,
where h, k, and l are fractions of the lattice parameters a, b, and c. Recall that a, b, and c
are the lengths of the edges of the crystal's unit cell in the x, y, and z directions.

18
Fig.2.1: seven crystals in three dimensions

19
A plane oriented with respect to the rectangular coordinate system, which intercepts the
x-,y-and z-coordinates
coordinates at distance a, b and c respectively is represented by the equation

(2.2)

Denoting the reciprocal of axial intercepts as and ,


Eq. (2.2) becomes

(2.3)

Fig.2.2:: Construction for description of a plane. This plane intercepts the a, b, c


axes at 3a, 2b, 2c.. (After Kittel, 1979)

3.4 Determination of Miller Indices


The Rules for Miller Indices are:

• Determine the intercepts of the plane along the three crystallographic


axes, in terms of unit cell dimensions. Coordinates of the points of
interception are expressed as integral multiples of the axial lengths in the
respective directions. The integers p, q and r are the multiples of axial
lengths a, b and c respectively
• Take the reciprocals of the integers p, q and r
• The reciprocals are reduced to the smallest set of integers h, k and l by
taking LCM
• The integers are written as (hkl) by enclosing in parenthesis

For example, if the x-, y-,, and z-


z intercepts are 2, 1, and 3, the Miller indices are
calculated as:

• Thee integers are 2, 1, 3


• Take reciprocals: 1/2, 1/1, 1/3
• Clear fractions (multiply by 6): 3, 6, 2
• Reduce to lowest terms (already there)

20
Thus, the Miller indices are 3, 6, 2. If a plane is parallel to an axis, its intercept is at
infinity and its Miller index is zero. A generic Miller index is denoted by (hkl). If a plane
has negative intercept, the negative number is denoted by a bar above the number. Never
alter negative numbers. For example, do not divide -1, -1, -1 by -1 1 to get 1, 1, 1. This
implies symmetry
ymmetry that the crystal may not have!

3.5 General Principles of Miller Indices

• If a Miller index is zero, the plane is parallel to that axis.


• The smaller a Miller index, the more nearly parallel the plane is to the
axis.
• The larger a Miller index, the more nearly perpendicular a plane is to that
axis.
• Multiplying or dividing a Miller index by a constant has no effect on the
orientation of the plane
• Miller indices are almost always small.

Fig.2.3 shows some planes for cub


cubic
ic lattices with their Miller notations.

Fig.2.3: Some of the prominent planes for cubic lattices with their Miller indices
(After Kachhava, 1992)
Note that:
 Miller indices are proportional to the direction cosines of the normal to the
correspondingg plane. Direction cosines are given as

 The normal to the plane with index numbers is the direction


 The purpose of taking reciprocals is to bring all the planes inside a single unit cell
 Assume repre
resent the distance between two adjacent parallel planes having
miller indices , then

21
567 
8
√6 . /7 . / .
(2.4)

Where 567 = distance between planes

a = lattice constant (edge of unit cell)


h, k,l = Miller indices of planes being considered

Figure 2.4 shows inter planer spacing in terms of the cube edge, a.

Fig.2.4: inter planer spacing (After Kachhava, 1992)

4.0 Conclusion
Miller indices are the most commonly used notation for specifying points, directions,
and planes in crystal lattice systems. Not only do they simplify the description of
locations and directions within the lattice, but they also allow vector operations like dot
and cross products.

5.0 Summary

• In a well defined crystal, the various symmetry elements (rotation, reflection,


inversion etc.) intersect at a point.
• Each set of symmetry elements intersecting at a point (the centre of unit cell) is
called a point-group.
• The Miller indices are defined as the reciprocals of the fractional intercepts which
the plane makes with the crystallographic axes.
• The angle θ between two crystallographic direction[u1v1w1] and [u2v2w2] in a
cubic system is given by
u1u 2 + v1 v 2 + w1 w2
cos θ =
(u12 + v12 + w12 )1 2 (u 22 + v22 + w22 )1 / 2
22
• The distance 567 between neighboring planes of the family:; ), is given in
terms of the cube edge a as

a
d hkl =
(h
+ k +l2 2 2
)1/ 2

6.0. Tutor marked Assignment

:; < in a cubic lattice of lattice constant a is


Q1 (a). Show that the perpendicular distance between two adjacent planes of a set

a
d hkl =
(h 2
+ k +l2
2
)
1/ 2

(b). The Bragg angle corresponding to the first order reflection from plane
(111) in a crystal is 300 when X-rays of wavelength 1.75Ǻ are used.
Calculate the interatomic spacing

Q2. If x, y and z axes intercept 3, 4, and 2, calculate the Miller indices

7.0 Further reading/References


Ashcroft, N.W., Mermin, D.N, Solid state physics, Saunders College Publishing,
1976.
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992 .
Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979.

23
UNIT3: SIMPLE LATTICES Page
1.0 Introduction - - - - - - - - 19
2.0 Objectives - - - - - - - - 19
3.0 Definition - - - - - - - - 19
3.1 Simple lattices- - - - - - - 19
3.2 Simple cubic lattice- - - - - - -- 19
3.3 Body –Centered Cubic (BCC - - - - - 20
3.4 Face-Centered Cubic (FCC) - - - - - 20
3.5 Hexagonal closed packed (HCP) - - - - - 20
3.6 Closed Packed Structure - - - - - - 21
4.0 Conclusion - - - - - - - - 23
5.0 Summary - - - - - - - - 23
6.0 Tutor Marked Assignment - - - - - - 23
7.0 Further Reading/References - - - - - - 24

24
1.0: Introduction
The most highly symmetrical lattices which occur naturally are cubic structure. These
are, therefore, of some practical interest and also provide useful simple examples which
help in visualizing the more general cases. About 90% of metallic lic crystal structures
structure
crystallize into 3 densely packed crystal structures vis-a-vis Body-Centered
Centered Cubic cell
(BCC), Face-Centered
Centered Cubic cell (FCC) and Hexagonal Close-Packed
Packed (HCP).

2.0 Objectives
nderstand metallic crystal structure such as:
The objectives are to understand
• Simple cubic
• Body centered cubic
• Face centered cubic
• Hexagonal Close packed
3.0 Definition of Simple lattices
Simple lattices are crystalline
rystalline solids that consist of a small group of atoms (unit
cells) that contains unique features.
3.1 Simple lattices
The simple lattices have the following elementary properties:
I. Effective no of atoms/ unit cell, Z,Z which defines the number of atom per
primitive cell
II. Atomic radius, R usually defines in terms of lattice constant (length of a
side of unit cell),
cell) a.
III. Nearest neighbor distance which defines the nearest distance between
atomic centers.
centers
IV. Coordinate number which defines the number of nearest neighbor of an
atom.
V. Atomic Packing Fraction (APF) defined as the fraction of volume v in a
crystal structure that is occupied by atoms.
3.2 The simple cubic lattice.
The simple cubic lattice has basis vectors
(3.1)

and the unit cell is a simple cube. The simplest crystal based on this lattice has single
atoms at the lattice points, Fig. 3.1. Each atom has six identical nearest neighbors.

Fig.3.1: Simple cubic lattice (After Kachhava, 1992)


25
3.2 Body-Centered cubic Lattice
The body-centered cubic (bcc) lattice may be regarded as two interpenetrating simple
cubic lattices with atoms at the centre of each cube as well as at the corners. The space
lattice may be taken with the basis vectors

(3.2)

Where a is the side of the cube and i, j, k are orthogonal unit vectors parallel to the cube
edges. The primitive cell of the bcc lattice has a volume one-half that of the unit cube.
By elementary vector analysis the volume is given by

V = │a1.a2 x a3│ (3.3)

3.3 Face- Centered Cubic Lattice


The face centered cubic lattice can be considered as four interpenetrating simple cubic
lattices giving a cubic unit cell with extra lattice points at the centers of the faces of the
fundamental cube. Each point has 12 nearest neighbours. The full translational
symmetry has basis vectors.

(3.4)

The primitive cell of the lattice is shown in Fig.3.4 and is a rhombohedron of


volume one quarter that of the unit cube. The translation vectors a1, a2 and a3 connect
the lattice point at the origin with the lattice points at the face centers. The angles
between the axes are 600 .
k

i
j

Fig.3.2: Face- centered cubic lattice

3.4 Hexagonal Close-Packed (HCP)


In the hexagonal closed packed structure, Fig 3.5 the unit cell is a rhombic and
the basis vectors are

26
=  *>, ?  , A  C

?2>/@AB3
(3.5)

Fig.3.3: Hexagonal Close-packed structure (After Kittel, 1979)

In this structure, there are two atoms per unit cell separated by the vector

 F*>  B  GCH
D 
E √A
(3.6)

Here, as in IGG structure, each atom has twelve neighbours, but the arrangement is
slightly different.

3.5 Closed-packed Structures


If the atoms are considered as hard spheres, then the most efficient packing in one plane
is the closed–packed arrangement shown in Fig 3.6.There are two simple ways in which
such planes can be laid on top of one another to form a three-dimensional structures.

hexagonal symmetry and is called the hexagonal closed packed :GJ< structure (Fig3.7).
One leads to the face-centered cubic (cubic close-packed) structure, while the other has

The fraction of the total volume filled by the spheres is 0.74 for both the IGG and hcp
structures.

Fig.3.4: A closed- packed layer of spheres (After Kittel, 1979)

27
Fig.3.6: The Hexagonal closed packed structures (After Kittel, 1979)

contact with six others. Such a layer can either be the basal plane of a GJ structure or
Spheres may be arranged in a single closest-packed layer by placing each sphere in

the (111) plane of IGG structure. A second similar layer is added by placing each sphere

in two ways: in the IGG structure the spheres in the third layer are placed over the holes
in contact with three spheres of the bottom layer as in Fig.3.6. A third layer can be added

in the first layer not occupied by the second layer; in the hexagonal structure the spheres

packing in the IGG structure is ABCABC. ….. , whereas in the GJ structure the packing
in the third layer are placed directly over the spheres in the first layer. We say that the

is ABABAB….. The GJ structure has a hexagonal primitive cell; the basis contains two
atoms. The IGG primitive cell contains one atom. The c/a ratio for hexagonal closest-
packing of spheres is (8/3)1/2 = 1.633. We refer to crystals as GJ even if the actual c/a

to commonly as GJ. Magnesium with c/a = 1.623 is close to ideal GJ. Many metals
ratio departs somewhat from the theoretical value. Thus zinc with c/a = 1.86 is referred

transform easily at appropriate temperatures between IGG and GJ. The coordination
number, defined as the number of nearest-neighbor atoms, is 12.
A quantitative measure of the closeness of packing in a crystal structure is provided by
the packing fraction, f, defined as

I
0L ,M LNN,OPMQ RS 8TLU:68VQ UO6MVMU<
0L ,M LW T6M ,PT NM LW T6M XYZGYZ
(3.5)

The theoretical calculations of f requires the knowledge of number of atoms, N, per unit
cell and atomic radius, Ra, in terms of a, the length of a side of a cubic lattice. Table 3.1
as reported by (Kachhava, 1992) displayed the values of N, Ra and f along with number
(Nn) of nearest neighbors and that (Nnn) for next nearest neighbors for simple cubic (sc),
body centered cubic ([GG), face-centered cubic:IGG< and hexagonal close-packed(GJ)
structures.

28
\
] ^ _
Table 3.1 Data for common structures (modified after Kachhava, 1992)

N 1 2 4 2
Nn 6 8 12 12

   
√? √A
Nnn 12 6 6 6

? ? ` ?
Ra

√? = 0.74 √A c = 0.68
a a
b
f π/6 = 0.52 0.74 (ideal)

4.0 Conclusion
The ideal crystal of classical structures is formed by the repetition of identical units in
space. The most highly symmetrical lattices which occur naturally are cubic structures
which help in visualizing the more general case.

5.0 Summary

= = a> ?  dB A = ak
• The simple cubic lattice has basis vectors

• Important simple structures are the bcc, IGG and GJ

IGG have the sequence ABCABC…


• The structures differ in the stacking sequence of the planes

GJ have the sequence ABABAB…



6.0 Tutor Marked Assignment


Q1. Use elementary vector analysis to find the value of the angle between the body
diagonals of a cube shown in the Figure Q1
Q2 Show that the c/a ratio for an ideal hexagonal closed-packed structure is

Sodium transform from bcc to GJ at about T= 23K. Assuming that the density
(8/3)1/2 = 1.633.

remain fixed, and the c/a ratio is ideal, calculate the GJ lattice spacing a given
Q3.

that the cubic lattice spacing a′ = 4.23 .What is the difference in the cubic phase

Fig.Q1
29
7.0 Further Reading/ References
Ashcroft, N.W., Mermin, D.N, Solid state physics, Saunders College Publishing,
1976.
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979.

30
UNIT 4: CRYSTAL DIFFRACTION (I) Page
1.0 Introduction - - - - - - - - 26
2.0 Objectives - - - - - - - - 26
3.0 Definition - - - - - - - - 26
3.1 Bragg formulation of diffraction by a crystal - - - 26
3.2 Von Laue formulation of diffraction by a crystal - - 28
3.3 Diffraction of crystal by electrons - - - - 30
3.4 Diffraction of crystal by neutrons - - - - 31
4.0 Conclusion - - - - - - - - 31
5.0 Summary - - - - - - - - 31
6.0 Tutor Marked Assignment - - - - - - 32
7.0 Further Reading/References - - - - - - 32

31
1.0 Introduction
In order to explore the structure of crystals we require waves which interact with atoms

is, we require a wavelength of the order of 1 A: 10ef G<. The interaction should be
and which have a wavelength comparable with the inter atomic spacing in crystals; that

weak enough so that the wave can penetrate in a coherent fashion into the crystal for a
distance of the order of perhaps 1000 lattice constants. The most convenient waves
suitable for our purpose are those associated with x-rays, while the waves associated
with neutrons and electrons have found important special applications. When an atom is
exposed to electromagnetic radiation, the atomic electrons are accelerated, and they
radiate at the frequency of the incident radiation. The superposition of the waves
scattered by individual atoms in a crystal results in the ordinary optical refraction. If the
wavelength of the radiation is comparable with or smaller than the lattice constant, we
will also under certain conditions have diffraction of the incident beam.

2.0 Objectives
• To study the use of X-ray as a tool for investigating the structure of crystals.

3.0. Definition
When a monochromatic beam of x-rays is shone upon a regular crystalline material then
the beam will be scattered from the material at definite angles. This produced an
interference effect called diffraction between the X-rays from different layers within the
crystal.

3.1 Bragg formulation of diffraction by a crystal

W. L. Bragg (1913) found that one could account for the position of the diffracted
beams produced by a crystal in an x-ray beam by a very simple model according to
which x-rays are reflected from various planes of atoms in the crystal. The diffracted
beams are found for situations in which the reflections from parallel planes of atoms
interfere constructively. The derivation of the Bragg law is indicated in Fig. 4.1. We
consider in the crystal a series of atomic planes which are partly reflecting for radiation
of wavelength X and which are spaced equal distances d apart. The radiation is incident

is 25Xg +. Reinforcement of the radiation reflected from successive planes will occur
in the plane of the paper. The path difference for rays reflected from adjacent planes

when the path difference is an integral number n of wavelengths. The condition for
constructive reflection is that

25 Xg +  h (4.1)

Equation (4.1) represents the Bragg law. The integer n represents the order of
corresponding reflection. It should be emphasized that the Bragg equation results from
the periodicity of the structure, without reference to the composition of the unit of
repetition.

32
Fig. 4.1: Derivation of the Bragg equation 25 Xg +  h ; here d is the spacing of
parallel atomic planes (After Ashcroft and Mermin, 1976).

Worked example:
(a) State Bragg’s law of diffraction and give two geometrical facts that are
necessary for the derivation of the law.

crystal structure, shows diffraction peaks at the following 2+:40, 58, 73, 86.8,
(b) An X-ray Diffractometer recorder chat for an element, which has a cubic

100.4 and 114.7. The wavelength of the incoming X-rays used was1.540 k.
(i) determine the type of the cubic structure possessed by the element
(ii) Determine the lattice constant of the element.

Solution:

(a) Bragg’s law of diffraction states that the path difference between two X-rays
which are reflected from adjacent planes is an integral multiple of its wavelength

25 sin +  h
i.e.,

Where; +  Bragg’s angle

h  Wavelength of the X-rays


d= interatomic plane spacing

n = order of diffraction
The two geometrical facts are:
(i) The incident beam, the normal to the diffraction plane and the diffracted beam are

(ii) The angle between the diffracted beam is always 2+. this is known as the
always coplanar.

(b) (i).The values of the angles given are2+. Therefore, + is equal to the half the
diffraction angle.

2+ values. The ratio of the square of the sine of the 1st two planes gives the

 Structure type
UP. l-
true structure of the element, i.e.,

UP. l.
 If the ratio is 0.5, the structure is [GG.
 If the ratio is 0.75, it is IGG

33
The 1st two planes have + values being 20, and 29 and the sines of these angles are

Xg E +D 0.117
0.3420 and 0.4848 respectively, therefore,
 n 0.5
Xg E +E 0.235

Hence, the crystal structure is bcc.

(ii). The relationship between Miller indices:; < of the Bragg plane and the Bragg

hE
angle is given by
Xg E +  E :E  ; E  E <
4*

For a bcc lattice, the sum   ;  must be even, hence the 1st set of principal
Where; a is the lattice constant.

diffraction plane for the bcc structure is o110p and the corresponding value for Xg E +
is 0.117, then,
h E  ; E  E
* q
2 Xg E +

This implies, *  t  0.318 nm


%.Drs  D. /D. /%
E %.DDu
w *  3.18k

3.2 Von Laue formulation of diffraction by a crystal


Considering the nature of the x-ray diffraction pattern produced by identical atoms
located at the corners (lattice points) of primitive cells of a space lattice to investigate

unit incident wave normal is XL and the unit scattered wave normal is s. let us examine at
scattering from any two lattice points, P1 and P2 (Fig. 4.2) separated by the vector r. The

a point a long distance away the difference in phase of the radiation scattered by P1 and
P2. If P1B and P2A are the projections of r on the incident and scattered wave directions,
the path difference between the two scattered waves is

xD  xE    · \   · \z{  · :\  \z < (4.2)

Fig.4.2: Calculation of the Phase difference of the waves scattered from two
lattice points (After Kittel, 1979)
34
The vector \  \z = S has a simple interpretation (Fig. 4.3) as the direction of the
normal to a plane that would reflect the incident direction into the scattering direction.

2+ is the angle s makes with\z , then + is the angle of incidence, and from the figure
This plane is a useful mathematical construction and this is called the reflecting plane. If

(4.3), we see that |}|  2Xg + , as s and s0 are unit vectors.

Fig.4.3: Construction of the normal the reflecting plane (After Kittel, 1979)

The phase difference ~ gX 2)h times the path difference. We have

~  F2)hH: · \< (4.3)

The amplitude of the scattered wave is a maximum in a direction such that the
contributions from each lattice point differ in phase only by integral multiplies of2.

multiple of 2 . If a, b, c are the basis vectors, we must have for the diffraction maxima
This is satisfied if the phase difference between adjacent lattice points is an integral

~8{ F2)hH: · \<  2;


~R{ F2)hH:
· \<  2;;
~N{ F2)hH: · \<  2 ;
(4.4)

where h, k; l are integers.

If , ,  are the direction cosines of S with respect to a, b, c, we have

 · \  2* Xg +  h

· \  2[ Xg +  ;h
· \  2G Xg +  h
(4.5)

values of + and the Wavelength h. The Laue equations (4.5) have a simple geometrical
Equations (4.4 & 4.5) are the Laue equations. They have solutions only for special

interpretation. The Laue equations state that in a diffraction direction the direction
cosines are proportional to h/a, k/b, l/c, respectively and the adjacent lattice planes:; )

35
direction cosines of the normal to :; < are proportional to h/a, k/b, l/c respectively
intersect the axes at intervals a/h, b/k, c/l so that by elementary plane geometry the

Therefore the lattice planes :; < must be parallel to the reflecting plane. If 5:; ) is
the spacing between two adjacent planes of a set :; <, we have by projection

[) G
5:; <  *) = ;  ) (4.6)

Then, from (4.5), we have

25 :; < Xg +  h (4.7)

We may interpret (4.7) by giving an extended meaning to the spacing d(; ) when h, k, l
have a common factor n: the diffracted wave actually arises from the nth order reflection
from the true lattice planes, but we may as a mathematical device think of the diffracted
wave as a first order reflection from a set of planes parallel to the true lattice planes but
with a spacing d(; ) equal to l/n of the true spacing.

3.3 Diffraction of crystals by electrons


de Broglie in 1924 predicted that the wavelength associated with a particle of
momentum p= mv is given by

h  )J (4.8)
where h is plank’s constant. One of the most direct pieces of evidence of the wave
aspect of particles was provided by the electron diffraction experiments of Davisson and
Germer in 1972. They concluded that if one associates a wavelength with the electrons
given by (4.9), the diffraction pattern obtained can be interpreted in exactly the same
way as the X-ray diffraction patterns. As long as the velocity of the electrons is small
compared with the velocity of light, the wavelength of the electrons may be expressed in
terms of the accelerating voltage V as follows
1 mv 2 = eV Or λ = h
(2meV )
(
150 )
1/ 2
(4.9)

h is obtained in Angstroms if V is expressed in volts. Note that only 150 volts are
1/ 2
2 V

required to produce electrons of a wavelength of 1 compared with X-rays, which


require approximately 12,000 volts for 1 . Electrons are scattered by the nucleus as well
as by the electrons in the atoms. For spherical charge distribution one can show that the
scattering factor is given by

 :+ <   )2E :‚  IU < .


E ƒ .

UP l
(4.10)

Here IU is the scattering factor for X-rays, Z is the nuclear charge, and + is the Bragg
angle. As for X-rays the scattering factor decreases with increasing values of +.
However, there is a considerable difference between X-rays and electrons in that,
electrons are scattered much more efficiently by atoms than are X-rays. In fact, atoms
scattered electrons more strongly by several powers of ten for the energy involved. At
normal incidence an electron of about 50 keV has a penetration depth for elastic

36
scattering of only about 500 k, while for the small angles of incidence used in reflection
techniques this may be about 50 k measured perpendicularly to the surface. It is evident,
therefore, that electron diffraction is particularly useful in investigating the structure of
thin surface layers such as oxide on metals. Such layers would not be detected by X-rays
diffraction because the patterns obtained are characteristics for the bulk material.

3.4 Diffraction of crystals by neutrons


The mass of a neutron is about 2000 as large as that of an electron, so that according to

of the same velocity. Thus the energy of a neutron required to give 1 k is of order of
Eqn.(4.8) the wavelength associated with a neutron is about 1/2000 that for an electron

only 0.1eV. Such neutrons can be obtained from a chain-reacting pile, and diffraction
from crystals may be observed. Neutrons are scattered essentially by the nuclei of the
atoms, except when they are magnetic. The radius of an atomic nuclei is of the order of

scattering angle, because h „ 10eD… G. Also, the scattering power does not vary in a
10-13cm, and as a consequence, the atomic scattering factor is nearly independent of the

regular manner with the atomic number, so that light elements such as hydrogen and
carbon still produce relatively strong scattering. The scattering of X-rays by light
element is in contrast, of course, relatively weak. Thus the positions of such atoms in
crystalline solids may be determined from neutron diffraction experiments. Another
important aspect of neutron diffraction is the fact that scattering from neighboring
elements in the periodic system may differ appreciably. For example, neutron diffraction
allows one to detect with relative ease ordered phases of an alloy such as FeCo, whereas
their detection by X-rays is difficult. A particularly important aspect of neutron
diffraction is their use in investigating the magnetic structure of solids. This is a result of
the interaction between the magnetic moment of the neutron and that of the atoms
concerned. In a paramagnetic substance, in which the magnetic moments are randomly
oriented in space, this leads to incoherent scattering, resulting in a diffuse background.
This diffuse background of magnetic scattering is then superimposed on the lines
produced by the nuclear scattering mentioned above. In a ferromagnetic substance in
which the magnetic moments within a domain are lined up in parallel, this diffuse
background is absent. In an antiferromagnetic solid, the magnetic moments of particular
pairs of atoms are aligned antiparallel and hence, from the point of view of the neutron,
such atoms would appear to be different.

4.0 Conclusions

of the structure of solid crystals that given X-rays of 1 k it requires energy of the order
From the discussions of the application of scattering diffraction techniques to the study

of 104 eV, for electrons of 1 k it needs 102 eV while the energy of a neutron required to
give 1 k is of the order of 0.l eV. Thus the diffraction technique is a useful tool in the
investigation of the structure of solid crystal from surface thin layers to bulky materials.

Bragg condition for crystal diffraction is given by 25 Xg +  h


5.0 Summary

~8{ F2)hH: · \<  2;


• Laue condition for diffraction is given by

37
~R{ F2)hH:
· \<  2;;
~N{ F2)hH: · \<  2 ; and
 · \  2* Xg +  h

· \  2[ Xg +  ;h
· \  2G Xg +  h
• de Broglie Wavelength equation is given by h  )J

150 D
⁄E
• wavelength of electron associated with accelerating velocity is given by

h† ˆ
V
• Scattering factor of electron by neutron is obtained by

 :+ <   )2E :‚  IU < UP. l


E ƒ .

6.0. Tutor Marked Assignment

Q1. (a) Discuss the major experimental differences between x-ray, electron, and
neutron diffraction from the standpoint of the observed diffraction patterns
(b) Show that the Laue equations for the incident beam parallel to the z cube edge of
a simple cubic crystal give diffracted rays in the yz plane when

h) = 2 : E  ; E < : E  ; E <
* ‰{ Š: E
and  ;E<
Where l and k are integers and ‰ is the direction cosine of the diffracted ray
relative to the z axis.

Q2. While sitting in front of a color TV with a 25Kv picture tube potential, you have an

(a) Calculate the shortest wavelength (maximum energy) X-ray. (  6.6 Œ 10e…s X,
excellent chance of being irradiated with X-rays.

G  3 Œ 10f ⁄X, 1 Ž  1.6 Œ 10eD ,

for a first order reflection maximum at h  0.5 k. :’“8”  2.165 •⁄G… <
(b) For a rock salt:*‘ < crystal placed in front of the tube, calculate the Bragg angle

7.0 Further readings / References


Ashcroft, N.W., Mermin, D.N, Solid state physics, Saunders College Publishing,
1976.
Carpenter, GB., Principles of crystal structure determination, Benjamin, 1969.
Brown,J.G., X-rays and their applications, Plenum, 1975.
Kittel,C., Introduction to solid state physics, Wiley eastern limited,1979.
Stokes, H.T., Solid state physics, Allyn and Bacon, Inc,1987.
Gay,P., Crystalline State, Olover and Boyd,1972.

38
UNIT 5: CRYSTAL DIFFRACTION (II) Page
1.0 Introduction - - - - - - - - 34
2.0 Objectives - - - - - - - - 34
3.0 Definition - - - - - - - - 34
3.1 Reciprocal Lattice - - - - - - 34
3.2 Ewald’s Construction - - - - - 36
3.3 Brillouin Zones - - - - - - 37
4.0 Conclusion - - - - - - - - 38
5.0 Summary - - - - - - - - 38
5.0 Tutor Marked Assignment - - - - -- 38
7.0 Further Reading/References - - - - - - 39

39
1.0 Introduction
To explain the theory of X-ray diffraction by crystal planes, Ewald introduced the
concept of reciprocal lattice. According to this concept, the description of
interpenetrating planes inside a crystal could be obtained in space by means of a set of
points. Thus the properties of planes and points are interchangeable. The space
constructed from these points is called reciprocal lattice.
2.0 Objective
The objectives of this unit is to explain
 Reciprocal lattice
 Ewald’s construction
 Brillouin zones

3.0 Definition
The reciprocal space lattice is a set of imaginary points constructed in such a way that
the direction of a vector from one point to another coincides with the direction of a
normal to the real space planes and the separation of those points (absolute value of the
vector) is equal to the reciprocal of the real inter planar distance

3.1 Reciprocal Lattice


For a perfect single crystal, the reciprocal lattice is an infinite periodic three-
dimensional array of points whose spacing is inversely proportional to the distances
between the planes in the direct lattice. The axis vectors of the reciprocal lattice is given
by Eqn. (5.1)

–  ?a ˜  ?a ; ™  ?a

— — —


— ·
— ·
—
; ; (5.1)

If a, b, c are primitive vectors of the crystal lattice, then A, B, C are primitive vectors of
the reciprocal lattice. Each vector is orthogonal to two of the axis vectors of the crystal
lattice. Thus A, B, C has property:

– ·   ?a, ˜ ·   z, ™ ·   z;
– ·
 z, ˜ ·
 ?a, C·
 z,
–· z, ˜ ·  š, ™ ·  ?a,
(5.2)

Any arbitrary set of primitive vectors a, b, c of a given crystal lattice leads to the same
set of reciprocal lattice points.

G = hA + kB+ lC, (h, k, l are integers) (5.3)

Any vector G of the form in Eq. (5.3) is called a reciprocal lattice vector. Every crystal
structure has two lattices associated with it, the crystal lattice and the reciprocal lattice.
A diffraction pattern of a crystal is a map of the reciprocal lattice of the crystal; a
microscopic image, if it could be resolved on a fine scale, represents a map of the crystal
structure in real space. When we rotate a lattice crystal, we rotate both the direct lattice
and the reciprocal lattice. Vectors in the crystal lattice have the dimensions of [length];
vectors in the reciprocal lattice have the dimensions of [length]-1. In dealing with wave
properties of crystals, it is convenient to define the reciprocal lattice vector G as
40
G = ?a:–  ;˜  ™< (5.4)

This in conjunction with equation (1.1) yields

G Rn=2: D  ; E  … <=2 Œ g Y• (5.5)

Thus every vector of the equation (5.3) satisfies the condition

›Jœ> ·  ž  1 (5.6)

Some of the elementary properties of the reciprocal lattice are as follows:


I. The unit cell of the reciprocal lattice need not be a parallelepiped.
II. Simple cubic lattice is its own reciprocal, so is the hcp. On the other hand,
bcc and fcc are reciprocal of each other.
III. The volume of a unit cell of the reciprocal lattice is inversely proportional to
the volume of a unit cell of the direct lattice.
IV. If A is the matrix of the components of A1, B1, C1 and B for those of A2, B2,
C2 then B =A-1

The properties of the reciprocal lattice that make it of importance in the diffraction
theory are:
i. The vector G (hkl) from the origin to the point (h, k, l) of reciprocal lattice is
normal to the (hkl) plane of the crystal lattice.
ii. The length of the vector G(hkl) is equal to the reciprocal of the spacing of the
planes(hkl) of the crystal lattice

Worked example:

8 …
Prove that the reciprocal lattice vectors as defined in equation (5.1) satisfy:
–. ˜ Œ ™ 
.
Œ
Solution:

.
Œ 
. Œ   .  Œ

To solve the problem, we need to use the vector identities:

and  Œ :
Œ <  :. <
 :.
<

:2<?
From Eq. (5.1)
–. ˜ Œ ™  – · 2:
. Œ <  :. Œ <
3
|.
Œ |?

:?a<?
–· 2:
. Œ <  z3
|.
Œ |?

:?a<A
–· :.
Œ <:.
Œ <
|.
Œ |A

–. ˜ Œ ™  .
Œ
fŸ  
Then,

41
3.2 Ewald’s Construction in the reciprocal lattice
For simplicity, we draw the Ewald construction in two dimensions. Ewald put the
information about the wavelength and direction for the incident X-ray beam into
reciprocal lattice as follows (Fig.5.1). Draw a vector AO in the incident direction of
length terminating at the origin O. Construct a circle of radius (a sphere, called
reflex sphere, of radius in three dimensions) with centre at A. Two possibilities
arise:
1. The circle does not pass through any reciprocal point. This implies that the particular
wavelength in question would not be diffracted by that crystal in the orientation. Further,
if the magnitude of the vector where a is the lattice constant), the circle
would not pass through any point, showing that X-ray diffraction cannot occur if
. It may also be noticed that the longer the vector AO (the shorter the wavelength),
the greater is the likelihood of the circle’s intersecting a point, and hence of diffraction.
2. The circle passes through any point B of the reciprocal lattice. Join A and O to B.
Thus, OB is a reciprocal lattice vector, G and is normal to some set of lattice planes,
e.g., AE. Hence, , d is the interplanar for the set.

k′
E

Fig.5.1: Ewald’s construction in the reciprocal lattice (After Ashcroft & Mermin,
(1976)

Let k = OA and k′ = AB respectively be the incident wave vector and the reflected wave
vector. Thus,
k′ = k + G (5.7)

which shows that (i) scattering changes only the diffraction of k and (ii) the scattered
wave differs from the incident wave by a reciprocal lattice vector G. for diffraction, it is
necessary that the vector k′, that is the vector AB, equal in magnitude to the vector k:
2
= k2 (5.8)
42
2k·G + G2 = 0 (5.9)
(k + G/2)·G = 0 (5.10)

that AE = k + G/2) is perpendicular to OB. Thus OB = 2OE = (2 sin θ)/ h . Also OB


Equation (5.10) is Bragg’s law in vector form. Its scalar form can be obtained by noting

=1)5 . Hence,
(2 sin θ)/ h = 1)5
2d sin θ = h
This shows that the Bragg equation has a simple geometrical significance in the
reciprocal lattice.

3.3 Brillouin Zones


For solid state physics the most important statement of the diffraction condition was
given by Brillouin. Fig. 5.1 shows that incident wave and reflected wave make an equal
angle with the lattice plane AE, which is, therefore, a reflecting plane. The reciprocal
lattice vector G = OB is perpendicular to the reflecting plane AE. Thus, corresponding
to G = OB, the reflecting plane is AE (produced). From the relation k′ = k + G, we see
that (AO +OE) OB = 0. That is AE OB = 0. Thus, AE is perpendicular to OB and also
bisects it, since E is the midpoint of OB by construction. Hence, for a given reciprocal
lattice vector, its right bisector is the reflecting plane. One can extend the procedure for
finding the reflecting planes corresponding to reciprocal lattice vectors connecting the
reciprocal lattice point O (origin) with its neighbours in reciprocal space. The volume
bounded by these planes is referred to as the geometrical definition of the first Brillouin
zone (BZ).
Figure 5.2 gives a portion of reciprocal space for a two dimensional oblique lattice
showing the lines bisecting some reciprocal lattice from O. The six shortest of these
vectors can be right bisected to produce the first BZ centered on the reciprocal point O.

