Unit 2-1111

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

P H 1 0 1 — PHYSICS

semiconductors
durjoy roy

CSE Lecture hall, University of Kalyani

SYLLABUS: Intrinsic and extrinsic semiconductors, Dependence of Fermi level on


carrier-concentration and temperature (equilibrium carrier statistics), Carrier
generation and recombination, Carrier transport: diffusion and drift, p-n junction,
Metal-semiconductor junction (Ohmic and Schottky), Semiconductor materials of
interest for optoelectronic devices.

contents

1 Introduction 2
1.0.1 Effective Mass and Statistical Considerations . . . . . . 2
1.0.2 Field Current and Mobility . . . . . . . . . . . . . . . . . 3
1.0.3 Lattice Scattering . . . . . . . . . . . . . . . . . . . . . . 5
1.0.4 Ionized impurity scattering . . . . . . . . . . . . . . . . 5
1.1 Types of crystalline solids . . . . . . . . . . . . . . . . . . . . . . 6

2 Types of Semiconductors 7
2.1 Intrinsic Semiconductors . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Extrinsic Semiconductors . . . . . . . . . . . . . . . . . . . . . . 8
2.2.1 p-type and n-type semiconductors . . . . . . . . . . . . 9

3 Carrier generation and recombination 11


3.1 Fermi-Dirac Distribution . . . . . . . . . . . . . . . . . . . . . . 11
3.2 Density of States . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Carrier Concentration Equations . . . . . . . . . . . . . . . . . 12
3.4 Temperature Effects on Carrier Concentration . . . . . . . . . . 13
3.4.1 Thermal Equilibrium & Law of Mass Action . . . . . . . 13
3.5 Doping and Its Impact . . . . . . . . . . . . . . . . . . . . . . . 14
3.5.1 Equilibrium Carrier Concentration with Impurities . . 14
[email protected] (scribe)

1
1. introduction

3.5.2 Equilibrium vs. Non-Equilibrium States . . . . . . . . . 15


3.6 Applications and Real-World Examples . . . . . . . . . . . . . . 15

4 Carrier Transport 16
4.1 Diffusion and Drift Current . . . . . . . . . . . . . . . . . . . . 16

5 The P-N Junction 16


5.1 Forward-Bias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.2 Reverse-Bias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
5.3 The Shockley Equation . . . . . . . . . . . . . . . . . . . . . . . 19
5.4 The Bands at operation . . . . . . . . . . . . . . . . . . . . . . . 21
5.5 Biased P-N Junction . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.6 The phenomenon of Breakdown . . . . . . . . . . . . . . . . . . 30
5.7 Zener diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

1 introduction
Lecture 1 (2 hours)
30th September 2024 In the previous unit, we have discussed about the movement of electrons in
periodic crystal lattice, and acquired the idea of band gap, or energy gap. We
will recap some concepts that will be required in this unit.

1.0.1 Effective Mass and Statistical Considerations

When a free electron is perturbed by an electric field, it will be subject to forces


that cause it to accelerate; it moves opposite the direction of the electric field,
and would speed up with time. However, the situation in a crystal is different,
because the electron is actually moving through a lattice of jiggling atoms that
all exert electromagnetic forces. We cannot use the standard electron mass;
we must use an effective mass for the electron in the crystal, a result of the
periodic forces of the host atoms in the crystal. The wonderful thing is that
in the simple picture, we can view the electron as moving as though it were
in a vacuum, but with this new effective mass that varies from material to
material.

Another difference is that inside the crystal, a moving electron will not
travel far before colliding with a host atom or impurity. These collisions
randomize the electron’s motion; therefore, it is useful to use an average
time, the relaxation time τ, which is based on the random thermal motion
of the electrons. In fact, the scattering processes of the electron bouncing
around causes it to lose energy, which is given off as heat. With the addition
of an applied electric field, we also have a mean free path length λ, or a net
displacement on average for a given electron.
This means free charge carriers have a drift velocity, an average speed at
which they travel through the material. The average drift velocity for a single
electron is the same as the average of all drift velocities of all the electrons,

2
Figure 1: These pictures represent the drift of an electron as a result of thermal
motion. In figure (a) where there is no electric field, the electron jumps around
but ends up covering no net distance; in figure (b) where an electric field is
present, the electron drifts opposite the direction of the field and has a net
displacement (and therefore a drift velocity).

and is given by the following equation:

1 1 qτ
vd = aτ = E (1)
2 2 m∗c

where a is the average acceleration of the carrier, q is the charge of the


carrier (including charge), m∗ is the effective mass of the charge carrier, τ is
the carrier lifetime, and E is the electric field strength.

1.0.2 Field Current and Mobility

The movement of charge carriers in an electric field results in an electric


current. We will call the current resulting from drifting carriers our field
current. The current density J, or the current flow of electrons per unit volume,
is given by the following:

Jn = nqvd (2)

1 qτ
Jn = nq E (3)
2 m∗

where n is the carrier concentration (per unit volume). Furthermore, we


can get rid of the factor of 2 in this equation by averaging the lifetime τ
over all carrier velocities. Therefore, we can now define a quantity called

3
1. introduction

mobility, in this case electron mobility. Carrier mobility is useful as it is the


ratio of drift velocity to the electric field strength. Below we will give the
mathematical definition and substitute mobility (given as µn ) into the current
density equation.

νd qτ
µn = = ∗ (4)
E m

Jn = nqµn E (5)

From these equations we can then obtain the conductivity of the material
in terms of the mobility as:

Jn = σE (6)

σ = nqµn (7)

The same conditions hold for hole mobility and conductivity, and therefore
the total conductivity, which is directly inversely related to the resistivity (the
material’s resistance to being conductive, so to speak), is given below:

1 J
σ= = qµe n + qµh p (8)
ρ E

where ρ is the resistivity, n and p are the concentrations of electrons


and holes respectively, and µe and µh are the electron and hole mobilities
respectively.

