Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
0% found this document useful (0 votes)
50 views

Module 1

Uploaded by

lemmefind45
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
50 views

Module 1

Uploaded by

lemmefind45
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 30

Module 1

Electrical conductivity

Classical free electron theory

Electrical conductivity in metals

Fermi Dirac distribution

Variation of Fermi function with temperature

Fermi Energy

Energy bands

Classification of materials into conductor, semiconductor and insulator.

Superconductivity

Superconductivity

Transition temperature

Critical field

Meissner effect

Type I and Type II Super conductors.

BCS Theory

Applications of superconductors.
Classical Free Electron Theory (CFET)
The Classical Free Electron Theory (CFET) was one of the earliest models proposed to explain the
electrical and thermal properties of metals. It was developed by Paul Drude in 1900 and later refined
by Hendrik Lorentz. The theory is based on the assumption that electrons in a metal behave similarly
to molecules in an ideal gas. This model laid the foundation for understanding metal conductivity, but
it also had significant limitations, particularly in light of quantum mechanics.

Key Assumptions of Classical Free Electron Theory

• Electrons in a metal are considered free to move throughout the metal, and they are not bound
to any specific atom.

• These electrons are treated as independent particles (like gas molecules in a box), and the
positive ions are treated as immobile, fixed in a periodic lattice.

• The behaviour of these free electrons is governed by Newtonian mechanics. Concepts like
velocity, momentum, and force are used to describe electron motion.

• The velocity of electrons in a metal is determined by the applied electric field and the random
thermal motion due to temperature.

• In the absence of an external electric field, electrons move randomly in all directions due to
their thermal energy. The net flow of charge (current) is zero in this case.

• Electrons collide elastically with ions and with each other. These collisions cause them to lose
energy, but between collisions, they travel in straight paths. The average time between
collisions is called the relaxation time τ.

• When an electric field E is applied to the metal, it causes the electrons to drift in the direction
opposite to the field (since electrons are negatively charged).

• Electrons not only carry charge but also transport heat. In the Drude model, the free electrons
are also responsible for the thermal conductivity of metals.

Failures of Classical Free Electron Theory

1. Over estimation of electronic contribution to heat capacity.


3
According to classical theory, the specific heat of free electrons should be 𝐾 𝑇 per electron.
2 𝐵

However, experiments show that the electronic contribution to the specific heat is much
smaller.
2. Incorrect temperature dependence of electrical conductivity.
The theory predicts that the electrical conductivity should decrease with increasing
temperature due to increased electron collisions, which aligns with experimental observations.
However, the temperature dependence predicted by the classical theory does not match
experimental results quantitatively.
3. Failed to explain magnetic properties like diamagnetism and ferromagnetism
4. Predicts incorrect values for the magnetic susceptibility of metals.
5. It cannot explain Pauli paramagnetism (where free electrons respond weakly to magnetic
fields) and other quantum effects in metals.
6. Fails to explain the quantum nature of electron behaviour, such as quantized energy levels in
solids, the existence of energy bands, and phenomena like photoelectric effect and
superconductivity.
7. Inaccurate prediction of Wiedmann-Franz law
Wiedemann-Franz law states that the ratio of thermal conductivity (κ) to electrical
conductivity (σ) is proportional to the temperature
κ/σ=LT
While CFET gets the general form of the Wiedemann-Franz law correct, it fails to explain the
correct value of the Lorenz number. The experimentally observed Lorenz number is
significantly different from the value predicted by classical theory.

Transition to Quantum Models

The Drude-Lorentz model was a stepping stone towards modern quantum theory of metals.
In 1928, Arnold Sommerfeld improved the classical theory by incorporating quantum
statistics (Fermi-Dirac statistics), leading to the development of the Quantum Free Electron
Theory. In this theory, electrons are treated as fermions obeying the Pauli exclusion
principle and moving according to quantum mechanics. The distribution of electrons is
described by the Fermi-Dirac distribution rather than the classical Maxwell-Boltzmann
distribution.

Importance of Classical Free Electron Theory

Despite of its limitations, the Classical Free Electron Theory played a crucial role in the development
of solid-state physics.

• It laid the groundwork for understanding the electrical and thermal properties of metals.

• The model was relatively simple and could provide insights into basic phenomena like
conductivity and Ohm’s law.
• It highlighted the need for quantum mechanics in explaining the behaviour of electrons in
solids, which led to more advanced theories like the Band Theory of Solids.

Terms need to familiarise


1. Electric Field (E)

The electric field is a vector field that represents the force experienced by a charge in space due to
electric forces. It is defined as the force per unit charge.