Fig.5.2: Construction of first BZ for a two-dimensional oblique lattice


(After Kittel.1979).

Mathematically the reflecting planes and hence the Brillouin zones could be calculated
from equation (5.9). For the simple square lattice (of lattice constant a), the reciprocal

G = : D >  E B<

lattice vectors are
8
(5.11)
43
The wave vector for an X-ray measured from the origin of the reciprocal lattice is

C  ;¡ >  ;S B (5.12)

Use of Eq. (5.11) and Eq. (5.12) in Eq.(5.9) gives

D ;¡  E ;S  : DE  EE <
Ÿ
8
(5.13)

By assigning different value to D , E , we can obtain various reflection lines. So all C-


vectors originating at the origin and ending on these lines, will produce Bragg reflection.

4.0 Conclusion
The reciprocal lattice explains the theory of X-ray diffraction by crystal planes while the
Brillouin zone gives a vivid interpretation of the diffraction condition.

A wavelength of the order of 1 (1 k =10ef cm) is require to explore the


5.0 Summary

structure of crystals
• The concept of reciprocal lattice explained the theory of X-ray diffraction by
crystal planes
• The reciprocal lattice is an infinite periodic three-dimensional array of points
whose spacing is inversely proportional to the distance between the planes in the
direct lattice.
• Brillouin zone gives a vivid interpretation of the diffraction condition.

6.0 Tutor marked assignment

Q1. Prove that the volume of the unit cell of the reciprocal lattice is proportional to
that of the corresponding direct lattice.

–  F3 . *)2H >  2*)23B ; ˜   F3 . *)2H >  2*)23B ; ™  GC


Q2. The primitive translational vectors of the hexagonal space lattice may taken as
- -
F H F H

Show that the volume of the primitive cell is F3F.H *)2H *E c


-
(a)

–  F2) D/E H >  22)*3B; ˜   F2) D/E H >  22)*3B ; ™  22)G3C


(b) Show that the primitive translations of the reciprocal lattice are

3 3
So that the lattice is its own reciprocal, but with a rotation axes.

Q3. Show that the volume of the first Brillouin zone is given by:2< ŠŽN . Where ŽN the
E

volume is is of a crystal primitive cell

44
7.0 Further reading/References
Brown, J.G., X-rays and their applications, Plenum, 1975
Carpenter, G.B., Principles of crystal structure determination, Benjamin, 1969
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992
Kittel, C., Introduction to solid state physics, Wiley eastern limited, 1976
Gay, P., Crystalline State, Olover and Boyd, 1972
Stokes, H.T., Solid state physics, Allyn and Bacon, Inc, 1987

45
UNIT 6: EXPERIMENTAL CRYSTAL STRUCTURE DETERMINATION Page
1.0 Introduction - - - - - - - - 41
2.0 Objectives - - - - - - - - 41
3.0 Definition - - - - - - - - 41
3.1 Laue method - - - - - - - 41
3.2 Rotating Crystal Technique - - - - - 42
3.3 Powder method - - - - - - 43
4.0 Conclusion - - - - - - - - 44
5.0 Summary - - - - - - - - 44
6.0 Tutor Marked Assignment - - - - - - 45
7.0 Further Reading/References - - - - - - 45

46
1.0 Introduction
In practice, to satisfy Bragg’s law for X
X-ray
ray diffraction, it is necessary to vary either the
angle of inclination of the specimen to the beam or the wavelength of radiation. The
three standard methods of X-ray
X ray crystallography to be discussed are the Laue method,
the Rotating crystal technique and the Powder method.

2.0 Objective
To explain experimental crystal structure determination according to:
• Laue method
• Rotating crystal technique
• Powder method

3.0 Definition
Experimental crystal structure determination is an experimental method to study
scattering of crystal based on Ewald’s simple geometric construction.
3.1 Laue method
In the Laue method (Fig 6.1), a single crystal is mounted on a gonimeter, which enables
the crystal to be rotated through known angles in two two perpendicular planes, and
maintained stationary in a beam of X-rays
X rays ranging in wavelength from about 0.1 to 2.0
A. The crystal selects out and diffracts those values of for which planes exits, of
spacing d and glancing angle θ , satisfying the Bragg equation. A flat photographic film
is placed to receive either the transmitted diffracted beam or the reflected diffracted
beam.
As shown in the figure (6.1), the resulting Laue pattern consists of a series of
spots. Sharp well-defined
defined spots on the film aare re good evidence of a perfect crystal
structure, whereas diffuse, broken or extended spots indicate lattice distortion, defects or
other departures from the perfect crystal lattice. The Laue pattern reveals the symmetry
of the crystal structure in the orie
orientation
ntation used; for example, if a cubic crystal is oriented
with a cube edge, i.e., a [100] axis, parallel the incident beam, the Laue pattern will
show the four fold symmetry appropriate to this axis.

Fig.6.1: Schematic representation of Laue technique (after Kacchava, 1990)

47
3.2 Rotating Crystal Technique
A small single crystal (1 mm dimension) is mounted on a goniometer which, in turn, is
rigidly fixed to a spindle so that the crystal can be rotated about a fixed axis in a beam of
monochromatic radiation. The specimen is usually oriented with one of the
graphic axes parallel to the axis of rotation. The resulting variation in θ brings
crystallographic
different lattice planes into position for reflection and diffracted images are recorded on
a photographic film placed cylindrically, coaxial with the rotating spindle (Fig.6.2).
(Fig.

Fig.6.2: Rotating crystal technique (after Kachhava, 1992)

To explain the general nature of the diffraction, consider a crystal mounted so that one
of the axes (e.g. is parallel to the axis of rotation, then diffraction cannot occur from
the planes of atoms parallel to this axis unless

(6.1)

where n is an integer [Fig.6.2 (a)]. The diffracted beam will, therefore, be along the
surface of a family of cones whos
whosee vertices are at the crystal, and whose semi-vertical
semi
angles are given by the above equation [Fig.6.2 (b)].

Fig.6.2: Diffraction pattern in rotating crystal technique


(a)Diffraction condition (b) Cones of diffraction (After Kachhava, 1992)

The diffracted
ffracted beams will only occur along those specific directions lying on the cones
for which the correct phase relationship also holds for planes parallel to the other two
coordinate axes. When the film is flattened out after development, these diffraction
images will lie on a series of lines called layer lines, as illustrated in Fig.6.3. All the
48
images on the zero layer line come from planes parallel to the axis of rotation, i.e.,
planes with l = 0, and the other layer lines arise from planes with l . . . . ., etc.
diffraction images from planes with the same values of h and k but different values of l,
all lie on one of a series of curves known as row lines which are transverse to the layer
lines and in the particular case when the A and B axes are perpendicular to C, they
intersect with the zero layer line at right angles.

Fig.6.3: Typical Rotation photograph (After Kachhava, 1992)

If is the separation of these layer lines and R is the radius of the camera, then from
Fig.6.2 (b),

(6.2)

From equation (6.1) and equation (6.2)

R (6.3)

(6.4)

By subsequent orientation of the crystal with A and B axes parallel to the axis of
rotation, the other unit cell parameters may be determined.

3.3: Powder method


In this technique, a monochromatic X X-ray
ray beam is allowed to irradiate a small specimen
of the substance grinded to a fine powder and contained in a thin-walled
thin walled glass capillary
tube. Since the orientation of the minute crystal fragments is completely random, a
certain number of them will lie with any set of lattice planes making exactly the correct
angle with the incident beam for reflection to occur. Further, these planes in the different
crystallites are randomly distributed about the axis of the incident beam so that the
corresponding reflections from all the crystallites in the specimen lie on a cone coaxial
with the axis and with a semi-apex
semi angle of twice the Bragg angle (i.e.2θ
e.2θ). The specimen

49
is surrounded by a cylindrical film and two small portions of each cone are recorded as
lines on the film (Fig.6.3). If the grain size is fairly large (> 10-6 m), there is insufficient
room within the irradiated volume for enough crystallites crystallites to be in all possible
orientations and the resultant powder lines will be rather ‘spotty’. This spottiness can be
eliminated by rotating the specimen during exposure this considerably increases the
number of crystallites which can contribute to each powder line.

Fig.6.3: Schematic of powder method (a) experimental arrangement


(b)Diffraction geometry (c) Developed films (After Kachhava, 1992)

The Bragg angle θ of the various reflections can be calculated by measuring the
separation of the pairs
irs of lines since, from the geometry of Fig.6.2 (b)

(6.5)

where R is the radius of the camera. The reflections can be indexed and the unit cell
parameters evaluated.

4.0 Conclusion
This unit showed that the three methods discussed are tools for better understanding of
diffraction phenomena in crystalline samples.
samples

5.0 Summary
• Variation of the angle of inclination of the specimen to the beam or the
wavelength of radiation allows better understanding of Bragg’s law.
• In the Laue technique , a single stationary crystal is irradiated by a range of X-ray
X
wavelengths
• in the Rotational crystal method, a single crystal specimen is rotated in a beam of
monochromatic x-rays
rays wavelength
• in the Powder technique, a polycrystalline powder sspecimen
pecimen is kept stationary in a
beam of monochromatic radiation.

50
6.0 Tutor marked Assignment

lines on the powder photographs of IGG crystal: a = 6.0 and h  1.54k


Q1. Find the Bragg angles and the indices of diffraction for the three lowest angle

Cobalt has two forms: α-Co, with GJ structure (lattice spacing of *= 2.15 ) and
-Co, with IGG structure (lattice spacing of *N,RPN = 3.55 k). Assume that the GJ
Q2.

structure has an ideal G⁄* ration. Calculate and compare the position of the first
five X-ray powder diffraction peaks. The quantity £  4)h sin + can be used to
characterize the peak positions (here h is the wavelength of the X-ray radiation
and 2 + is the scattering angle)

7.0 Further Reading/References

Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company


Limited, New Delhi, 1992
Carpenter, G.B., Principles of crystal structure determination, Benjamin, 1969
Brown, J.G., X-rays and their applications, Plenum, 1975
Kittel, C., Introduction to solid state physics, Wiley eastern limited, 1996
Stokes, H.T., Solid state physics, Allyn and Bacon, Inc, 1987
Gay, P., Crystalline State, Olover and Boyd, 1972

51
Module 2:

CRYSTAL ELASTIC CONSTANTS AND


VIBRATIONS

52
UNIT 1: ELASTIC CONSTANTS OF CRYSTALS Page
1.0 Introduction - - - - - - - - 48
2.0 Objectives - - - - - - - - 48
3.0 Definition - - - - - - - - 48
3.1 Analysis of elastic strains and stresses - - - 48
3.2 Dilation - - - - - - - 50
3.3 Shearing strain - - - - - 50
3.4 Stress components - - - - - - 50
3.5 Elastic compliance and stiffness constants - - 51
3.6 Energy density - - - - - - - 52
3.7 Cubic crystals - - - - - - 53
4.0 Conclusion - - - - - - - - 54
5.0 Summary - - - - - - - - 54
6.0 Tutor Marked Assignment - - - - - - 54
6.0 Further Reading/References - - - - - - 55

53
1.0 Introduction
The study of the elastic behavior of solids is very important in the fundamental and
technical researches. In technology, it would tell us about the strength of the materials. In
fundamental research, it is of interest because of the insight it provides in to the nature of
binding forces in solids. They are also of importance for the thermal properties of solids.

2.0 Objective

• To explain elastic constant in solids


• To explain strength of solid materials
• To understand fully the binding forces in solids

3.0 Definition
Elasticity is the study of the ability of crystals to incorporate changes or adapt to new
circumstances easily

The local elastic strain of a body may be specified by six numbers. If , ,  are the
3.1 Analysis of elastic strains and stresses

Δ, Δ, ā; Δ*, Δ[, ΔG resulting from the deformation. This is a good physical
angles between the unit cell axes a, b, c, the strain may be specified by the changes

specification of strain, but for non-orthogonal axes it leads to mathematical

components ¡¡. , SS ,‰‰ , ¡S , S‰, ‰¡ which are defined below. We imagine that three
complications. The strain may be specified in terms of the six

orthogonal axes f, g, h of unit length are embedded securely in the unstrained solid, as

the axes, which we now label, ]¥, ¦¥, ^¥, are distorted in orientation and in length, so that
shown in Fig. 1.1(a). We suppose that after a small uniform deformation has taken place

with the same atom as origin we may write.

]¥  :1  §¡¡ <]  §¡S ¦  §¡‰ ^;


¦¥  §S¡ ]  21  §SS 3¦  §S‰ ^;
^¥  §‰¡ ]  §‰S ¦  :1  §‰‰ <^
(1.1)

The fractional changes of length of the f, g, and h. axes are §¡¡ , §SS , §‰‰ respectively, to
the first order. We define the strain components ¡¡. , SS ,‰‰ by the relations

¡¡  §¡¡ ; SS  §SS ; ‰‰  §‰‰ ; (1.2)

The strain components ¡S , S‰, ‰¡ may be defined as the changes in angle between

¡S  ]¥ · ¦¥  ¨©— +¨—© ;


the axes, so that to the first order

S‰  ¦¥ · ^¥  ¨ª© +¨©ª ;


‰¡  ^¥ · ]¥  ¨ª— +¨—ª ;
(1.3)

54
This completes the definition of the six strain components. A deformation is uniform if
the values of the strain components are independent of the choice of origin.

Fig.1.1 Coordinate axes for the description of the state of strain; the
orthogonal unit axes in the unstrained state (a) are deformed in the strained
state (b).

pure rotation ¨©—  ¨—© ; ¨ª©  ¨©ª ; ¨ª—  ¨—ª . If we exclude pure rotations,
We note that merely rotating the axes does not change the angle between them, so for a

we may without further loss of generality take ¨©—  ¨—© ; ¨ª©  ¨©ª ; ¨ª— 
¨—ª . so that in terms of the strain components we have
]¥  ]  ¡¡  ¡S ¦  ‰¡ ^ ;
D D
E E
¦ ¦ ›¬ ]  SS ¦  S‰ ^;
« D D
E E
^¥  ^  ‰¡ ]  S‰ ¦  ‰‰ ^ ;
D D
(1.4)

E E

We consider under a deformation which is substantially uniform near the origin a particle

  ›]  ¬¦  ­^
originally at the position
(1.5)

¥  ›]¥  ¬¦¥  ­^¥


After deformation the particle is at
(1.6)

®  «    ›:]«  ]<  ¬ :¦«  ¦<  ­:^«  ^<


so that the displacement is given by
(1.7)

®  Z]  ¯¦  ^
If we write the displacement as
(1.8)

55
we have from Eq.(1.4) and Eq.(1.7) the following expressions for the strain components:

¡¡ = °Z
°›
¬¬ = °¬
°¯
‰‰ = °
°­
±0 °Z
¡S = + ; ¬­ = ­› =
; ; ; (1.9)
±1 ±0 °Z °
±¡ ±S ±S ±‰ °­ °›
+ ; +

We have written derivatives for application to non-uniform strain. The expressions (1.9)

definitions of ¡S , ¬­ , and ­› are given which differ by a factor ½ from those given
are frequently used in the literature to define the strain components. Occasionally

here. For a uniform deformation the displacement ® has the components

 ¡¡ ›  ¡S ¬  ‰¡ ­ ;


D D
E E
u

 1
 ›  ¬¬ ¬  12 ¬­­;
2 ›¬
 ‰¡ ›  S‰ ¬  ‰‰ ­ ;
D D
v (1.10)

E E

3.2 Dilation
The fractional increment of volume caused by a deformation is called the dilation. The

››  ¬¬  ­­


unit cube of edges f, g, and h. after deformation has a volume

Ž «  ]« · ¦« ² ^« ³ 1  (1.11)

where squares and products of strain components are neglected. Thus the dilation is

´=  ››  ¬¬  ­­


µ¶
¶·
(1.12)

3.3 Shearing strain


We may interpret the strain components of the type

°Z
¡S =
±0
±¡ ±S
+

as made up of two simple shears. In one of the shears, planes of the material normal to
the x axis slide in the y direction; in the other shear, planes normal to y slide in the x
direction.

3.4 Stress Components


The force acting on a unit area in the solid is defined as the stress. There are nine stress
components: Xx, Xy, Xz, Yx, Yy, Yz, Zx, Zy, Zz. The capital letter indicates the direction of the
force, and the subscript indicates the normal to the plane to which the force is applied.
Thus the stress component Xx represents a force applied in the x direction to a unit area of
a plane whose normal lies in the x direction; the stress component Xy represents a force

56
applied in the x direction to a unit area of a plane whose normal lies in the y direction.
The number of independent stress components is reduced to six by applying to an
elementary cube as in Fig. 1.2 the condition that the angular acceleration vanish, and
hence that the total torque must be zero. It follows that

Yz = Zy , Zx = Xz,, Xy = Yx
and the independent stress components may be taken as Xx, Yy, Zz, Yz, Zx, Xy The stress
components have the dimensions of force per unit area or energy per unit volume, which
the strain components are dimensionless

Fig. 1.2: Demonstration that number of independent stress components Yx


= Xy order that the body may be in equilibrium.

3.5 Elastic Compliance and Stiffness Constants


Hooke's law states that for small deformations the strain is proportional to the stress, so
that the strain components are linear functions of the stress components:

¸——  \== ²—  \=? ¹©  \=A ºª  \=` ¹ª  \=» º—  \=b ²© ;

¸©©  \?= ²—  \?? ¹©  \?A ºª  \?` ¹ª  \?» º—  \?b ²© ;

¸ªª  \A= ²—  \A? ¹©  \AA ºª  \A` ¹ª  \A» º—  \Ab ²© ; (1.13)

¸©ª  \`= ²—  \`? ¹©  \`A ºª  \`` ¹ª  \`» º—  \`b ²© ;

¸ª—  \»= ²—  \»? ¹©  \»A ºª  \»` ¹ª  \»» º—  \»b ²© ;

¸—©  \b= ²—  \b? ¹©  \bA ºª  \b` ¹ª  \b» º—  \bb ²©

Conversely, the stress components are linear functions of the strain components:

²—  == ¸——  =? ¸©©  =A ¸ªª  =` ¸©ª  =» ¸ª—  =b ¸—© ;

¹©  ?= ¸——  \?? ¸©©  ?A ¸ªª  ?` ¸©ª  ?» ¸ª—  ?b ¸—© ;


57
ºª  A= ¸——  A? ¸©©  AA ¸ªª  A` ¸©ª  A» ¸ª—  Ab ¸—© ;

¹ª  `= ¸——  `? ¸©©  `A ¸ªª  `` ¸©ª  `» ¸ª—  `b ¸—© ;

º—  »= ¸——  »? ¸©©  »A ¸ªª  »` ¸©ª  »» ¸ª—  »b ¸—© ; (1.14)

²©  b= ¸——  b? ¸©©  bA ¸ªª  b` ¸©ª  b» ¸ª—  bb ¸—©

The quantities \== … . . \=? are called the elastic constants or elastic compliance
constants; the quantities == … … . . == are called the elastic stiffness constants or moduli
of elasticity. Other names are also current. The S’s and C’s have the dimension of area
per unit force or volume per unit energy and force per unit area or energy per unit volume
respectively

We calculate the increment of work ´½ done by the stress system in straining a small
3.6 Energy Density

cube of side L, with the origin at one corner of the cube and the coordinate axes parallel
to the cube edges. We have

´½  ¾ · ¿® (1.15)

where F is the applied force and

´® = ]´Z  ¦´¯  ^´ (1.16)

is the displacement. If X, Y, Z denote the components of F per unit area, then

´½  ÀE :Á´Z  ´¯  ‚´< (1.17)

We note that the displacement of the three cube faces containing the origin is zero, so that
the forces all act at a distance L from the origin. Now by definition of the strain
components

´Z  À F´¡¡  ´¡S  ´‰¡ H


D D
E E
(1.18)
etc., so that

´½  À… 2Á¡ ´¡¡  ÂS ´SS  ‚‰‰ ´‰‰  ‰ ´S‰  ‚‰¡ ´‰¡  ÁS ´¡S 3 (1.19)

The increment ´U of elastic energy per unit volume is

´Ã  Á¡ ´¡¡  ÂS ´SS  ‚‰ ´‰‰  ‰ ´S‰  ‚¡ ´‰¡  ÁS ´¡S (1.20)

We have ´Ã)´ = Á¡ and ´Ã)´ = ÂS and on further differentiation


¡¡ SS
58
´Á¡ ´ÂS
Š´ = Š
SS ´¡¡

=?  ?=
This leads from Eq. (1.14) to the relation

and in general we have

>B { B> (1.21)

giving fifteen relations among the thirty non-diagonal terms of the matrix of the Cs. The
thirty-six elastic stiffness constants are in this way reduced to twenty-one coefficients.
Similar relations hold among the elastic compliances. The matrix of the Cs or S's is
therefore symmetrical.

3.7 Cubic crystal


The number of independent elastic stiffness constants is usually reduced if the crystal
possesses symmetry elements, and in the important case of cubic crystals there are only
three independent stiffness constants, as we now show. We suppose that the coordinate
axes are chosen parallel to the cube edges. In Eq. (1.14) we must have

GDs  GDr  GDÄ  GEs  GEr  GEÄ  G…s  G…r  G…Ä  0

Since the stress must not be altered by reversing the direction of one of the other

GDD  GEE  G…… ,


coordinate axes. As the axes are equivalent, we also have

and GDE  GD…  GED  GE…  G…D  G…E ,

GDD and GDE . The last three lines of Eq.(1.14) are described by the independent
so that the first three lines of Eq.(1.14) are described by the two independent constants

constant Gss , as

Gss  GDr  GÄÄ

by equivalence of the axes, and the other constants all vanish because of their behavior on
reversing the direction of one or other axis. The array of values of the elastic stiffness
constant is therefore reduced for a cubic crystal to the matrix below:

Á¡ ‘DD ‘DE ‘DE 0 0 0


ÂS ‘DE ‘DD ‘DE 0 0 0
Ç Ç
‚ ‘DE ‘DE ‘DD 0 0 0
ÅGPÆ Å  ‰
 0 0 0 ‘ss 0 0
Ç ‰ Ç
(1.22)
‚¡ 0 0 0 0 ‘ss 0
ÁS 0 0 0 0 0 ‘ss

It is readily seen that for a cubic crystal


59
à  1)2 ‘DD 2¡¡
E
 SS
E
 ‰‰
E
3  ‘DE 2SS ‰‰  ‰‰ ¡¡  ¡¡ SS 3
1
 )2 ‘ss 2S‰  ‰¡
E E
 ¡SE
3 (1.23)

satisfies the Eq.(1.19); for the elastic energy density function.

°Ã)
For example, °SS  ‘DD SS  ‘DE ‰‰  ‘DE ¡¡  ÂS,
Using Eq. (1.22).
For cubic crystals the compliance and stiffness constants are related by

‘DD  :È ;
È-- /È-.
-- eÈ-. <:È-- /EÈ-. <

‘DE  :È ;
eÈ-.
-- eÈ-. <:È-- eEÈD-. <
(1.24)

‘Ds  1)}
ss

A general review of elastic constant data and of relationships among various coefficients
for the crystal classes has been given by Hearmon (1946).

4.0 Conclusion
The elastic properties of a crystal considered as homogeneous continuous medium rather
than a periodic array of atoms is obtained by Hook’s law and Newton second law.

5.0 Summary

¡¡., SS ,‰‰ , ¡S , S‰, ‰¡


• The local elastic strain of a body is specified by six component numbers:

• There are nine stress components: Xx, Xy, Xz, Yx, Yy, Yz, Zx, Zy, Zz
• A deformation is uniform if the values of the strain components are independent
of the choice of origin
• The fractional increment of volume caused by a deformation is called the dilation
• Cubic crystals have only three independent stiffness constants.

Q1. Show that the shear constant E :‘DD  ‘DE < in a cubic crystals defined by
D
6.0 Tutor marked assignment

setting ¡¡   SS   and all other strains equal to zero.


D
E

:110< plane in the œ11É0ž direction is equal :‘DD  ‘DE </2.


Q2. Prove that in a cubical, the effective elastic constant for a shear across the

60
7.0 Further Reading/ References
C. Kittel, Introduction to solid state physics, Wiley Eastern Limited, 1979
R. F. S. Hearmon, "Elastic constants of anisotropic materials," Revs. Modern
Phys. 18, 409-440 (1946).
A. E. H. Love, A treatise on the mathematical theory of elasticity, Dover
Publications, New York, 1944.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
C. Zener, Elasticity and anelasticity of metals, University of Chicago Press, Chicago,
1948.

61
UNIT 2: ELASTIC CONSTANTS OF CRYSTALS (II) Page
1.0 Introduction - - - - - - - - 57
2.0 Objectives - - - - - - - - 57
3.0 Definition - - - - - - - - 57
3.1 Elastic waves in cubic crystals - - - - 57
3.2 Elastic isotropy - - - - - - 58
3.3 Cauchy relations - - - - - - 59
4.0 Conclusion - - - - - - - - 60
5.0 Summary - - - - - - - - 60
6.0 Tutor Marked Assignment - - - - - - 60
7.0 Further Reading/References - - - - - - 61

62
1.0 Introduction
The elastic properties of a homogeneous crystal are generally anisotropic. Even in a cubic
crystal, the relationship between stress and strain depends on the orientation of the crystal
axes relative to stress. In general, the number of elastic constants characterizing a body is
large. However, this number is considerably reduced due to the symmetric nature of both
strain and stress tensors.

2.0 Objective
The objectives of this unit are to describe:
• Elastic waves in cubic crystals
• Elastic isotropy
• Cauchy relations
• Lattice theory of elastic coefficients
3.0 Definition
Same as in unit 1
3.1 Experimental determination of elastic constants.
The classic methods for the measurement of the elastic constants of crystals are
described in the review by Hearmon (1964). In this method, quartz transducer is
transmitted through the test crystal and reflected from the rear surface of the crystal back
to the transducer. The elapsed time between initiation and receipt of the pulse is
measured by standard electronic methods. The velocity is obtained by dividing the round
trip distance by the elapsed time. In a representative arrangement the experimental
frequency may be 15 s-1, and the pulse length 1 µsec. The wavelength is of the order of 3

stiffness constants ‘DD , ‘DE,; ‘ss of a cubic crystal may be determined from the
x 10-4 cm. The crystal specimen may be of the order of 1 cm in length. The elastic

velocityF‘DD)’H , where ’ is the density. A shear wave propagates along a cube axis
velocities of three waves. A longitudinal wave propagates along a cube axis with
D/E

with Velocity F )’H , while a shear wave with particle motion along a 11É0 direction
‘44 1/2

propagates along a 110 direction with velocity:‘DDe ‘DE<)2’


D/E
.

3.2 Elastic waves in cubic crystals


By considering the forces acting on an element of volume in the crystal we find for the
equation of motion in the x direction

’ZÊ   
±ËÌ ±ËÍ ±ËÎ
±¡ ±S ±‰
(2.1)

With similar equations for the y and z directions; ’ is the density and u is the
displacement and ZÊ is . . From Eq. (1.21) in unit 1, it follows, taking the cube edges as
Q. ,
QT
the x, y, z directions, that

’ZÊ  ‘12 F H ‘44 F  H


°›› °¬¬ °­­ ±MÌÍ ±MÎÌ
 ‘11 
°› °› °› ±S ±‰
63
’ZÊ F H :‘12  ‘44 < F H
This reduces, using Eq. (1.9) of unit 1, to
2 2 2 2 2
° Z ° Z ° Z ° ¯ ° 
 ‘11  ‘44  
°›2 °¬2 °­2 °›°¬ °›°­
(2.2)

Here u, v and w are components of displacement


One solution is given by a longitudinal wave,

Z  Z%  P:ÏTe7¡<
moving along the x cube edge; from (2.2)

ÐE ’  ; E ‘DD
Here ;  where h is wave vector and Ð  2¯ is the angular frequency

ƒ
So that the velocity is

D/E
‘
¯  Ð);  Ñ DD)’Ò (2.3)

Another solution is given by a transverse or shear wave moving along the y cube edge
with the particle motion in the x direction:

¯  ¯%  œP:ÏTe7S<ž
which gives, on substitution in Eq. (1.2)

ÐE ’  ; E ‘ss
so that;

D/E
‘
¯  Ñ ss)’Ò (2.4)

There is also a solution given by a shear wave moving in the z direction with particle
motion in the x direction. In general there are three types of wave motion for a given
direction of propagation in the crystal, but only for a few special directions can the
waves be classified as pure longitudinal or pure transverse.

3.3 Elastic isotropy

Ê DE  2‘ss < ± ,  ‘ss ÔE Z  :‘DE  ‘ss < ± 5g¯ ®


Óz ZÊ  :‘DD  ‘
By minor manipulations we may rewrite Eq. (2.2) as
.

±¡ . ±¡
(2.5)

64
where the displacement Õ  Z>  ¯B  C is not to be confused with density now
written as Óz . if

‘DD  ‘DE  2‘ss (2.6)

the first term on the right in (2.5) drops out, and we can write on summing with the
equations for the y and z motions:

®Ê  ‘ss ÔE ®  :‘DE  ‘ss < •*5 5g¯ ® (2.7)

This equation has the important property that it is invariant under rotations of the
reference axes, as each term in the equation is an invariant. Thus the relation (1.6) is the
condition that the crystal should be elastically isotropic; that is, that waves should
propagate in all directions with equal velocities. However, the longitudinal wave
velocity is not necessarily equal to the transverse wave velocity.
The anisotropy factor A in a cubic crystal is defined as

2‘ss
 ):‘  ‘ <
DD DE
(2.8)

and is unity for elastic isotropy.

3.4 Cauchy relation


There are among the elastic stiffness constants certain relations first obtained by
Cauchy. The relations reduce to

‘DE  ‘ss

‘DD  3‘ss . If then a cubic crystal were elastically isotropic and the Cauchy relation is
in a crystal of cubic symmetry. If this is satisfied, the isotropy condition (2.6) becomes

satisfied, the velocity of the transverse waves would be equal to the velocity of the
longitudinal waves.
The conditions for the validity of the Cauchy relations are:
I. All forces must be central, i.e., act along lines joining the centers of the atoms.
This is not generally true of covalent binding forces, nor of metallic binding
forces.
II. Every atom must be at a center of symmetry; that is, replacing every inter atomic
vector should not change the structure.
III. The crystal should be initially under no stress. In metallic lattices the nature of
the binding is not such that we would expect the Cauchy relation to work out
well. In ionic crystals the electrostatic interaction of the ions is the principal
interaction and is central in nature. It is not surprising that the Cauchy relation is
moderately well satisfied in the alkali halides

Show that the velocity of a longitudinal wave in theœ111ž direction of a cubic


Worked example:

D⁄E
crystal is given by ¯U   :‘DD  2‘DE  4‘ss <⁄’
D
…
.

65
For a longitudinal phonon in the œ111ždirection, u = v = w.
Solution:

Let Z  Z%  P7œ¡/S/‰ž⁄√…  ePÏT


Where ;  is the wave number and Ð  2Ö is angular frequency. From Eq. (2.2),

ƒ

ÐE ’  œ‘DD  2‘ss  :‘DE ‘ss <ž ; E ⁄3

Thus, the velocity Ð⁄; of the longitudinal wave in the œ111ž direction is given by

¯U  Ð⁄;  œ:‘DD  2‘DE  4‘ss ⁄3’<žD⁄E

4.0 Conclusion

central inters atomic forces leads to the well known Cauchy relation, ‘DE  ‘ss .
The existence of the centre of symmetry of a cubic crystal stable under the

This reduces the number of independent elastic constants of a cubic crystal to


two only.

5.0 Summary

‘
¯  Ð);  F DD)’H
D/E
The longitudinal wave velocity along the x cube edge is given by

• The transverse wave velocity along the y cube edge with the particle motion in

‘
¯  Ð);  F ss)’H
D/E
the x direction is given by

• The Cauchy relation is ‘DE  ‘ss

• Cauchy relation does not work well for metallic lattices while it is moderately
well satisfied in the alkali halides.

Q1. Show that the velocity of a longitudinal wave in the œ111ž direction of a
6.0 Tutor marked assignment

cubic crystal is given by

D/E
¯   :‘DD  2‘DE  4‘ss </’
D
…

Q2. Show that the velocity of a transverse wave in the œ111ž direction of a
cubic crystal is given by

D/E
¯   :‘DD  ‘DE  ‘ss </’
D
…

66
7.0 Further reading/references
A. E. H. Love, A treatise on the mathematical theory of elasticity, Dover
Publications, New York, 1944.
C. Kittel, Introduction to solid state physics, Wiley Eastern Limited, 1979
C. Zener, Elasticity and anelasticity of metals, University of Chicago Press,
Chicago,1948
R. F. S. Hearmon, "Elastic constants of anisotropic materials," Revs. Modern
Phys. 18, 409-440 (1946).
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
.

67
UNIT 3: CRYSTALS BINDING Page
1.0 Introduction - - - - - - - - 63
2.0 Objectives - - - - - - - - 63
3.0 Definition - - - - - - - - 63
3.1 Inter atomic force - - - - - - 63
3.2 Vander Waals (Molecular) bonding - - - 65
3.3 Ionic bonding - - - - - - 66
3.4 Covalent bonding - - - - - - 68
3.5 Metallic Bonding - - - - - - 68
4.0 Conclusion - - - - - - - - 69
5.0 Summary - - - - - - - - 69
6.0 Tutor Marked Assignment - - - - - - 69
7.0 Further Reading/References - - - - - - 69

68
1.0 Introduction
The attractive electrostatic interaction between the negative charges of the electrons and the
positive charges of the nuclei is entirely responsible for the cohesion of solid. As the atoms
come close together their closed electron shells will start to overlap. The Pauli principle
states that each electron state can be occupied by only one electron. In order to have overlap
of closed shells, electrons have to be excited to higher states. This costs energy and leads to a
repulsive interaction between the atoms. The repulsive interaction dominates for short
distances between atoms, while the attractive interaction dominates at large distances. The
actual atomic spacing in a crystal is defined by the equilibrium where the potential energy
exhibits a minimum.