We see that conductivity in a material is directly related to the mobility,


which depends on the density of dopants, temperature, and electric field
strength. Thus, as mobility decreases conductivity decreases. As materials
become more heavily doped, mobility decreases because dopant atoms are
very effective scatterers, and therefore decrease the average time between
collisions. Similarly, as temperature increases, mobility decreases, however
this effect becomes insignificant in heavily doped materials. As electric field
increases, the drift velocities of carriers will eventually become comparable
to the random thermal velocities. Therefore, high field strength decreases
mobility; semiconductors in this way differ from conductors, which so easily
generate current that only a low field strength occurs during current flow. It is
worth noting that less specialized impurities and crystal defects in the semi-
conductor material will also decrease the mobility, because of the scattering
effects mentioned above.

4
1.0.3 Lattice Scattering
Lattice scattering is the scattering of ions by interaction with atoms in a lattice.
This effect can be qualitatively understood as phonons colliding with charge
carriers.

In the current quantum mechanical picture of conductivity the ease with


which electrons traverse a crystal lattice is dependent on the near perfectly
regular spacing of ions in that lattice. Only when a lattice contains perfectly
regular spacing can the ion-lattice interaction (scattering) lead to almost
transparent behavior of the lattice.

Figure 2: Ions scattering off lattice atoms while diffusing in a lattice.

In the quantum understanding, an electron is viewed as a wave traveling


through a medium. When the wavelength of the electrons is larger than the
crystal spacing, the electrons will propagate freely throughout the metal
without collision.

1.0.4 Ionized impurity scattering


In quantum mechanics, ionized impurity scattering is the scattering of charge
carriers by ionization in the lattice. The most primitive models can be concep-
tually understood as a particle responding to unbalanced local charge that
arises near a crystal impurity; similar to an electron encountering an electric
field. This effect is the mechanism by which doping decreases mobility.

In the current quantum mechanical picture of conductivity the ease with


which electrons traverse a crystal lattice is dependent on the near perfectly
regular spacing of ions in that lattice. Only when a lattice contains perfectly
regular spacing can the ion-lattice interaction (scattering) lead to almost
transparent behavior of the lattice. Impurity atoms in a crystal have an effect
similar to thermal vibrations where conductivity has a direct relationship
with temperature.

5
1. introduction

A crystal with impurities is less regular than a pure crystal, and a reduction
in electron mean free paths occurs. Impure crystals have lower conductivity
than pure crystals with less temperature sensitivity in that lattice.

1.1 Types of crystalline solids


We also have learnt that, based upon the value of band gap, crystallite solids
can be categorized in three electrically conducting types namely,

1. Insulator: An insulator is type of material that does not allow the elec-
tric current to pass through it, due to its high electrical resistance. In
the insulators, the energy gap between valance and conduction bands
is very large (about 15 eV). Therefore, a very high electric field is re-
quired to push the valance electrons to the conduction band. Due to this,
there are no free electrons in the conduction band. For this reason, the
electrical conductivity of insulators is very low and may considered nil
under ordinary conditions. At room temperature, the valance electrons
of the insulator do not have enough energy to cross over to the forbidden
energy gap. But, if the temperature is raised, some of the electrons may
acquire enough energy to cross over to the forbidden energy gap. Hence,
the resistance of the insulator decreases with the rise in temperature.
Therefore, the insulators have negative temperature co-efficient of re-
sistance. Due to high electrical resistance, the insulators are used for
protection against electric shocks.

2. Conductor: A conductor is a type of material that allows the electric


current to flow through it i.e., it possesses least resistance in the path of
free electrons. In case of conductor, the valance and conduction bands
overlap. Due to this overlapping, a small potential difference across a
conductor causes the free electrons to constitute electric current. All
the metals are conductors. The resistance of the conductors increases
with the increase in the temperature. Hence, the conductor have positive
temperature co-efficient of resistance.

3. Semiconductor: The semiconductors are the materials having conduc-


tivity in-between conductors and insulators. In a semiconductor, the for-
bidden energy gap between valance and conduction bands is very small
(about 1 eV) as compared to insulators. Therefore, a smaller electric
field (smaller than insulators but greater than conductors) is required
to push the free electrons from valance band to the conduction band.
At low temperature, the valance band of semiconductor is completely
full and the conduction band is completely empty. Thus, a semicon-
ductor behaves as an insulator at low temperature. However, at room
temperature, some electrons can cross the forbidden energy gap, im-
parting a little conductivity to the semiconductor. As temperature is
increased, more valance electrons cross over to the energy gap to reach

6
to the conduction band and the conductivity increases. This shows that
electrical conductivity of semiconductor increases with the rise in tem-
perature. Hence, a semiconductor has negative temperature coefficient
of resistance. The conductivity of semiconductors can also be increased
by adding some impurity in the pure semiconductor material, called
doping. The semiconductors are commonly used in manufacturing of
solid state electronic devices.

2 types of semiconductors

In most pure semiconductors at room temperature, the population of ther-


mally excited charge carriers is very small. Often the concentration of charge
carriers may be orders of magnitude lower than for a metallic conductor. For
example, the number of thermally excited electrons cm–3 in silicon (Si) at 298
K is 1.5 × 1010 . In gallium arsenide (GaAs) the population is only 1.1 × 106 elec-
trons cm–3 . This may be compared with the number density of free electrons
in a typical metal, which is of the order of 1028 electrons cm–3 .

Given these numbers of charge carriers, it is no surprise that, when they


are extremely pure, silicon and other semiconductors have high electrical
resistivities, and therefore low electrical conductivities. This problem can be
overcome by doping a semiconducting material with impurity atoms. Even
very small controlled additions of impurity atoms at the 0.0001% level can
make very large differences to the conductivity of a semiconductor.

Figure 3: A Silicon wafer, ready for fabrication of devices. The flat cut is to
show their crystallographic orientation.

7
2. types of semiconductors

2.1 Intrinsic Semiconductors

Semiconductors are one of the three classes of electrical materials and are the
foundation of every solid-state electronic device which is in use today. Intrin-
sic semiconductors, also known as pure or undoped semiconductors, describe
perfect semiconductor crystals which are free from defects and impurities of
other elements. Intrinsic semiconductors which are intentionally doped with
other elements are referred to as Extrinsic Semiconductors. Intrinsic proper-
ties are found in all semiconductor materials, even those doped with other
elements, with the doping elements introducing other desired properties.