𝐸 = 𝐹/𝑒

Where:

• E is the electric field,

• F is the force experienced by the charge,

• q is the charge.

In a conductor, the electric field causes free electrons to accelerate and flow, creating an
electric current.

2. Current Density (J)

The current density J represents the amount of electric current flowing per unit area of the
conductor's cross-section.

𝐽 = 𝐼/𝐴

Where:

• J is the current density (measured in amperes per square meter, A/m²),

• I is the electric current,

• A is the cross-sectional area of the conductor.

3. Drift Velocity (vd)

The drift velocity is the average velocity attained by charged particles, such as electrons, in a
conductor when subjected to an electric field. It is the net velocity at which electrons move,
superimposed on their random thermal motion.
Drift velocity is proportional to the electric field and plays a direct role in determining the current in a
conductor.

4. Relaxation Time (τ)

The relaxation time τ is the average time interval between two successive collisions of an electron in
a conductor. After each collision, the electron loses its velocity and starts accelerating again due to the
electric field.

Relaxation time influences how freely electrons can move, and longer relaxation times lead to higher
conductivity.

5. Mean Free Path (λ)

The mean free path is the average distance an electron travels between collisions. It is related to the
relaxation time and the electron’s average velocity (typically the thermal velocity).

The mean free path gives insight into the distance an electron travels before scattering, affecting the
material's overall conductivity.

6. Mobility (μ)

The mobility μ of charge carriers (such as electrons) measures how quickly they respond to an
electric field. It is defined as the drift velocity per unit electric field:

μ= drift velocity/ E

From the drift velocity equation:

𝜇 = −𝑒𝜏/𝑚

Where:

• μ is the mobility of the electrons,

• τ is the relaxation time,

• m is the mass of the electron.

Mobility depends on the relaxation time, and higher mobility implies that the electrons can move
more freely when an electric field is applied.

7. Conductivity (σ)

Electrical conductivity σ quantifies how easily a material can conduct electric current. It depends on
the number of free electrons, their charge, and their mobility:
𝜎 = 𝑛𝑒𝜇

The higher the conductivity, the easier it is for the material to allow current flow.

8. Resistivity (ρ)

Resistivity ρ is the inverse of conductivity and represents how strongly a material opposes the flow of
electric current:

ρ=1/σ

Resistivity depends on the material's properties and temperature. Materials with low resistivity (such
as metals) are good conductors, while those with high resistivity (such as insulators) resist current
flow.

Electrical Conductivity in Metals

Conductivity is a fundamental property of materials that measures their ability to conduct electric
current. It plays a crucial role in understanding how charges move through different media and is
essential in fields such as solid-state physics, materials science, and electrical engineering. The
derivation of the equation for conductivity involves examining the microscopic behaviour of charge
carriers, such as electrons in a conductor, and their interactions within the material.

Consider a metal full of free electrons. When an electric field E is applied, each free electron
experiences a force,
𝐹 = −𝑒𝐸
Where:
• e is the charge of the electron (e=1.6×10-19C),
• E is the electric field vector.
This force can cause the electrons to accelerate. This motion of electrons will be in a direction
opposite to the applied electric field. According to Newton's second law:
𝐹 = 𝑚𝑎
So, the acceleration a is,
𝑎 = 𝐹/𝑚 = −𝑒𝐸/𝑚
Where m is the mass of the electron.
According to newtons equation of motion,
𝑣 = 𝑢 + 𝑎𝑡
𝑣𝑑 = 0 + (−𝑒𝐸/𝑚) 𝜏
𝑣𝑑 = −𝑒𝐸𝜏 /𝑚
Where, τ is the relaxation time, the average time between collisions.
The current density J is the amount of electric current flowing per unit area of the material. It is
related to the number of free electrons and their drift velocity. Mathematically, it is expressed as:
𝐽 = −𝑛𝑒𝑣𝑑

Where:
n is the number of free electrons per unit volume,
e is the charge of the electron,
vd is the drift velocity of the electrons.
Substituting the expression for vd,
𝐽 = −𝑛𝑒(−𝑒𝐸𝜏/𝑚)

ne2 τ
J= E
m
We can derive another expression for conductivity. Current density J is defined as the amount of
current per unit area
𝐼
𝐽=
𝐴
where I is the current flowing through the conductor and A is the cross-sectional area of the conductor.
Ohm's law states that the current I flowing through a conductor is proportional to the voltage V
applied across it.
𝑉
𝐼=
𝑅
where R is the resistance of the conductor.
The resistance R of a conductor is related to its resistivity ρ, length L, and cross-sectional area A by
the following formula:
𝐿
𝑅=ρ
𝐴
Substituting this into Ohm’s Law,
𝑉𝐴
𝐼=
ρ𝐿