2.0 Objective
To explain:
• Inter atomic forces
• Vander Waals bonding
• Ionic bonding
• Covalent bonding
• Metallic bonding

3.0 Definition
Crystal binding is the attractive inter atomic force that hold atom together in a crystal.

3.1 Inter atomic forces


Solids are stable structures, and therefore there exist interactions holding atoms in a
crystal together. For example a crystal of sodium chloride is more stable than a
collection of free Na and Cl atoms. This implies that the Na and Cl atoms attract each
other, i.e. there exist an attractive inter atomic force, which holds the atoms together.
This also implies that the energy of the crystal is lower than the energy of the free atoms.
The amount of energy which is required to pull the crystal apart into a set of free atoms
is called the cohesive energy of the crystal.

Cohesive energy = energy of free atoms – crystal energy

Magnitude of the cohesive energy varies for different solids from 1 to 10 eV/atom,
except inert gases in which the cohesive energy is of the order of 0.1eV/atom. The
cohesive energy controls the melting temperature. A typical curve for the potential
energy (binding energy) representing the interaction between two atoms is shown in
Fig.1.1 It has a minimum at some distance R=R0. For R>R0 the potential increases
gradually, approaching 0 as R→∞, while for R<R0 the potential increases very rapidly,
tending to infinity at R=0. Since the system tends to have the lowest possible energy, it
is most stable at R=R0, which is the equilibrium inter atomic distance. The
corresponding energy U0 is the cohesive energy. A typical value of the equilibrium
distance is of the order of a few angstroms (e.g. 2-3Å), so that the forces under
consideration are short range. The inter atomic force is determined by the gradient of the
potential energy, so that

69
× :Ø<  
±Ù
±Ú
(3.1)

If we apply this to the curve in Fig.3.1, we see that F(R) <0 for R>R0. This means that
for large separations the force is attractive, tending to pull the atoms together. On the
other, hand F(R)>0 for R<R0, i.e. the force becomes repulsive at small separations of the
atoms, and tends to push the atoms apart. The repulsive and attractive forces cancel each
other exactly at the point R0, which is the point of equilibrium. The attractive inter
atomic forces reflect the presence of bonds between atoms in solids, which are
responsible for the stability of the crystal. There are several types of bonding, depending
on the physical origin and nature of the bonding force involved.

Fig.3.1. A typical curve for the potential energy (binding energy)


representing the interaction between two atoms (After Kittel.1979)

Although the nature of the attractive energy is different in different solids, the origin of
the repulsive energy is similar in all solids and it is mainly due to the Pauli Exclusion
Principle. The elementary statement of this principle is that two electrons cannot occupy
the same orbital. As ions approach each other close enough, the orbits of the electrons
begin to overlap, i.e. some electrons attempt to occupy orbits already occupied by
others. This is, however, forbidden by the Pauli Exclusion Principle. As a result,
electrons are excited to unoccupied higher energy states of the atoms. Thus, the electron
overlap increases the total energy of the system and gives repulsive contribution to the
interaction. The repulsive interaction is not easy to treat analytically from first
principles. In order to make some quantitative estimates it is often assumed that this

/ρ ), where λ and ρ are some constants or of the form /Ø , where n is sufficiently large
interaction can be described by a central field repulsive potential of the form λ exp (−r

and B is some constant.

70
3.2 Vander Waals (Inter atomic) bonding
This type of binding is exhibited by solid noble gas crystals. The outermost electron shell is
completely filled and the electron distribution is spherically symmetric. Each atom is neutral
and has no permanent dipole moment. The attractive forces between the atoms arise from
fluctuations in the electron distribution. These give an instantaneous fluctuating dipole
moment in the atom. Its interaction with induced dipole moments in the neighboring atom

free atoms. The noble gases such as neon (Ne), argon :<, krypton (Kr) and xenon (Á)
leads to a weak interaction. The electron distribution in inert gases is very close to that in

are characterized by filled electron shells and a spherical distribution of electronic

IGG structure. Consider two inert gas atoms (1 and 2) separated by distance R. The
clouds in the free atoms. In the crystal the inert gas atoms pack together within the cubic

the average dipole moment of atom 1 is zero: ÛÜ1Ý 0. Here the brackets denote the time
average charge distribution in a single atom is spherically symmetric, which implies that

average of the dipole moment. However, at any moment of time there may be a non-zero
dipole moment caused by fluctuations of the electronic charge distribution. We denote
this dipole moment by d1. From electrostatics consideration, this dipole moment
produces an electric field, which induces a dipole moment on atom 2. This dipole
moment is proportional to the electric field which is in its turn proportional to the d1/R3
so that

5E Þ  Þ
Q-
Ú 
(3.2)

The dipole moments of the two atoms interact with each other. The energy is therefore
reduced due to this interaction. The energy of the interaction is proportional to the
product of the dipole moments and inversely proportional to the cube of the distance
between the atoms, so that

 Þ
Q- Q. Q-.
Ú  Úß
(3.3)

and that the coupling between the two dipoles, one caused by a fluctuation, and the other
induced by the electric field produced by the first, results in the attractive force, which is

average value of Û5DE Ý which is not vanish, even though ÛÜ1Ýis zero.
called the Van der Waals force. The time averaged potential is determined by the

à Þ
ÛQ-. Ý
Úß
(3.4)

The respective potential decreases as Ø Ä reduces with the separation between the atoms.
Van der Waals bonding is relatively weak; the respective cohesive energy is of the order
of 0.1eV/atom. This attractive interaction described by Eq. (3.4) holds only for a
relatively large separation between atoms. At small separations a very strong repulsive
forces cause by the overlap of the inner electronic shells start to dominate. It appears that
for inert gases this repulsive interaction can be fitted quite well by the potential of the

71
form )ØDE where B is a positive constant. Combining this with Eq. (3.4) we obtain the
total potential energy of two atoms at separation R which can be represented as

à DE à Ä
à  4§ F H  F H 
Ú Ú
(3.5)

where 4§á Ä ≡ A and 4§á EÄ≡ B. This potential is known as Lennard-Jones potential.

3.3 Ionic bonding


The ionic bond results from the electrostatic interaction of oppositely charged ions. Let

single valence electron to a neighboring Cl atom, producing Na+ and ‘ e− ions which
us take sodium chloride as an example. In the crystalline state, each Na atom loses its

have filled electronic shells. As a result an ionic crystal is formed containing positive

*  5.1Ž :âã g­*Ygã  •¬< ä * / e


and negative ions coupled by a strong electrostatic interaction.

 e  ‘ ä ‘ e  3.6Ž : GYã *IIg gY¬<


*/  ‘ e ä *‘  7.9Ž : GYãXY*YgG  •¬<

3.6eV, i.e. Na + Cl → *‘ + 6.4 eV (cohesive energy). The structure of *‘ is two
The cohesive energy with respect to neutral atoms can be calculated as 7.9eV - 5.1eV +

interpenetrating fcc lattices of Na+ and ‘ e ions as shown in Fig.3.2

Fig 3.2 structure of *‘ (After Kachhava, 1992)

Thus each Na+ ion is surrounded by 6 ‘ e ions and vice versa. This structure suggests
that there is a strong attractive Coulombic force between nearest-neighbors ions, which
is responsible for the ionic bonding. To calculate binding energy we need to include
Coulomb interactions with all atoms in the solid. Also we need to take into account the
repulsive energy, which we assume to be exponential. Thus the interaction between two
atoms i and j in a lattice is given by
eVæç
ê EŠ
ÃPÆ  h é
F )èH
PÆ (3.6)

Here PÆ is the distance between the two atoms, q is the electric charge on the atom, the
(+) sign is taken for the like charges and the (–) sign for the unlike charges. The total
energy of the crystal is the sum over i and j so that

72
(3.7)

In this formula ½ is due to the fact that each pair of interactions should be counted only
once. The second equality results from the fact in the structure the sum over j does
not depend on whether the reference ion i is positive or negative, which gives the total
number of atoms. The latter divided by two gives the number of molecules N, composed
of a positive and a negative ion. We assume for simplicity that the repulsive interaction
is non-zero only for the nearest neighbors (because it drops down very quickly with the
distance between atoms). In this case we obtain

(3.8)

Here R is the distance between the nearest neighbors; z is the number of the nearest
neighbors, and α is the Madelung constant:

(3.9)

where is defined by .The value of the Madelung constant plays an


important role in the theory of ionic crystals. In general it is not possible to compute the
Madelung constant analytically. A powerful method for calculation of lattice sums was
developed by Ewald, which is called Ewald summation. This method can be used for the
numerical evaluation of the Madelung constants in solids. Example considers a one-
dimensional lattice of ions of alternating sign as shown in Fig.3.3 below.

Fig.3.3: 1-D lattice of ions of alternating sign.

In this case

Or
(3.10)

The factor 2 occurs because there are two ions, one to the right and one to the left at
equal distances .
we sum this series by the expansion
73
í
Á
ln:1  › <  ì:1<eD

{D

Thus the Madelung constant for 1-dimensional chain is   2 ln2.


In three dimensions calculation of the series is much more difficult and cannot be
performed so easy. The values of the Madelung constants for various solids are
calculated, tabulated and can be found in literature (e.g.Kittel, 1996).

3.4 Covalent bonding


The covalent bond is another important type of bond which exits in many solids. The
covalent bond between two atoms is usually formed by two electrons, one from each
atom participating in the bond. The electrons forming the bond tend to be partly
localized in the region between the two atoms joined by the bond. Normally the covalent
bond is strong: for example, it is the bond, which couples carbon atoms in diamond. The
covalent bond is also responsible for the binding of silicon and germanium crystals. In a
two-atomic molecule (one electron per atom) the energy levels are split into a binding and
an antibinding one. The two electrons are shared between the two atoms and fill the lowest,
binding, molecular orbital. In a solid the energy levels are no longer discrete but the binding
and antibinding levels become broad energy bands. The structure of covalent crystals is
determined by the direction of the bonds, they have often fewer nearest neighbor atoms
(lower coordination number).

of ionic and covalent binding. Ex: î*X. î* has 3 valence electrons and X has 5. On the
Compounds where the atoms have different number of valence electrons exhibit a mixture

neighboring atoms. However if the bonds are to be symmetrical the î* will be negatively
average we have 4 electrons per atom which can be shared in tetrahedral bonds with

charged and X positively charged. Hence partial ionic binding cannot be avoided in this and
similar cases.

3.5 Metallic bonding


Metals are characterized by a high electrical conductivity, which implies that a large
number of electrons in a metal are free to move. The electrons capable to move
throughout the crystal are called the conductions electrons. Normally the valence
electrons in atoms become the conduction electrons in solids. The main feature of the
metallic bond is the lowering of the energy of the valence electrons in metal as compared
to the free atoms. Below, some qualitative arguments are given to explain this fact.

the momentum are related to each other so that ∆›∆J  ð . In a free atom the valence
According to the Heisenberg uncertainty principle the indefiniteness in coordinate and in

electrons are restricted by a relatively small volume. Therefore, ∆p is relatively large


which makes the kinetic energy of the valence electrons in a free atom large. On the other
hand in the crystalline state the electrons are free to move throughout the whole crystal,
the volume of which is large. Therefore the kinetic energy of the electrons is greatly
reduced, which leads to diminishing the total energy of the system in the solid. This
mechanism is the source of the metallic bonding. Figuratively speaking, the negatively
charged free electrons in a metal serve as glue that holds positively charged ions together.
The metallic bond is somewhat weaker than the ionic and covalent bond. For instance the
melting temperature of metallic sodium is about 4000 which is smaller than 11000 in
74
*‘ and about 4000 in diamond. Nevertheless, this type of bond should be regarded as
strong. In transition metals like Fe, Ni, Ti, Co the mechanism of metallic bonding is more
complex. This is due to the fact that in addition to s electrons which behave like free
electrons we have 3d electrons which are more localized. Hence the d electrons tend to
create covalent bonds with nearest neighbors. The d electrons are normally strongly
hybridized with s electrons making the picture of bonding much more complicated.

4.0 Conclusion
Solids are stable structures, and therefore there exist interactions holding atoms in a
crystal together. Depending on the distribution of the outer electrons with respect to the
ions, different binding types can occur.

5.0 Summary
• The cohesive energy is the energy that must be added to the crystal to separate it to
neutral free atoms at rest, at infinite separation.
• Crystals of inert gas atoms are bound by Vander Waals interaction.
• Ionic crystals are bound by electrostatic attraction of charged ions of separate signs.
• A covalent bound is characterized by the overlap of charge distributions of
antiparallel spin.
• Metals are bound by reduction in kinetic energy of the valence electrons in the metal
as compared with the free atom.

Repulsive potential between two atoms is represented by )Ø  , where


6.0 Tutor marked assignment
Q1.
constants A and n are phenomenological parameters.

6  eD
(a) Show that the equilibrium inter atomic distance is given by

Ø%  † E ˆ

êE 1
(b) Demonstrate that the cohesive energy per molecule at equilibrium is
Ã%   †1  ˆ
Ø%

(c) Calculate the constant n for *‘ , taking into account that the lattice
constant is a=5.63Å, α=1.75, q=e and the measured binding energy per
molecule for this crystal is −7.94 eV.
-2
Q2. Using the Lennard-Jones potential with e=1.04ÿ10 eV and s=3.40Å and
taking into account only nearest-neighbor atoms, calculate the lattice
parameter and the cohesive energy of the fcc crystal of Ar.

7.0 Further reading/References.


C. Kittel, Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992
R. F. S. Hearmon, "Elastic constants of anisotropic materials," Revs. Modern
75
Phys. 18, 409-440 (1946).
A. E. H. Love, A treatise on the mathematical theory of elasticity, Dover
Publications, New York, 1944.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
1938.
C. Zener, Elasticity and anelasticity of metals, University of Chicago Press,
Chicago, 1948.

76
UNIT 4: LATTICE VIBRATION Page
1.0 Introduction - - - - - - - - 72
2.0 Objectives - - - - - - - - 72
3.0 Definition - - - - - - - - 72
3.1 One-dimensional lattice - - - - - 72
3.2 Diatomic one-dimensional lattice - - - - 76
3.3 Three –dimensional lattice - - - - - 79
3.4 Phonon - - - - - - - - 81
4.0 Conclusion - - - - - - - - 83
5.0 Summary - - - - - - - - 83
6.0 Tutor Marked Assignment - - - - - - 83
7.0 Further Reading/References - - - - - - 84

77
1.0 Introduction
An important aspect of the study of solid state physics is the lattice dynamics, which
concerns itself with the vibrations of atoms about their equilibrium sites in a solid. These
vibrations occur at any temperature, even at absolute zero. They are responsible for the
thermal properties - heat capacity, thermal conductivity, thermal expansion, etc. of
insulators and contribute the greater part of the heat capacity of metals.

2.0 Objective
To describe:
• One-dimensional monatomic lattice.
• One-dimensional diatomic lattice
• Three- dimensional lattice.

3.0 Definition
Lattice vibration is a continuing periodic oscillation relative to a fixed reference point, or
a single complete oscillation.

3.1 One-dimensional monatomic lattice


Consider one-dimensional crystal lattice and assume that the forces between the atoms
in this lattice are proportional to relative displacements from the equilibrium positions.
This is known as the harmonic approximation, which holds well provided that the
displacements are small. One might think about the atoms in the lattice as interconnected
by elastic springs (Fig.4.1).

Fig.4.1: Lattice vibration of monatomic lattice


(After www.pa.uk.edu/kwng/phy/525/lec)

Therefore, the force exerted on nth atom in the lattice is given by

×  ‘ :Z/D  Z <  ‘ :ZeD  Z < (4.1)

motion of the  Y atom we obtain


where C is the interatomic force (elastic) constant. Applying Newton’s second law to the

ñ  ×  ‘:Z 1  Z <  ‘:Z 1  Z <  ‘:2Z  Z 1  Z 1 <(4.2)


Q . ,ò
QT .

where M is the mass of an atom. Note that we neglected here the interaction of the  Y
atom with all but its nearest neighbors. A similar equation should be written for each
atom in the lattice, resulting in N coupled differential equations, which should be solved
78
simultaneously (N is the total number of atoms in the lattice). In addition the boundary
conditions applied to the end atom in the lattice should be taken into account.
Now let us attempt a solution of the form

à   P:ó¡òe ÏT< (4.3)

where › is the equilibrium position of the  Y atom so that ›  *. This equation


represents a traveling wave, in which all the atoms oscillate with the same frequency Ð
and the same amplitude A and have wave vector q. Note that a solution of the form Eq.
(4.3) is only possible because of the transnational symmetry of the lattice. Now
substituting Eq. (4.3) into Eq.(4.2) and canceling the common quantities (the amplitude
and the time-dependent factor) we obtain

ñ:Ð E < Pó8  ‘ô2 Pó8   Pó:/D<8   Pó:eD<8 õ (4.4)

This equation can be further simplified by canceling the common factor gê * which
leads to

ñÐE  ‘22  gê*  gê* 3  2‘ :1  cos ê*<  4‘Xg E


ó8
E
(4.5)

We find therefore the dispersion relation for the frequency

Ð t ÷Xg ÷
s” ó8
ö E
(4.6)

which is the relationship between the frequency of vibrations and the wave vector q. This
dispersion relation has a number of important properties.

(i) Reducing to the first Brillouin zone. The frequency (4.6) and the displacement of the
atoms (5.3) do not change when we change q by q+2π/a. This means that these solutions
are physically identical. This allows us to set the range of independent values of q
within the first Brillouin zone, i.e.

8 ø ê ø 8
Ÿ Ÿ

Within this range of ê the Ð versus ê is shown in Fig.4.2. The maximum frequency
(4.7)

is@4‘ /ñ. The frequency is symmetric with respect to the sign change in q, i.e.
Ð:ê<=Ð:ê<). This is not surprising because a mode with positive q corresponds to the
wave traveling in the lattice from the left to the right and a mode with a negative q

equivalent in the lattice the frequency does not change with the sign change in ê. At the
corresponds to the wave traveling from the right to the left. Since these two directions are

boundaries of the Brillouin zone q=±π/a the solution represents a standing waveà 

79
:1<  ePÏT : atoms oscillate in the opposite phases depending on whether n is even or
odd. The wave moves neither right nor left.

Fig.4.2: Dispersion curve of a one-dimensional monatomic lattice


representing the First Brillouin Zone.
www.pa.uk.edu/kwng/phy/525/lec)

(ii) Phase and group velocity. The phase velocity is defined by

¯O 
Ï
ó
(4.8)
and the group velocity by

¯ù 


¯O is the velocity of the
(4.9)

propagation of the plane wave, whereas the ¯ù is the velocity of the propagation of the
The physical distinction between the two velocities is that

wave packet. The latter is the velocity for the propagation of energy in the medium. For
the particular dispersion relation Eq. (4.6) the group velocity is given by

¯ù  t cos
”8. ó8
ö E
(4.10)

As is seen from Eq. (4.10) the group velocity is zero at the edge of the zone where
q=±π/a. Here the wave is standing and therefore the transmission velocity for the energy
is zero.

ê*<<1. We can then expand the sine in Eq. (4.6) and obtain for the positive frequencies:
(iii) Long wavelength limit. The long wavelength limit implies that λ>>a. In this limit

80
(4.11)

We see that the frequency of vibration is proportional to the wave vector. This is
equivalent to the statement that velocity is independent of frequency. In this case

(4.12)
Worked example:
Atoms in crystals are held together by chemical bonds. Consider these bonds to be elastic
springs of the same force constants for one-dimensional crystal lattice. Suppose one of
the atoms is displaced from its mean position by an external force and then released;
(a) derive an expression for its periodic motion with respect to its nearest
neighbours
(b) prove that these atoms can vibrate with a number of discrete frequencies up to a
maximum value given by

Solution:
(a) Consider a linear chain of atoms connected by elastic springs, each of spring constants
β (Fig below). If the atoms are each of mass m and the distance between any two
consecutive atoms is ‘a’, then a small displacement by some external force on one of
them will result into an oscillatory motion?

FL FR
n-1 n n+1

Un-1 Un Un Un+1
a

The displacements of and atoms from their mean positions at


any instant will be and respectively. Also, the extension of the spring
between and atoms will be and therefore, the restoring
force FL on the atom due to the left spring will be
(i)
Similarly, the extension of the spring on the right of atom will be
and restoring force is given by
(ii)
The net force on the atom will be

(iii)
Applying Newton’s second law of motion to the displacement of the atom, we obtain

81
w  .ò  :Ã/D  ÃeD  2Ã <  0
Q. Ù
QT
Hence, Eq.(iv) is the equation of periodic motion of the Y atom with respect to
(iv)

:  1<Y and :  1<Y atoms.

à  à P:ÏT/7Ëò <
(b)The general solution of Eq. (iv), if the amplitude of this motion is U, is given by

Where Á is the distance of the Y atom from the origin i.e. Á  *. Similarly, if ÁeD
(v)

and Á/D are the distances of :  1<Y and :  1<Y atom from the origin, then
ÁeD  :  1<* and Á/D  :  1<*. Thus, we have

ÃeD  Ã P:ÏT/7Ëòú-<

Ã/D  Ã P:ÏT/7Ëòû- <


(vi)

Where Ð is the angular frequency and ; 



ƒ
Substituting Eq.(v) and Eq.(vi) into Eq.(iv) with Á  *, ÁeD  :  1<* and
Á/D  :  1<*, gives,

ÐE Ã PÏT  P78  Ã PÏT  P78 ô P78   e78  2õ

P78 E P78 E P78 P78 E


ÐE à PÏT  P78   ü† E ˆ  † eE ˆ  2ý    E  e E 

P78 78 E
 E  e E ;* E
ÐE  4g E þ   4 sin 
2g 2
4 ;* E
Ð 
E
sin 
 2

w Ð t sin
s 78
 E
(vii)

can vibrate. When sin  é1 i.e. when  , the maximum frequency is obtained
78 78 Ÿ
E.(vii) gives a number of frequencies with which the atoms of the 1-dimensional lattice
E E E

from Eq.(vii) as Ð  ét
s


3.2 Diatomic one-dimensional lattice

Fig.4.3 shows a diatomic lattice with the unit cell composed of two atoms of masses ñD
Now we consider a one-dimensional lattice with two non-equivalent atoms in a unit cell.

and ñE with the distance between two neighboring atoms a.

82
Fig.4.3: Lattice vibration of diatomic lattice (After
www.pa.uk.edu/kwng/phy/525/lec)

We can treat the motion of this lattice in a similar fashion as for monatomic lattice.
However, in this because we have two different kinds of atoms, we should write two

5E Ã
equations of motion:
ñD  ‘ :2Ã  Ã/D  ÃeD <
5Y E

ñE  ‘ :2Ã/D  Ã/E  Ã <


Q. Ùòû-
(4.13)

QT .

In analogy with the monatomic lattice we are looking for the solution in the form of
traveling mode for the two atoms:

à   Pó8
    DPó:/D<8   egÐY
Ã/D E 
(4.14)

2‘  ñD ÐE 2 cos ê* D
E      0
2‘ cos ê* 2‘  ñE Ð

E
(4.15)

This is a system of linear homogeneous equations for the unknowns A1 and A2. A
nontrivial solution exists only if the determinant of the matrix is zero. This leads to the

:2‘  ñD ÐE <:2‘  ñE ÐE <  4‘ E cos E ê*  0


secular equation
(4.16)

This is a quadratic equation, which can be readily solved

E
Ð  ‘F  H é ‘ tF  H 
E D D D D sUP. ó8
ö- ö. ö- ö. ö- ö.
(4.17)

Depending on sign in this formula there are two different solutions corresponding to two
different dispersion curves, as is shown in Fig.4.4:

83
Fig.4.4: Dispersion Curve for one-dimensional diatomic lattice.
(After www.pa.uk.edu/kwng/phy/525/lec)

The lower curve is called the acoustic branch, while the upper curve is called the optical
branch. The optical branch begins at q=0 and ω=0. Then with increasing q the frequency
increases in a linear fashion. This is why this branch is called acoustic: it corresponds to
elastic waves or sound. Eventually this curve saturates at the edge of the Brillouin zone.
On the other hand, the optical branch has a nonzero frequency at zero q

Ð%  t2‘ †ñ  ñ ˆ
1 1
1 2
(4.18)

and it does not change much with q.


The distinction between the acoustic and optical branches of lattice vibrations can be seen

acoustic branch ω=0 and D =E . So in this limit the two atoms in the cell have the same
most clearly by comparing them at q=0 (infinite wavelength). From Eq. (4.15), for the

amplitude and the phase. Therefore, the molecule oscillates as a rigid body, as shown in
Fig.4.5 for the acoustic mode. On the other hand, for the optical vibrations, substituting
Eq. (4.18) to Eq. (4.15), we obtain for q=0:

ñD D  ñE E  0 (4.19)

It implies that the optical oscillation takes place in such a way that the center of mass of a
molecule remains fixed. The two atoms move in out of phase as shown in Fig.4.5. The
frequency of these vibrations lies in infrared region which is the reason for referring to
this branch as optical.

84
Fig.4.5: Distinction between Acoustic and Optical wave (After Kittel,
1976)

3.3 Three- dimension


The concept of the division of the vibrational modes into acoustic and optical branches
can be generalized to be applicable to three-dimensional structure. To avoid mathematical
details we shall present only a qualitative discussion. Consider, first, the monatomic
Bravais lattice, in which each unit cell has a single atom. The equation of motion of each
atom can be written in a manner similar to that of Eq. (4.2). The solution of this equation
in three dimensions can be represented in terms of normal modes.

  – P:óVe ÏT< (4.20)

where the wave vector q specifies both the wavelength and direction of propagation. The
vector A determines the amplitude as well as the direction of vibration of the atoms. Thus
this vector specifies the polarization of the wave, i.e., whether the wave is longitudinal
(A parallel to q) or transverse (A perpendicular to q). When we substitute Eq.(5.20) into
the equation of motion, we obtain three simultaneous equations involving Ax, Ay. and Az,
the components of A. These equations are coupled together and are equivalent to a 3 x 3
matrix equation. The roots of this equation lead to three different dispersion relations, or
three dispersion curves, as shown in Fig.4.6. All the three branches pass through the
origin, which means all the branches are acoustic. This is of course to be expected, since
we are dealing with a monatomic Bravais lattice.

85
Fig.4.6: Dispersion curve.

The three branches in Fig.4.6 differ in their polarization. When q lies along a direction of
high symmetry - for example, the [100] or [110] directions − these waves may be
classified as either pure longitudinal or pure transverse waves. In that case, two of the
branches are transverse and one is longitudinal. One usually refers to these as the TA -
transverse acoustic and LA − longitudinal acoustic branches, respectively. However,
along non-symmetry directions the waves may not be pure longitudinal or pure
transverse, but have a mixed character.

Fig.4.7: Dispersion curve for Al in the [100] and [110] directions (After
Kittel, 1979)

Figure 4.7 shows the dispersion curves for Al in the [100] and [110] directions. Note that
in certain high-symmetry directions, such as the [100] in Al, the two transverse branches
coincide. The branches are then said to be degenerate.
We turn our attention now to the non-Bravais three-dimensional lattice. Here the unit
cell contains two or more atoms. If there are s atoms per cell, then on the basis of our
previous experience we conclude that there are 3s dispersion curves. Of these, three
branches are acoustic, and the remaining (3s −3) are optical. The mathematical
justification for this assertion is as follows: We write the equation of motion for each
86
atom in the cell, which results in s equations. Since these are vector equations, they are
equivalent to 3s scalar equations, which have 3s roots. It can be shown that three of these
roots always vanish at q = 0, which results in three acoustic branches. The remaining (3s
−3) roots, therefore, belong to the optical branches, as stated above. The acoustic
branches may be classified, as before, by their polarizations as TA1, TA2, and LA. The
optical branches can also be classified as longitudinal or transverse when q lies along a
high symmetry direction, and one speaks of LO and TO branches. As in the one-
dimensional case, one can also show that, for an optical branch, the atoms in the unit cell
vibrate out of phase relative to each other. As an example of a non-Bravais lattice, the
dispersion curves for Ge are shown in Fig.4.8. Since there are two atoms per unit cell in
germanium, there are six branches: three acoustic and three optical. Note that the two
transverse branches are degenerate along the [100] direction, as indicated earlier.

Fig.4.8: Dispersion curve for Ge along [100] and [110] directions


(After Kittel, 1979)
.
3.4 Phonons
So far we discussed a classical approach to the lattice vibrations. As we know from
quantum mechanics the energy levels of the harmonic oscillator are quantized. Similarly
the energy levels of lattice vibrations are quantized. The quantum of vibration is called a
phonon in analogy with the photon, which is the quantum of the electromagnetic wave.
We know that the allowed energy levels of the harmonic oscillator are given by

  2  1)23ðÐ (4.21)

where n is the quantum number. A normal vibration mode in a crystal of frequency ω is


given by Eq. (4.20). If the energy of this mode is given by Eq. (4.21) we can say that this
mode is occupied by n phonons of energy ω . The term ½ ω is the zero point energy of
the mode.
Let us now make a comparison between the classical and quantum solutions in one-
dimensional case. Consider a normal vibration

87
   P:ó¡e ÏT< (4.22)

where u is the displacement of an atom from its equilibrium position x and A is the
amplitude. The energy of this vibrational mode averaged over time is

  1)2 ñÐE E  2  1)23ðÐ (4.23)

It is evident from Eq.(4.23) above that there is a relationship between the amplitude of
vibration and the frequency and the phonon occupation of the mode. In classical
mechanics any amplitude of vibration is possible, whereas in quantum mechanics only
discrete values are allowed. This is shown in Fig.4.9.

Fig.4.9: Relation between amplitude and frequency (After Kittel, 1979)

The lattice with s atoms in a unit cell is described by 3s independent oscillators. The

equations as we discussed before. They are ÐO :< , where p denotes a particular mode,
frequencies of normal modes of these oscillators will be given by the solution of 3s linear

i.e. p = 1,…3s. The energy of this mode is given by

óO  F óO  H ðÐO :<


D
E
(4.24)

where óO the occupation is number of the normal mode and is an integer. A vibrational
state of the entire crystal is specified by giving the occupation numbers for each of the 3s
modes. The total vibrational energy of the crystal is the sum of the energies of the
individual modes, so that

  ∑óO óO  ∑óO F óO  EH ðÐJ :<


D
(4.25)

interaction occurs such as if photon had a momentum ðq. However, a phonon does not
Phonons can interact with other particles such as photons, neutrons and electrons. This

carry real physical momentum. The reason is that the center of mass of the crystal does
not change it position under vibrations (except q=0). In crystals there exist selection
rules for allowed transitions between quantum states. We saw that the elastic scattering of
88
an x-ray photon by a crystal is governed by the wave vector selection rule k′ = k +G,
where G is a vector in the reciprocal lattice; k is the wave vector of the incident photon
and k′ is the wave vector of the scattered photon. This equation can be considered as

acquires a momentum ðG. If the scattering of photon is inelastic and is accompanied


condition for the conservation of the momentum of the whole system, in which the lattice

by the excitation or absorption of a phonon the selection rule becomes

k′ = k ± q +G (4.26)

of phonon. Phonon dispersion relations ÐO q can be determined by the inelastic


where sign (+) corresponds to creation of phonon and sign (–) corresponds to absorption

scattering of neutrons with emission or absorption of phonons. In this case in addition to


the condition of the momentum conservation we have the requirement of conservation of
energy. The latter condition can be written as

 é ðÐ
ð. 7 . ð. 7 .
Eö Eö
(4.27)

where M is the mass of the neutron and k and k′ are the momenta of the incident and
scattered neutron. Once we know in experiment the kinetic energy of the incident and
scattered neutrons from Eq. (4.27) we can determine the frequency of the emitted or
absorbed phonon. Then experimentally we need to determine those directions, which
characterized by highest intensity of the scattered beam. For these directions the
conditions (5.26) are satisfied and therefore from Eq. (4.26) we can find the wave vector
of the phonon. Therefore, this is the way to obtain the dispersion conditions for the
frequency of phonons which we discussed before.

4.0 Conclusion
Lattice vibrations are elastic waves propagating within crystals and the quantum unit of

phonon spectrum, Ð.
vibration is a phonon. The general equation of motion provides the phonon dispersion or

5.0 Summary
• All lattice waves can be described by wave vectors that lie within the first
Brillouin zone

The energy of the phonon is ðÐ


• The quantum unit of vibration is a phonon.

6.0 Tutor marked assignment


Q1. Consider a linear chain in which alternative ions have masses M1 and M2
and only nearest neighbors interact.
(a) Discuss the form of the dispersion relation and the nature of the
vibrational modes when M1 >> M2.
(b) Show that for M1=M2 the dispersion relation becomes identical to that
for the monatomic lattice

89
Q2. Consider the normal modes of a linear chain in which the force constants
between nearest-neighbor atoms are alternatively C and 10C. Assuming
that the masses are equal and the nearest neighbor separation is a/2 find
ω(q) at q=0 and q=π/a. Sketch the dispersion curve. This problem
simulates a crystal of diatomic molecules such as H2.

7.0 Further reading/References


C. Kittel, Introduction to solid state physics, Wiley Eastern Limited, 1979
A. E. H. Love, A treatise on the mathematical theory of elasticity, Dover
Publications, New York, 1944.
Ashcroft, N.W., Mermin, D.N, Solid state physics, Saunders College Publishing,
1976
W. A. Wooster, A textbook on crystal physics, Cambridge University Press, 1938.
C. Zener, Elasticity and an elasticity of metals, University of Chicago Press,
chicago, 1948.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992

90
UNIT 5: THERMAL PROPERTIES Page
1.0 Introduction - - - - - - - -- 86
2.0 Objectives - - - - - - - -- 86
3.0 Definition - - - - - - - -- 86
3.1 Heat Capacity - - - - - - 86
3.2 Density of states - - - - - - 88
3.2.1 Debye model - - - - - - - 91
3.2.2 Einstein model - - - - - - 93
3.3 Thermal conductivity - - - - - 94
4.0 Conclusion - - - - - - - - 96
5.0 Summary - - - - - - - - 96
6.0 Tutor Marked Assignment - - - - - - 97
7.0 Further Reading/References - - - - - 97

91
1.0 Introduction
This unit is devoted to the description of certain thermal properties of solid
materials. The properties considered on the basis of atomic point of view are specific
heat, thermal expansion, equation of state and thermal conductivity. The most
fundamental approach for the theoretical evaluation of these characteristics for a solid is
to relate them to the internal energy, the total kinetic energy and potential energy of its
constituents.