2.2 Extrinsic Semiconductors

An extrinsic semiconductor is a material with impurities introduced into


its crystal lattice. The goal of these impurities is to change the electrical
properties of the material, specifically (increasing) its conductivity. Today,
extrinsic semiconductors are a part of innovative, modern technology devices
(including efficient solid state lighting and renewable energy) such as light
emitting diodes, solar cells, lasers, and transistors.
Extrinsic semiconductors are just intrinsic semiconductors that have been
doped with impurity atoms (one dimensional substitutional defects in this
case). Doping is the process where semiconductors increase their electrical
conductivity by introducing atoms of different elements into their lattice. A
semiconductor can be doped by vapor phase epitaxy, where some concentra-
tion of impurities in their gaseous phase is contacted with the semiconductor
wafer, or by being grown in with the wafer itself with the help of photolithogra-
phy (microprocessing areas of a wafer), diffusion (gradient controlled particle
motion), and ion implantation (utilizing an electric field to contact an ion
with a solid) to increase the dopant concentration in certain parts of the wafer.
Based on whether the added impurities are electron donors or acceptors, the
semiconductor’s Fermi level (the energy state below which all allowable en-
ergy states are filled and above which all states are empty as the temperature
approaches 0 Kelvin) is able to move either up or down from its original
position in the center of the energy band gap (the gap between the valence
and conduction bands of the semiconductor).
If a semiconducting material is doped with atoms that can donate elec-
trons, it is known as an n-type semiconductor which uses those donated
electrons to increase its conductivity. If it is doped with with atoms that can
accept electrons, it is known as a p-type semiconductor which uses the lack
of electrons in the lattice, called holes (can be considered as positive charges
as well), to increase its conductivity as well. Aside from being dependent
on the concentration of dopants, both n-type and p-type semiconductors are
dependent on temperature changes, especially their electrical conductivities,

8
2.2 . extrinsic semiconductors

carrier mobilities (how freely a charge carrier can move), and even the Fermi
levels. The mobilities are based off of two different temperature dependent
scattering effects, called lattice scattering and ionized impurity scattering.
Lattice scattering, dominant at higher temperatures, is based on the thermal
vibrations of the semiconductor atoms which act as obstacles to mobile charge
carriers. Ionized impurity scattering, dominant at lower temperatures, de-
pends on the number of dopant ions, which all behave as scattering centers,
and their ability to negate charge carriers from being moving to different
energy levels because of the electrostatic attraction between ion and carrier
known as Coulomb’s Law, described as

qQ
F= (9)
4πϵ0⃗r2

where q and Q are the charges (opposite charges) on the charge carrier
and dopant ion, ϵ0 is the permittivity of free space, ⃗r is the distance between
ion and carrier, and F is the electrostatic attractive force that each of them
experience.

np = n2i (10)

where n is the electron concentration, p is the hole concentration, and ni is


the intrinsic carrier concentration (the concentration had the semiconductor
not been doped).

2.2.1 p-type and n-type semiconductors

If a very small number of atoms of a group V element such as phosphorus


(P) are added to the silicon as substitutional atoms in the lattice, additional
valence electrons are introduced into the material because each phosphorus
atom has 5 valence electrons. (See Fig. (4)) These additional electrons are
bound only weakly to their parent impurity atoms (the bonding energies are
of the order of hundredths of an eV), and even at very low temperatures these
electrons can be promoted into the conduction band of the semiconductor.
This is often represented schematically in band diagrams by the addition
of ’donor levels’ just below the bottom of the conduction band, as in the
schematic shown in Fig. (5a).

The presence of the dotted line in this schematic does not mean that
there now exist allowed energy states within the band gap. The dotted line
represents the existence of additional electrons which may be easily excited
into the conduction band. Semiconductors that have been doped in this way
will have a surplus of electrons, and are called n-type semiconductors. In
such semiconductors, electrons are the majority carriers.

Conversely, if a group III element, such as Boron (B), is used to substitute

9
2. types of semiconductors

(a) A Silicon atom is replaced by a Boron (b) A Silicon atom is replaced by a Phos-
atom. The absense of electron creates free phorus atom. The extra electron creates
carrier in the p-type semiconductor. free carrier in the n-type semiconductor.

Figure 4: Simplified crystal strucures of Extrinsic Semiconductors

(a) Band diagram of an n-type semicon- (b) Band diagram of an p-type semicon-
ductor. The dashed line is the Fermi Level, ductor. The dashed line is the Fermi Level,
and is called here donor level. and is called here acceptor level.

Figure 5: Band diagrams of Extrinsic Semiconductors

for some of the atoms in silicon, there will be a deficit in the number of valence
electrons in the material. (See Fig. (4a)) This introduces electron-accepting
levels just above the top of the valence band, and causes more holes to be
introduced into the valence band. Hence, the majority charge carriers are
positive holes in this case. Semiconductors doped in this way are termed
p-type semiconductors.

Doped semiconductors (either n-type or p-type) are all extrinsic semicon-


ductors. The activation energy for electrons to be donated by or accepted to
impurity states is usually so low that at room temperature the concentration of
majority charge carriers is similar to the concentration of impurities. It should
be remembered that in an extrinsic semiconductor there is an contribution to
the total number of charge carriers from intrinsic electrons and holes, but at
room temperature this contribution is often very small in comparison with
the number of charge carriers introduced by the controlled impurity doping
of the semiconductor.

10
3 carrier generation and recombination

The carriers in a semiconductor goes through a continuous process of genera-


tion and recombination. Let us discuss on the phenomenon here.

3.1 Fermi-Dirac Distribution


The Fermi-Dirac distribution describes the probability of an electron occupy-
ing an energy state at thermal equilibrium. It is represented by the equation:

1
f (E) = (11)
1+ e(E−EF )/kT

where:
• E is the energy of the state,

• EF is the Fermi level,

• k is the Boltzmann constant, and

• T is the absolute temperature.


At T = 0 K, the distribution becomes a step function, with all states below
EF occupied and all states above EF empty. As the temperature increases,
the distribution smoothens, with some states above EF becoming occupied
and some below EF becoming empty. The Fermi level plays a crucial role in
determining carrier concentrations in semiconductors, affecting their electrical
properties.