The electric field E in a conductor is related to the applied voltage V and the length L of the
conductor:
𝑉
𝐸=
𝐿
Substituting this into the expression for current I
𝐴𝐸
𝐼=
ρ
Since current density J is 𝐼/𝐴, we can express it as,
𝐼 𝐸
𝐽= =
𝐴 ρ
Conductivity 𝜎 is defined as the reciprocal of resistivity ρ
1
σ=
ρ
Thus, the relation between current density J and the electric field E becomes:
𝐽 = σ𝐸
𝑛𝑒 2 𝜏
From part 1, current density is 𝐉 = 𝑚
𝐄. Comparing these two equations,

𝑛𝑒 2 𝜏
𝜎=
𝑚
Factors Affecting Electrical Conductivity in Metals

• Number of Free Electrons (n): Metals typically have a high density of free
electrons. Conductivity increases with the number of free electrons available to
participate in current flow. For example, silver (Ag) and copper (Cu) are highly
conductive metals due to their high electron density.
• Mean Free Path: The mean free path is the average distance an electron travels
between collisions with the atoms in the metal. At high temperature, electrons get
scattered due to the of collisions between them, between electrons and positive ions
and also due to the thermal vibration. This effect becomes more pronounced at higher
temperatures, reducing conductivity.
• Relaxation Time (τ): The relaxation time τ is the average time between two
collisions. The longer the relaxation time, the greater the electron mobility and the
higher the conductivity.
• Temperature: At low temperatures, electrons face fewer collisions because the
thermal vibrations (phonons) of the lattice are reduced. As a result, conductivity
increases with a decrease in temperature. At high temperatures: As temperature rises,
lattice vibrations increase, causing more frequent electron-phonon scattering. This
reduces the conductivity.
• Impurities and defects: Impurities introduce foreign atoms into the metal lattice.
These foreign atoms and crystal defects scatter free electrons, reducing the relaxation
time τ hence reducing conductivity.
Energy Bands
Energy Band Theory in Solids

In solids, atoms are closely packed together, and the energy levels of their electrons interact and form
continuous ranges of energy called energy bands. These bands determine the electrical properties of
the material. The two most important energy bands are the valence band and the conduction band.

1. Valence Band: The valence band is the range of energy levels where electrons are bound to
atoms. These electrons are not free to move because they are tightly attached to their
respective atoms. The valence band is typically full or nearly full of electrons, which limits
their ability to conduct electricity.

2. Conduction Band: The conduction band is the range of higher energy levels where electrons
are free to move throughout the solid. When electrons have enough energy to reach the
conduction band, they can move freely and thus conduct electric current. For a material to
conduct electricity, electrons must jump from the valence band to the conduction band.

3. Band Gap: The band gap is the energy difference between the conduction band and the
valence band. It plays a crucial role in determining the electrical properties of the material. If
the band gap is small or non-existent, electrons can easily move to the conduction band, and
the material conducts electricity. Conversely, a large band gap makes it harder for electrons to
move to the conduction band, limiting the material's conductivity.

Ec refers to the energy at the edge of the conduction band. It is the minimum energy that an
electron must have to enter the conduction band, where it can move freely through the material and
contribute to electrical conduction. Electrons that gain enough energy to reach Ec are no longer bound
to atoms and can travel under the influence of an electric field, making the material conductive. Ev
refers to the energy at the edge of the valence band. It is the highest energy an electron can have
while still being bound to the atoms of the material. Electrons in the valence band are not free to
move, as they are tightly bound to their respective atoms, and their movement is restricted. For an
electron to contribute to conduction, it must jump from Ev, the top of the valence band, to Ec, the
bottom of the conduction band.
The energy difference between Ec and Ev is known as the band gap (Eg). This band gap defines the
energy that an electron must gain to move from the valence band to the conduction band, and it plays
a critical role in determining the electrical properties of the material. In conductors, Ec and Ev may
overlap, allowing
electrons to move
freely between the
bands, resulting in
high conductivity. In
insulators, the gap
between Ec and Ev
is large, making it
nearly impossible
for electrons to
move to the
conduction band. In semiconductors, the band gap is moderate, and electrons can be excited to move
from Ev to Ec under certain conditions, such as heat or light, enabling the material to conduct
electricity.