2.0 Objective
To explain
• Lattice specific heats
• Debye model
• Einstein model
• Lattice thermal conductivity

3.0 Definition
Specific heat is a measure of the number of degrees of freedom of oscillating lattice.

3.1 Heat capacity


The heat capacity C is defined as the heat ∆Q which is required to raise the temperature by

‘

∆T, i.e.

(5.1)

If the process is carried out at constant volume V, then ∆Q = ∆E, where ∆E is the increase in

‘0  F H
±
internal energy of the system. The heat capacity at constant volume C is therefore given by
V

±
0
(5.2)

The contribution of the phonons to the heat capacity of the crystal is called the lattice heat
capacity.
The total energy of the phonons at temperature T in a crystal can be written as the sum of the
energies over all phonon modes, so that

  ìÛ óO ÝðÐ:< (5.3)
óO

Where Û óO Ý is the thermal equilibrium occupancy of phonons of wave vector q and mode p
(p = 1…3s, where s is the number of atoms in a unit cell). The angular brackets denote the
average in thermal equilibrium. Note that we assumed here that the zero-point energy is
chosen as the origin of the energy, so that the ground energy lies at zero. The average thermal
equilibrium can be calculated.

92
Consider a harmonic oscillator in a thermal bath. The probability to find this oscillator in
an excited state, which is characterized by a particular energy E is given by the
n
Boltzmann distribution

FeðÏ)7
H
x  x%  (5.4)

where the constant P is determined from the normalization condition.


í
0

ì x  1
{%
(5.5)

so that

í eD
FeðÏ)7
H
x% { Ñì  Ò :». b<
{%

The average excitation number of the oscillator is given by

í
{% 
∑í eðÏ⁄7

Û Ý  ì x 
∑í{% 
eðÏ⁄7

(5.7)
{%

The summation in the numerator can be performed using the known property of geometrical

í
progression:

ì ›  1 (5.8)
{%

Using this property we find:


í í
5 5 1 ›
ì ›  › 
ì ›  › :». <
5› 5› 1  › :1  ›<E
{% {%

Where›   e  , then we obtain


ð

Û Ý   
¡ D D
De¡ ¡ ú- eD M 2ð⁄ 3 eD
(5.10)

The distribution given by Eq. (5.10) is known as the Planck distribution. Coming back to the
expression for the total energy of the phonons, we find that

93
í
 ðÏ:ó<
ì
 ðÏ:ó<  1
(5.11)
{%

order to do this we need to introduce the density of modes or the density of states  :Ð).
Usually it is convenient to replace the summation over q by an integral over frequency. In

O :Ð<5Ð represents the number of modes of a given number s in the frequency range (ω, ω
+ 5Ð . Then the energy is

ðÐ
  ì  5ÐO :Ð<
 :ðÏ⁄7 < 1
(5.12)
O

The lattice heat capacity can be found by differentiation of this equation with respect to
temperature, so that
. ðЊ
†
ðÐ
ˆ  ; 
‘0   ; ∑  5ÐO :Ð<
±  
±
2
FðÐ); H
† 1ˆ
(5.13)

We see that the central problem is to find the density of states :Ð), the number of modes
per unit frequency range.
3.2 Density of state
Consider the longitudinal waves in a long bar. The solution for the displacement of atoms is

Z   Pó¡
given by
(5.14)
where we omitted a time-dependent factor it is irrelevant for the present discussion. We shall
now consider the effects of the boundary conditions on this solution. These boundary
conditions are determined by the external constraints applied to the ends of the bar. The most
convenient type of boundary condition is known as the periodic boundary condition. By this
we mean that the right end of the bar is constrained in such a way that it is always in the same
state of oscillation as the left end. It is as if the bar were deformed into a circular shape so
that the right end joined the left. Given that the length of the bar is L, if we take the origin as
being at the left end, the periodic condition means that

Z:›  0<  Z :›  À< (5.15)

where u is the solution given by Eq.(5.14). If we substitute (5.14) into (5.15), we find that

 Pó  1 (5.16)
This equation imposes a condition on the admissible values of q:

94
ê


(5.17)
where n = 0, + 1, ±2, etc. When these values are plotted along q-axis, they form a one-
dimensional mesh of regularly spaced points. The spacing between the points is 2π/L. When
the bar length is large, the spacing becomes small and the points form a quasi-continuous

arbitrary interval 5ê in q-space, and look for the number of modes whose q’s lie in this
mesh. Each q-value of Eq. (5.17) represents a mode of vibration. Suppose we choose an

interval. We assume here that L is large, so that the points are quasi-continuous, which is true
for the macroscopic objects. Since the spacing between the points is 2π/L, the number of


modes is


(5.18)
We are interested in the number of modes in the frequency range 5Ð lying betwe:Ð, Ð 
5Ð. The density of states (Ð) is defined such that (Ð)5Ð gives this number. Comparing
this definition with Eq.(5.18), one may write :Ð<5Ð  :À/2< 5ê, ã :Ð< 
:À/2</:5Ð/5ê<. We note from Fig.5.1, however, that in calculating :Ð< we must

is to multiply the above expression for :Ð< by a factor of two. That is,
include the modes lying in the negative q-region as well as in the positive region. The effect

:Ð< 
 D
Ÿ QÏ)Qó
(5.19)

We see that the density of states :Ð< is determined by the dispersion relation
Ð  Ð:ê<.
Now we extend these results to the 3D case. The wave solution analogous to (5.14) is

  –¸>2— —/© ©/ª ª3 (5.20)

Fig.5.1: Density of state


where the propagation is described by the wave vector q = (q q q ), whose direction
x, y, z
specifies the direction of wave propagation. Here again we need to take into account the
boundary conditions. For simplicity, we assume a cubic sample whose edge is L. By
imposing the periodic boundary conditions, one finds that the allowed values of q must
satisfy the condition
95
 PóÌ =  PóÍ   PóÎ (5.21)

:ê› ê¬ ê­<  : , , <


Therefore, the values are given by
EŸ EŸ EŸ
, ,   
(5.22)
where l, m, n are some integers
if we plot these values in a q-space, as in Fig.5.2, we obtain a three-dimensional cubic mesh.
3
The volume assigned to each point in this q-space is (2π/L) .

Fig.5.2:Three-dimensional cubic mesh (After Kittel, 1979)

in the spherical shell between the radii q and ê  5ê, as shown in Fig.2. The volume of this
Each point in Fig.5.2 determines one mode. We now wish to find the number of modes lying
3
shell is, 4πq2dq and since the volume per point is (2π/L) , it follows that the number we seek

…
FEŸH 4êE 5ê  4êE 5ê
 ¶
:EŸ< 
(5.23)

is equal to :Ð<5Ð . Thus, we arrive at


where V = L is the volume of the sample. By definition of the density of modes, this quantity

 :Ð < 
¶ó . D
EŸ. QÏ)

(5.24)

We note that Eq. (5.24) is valid only for an isotropic solid, in which the vibrational
frequency, ω, does not depend on the direction of q. Also we note that in the above
discussion we have associated a single mode with each value of q. This is not quite true for
the 3D case, because for each q there are actually three different modes, one longitudinal and
two transverse, associated with the same value of q. In addition, in the case of non-Bravais
lattice we have a few sites, so that the number of modes is 3s, where s is the number of non-
equivalent atoms. This should be taken into account by index p=1…3s in the density of states
because the dispersion relations for the longitudinal and transverse waves are different, and
acoustic and optical modes are different.

96
3.2.1 Debye model
The Debye model assumes that the acoustic modes give the dominant contribution to the heat
capacity. Within the Debye approximation the velocity of sound is taken a constant
independent of polarization as it would be in a classical elastic continuum. The dispersion

Ð  ¯ê
relation is written as
(5.25)
where v is the velocity of sound.
In this approximation the density of states is given by

:Ð< 
¶Ï.
EŸ. 0  
(5.26)
i.e. the density of states increases quadratically with the frequency.
The normalization condition for the density of states determines the limits of integration over
ω. The lower limit is obviously ω=0. The upper limit can be found from the condition that the
number of vibrational modes in a crystal is finite and is equal to the number of degrees of
freedom of the lattice. Assuming that there are N unit cells is the crystal, and there is only
one atom per cell (so that there are N atoms in the crystal), the total number of phonon modes
is 3N. Therefore, we can write

: < :». ?<
ì   Ð ÜÐ  3
_ z

where the cutoff frequency ω is known as Debye frequency. Assuming that the velocity of
D
the three acoustic modes is independent of polarization and substituting Eq.(5.26) in
Eq.(5.27) we obtain
D/…
Ð  F H
ğ. 0   “

(5.28)

ê
The cutoff wave vector which corresponds to this frequency is given by

-/ 
ß .!
{ {† ˆ
(5.29)

 "

number of modes with ê ø ê exhausts the number of degrees of freedom of the lattice.
so that modes of wave vector larger than q are not allowed. This is due to the fact that the
D

The thermal energy is given by Eq. (5.12), so that


ðÐ
  3 % 5Ð
Ï ¶Ï.
EŸ. 0   M 2ð⁄ 3 eD
(5.30)
where a factor of 3 is due to the assumption that the phonon velocity is independent of
polarization. This leads to

97
¡
3Žð Ï Ѕ 3Ž;s  s ›…
  E …  5Ð :ðÏ⁄7
<  E … …  5› ¡
2 ¯ %   1 2 ¯ ð  1

(5.31)
% 1

Where ›  ðÐ);  and




ðÐ +
›  );   )

(5.32)

The latter expression defines the Debye temperature


D/…
+  F H
ð0 ğ. “
7 ¶
(5.33)
The total phonon energy is then
…
  9;  F H %  5› Ì

¡ ¡ 
l M eD 
(5.34)

+)
where N is the number of atoms in the crystal and › = .

The heat capacity is most easily found by differentiating the middle expression of Eq.(5.31)
with respect to the temperature so that

ð)
…
‘0  5Ð  9; F H %  5› :M Ì
…¶ð. Ï Ï#M ¡ ¡ #M Ì

 

EŸ 0 7

.   . % ð) .
l e D<.
(5.35)
M   eD 

In the limit T>>θ, we can expand the expression under the integral and obtain:‘0  3; .
This is exactly the classical value for the heat capacity, which is known from the elementary

per a degree of freedom is equal to   ;  . Therefore for a system of N atoms  =3; 


physics. Recall that, according to the elementary thermodynamics the average thermal energy

which results in ‘¯  3; . This is known as the Dulong-Petit law.


Now consider an opposite limit, i.e. T<<θ. At very low temperatures we can approximate
(5.34) by letting the upper limit go to infinity. We obtain

… … Ÿ# …
  9;  F H % 5› Ì  9;  F H  ;  F H

í ¡ 
…Ÿ#

l M eD l Dr r l
(5.36)
  

and therefore

…
‘0  ;  F H
DEŸ#

r l
(5.37)

98
to  …. The cubic dependence may be understood from the following qualitative argument. At
We see that within the Debye model at low temperatures the heat capacity is proportional

low temperature, only a few modes are excited. These are the modes whose quantum energy
ω is less than k T. The number of these modes may be estimated by drawing a sphere in the
q-space whose frequency Ð  ð);  , and counting the number of points inside, as shown in
B

Fig. 5.3. This sphere may be called the thermal sphere, in analogy with the Debye sphere

to ê… ~Ѕ ~ … . Each mode is fully excited and has an average energy equal to k T. Therefore
discussed above. The number of modes inside the thermal sphere is proportional

the total energy of excitation is proportional to  … , which leads to a specific heat proportional
B

to  …, in agreement with Eq. (5.37).

Fig.5.3: The thermal sphere (After Kittel, 1979)


3.2.2 Einstein model

frequency Ð i.e.
Within the Einstein model the density of states is approximated by a delta function at some

:Ð<  ´:Ð  Ð <


where N is the total number of atoms (oscillators). Ð is known as the Einstein frequency.
(5.38)


3ðÏ%
The thermal energy of the system is then

  ⁄; < 1
:ðÐ
(5.39)

where a factor of 3 reflects the fact that there are three degree of freedom for each oscillator.
The heat capacity is then

‘0  F H  3;
± ðÏ% M ð%⁄ 
±
¶ 7
FM 2ð%⁄ 3 eDH
(5.40)

i.e. ‘0  3; , which is the Dulong-Petit law. At low temperatures however Eq.(5.40)
The high temperature limit for the Einstein model is the same as that for the Debye model,

ðÐ

decreases as ‘0 ~  ; , while the experimental form of the phonon is known to be T 3 as




given by the Debye model. The reason for this disagreement is that at low temperatures only
acoustic phonons are populated and the Debye model is much better approximation that the
Einstein model. The Einstein model is often used to approximate the optical phonon part of
the phonon spectrum. Concluding our discussion about the heat capacity we note that a real
99
density of vibrational modes could be much more complicated than those described by the
Debye and Einstein models.

3.3 Thermal conductivity


When the two ends of a given sample material are at two different temperatures, T and T
1 2
(T >T ), heat flows down the thermal gradient, i.e. from the hotter to the cooler end.
2 1
Observations show that the heat current density j (amount of heat flowing across unit area

&  £
Q

per unit time) is proportional to the temperature gradient (5/5›<. That is,


(5.41)

The proportionality constant K, known as the thermal conductivity, is a measure of the ease
of transmission of heat across the bar (the minus sign is included to make a positive
quantity).
Heat may be transmitted in the material by several independent agents. In metals, for
example, the heat is carried by both electrons and phonons, although the contribution of the
electrons is much larger. In insulators, on the other hand, heat is transmitted entirely by
phonons, since there are no mobile electrons in these substances. Here we consider only
transmission by phonons.
When we discuss transmission of heat by phonons, it is convenient to think of these as
forming a phonon gas. In every region of space there are phonons traveling randomly in all
directions, corresponding to all the q's in the Brillouin zone (BZ), much like the molecules in
an ordinary gas. The concentration of phonons at the hotter end of the sample is larger and
they move to the cooler end. The advantage of using this gas model is that many of the
familiar concepts of the kinetic theory of gases can also be applied here. In particular,
thermal conductivity is given by
£ ‘ ¯
D
… 0
(5.42)
where C is the specific heat per unit volume, v the velocity of the particle, and l its mean
V
free path. In the present case, v and l refer, of course, to the velocity and the mean free path

two consecutive scattering events, so that  ¯', where τ is the average time between
of the phonon, respectively. The mean free path is defined as the average distances between

collisions which is called collision time or relaxation time.


Let us give a qualitative explanation for Eq. (5.42). For simplicity we consider a one-
dimensional picture, in which phonons can move only along the x axis. We assume that a
temperature gradient is imposed along the x axis. We also assume that collisions between

energy density to a particular point of the sample œ:›<ž. The phonons which originate
phonons maintain local thermodynamic equilibrium; so that we can assign local thermal

from this point have this energy on average. At a given point x half the phonons come from
the high temperature side and half phonons come from the low temperature side. The
phonons which arrive to this point from the high-temperature side will, on the average, have

œ:›  <ž. Their contribution to the thermal current density at point x will therefore be the
had their last collision at point x−l, and will therefore carry a thermal energy density of

½¯œ:›  <ž. The phonons arriving at x from the low temperature side, on the other hand,

100
will contribute ½¯œ:›  <ž, since they come from the positive x-direction and are

& ½¯œ:›  <ž  ½¯œ:›  <ž


moving toward negative x. Adding these together gives

(5.43)
Provided that the variation in the temperature over the mean free path is very small we may

&  ¯ Q
F Q¡ H  ¯ E ' Q
F Q¡ H
Q Q
Q Q

expand this about the point x to find:

(5.44)

the x-component ¯¡ , and then average over all the angles. Since Û¯¡E Ý  Û¯SE Ý  Û¯‰E Ý 1)3 ¯ E
This result can be easily generalized to the three dimensional case. We need to replace v by

‘0  is the heat capacity we obtain,


Q
Q

and since

& ‘¯ ¯ F H
D Q

… 5›
(5.45)

where v is the phonon velocity.


Let us now discuss the dependence of the thermal conductivity j on temperature. The
dependence of C on temperature has already been studied in detail, while the velocity v is
V
found to be essentially insensitive to temperature. The mean free path l depends strongly on
temperature. Indeed, l is the average distance the phonon travels between two successive
collisions. Three important mechanisms may be distinguished: (a) The collision of a phonon
with other phonons, (b) the collision of a phonon with imperfections in the crystal, such as
impurities and dislocations, and (c) the collision of a phonon with the external boundaries of
the sample.
Consider a collision of type (a). The phonon-phonon scattering is due to the anharmonic
interaction between them. When the atomic displacements become appreciable, this gives
rise to anharmonic coupling between the phonons, causing their mutual scattering. Suppose
that two phonons of vectors q and q collide, and produce a third phonon of vector q . Since
1 2 3
momentum must be conserved, it follows that q = q + q . Although both q and q lie inside
3 1 2 1 2
the Brillouin zone (Brillouin zones are primitive cells that arise in the theories of
electronic levels - Band Theory), q may not do so. If it does, then the momentum of the
3
system before and after collision is the same. Such a process has no effect at all on thermal
resistivity, as it has no effect on the flow of the phonon system as a whole. It is called a
normal process. By contrast, if q lies outside the BZ, such a vector is not physically
3
meaningful according to our convention. We reduce it to its equivalent q inside the first BZ,
4
where q = q + G and G is the appropriate reciprocal lattice vector. As is seen from Fig.5.6,
3 4
the phonon q produced by the collision travels in a direction almost opposite to either of the
4
original phonons q and q . The difference in momentum is transferred to the center of mass
1 2
of the lattice. This type of process is highly efficient in changing the momentum of the
phonon, and is responsible for phonon scattering at high temperatures. It is known as the
umklapp process (German for "flipping over").

101
Fig.5.6: Umklapp process(After Kittel, 1979)
Phonon-phonon collisions become particularly important at high temperature, at which the

proportional to the temperature, that is, Þ 1⁄. This is reasonable, since the larger T is, the
atomic displacements are large. In this region, the corresponding mean free path is inversely

greater the number of phonons participating in the collision.


The second mechanism (b) which results in phonon scattering results from defects and
impurities. Real crystals are never perfect and there are always crystal imperfections in the

destroy the perfect periodicity of the crystal. At very low temperature (say below 10% £), both
crystal lattice, such as impurities and defects, which scatter phonons because they partially

phonon-phonon and phonon-imperfection collisions become ineffective, because, in the


former case, there are only a few phonons present, and in the latter the few phonons which
are excited at this low temperature are long-wavelength ones. These are not effectively
scattered by objects such as impurities, which are much smaller in size than the wavelength.
In the low-temperature region, the primary scattering mechanism is the external boundary of
the specimen, which leads to the so-called size or geometrical effects. This mechanism
becomes effective because the wavelengths of the excited phonons are very long -
comparable, in fact, to the size of the specimen. The mean free path here is l ~ L, where L is
roughly equal to the diameter of the specimen, and is therefore independent of temperature.

4.0 Conclusion
There are two contributions to thermal properties of solids: one comes from phonons (or
lattice vibrations) and another from electrons. In most solids, the energy given to lattice
vibrations is the dominant contribution to specific heat.

5.0 Summary
• Lattice heat capacity is the contribution of phonon to heat capacity

Dulong Petit law results in ‘¯  3; for N atoms


• Debye model at low temperature is proportional to T3

• Einstein model is used to approximate the optical part of the phonon spectrum
• Changing the momentum of the phonon which is responsible for phonon scattering at
high temperatures is known as the umklapp process

102
6.0 Tutor marked assignment

Q1. Using the dispersion relation for the monatomic linear lattice of N atoms with
nearest neighbor interactions, show that the density of vibrational modes is given by

 :Ð <  were Ð is the maximum frequency


E“ D
Ÿ
tÏ)
. eÏ.

Q2. In the Debye approximation, show that the mean square displacement of an

ÛØ E Ý 
.
…ðÏ
atom at absolute zero is

fŸ . è0 .
where v is the velocity of sound. Estimate this value
ðÐ2
for Cu (+   343£, ρ = 8920 kg/m , v = 3570 m/s).
3

7.0 Further reading/References


Ashcroft, N.W., Mermin, D.N, Solid state physics, Saunders College Publishing,
1976
C. Kittel, Introduction to solid state physics, Wiley Eastern Limited, 1979
Choquard, P. The Anharmonic Crystal, Benjamin, 1967
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992
W. A. Wooster, A textbook on crystal physics, Cambridge University Press, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8

103
Module 3:
FREE ELECTRON FERMI GAS

104
UNIT 1: FREE ELECTRON THEORY OF METALS page
1.0 Introduction - - - - - - - - - 100
2.0 Objectives - - - - - - - - 100
3.0 Definition - - - - - - - - 100
3.1 Free electron model- - - - - - - 100
3.2 One- dimension - - - - - - 100
3.3 Fermi distribution - - - - - - 102
3.4 Three -dimension - - - - - - 103
3.5 Heat capacity - - - - - - - 106
4.0 Conclusion - - - - - - - - 106
5.0 Summary - - - - - - - - 108
6.0 Tutor Marked Assignment - - - - - - 108
7.0 Further Reading/References - - - - - - 108

105
1.0 Introduction
The free electron theory of metals refers to the case in which the atomic valance electrons
are treated as if they are free rather than being bound to the lattice points. Our assumption
amounts to supposing that the electrons move in a uniform potential rather than the true
periodic potential provided by the positive ions. The basic assumption of the theory is
that a metal is equivalent to a gas of free electrons in an otherwise empty box.

2.0 Objective
• To revise the free electron gas (FEG) model and assumptions made.
• To understand how this simple model can be used to derive equations heat
capacity of the free electron.
• To employ the time-independent Schrodinger equation to derive the electron wave
functions and energies.

3.0 Definition
A free electron model is the simplest way to represent the electronic structure of metals.

3.1 Free electron model


A free electron model is the simplest way to represent the electronic structure of metals.
Although the free electron model is a great oversimplification of the reality, surprisingly
in many cases it works pretty well, so that it is able to describe many important properties
of metals. According to this model, the valence electrons of the constituent atoms of the
crystal become conduction electrons and travel freely throughout the crystal. Therefore,
within this model we neglect the interaction of conduction electrons with ions of the
lattice and the interaction between the conduction electrons. In this sense we are talking
about a free electron gas. However, there is a principle difference between the free
electron gas and ordinary gas of molecules. First, electrons are charged particles.
Therefore, in order to maintain the charge neutrality of the whole crystal, we need to
include positive ions. This is done within the jelly model, according to which the positive
charge of ions is smeared out uniformly throughout the crystal. This positive background
maintains the charge neutrality but does not exert any field on the electrons. Ions form a
uniform jelly into which electrons move. Second important property of the free electron
gas is that it should meet the Pauli Exclusion Principle, which leads to important
consequences.

3.2 One-dimension
We consider first a free electron gas in one dimension. We assume that an electron of

* :›< of the electron is a solution of the Schrödinger equation, +*:› < = *:›< where En
mass m is confined to a length L by infinite potential barriers. The wave function

is the energy of electron in the orbital. Since w can assume that the potential lies at zero,
the Hamiltonian H includes only the kinetic energy so that

+* :›<  * :›<   * :›<   * :› <


O. ð. Q .
E E Q¡ .
1.1
106
Note that this is a one-electron equation, which means that we neglect the electron-
electron interactions. We use the term orbital to describe the solution of this equation.
Since the is a continuous function and is equal to zero beyond the length L, the
boundary conditions for the wave function are . The solution of Eq.
(1.1) is therefore

1.2

where A is a constant and n is an integer. Substituting (1.2) into (1.1) we obtain the
eigenvalues

1.3

These solutions correspond to standing waves with a different number of nodes within
the potential well as is shown in Fig.1.1

Fig.1.1 First three energy levels and wave-functions of a free electron of mass m
confined to a line of length L.(Kittel, 1979).

Now we need to accommodate N valence electrons in these quantum states. According to


the Pauli Exclusion Principle no two electrons can have their quantum number identical.
That is, each electronic quantum state can be occupied by at most one electron. The
electronic state in a 1D solid is characterized by two quantum numbers that are n and ms,
where n is the positive integer and ms is the magnetic quantum number such that ms = ±½
according to spin orientation.
Therefore, each orbital labeled by the quantum number n can accommodate two
electrons, one with spin up and one with spin down orientation.
Let denote the highest filled energy level, where we start filling the levels from the
bottom (n = 1) and continue filling higher levels with electrons until all N electrons are

107
accommodated. It is convenient to suppose that N is an even number. The condition
= N determines the value of n for the uppermost filled level. The energy of the highest
occupied level is called the Fermi energy . For one -dimensional system of N electrons
we can define , using Eq. (1.3),

1.4

In metals the value of the Fermi energy is of the order of 5 eV. The ground state of the N
electron system is illustrated in Fig.1.2 a: All the electronic levels are filled up to the
Fermi energy. All the levels above are empty.

Fig. 1.2 (a) Occupation of energy levels according to the Pauli


exclusion principle, (b) The distribution function f(E), at T = 0°K
and T> 0°K.

3.3 Fermi distribution


This is the ground state of the N electron system at absolute zero. What happens if the
temperature is increased? The kinetic energy of the electron gas increases with
temperature. Therefore, some energy levels become occupied which were vacant at zero
temperature, and some levels become vacant which were occupied at absolute zero. The
distribution of electrons among the levels is usually described by the distribution
function which is defined as the probability that the level E is occupied by an
electron. Thus if the level is certainly empty, then 0, while if it is certainly full,
then In general, has a value between zero and unity. It follows from
the preceding discussion that the distribution functions for electrons at T = 0°K has the
form

108
1,  . /
I : <  , -
0,  . /
(1.5)

That is, all levels below / are completely filled, and all those above / are completely
empty. This function is plotted in Fig. 1.2(b), which shows the discontinuity at the Fermi
energy.
When the system is heated (T>0°K), thermal energy excites the electrons. However, all

treatment, because the electrons lying well below the Fermi level / cannot absorb
the electrons do not share this energy equally, as would be the case in the classical

energy. If they did so, they would move to a higher level, which would be already

the energy which an electron may absorb thermally is of the order ;  (= 0.025 Ž at
occupied, and hence the exclusion principle would be violated. Recall in this context that

room temperature), which is much smaller than / , this being of the order of 5 eV.

above / are empty, and hence when those electrons move to a higher level there is no
Therefore only those electrons close to the Fermi level can be excited, because the levels

violation of the exclusion principle. Thus only these electrons which are small fraction of
the total number - are capable of being thermally excited. The distribution function at
non-zero temperature is given by the Fermi distribution function. The Fermi distribution
function determines the probability that an orbital of energy E is occupied at thermal
equilibrium.

I : < 
D
M 2Û%ú 0Ý⁄ 3 /D
(1.6)

This function is also plotted in Fig.1.2(b), which shows that it is substantially the same as

electrons are excited from below / to above it. The quantity µ is called the chemical
the distribution at T = 0°K, except very close to the Fermi level, where some of the

electrons in the system is equal to N. At absolute zero µ = / .


potential. The chemical potential can be determined in a way that the total number of

3.3 Three – dimension


The Schrödinger equation in the three dimensions takes the form

+*:<  *:<   E ÔE *:<   E F±¡ .  ±S.  ±‰ .H *:<  *:< (1.7)


O. ð. ð. ±. ±. ±.
E

If the electrons are confined to a cube of edge L, the solution is the standing wave

*:<   sin F ›H sin F ¬H sin F ­H


ŸÌ ŸÍ ŸÎ
(1.8)
  

where ¡, S, , and ‰, are positive integers.

In many cases, however, it is convenient to introduce periodic boundary conditions, as


we did for phonons. The advantage of this description is that we assume that our crystal
is infinite and disregard the influence of the outer boundaries of the crystal on the
109
solution. We require then that our wave function is periodic in x, y, and z directions

*:›  À, ¬, ­<  *:›, ¬, ­<,


with period L, so that
(1.9)

and similarly for the y and z coordinates. The solution of the Schrödinger equation Eq.
(1.7) which satisfies these boundary conditions has the form of the traveling plane wave:

*C :<  ›J:gC. <, (1.10)

provided that the component of the wave vector k are determined from

;¡  ; ;S  ; ;‰ 
EŸÌ EŸÍ EŸÎ
  
(1.11)

where ¡, , S, and ‰, are positive or negative integers.


If we now substitute this solution to Eq. (1.7) we obtain for the energies of the orbital
with the wave vector k

C   2CE¡  CES  CE‰ 3


ð. C. ð.
E E
(1.12)

1  gðÔ this can be readily seen by differentiating (1.10):


The wave functions equations (1.10) are the eigenfunctions of the momentum

J*7 :< = gðÔ*7 :<  ðC2C :< (1.13)

The eigenvalues of the momentum is ðC . The velocity of the electron is defined by v = p


/m = ðC /m.

Therefore all the occupied states lie inside a in k space, ;/ . The energy at the surface of
In the ground state a system of N electrons occupies states with lowest possible energies.

this sphere is the Fermi energy/ . The magnitude of the wave vector ;/ and the Fermi
energy are related by the following equation:

/ 
.
ð. 73
E
(1.14)
)

of valence electrons in the system. In order to find the relationship between N and ;/ we
The Fermi energy and the Fermi wave vector (momentum) are determined by the number

need to count the total number of orbitals in a sphere of radius ;/ which should be equal
to N. There are two available spin states for a given set of ;¡ , ;S and ;‰ . The volume in
the k space which occupies this state is equal to:2 / À<… . Thus in the sphere of †2;/)3ˆ
…

2 :EŸ⁄ < = …Ÿ. C/  


sŸCA¾ ⁄… ¶ …
the total number of states is

 (1.15)

110
where the factor 2 comes from the spin degeneracy. Then

D⁄…
;/ =F H
…Ÿ. “

(1.16)

this depends only of the particle concentration. We obtain then for the Fermi energy:

E⁄…
/  F H
𠅟. “
E ¶
(1.17)

and the Fermi velocity

ð 3 E 
D⁄…
¯/   
 Ž
(1.18)

An important quantity which characterizes electronic properties of a solid is the density of

Eq.(1.17) and write the total number of orbitals of energy ø E :


states, which is the number of electronic states per unit energy range. To find it we use

E …⁄E
 : <  .F H

…Ÿ ð.
(1.19)

The density of states is then

E …⁄E
: <   F H .  D⁄E
Q“ ¶
Q EŸ. ð.
(1.20)

or equivalently

 : < 

E
(1.21)

the Fermi energy :/ <, is the total number of conduction electrons divided by the Fermi
So within a factor of the order of unity, the number of states per unit energy interval at

energy.
The density of states normalized in such a way that the integral

   :<5 (1.22)
%

111
gives the total number of electrons in the system. At non-zero temperature we should take
into account the Fermi distribution function so that
í

   : <I: <5 (1.23)


%

This expression also determines the chemical potential.

3.5 Heat capacity

The question that caused the greatest difficulty in the early development of the electron

statistical mechanics predicts that a free particle should have a heat capacity of 3)2 ; ,
theory of metals concerns the heat capacity of the conduction electrons. Classical

where ; is the Boltzmann constant. If N atoms each give one valence electron to the

heat capacity should be3)2 Nk6 , just as for the atoms of a monatomic gas. But the
electron gas and the electrons are freely mobile, then the electronic contribution to the

observed electronic contribution at room temperature is usually less than 0.01 of this
value. This discrepancy was resolved only upon the discovery of the Pauli Exclusion

zero not every electron gains an energy ~; as expected classically, but only those
Principle and the Fermi distribution function. When we heat the specimen from absolute

electrons, which have the energy within an energy range ;  of the Fermi level, can be
excited thermally. These electrons gain an energy, which is itself of the order of; , as
in Fig. 3. This gives a qualitative solution to the problem of the heat capacity of the

of ; T// can be excited thermally at temperature T, because only these lie within an
conduction electron gas. If N is the total number of electrons, only a fraction of the order

energy range of the order of ;  of the top of the energy distribution. Each of these
; ) electrons has a thermal energy of the order of ; T. The total electronic thermal
/

kinetic energy U is of the order of à ³ F; ) H ; . The electronic heat capacity is
×
‘M   ; F; ) Hand is directly proportional to T, in agreement with the
Q,
Q
/

than the classical value ≈ N; by a factor 0.01 or less, for / ~5 Œ 10s ;
experimental results discussed in the following section. At room temperature C is smaller

temperatures ;  7 / . The total energy of a system of N electrons at temperature T is


We now derive a quantitative expression for the electronic heat capacity valid at low

à   : <I:,  <5 (1.24)


%
Where f (E, T) is the Fermi distribution function and D (E) is the density of states. The
heat capacity can be found by differentiating this equation with respect to temperature.
í
5Ã 5I:, <
Since only the distribution function depends on temperature we obtain
‘M    :< 5
5 5 (1.25)
%
It is more convenient to represent this result in a different form:

112
í
5I:, <
‘M   :  / <: < 5
5
(1.26)

%
Eq. (1.26) is equivalent to Eq. (1.25) due to the fact which follows from Eq. (1.22):
í
5 5I:, <
0  /  /  :< 5
5 5
(1.27)
%

Since we are interested only temperatures for which ;  7 / the derivative 5I⁄5 is
large only at the energies which lie very close to the Fermi energy. Therefore, we can
ignore the variation of D (E) under the integral and take it outside the integrand at the
Fermi energy, so that
í
5I:, <
‘M   :/ <  :  / < 5
5
(1.28)
%

that 8  / , which is good approximation at room temperature and below. Then


We also ignore the variation of the chemical potential with temperature and assume


9::;,<< ;e ;= @ô2Aú A=3)B? Cõ
>? <. @ô2AúA=3)B? Cõ .
(1.29)
9<

Eq. (1.28) can then be rewritten as

‘M  :/ < % 5  :/ < – 5›


í : e 3 <. M 2%ú %3 ⁄ 3 í ¡ . :7
<  MÌ
3 )7
:M Ì /D<.
.
7
. M 2%ú %3 3)  /D 7
.
(1.30)

Taking into account that / >>; T, we can put the low integration limit to minus infinity
and obtain

‘M  :/ <;E  eí :M Ì 5›   :/ <;E 


í ¡ .M Ì Ÿ.
/D< . … (1.31)

‘M  ; ) ,
Ÿ.
For a free electron gas we should use Eq. (1.21) for the density of states to finally obtain

E
(1.32)
3

where we defined the Fermi temperature / 


3
7
. This is similar to what we expected to
obtain according to the qualitative arguments given in the beginning of this section.
Experimentally the heat capacity at temperatures much below both the Debye
temperature and the Fermi temperature can be represented in the form:

113
‘  ‘M  ‘O6     … (1.33)

The electronic term is dominant at sufficiently low temperatures. The constants α and
β can be obtained by fitting the experimental data.