Figure 6: Fermi-Dirac Distribution Function

11
3. carrier generation and recombination

3.2 Density of States

The density of states (DOS) quantifies the number of available energy states
per unit volume and energy interval in a semiconductor. For a parabolic band,
the density of states is proportional to the square root of energy:

DOS(E) ∝ E (12)

The conduction band density of states is given by:


!3/2
2πm∗e kT
NC = 2 (13)
h2

where m∗e is the effective mass of electrons, and h is Planck’s constant.

Similarly, the valence band density of states is given by:


!3/2
2πm∗h kT
NV = 2 (14)
h2

where m∗h is the effective mass of holes.

3.3 Carrier Concentration Equations


Lecture 2 (2 hours)
4th October 2024 Carrier concentrations in semiconductors are determined by the Fermi-Dirac
distribution and the density of states. For intrinsic semiconductors, the elec-
tron and hole concentrations are equal:

ni = pi (15)

The intrinsic carrier concentration is given by:

p
ni = NC NV e−Eg /2kT (16)

where Eg is the band gap energy.

For extrinsic semiconductors, the majority carrier concentration is deter-


mined by the doping level. For n-type semiconductors:

n ≈ ND (17)

where ND is the donor concentration. For p-type semiconductors:

12
3.4 . temperature effects on carrier concentration

p ≈ NA (18)

where NA is the acceptor concentration. The minority carrier concentration


can be calculated using the mass action law:

The law of mass action


states that the rate of a re-
action is proportional to the
np = n2i (19) product of the concentra-
tions of each reactant.

3.4 Temperature Effects on Carrier Concentration


Temperature significantly affects the carrier concentrations in semiconductors.
As the temperature increases, more electrons are excited from the valence band
to the conduction band, leading to an exponential increase in the intrinsic
carrier concentration:

ni ∝ e−Eg /2kT (20)

In extrinsic semiconductors, the majority carrier concentration is relatively


insensitive to temperature, as it is determined by the doping level. However,
the minority carrier concentration increases with temperature.

3.4.1 Thermal Equilibrium & Law of Mass Action

Thermal equilibrium is described by a relationship between the concentrations


of charge carriers at a given temperature. In other words, at a set temperature,
there is equilibrium between generation (the formation of electron-hole pairs
as a result of either thermal excitation or photon absorption) and recombi-
nation (the natural destruction of pairs that occurs when electrons from the
conduction band fall back into the valence band).

Let’s discuss the intrinsic case first (undoped semiconductor). In the in-
trinsic case, the number of bound electrons is much greater than the number
of free electrons, therefore the generation of charge carriers is independent
of the number of electron-hole pairs that have already been formed. As soon
as generation occurs though, recombination comes into play and depends on
the concentration of charge carriers in the material. In fact, according to the
Law of Mass Action, recombination is directly proportional to the number of
charge carriers. This gives

R = rpn (21)

13
3. carrier generation and recombination

Where R is the recombination rate per unit volume per unit time and r is the
recombination probability.
Now, in thermal equilibrium the generation rate G is equal to the recom-
bination rate R, and therefore because of the conditions outlined in Intrinsic
Carrier Concentration, we have

G = R = rnp (22)

np = n2i (23)
G
= n2i (24)
r
where ni is the intrinsic carrier concentration. This means that the intrinsic
carrier density in a material is determined by the ratio of the generation
rate and recombination probability. Therefore the only variable dependence
here is on G, which in turn depends on the temperature. It turns out this
important relationship holds not only for intrinsic materials, but also for
doped semiconductors.

3.5 Doping and Its Impact


Doping involves the intentional introduction of impurities into a semiconduc-
tor to modify its electrical properties.

• n-type doping: introduces donor impurities (e.g., phosphorus) that


provide extra electrons to the conduction band.

• p-type doping: introduces acceptor impurities (e.g., boron) that create


holes in the valence band.

Doping shifts the Fermi level towards the corresponding band edge. In
n-type doping, the Fermi level moves closer to the conduction band, while
in p-type doping, it moves closer to the valence band. Heavily doped semi-
conductors have higher conductivity compared to lightly doped or intrinsic
semiconductors.

3.5.1 Equilibrium Carrier Concentration with Impurities


Now let’s consider the case with impurities. If we add the intrinsic carrier
concentration to the number of free carriers we obtain by doping the semicon-
ductor, we have the equilibrium carrier concentration of the majority carrier
in the material. However, it turns out that in most conditions, the doping of
the semiconductor is several orders of magnitude greater than the intrinsic
carrier concentration. In fact, for Silicon doped with Phosphorous, even at
low temperatures (< 70K) almost all the electrons will be free and located in
the conduction band; furthermore, at room temperature, the concentration

14
3.6 . applications and real-world examples

of conduction electrons in doped Silicon is six orders of magnitude greater


than in the intrinsic case, and conductivity increases to the same degree. This
means that conductivity becomes dependent on the number of dopant atoms
rather than the temperature (doped semiconductors are conductive over a
wide temperature range).

The above conditions (namely that the concentration of conductive elec-


trons from doping is much greater than that from the intrinsic, and that nearly
all the dopant atoms are ionized) imply the number of majority carriers is
approximately equal to the doping, as given in the condition below.

+
ND ≈ ND ≈ n (N − Type) (25)
+
NA ≈ NA ≈ p (P − Type) (26)

In the equations above, ND and NA are the concentration of donated


electrons and concentration of accepting holes respectively. Using the Law of
Mass Action from the section above, we get the following equations

Ni2
p≈ << n (N − Type) (27)
ND
Ni2
n≈ << p (P − Type) (28)
NA
These relations just mean that in an doped semiconductor, we have that
the concentration of minority carriers is much smaller than the concentration
of majority carriers.

3.5.2 Equilibrium vs. Non-Equilibrium States

In an equilibrium state, no external forces act on the semiconductor, and the


carrier concentrations are determined by the Fermi-Dirac distribution. In a
non-equilibrium state, external forces such as electric fields or light disturb
the equilibrium, leading to a non-constant Fermi level or quasi-Fermi levels
for electrons and holes.