Conductors

Conductors are materials that easily allow the flow of electric current. This is because, in conductors,
the valence band and conduction band either overlap or have no band gap at all. In other words, there
is no significant energy barrier that prevents electrons from moving into the conduction band. Since
many electrons are already in the conduction band, they can move freely under the influence of an
external electric field, allowing the material to conduct electricity effectively.

Common examples of conductors include metals like copper, silver, and aluminium. In these
materials, even at room temperature, there are abundant free electrons available for conduction, which
is why metals are excellent electrical conductors. The absence of a band gap allows continuous flow
of electricity without requiring additional energy to excite electrons from the valence band to the
conduction band.

Examples: Copper, silver, aluminium.

Semiconductors

Semiconductors are materials with a small but significant energy band gap, usually in the range of 1 to
2 electron volts (eV). At absolute zero, the conduction band in semiconductors is empty, and all
electrons are bound in the valence band. However, at room temperature or when exposed to heat or
light, some electrons gain enough energy to cross the small band gap and move from the valence band
to the conduction band.

When electrons jump to the conduction band, they leave behind "holes" in the valence band. These
holes act as positive charge carriers. The movement of both the free electrons in the conduction band
and the holes in the valence band enables semiconductors to conduct electricity. However, this
conduction is weaker compared to conductors because fewer electrons are present in the conduction
band.

Semiconductors are highly useful in electronic devices because their conductivity can be precisely
controlled through doping—the process of adding small amounts of impurities. Doping can either
increase the number of free electrons (n-type doping) or increase the number of holes (p-type doping),
thus enhancing the material's ability to conduct electricity.

Examples: Silicon, germanium.

Insulators

Insulators are materials with a large band gap, typically greater than 5 eV. The large energy difference
between the valence and conduction bands means that electrons in the valence band cannot gain
enough energy under normal conditions (like room temperature) to jump to the conduction band. As a
result, very few or no electrons are available in the conduction band to carry current, making
insulators poor conductors of electricity.

The lack of free charge carriers in insulators makes them ideal materials for preventing the flow of
electric current. Common examples of insulators include materials like rubber, plastic, and glass,
which are used to coat or enclose electrical wiring to prevent accidental current flow.

In insulators, the electrons are tightly bound to their atoms, and even when an electric field is applied,
the energy provided by the field is not enough to move the electrons into the conduction band. This
makes insulators highly resistant to electric current, which is why they are used as protective materials
in electrical systems.

Examples: Rubber, plastic, glass.

Fermi energy
The Fermi level is a crucial concept in understanding the electronic properties of solids, especially in
materials like conductors, semiconductors, and insulators. It represents the highest energy level that
electrons occupy at absolute zero temperature (0 Kelvin).
Fermi Level is a more general term used at non-zero temperatures, representing the energy level
where the probability of finding an electron is 50%.

Conductors (Metals): In conductors, like metals, the Fermi level lies within the conduction band
or at the overlap between the conduction and valence bands. This means that at absolute zero, there
are already free electrons available in the conduction band, allowing metals to conduct electricity
easily.

Semiconductors: In semiconductors, the Fermi level lies within the band gap, between the
conduction band and the valence band. The exact position of the Fermi level within the band gap can
shift depending on whether the semiconductor is intrinsic (pure) or extrinsic (doped).

• In intrinsic semiconductors (pure), the Fermi level is typically located near the middle of the
band gap, equidistant from the conduction and valence bands.

• In n-type semiconductors (doped with extra electrons), the Fermi level shifts closer to the
conduction band.

• In p-type semiconductors (doped with holes or electron acceptors), the Fermi level shifts
closer to the valence band.

Insulators: In insulators, the Fermi level is also within the band gap, similar to semiconductors, but
the band gap is much larger. The Fermi level is located far from both the conduction and valence
bands, making it extremely difficult for electrons to jump from the valence band to the conduction
band. As a result, insulators have very few, if any, free electrons at normal temperatures, and they do
not conduct electricity.

The Fermi energy in a 3D system of free electrons (like in a metal) can be approximated as:

ℏ2 2
𝐸𝑓 = (3𝜋 2 𝑛) ⁄3
2𝑚

where:

ℏ is the reduced Planck constant,

m is the mass of the electron,

n is the electron density.

Fermi-Dirac Distribution
Statistical methods in physics are used to describe how particles are distributed among various
energy states, particularly when dealing with large systems of particles, such as gases or solid
materials. There are 3 major statistical methods namely,
1. Maxwell-Boltzmann Statistics which is applicable to classical, distinguishable (for e.g., air
molecules).
2. Fermi-Dirac Statistics which is applicable to fermions (e.g., electrons, protons, neutrons etc.)
particles with half-integer spin.
3. Bose-Einstein Statistics which is applicable to bosons (e.g., photons), particles with integer
spin.