4.0 Conclusion
The classical free electron theory is an attempt to regard the valence electrons in metal as
the non-interacting particles of an ideal gas. The only difference between this gas of
electrons and any other ideal gas defined by kinetic theory is that the particles are
charged.

5.0 Summary
• The energy of the highest occupied level is called the Fermi energy
• Various electronics states of the crystals can be obtained through the application
of Schrodinger’s wave equation.
• The total energy of a system of N electrons at temperature T is
í

U   ED:E<f:E, T<dE
%

6.0 Tutor marked assignment

Q1. Consider the free electron energy bands of an fcc crystal lattice in the
reduced zone scheme in which all k's are transformed to lie in the first
Brillouin zone. Plot roughly in the [111] direction the energies of all bands
up to six times the lowest band energy at the zone boundary at k =
(2π/a)(½,½,½). Explain what happens with these bands in the presence of
a weak crystal potential.

constant * is composed of a series of rectangular wells which surround the


Q2. Suppose that the crystal potential in a one-dimensional lattice of lattice

atom. Suppose that the depth of each well is U and its width a/5.
0
a. Calculate the values of the first three energy gaps. Compare the

b. Evaluate these gaps for the case of U = 5 Ž and a = 4Å.


magnitudes of these gaps.
0

7.0 Further reading/ References

Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8

114
UNIT 2: ELECRONIC TRANSFER page
1.0 Introduction - - - - - - - - 110
2.0 Objectives - - - - - - - - 110
3.0 Definition - - - - - - - - 110
3.1 Drude model - - - - - - - 110
3.2 The origin of collision time - - - - - 113
3.3 Thermal conductivity - - - - - - 116
3.4. Motion in a magnetic field - - - - - 119
3.4.1 Cyclotron resonance - - - - - - 119
3.4.2 The Hall Effect - - - - - - 119
4.0 Conclusion - - - - - - - - 122
5.0 Summary - - - - - - - - 122
6.0 Tutor Marked Assignment - - - - - - 123
7.0 Further Reading/References - - - - - - 123

115
1.0 Introduction
In this unit we are going to study how the classical free electron theory developed by
Lorentz, Drude and Debye uses kinetic theory to calculate the transport properties of the
free electron of a gas including electrical and thermal conductivity.
2.0 Objective
• To explain the Drude model of the thermal conductivity of solid
• To explain motion in Magnetic field in terms of Cyclotron resonance and Hall Effect

3.0 Definition
Electronic transfer is the determination of the thermal conductivity of electrons treated as
classical particles.

3.1 Drude model

The simplest treatment of the electrical conductivity was given by Drude. There are four
major assumptions within the Drude model.
i. Electrons are treated as classical particles within a free-electron approximation.
Thus, in the absence of external electromagnetic fields each electron is taken to
move uniformly in a straight line, neglecting the interactions with other electrons
and ions. In the presence of external fields each electron is taken to move
according to Newton's laws of motion.
ii. Electrons move free only between collisions with scattering centers. Collisions, as
in kinetic theory, are instantaneous events that abruptly alter the velocity of an
electron. Drude attributed them to the electrons scattering by ion cores. However,
as we will see later, this is not a correct picture of electron scattering on ordered
periodic structures. A particular type of scattering centers does not matter in the
Drude model. An understanding of metallic conduction can be achieved by simply
assuming that there is some scattering mechanism, without inquiring too closely
into just what that mechanism might be.
iii. An electron experiences a collision, resulting in an abrupt change in its velocity,
with a probability per unit time1/'. This implies that the probability of an
electron undergoing a collision in any infinitesimal time interval of length 5Y is
just 5Y/'. The time ' is therefore an average time between the two consecutive
scattering events. It is known as, the collision time (relaxation time), it plays a
fundamental role in the theory of metallic conduction. It follows from this
assumption that an electron picked at random at a given moment will, on the
average, travel for a time t before its next collision. The relaxation time t is taken
to be independent of an electron's position and velocity.
iv. Electrons are assumed to achieve thermal equilibrium with their surroundings
only through collisions. These collisions are assumed to maintain local thermo-
dynamic equilibrium in a particularly simple way: immediately after each
collision an electron is taken to emerge with a velocity that is not related to its
velocity just before the collision, but randomly directed and with a speed
appropriate to the temperature prevailing at the place where the collision
occurred.

116
Now we consider the application of the Drude model for electrical conductivity in a
metal. According to Ohm's law, the current I flowing in a wire (Fig 2.1) is proportional
to the potential drop V=V2−V1 along the wire: V = IR, where R, the resistance of the
wire, depends on its dimensions. It is much more convenient to express the Ohm's law in
a form which is independent of the dimensions of the wire because these factors are
irrelevant to the basic physics of the conduction We define the conductivity which is the
proportionality constant between the current density j and the electric field E at a point
in the metal:

Fig. 2.1: Current flowing in a wire (After


www.pa.uk.edu/kwng.phy/525/lec/lecture-8)

B  áK (2.1)

The current density j is a vector, parallel to the flow of charge, whose magnitude is the
amount of charge per unit time crossing a unit area perpendicular to the flow. Thus if a

density will be B  â⁄ Since the potential drop along the wire will be Ž  À Eq. (2.1)
uniform current I flows through a wire of length L and cross-sectional area A, the current

Unlike R, σ and ’ is a property of the material, since it does not depend on the shape and
gives I/A=σV/L, and hence R = L/σA =ρL/A, here we have introduced resistivity ρ=1/σ.

size. Now we want to express σ is terms of the microscopic properties using the Drude

they give rise to will be parallel to v. Furthermore, in a time 5Y the electrons will advance
model. If n electrons per unit volume all move with velocity v, then the current density

by a distance ¯5Y in the direction of v, so that :¯5Y< electrons will cross an area A

crossing A in the time 5Y will be – ¯5Y and hence the current density is
perpendicular to the direction of flow. Since each electron carries a charge -e, the charge

B  L. (2.2)

At any point in a metal, electrons are always moving in a variety of directions with a
variety of thermal energies. The net current density is thus given by Eq. (2.2), where v is
the average electronic velocity or drift velocity. In the absence of an electric field,
electrons are as likely to be moving in any one direction as in any other, v averages to
zero, and, as expected, there is no net electric current density. In the presence of a field
E, however, there will be a drift velocity directed opposite to the field (the electronic
charge being negative), which we can compute as follows. Consider a typical electron at

be its velocity Ž% immediately after that collision plus the additional velocity  KY) it
time zero. Let t be the time elapsed since its last collision. Its velocity at time zero will

has subsequently acquired. Since we assume that an electron emerges from a collision in
a random direction, there will be no contribution from Lz to the average electronic

117
velocity, which must therefore be given entirely by the average of However,
the average of t is the relaxation time τ. Therefore
(2.3)

(2.4)
E

The conductivity is, therefore, given by


(2.5)

We see that the conductivity is proportional to the density of electrons, which is not
surprising since the higher the number of carriers, the more the current density. The
conductivity is inversely proportional to the mass because the mass determine the
acceleration of an electron in electric field. The proportionality to τ follows because τ is
the time between two consecutive collisions. Therefore, the larger τ is, the more time for
electron to be accelerated between the collisions and consequently the larger the drift
velocity. The values of relaxation time can be obtained from the measured values of
electrical conductivity. For example at room temperature the resistivity of many metals
lies in the range of 1-10 µΩcm. The corresponding relaxation time is of the order
of . In this discussion of electrical conductivity we treated electrons on a
classical basis. How are the results modified when the quantum mechanics is taken into
account? Let us refer to Fig.2.3. In the absence of an electric field, the Fermi sphere is
centred at the origin (Fig. 2.3a). The various electrons are all moving - some at very high
speeds - and they carry individual currents. But the total current of the system is zero,
because, for every electron at velocity v there exists another electron with velocity -v and
the sum of their two currents is zero. Thus the total current vanishes due to pair wise
cancellation of the electron currents.

Fig.2.2: (a) The Fermi sphere at equilibrium, (b) Displacement of the


Fermi sphere due to an electric field (After
www.pa.uk.edu/kwng.phy/525/lec/lecture-8)

118
The situation changes when a field is applied. If the field is in the positive x-direction,
each electron acquires a drift velocity, as given by Eq. (2.2). Thus the whole Fermi
sphere is displaced to the left, as shown in Fig.2.2 (b). Although the displacement is very
small and although the great majority of the electrons still cancel each other pair wise,
some electrons - in the shaded crescent in the figure -remain uncompensated. It is these
electrons which produce the observed current. The very small displacement is due to a
relatively small drift velocity. If we assume that the electric field is 0.1V/cm, we obtain
the drift velocity of 1cm/s, which is by 8, orders in magnitude smaller the Fermi velocity
of electrons.
Let us estimate the current density. The fraction of electrons which remain
uncompensated is approximately ¯⁄¯/ . The concentration of these electrons is therefore
:¯ ⁄¯/ <and since each electron has a velocity of approximately¯/ , the current density is
given by

B   :¯⁄¯/ <Ž/   ¯ (2.6)

This is the same expression we obtained before. Therefore, formally the conductivity is
expressed by the same formula (2.5). However, the actual picture of electrical conduction
is thus quite different from the classical one. In the classical picture, we assumed that the
current is carried equally by all electrons, each moving with a very small drift velocity v.
In the quantum-mechanical picture the current is carried only by very small fraction of
electrons, all moving with the Fermi velocity. The relaxation time is determined only by
electrons at the Fermi surface, because only these electrons can contribute to the transport
properties. Both approaches lead to the same result, but the latter is conceptually the more
accurate. Since only electrons at the Fermi surface contribute to the conductance, we can
define the mean free path of electrons as  M¯/ . We can make an estimate of the mean
free path for metal at room temperature. This estimate gives a value of 100Å. So it is of
the order of a few tens inter atomic distances. At low temperatures for very pure metals
the mean free path can be made as high as a few cm.

3.2 The origin of collision time


We see that between two collisions, the electron travels a distance of more than 20 times
the inter atomic distance. This is much larger than one would expect if the electron
really did collide with the ions whenever it passed them. This paradox can be explained
only using quantum concepts according to which an electron has a wave character. It is
well known from the theory of wave propagation in periodic structures that, when a
wave passes through a periodic lattice, it continues propagating indefinitely without
scattering. The effect of the atoms in the lattice is to absorb energy from the wave and
radiate it back, so that the net result is that the wave continues without modification in
either direction or intensity. Therefore we see that, if the ions form a perfect lattice, there
is no collision at all - that is, l = ∞ - and hence τ = ∞, which in turn leads to infinite
conductivity. It has been shown, however, that the observed l is about 10E A. The
finiteness of σ must thus be due to the deviation of the lattice from perfect periodicity;
this happens either because of (1) thermal vibration of the ions, or because of (2) the
presence of imperfections or foreign impurities.

119
In order to consider their contribution we examine the temperature dependence of the
electrical conductivity. The electrical conductivity of a metal varies with temperature in a
characteristic manner. This variation is usually discussed in terms of the behavior of the
resistivity ρ versus T. Figure 2.3 shows the observed curve for Na. At T ~ 0°K, ρ has a
small constant value; above that, ρ increases with T, slowly at first, but afterward
ρ increases linearly with T. The linear behavior continues essentially until the melting
point is reached. This pattern is followed by most metals, and usually room temperature
falls into the linear range.

Fig. 2.3 The normalized resistivityρ (T)/ρ (290°K) versus T for Na in the
low-temperature region (a), and at higher temperatures (b) (After Kittel,
1979)

We want to explain this behavior in terms of the Drude formula. Recalling that
we have
(2.7)

As we have discussed earlier 1/τ which enters equation (2.7), is the probability of the
electron scattering per unit time. Thus, if , then the electron undergoes
collisions in one second. We found that the electron undergoes collisions only because
the lattice is not perfectly regular. We group the deviations from a perfect lattice into
two classes. a) Lattice vibrations (phonons) of the ions around their equilibrium position
due to thermal excitation of the ions. (b) All static imperfections, such as impurities or
crystal defects. Of this latter group we shall take impurities as an example. The total
probability for an electron to be scattered in a unit time is the sum of the probabilities of
scattering by phonons and by impurities. This is because these two mechanisms are
assumed to act independently. Therefore we may write

(2.8)

120
Where the first term on the right is due to impurities and the second is due to phonons.
The scattering by impurities is essentially independent of temperature, whereas the
scattering by phonons is temperature dependent because the number of phonons increases
with temperature. When equation (2.8) is substituted into equation (2.7), we readily find
 
’  ’P  ’O6  
M . Næ M . NOP (2.9)

We see that ρ has split into two terms. A term ’P due to scattering by impurities, which is
independent of T, is called the residual resistivity. Another term ’O6 :<is due to
scattering by phonons; hence it is temperature dependent. Sometimes it is called the
lattice resistivity.
At very low T, scattering by phonons is negligible because the amplitudes of oscillation
are very small; in that region MO6 Q ∞, ’O6 Q 0and hence ’  ’P is a constant. This is in
agreement with Fig.2.3. As T increases, scattering by phonons becomes more effective,
and ’J:<,) increases; this is why ρ increases. When T becomes sufficiently large,
scattering by phonons dominates and ρ ~’J:<. The statement that ρ can be split into
two parts, is known as the Matthiessen rule. This rule is embodied in (2.9). In general, the
Matthiessen rule predicts that if there are two distinguishable sources of scattering (like
in the case above – phonons and impurities) the resistivity is the sum of the resistivities
due to the first and the second mechanism of scattering. The Matthiessen rule is sort of
empirical observation which can be used for a qualitative understanding of the
contribution from different scattering mechanisms. However, one must always bear in
mind the possibility a failure of this rule. In particular, in the case when the relaxation
time depends on the wave vector k, the Matthiessen rule becomes invalid.
Now let us derive approximate expressions for MP and MO6 using arguments from the
kinetic theory of gases. Consider first the collision of electrons with impurities. We write

æ
MP  (2.10)
03

Where P is the mean free path for collision with impurities. In order to find the mean free
path we shall assume, for simplicity, that the collision is of the hard-spheres (billiard-
ball) type and introduce the scattering cross section of an impurity Σi which is the area an
impurity atom presents to the incident electron. Then, we can argue that the product of
the mean free path and the cross section of impurity liΣi , is equal to the average volume
per impurity,1) P , where P is the impurity concentration, i.e.

D
P ∑P  æ
(2.11)

and therefore

P 
D
æ ST
(2.12)

121
The scattering cross section Σi is of the same magnitude as the actual geometrical area of
the impurity atom. That is, Σi ~ lÅ2. Calculations of the exact value of Σi require quantum
scattering theory. By substituting Eqs. (2.12) and Eqs. (2.10) into (2.9), we find
03
’P  P ΣP
M . (2.13)

As expected, ’P is proportional to P the concentration of impurities. Calculating ’O6 is


much more difficult, but equations similar to (2.10) and (2.12) still hold. In particular,
one may write
1
âO6 
(2.14)
8 Σ8

where 8 is the concentration of the host atoms in the lattice, and Σa is the scattering cross
section per atom. We should note here that Σa has no relation to the geometrical cross
section of the atom. Rather it is the area presented by the thermally fluctuating atom to
the passing electron. Suppose that the distance of deviation from equilibrium is x, then
the average scattering cross section is

ì Þ Û› E Ý (2.15)
8

where ۛ E Ý is the average of › E . We can easily estimate this value at high temperatures,
when the classical approach is valid. Since the ion is a harmonic oscillator, the value ۛ E Ý
is proportional to the average of its potential energy is equal to half the total energy.
Thus,
∑8 Þ Û› E Ý Þ 
7
E” (2.16)
where C is inter atomic force constant introduced earlier and we used the formula for the
energy of a classical oscillator. We see therefore that at high temperatures the resistivity
is linear in T,

¯/ 8 ;

(2.17)
’O6 Þ
 E 2‘
which is in agreement with experiment.
In the low-temperature range the lattice resistivity varies with temperature in a different
way. Using the Debye model at low temperature range one can find that ’O6 ~ r .

3.3 Thermal conductivity


When the ends of a metallic wire are at different temperatures, heat flows from the hot to
the cold end. The basic experimental fact is that the heat current density, & i.e. the
amount of thermal energy crossing a unit area per unit time is proportional to the
temperature gradient

122
( 2.18)

where K is the thermal conductivity. In insulators, heat is carried entirely by phonons, but
in metals heat may be transported by both electrons and phonons. The thermal
conductivity K is therefore equal to the sum of the two contributions

(2.19)

where and refer to electrons and phonons, respectively. In most metals, the
contribution of the electrons greatly exceeds that of the phonons, because of the great
concentration of electrons. Typically

Fig.2.4: Heat conduction process (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8)

The physical process by which heat conduction takes place via electrons is illustrated in
Fig.2.4. Electrons at the hot end (to the left) travel in all directions, but a certain fraction
travel to the right and carry energy to the cold end. Similarly, a certain fraction of the
electrons at the cold end (on the right) travel to the left, and carry energy to the hot end.
Since on the average electrons at the hot end are more energetic than those on the right, a
net energy is transported to the right, resulting in a current of heat. Note that heat is
transported entirely by electrons having the Fermi energy, because those well below this
energy cancel each other's contributions.
To evaluate the thermal conductivity K quantitatively, we use the formula
here is the electronic specific heat per unit volume, v is the Fermi velocity
of electrons; l is the mean free path of electrons at the Fermi energy. Using expression for
the heat capacity derived earlier, we find

(2.20)

Noting that and that we can simplify this expression for K to


(2.21)

This expresses thermal conductivity in terms of the electronic properties of the metal.
Many of the parameters appearing in the expression for K were also included in the
expression for electrical conductivityσ. Recalling that we find

123
D Ÿ7 E
 F H   À
V
… M
(2.22)
à

We see from here that the ratio of the thermal conductivity to the electrical conductivity
is directly proportional to the temperature. This is called the Wiedemann-Franz law. The

particular metal. It depends only on the universal constants ; and e, should be the same
constant of proportionality L, which is called the Lorentz number, is independent of the

for all metals. The Lorentz number numerical value is2.45 Œ 10ef ½Ω/£ E . This
conclusion suggests that the electrical and thermal conductivities are intimately related,
which is to be expected, since both electrical and thermal current are carried by the same
agent: electrons.

Solid  has an IGG structure with cubic lattice constant *  5.26k, atomic mass
Worked example:

XÚ  6.67 Œ 10eEÄ kg and a Debye temperature +  92% £.

a) Estimate the phonon velocity using the Young modulus of , ‘DD  1.6 Œ
10 /E .
b) Using the expression £  ‘¯ in which C is the phonon heat capacity per unit
D
…
volume. Find the thermal conductivity, K (in unit of eD X eD £ eD ) of a 1 …
crystal of  at 1% £, assuming that phonon scattering occurs only at the
boundaries of the sample.

Solution:
a) The phonon velocity is estimated from the velocity of sound which is
‘ *… ‘DD)
¯  q DD)’ =q 4XÚ since in IGG structure there are 4 atoms in a cubic

unit cell and hence, ’  8  


 s
)s 8 
, then,

¯  q5. 26 Œ 10
e…% Œ 1.6 Œ 10 E
)4 Œ 6.67 Œ 10eEÄ ;•… Y 934 X eD
…

b) Since   1% £ 7 +  92% £, we can use the low temperature approximation


for heat capacity. Recall Eq.(5.37) in Module 2, unit 5, the heat capacity of a solid

12 E  …
which contains N atoms is given,

‘0  ; † ˆ
5 +

dividing the latter by * )4, we obtain the heat capacity per unit volume, C.
…
Dividing the expression by N, we obtain the heat capacity per unit atom and

Therefore,

124
48 E ;  … 48 Œ 3.143 Œ 1.38 Œ 1023 1 3  
‘0  † ˆ  † ˆ )3 £ Y 1.14 Œ 102 )3 £
5* … + 3
5 Œ 5.26 Œ 10 30
92
Since the scattering of phonons is determined by the boundaries of the sample we can
assume that the mean free path is l= 1 mm and the thermal conductivity is
1  
£  ‘0 ¯  0.33 Œ 1.14 Œ 10E 934 Œ 10e… )… £ ⁄X  n 35 )X£
3
3.4 Motion in a magnetic field
The application of a magnetic field to a metal gives rise to several interesting phenomena
due to conduction electrons. The cyclotron resonance and the Hall Effect are to be
considered

3.4.1 Cyclotron resonance


If a magnetic field is applied to a metal the Lorentz force F = −e[E+(v × B)] acts on each
electron. For a perfect metal in the absence of electric field the equation of motion takes
the form
ÜL
 ÜZ -evŒ ˜ (2.23)

If the magnetic field lies along the z-direction this results in

Q0Ì
 ÐN ¯S, (2.24)
QT

5¯S
 ÐN ¯¡
5Y

where

˜ (2.25)
ÐN 

is the cyclotron frequency in SI system of units :g ‘î} ÐG   )G <. For magnetic
fields of the order of a few ;î the cyclotron frequencies lie in the range of a few GHz.
For example for B=1kG, the cyclotron frequency is ¯G  ÐGŠ  2.8î+­. Therefore, the
2
magnetic field causes electrons to move in a counterclockwise circular fashion with the
cyclotron frequency in a plane normal to the field.
Suppose now that an electromagnetic signal is passed through the slab in a direction
parallel to B, as shown in figure 2.5. The electric field of the signal acts on the electrons,
and some of the energy in the signal is absorbed. The rate of absorption is greatest when
the frequency of the signal is exactly equal to the frequency of the cyclotron (see
Fig.2.5b), i.e.

Ð  ÐN (2.26)

125
Fig. 2.5 (a) Cyclotron motion, (b) The absorption coefficient versus
ω(After www.pa.uk.edu/kwng.phy/525/lec/lecture-8)

This is so because, when this condition holds true, each electron moves with the wave
throughout the cycle, and therefore the absorption continues all through the cycle. Thus,
Eq. (2.26) is the condition for cyclotron resonance. On the other hand, when Eq. (2.26) is
not satisfied, the electron is in phase with the wave through only a part of the cycle, during
which time it absorbs energy from the wave. In the remainder of the cycle, the electron is
out of phase and returns energy to the wave. Cyclotron resonance is commonly used to
measure the electron mass in metals and semiconductors. The cyclotron frequency is
determined from the absorption curve, and this value is then substituted in Eqs. (2.25) to
evaluate the effective mass.

3.4.2 Hall effect

First we derive an equation of motion of an electron in applied magnetic and electric field
in the presence of scattering. Assume that that the momentum of an electron is at
time t, let us calculate the momentum per electron an infinitesimal time
later. An electron taken at random at time t will have a collision before time with
probability and will therefore survive to time without suffering a collision
with probability If it experiences no collision, however, it simply evolves under
the influence of the force F (due to the spatially uniform electric and/or magnetic fields)
and will therefore acquire an additional momentum . The contribution of all those
electrons that do not collide between t and to the momentum per electron at time
is the fraction they constitute of all electrons, times their average
momentum per electron . Thus, neglecting the moment the contribution to
from those electrons that do undergo a collision in the time between t and
, we have

(2.27)

Note that if the force is not the same for every electron it should be averaged.

The correction to (2.27) due to those electrons that have had a collision in the interval t to
is only of the order of . To see this, first note that such electrons constitute a
fraction /τ of the total number of electrons. Furthermore, since the electronic velocity
126
(and momentum) is randomly directed immediately after a collision, each such electron
will contribute to the average momentum only to the extent that it has
acquired momentum from the force F since its last collision. Such momentum is acquired
over a time no longer than and is therefore of order . Thus the correction to (2.27)
is of order , and does not affect the terms of linear order in We may
therefore write

(2.28)

This simply states that the effect of individual electron collisions is to introduce a
damping term into the equation of motion for the momentum per electron. We apply this
equation to discuss the Hall Effect in metals using a free electron model. The physical
process underlying the Hall Effect is illustrated in Fig.2.6. Suppose that an electric
current is flowing in a wire in the x-direction, and a magnetic field is applied normal
to the wire in the z-direction. We shall show that this leads to an additional electric field,
normal to both and , that is, in the y-direction. Before the magnetic field is applied,
there is an electric current flowing in the positive x direction, which means that the
conduction electrons are drifting with a velocity v in the negative x-direction. When the
magnetic field is applied, the Lorentz force causes the electrons to bend
downward, as shown in the figure. As a result, electrons accumulate on the lower surface,
producing a net negative charge there. Simultaneously a net positive charge appears on
the upper surface, because of the deficiency of electrons there. This combination of
positive and negative surface charges creates a downward electric field , which is
called the Hall field.

Fig. 2.6: Origin of the Hall field and Hall Effect


(After www.pa.uk.edu/kwng.phy/525/lec/lecture-8)

Let us evaluate this Hall field. We start from the Lorentz force acting on each electron F
= −e [E+ (v ×B)]. According to (2.28) we find

(2.29)

127
where τ is the relaxation time. Note that the Lorentz force is not the same for all electrons
because they move with different velocities; therefore it is averaged over ensemble. We
are looking for the solution of this equation in the steady state when the current is
independent of time and therefore 5L)5Y = 0.


0  ¡   vS  
M
(2.30)

0  S   v¡  
M

We multiply these equations by −neτ/m to introduce current densities components


&¡   ¯¡ and &S   ¯S , so that

σE^  ω` ajc  j^
(2.31)
σES  ÐN M&¡  &S

Where σ is the Drude conductivity in the absence of a magnetic field. In the steady state
there is no electric current flowing perpendicular to the wire. Therefore the Hall field d
=e can be determined by the requirement that there be no transverse current&¡ .
Setting &¡ to zero in the second equation of (2.31) we find that

S   F H &¡  
Ïf M D
 ¡
&
à
(2.32)

The proportionality constant 1)ne , is known as the Hall constant, and is usually
denoted by Ød .. Therefore,

Ød  
D
M
(2.33)

parameters of the metal except the density of carriers. Since Ød is inversely proportional
This is a very striking result, which predicts that the Hall coefficient depends on no

to the electron concentration n, it follows that we can determine n by measuring the Hall
field. Since we have already calculated n assuming that the atomic valence electrons
become the metallic conduction electrons, a measurement of the Hall constant provides a
direct test of the validity of this assumption.

4.0 Conclusion
The electrical and thermal conductivity of the free electron were obtained through the
Drude model.

5.0 Summary
• Drude model provided the simplest treatment of electrical conduction of a metal

128
• The splitting up of resistivity to two terms (due to impurities and phonon) is
known as Matthiessen rule
• Resistivity (:’O6 < due to scattering of phonons which is independent of
temperature is known as lattice resistivity
• Resistivity :’P ) due to scattering by impurities which is independent of
temperature is known as residual resistivity
• The cyclotron resonance and the Hall Effect are phenomena due to application of
a magnetic field to a metal.

6.0 Tutor marked assignment


Q1. A Cu wire of diameter 2mm carries 10A of current. Find the drift velocity
Q2. If the Fermi energy of Na is 3.1 eV and the electrical conductivity is
2.1x1017 esu at 0K, calculate the relaxation time.
Q3. Using the Drude formula, calculate the mean free path of K, if its lattice
parameter a = 4.2Å. Also calculate the Hall coefficient.

7.0 Further reading/References


Animalu, A.O.E, intermediate quantum theory of crystalline solids, Prentice-Hall
of India, New Delhi, 1978
Blakemore ,J.S., Solid State Physics, W.B. Saunders Co.,1974
C. Kittel, Introduction to solid state physics, Wiley Eastern Limited, 1979
Hurd, C.M., The hall effect in metals and alloys, Plenum,1972
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8

129
UNIT 3: ENERGY BAND THEORY page
1.0 Introduction - - - - - - - - 124
2.0 Objectives- - - - - - - - - 124
3.0 Definition- - - - - - - - - 124
3.1 Energy bands - - - - - - - 124
3.2 Periodic potential- - - - - - - 126
3.3 Weak potential- - - - - - - 131
3.4. Metal and Insulators - - - - - - 138
4.0 Conclusion - - - - - - - - 139
5.0 Summary - - - - - - - - 139
6.0 Tutor Marked Assignment - - - - - - 140
7.0 Further Reading/References- - - - - - - 140

130
1.0 Introduction

The free electron model gives us a good insight into many properties of metals, such as
the heat capacity, thermal conductivity and electrical conductivity. However, this model
fails to help us with other important properties. For example, it does not predict the
difference between metals, semiconductors and insulators. It does not explain the
occurrence of positive values of the Hall coefficient. Also the relation between
conduction electrons in the metal and the number of valence electrons in free atoms is not
always correct. We need a more accurate theory, which would be able to answer these
questions.

2.0 Objective
The objectives of this unit is
• To explain the general features of band levels
• To explain the periodic potential of an electron
• To explain the properties of the Bloch electron
• To explain the difference between Metals and Insulators.

3.0 Definition
Energy band is the range of energies possessed by electrons in a solid
3.1 Energy band
It is customary to visualize the existence of bands on an energy scale of band structure
scheme, according to which, the energy bands for the most tightly bound electrons lie at
the bottom, followed by the band of the second most tightly bound electrons, and so on,
till we reach the top of the set of completely full energy bands. The top of the band of the
set is known as the valence band. Next higher energy band is referred to as conduction
band, which might be completely empty. The characteristic energy that separate the
occupied from empty states is called Fermi energy EF and is characterized by Fermi level
existing between the conduction band and the valence band. The two bands are separated
by energy gap ù , defined by

ù  N  0 (3.1)

Where N and 0 are respectively the energy of the bottom of the conduction band and
the top of the valence band. The ù value for a semiconductor is typically of the order of
1 Ž and that for an insulator is 5 Ž. based o the relative positions of conduction and
valence bands, metals may be classified into two categories. In one, valence band is
completely full and conduction band is partially filled, e.g., Na, 2p (valence) band is
completely full and conduction (3s) band is half filled. In the other, conduction and
valence bands overlap each other. For example, Mg:1X E , 2X E , 2JÄ , 3X E <, 3X E (valence)
and 3J(conduction) bands overlap in energy.

131
3.2 Periodic Potential
The potential seen by an electron due to the nucleus of an isolated atom of valence z is
­ E) , where e is the electronic charge and r the nucleus –electron distance. However ,


the atom in a perfect crystal are arranged in a regular periodic array, therefore, we are led
to consider the problem of an electron in a potential U(r) with the periodicity of the
under-lying Bravais lattice i.e.

U(r) = U(r + T) (3.2)

where T is a lattice vector. Qualitatively, a typical crystalline potential might be expected


to have a form shown in Fig.3.1, resembling the individual atomic potentials as the ion is
approached closely and flattening off in the region between ions.

Fig.3.1: The crystal potential seen by the electron (After Kittel, 1979)

Since the scale of periodicity of the potential U (~ 10-8 cm) is the size of a typical de
Broglie wavelength of an electron, it is essential to use quantum mechanics in accounting
for the effect of periodicity on electronic motion. Thus we consider the Hamiltonian.

ð.
+ :<   E h E  Ã:< (3.3)

Using Eq. (3.2) in Eq. (3.3) leads to

+ :  i<  +:< (3.4)

This shows that the Hamiltonian also has the lattice periodicity. Hence, to predict the
physical properties of the crystal, one should solve the following Schrodinger equation
for a single electron

+ :2<   h E  Ã:< *:<  *:<


ð.
E
(3.5)

in which ψ (r) is a wave function for one electron. Independent electrons, which obey a
one electron Schrödinger equation (3.5) with a periodic potential, are known as Bloch
electrons, in contrast to "free electrons," to which Bloch electrons reduce when the
periodic potential is identically zero.
Now we discuss general properties of the solution of the Schrödinger equation (3.5)
taking into account periodicity of the effective potential (3.2) and discuss main properties
132
of Bloch electrons, which follow from this solution. We represent the solution as an
expansion over plain waves.

*:<  ì G7  PC (3.6)


>

This expansion in a Fourier series is a natural generalization of the free-electron solution


for a zero potential. The summation in (3.6) is performed over all k vectors, which are
permitted by the periodic boundary conditions. According to these conditions the wave
function (3.6) should satisfy

*:›, ¬, ­<  *:›  À, ¬, ­<  *:›, ¬  À, ­<  *:›, ¬, ­  À< (3.7)

k¡  ; kS  ; k‰ 
So that
EŸÌ EŸÍ EŸÎ
  
(3.8)

where ¡ , S , and ‰ are positive or negative integers. Note that in general ψ (r) is not
periodic in the lattice translation vectors. On the other hand, according to Eq. (3.2) the
potential energy is periodic, i.e. it is invariant under a crystal lattice translation.
Therefore, its plane wave expansion will only contain plane waves with the periodicity of
the lattice. Therefore, only reciprocal lattice vectors are left in the Fourier expansion for
the potential:

Ã:<  ì Ý  PV (3.9)




where the Fourier coefficients UG are related to U(r) by

NM  ePV Ã:<5


D
Ý 
¶f (3.10)

where ŽN is the volume of the unit cell. It is easy to see that indeed the potential energy
represented by (3.9) is periodic in the lattice:

U:j  k<  ì Ul eml:n/<<  eml< ì Ul emln  U:j<


(3.11)
l
where the last equation comes from the definition of the reciprocal lattice vectors  Po

l

1. The values of Fourier components Ão for actual crystal potentials tend to decrease
rapidly with increasing magnitude of G. For example, for a Coulomb potential Ão
decreases as 1)î E .Note that since the potential energy is real the Fourier components
should satisfy Ãeo  Ãop .
We now substitute (3.6) and (3.9) in Eq. (3.5) and obtain:

133
ðE
ì ; E ‘7  P7V  ì ì Ão ‘7  P:7/o<V   ì ‘7  P7V
2
(3.12)
7 7 o V

changing the summation index in the second sum on the left from k to k +G this equation
can be rewritten in a form:

; E  H ‘7  ∑o Ý ‘Ce q  0
ð.
∑r  PCV ,F
E
(3.13)

Since this equation must be satisfied for any r the Fourier coefficients in each separate
term of (3.13) must vanish and therefore

ðE E
;   ‘7  ì Ão ‘7eo  0
2

(3.14)
o

This is a set of linear equations for the coefficients Ck. These equations are nothing but
restatement of the original Schrödinger equation in the momentum space, simplified by
the fact that the potential is periodic. This set of equations does not look very pleasant
because, in principle, an infinite number of coefficients should be determined. However,
a careful examination of Eq. (3.14) leads to important consequences.
First, we see that for a fixed value of k the set of equations (3.14) couples only those
coefficients, whose wave vectors differ from k by a reciprocal lattice vector. In the one-
dimensional case these are k, k±2π/a, k±4π/a, and so on. We can therefore assume that
the k vector belongs to the first Brillouin zone. The original problem is decoupled to N
independent problems (N is the total number of atoms in a lattice): for each allowed value
of k in the first Brillouin zone. Each such problem has solutions that are superposition of
plane waves containing only the wave vector k and wave vectors differing from k by the
reciprocal lattice vector.
Putting this information back into the expansion (3.6) of the wave function ψ (r), we see that
the wave function will be of the form

*C :<  ì ‘7eo  P:Ce< (3.15)




where the summation is performed over the reciprocal lattice vectors and we introduced
index k for the wave function. We can rearrange this so that

*7 :<   PC ì ‘Ce  eP


(3.16)


*C :<   PC ZC :<


Or
(3.17)

where Z7 :<  Z7 :  i< is a periodic function which is defined by

ZC :<  ì ‘Ce  eP (3.18)



134
Equation (3.17) is known as Bloch theorem, which plays an important role in electronic
band structure theory. Now we discuss a number of important conclusions which follow
from the Bloch theorem.