3.6 Applications and Real-World Examples

Semiconductors are fundamental to various devices:

• Solar cells: convert light into electrical energy using p-n junctions.

• LEDs: emit light when forward biased, using minority carrier injection.

• Transistors: function as amplifiers and switches, including bipolar junc-


tion transistors (BJTs) and field-effect transistors (FETs).

15
5. the p-n junction

• Semiconductor lasers: produce coherent light through stimulated emis-


sion.

• Photodetectors: convert light into electrical signals.

4 carrier transport

The transport of carriers, both majority and minority, are governed by certain
rules. Transport of carriers causes current, and we have two types of it, namely,
diffusion, and drift currents.

4.1 Diffusion and Drift Current


Diffusion current is a current in a semiconductor caused by the diffusion of
charge carriers (electrons and/or electron holes). This is the current which is
due to the transport of charges occurring because of non-uniform concentra-
tion of charged particles in a semiconductor.
The drift current, by contrast, is due to the motion of charge carriers due
to the force exerted on them by an electric field. Diffusion current can be in
the same or opposite direction of a drift current. The diffusion current and
drift current together are described by the drift–diffusion equation.

Table 1: Comparison between Drift & Diffusion Currents

Drift current Diffusion current


Drift current is caused by electric Diffusion current is caused by varia-
fields. tion in the carrier concentration.
Direction of the drift current is al- Direction of the diffusion current
ways in the direction of the electric depends on the gradient of the car-
field. rier concentration.

Obeys Ohm’s law: J = qρµE Obeys Fick’s law: J = qd dx E

It is necessary to consider the part of diffusion current when describing


many semiconductor devices. For example, the current near the depletion
region of a p–n junction is dominated by the diffusion current. Inside the
depletion region, both diffusion current and drift current are present. At
equilibrium in a p–n junction, the forward diffusion current in the depletion
region is balanced with a reverse drift current, so that the net current is zero.
A comparison between the drift & Diffusion currents are listed in the
Table (1).

5 the p - n junction

If we were to create a junction between a region of N-type material and a


region of P-type material within a single crystal, an intriguing phenomenon

16
would occur. Assuming the crystal is not at absolute zero temperature, the
thermal energy within the system would lead some of the free electrons in
the N-type material to move into the surplus holes present in the adjacent
P-type material. This would result in the formation of a region that lacks
any charge carriers. It’s important to note that electrons are the predominant
charge carriers in N-type material, while holes are the predominant charge
carriers in P-type material. In essence, the region where the N and P materials
meet becomes depleted of available electrons and holes, and is commonly
referred to as a ”depletion region.” This concept is illustrated in Fig. (7a),
where the excess electrons in the N-type material are represented by minus
signs, and the excess holes in the P-type material are represented by plus
signs. At the junction, the free electrons combine with the holes, and when an
electron recombines, it leaves a positive ion in the N-type material (depicted
as a circled plus sign) and generates a negative ion in the P-type material
(depicted as a circled minus sign).

(a) PN junction. (b) Energy bands in PN junction.

Figure 7: A PN junction, and its band diagram

Now, we have established a zone devoid of charge carriers, and this will
impact the ability to initiate a current flow through the device. Essentially,
we have introduced a barrier in the form of an energy elevation that must be
surmounted.

(b) A PN junction diode in DO-204


(a) Diode schematic symbol (ANSI). case.

Figure 8: PN junction diode symbol, and a common packaging of a PN junction


diode.

17
5. the p-n junction

To grasp the concept of this energy barrier, let’s recall that in the previous
chapter, we learned that doping an intrinsic crystal results in a shift of the
Fermi level. In N-type material, the Fermi level shifts upward, closer to the
conduction band, whereas in P-type material, it shifts downward, approaching
the valence band. When two dissimilar regions meet, as is the case here, the
energy bands adjust so that the Fermi levels align. This alignment effectively
causes the energy bands of the P-type material to rise relative to those of the
N-type material. The junction between the two regions appears as an elevated
region, and this corresponds to the previously mentioned depletion region.
This situation is depicted graphically in Fig. (7b). You can compare this energy
diagram to the energy diagrams for N-type and P-type materials presented in
the preceding chapter. By simply aligning the Fermi levels, it becomes evident
how we arrive at this new energy diagram.

Now let’s consider what happens if we were to connect this device to an


external voltage source as shown in Fig. (9). Obviously, there are two ways
to orient the PN junction with respect to the voltage source. This version is
termed forward-bias.

Figure 9: Biased PN junction.

5.1 Forward-Bias
In Fig. (9), the dashed line illustrates the path of electron flow, which is
opposite to the direction of conventional flow. Initially, electrons move from
the negative terminal of the battery towards the N material. In the N material,
electrons, which are the majority carriers, can move through it with ease.
When they enter the depletion region, if the applied potential is sufficiently
high, these electrons can diffuse into the P material, where numerous lower-
energy holes are present. Subsequently, the electrons can travel through to
the positive terminal of the source, thus completing the circuit. A resistor is
added to limit the maximum current flow.

The key to this process is ensuring that the applied potential is substantial
enough to overcome the influence of the depletion region. This implies that

18
5.2 . reverse-bias

a specific voltage must be applied across the depletion region to facilitate


current flow. This required potential is referred to as the barrier potential or
forward voltage drop. The exact value varies depending on the material used.
For silicon devices, the barrier potential is typically estimated to be around
0.7 volts, while for germanium devices, it is closer to 0.3 volts. LEDs, on the
other hand, may exhibit barrier potentials in the range of 1.5 to 3 volts, which
partly depends on their color.

Another way to conceptualize this is that the addition of the voltage source
effectively ”flattens” the inherent energy barrier of the junction. Once the
applied forward-bias voltage equals or exceeds this barrier, current can flow
freely.

5.2 Reverse-Bias
If the polarity of the voltage source is reversed in Fig. (9), it significantly
alters the behavior of the PN junction. In this scenario, the electrons in the
N material are attracted towards the positive terminal of the source, while
the holes in the P material are pulled towards the negative terminal. This
creates a brief, small current flow. This process widens the depletion region,
and once it matches the supplied potential, the current flow comes to a halt.
Essentially, it’s like enlarging the energy barrier. Increasing the source voltage
only exacerbates the situation, causing the depletion region to expand to
accommodate the change. Ideally, with a reverse-bias voltage applied, the PN
junction behaves as if it were an open circuit.