Particles

Indistinguishable Distinguishable

Fermions Mesons

In solid-state physics, this distribution is crucial for understanding the behaviour of electrons in a
solid, especially at different temperatures. So, we choose Fermi-Dirac distribution, since we are
studying about electrons.

The Fermi-Dirac distribution function gives the probability f(E), that a quantum state with energy E
is occupied by an electron at a given temperature T. It is expressed as:

1
𝑓(𝐸) = (𝐸−𝐸𝑓 )
𝑒 𝐾𝐵 𝑇 +1
f(E): The probability that an energy state E is occupied by an electron.
E: The energy of the state in question.
Ef: Fermi energy, the highest energy level occupied by electrons at absolute zero (0 K).
KB: Boltzmann constant 1.38×10−23 J/K.
T: Temperature (in Kelvin).
Variation of Fermi function with Temperature
The Fermi function depends on temperature, and its behaviour changes as the temperature changes.
Let’s divide it into 3 cases and analyse them separately.

Case-1: At Absolute Zero (T = 0 K)


The Fermi function f(E) is given by,
1
𝑓(𝐸) = (𝐸−𝐸𝑓 )
𝑒 𝐾𝐵 𝑇 +1
There are two possibilities. Energy E can be greater than fermi energy Ef or less than fermi energy Ef.
For Energy E greater than fermi energy Ef, 𝐸 − 𝐸𝑓 will be positive.

For Energy E less than fermi energy Ef, 𝐸 − 𝐸𝑓 will be negative.

If Energy E is equal to fermi energy Ef, 𝐸 − 𝐸𝑓 will be zero.

• When T=0K, for 𝐸 > 𝐸𝑓 ,

(𝐸 − 𝐸𝑓 ) (𝐸 − 𝐸𝑓 )
= = ∞ (𝑖𝑛𝑓𝑖𝑛𝑖𝑡𝑦)
𝐾𝐵 𝑇 𝐾𝐵 × 0
Then the Fermi function f(E) become,
1
𝑓(𝐸) =
(𝐸−𝐸𝑓 )
𝑒 𝐾𝐵 𝑇 +1
1 1 1
= = = =0
𝑒∞ +1 ∞+1 ∞

• When T=0K, for 𝐸 < 𝐸𝑓 ,

(𝐸 − 𝐸𝑓 ) (𝐸 − 𝐸𝑓 )
= = −∞
𝐾𝐵 𝑇 𝐾𝐵 × 0
Then the Fermi function f(E) become,
1
𝑓(𝐸) =
(𝐸−𝐸𝑓 )
𝑒 𝐾𝐵 𝑇 +1
1 1 1
= = = =1
𝑒 −∞ + 1 0+1 1
• When T=0K, for 𝐸 = 𝐸𝑓 ,

(𝐸 − 𝐸𝑓 ) (𝐸𝑓 − 𝐸𝑓 )
= = 0
𝐾𝐵 𝑇 𝐾𝐵 × 0
Then the Fermi function f(E) become,
1
𝑓(𝐸) =
(𝐸−𝐸𝑓 )
𝑒 𝐾𝐵 𝑇 +1
1 1 1
= = = = 0.5
𝑒0 + 1 1 + 1 2
In short,
For Energy E greater than fermi energy Ef, 𝑓(𝐸) = 0
For Energy E equal to fermi energy Ef, 𝑓(𝐸) = 0.5
For Energy E less than fermi energy Ef, 𝑓(𝐸) = 1
When dealing with probability functions, remember that the maximum value of probability is 1 and
minimum value of probability is 0. Thus, the probability of electron to be occupied in an energy level
below the fermi energy is 100 percent, while the same for levels having energy greater than fermi
level is zero percentage. Also, the probability of an electron to be occupied in the fermi level is 50
percent. So,

• All energy states below the Fermi energy are fully occupied (probability f(E)=1).
• All energy states above the Fermi energy are completely empty (probability f(E)=0).

Case 2: At Finite Temperature (T > 0 K)

As temperature increases from absolute zero, some electrons are thermally excited from some levels
just below fermi energy to higher energy levels just above fermi energy. There is a small, but nonzero
probability that electrons will have enough thermal energy such that states above the Fermi energy
will be partially occupied and some states below will be partially empty. That is, the probability of
occupying states above Ef is no longer zero, and the probability of occupying states below Ef is no
longer one.