1. Bloch's theorem introduces a wave vector k, which plays the same fundamental
role in the general problem of motion in a periodic potential that the free electron
wave vector k plays in the free-electron theory. Note, however, that although the
free electron wave vector is simply _)ð, where p is the momentum of the electron,
in the Bloch case k is not proportional to the electronic momentum. This is clear
on general grounds, since the Hamiltonian does not have complete translational
invariance in the presence of a non-constant potential, and therefore its eigenstates
will not be simultaneous eigenstates of the momentum operator. This conclusion
is confirmed by the fact that the momentum operator, _  >ðh, when acting
on Ð7 :< gives

gðÔ*7 :<  gðÔô PCZC :<õ  ðC*7 :<  gð PCÔZC :< (3.19)

Which is not, in general, just a constant time Ð7 :<; i.e., Ð7 :< is not a
momentum eigenstate. Nevertheless, in many ways ðk is a natural extension of p
to the case of a periodic potential. It is known as the crystal momentum or
quasimomentum of the electron, to emphasize this similarity, but one should not
be misled by the name into thinking that ðk is a momentum.

2 The wave vector k appearing in Bloch's theorem can always be confined to the
first Brillouin zone (or to any other convenient primitive cell of the reciprocal
lattice). This is because any k' not in the first Brillouin zone can be written as

r«  r  l (3.20)

where G is a reciprocal lattice vector and k does lie in the first zone. Since
 Po
1 for any reciprocal lattice vector, if the Bloch form Eq. (3.17) holds for
k', it will also hold for k. An example is given below for a nearly free electron
model.
The energy E of free electrons which is plotted versus k in Fig 3.2a exhibits a
curve in the familiar parabolic shape. Figure 3.2b shows the result of translations.
Segments of the parabola of Fig.3.2a are cut at the edges of the various zones, and
are translated by multiples of G = 2π/a in order to ensure that the energy is the
same at any two equivalent points. Fig.3.2c displays the shape of the energy
spectrum when we confine our consideration to the first Brillouin zone only. The
type of representation used in Fig.3.2c is referred to as the reduced-zone scheme.
Because it specifies all the needed information, it is the one we shall find most
convenient. The representation of Fig.3.2 a, known as the extended-zone scheme
is convenient when we wish to emphasize the close connection between a
crystalline and a free electron. Fig.3.2b employs the periodic-zone scheme, and is
sometimes useful in topological considerations involving the k space. All these
135
representations are strictly equivalent; the use of any particular one is dictated by
convenience, and not by any intrinsic advantages it has over the others.

Fig.3.2 Free electron bands within reduced- (a), extended- (b) and periodic-
zone (c) scheme (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8)

3 An important consequence of the Bloch theorem is the appearance of the energy


bands. All solutions to the Schrodinger equation (3.5) have the Bloch form
where k is fixed and has the periodicity of the Bravais
lattice. Substituting this into the Schrodinger equation, we find that is
determined by the eigenvalue problem

(3.21)

With boundary condition

(3.22)
Because of the periodic boundary condition we can regard (3.21) as an eigenvalue
problem restricted to a single primitive cell of the crystal. Because the eigenvalue
problem is set in a fixed finite volume, we expect on general grounds to find an
infinite family of solutions with discretely spaced eigenvalues, which we label
with the band index n. The Bloch function can therefore be denoted by
which indicates that each value of the band index n and the vector k specifies an
electron state, or orbital with energy . Note that in terms of the eigenvalue
problem specified by (3.21) and (3.22), the wave vector k appears only as a
136
parameter in the Hamiltonian H (k). We therefore expect each of the energy
levels, for given k, to vary continuously as k varies. In this way we arrive at a
description of the levels of an electron in a periodic potential in terms of a family
of continuous functions  :<. For each n, the set of electronic levels specified by
 :< is called an energy band. The information contained in these functions for
different n and k is referred to as the band structure of the solid.

4 Number of states in a band.


The number of orbitals in a band within the first Brillouin zone is equal to the
number of unit cells N in the crystal. This is much the same as the statement made
in connection with the number of lattice vibrational modes, and is proved in a like
manner, by appealing to the boundary conditions. Consider first the one-
dimensional case. The allowed values of k form a uniform mesh whose unit
spacing is 2π/L. The number of states inside the first zone, whose length is 2π/a,
is therefore equal to (2π/a)/ (2π/L) = L/a = N, where N is the number of unit cells,
in agreement with the assertion made earlier. A similar argument may be used to
establish the validity of the statement in two- and three-dimensional lattices. It has
been shown that each band has N states inside the first zone. Since each such state
can accommodate at most two electrons, of opposite spins, in accordance with the
Pauli Exclusion Principle, it follows that the maximum number of electrons that
may occupy a single band is 2N. This result is significant, as it will be used in a
later section to establish the criterion for predicting whether a solid is going to
behave as a metal or an insulator.

5. Now we show that an electron in a level specified by band index n and wave vector k
has a nonvanishing mean velocity, given by

9;s :C<
¯ :C<  (3.23)
ð9C

To show this we calculate the expectation value of the derivative of the Hamiltonian
H (k) in Eq. (3.21) with respect to k:

:gC  Ô<÷ Z Ý  Û* ÷ð F ÔH÷ * Ý


Qd:C< ð. Pð
ÛZ ÷ ÷ Z Ý  ÛZ ÷g

(3.24)
QC 

Since ¯  :gð⁄<Ô is the velocity operator, this establishes (3.23).


This is a remarkable fact. It asserts that there are stationary levels for an electron
in a periodic potential in which, in spite of the interaction of the electron with the
fixed lattice of ions, it moves forever without any degradation of its mean
velocity. This is in striking contrast to the idea of Drude that collisions were
simply encounters between the electron and a static ion.

3.3 Weak potential

 z :C<  ,
ð. >.
When the potential is zero the solutions of the Schrödinger equation (3.14) are plane waves

Et
(3.25)
137
(3.26)

Where the wave function is normalized to the volume of unit cell . In the reduced-zone
representation shown in Fig.3.3, for each k there is an infinite number of solutions which
correspond to different G (and can be labeled by index n), as we have already discussed.
Each band in Fig.3.3 corresponds to a different value of G in the extended scheme.

Fig.3.3: Only those states which have the same k in the First Brillouin
zone are coupled by perturbation (After Kittel, 1979)

Suppose now that a weak potential is switched on. According to the Schrödinger equation
(3.14) only those states, which differ by G, are coupled by a perturbation. In the reduced
zone scheme those states have same k and different n (see Fig.3.3). From quantum
mechanics, if the perturbation is small compared to the energy difference between the
states, which are coupled by the perturbation, we can use the perturbation theory to
calculated wave functions and energy levels. Assuming for simplicity that we are looking
for the correction to the energy of the lowest band , the condition for using the
perturbation theory is

(3.27)

For any G ≠ 0. According to the perturbation theory the energy is given by

(3.28)

The first term in Eq. (3.28) is the undisturbed free-electron value for the energy. The second
term is the mean value of the potential in the state :

(3.29)

138
This term gives a constant independent of k. Its effect on the spectrum is a rigid shift by a
constant value without causing any change in the shape of the energy spectrum. This term
can be set equal to zero. The third term can be rewritten as

(3.30)

Finally we obtain for the energy:

(3.31)

The perturbation theory breaks down, however, in those cases when the potential cannot
be considered as a small perturbation. This happens when the magnitude of the potential
becomes comparable with the energy separation between the bands, i.e.

(3.32)

In this case we have to include these levels in the Schrödinger equation and solve it explicitly
There are special k points for which the energy levels become degenerate and the
relationship (3.32) holds for any non-zero value of the potential. For these k points

(3.33)

and consequently

(3.34)

The latter conduction implies that k must lie on a Bragg plane bisecting the line joining
the origin of k space and the reciprocal lattice point G, as is shown in Fig.3.4.

Fig. 3.4 If |k| = |k – G|, then the point k must lie in the Bragg plane
determined by G. (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8)

Therefore, a weak periodic potential has its major effect on those free electron levels
whose wave vectors are close to ones at which the Bragg reflection can occur. In order to
139
find the energy levels and the wave functions of near these points we include to the
equation (3.14) only the two levels: one which corresponds to k and the other which
corresponds to assuming that k lies near the Bragg plane:

(3.35)

These equations have the solution when the determinant is equal to zero, i.e.

(3.36)

this leads to the quadratic equation


(3.37)

The two roots are

(3.38)

These solutions are plotted in Fig.3.4 for k parallel to G.

Fig.3.4: Plot of the energy bands given by Eq. (3.38) for k parallel
to G. (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8)

This results is particularly simple for point lying on the Bragg plane, since in this case
we find from (3.38) then that

(3.39)

140
Thus, at all points in the Bragg plane, one level is uniformly raised by and the other
is uniformly lowered by the same amount. This means that there are no states in the
energy interval between and which
implies the creation of the band gap. The magnitude of the band gap is equal to twice the
Fourier component of the crystal potential. We illustrate this behavior using a one-
dimensional lattice shown in Fig.3.5. We see the splitting of the bands at each Bragg
plane in the extended-zone scheme (Fig.3.5b). This results in the splitting of the bands
both at the boundaries and at the centre of the first Brillouin zone (Fig.3.5a). There are
two important points to note. First, since the energy there increases as , the higher the
band, the greater its width. Second, the higher the energy, the narrower the gap; this
follows from the fact that the gap is proportional to a Fourier component of the crystal
potential and that the order of the component increases as the energy rises. Since the
Fourier components of the potential decrease rapidly as the order increases, this leads to a
decrease in the energy gap. It follows therefore that, as we move up the energy scale, the
bands become wider and the gaps narrower; i.e., the electron behaves more and more like
a free particle.

Fig. 3.5 (a) Dispersion curves in the nearly-free-electron model, in the


reduced-zone scheme; (b) The same dispersion curves in the extended-
zone scheme. (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8)

Now we discuss the origin of the appearance of the band gaps at the Bragg planes. When
k lies on a Bragg plane we can easily find the form of the wave function corresponding to
the two solutions (3.39). Assuming for simplicity that the potential is real we obtain from
Eq. (3.35)
(3.40)

For simplicity we consider a one-dimensional lattice, for which the Bragg reflection
occurs at k=½G. We have then
141
(3.41)

We see that at the zone edge, the scattering is so strong that the reflected wave has the
same amplitude as the incident wave. The electron is represented there by a standing
wave, very unlike a free particle.

The distribution of the charge density is proportional to |ψ|2, so that

(3.42)

Since the origin lies at the ion, the ψ − state distributes the electron so that it is piled
predominantly at the nuclei (see Fig.3.6). Since the potential is most negative there, this
distribution has a low energy. The function ψ − therefore corresponds to the energy at the
top of band 1, that is, point A1 in Fig. 3.5a.

Fig.3.6: Spatial distributions of the charge density described by the


functions ψ + and ψ − (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8).

By contrast, the function ψ + deposits its electron mostly between the ions (as shown in
Fig.3.6), corresponds to the bottom of band 2 in Fig.3.5a, that is, point A2. The gap
arises, therefore, because of the two different distributions for the same value of k, the
distributions having different energies.

Worked Example:

Consider two-dimensional electrons subjected to a weak periodic potential coming from a


square lattice of spacing . For a k vectors far away from the Brillouin zone
boundary, the wavefunction can be well described by planes waves. Assume we want to
write the wavefunction in the Bloch form, and considering a state of
energy E and wavevector ,

142
a) What will the three lowest energies be at this wavenumber?
b) What are the corresponding Z:< functions
Note that ð )2  3.806 ŽkE .
E

Solution:

a) Recall Eq.(3.5), the Schrodinger equation is


ð. u. v
  Ž:<*  *:<
E uC?

u. v
 ∆E
u7 .
Where

If the potential is weak, the solutions will be plane waves:


From Eq. (3.25)

ðE |; p|E
7 
2
and Eq.(3.25)

*:Cp <   PC
1 p

√Ž
where Cp extends over the entire k space. We can transform the Cp wavevector into the

Cp  C  
first Brillouin zone by using Eq. (3.20) i.e.

Let C  [D  [E ,
Where [D and [E are primitive reciprocal lattice vectors and n and m are integers. The

0
[D  †1.256 k ˆ and [E  F
eD
H.
primitive reciprocal lattice vectors are given by

0 1.256 keD

With the value of ;  0.5 keD , the length of the Cp vector for several values of n and m is

|Cp |
shown in the table below
n m
0 0 0.5
-1 0 0.756
1 0 1.756
0 1 1.351
0 -1 1.351
1 1 2.159

Since the energies increases with |Cp |, the three lowest energies obtained using Eq. (3.25)

  0.95 Ž :  0,   0<
are:

  2.17 Ž :  1,   0<


I.

  6.96 Ž :  0,  é 1<
II.
III.
143
›
  F¬ H
(b) From
p
 >C   >C Z:< ,
Solving for Z :<, we have:
(a) Z:<  1
(b) Z :<   ePR- ¡   ePD.ErÄ¡
(c) Z:<   éR. S   éD.ErÄS
Note that from Eq. (3.2), the function Z:< has the periodicity of the lattice,Z:< 
Z:  i<. The third energy level is degenerate; there are two corresponding
wavefunctions.

3.4 Metals and Insulators


Solids are divided into two major classes: metals and insulators. A metal – or a conductor
– is a solid in which an electric current flows under the application of electric field. By
contrast, application of an electric field produces no electric current in an insulator. There
is a simple criterion for distinguishing between the two classes on the basis of the band
structure. If the valence electrons exactly fill one or more bands, leaving others empty,
the crystal will be an insulator. An external electric field will not cause current flow in an
insulator. Provided that a filled band is separated by energy gap from the next higher
band, there is no continuous way to change the total momentum of the electrons if every
accessible state is filled. Nothing changes when the field is applied.
On the contrary if the valence band is not completely filled the solid is a metal. In a metal
there are empty states available above the Fermi level like in a free electron gas. An
application of an external electric field results in the current flow. It is possible to
determine whether a solid is a metal or an insulator by considering the number of valence
electrons. A crystal can be an insulator only if the number of valence electrons in a
primitive cell of the crystal is an even integer. This is because each band can
accommodate only two electrons per primitive cell. For example, diamond has two atoms
of valence four, so that there are eight valence electrons per primitive cell. The band gap
in diamond is 7eV and this crystal is a good insulator. However, if a crystal has an even
number of valence electrons per primitive cell, it is not necessarily an insulator. It may
happen that the bands overlap in energy. If the bands overlap in energy, then instead of
one filled band giving an insulator, we can have two partly filled bands giving a metal
(Fig.3.7b). For example, the divalent metals, such as Mg or Zn, have two valence
electrons per cell. However, they are metals, although a poor ones – their conductivity is
small.

144
Fig.3.7: Occupied states and band structures giving (a) an insulator, (b) a
metal or a semimetal because of band overlap, and (c) a metal because of
electron concentration (After Kittel, 1979)

If this overlap is very small, we deal with semimetals. The best known example of a
semimetal is bismuth (Bi). If the number of valence electrons per cell is odd the solid is
a metal. For example, the alkali metals and the noble metals have one valence electron
per primitive cell, so that they have to be metals. The alkaline earth metals have two
valence electrons per primitive cell; they could be insulators, but the bands overlap in
energy to give metals, but not very good metals. Diamond, silicon, and Germanium each
have two atoms of valence four, so that there are eight valence electrons per primitive
cell; the bands do not overlap, and the pure crystals are insulators at absolute zero. There
are substances, which fall in an intermediate position between metals and insulators. If
the gap between the valence band and the band immediately above it is small, then
electrons are readily excitable thermally from the former to the latter band. Both bands
become only partially filled and both contribute to the electric condition. Such a
substance is known as a semiconductor. Examples are Si and Ge, in which the gaps are
about 1 and 0.7 eV, respectively. Roughly speaking, a substance behaves as a
semiconductor at room temperature whenever the gap is less than 2 eV. The
conductivity of a typical semiconductor is very small compared to that of a metal, but it
is still many orders of magnitude larger than that of an insulator. It is justifiable,
therefore, to classify semiconductors as a new class of substance, although they are,
strictly speaking, insulators at very low temperatures.

4.0 Conclusion
Solution of Schrodinger equation for a single electron allows the prediction of the physical
properties of a crystal while the Bloch theorem plays an important role in electronic band
structure theory.

5.0 Summary
• Separation of the valence and conduction band : ù  N  0
• Periodic potential of an electron is in the form: Ã :<  Ã:  i<
• One electron Schrödinger equation with a periodic potential, are known as Bloch
electrons

145
• From the Bloch theorem, The number of orbitals in a band within the first
Brillouin zone is equal to the number of unit cells N in the crystal
• Solids are divided into two major classes: metals and insulators which can be
distinguished on the basis of band structure.

6.0 Tutor marked assignment


Q1. Using the solution for the energy bands near the zone boundary in the
presence of a weak crystal potential. Show that the electron velocity is
parallel to the Bragg plane.

Q2. Prove that the current carried by Bloch electrons is given by

ð¸?
B  † C
ˆC


7.0 Further reading/References


Animalu, A.O.E, intermediate quantum theory of crystalline solids, Prentice-Hall
of India, New Delhi, 1978
Blakemore, J.S., Solid State Physics, W.B. Saunders Co., 1974
Callaway, J., Energy band theory, Academic Press, New York, 1958
Kittel C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kittel, C., Quantum theory of solids, John Wiley, 1963
Hurd, C.M., The Hall Effect in metals and alloys, Plenum, 1972
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge,1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8
Ziman, J.M., Electrons and Phonons, Cambridge University Press, 1960

146
UNIT 4: ELECTRON DYNAMICS page
1.0 Introduction - - - - - - - - 142
2.0 Objectives - - - - - - - - 142
3.0 Definition - - - - - - - - 142
3.1 Electron dynamics - - - - - - 142
3.2 Effective mass - - - - - - - 143
3.3 Current density - - - - - - 144
3.4. Hole - - - - - - - - 145
4.0 Conclusion - - - - - - - - 146
5.0 Summary - - - - - - - - 146
6.0 Tutor Marked Assignment - - - - - - 146
7.0 Further Reading/References - - - - - - - 146

147
1.0 Introduction
The Fermi surfaces (FS) concept enables to visualize the relative fullness or occupation
of the allowed empty lattice bands geometrically in k-space and thus helps in the
theoretical determination of the electronic properties of a solid- metal, semiconductor or
insulator. In fact, the purpose of the FS construction is to know about the details of the
motion of an itinerant electron in three-dimension.

2.0 Objective
• to understand the concept of Fermi surfaces
• to revise the concept of electron dynamic
• to revise the concept of effective mass
• to revise the concept of hole

3.0 Definition
Electron dynamics is using classical equations of motion in a classical way to describe
electronic structure quantum-mechanically, i.e. standing waves that distribute electrons to
different regions of the bands.

3.1 Electro dynamics


Given the functions En(k) the semiclassical model associates with each electron a
position, a wave vector and a band index n. In the presence of applied fields the
position, the wave vector, and the index are taken to evolve according to the following
rules:
(i) The band index is a constant of the motion. The semiclassical model ignores
the possibility of interband transitions. This implies that within this model it
assumed that the applied electric field is small.
(ii) The time evolution of the position and the wave vector of an electron with band
index n are determined by the equations of motion:

Q D Q ò :C<
 ¯ :C< 
ð QC
(4.1)
QT

ð  ¾:, Y<  K:, Y<


QC
QT (4.2)

Strictly speaking Eq. (4.2) has to be proved. It is identical to the Newton’s second law if
we assume that the electron momentum is equal toðk. The fact that electrons belong to
particular bands makes their movement in the applied electric field different from that of
free electrons. For example, if the applied electric field is independent of time, according
to Equation (4.2) the wave vector of the electron increases uniformly with time.

C:Y <  C:0< 


MKT
ð
(4.3)

148
Since velocity and energy are periodic in the reciprocal lattice, the velocity and the
energy will be oscillatory. This is in striking contrast to the free electron case, where v is
proportional to k and grows linearly in time. The k dependence (and, to within a scale
factor, the t dependence) of the velocity is illustrated in Fig 4.1, where both E(k) and v(k)
are plotted in one dimension. Although the velocity is linear in k near the band minimum,
it reaches a maximum as the zone boundary is approached, and then drops back down,
going to zero at the zone edge. In the region between the maximum of v and the zone
edge the velocity actually decreases with increasing k, so that the acceleration of the
electron is opposite to the externally applied electric force! This extraordinary behavior is
a consequence of the additional force exerted by the periodic potential, which is included
in the functional form of E (k). As an electron approaches a Bragg plane, the external
electric field moves it in the opposite direction due to the Bragg-reflection.

Fig.4.1. E(k) and v(k) vs. k in one dimension (After


www.pa.uk.edu/kwng.phy/525/lec/lecture-8)

3.2 Effective mass


When discussing electron dynamics in solids it is often convenient to introduce the
concept of effective mass. If we differentiate Eq. (4.1) with respect to time we find that

(4.4)

Where the second derivative with respect to a vector should be understood as a tensor.
Using Eq. (4.2) we find that

(4.5)

In one dimensional case this reduces to

(4.6)

149
This has the same form as the Newton’s second law, provided that we defined an effective
mass by the relation:

(4.7)

The mass m* is inversely proportional to the curvature of the band; where the curvature is
large - that is, is large - the mass is small; a small curvature implies a large mass
(Fig.4.2).

Fig: 4.2. The inverse relationship between the mass and the
curvature of the energy band
(After www.pa.uk.edu/kwng.phy/525/lec/lecture-8).

In a general case the effective mass is a tensor which is defined by

(4.8)

Where and are Cartesian coordinates. The effective mass can be different
depending on the directions on the crystal.

3.3 Current density


The current density within a free electron model was defined as
where n is the number of valence electrons per unit volume, and v is the velocity of
electrons. This expression can generalize to the case of Bloch electrons. In this case the
velocity depends on the wave vector and we need to sum up over k vectors for which
there are occupied states available:

(4.9)

Here the sum is performed within the extended zone scheme and V is the volume of the
solid. It is often convenient to replace the summation by the integration. Because the
volume of k-space per allowed k value is we can write the sum over k as

150
∑w 

fŸ  
 5C (4.10)

Taking into account the spin degeneracy we obtain for the current density:

5C
B  
4 …
 L:C<
(4.11)

}NN,OPMQ

Using this expression we show now that completely filled bands do not contribute to the
current. For the filled bands Eq. (4.11) should be replace by

dr dE:r<
x  e 
4π… dr
(4.12)
z{|@

This vanishes as a consequence of the theorem that the integral over any primitive cell of
the gradient of a periodic function must vanish.

3.4 Hole
One of the most impressive achievements of the semiclassical model is its explanation for
phenomena that free electron theory can account for only if the carriers have a positive
charge. We now introduce the concept of a hole.
The contribution of all the electrons in a given band to the current density is given by Eq.
(4.11), where the integral is over all occupied levels in the band. By exploiting the fact
that a completely filled band carries no current, thus we have

5; 5C 5C
0  L:C<  L:C<  L:C<
4 … 4 … 4 …
  (4.13)
‰LM LNN,OPMQ ,LNN,OPMQ

we can equally well write Eq. (4.11), in the form:

5C
B    ¯ :C<
4 … (4.14)
,LNN,OPMQ

Thus the current produced by electrons occupying a specified set of levels in a band is
precisely the same as the current that would be produced if the specified levels were
unoccupied and all other levels in the band were occupied with particles of charge +e
(opposite to the electronic charge).
Thus, even though the only charge carriers are electrons, we may, whenever it is
convenient, consider the current to be carried entirely by fictitious particles of positive
charge that fill all those levels in the band that are unoccupied by electrons. The fictitious
particles are called holes. It must be emphasized that pictures cannot be mixed within a

151
given band. If one wishes to regard electrons as carrying the current, then the unoccupied
levels make no contribution; if one wishes to regard the holes as carrying the current,
then the electrons make no contribution. One may, however, regard some bands using the
electron picture and other bands using the hole picture, as suits one's convenience.
Normally it is convenient to consider transport of the holes for the bands which are
almost occupied, so that only a few electrons are missing. This happens in
semiconductors in which a few electrons are excited from the valence to the conduction
bands. Similar to electrons we can introduce the effective mass for the holes. It has a
negative sign.

4.0 Conclusion
The electron dynamics in metals is the electronic structure described by quantum
mechanics based on semiclassical model

5.0 Summary
• Effective mass of an electron is defined by
1 1 5E 

p ðE 5C?
• Current density is defined by

5C
L:C<
4 …
B  LNN,OPMQ

6.0 Tutor marked assignment

Q1. Consider a slab of Cu 0.1mm thick, 10.0 mm wide and 10.0mm long.
(a) If a current of 1A is driven down the length of the slab, what is the
current density?
(b) If we put the slab in the magnetic field of 1 T with the field
perpendicular to the 1 mm x10 mm face, what Hall Effect will be
produced, if the Hall coefficient is -0.55x10-10 m3/C.
(c) What Hall voltage will be observed across the slab?

7.0 Further reading/References


Animalu, A.O.E, intermediate quantum theory of crystalline solids, Prentice-Hall
of India, New Delhi, 1978
Blakemore,J.S., Solid State Physics, W.B. Saunders Co.,1974
Callaway, J., Energy band theory, Academic Press, New York, 1958
Kittel C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kittel, C., Quantum theory of solids, John Wiley, 1963
Hurd,C.M., The hall effect in metals and alloys, Plenum,1972
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge,1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8
Ziman, J.M., Electrons and Phonons, Cambridge University Press, 1960
152
UNIT 5: FERMI SURFACES page
1.0 Introduction - - - - - - - - 148
2.0 Objectives - - - - - - - - 148
3.0 Definition - - - - - - - - 148
3.1 Fermi surfaces - - - - - - - 148
3.2 Brillouin zone - - - - - - - 149
3.3 Effect of crystal potential - - - - - 152
3.3.1 Alkali metals - - - - - - - 153
3.3.2 Noble metals - - - - - - 153
3.3.3 Cubic divalent metals - - - - - 153
3.3.4 Trivalent metal - - - - - - 154
4.0 Conclusion - - - - - - - - 155
5.0 Summary - - - - - - - - 155
6.0 Tutor Marked Assignment - - - - - - 155
7.0 Further Reading/References - - - - - - 156

153
1.0 Introduction
The Fermi surface is the surface of constant energy ~/ in k space. The Fermi surface
separates the unfilled orbitals from the filled orbitals, at absolute zero. Quantum
mechanics showed that the occupation of electron states is governed by the Pauli
exclusion and that the chemical potential, 8 is equal to ~/ . The shape of the Fermi surface
may be very intricate but the constructions required the applications of the reduced and
the periodic zone schemes. In the reduced zone scheme, it is always possible to select the
wavevector index k of any Bloch function to lie within the first Brillouin zone. This
procedure is known as mapping the band in the reduced zone scheme. In the periodic
zone, a given Brillouin zone is repeated periodically through all of the wavevector space.
This is achieved by translating the zone by a reciprocal lattice.

2.0 Objective
• to understand Fermi surfaces
• to explain the Brillouin zone
• to explain effect of crystal potential
3.0 Definition
Fermi energy surface is the energy distribution of particles that obey the Pauli Exclusion
Principle.

3.1 Fermi surface


The ground state of N Bloch electrons is constructed in a similar fashion as that for free
electrons, i.e. by occupying all one-electron energy levels with band energies  :C< less
than / , where / is determined by requiring the total number of levels with energies
less than / to be equal to the total number of electrons. The wave vector k must be
confined to a single primitive cell of the reciprocal lattice. When the lowest of these
levels are filled by a specified number of electrons, two quite distinct types of
configuration can result:
1. A certain number of bands may be completely filled, all others remaining empty.
Because the number of levels in a band is equal to the number of primitive cells in
the crystal (and because each level can accommodate two electrons (one of each
spin), a configuration with a band gap can arise only if the number of electrons
per primitive cell is even.
2 A number of bands may be partially filled. When this occurs, the energy of the
highest occupied level, the Fermi energy/ , lies within the energy range of one or
more bands. For each partially filled band there will be a surface in k-space
separating the occupied from the unoccupied levels. The set of all such surfaces is
known as the Fermi surface, and is the generalization to Bloch electrons of the
free electron Fermi sphere. The parts of the Fermi surface arising from individual
partially filled bands are known as branches of the Fermi surface.
Analytically, the branch of the Fermi surface in the n-th band is that surface in k-
space determined by

 :;<  / (5.1)

Thus the Fermi surface is a constant energy surface (surfaces) in k-space.

154
Since the are periodic in the reciprocal lattice, the complete solution to Eq. (5.1)
for each n is a k-space surface with the periodicity of the reciprocal lattice. When a
branch of the Fermi surface is represented by the full periodic structure, it is said to be
described in a repeated zone scheme. Often, however, it is preferable to take just enough
of each branch of the Fermi surface so that every physically distinct level is represented
by just one point of the surface. This is achieved by representing each branch by that
portion of the full periodic surface contained within a single primitive cell of the
reciprocal lattice. Such a representation is described as a reduced zone scheme. The
primitive cell chosen is often, but not always, the first Brillouin zone.

3.2 Brillouin Zone

We consider now an example of building of a Fermi surface. We start from considering


the Fermi surface for free electrons and then investigate the influence of the crystal
potential. The Fermi surface for free electrons is a sphere centered at k = 0. To construct
the Fermi surface in the reduced-zone scheme, one can translate all the pieces of the
sphere into the first zone through reciprocal lattice vectors. This procedure is made
systematically through the geometrical notion of the higher Brillouin zones

a b
Fig. 5.1: (a) Construction in k space of the first three Brillouin
zones of a square lattice. (b) On constructing all lines equivalent by
symmetry to the three lines in (a) we obtain the regions in k space
which form the first three Brillouin zones (After Kittel, 1979).

We illustrate this construction for the two dimensional cubic lattice shown in Fig.5.1.
Recall that the boundaries of the Brillouin zones are planes normal to G at the midpoint
of G. The first Brillouin zone of the square lattice is the area enclosed by the
perpendicular bisectors of and of the three reciprocal lattice vectors equivalent by
symmetry to in Fig. 5.1a. These four reciprocal lattice vectors are and
The second zone is constructed from and the three vectors equivalent to
it by symmetry, and similarly for the third zone. The pieces of the second and third zones
are drawn in Fig. 5.1b.
In general, the first Brillouin zone is the set of points in k-space that can be reached from
the origin without crossing any Bragg plane. The second Brillouin zone is the set of
points that can be reached from the first zone by crossing only one Bragg plane.
155
Brillouin zone is the set of points not in the zone that can
be reached from the zone by crossing only one Bragg plane. The free electron
Fermi surface for an arbitrary electron concentration is shown in Fig.5.2.

Fig.5.2: Brillouin zones of a square lattice in two dimensions (After Kittel,


1979).

Now we perform a transformation to the reduced zone scheme as is shown in Figs.5.3


and 5.4. We take the triangle labeled 2a (Fig 5.2) and move it by a reciprocal lattice
vector such that the triangle reappears in the area of the first Brillouin
zone (Fig.5.3). Other reciprocal lattice vectors will shift the triangles 2b, 2c, 2d to other
parts of the first zone, completing the mapping of the second zone into the reduced zone
scheme. The parts of the Fermi surface falling in the second zone are now connected, as
shown in Fig. 5.4.

a b c
Fig.5.3 Mapping of the first, second, and third Brillouin zones in
the reduced zone scheme. The sections of the second zone in Fig.
5.1 are put together into a square by translation through an
appropriate reciprocal lattice vector (After Kittel, 1979).

156
Fig.5.4: The free electron Fermi surfaces of Fig.5.3, as viewed in
the reduced zone scheme. The shaded areas represent occupied
electron states. Parts of the Fermi surface fall in the second and
third zones. The first zone is entirely occupied (After Kittel, 1979).

Construction of Brillouin zones and Fermi surfaces in three-dimensions is more


complicated. Fig5.5 shows the first three Brillouin zones for bcc and fcc structures.

a b
Fig.5.5:Surfaces of the first, second, and third Brillouin zones for
(a) body-centered cubic and (b) face-centered cubic crystals. (Only
the exterior surfaces are shown (After Kittel, 1979)..

157
The free electron Fermi surfaces for IGG cubic metals of valence 2 and 3 are shown in
Fig.5.6.

Fig.5.6: The free electron Fermi surfaces for face-centered cubic metals of
valence 2 and 3(After Kittel, 1979).