This asymmetric response to the applied potential proves to be incredibly


useful. One of the simplest semiconductor devices is the diode, which, at its
core, is essentially a PN junction. It serves as a component that readily allows
current to flow in one direction while effectively blocking current flow in the
opposite direction.

5.3 The Shockley Equation


Lecture 3 (2 hours)
We can quantify the behavior of the PN junction through the use of an equation 11th November 2024
derived by William Shockley.
 VD q 
I = IS e ηk B T
−1 (29)

where,

I is the diode current, I S is the reverse saturation current, VD is the voltage


across the diode, q is the charge on an electron, 1.6 × 10−19 coulombs, η is
the quality factor (typically between 1 and 2), k B is the Boltzmann constant,
1.38 × 10−23 joules/kelvin, and T is the temperature in kelvin.

19
5. the p-n junction

q
At 300 kelvin, the value of k T is approximately 38.6. Consequently, for
B
even very small forward (positive) voltages, the “-1” term can be ignored. Also,
I S is not a constant. It increases with temperature, approximately doubling
for each 10◦ C rise in T.

Plotting the Shockley equation with typical values for a silicon device
results in the curve depicted in Fig. (10. This graph illustrates the junction
current in relation to the forward (positive) voltage applied to the device. It’s
important to note that this curve serves as a representative example. While
all silicon diodes exhibit a similar overall shape, the exact current value for
a specific voltage will vary based on the device’s design. At voltages below
approximately 0.5 volts, the current is virtually negligible. However, beyond
this threshold, the current experiences a rapid increase, almost reaching a
vertical slope at around 0.7 volts. If the graph were recreated at a higher
temperature, it would shift the curve to the left, resulting in higher current
levels for a given voltage.

Figure 10: V-I Characeristics of a forward biased PN junction.

When it comes to negative voltages (reverse-bias), the Shockley equation


predicts minimal diode current, up to a certain point. It’s important to note
that this equation does not account for breakdown effects. When the reverse
voltage exceeds a certain threshold, the diode will begin to conduct, as demon-
strated in Fig. (11). In the first quadrant of the graph, we observe a similar

20
5.4 . the bands at operation

overall shape to what was seen in Fig. (10). Here, VF represents the forward
”knee” voltage, which is approximately 0.7 volts for silicon. I R stands for the
reverse saturation current, ideally close to zero but with a very small real-
world current flow. VR signifies the reverse breakdown voltage, and it’s worth
noting that the current experiences a rapid increase once this reverse voltage
is reached.

Figure 11: Simplified forward and reverse I-V curve for diode.

5.4 The Bands at operation


To help in our ability to picture what is going on, we will often add to this
band diagram, some small signed circles to indicate the presence of mobile
electrons and holes in the material. Note that the electrons are spread out in
energy. We know they like to stay in the lower energy states if possible, but
some will be distributed into the higher levels as well. What is distorted here
is the scale. The band-gap for silicon is 1.1 eV, while the actual spread of the
electrons would probably only be a few tenths of an eV, not nearly as much
as is shown in Figure (12a). Let’s look at a sample of p-type material, just for
comparison. Note that for holes, increasing energy goes down, not up, so their
distribution is inverted from that of the electrons. You can kind of think of
holes as bubbles in a glass of soda; they want to float to the top if they can.
Note also for both n-type and p-type material there are also a few ”minority”
carriers, or carriers of the opposite type, which arise from thermal generation
across the band-gap.
We are now ready to make an actual useful device! Let’s take a piece of
n-type material, and a piece of p-type material, and stick them together, as
shown in Figure (13). This way we will be making a p-n junction, or diode,
which will be our first real electric device other than a simple resistor.

21
5. the p-n junction

(a) Band diagram for an n-type semi- (b) Band diagram for an p-type semi-
conductor. conductor.

Figure 12: Career Concentration approximations in P and N type semiconduc-


tors

Figure 13: Career Concentration approximations in a P-N Junction. The


Dashed line shows the position of the Fermi Level.

There are a couple of things wrong with Figure (13). First of all, one of the
rules regarding the Fermi level is that when you have a system at equilibrium
(that is, when it is at rest, and is not being influenced by external forces such
as thermal gradients, electrical potentials etc.), the Fermi level must be the
same everywhere. Secondly, we have a big bunch of holes on the right and a
big bunch of electrons on the left, and so we would expect, that in the absence
of some force to keep them this way, they will start to spread out until their
distribution is more or less equal everywhere. Finally, we remember that a hole
is just an absence of an electron, and since an electron in the conduction band
can lower the system energy by falling down into one of the empty hole states,
it seems likely that this will happen. This process is called recombination.
The place where this is most likely to occur, of course, would be right at the

22
5.4 . the bands at operation

junction between the n and p regions. This is shown in Figure (14).

Figure 14: Recombination of holes and electrons.

Now is might seem that this recombination effect might just go on and on,
until there are no carriers left in the sample. This is not the case, however. In
order to see what brings everything to a halt, we need yet another diagram.
Figure (15) is more physical than what we have been looking at so far. It is a
picture of the actual p-n junction, showing both the holes and the electrons.
We also need to put in the donors and acceptors, however, if we want to see
what goes on. The fixed (meaning they can’t move around) charges of the
donors and acceptors are represented by simple ”+” and ”-” signs. They are
arranged in a nice lattice-like arrangement to remind us that they are stuck
to the crystal lattice. (In reality however, even though they are stuck in the
crystal lattice, there are so few of them compared to the silicon atoms that their
distribution would be quite random.) For the mobile holes and electrons, we
will stay with the little circles with charge signs in them. These are randomly
distributed, to remind us that they are free to move about the crystal.

Figure 15: Spatial schematic of a p-n junction.