Fermi level

For higher energy states, E >>> Ef , the Fermi function decays rapidly, as f(E)≈ 0 (low probability of
occupation).

For lower energy states, E <<< Ef , the Fermi function approaches f(E)≈ 1 (high probability of
occupation).

Case 3: High Temperatures (T ≫ 0 K)


As the temperature rises further, the transition around fermi level becomes even more gradual.
Electrons have more thermal energy, and the distribution of electrons among energy levels spreads
out.

The Fermi function becomes less sharp, meaning that:

• A larger fraction of electrons occupies higher energy states above fermi level.
• The probability of finding electrons in states below fermi level decreases (some electrons are
thermally excited to higher energy levels).

At very high temperatures, the Fermi-Dirac distribution starts to resemble the Maxwell-Boltzmann
distribution.

Fermi level

Graphical Representation of the Fermi Function


• At T = 0 K: The graph shows a step at fermi level, with a sudden drop from 1 (fully occupied)
to 0 (completely empty).

• At T > 0 K: The step becomes a curve that transitions more gradually from 1 to 0 around
fermi level.

• At high temperatures: The curve smooths out even more, and it starts to resemble an
exponential decay, similar to the classical Maxwell-Boltzmann distribution.
Superconductivity

• Superconductivity is a phenomenon observed in certain materials where they exhibit


zero electrical resistance and the expulsion of magnetic fields when cooled below a
certain temperature called critical temperature (Tc).
• Superconductivity was first discovered in 1911 by Dutch physicist Heike Kamerlingh
Onnes while studying the electrical resistance of mercury at very low temperatures.
• He observed that at a temperature of 4.2 K (Kelvin), the electrical resistance of
mercury suddenly dropped to zero, indicating the onset of superconductivity.
• When a material becomes superconducting, it can conduct electric current without any
energy loss, which is drastically different from normal conductive materials.

Critical Temperature

• The transition temperature, also known as the critical temperature (Tc), is the
temperature below which a material transitions from a normal conductive state to a
superconducting state.
• Above Tc: The material behaves like a normal conductor, with finite electrical
resistance and no expulsion of magnetic fields.
• Below Tc: The material enters the superconducting state, characterized by zero
resistance and the Meissner effect.
• Each superconducting material has its own specific critical temperature, which can
range from near absolute zero (0 K) to above 100 K for high-temperature
superconductors.
• High-temperature superconductors, such as certain ceramic materials, have higher
transition temperatures (above 77 K), making them easier to use in practical
applications because they can be cooled with liquid nitrogen instead of liquid helium.
• If the material is then warmed above the critical temperature, it returns to its
normal conductive state with finite electrical resistance, and any expelled magnetic
fields are re-admitted into the material. i.e., the superconducting state is reversibly
achieved.
Critical Field
• The critical field (Hc) in superconductivity refers to the maximum magnetic field
strength that a superconductor can withstand while remaining in the superconducting
state.
• When the applied magnetic field exceeds this critical value, the superconducting state
is destroyed, and the material transitions back to the normal (non-superconducting)
state with finite electrical resistance.
• The critical field varies with temperature and decreases as the temperature approaches
the critical temperature (Tc).
• The critical field is a crucial parameter for determining the applications of
superconductors. For instance, superconducting magnets used in MRI machines and
particle accelerators require materials with high critical fields.
• Superconductors with higher critical fields can be used in stronger magnetic field
environments without losing their superconducting properties
𝑇 2
• The relationship can be approximated by: 𝐻𝑐 (𝑇) = 𝐻𝑐0 (1 − (𝑇 ) ), where 𝐻𝑐0 is the
𝑐

critical field at absolute zero (0 K), and T is the temperature.

Review of Magnetism
1. The region of space around a magnet, current carrying conductor or a moving charge in
which magnetic effect can be experienced is called magnetic field.

2. If the field is strong, these field lines are more closely spaced. The magnetic field strength
is called magnetic induction (B). Magnetic induction (magnetic flux density) at a point is
defined as the flux (field lines) passing through unit area around the point. The unit of
magnetic induction is tesla (T).

3. When a magnetic field generated by current passes through a magnetic material, the
material itself contributes an internal magnetic field called magnetizing field (𝐇). Its SI
unit is ampere/metre (A/m).
4. When a material is placed in a magnetic field, it gets magnetised. The magnetic moment
per unit volume of the material is called intensity of magnetisation (M). It is the extent to
which a specimen is magnetised. Unit is ampere/metre (A/m).