3.3 Effect of a crystal potential


How do we go from Fermi surfaces for free electrons to Fermi surfaces in the presence of
a weak crystal potential? We can make approximate constructions freehand by the use of
the following facts:
(i) The interaction of the electron with the periodic potential of the crystal causes
energy gaps at the zone boundaries.
(ii) Almost always the Fermi surface will intersect zone boundaries perpendicularly.

that QC   FC  E H which implies that on the Bragg plane the gradient of


D
Using the equation for the energy near the zone boundary it is easy to show
Q ð.

energy is parallel to the Bragg plane. Since the gradient is perpendicular to the
surfaces on which function is constant, the constant energy surfaces at the Bragg
plane are perpendicular to the plane.
(iii) The crystal potential will round out sharp corners in the Fermi surfaces.
(iv) The total volume enclosed by the Fermi surface depends only on the electron
concentration and is independent of the details of the lattice interaction.
(v) If a branch of the Fermi surface consists of very small pieces of surface
(surrounding either occupied or unoccupied levels, known as "pockets of
electrons" or "pockets of holes"), then a weak periodic potential may cause these
to disappear. In addition, if the free electron Fermi surface has parts with a very
narrow cross section, a weak periodic potential may cause it to become
disconnected at such points.

Below we give a few examples for real metals.

158
3.3.1. Alkali metals
The radius of the Fermi sphere in bcc alkali metals is less than the shortest distance from
the center of the zone to a zone face and therefore the Fermi sphere lies entirely within
the first Brillouin zone. The crystal potential does not distort much the free electron
Fermi surface and it remains very similar to a sphere. Fig 5.7 shows Fermi surface for
sodium.

Fig.5.7: Fermi surface of sodium (After www.pa.uk.edu/kwang.phy/525/lec-8)

3.3.2. Noble metals


The Fermi surface for a single half-filled free electron band in fcc Bravais lattice is a
sphere entirely contained within the first Brillouin zone, approaching the surface of the
zone most closely in the [111] directions, where it reaches 0.903 of the distance from the
origin to the center of the hexagonal face. For all three noble metals therefore their Fermi
surfaces are closely related to the free electron sphere. However, in the [111] directions
contact is actually made with the zone faces, and the measured Fermi surfaces have the
shape shown in Fig.5.8. Eight "necks reach out to touch the eight hexagonal faces of the
zone, but otherwise the surface is not grossly distorted from spherical.

Fig. 5.8: In the three noble metals the free electron sphere bulges
out in the [111] directions to make contact with the hexagonal zone
faces.

3.3.3. Cubic divalent metals


With two electrons per primitive cell, calcium, strontium, and barium could, in principle,
be insulators. In the free electron model, the Fermi sphere has the same volume as the
first zone and therefore intersects the zone faces. The free electron Fermi surface is thus a
fairly complex structure in the first zone, and pockets of electrons in the second. The

159
question is whether the effective lattice potential is strong enough to shrink the second-
zone pockets down to zero volume, thereby filling up all the unoccupied levels in the first
zone. Evidently this is not the case, since the group II elements are all metals.
Calculations show that the first Brillouin zone is completely filled and a small number of
electrons in the second zone determine the non-zero conductance.

Fig.5.9: Fermi surface of calcium (Afterwww.pa.uk.edu/kwng.phy/525/lec/lecture-8)


.

3.3.4. Trivalent metals


The Fermi surface of aluminum is close to that of the free electron surface for fcc cubic
monatomic lattice with three conduction electrons per atom. The first Brillouin zone is
filled and the Fermi surface of free electrons is entirely contained in the second, third and
fourth Brillouin zones. When displayed in a reduced-zone scheme the second-zone
surface is a closed structure containing unoccupied levels, while the third-zone surface is
a complex structure of narrow tubes (Fig.5.6). The amount of surface in the fourth zone is
very small, enclosing tiny pockets of occupied levels. The effect of a weak periodic
potential is to eliminate the fourth-zone pockets of electrons, and reduce the third-zone
surface to a set of disconnected "rings" (Fig.5.10). Aluminum provides a striking
illustration of the theory of Hall coefficients. The high-field Hall coefficient should
be,Ød  1: M  6 < where M and 6 are the number of levels per unit volume
enclosed by the particle-like and hole-like branches of the Fermi surface. Since the first
zone of aluminum is completely filled and accommodates two electrons per atom, one of
the three valence electrons per atom remains to occupy second- and third-zone levels.
Thus


M  M  (5.2)
…

where n is the free electron carrier density appropriate to valence 3. On the other hand,
since the total number of levels in any zone is enough to hold two electrons per atom, we
also have


M  6  2 (5.3)
…

Subtracting (5.3) from (5.2) gives

160

M  6   (5.4)
…

Thus the high-field Hall coefficient should have a positive sign and yield an effective
density of carriers a third of the free electron value. This is precisely what is observed.

Fig.5.10: Fermi surface of aluminum (After www.pu. uk.edu/kwang.phys/525/lecture8)

4.0 Conclusion
The Fermi surfaces (FS) concept enables to visualize the relative fullness or occupation
of the allowed empty lattice bands geometrically in k-space and thus helps in the
theoretical determination of the electronic properties of a solid.

5.0 Summary
• The N Bloch electron is constructed when the wave vector k is confined to single
primitive cell.
• In Alkali metals, the Fermi surface is very much like a sphere
• In Noble metals, the Fermi surface is a sphere entirely contained within the first
Brillouin zone.
• In Cubic divalent metals, the Fermi surface has the same volume as the first
Brillouin zone.
• In Trivalent metals, the Fermi surface is entirely contained in the 2nd , 3rd and the
4th Brillouin zone.

6.0 Tutor marked assignment


Q1. A two-dimensional metal has one atom of valence one in a simple
rectangular primitive cell of a = 2Å and a = 4Å.
1 2
(a) Draw the first and the second Brillouin zones.
(b) Calculate the radius of the free electron Fermi sphere and draw this
sphere to scale on the drawing of the Brillouin zones.
(c) Draw the Fermi surface in reduced zone scheme and show
schematically the effect of a weak crystal potential.
Q2 Suppose that some atoms in a Cu crystal, which has an IGG lattice, are
gradually replaced by Zn atoms. Considering that Zn is divalent while Cu
is monovalent, calculate the atomic ratio of Zn to Cu in a ‘Z‚ alloy
161
(brass) at which the Fermi sphere touches the zone faces. Use the free-
electron model. This particular alloy is interesting because the solid
undergoes a structural phase change at this concentration ratio.

7.0 Further reading/References


Animalu, A.O.E, intermediate quantum theory of crystalline solids, Prentice-Hall
of India, New Delhi, 1978
Blakemore, J.S., Solid State Physics, W.B. Saunders Co., 1974
Callaway, J., Energy band theory, Academic Press, New York, 1958
Kittel C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kittel, C., Quantum theory of solids, John Wiley, 1963
Hurd,C.M., The hall effect in metals and alloys, Plenum,1972
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
W. A. Wooster, A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8
Ziman, J.M., Electrons and Phonons, Cambridge University Press, 1960

162
Module 4:

SEMICONDUCTORS AND SUPERCONDUCTORS

163
UNIT 1: STRUCTURE AND BONDING (SEMICONDUCTORS) page
1.0 Introduction - - - - - - - - - 159
2.0 Objectives - - - - - - - - 159
3.0 Definition - - - - - - - - 159
3.1 Crystal structure and bonding - - - - - 159
3.2 Bonding structure - - - - - - 160
3.3 intrinsic semiconductor- - - - - - 161
3.4 Impurities states - - - - - - 165
3.5 Acceptors - - - - - - - 167
4.0 Conclusion - - - - - - - - 168
5.0 Summary - - - - - - - - 168
6.0 Tutor Marked Assignment - - - - - - 169
7.0 Further Reading/References - - - - - - 169

164
1.0 Introduction
In a semiconductor the valence band is almost completely filled while the conduction
band is empty. Thermal excitation or (energy) absorption processes may cause some
electrons to cross the band gap, making it similar to semimetals. Semiconductors tend to
be bonded tetrahedrally and covalently, although binary semiconductors may have polar,
as well as covalent character.
2.0 Objective
• The objective of this unit is to
• Understand the structure and bonding in semiconductors.
• Explain intrinsic semiconductors.
• Understand the importance of impurity states of semiconductors.

Definition
Semiconductors are electronic conductors with electrical resistivity values generally in
the range of 10-2 to 109 ohm-cm at room temperature, intermediate between good
conductors(10-6 ohm-cm) and insulators (1014 to 1022 ohm-cm).

3.1 Crystal structure and bonding

Semiconductors include a large number of substances of widely different chemical and


physical properties. These materials are grouped into several classes of similar behavior,
the classification being based on the position in the periodic table of the elements.
The best-known class is the Group IV semiconductors - C (diamond), Si, î, - all of
which lie in the fourth column of the periodic table. They have been studied intensively,
particularly Si and Ge, which have found many applications in electronic devices. The
elemental semiconductors all crystallize in the diamond structure. The diamond structure
has an fcc lattice with a basis composed of two identical atoms, and is such that each
atom is surrounded by four neighboring atoms, forming a regular tetrahedron. Group IV
semiconductors are covalent crystals, i.e., the atoms are held together by covalent bonds.
These bonds consist of two electrons of opposite spins distributed along the line joining
the two atoms. The covalent electrons forming the bonds are hybrid XJ… atomic orbitals.
Another important group of semiconductors is the Group III-V compounds, so named
because each contains two elements, one from the third and the other from the fifth
column of the periodic table. The best-known members of this group are î*X and InSb
(indium antimonite), but the list also contains compounds such as î*x, InAs, î*}[, and
many others. These substances crystallize in the zinc blend structure which is the same as

different. Thus, in î*X, the basis of the fcc lattice consists of two atoms, Ga and As.
the diamond structure, except that the two atoms forming the basis of the lattice are now

Because of this structure, each atom is surrounded by four others of the opposite kind,
and these latter atoms form a regular tetrahedron, just as in the diamond structure.
The bonding in the III-V compounds is also primarily covalent. The eight electrons
required for the four tetrahedral covalent bonds are supplied by the two types of atoms,
the trivalent atom contributing its three valence electrons, and the pentavalent atom five
electrons. The bonding in this group is not entirely covalent. Because the two elements in
the compound are different, the distribution of the electrons along the bond is not
symmetric, but is displaced toward one of the atoms. As a result, one of the atoms

165
acquires a net electric charge. Such a bond is called heteropolar, in contrast to the purely
covalent bond in the elemental semiconductors, which is called homopolar.
The distribution of electrons in the bond is displaced toward the atom of higher
electronegativity. In î*X for instance, the As atom has a higher electronegativity than
the Ga, and consequently the As atom acquires a net negative charge, whose value is
−0.46e per atom (a typical value in Group III-V compounds). The Ga atom
correspondingly acquires a net positive charge of 0.46e. Charge transfer leads to an ionic
contribution to the bonding in Group III-V compounds. Their bonding is therefore
actually a mixture of covalent and ionic components, although covalent ones predominate
in most of these substances.

3.2 Bonding structure

A semiconductor is a solid in which the highest occupied energy band, the valence band,
is completely full at T = 0°K, but in which the gap above this band is also small, so that
electrons may be excited thermally at room temperature from the valence band to the
next-higher band, which is known as the conduction band. Generally speaking, the
number of excited electrons is appreciable (at room temperature) whenever the energy
gap E is less than 2 eV. The substance may then be classified as a semiconductor. When
F
the gap is larger, the number of electrons is negligible, and the substance is an insulator.
When electrons are excited across the gap, the bottom of the conduction band (CB) is
populated by electrons, and the top of the valence band (VB) by holes. As a result, both
bands are now only partially full, and would carry a current if an electric field were
applied. The conductivity of the semiconductor is small compared with the conductivities
of metals of the small number of electrons and holes involved, but this conductivity is
nonetheless sufficiently large for practical purposes. The simplest band structure of a
semiconductor is indicated in Fig.1.1. Since we are interested only in the region which
lies close to the band gap, where electrons and holes lie, we can ignore a more complex
variation of the energy bands far away from the gap. The energy of the CB has the form.

ð. 7 .
N :; <  N  (1.1)
E€

where k is the wave vector and m the effective mass of the electron. The energy E
e g
represents the energy gap. The zero-energy level is chosen to lie at the top of the VB.
The energy of the VB (Fig.1.1) may be written as
ð. 7 .
0 :;<  0  E (1.2)
P

Where m is the effective mass of the hole which is positive. (Because of the inverted shape
h
of the VB, the mass of an electron at the top of the VB is negative, but the mass of a hole is
positive).

166
Fig. 1.1: Band structure in a semiconductor.

Within this simple picture of the semiconductor, the primary band-structure parameters
are thus the electron and hole masses m and m , and the band gap E . Table 1.1 gives
e h g
these parameters for various semiconductors. Note that the masses differ considerably
from the free-electron mass. In many cases they are much smaller than the free-electron
mass. The energy gaps range from 0.18 eV in â }[ to 3.7 eV in ZnS. The table also
shows that the wider the gap, the greater the mass of the electron. The energy gap for a
semiconductor varies with temperature, but the variation is usually slight. That a variation
with temperature should exist at all can be appreciated from the fact that the crystal, when
it is heated, experiences a volume expansion, and hence a change in its lattice constant.
This, in turn, affects the band structure, which is a sensitive function of the lattice
constant. The band structure in Fig 1.1 is the simplest possible structure. Band structures
of real semiconductors are somewhat more complicated, as we shall see later.

3.3 Intrinsic Semiconductors


In the field of semiconductor, electrons and holes are usually referred to as free carriers,
or simply carriers, because it is these particles which are responsible for carrying the
electric current. The number of carriers is an important property of a semiconductor, as
this determines its electrical conductivity. Intrinsic semiconductors are semiconductors in
which the number of carries and the conductivity is not influenced by impurities. Intrinsic
conductivity is typical at relatively high temperatures in highly purified specimens. In
order to determine the number of carriers, we need some of the basic results of statistical
mechanics.

167
Table 1.1. Band Structure parameters of Semiconductors

The most important result in this regard is the Fermi-Dirac (FD) distribution function.
D
I: <  :%ú0<
 Š 
(1.3)
M /D

This function, gives the probability that an energy level E is occupied by an electron
when the system is at temperature T. The function is plotted versus E in Fig.1.2. Here we
see that, as the temperature rises, the unoccupied region below the Fermi level E becomes
F
longer, which implies that the occupation of high energy states increases as the temperature is
raised, a conclusion which is most plausible, since increasing the temperature raises the
overall energy of the system.

Fig. 1.2: The Fermi-Dirac distributions function (After Kittel, 1979)


.
We will see later that the Fermi level in intrinsic semiconductors lies close to the middle
of the band gap. Therefore we can represent the distribution function and the conduction
and valence bands of the semiconductor as shown in Fig.1.3.

168
Fig.1.3: (a) conduction and valence bands (b) the distribution function
(c) Density of states for electrons and holes (After Kittel, 1979)
.
First we calculate the concentration of electrons in the CB. The number of states in the
energy range (E, E + dE) is equal toM : <5, whereM : < is the density of electron
states. Since each of these states has an occupation probability f (E), the number of
electrons actually found in this energy range is equal toI:<M :<5. The concentration
of electrons throughout the CB is thus given by the integral over the conduction band.

(1.4)

where N is the bottom the conduction band, as shown in Fig.1.3.

The band gap in semiconductors is of the order of 1eV, which is much larger than kT.
Therefore (E−µ) >> k T and we can neglect the unity term in the denominator of the
B
distribution function Eq. (1.3), so that

IM : < Y  e: e<⁄7
(1.5)

The density of the conduction band is given by

Etƒ …⁄E
.F H :E  E` <D⁄E
D
D@ :E< 
E‚ ð.
(1.6)

Note that M : < vanishes for  . N and is finite only for  Q N as shown in Fig.1.3.
When we substitute equations for f (E) and M : < into Eq. (1.4), we obtain

169
(1.7)

By changing the variable, and using the result


(1.8)

one can readily evaluate the integral in (1.7). The electron concentration then reduces to
the expression

⁄E
M ; … :e f <⁄7

(1.9)
 2† ˆ 
2ðE
The electron concentration is still not known explicitly because the Fermi energy µ is so
far unknown. Essentially the same ideas employed above may also be used to evaluate
the number of holes in the VB. The probability that a hole occupies a level E in this band
is equal to 1−f (E), since f (E) is the probability of electron occupation. Assuming that the
Fermi level lies close to the middle of the band gap, i.e. (µ−E)>>k T for the valence
B
band, we find for the distribution function of holes

1 1
I6 : <  1   Y  e:e <⁄7
(1.10)
 œ: e<⁄7
ž 1  œ:e <⁄7
ž 9)

The density of states for the holes is

EP …⁄E
F H :0   <D⁄E
D
6 : < 
EŸ . ð.
(1.11)

where 0 is the energy of the valence band edge. Proceeding in a similar fashion as we
did for electrons we find for the concentration of holes in the valence band

(1.12)

The electron and hole concentrations have thus far been treated as independent quantities.
For intrinsic semiconductors the two concentrations are, in fact, equal, because the
electrons in the CB are due to excitations from the VB across the energy gap, and for
each electron thus excited a hole is created in the VB. Therefore,

n=p (1.13)
and

170
:M <…⁄E  :e f <⁄7
 :6 <…⁄E  :  e<⁄7

(1.14)

We obtain then, for the Fermi energy

0  N 3 6
8  ; ln
(1.15)
2 4 M
The second term on the right of (1.15) is very small compared with the first, and the
energy level is close to the middle of the energy gap. This is consistent with earlier
assertions that both the bottom of the CB and the top of the VB are far from the Fermi
level. The concentration of electrons may now be evaluated explicitly by using the above
value of µ. Substitution of Eq. (1.15) into Eq. (1.9) yields

7
…⁄E
 2F H :M 6 <…⁄s  e „⁄E7

EŸð.
(1.16)

where ù  N  0 is the band gap. The important feature of this expression is that n
increases very rapidly - exponentially - with temperature, particularly by virtue of the
exponential factor. Thus as temperature is raised, a vastly greater number of electrons is
excited across the gap. Our discussion of carrier concentration in this section is based on
the premise of a pure semiconductor. When the substance is impure, additional electrons
or holes are provided by the impurities. In that case, the concentrations of electrons and
holes may no longer be equal, and the amount of each depends on the concentration and
type of impurity present. When the substance is sufficiently pure so that the
concentrations of electrons and holes are equal, we speak of an intrinsic semiconductor.
That is, the concentrations are determined by the intrinsic properties of the semiconductor
itself. On the other hand, when a substance contains a large number of impurities which
supply most of the carriers, it is referred to as an extrinsic semiconductor.

3.4 Impurity states


A pure semiconductor has equal numbers of both types of carriers, electrons and holes. In
most applications, however one needs specimens which have one type of carrier only,
and none of the other. By doping the semiconductor with appropriate impurities, one can

specimen of Si which has been doped by As. The X atoms (the impurities) occupy some
obtain samples which contain either electrons only or holes only. Consider, for instance, a

of the lattice sites formerly occupied by the Si host atoms. The distribution of the

very important respect. The X atom has valence 5 while Si has valence 4. Of the five
impurities is random throughout the lattice. But their presence affects the solid in one

electrons of X, four participate in the tetrahedral bond of Si, as shown in Fig. 1.4. The
fifth electron cannot enter the bond, which is now saturated, and hence this electron
detaches from the impurity and is free to migrate through the crystal as a conduction

X / (since it has lost one of its electrons), and thus it tends to capture the free electron,
electron, i.e., the electron enters the CB. The impurity is now actually a positive ion,

the electron in most circumstances. The net result is that the X impurities contribute
but we shall show shortly that the attraction force is very weak, and not enough to capture

171
electrons to the CB of the semiconductors, and for this reason these impurities are called
donors. Note that the electrons have been created without the generation of holes.

Fig.1.4: An As impurity in a Si crystal. The extra electron migrates


through the crystal.

When an electron is captured by an ionized donor, it orbits around the donor much like
the situation in hydrogen. We can calculate the binding energy by using the familiar Bohr
model. However, we must take into account the fact that the coulomb interaction here is
weakened by the screening due to the presence of the semiconductor crystal, which
serves as a medium in which both the donor and ion reside. Thus the coulomb potential is
now given by

M.
Ž :<   (1.17)
…V

where ε is the reduced dielectric constant of the medium . The dielectric constant ε = 11.7
in Si, for example, shows a substantial decrease in the interaction force. It is this
screening which is responsible for the small binding energy of the electron at the donor
site. Using this potential in the Bohr model, we find the binding energy, corresponding to the
ground state of the donor, to be

M # €
Q   E…. ð.
(1.18)

Note that binding energy of the hydrogen atom, which is equal to 13.6 eV. The binding
energy of the donor is reduced by the factor 1⁄…2 , and also by the mass factor
M ⁄ which is usually smaller than unity. Using the typical values ε ~ 10
and M ⁄ ~0.1, we find that the binding energy of the donor is about 10 of the
-3

hydrogen energy, i.e., about 0.01 eV. This is indeed the order of the observed values.
The donor level lies in the energy gap, very slightly below the conduction band, as shown
in Fig.1.5. Because the level is so close to the CB, almost all the donors are ionized at
room temperature, their electrons have been excited into CB.

172
Fig. 1.5: The donor level in a semiconductor

It is instructive to evaluate the Bohr radius of the donor electron. Straightforward


adaptation of the Bohr result leads to

Q  …

*% (1.19)

where *% is the Bohr radius, equal to 0.53 Å. The radius of the orbit is thus much larger
than *% , by a factor of 100, if we use the previous values for ε and M . A typical radius is
thus of the order of 50 Å. Since this is much greater than the inter atomic spacing, the
orbit of the electron encloses a great many host atoms, and our picture of the lattice
acting as a continuous, polarizable dielectric is thus a plausible one. Since the donors are

Typical concentrations are about 10Dr G… . But sometimes much higher concentrations
almost all ionized, the concentration of electrons is nearly equal to that of the donors.

are obtained by doping of the sample, for example, 10Df G … or even more.

3.5 Acceptors

An appropriate choice of impurity may produce holes instead of electrons. Suppose that
the Si crystal is doped with Ga impurity atoms. The Ga impurity resides at a site
previously occupied by a Si atom, but since Ga is trivalent; one of the electron bonds
remains vacant (Fig.1.6). This vacancy may be filled by an electron moving in from
another bond, resulting in a vacancy (or hole) at this latter bond. The hole is then free to
migrate throughout the crystal. In this manner, by introducing a large number of trivalent
impurities, one creates an appreciable concentration of holes, which lack electrons. The
trivalent impurity is called an acceptor, because it accepts an electron to complete its
tetrahedral bond. The acceptor is negatively charged, by virtue of the additional electron
it has entrapped. Since the resulting hole has a positive charge, it is attracted by the
acceptor. We can evaluate the binding energy of the hole at the acceptor in the same
manner followed above in the case of the donor. Again this energy is very small, of the
order of 0.01 eV. Thus essentially all the acceptors are ionized at room temperature.

173
Fig. 1.6: A Ga impurity in a Si crystal. The extra hole migrates through the crystal

The acceptor level lies in the energy gap, slightly above the edge of the VB, as shown in
Fig.1.7. This level corresponds to the hole being captured by the acceptor. When an
acceptor is ionized (an electron excited from the top of the VB to fill this hole), the hole
falls to the top of the VB, and is now a free carrier. Thus the ionization process, indicated
by upward transition of the electron on the energy scale, may be represented by a
downward transition of the hole on this scale.

Fig.1.7: The acceptor level in a semiconductor.


4.0 Conclusion
Semiconductors include a large number of substances of widely different
chemical and physical properties. The number of carriers (electrons and holes) is
an important property of a semiconductor, as this determines its electrical
conductivity.
5.0 Summary

• The best-known class of semiconductors is the Group IV (diamond, Silicon,


Germanium).
• The valence band is completely full at T = 0°K.
• Electrons at room temperature may be excited thermally from the valence band to
the next-higher band, known as the conduction band.
• The energy of the CB has the form.

ð. 7 .
N :; <  N  .
E€
• The energy of the VB

174
ðE ; E
0 :;<  0 
2M
• In an intrinsic semiconductor the number of electrons is equal to the number of
holes.

6.0 Tutor marked assignment

Q1. For the nondegenerate case where E − µ >> kT, calculate the number of
electrons per unit volume in the conduction band from the integral
í

  : < I: <5


f
D (E) is the density of states, f (E) is the Fermi function
Q2. (a) Compute the concentration of electrons and holes in an intrinsic
semiconductor â }[ at room temperature (ù =0.2eV,  = 0.01m
and  = 0.018 m).
(b) Determine the position of the Fermi.

7.0 Further readings/Reference


Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing
Company Limited, New Delhi, 1992.
Long, D., Energy bands in semiconductors, Wiley, 1968
Smith, R.A., Semiconductors, Cambridge, 1959
Wooster, W. A., A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8

175
UNIT 2: SEMICONDUCTOR STATISTICS page
1.0 Introduction - - - - - - - - - 171
2.0 Objectives - - - - - - - - 171
3.0 Definition - - - - - - - - 171
3.1 Semiconductor statistics - - - - - 171
3.1.1 Intrinsic region - - - - - - 171
31.2 extrinsic region- - - - - - - 172
4.0 Conclusion - - - - - - - - 173
5.0 Summary - - - - - - - - 173
6.0 Tutor Marked Assignment - - - - - - 173
7.0 Further Reading/References - - - - - - 173

176
1.0 Introduction

In this unit, we are going to study the concentration of the carriers both in the conduction
and valence bands and the difference between intrinsic region and the extrinsic region.

2.0 Objective
The objective of this unit is to differentiate
• the intrinsic region from
• the extrinsic region

3.0 Definition
3.1 Semiconductor statistics
Semiconductors usually contain both donors and acceptors. Electrons in the CB can be
created either by thermal excitation or by thermal ionization of the donors. Holes in the
VB may be generated by interband excitation or by thermal excitation of electrons from
the VB into the acceptor level. And in addition, electrons may fall from the donor levels
to the acceptor level. Figure 2.1 indicates these various processes.

Fig. 2.1: The various electronic processes in a semiconductor

Finding the concentrations of carriers, both electrons and holes, taking all these processes
into account, is quite complicated. We shall treat a few special cases, which are often
encountered in practice. Two regions may be distinguished, depending on the physical
parameters involved: The intrinsic and the extrinsic regions.

3.1.1. Intrinsic region


The concentration of carriers in the intrinsic region is determined primarilyby thermally
induced interband transitions. In this region n=p. The intrinsic region obtains when the
impurity doping is small. When we denote the concentrations of donors and acceptors by
Q and 8, the requirement for the validity of the intrinsic condition is

(2.1)

Since n increases rapidly with temperature, the intrinsic condition becomes more
favorable at higher temperatures. All semiconductors, in fact, become intrinsic at
sufficiently high temperatures (unless the doping is unusually high).

177
3.1.2 Extrinsic region

about 10Dr Ge… , the number of carriers supplied by the impurities is large enough to
Quite often the intrinsic condition is not satisfied. For the common dopings encountered,

change the intrinsic concentration appreciably at room temperature. The contribution of


impurities, in fact, frequently exceeds those carriers that are supplied by interband
excitation. When this is so, the sample is in the extrinsic region.

donor concentration greatly exceeds the acceptor concentration, that is, when Q „
Two different types of extrinsic regions may be distinguished. The first occurs when the

8 . In this case; the concentration of electrons may be evaluated quite readily. Since the
donor's ionization energy (i.e. the binding energy) is quite small all the donors are
essentially ionized, their electrons going into the CB. Therefore, to a good approximation,

 Q (2.2)

A semiconductor in which n >> p is called an n-type semiconductor (n for negative).

The other type of extrinsic region occurs when 8 „ Q that is, the doping is primarily
Such a sample is characterized, as we have seen, by a great concentration of electrons.

by acceptors. Using an argument similar to the above, one then has,

J  8 (2.3)

i.e., all the acceptors are ionized. Such a material is called a p-type semiconductor. It is
characterized by a preponderance of holes. In discussing ionization of donors (and
acceptors), we assumed that the temperature is sufficiently high so that all of these are
ionized. This is certainly true at room temperature. But if the temperature is progressively
lowered, a point is reached at which the thermal energy becomes too small to cause
electron excitation. In that case, the electrons fall from the CB into the donor level, and
the conductivity of the sample diminishes dramatically. This is referred to as freeze-out,

freeze-out takes place is Q ~ kT, which gives a temperature of about 100°K. The
in that the electrons are now "frozen" at their impurity sites. The temperature at which

variation of the electron concentration with temperature in an n-type sample is indicated


schematically in Fig. 2.2.

Fig.2.2: Variation of electron concentration n with temperature in an n-type


semiconductor.

178
4.0 Conclusion
Both holes and electrons contribute to conductivity.

5.0 Summary
• Thermal vibration or energy can be used to create a hole by exciting an electron
from the valence band to the conduction band.
• In an intrinsic semiconductor (undoped), the number of holes in the valence band
is equals the number of electrons in the conduction band.
• an n-type semiconductor is one characterized by a great concentration of
electrons.
• a p-type semiconductor is one characterized by a preponderance of holes.

6.0 Tutor marked assignment


Q1. Indium antimonide has ù = 0.23 eV; dielectric constant ε = 18;
electron effective mass m = 0.015 m. Calculate
e
(a) the donor ionization energy and
(b) the radius of the ground state orbit.

Q2. In a particular semiconductor there are 1013donor/cm3 with an ionization


energy Ed of 1 meV and an effective mass 0.01 m.
Estimate the concentration of conduction electrons at 4 K
What is the value of the Hall coefficient? Assume no acceptor atoms are
present and that ù „ ; .

7.0 Further readings/References


Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
Long, D., Energy bands in semiconductors, Wiley, 1968
Smith, R.A., Semiconductors, Cambridge, 1959
Wooster, W. A., A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8

179
UNIT 3: ELECTRICAL CONDUCTIVITY AND REAL SEMICONDUCTORS Page
1.0 Introduction - - - - - - - - - - 175
2.0 Objectives - - - - - - - - - 175
3.0 Definition - - - - - - - - - 175
3.1 Electrical conductivity - - - - - - 175
3.1.1 Dependence on temperature- - - - - - - 177
3.1.2 Mobility Vs Temperature- - - - - - - 178
3.2 Band structure of real semiconductors- - - - - 179
3.3 Excitons - - - - - - - - 180
4.0 Conclusion - - - - - - - - - 183
5.0 Summary - - - - - - - - - 183
6.0 Tutor Marked Assignment - - - - - - - 183
7.0 Further Reading/References - - - - - - - 183

180
1.0 Introduction
In this unit, we are going to study the electrical conductivity and mobility which are the
primary interest in semiconductors, the band structure so that the observed phenomenon
in the model structure can be used to obtain quantitative agreement between experiments
and theoretical analysis.

2.0 Objective
The objectives of this unit is to
• Understand electrical conductivity which measures both scattering and electron
concentration
• Understand electrical mobility which measures scattering
• Understand band structure of real semiconductor

3.0 Definition
Electrical conductivity is the ability of a material to conduct electrical current.

3.1 Electrical conductivity


Electrical conductivity is, of course, the quantity of primary interest in semiconductors.
Both electrons and holes contribute to electric current. Assume first that a sample is
strongly n-type and contains only one type of carrier: electrons. The conductivity can be
treated according to the free- electron model:

M . N€
áM  €
(3.1)

where M is an effective mass and 'M is the lifetime of the electron. To estimate the value
for áM , we substitute n =10Ds Ge… , which is eight orders in magnitude less than that in
metals, and M = 0.lm. This leads to áM ~10eu :8 ã ˆ G<eD which is a typical

value in a typical metal, where áM ~ 1:8 ã ˆ G<eD the conductivity in a


figure in semiconductors. Although this is many orders of magnitude smaller than the

physicists often use another transport coefficient: mobility. The mobility 8M is defined
semiconductor is still sufficiently large for practical applications. Semiconductor

as the proportionality coefficient between the electron drift velocity and the applied
electric field, i.e.

|V@ |  µ@ E (3.2)

Where |V@ | is the absolute value of the velocity. Taking into account that
BM   M ŠM and BM  áM K we find that
'
8M 

(3.3)
As defined, the mobility is a measure of the rapidity of the motion of the electron in the
field. The longer the lifetime of the electron and the smaller its mass, the higher the
mobility. We can now express electrical conductivity in terms of mobility. We can write

181
áM  8M (3.4)

Indicating that áM is proportional to 8 . A typical value for 8M may be obtained by


substituting áM = :8 ã ˆ G<eD and  10Ds Ge… in Eq. (3.4). This yield

8M ~10…G… Ž1 X1 (3.5)

What we have said about electrons in a strongly n-type substance can be carried over to
a discussion of holes in a strongly p-type substance. The conductivity of the holes is
given by

á6   J86
OM . N€
P
(3.6)

where 86 is the hole mobility.

Let us now treat the general case, in which both electrons and holes are present. When a
field is applied, electrons drift opposite to the field and holes drift in the same direction as
the field. The currents and conductivities of the two carriers are both additive. Therefore

á  áM  á6 (3.7)

i.e., both electrons and holes contribute to the currents. In terms of the mobilities, one
may write

á  8M  J86 (3.8)

The carriers' concentrations n and p may be different if the sample is doped, as discussed

semiconductor is  ã J  Y¬J. When the substance is in the intrinsic region,


before. And one or the other of the carriers may dominate, depending on whether the

however, n = p, and Eq. (3.8) becomes

á  :8M  86 < (3.9)

where n is the intrinsic concentration. Even now the two carriers do not contribute
equally to the current. The carrier with the greater mobility usually the electron
contributes the larger share.

182
3.1.1 Dependence on temperature

Conductivity depends on temperature, and this dependence is often pronounced. Consider


a semiconductor in the intrinsic region. Its conductivity is expressed by (3.9). But in this
situation the concentration n increases exponentially with temperature, as may be recalled
from Eq. (1.16). We may write the conductivity in the form

σ=F:T< 2e „ ⁄E7
3 (3.10)

where F(T) is a function which depends only weakly on the temperature. (This function
depends on the mobilities and effective masses of the carriers.) Thus conductivity
increases exponentially with temperature as shown in Fig.3.1.

Fig. 3.1: Conductivity of Si versus 1/T in the intrinsic range.

This result can be used to determine the energy gaps in semiconductors. In the early days
of semiconductor this was the standard procedure for finding the energy gap. Nowadays,
however, the gap is often measured by optical methods. When the substance is not in the
intrinsic region, its conductivity is given by the general expression (3.8). In that case the
temperature dependence of the conductivity on T is not usually as strong as indicated
above. To see the reason for this, suppose that the substance is extrinsic and strongly n-

á  8M
type. The conductivity is
(3.11)

But the electron concentration n is now a constant equal to Q , the donor (hole)
concentration. And any temperature dependence present must be due to the mobility of
electrons or holes.