We will now have to allow some of the holes and electrons (again near the
junction) to recombine. Remember, when an electron and a hole recombine,
they both are annihilated and disappear. Note that this process conserves

23
5. the p-n junction

charge and (if we could calculate it) momentum as well. There is obviously
some energy lost, but this will simply show up as vibrations, or heat, within
the crystal lattice — or, in the case of an LED, as light emitted from the
device. See, already we know enough about semiconductors to understand
(somewhat) how an actual device works. Light coming from an LED is simply
the energy which is released when an electron and hole recombine. We will
take a look at this in more detail later. Let’s allow some recombination to
occur, as shown in Figure (16).

Figure 16: The junction after some recombination has occurred.

If you look closely at these pictures, you will notice something. As we


remove more and more electrons and holes, we are starting to ”uncover” the
fixed charges associated with the donors and acceptors. We are making what is
known as a depletion region, so named because it is depleted of mobile carriers
(holes and electrons). The uncovered net charge in the depletion region is
separated, with negative charge in the p-region, and positive charge in the
n-region. What will such a charge separation give rise to? Why, an electric
field! Of course! Which way will the field point? The electric field which arises
from a separation of charges always goes from the positive charge, towards
the negative charge. This is shown in Figure (17).

Figure 17: The p-n junction with the resultant built-in electric field.

What effect will this field have on our device? It will have the tendency to
push the holes back into the p-region and the electrons into the n-region. This
is just what we need to counteract the recombination which has been going

24
5.4 . the bands at operation

on, and hopefully bring it to a stop.

Now try to think through what effect this field could have on our energy
band diagram. The band diagram is for electrons, so if an electron moves from
the right hand side of the device (the n-region) towards the left hand side (the
p-region), it will have to move through an electric field which is opposing
its motion. This means it has do some work, or in other words, the potential
energy for the electron must go up. We can show this on the band diagram by
simply shifting the bands on the left hand side upward, to indicate that there
is a shift in potential energy as electrons move from right to left across the
junction.

Figure 18: Energy band diagram for a p-n junction at equilibrium.

The shift of the bands, which is just the difference between the location of
the Fermi level in the n-region and the Fermi level in the p-region, is called
the built-in potential, VBI . This built-in potential keeps the majority of holes in
the p-region, and the electrons in the n-region. It provides a potential barrier,
which prevents current flow across the junction. (On the band diagram we
have to multiply the built-in potential VBI by the charge of an electron, q, so
that we can represent the shift in energy in terms of electron volts, the unit of
potential energy used in band diagrams.)

How big is Vbi ? This is not too hard to figure out. Let’s look at Figure (18)
a little more carefully. Remember, we know that since n = Nd in the n-region
and n = Na in the p-region, we can relate the distance of the Fermi level from
Ec and Ef by
!
Nc
Ec − Ef = kT ln (30)
Nd

and
!
Nv
Ef − Ev = kT ln (31)
Na

25
5. the p-n junction

Look at Figure (18) and see if you can agree that


   
qVBI = Eg − Ec − Ef − Ef − Ev
! !
Nc Nv
= Eg − kT ln − kT ln
Nd Na
!
Nc Nv
= Eg − kT ln .
Nd Na

where Nd and Na are the doping densities in the n and p regions, respec-
tively. Remember that kT = 1/40 eV = 0.025 eV, Eg = 1.1 eV, and both Nc and
Nv are approximately 1019 .

1038
!
qVBI = 1.1 eV − 0.025 eV · ln (32)
Nd Na

Here the q in front of the VBI and the e in eV are both the charge of 1
electron, and they cancel out, making

1038
!!
VBI = 1.1 − 0.025 ln volts (33)
Nd Na

Assuming the dimension of Nc and Nd in the order of 1015 , one can easily
find the dimention of the Built in Potential.

5.5 Biased P-N Junction

Now let’s take a look at what happens when we apply an external voltage to
this junction. First we need some conventions. We make connections to the
device using contacts, which we show as cross-hatched blocks. These contacts
allow the free passage of current into and out of the device. Current usually
flows through wires in the form of electrons, so it is easy to imagine electrons
flowing into or out of the n-region. In the p-region, when electrons flow out
of the device into the wire, holes will flow into the p-region (so as to maintain
continuity of current through the contact.) When electrons flow into the p-
region, they will recombine with holes, and so we have the net effect of holes
flowing out of the p-region.

With the convention that a positive applied voltage means that the termi-
nal connected to the p-region is positive with respect to the terminal connected
to the n-region. This is easy to remember: ”p is positive, n is negative”. Let us
try to figure out what will happen when we apply a positive applied voltage
Va . If Va is positive, then that means that the potential energy for electrons
on the p-side must be lower than it was under the equilibrium condition. We
reflect this on the band diagram by lowering the bands on the p-side from

26
5.5 . biased p-n junction

Figure 19: A p-n diode with contacts and external bias.

where they were originally. This is shown in Figure (20).

Figure 20: A p-n junction under forward bias.

As we can see from Figure (20), when the p-region is lowered a couple of
things happen. First of all, the Fermi level (the dotted line) is no longer a flat
line, but rather it bends upward in going from the p-region to the n-region.
The amount it bends (and hence the amount of shift of the bands) is just
given by qVa , where the energy scale we are using for the band diagram is in
electron-volts which, as we said before, is a common measure of potential
energy when we are talking about electronic materials. The other thing we can
notice is that the electrons on the n-side and the holes on the p-side now ”see”
a lower potential energy barrier than they saw when no voltage was applied.
In fact, it looks as if a lot of electrons now have sufficient energy such that they
could move across from the n-region and flow into the p-region. Likewise, we
would expect to see holes moving across from the p-region into the n-region.