5. The magnetic permeability of  of a material is defined as the ratio of resultant magnetic


field B inside the material to the magnetizing field H. Its unit is henry/metre (H/m). The
permeability of free space is a constant 𝜇0 = 4𝜋 × 10−7 𝐻/𝑚

𝐵
𝜇=
𝐻

6. The ratio of magnetization M to the magnetizing field H is called magnetic susceptibility.


It is denoted by the letter . It doesn’t have any unit.

𝑀
𝜒=
𝐻

7. The key difference between magnetic permeability and susceptibility is that magnetic
permeability describes the ability of a material to support the formation of a magnetic
field inside itself whereas susceptibility describes whether a material is attracted to a
magnetic field or is repelled from it.

8.  is –ve for diamagnetic substances and  is + ve for paramagnetic substances

9. Relation connecting magnetization(M), magnetizing field(H) and magnetic induction(B)


is,

𝐵 = 𝜇0 (𝑀 + 𝐻)

Meissner effect
The Meissner Effect is a fundamental phenomenon observed in superconductors, where a
superconductor expels all magnetic fields from its interior when it transitions into the
superconducting state below a critical temperature (Tc). This effect is one of the defining
characteristics of superconductivity and distinguishes superconductors from perfect
conductors.
When a superconducting material is cooled below its critical temperature in the
presence of a magnetic field, it will expel the magnetic field lines from its interior.
This expulsion results in zero magnetic flux density (B=0) inside the superconducting
material.
For a material, we know that,
𝐵 = 𝜇0 (𝑀 + 𝐻)
In the case of a superconductor, according to Meissner effect, total magnetic field
inside is zero.
i.e., 𝐵=0
0 = 𝜇0 (𝑀 + 𝐻)
0 = (𝑀 + 𝐻)
𝐻 = −𝑀
Magnetic susceptibility,  is
𝑀
𝜒=
𝐻
𝑀
𝜒=
−𝑀
𝜒 = −1

The Meissner Effect causes superconductors to behave as perfect diamagnets, meaning they
exhibit a magnetic susceptibility (χ) of −1. This complete diamagnetism is a unique feature
that occurs only in superconductors.
Type 1 superconductors

• Type I superconductors are materials that exhibit superconductivity and completely

expel magnetic fields below a critical temperature (Tc) and a critical magnetic field

(Hc).

• Exhibit complete Meissner effect.

• They are typically pure elemental superconductors.

• They have only a single critical magnetic field (Hc). When the applied magnetic field

exceeds this value, the magnetization of the material abruptly falls to the zero.

• It loses magnetization abruptly.

• There is no mixed state is present.

• Type I superconductors are generally limited to low magnetic fields, making them

less useful in applications requiring high magnetic fields.

• Its critical magnetic field is usually low. i.e., low magnetic field is required to destroy

the superconductivity of the material. Therefore, they are known as soft

superconductors.

• Lead (Pb): Critical temperature 𝑇𝑐 ≈ 7.2 K

• Mercury (Hg): Critical temperature 𝑇𝑐 ≈ 4.2 K

• Tin (Sn): Critical temperature 𝑇𝑐 ≈ 3.7 K

Type 2 superconductors

• Type II superconductors are characterized by their ability to partially allow magnetic

fields to penetrate their surface while maintaining superconductivity.

• Exhibit partial Meissner effect.

• They are typically pure elemental as well as alloy superconductors.

• Has 2 critical fields, a lower critical field (𝑯𝒄𝟏 ) and the upper critical field (𝑯𝒄𝟐 ).
• Lower Critical Field (𝑯𝒄𝟏 ): The field strength below which the material exhibits

perfect diamagnetism. I.e., below the lower critical field (𝑯𝒄𝟏 ), Type II

superconductors exhibit the Meissner effect by completely expelling magnetic fields.

• Between 𝑯𝒄𝟏 and the upper critical field (𝑯𝒄𝟐 ), they enter a mixed state (or vortex

state) where magnetic flux lines penetrate into the material.

• Upper Critical Field (𝑯𝒄𝟐 ): The field strength above which superconductivity is

completely destroyed and will become a normal conductor.

• It loses magnetization gradually.

• Type II superconductors typically have higher critical temperatures and critical

magnetic fields compared to Type I superconductors. This makes it versatile for

various applications, such as Superconducting magnets, Quantum computing etc.

• Since the critical field is large, it requires a large magnetic field to destroy the

superconductivity of the material. Therefore, they are known as hard superconductors.