183
3.1.2 Mobility versus temperature
Mobility of electrons (or holes) varies with temperature. In n-type semiconductor

MN€ M €
8M   (3.12)
€ € ¶€

Since the lifetime of the electron, or its collision time, varies with temperature, its
mobility also varies with temperature. Normally, both lifetime and mobility diminish as
the temperature rises. The relaxation time is given by '  /Ž where M is the mean
free path of the electron and ŽM is the drift velocity. The velocity of electrons is different
depending on their location in the conduction band. Electrons at the bottom of the
conduction band in a semiconductor obey the classical statistics and not the highly
degenerate Fermi statistics prevailing in metals. The higher electrons are in the band, the
greater their velocity. We can evaluate the conductivity by assuming that ŽM is the average
velocity. The average velocity can be estimated using the procedure of the kinetic
theory of gases:

1)  Ž E  3) ;
2 M M 2 (3.13)

-1/2
This introduces a factor of T dependence in the mobility:
M €
8M  -/. (3.14)
€ :…7
<-/.

The mean free path M also depends on the temperature, and in much the same way as it
does in metals. M is determined by the various collision mechanisms acting on the
electrons. These mechanisms are the collisions of electrons with thermally excited
phonons and collisions with impurities. At high temperatures, at which collisions with

M Þ  eD . In that case, mobility varies as 8M Þ  …⁄E . Figure 3.2 shows this for î.
phonons is the dominant factor, le is inversely proportional to temperature, that is,

Another important scattering mechanism in semiconductors is that of ionized impurities.


When a substance is doped the donors (or acceptors) lose their electrons (or holes) to
the conduction band. The impurities are thus ionized, and are quite effective in scattering
the electrons (holes). At high temperatures this scattering is masked by the much stronger
phonon mechanism, but at low temperatures this latter mechanism becomes weak and the
ionized-impurity scattering gradually takes over.

184
Fig.3.2: Electron mobility versus T in î. The dashed curve represents pure phonon
scattering; numbers in parentheses refer to donor concentrations.

3.2 Band structure of real semiconductor

So far, we have assumed the simplest possible band structure, namely, a conduction
band of a standard form, centered at the origin, k = 0, and a valence band of a standard
inverted form, also centered at the origin. Such a simple structure is applicable for
elucidating many observed phenomena, but it does not represent the actual band
structures of many common semiconductors. Only when one uses the actual band
structure is it possible to obtain a quantitative agreement between experiments and
theoretical analysis.
A material whose band structure comes close to the ideal structure is î*X (Fig. 3.3).
The conduction band has a minimum at the origin k = 0 and the region close to the origin
ð. 7 .
is well represented by quadratic energy dependence, :;<  , where me = 0.072 m.

Since the electrons are most likely to populate this region, one can represent this band by
a single effective mass. Note, however, that as k increases, the energy E(k) is no longer
quadratic in k, and those states may no longer by represented by a single, unique
effective mass. In particular that the next-higher energy minimum occurs along the [100]
direction. The dependence of energy on kin the neighborhood of this secondary minimum
is quadratic, and hence an effective mass may be defined locally, but its value is much
greater than that of the primary minimum (at the center). The actual value is 0.36 m.
Due to cubic symmetry there are six equivalent secondary minima, or valleys, in all
along the [100] directions.

185
Fig. 3.3 Band structure of GaAs plotted along the [100] and [111] directions.

These secondary valleys do not play any role under most circumstances, since the
electrons usually occupy only the central or primary valley. In such situations, these
secondary valleys may be disregarded altogether. There are also other secondary
valleys in the [111] directions, as shown in Fig. 3.3. These are higher than the [100]
valleys, and hence are even less likely to be populated by electrons. The valence band is
also illustrated inFig.3.3. Here it is composed of three closely spaced subbands. Because
the curvatures of the bands are different, so are the effective masses of the corresponding
holes .One speaks of light holes and heavy holes. Other III- V semiconductors have band
structures quite similar to that of î*X.

Figure 3.4a shows the band structure of Si. An interesting feature is that the conduction
band has its lowest (primary) minimum not at k=0. The minimum lies along the [100]
direction, at about 0.85 the distance from the center to the edge of the zone. Note that
the bottom of the conduction does not lie directly above the top of the valence band. This
type of semiconductors is known as indirect gap semiconductors. These should be
distinguished from direct gap semiconductors such as î*s. Because of the cubic
symmetry, there are actually six equivalent primary valleys located along the [100]
directions. These are illustrated in Fig. 3.4b. The energy surfaces at these valleys are
composed of elongated ellipsoidal surfaces of revolution, whose axes of symmetry are
along the [100] directions. There are two different effective masses which correspond to
these surfaces: the longitudinal and the transverse effective masses. The longitudinal
mass is  = 0.97m, while the two identical transverse masses are T = 0.19m. The mass
anisotropy ratio is about 5. The valence band in silicon is represented by three different
holes (Fig.3.4a). One of the holes is heavy (6 = 0.5m), and the other two are light. The
energy gap in Si, from the top of the valence band to the bottom of the conduction band,
is equal to 1.08 eV. The fact that the bottom of the conduction does not lie directly above
the top of the valence band, is irrelevant to the definition of the band gap.

3.3 Excitons
An electron and a hole may be bound together by their attractive coulomb
interaction, just as an electron is bound to a proton to form a neutral
186
hydrogen atom. The bound electron-hole pair is called an Excitons, Fig.3.5.
Excitons can move through the crystal and transport energy; it does not transport charge
because it is electrically neutral. It is similar to positronium, which is formed from an
electron and a positron. Excitons can be formed in every insulating crystal. All Excitons
are unstable with respect to the ultimate recombination process in which the electron
drops into the hole. The binding energy of the Excitons can be measured by optical
transitions from the valence band, by the difference between the energy required to create
an Excitons and the energy to create a free electron and free hole, Fig.3.6.

Fig.3.4 (a) Band structure of Si plotted along the [100] and [111] directions, (b)
Ellipsoidal energy surfaces corresponding to primary valleys along the [100] directions
(After Kittel, 1979)
.

Fig.3.5: An Excitons, a bound electron-hole pair.

187
Fig.3.6: Energy levels of Excitons.

Energy levels of Excitons can be calculated as follows. Consider an electron in the


conduction band and a hole in the valence band. The electron and hole attract each other
by the Coulomb potential

M.
Ž :<   § (3.15)

where r is the distance between the particles and ε is the appropriate dielectric constant.
There will be bound states of the Excitons system having total energies lower than the
bottom of the conduction band. The problem is the hydrogen atom problem if the energy
surfaces for the electron and hole are spherical and nondegenerate. The energy levels are
given by

M #
  N  (3.16)
E‹ . ð. .

Here n is the principal quantum number and µ is the reduced mass:

 
D D D
€ P
(3.17)


formed from the effective masses of the electron and hole. The Excitons ground state
energy is obtained on setting n = 1 in Eq. (316); this is the ionization energy of the
Excitons.

At room temperature, ;  ⁄  26 Ž. A sample of cadmium sulfide displays a mobile


Worked example:

carrier density of 10DÄ cm-3 and a mobility coefficient 8  10E GE ⁄¯ã Y XG
(a) Calculate the electrical conductivity of this sample
(b) If the charge carriers have an effective mass equal to 0.1 times the mass of a
free electron, what is the average time between successive scatterings
Solutions:
(a) From Eq. (3.4), the electrical conductivity in terms of mobility is given by
188
áM  8M
With  10EE e… ,   1.6 Œ 10eD , 8M  10eE E Ž eD X eD ,

we have áM  16ΩeD eD

á  p , where p gX the effective mass of an electron is, then the average


M . N
(b) From Eq.(3.6), the free electron model of metals gives

time between successive scattering is

0.1á 
' F H  5.7 Œ 10eDr X
 
4.0 Conclusion
The number of carriers (electrons and holes) is an important property of a semiconductor,
as this determines its electrical conductivity. Both conductivity and mobility (a measure
of the rapidity of the motion of the electron in the field) depend on temperature.

5.0 Summary

áM 
M . N€

• defines electrical conductivity according to free electron model.

8M  '

electrical conductivity in terms of mobility is defined as áM  8M
• defines mobility

• á  8M  J86 defines contribution to the currents by both electrons and

A material whose band structure comes close to the ideal structure is GaAs
holes in terms of the mobilities

• The bound electron-hole pair is called an Excitons

6.0 Tutor marked Assignment


Q1. A sample of Si contains 10–4 atomic per cent of phosphorous donors that
are all singly ionized at room temperature. The electron mobility is 0.15
m2V– 1s–1. Calculate the extrinsic resistivity of the sample (for Si, atomic

Given the data for Si: 8M = 1350 cm /Vˆs, 86 = 475 cm /Vˆs, M = 0.19m,
weight = 28, density = 2300 kg/m3).
2 2

6 = 0.16m and ù = 1.1 eV, calculate


Q2.

(a) The lifetimes of electrons and holes.


(b) The intrinsic conductivity σ at room temperature

7.0 Further readings/References


Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.
Long, D., Energy bands in semiconductors, Wiley, 1968
Smith, R.A., Semiconductors, Cambridge, 1959

189
Wooster, W. A., A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.
www.pa.uk.edu/kwng.phy/525/lec/lecture-8

190
UNIT 4: SUPER CONDUCTIVITY (I): THE BASIC PHENOMENON page
1.0 Introduction - - - - - - - - - 186
2.0 Objectives - - - - - - - - 186
3.0 Definition - - - - - - - - 186
3.1 Empirical criteria - - - - - 187
3.2 Transition Temperature - - - - - 188
3.3 Energy Gap - - - - - - 188
3.4 Properties dependent on Energy gap- - - - 189
3.4.1 Microwave and Infrared absorption - - - 189
3.4.2 Density of states- - - - - - - 189
3.4.3 Specific Heat - - - - - - - 190
3.4.4 Acoustic attenuation - - - - - - 191
3.4.5 Thermal Conductivity - - - - - 192
4.0 Conclusion - - - - - - - - 193
5.0 Summary - - - - - - - - 193
6.0 Tutor Marked Assignment - - - - - - 193
7.0 Further Reading/References - - - - - - 193

191
1.0 Introduction
Superconductivity was first discovered and so named by Kamerlingh Onnes in 1911. In the
course of an investigation of the electrical resistance of various metals at liquid helium
temperatures, he observed that the resistance of a sample of mercury dropped from 0.08 Ω
at about 4 K to less than 3 x l 0 - 6 Ω over a temperature interval of 0.01K. Subsequent
attempts showed that the width of the transition region in a particular specimen depends
on a number of factors, such as the purity and metallurgical history and can be as sharp as
one millidegree or spread over several degrees. While the breadth of the transition may
increase if the sample is metallurgically imperfect, the extraordinary smallness of the
resistance in the superconducting state appears to hold for all superconductors. Thus, the
first characteristic property of a superconductor is that its electrical resistance, for all
practical purposes, is zero, below a well-defined temperature Tc, called the critical, or
transition temperature. Thus, the conductivity in this range of temperature is infinite; hence
the nomenclature of superconductivity.
Figure 4.1 shows how the electrical resistivity in a superconductor becomes
immeasurably small at the transition temperature. The figure also contrasts the behaviour
of a normal metal for which at very low temperatures, the remanent resistivity is
characteristic of residual impurities. The resistance of a superconductor is believed to be
zero rather than just very small.

Normal metal

Superconductor
SUPERCONDUCTOR

Temperature T (K)

Fig.4.1: Temperature dependence of the resistance: of a normal


and superconducting material (After Kachhava, 1992)

2.0 Objective
The objective of this unit is to revise the basics of Superconductors in terms of:
• Empirical criteria
• Transition temperature
• Energy gap

3.0 Definition
Superconductivity is the phenomenon on which the electrical resistivity of metals or
alloys drop to zero (infinite conductivity) when cooled into its critical temperature.

192
3.1 Empirical criteria
There are found to be a number of regularities in the appearance of superconductivity,
the principal of which are the following:
I. Superconductivity has been observed only for those metallic substances for which
the number of valence electrons Z lies between 2 and 8.
II. In all cases involving transition metals, the variation of Tc with number of valence
electrons shows sharp maxima for Z = 3, 5 and 7, as shown in Fig. 4.2.
III. A rather striking correlation (a straight line graph) exists between 3 and Z2 for
elements along given rows of periodic table (Fig. 4.3).
IV. For a given value of Z, certain crystal structures seem more favourable than others.
For example, β-tungsten and α-manganese structure are conductive to the
phenomenon of superconductivity.
V. Ferromagnetic and ferroelectric ordering are found to inhibit superconductivity.
VI. Tc increases with a high power of the atomic volume and inversely as the atomic
mass.
VII. Superconductivity occurs in materials having high normal resistivities. The condition
n p > 106 is a good criterion for the existence of superconductivity, where n is the
number of valence electrons per c.c. and p is the resistivity in electrostatic units at
20°C.
These empirical rules have played an important role in the discovery of new
superconductors.
Tc

0 2 4 8 10
Z
Fig 4.2: Variation of transition temperature with
number of valence electron

Fig 4.3: Empirical correlation between transition


temperature and Z2(After Kachhava, 1992)
193
3.2 Transition Temperature

The temperature at which the normal metal passes into superconducting state is called the
transition temperature, Tc. The transition temperature is generally affected by the application of
pressure though no specific regularity in the behaviour has been found. The value of Tc for most of
the metals lies below 4K; e.g., for Al, it is 1.20 K. For C-15 structure (e.g., V2 Hf), it is 10K; for B-
1 structure (e.g., NbN), it is near 13 K, whereas NbZr and NbT1 [BCC (A-2) structure] have the
values of Tc as 11.0 and 10.0 K respectively. For A-15 structure, the highest Tc = 23.2 K has been
observed in NB3Ge.

3,3 Energy Gap


Experiments have shown that in superconductors, for temperatures in the vicinity of
absolute zero, a forbidden energy gap just above the Fermi level is observed. Figure 4.4(a)
shows the conduction band in the normal state, while (b) depicts an energy gap equal to 2∆
at the Fermi level in the superconducting state. Thus, the Fermi level in a superconductor is
midway between the ground state and the first excited state so that each lies an energy dis-
tance = away from the Fermi level. Electrons in excited states above the gap behave as
normal electrons. At absolute zero, there are no electrons above the gap. 2∆is typically of
the order of 10- 4 eV.

Normal Superconducting
(a) (b)
Fig.4.4: (a) Conduction band in the normal metal (b) Energy gap at
the Fermi level in the superconducting state (After Kachhava,
1992)

∆ is found to be a function of temperature T. Thus, (T) represents energy gap at temperature T.


Figure 4.5 shows reduced values of observed energy gap (T)/ (0) as a function of the
reduced temperature T/Tc. Elementary theory predicts that

(4.1)

We observe that the energy gap decreases continuously to zero as the temperature is increased to
Tc . Numerically, experiments show that for most of the metals. The transition from the
superconducting state to the normal state is observed to be a second-order phase transition.
In such a transition, there is no latent heat, but there is a discontinuity in the heat
capacity.

194
02 <K 06 O8 10
T/Tc———>-
Fig. 4.5: Temperature dependence of the
superconducting energy gap (After Kachhava, 1992)

3.4. Properties Dependent on Energy Gap

3.4.1 Microwave and Infrared Absorptions


The response of a metal to electromagnetic radiation is determined by the frequency
dependent conductivity. This in turn depends on the available mechanisms for energy
absorption by the conduction electrons at the given frequency. Because the electronic
excitation spectrum in the superconducting state is characterized by an energy gap , one
would expect the AC conductivity to differ substantially from its normal state form at
frequencies small compared with , and to be essentially the same in the
superconducting and normal states at frequencies large compared with The value of
, is typically in the range between microwave and infrared frequencies. In the
superconducting state, an AC behaviour is observed which is indistinguishable from that in
the normal state at optical frequencies. Deviations from normal state behaviour first
appear in the infrared, and only at microwave frequencies does AC behaviour fully
displaying the lack of electronic absorption characteristic of an energy gap becomes
completely developed.

3.4.2 Density of States


The three parts of Fig. 4.6 give a highly exaggerated picture of the difference between the
spectrum and occupancy of states in a normal metal and those in a superconductor. Part
(a) considers the density of states at T = 0 in the absence of superconductivity (which can
be arranged by applying a suitable magnetic field). The superconducting ground state for
zero temperature is pictured in part (b). This shows a zero density of states for energies within
± on either side of the Fermi energy, and a piling up of the displaced states on either
side of the gap. At T — 0, no electrons are excited to higher states. Part (c) of the figure
imagines the consequences of a finite temperature less than Tc .The superconducting energy
gap is now smaller than . Fractions of number of electrons are in states above
leaving behind some unoccupied states below . Finally, the gap decreases to
zero when T reaches Tc and the corresponding density of states is the one depicted in part (a).

195
a

Fig. 4.6: Density and occupancy of states (D.O.S) for a normal and a
superconductor (After Kachhava, 1992)

3.4.3 Specific Heat


There is no heat of transformation associated with the superconducting-normal transition
in a metal, but there is an anomaly in the electronic component of the specific heat. An
example of this is illustrated in Fig. 4.7. The discontinuity in the specific heat reflects the
second-order transition from a relatively disordered (normal) state to a more highly ordered
(superconducting) state of lower entropy. At low temperatures, the specific heat of a normal
metal has the form

(4.2)

where the linear term is due to electronic excitations- and the cubic term is due to lattice
vibrations. Below the superconducting critical temperature, this' behaviour is substantially
altered. As the temperature drops below Tc, the specific heat jumps to a higher value and then
slowly decreases, eventually falling well below the value one would expect for a normal metal.
By applying a magnetic field to drive the metal into the normal state, one can compare the
specific heats of the superconducting and normal states below the critical temperature.

196
Fig.4.7: Specific heat of normal and superconductor (After Kachhava, 1992)

Such an analysis reveals that in the superconducting state, the linear electronic contribution
to the specific heat is replaced by term that vanishes much more rapidly at very low
temperatures, having dominant low-temperature behaviour of the form exp:∆⁄;  <. This
is the characteristic thermal behaviour of a system whose excited levels are separated from the
ground state by energy 2∆, thus, the total specific heat of the superconducting state is
(4.3)

Where

(4.4)

where is the low-temperature electronic specific heat of the normal state (obtained by
applying suitable magnetic field), and and .These parameters are themselves
weakly temperature dependent. In Fig.4.7 the size of the discontinuity in specific heat at T =Tc
is 2.5 in units of .The exponential decrease in specific heat below Tc can be interpreted as
follows. Because, of the energy gap, the number of electrons excited across the gap is given
roughly by a Boltzmann factor, exp:∆⁄;  <. Hence, the heat capacity varies exponentially
with temperature.

3.4.4 Acoustic Attenuation


When a sound wave propagates through a metal, the microscopic electric fields due to the
displacement of the ions can impart energy to electron near the Fermi level, thereby
removing energy from the wave. This is expressed by the attenuation coefficient, α, of
acoustic waves. The ratio of α for superconducting and normal state is given by

(4.5)

At low temperatures

(4.6)
197
The exponential decay ratio is represented in Fig. 4.8

Fig.4.8: Ratio of attenuation coefficients for acoustic waves


in superconducting and normal metal as a function of
temperature (After Kachhava, 1992).

3.4.5 Thermal Conductivity


In normal metals, the heat current is predominantly carried by the conduction electrons
and at low temperatures, the electronic contribution to the thermal conductivity Ken is
given by the Wiedemann-Franz law. In a superconductor, however, the electron pairs
have zero energy so they cannot contribute to energy transport and hence to the heat
current (but being charged, they can still contribute to the electric current). Hence, the
electronic contribution to the heat current depends on the number of normal electrons and
like the electronic specific heat represented by Eq. (4.4), we have the ratio of
superconducting to normal phase conductivities as

(4.7)

This is illustrated in Fig.4.9. When T , and the only thermal current will be
carried by the phonons (as in insulator). Under suitable conditions, may be very large

( ) and this property can be used to make a heat switch, the heat flow being
controlled by a magnet. The phonon contribution to thermal conduction will actually
increase in the superconducting state since the scattering of phonons by electrons is
reduced by the formation of pairs. In extreme cases when is made small by the
introduction of impurities, the increase in the phonon contribution to the thermal
conductivity below may outweigh the reduction in the electronic contribution so that
the total conductivity increases in the superconducting state. To achieve this condition, an
impurity of similar mass but different valence, which will reduce without greatly
affecting phonon transport, should be used. An example is Bi in Pb.

198
Fig.4.9: Ratio of the electronic contribution to the thermal
conduction of Al (After Kachhava, 1992)

4.0 Conclusion
At a critical temperature , many metals and alloy undergo a phase transition from a
state of normal electrical resistivity to a superconducting state.

5.0 Summary
• Superconductivity has been observed only for those metallic substances for which the
number of valence electrons Z lies between 2 and 8.
• The temperature at which the normal metal passes into superconducting state is called
the transition temperature, Tc
• In superconductors, for temperatures in the vicinity of absolute zero, a forbidden
energy gap just above the Fermi level is observed.
• The ratio of attenuating coefficient for superconducting and normal state is given by

• The ratio of superconducting to normal phase conductivities is given as

6.0 Tutor marked assignment


Q1. Prove that the Meissner effect is consistent with the disappearance of
resistivity in a super conductor.
Q2. Show that when superconductivity is destroyed by the of a magnetic
field, the magnet will cool.

7.0 Further readings/References


James, D.P., Bernard, C.B., Solid state physics: introduction to the theory,
Springer, 2005
Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992.

199
Wooster, W. A., A textbook on crystal physics, Cambridge University Press,
Cambridge, 1938.

200
UNIT 5: SUPERCONDUCTIVITY (II): EXPERIMENTS AND THOERIES page
1.0 Introduction - - - - - - - - - 196
2.0 Objectives - - - - - - - - 196
3.0 Definition - - - - - - - - 196
3.1 Meissner effect - - - - - 196
3.2 Critical field - - - - - 197
3.3 Type I & Type II Semiconductors - - - - 197
3.4 Critical Currents - - - - 198
3.5 London equation - - - 199
3.6 Thermodynamics of Superconducting transition - - 203
3.7 Isotope effect - - - - - - - 205
4.0 Conclusion - - - - - - - - 206
5.0 Summary - - - - - - - - 206
6.0 Tutor Marked Assignment - - - - - - 207
7.0 Further Reading/References - - - - - - 207

201
1.0 Introduction
In this unit, we are going to study both the experimental and theoretical situations
concerning superconductivity. The experimental survey includes the effects of magnetic
field on superconductivity(the Meissner effect), the minimum magnetic field (critical
field) necessary to destroy superconductivity as well as the minimum current (critical
current) that can be passed without destroying superconductivity. Thermodynamics,
London equation and type I and II of semiconductors constitute the theoretical surveys.

2.0 Objective
The objectives of this unit are:
• To survey the central experimental facts concerning superconductivity
• To discuss the theoretical situations of superconductivity
3.0 Definition
Superconductivity is the phenomenon on which the electrical resistivity of metals or
alloys drop to zero (infinite conductivity) when cooled into its critical temperature.
3.1 Meissner effect
Meissner and Ochsenfeld (1933) showed that, if a long superconductor is cooled in a
longitudinal magnetic field from above the transition temperature, the lines of induction
are pushed out (Fig. 5.1) at the transition. The Meissner effect shows that a super-
conductor behaves as if inside the specimen B = 0 or ; that is, a
superconductor exhibits perfect diamagnetism. This very important result cannot be
derived merely from the characterization of a superconductor as a medium of zero
resistivity : from E = j we see that, if is zero while j is finite, then E must be zero and
with it curl E must be zero. Therefore from Maxwell's equations

(5.1)

so that the flux through the metal cannot change on cooling through the transition. The
Meissner effect contradicts this result and suggests that perfect diamagnetism and zero
resistivity are two independent essential properties of the superconducting state.

Fig.5.1: Meissner effect in a sphere cooled in a constant applied magnetic


field; on passing below the transition temperature the lines of induction are
ejected from the sphere. (After Kittel,

202
3.2 Critical Field
The minimum applied magnetic field necessary to destroy superconductivity and restore
the normal resistivity is called the critical field, . depends on the temperature.
Fig.5.2 shows the critical field as a function of temperature. The curve is nearly parabolic
and can be reasonably well represented by the relation

(5.2)

Where is the critical field at absolute zero. This equation is really the equation of
phase boundary between the normal and superconducting state. The typical value of is
5000A/m.

Fig.5.2: Critical magnetic field as a function of temperature (After


James and Bernard, 2005).

3.3 Type I and Type II Superconductors


Superconductors may be divided into two classes which depend on the way in which the
transition from the superconducting to the normal state proceeds when the applied field
exceeds +N . In type-I materials, as +N is reached entire specimen enters the normal state

PTMV8 
practically simultaneously, the resistance returns, the diamagnetic moment becomes zero and
M¡TMV8 (Fig5.2a).

a b
Fig.5.2: Flux penetration as a function of magnetic filed in
(a) type-I superconductor and (b) type-II superconductor
203
In type-II superconductors, the transition to a completely normal specimen is much more
gradual. As shown in Fig. 5.2b, there is a partial penetration of the magnetic field between the
critical field Hc1 and Hc2. Small surface super currents may still flow up to an applied field Hc3.

3.4 Critical Currents


The minimum current that can be passed in a sample without destroying its superconductivity
is called critical current . If a wire (radius r) of a type-I superconductor carries a current I,
there is a surface magnetic field HI= I/2πr associated with the current. If HI exceeds , the
material will go normal. If in addition, a transverse magnetic field H is applied to the wire,
the condition for the transition to the normal state at the surface is that the sum of the
applied field and the field due to the current should equal the critical field. Thus, as seen
from Fig. (5.3b), we have

Hence (5. 3)

The critical current will decrease linearly with increase of the applied field until it reaches
zero at +  +N ⁄2. If the applied field is zero, similar considerations apply to
type-II superconductor for that is when the superconductor is not in the mixed state.

Fig.5.3 :(a) wire carrying current I subjected to transverse field H.(b) Cross-
section of wire showing fields at equatorial position on the surface(After
Kachhava,1990)

204
3.5 London Equations
In 1935, two brothers F. and H. London, proposed two equations to govern the
microscopic electric and magnetic fields (two basic electrodynamics properties) which
give superconductivity its unique interest. The London theory is based on rather old ideas
of the two –fluid model according to which a superconductor can be thought to be
composed of both normal and superfluid electrons. Let , be respectively
the density, and velocity of the normal and superfluid electrons. If is the number of
electrons per unit volume, then on the average

The equation of motion for the superfluid electrons is

(5.4)

The density of the superfluid electrons is

(5.5)

Then Eq. (5.4) and (5.5) yield

(5.6)

This is the first London equation.


Taking curl of Eq. (5.6)

and using Maxwell’s equation

we get
= (5.7)

Integrating this equation with respect to time, and choosing the constant of integration to
be zero consistent with the Meissner effect, we have

= (5.8)

This is the second London equation.

205
We may derive the Meissner effect from the second London equation by using the
Maxwell equation

= (5.9)

Taking curl of this equation

B= (5.10)

Then using the condition div B = 0 of a superconductor in the identity

B = grad div B B

We get
B= (5.11)

On combining Eq. (5.10) and (5.11),

(5.12)

This along with Eq. (5.8) gives

(5.13)

Where is called the London penetration depth and is defined by

(5.14)

For a superconductor to the right of the plane x = 0, Eq. (5.13) has the solution

(5.15)

This equation indicates that B does not penetrate very deeply into superconductor, and
therefore it implies the Meissner effect. The field penetrates only a distance within the
surface. is typically of the order of 1000Å. The graphical form of Eq. (5.15) is shown
in Fig.5.4. The penetration depth is also found to depend strongly on temperature and to
become much larger as T approaches . The observation can be fitted extremely well by
a simple expression of the form

(5.16)

This equation implies that


206
(5.17)

Fig.5.4: Magnetic field penetration at surface of a superconductor (After


Kachhava, 1992).

The density superconducting electrons increase from zero at to at absolute zero as


shown in Fig.5.5, which also depicts the temperature variation of . is called the order
parameter because it characterizes the order in the superconducting state.

Fig.5.5: Density of superconducting electrons as a function of temperature


(After Kachhava, 1992)

207
Worked example:

The London equation for simple superconductor is a phenomenological equation relating


the supercurrent B\ to the magnetic vector potential A:
e€ M .
B\  –
€ N

Where M is the electron mass. Using the appropriate Maxwell equation, show how the
above equation leads to Meissner effect.

Solution:

The Meissner effect refers to the fact that in the superconducting state magnetic induction
vanishes and materials become strongly diamagnetic. From London equation (Eq.58),

= (i)

Since hE 

 M .
, we get

=  Ž? ˜
=
(ii)

Inside a superconductor, the electrical field vanishes and we have the Maxwell equation

ã! ãã;*; ãY 5Ig 5. h Œ ˜ = B\


`a

˜  GhE Ô Œ B\   œÔ:Ô · ˜<  h ? ˜ž,


? Ž?

Hence

Or, using Maxwell’s equation ÔE ˜  ˜


D
ƒ.

D⁄E
Where h  F H

  M.

For a superconductor to the right of x= 0, Eq. (ii) has the solution

˜  ˜z
Ì
Fe H
‘

This shows that B decays exponentially such that ˜  ˜% at x = h.


D
M

For › „ h, Bä 0, indicating that the magnetic field exists only in a thin layer of
thicknessY h beneath the surface of the superconductor. Thus the magnetic field inside a
superconductor is zero. This is the Meissner effect.

208
3.6 Thermodynamics of Superconducting transition
It has been demonstrated experimentally that the transition between the normal and
superconducting states is thermodynamically reversible, in the same sense that with slow
evaporation the transition between liquid and vapor phases of a substance is reversible.
The Meissner effect also suggests that the transition is reversible and would not subsist if
the superconducting currents die away with the production of Joule heat when
superconductivity is destroyed. As the transition is reversible we may apply
thermodynamics to the transition, obtaining an expression for the entropy difference
between normal and superconducting states in terms of the critical field curve versus
.
The Gibbs free energy per unit volume in a magnetic field

(5.18)
Then the differential Gibbs free energy is

(5.19)
At constant T and P, the free energy difference, because of the presence of a magnetic
field, is found by integration. Thus

(5.20)

(5.21)

For superconductor, or and

(5.22)
Here is the free energy of a superconducting phase
Along the phase boundary between normal and superconducting state, the normal phase
must have a free energy indistinguishable from that of the superconducting phase.
Therefore

(5.23)

Where is the free energy of the normal phase. Fig.5.6 shows the variation of and
below , where the normal phase is obtained by applying the field in excess of .

209
Fig.5.6: Experimental values of free energy of Al in the normal
and superconducting state as a function of temperature (After
Kachhava, 1992)

Let us now calculate the difference in entropy of the two phases. For solids, the entropy S
is given by . Hence, differentiating Eq. (5.23) with respect to T, we have

= (5.24)
Where the entropies and refer to normal and superconducting phases respectively.
Thus as illustrated in Fig.5.7.

Fig.5.7: Entropy S of Al in the normal and superconducting state


as a function of temperature (After Kachhava, 1992).

As is always negative, is always positive and the superconducting state is


observed to be more ordered than the normal state. At the transition temperature
= 0 because = 0, and at 0K, = 0 from the third law of thermodynamics, which
is satisfied, because tends to zero. At some intermediate temperatures, has a
maximum. The latent heat absorbed when superconductivity is destroyed is

210
( )

= (5.25)

In the absence of a magnetic field, the transition occurs at and the latent heat is zero. If
and are respectively the normal and superconducting state internal energies, then
from Eq. (5.25)

= ( )

= (5.26)

From experiment, ) , which is extremely small compared to the band


energies. For a unit volume, the difference of the of the heat capacities, from Eq. (5.26),
will be

) = )

= (5.27)

On substituting , = 0 in this equation, we get the Rugers formula

) = (5.28)
This equation reproduces the experimental data very well.

3.7 Isotope effect

It has been found by early experimentalists that the transition temperature is strongly
dependent on the average isotopic mass, M, of the constituents of a superconductor. In
particular

(5.29)

More recent experiments have suggested the following general form

(5.30)
In which is called the isotope effect coefficient and is defined by

(5.31)

211
Recent theories lead to the result

(5.32)

where the parameter is the density of single states for one spin at the Fermi level
and V is the model potential between the electrons. The transition temperature can be
connected to the Debye temperature, because sound velocity . Hence,
from Eq. (5.30),

i.e. = constant (5.33)

The constant of implies that the lattice vibrations have an important bearing on
superconductivity, and gives a clear guide to the theory that electron-phonon interaction
must be the basis of the existence of superconductivity.

4.0 Conclusion

The magnetic properties exhibited by superconductors are as dramatic as their electrical


properties. The magnetic properties cannot be accounted for by the assumption that the
superconducting state is characterized properly by zero electrical resistivity.

5.0 Summary

• A bulk specimen of metal in the superconducting state exhibits perfect diamagnetism,


with the magnetic induction B = 0. This is Meissner effect.
• There are two types of superconductors, I and II
• In type I, the superconducting state is destroyed and the normal state is restored by
application of critical value .
• A type II superconductor has two critical fields
• The London 1st and 2nd equations

Or =

Leads to the Meissner effect through the penetration equation

212
6.0 Tutor marked assignment

Q1. A superconducting tin has a critical temperature of 3.7 K in zero magnetic


fields and a critical field of 0.0306 T at 0 K. Find the critical field at 2 K.
Q2. Estimate the London penetration depth from the following data:
Critical temperature = 3.7 K
Density = 7.3 g cm-3
Atomic weight = 118.7
Effective mass* = 1.9m, where m is the mass of a free electron

7.0 Further readings/References


Animalu, A.O.E, intermediate quantum theory of crystalline solids, Prentice-Hall
of India, New Delhi, 1978
Blakemore,J.S., Solid State Physics, W.B.Saunders Co.,1974
James, D.P., Bernard, C.B., Solid state physics: introduction to the theory,
Springer, 2005
Cuper, C.G., An introduction to the theory of Superconductivity, Oxford
University Press, 1968
Kittel, C., Introduction to solid state physics, Wiley Eastern Limited, 1979
Kachhava, C.M., Solid State physics, Tata McGraw-Hill Publishing Company
Limited, New Delhi, 1992

213

You might also like