This flow of carriers across the junction will result in a current flow across
the junction. In order to see how this current will behave with applied volt-
age, we have to use a result from statistical thermodynamics concerning the
distribution of electrons in the conduction band, and holes in the valence

27
5. the p-n junction

band. We know that the electrons tend to fill in the lowest states first, with
fewer and fewer of them as we go up in energy. For most situations, a very
good description of just how the electrons are distributed in energy is given
by a simple exponential decay. (This comes about from a statistical analysis
In particle physics, a of electrons, which belong to a class of particles called Fermions. Fermions
fermion is a subatomic
particle that has a half-odd-
have the properties that they are: (a) indistinguishable from one another; (b)
integer spin and follows the obey the Pauli Exclusion Principle, which says that two Fermions can not
Pauli exclusion principle
and Fermi–Dirac statistics.
occupy the same exact state (energy and spin); and (c) remain at some fixed
Some examples of fermions total number N.)
include: protons, neutrons,
electrons, neutrinos, quarks,
tritium, helium-3, and If n(E) tells us how many electrons there are with an energy greater than
uranium-233. Fermions are some value Ec , then n(E) is given simply as:
one of two types of particles
that make up matter, the
E−Ec
other being bosons. The
term ”fermion” comes from n(E) = Nd e− kT (34)
Enrico Fermi and was first
used in 1947.
The expression in the denominator is just Boltzmann’s constant times
the temperature in Kelvins. The value of Boltzmann’s constant is given by
1.380649 × 10−23 J · K−1 , or 8.617333262 × 10−5 eV · K−1 .
At room temperature, i.e., at 300 K, kT has a value of about:
1
8.617333262 × 10−5 eV · K−1 × 300 K ≈ eV or 25 meV.
40
This number is sometimes called the thermal voltage, VT , but it’s okay to
think of it as a constant that comes from the thermodynamics of the prob-
1
lem. Because kT ≈ 40 eV, you will sometimes see Equation (34) and similar
equations written as:

n(E) = Nd e−40(E−Ec ) (35)

This can look a little strange if you forget where the 40 came from, and
just see it sitting there.

If the energy E is Ec , the energy level of the conduction band, then n(Ec ) =
Nd , the density of electrons in the n-type material. As E increases above Ec ,
the density of electrons falls off exponentially, as depicted schematically in
Figure (21).

Now let’s go back to the unbiased junction. Remember, as we said before,


there are currents flowing across the junction, even if there is no bias. The
current we have shown as If is due to those electrons which have an energy
greater than the built-in potential. They are flowing from right to left, as
shown by the open arrow, which, of course, gives a current flowing from left
to right, as shown by the solid arrows. Based on Equation (34) the current
should be proportional to:

28
5.5 . biased p-n junction

Figure 21: Distribution of electrons in the conduction band with energy.

qVbi
I f ∝ Nd e − kT (36)

Figure 22: Balanced flow across a junction.

The principle of detailed balance says that at zero bias, If = −Ir , and so
 qVbi

Ir ∝ − Nd e− kT (37)

   qVbi
Ir = − If α − Nd e− kT (38)

Now, what happens when we apply the bias? For the electrons over on the
n-side, the barrier has been reduced from a height of qVbi to q (Vbi − Va ) and
hence the forward current will be significantly increased.
qVbi
I f ∝ Nd e − kT (39)

The reverse current, however, will remain just the same as it was before.

29
5. the p-n junction

Figure 23: Current when the junction is forward biased.

The total current across the junction is just


 qV 
  a
If + Ir ∝ Nd e− kT − 1 (40)

qVbi
where we have factored out the Nd e− kT term out of both expressions. We
are not prepared, with what we know at this point, to get the other terms in
the proportionality that are involved here. Also, the astute reader will note
that we have not said anything about the holes, but it should be obvious that
they will also contribute to the current, and the arguments we have made for
electrons will hold for the holes just as well.
We can take the effect of the holes, and the other unknowns about the
proportionality, and bind them all into one constant called Isat so that we
write:
 qV 
a
I = Isat e kT − 1 (41)

Which is, once again, the famous Shockley equation.

5.6 The phenomenon of Breakdown


As a general rule, diodes should not be operated in the breakdown region,
with the exception of Zener diodes. Two underlying mechanisms account
for this behavior. When doping levels are high, and breakdown voltages are
typically below five or six volts, the Zener effect, named after Clarence M.
Zener, takes precedence. It arises from the generation of an extremely high
electric field across the depletion region, leading to a significant current
through electron tunneling. Conversely, in devices with lower levels of doping,
avalanche breakdown becomes dominant. In this scenario, a strong electric
field accelerates free electrons to the extent that they collide with nearby
atoms, forming new electron-hole pairs, consequently generating additional

30
5.7 . zener diodes

free electrons that can perpetuate this process. This results in a rapid and
substantial increase in current.

5.7 Zener diodes


Zener diodes are essentially diodes that operate in a reverse-biased mode while
being able to withstand breakdown conditions. As the reverse bias voltage
increases, Zener diodes maintain a constant level of current flow (referred to
as the saturation current) until a specific voltage, known as the breakdown
voltage (VZ ), is attained. When this breakdown voltage is reached, the diode
enters a state of breakdown and permits virtually any amount of current to
pass through. Consequently, during breakdown, the magnitude of the current
is dictated by the other components of the circuit, such as effective resistance
and current sources. The breakdown voltages for Zener diodes can vary within
the range of 1 to 100 volts.
Breakdown in Zener diodes can be attributed to two distinct yet closely re-
lated mechanisms: the avalanche effect and the Zener effect. The avalanche ef-
fect occurs when the potential difference across the p-n junction reaches a crit-
ical level, causing free electrons traversing the junction to acquire enough en-
ergy to dislodge other covalently-bonded electrons through collisions, thereby
generating new electron-hole pairs. This process propagates in a chain reac-
tion, leading to a rapid and substantial surge in charge carriers, often likened
to an ”avalanche,” resulting in an abrupt increase in current flowing through
the diode.

Figure 24: Circuit symbol for a Zener diode.

Conversely, the Zener effect comes into play when the electric field created
by the space charge region attains a level of intensity that can forcefully extract
covalently-bonded electrons from their bonds. This also leads to the creation
of new electron-hole pairs, which are swiftly separated by the potent electric
field. When the electric field becomes powerful enough to simultaneously
separate numerous electrons and holes, it triggers a significant upsurge in
current.
Zener diodes find practical applications in the field of electronics due to
their ability to handle substantial current flows, enabling them to dissipate
substantial power (P=IV). Additionally, they serve a crucial role in voltage
regulation, a function where they maintain a constant output voltage even
when the input voltage varies. A basic voltage regulator can be created by
incorporating a Zener diode in series with a resistor.

31

You might also like