• Niobium (Nb): Critical temperature Tc≈9.2K

• Yttrium Barium Copper Oxide: Tc≈92K

• Magnesium Diboride: Critical temperature Tc≈39K


BCS Theory
The BCS Theory is a microscopic theory of superconductivity developed by John Bardeen,
Leon Cooper, and Robert Schrieffer in 1957. This groundbreaking theory explains how
superconductivity arises in certain materials at low temperatures and provides a framework
for understanding the behaviour of superconductors.

Normally, electrons, which are negatively charged, repel each other. This repulsion causes
them to scatter when moving through a material, creating resistance. At low temperatures,
electrons can attract each other in a special way. The pairing occurs due to an attractive
interaction between electrons, mediated by lattice vibrations (phonons). When an electron
moves through the lattice, it slightly distorts the lattice, attracting another electron to the
region. The first electron's interaction with the lattice causes a phonon to be emitted, and the
second electron absorbs this phonon, leading to an attractive force between the two electrons.
This leads to the formation of a Cooper pair. The electrons in these pairs have opposite
momenta and spin. In short, phonons act like the "glue" that helps electrons pair up to create
the special state needed for superconductivity.

When electrons form Cooper pairs, an energy gap opens up in the energy spectrum of the
material. This gap means that a certain minimum amount of energy is required to break a
Cooper pair. At temperatures below the critical temperature, thermal energy is not sufficient
to break the pairs, so the electrons continue to move without resistance, leading to
superconductivity. In this state, they move coherently as a single entity, without scattering,
because any disturbance would have to overcome the energy gap to disrupt the pairs. All
Cooper pairs in a superconductor are in the same quantum state, which means they exhibit
phase coherence. The movement of these pairs through the lattice does not involve scattering,
leading to zero electrical resistance.

Applications of Superconductors
Superconductors have numerous applications across various fields due to their unique
properties such as zero electrical resistance and the ability to expel magnetic fields. Here
are some of the key applications:
1. Magnetic Resonance Imaging (MRI):

Superconducting magnets are used in MRI machines to generate strong and


stable magnetic fields. The high field strength allows for better image
resolution and clarity.

2. Maglev Trains:

Magnetic levitation trains utilize superconductors to achieve frictionless


movement. The levitation effect produced by superconductors allows trains to
float above the tracks, resulting in higher speeds and energy efficiency.

3. Particle Accelerators:

Superconducting materials are used in the construction of magnets for particle


accelerators, such as the Large Hadron Collider (LHC). These magnets are
crucial for steering and focusing particle beams at high energies.

4. Quantum Computing:

Superconducting qubits are used in quantum computers. The unique properties


of superconductors allow for the creation of qubits that can exist in
superpositions of states, enabling advanced computation capabilities.

5. Superconducting Power Cables:

Superconducting materials can be used in power transmission lines, allowing


for efficient transmission of electricity without energy losses. This can lead to
significant improvements in grid efficiency and reliability.

6. Flux Leakage Sensors:

Superconducting materials are employed in devices known as SQUIDs


(Superconducting Quantum Interference Devices), which are extremely
sensitive magnetometers. SQUIDs are used for detecting weak magnetic fields
in various applications, including medical diagnostics and geological surveys.

7. Energy Storage:

Superconducting magnetic energy storage (SMES) systems use


superconductors to store energy in magnetic fields. This technology offers
rapid response times and high efficiency for stabilizing electrical grids and
providing backup power.

8. Telecommunications:

Superconducting components can enhance the performance of communication


systems by reducing noise and improving signal integrity. They are
particularly useful in radiofrequency (RF) applications and high-frequency
circuits.

Problems

Q. A semiconductor has an electrical conductivity of 1.2 × 105 S/m. The carrier

concentration is 2 × 1024 m−3, and the effective mass of the charge carriers is

0.2 × 9.11 × 10−31 kg. Find the relaxation time.

Answer

To find the relaxation time (τ) of the semiconductor, we can rearrange the formula for
electrical conductivity (σ):

𝑛𝑒 2 𝜏
𝜎=
𝑚

σ is the electrical conductivity 1.2 × 105 S/m

n is the carrier concentration 2 × 1024 m−3

e is the charge of an electron

m is the effective mass of the charge carriers 0.2 × 9.11 × 10−31 kg

τ is the relaxation time (what we need to find).


Q. A metal has an electron concentration of 8.5 × 1028 m−3 . The relaxation time for

electron collisions is 𝜏 = 2.5 × 10−14 s. Calculate the electrical conductivity of the

metal.

Answer
Q. Calculate the probability of an electron occupying an energy state that is 0.2 eV above the

Fermi level at a temperature of 300 K.


Q. Find the temperature at which the probability of occupying an energy state 0.1 eV above

the Fermi level is 0.1.

You might also like