Вы находитесь на странице: 1из 44

CHAPTER 1

INTRODUCTION

The knowledge behind the electric power generated by the direct conversion of solar

energy in the past was low simply because little attention was paid to what was space

technology, which was also utilising this solar energy conversion. But its potential is

gaining prominence and as a result, the physical processes involved are looked into.

Solar cells are basically solid-state (especially p-n diode) devices, which are made from

semiconductors. As such, the physics of the photovoltaic effect are linked with some

aspects of the semiconductor physics. The explanation used will be based on the classical

theory of a non-illuminated diode developed by Schottky and Spenke as the crystal

(material), the dopant and the junction are linked with the diode operation.

A quick look at the history shows that Becquerel first observed photovoltaic effect in an

electrolyte in 1839. This PV effect was then discovered in 1876, on a semiconductor

(selenium) by Adams and Day. But the PV effect could not be explained until in 1905,

when Einstein used the quantum theory developed by Planck. Light was regarded as

travelling in wave packets (photons), which upon striking a material will cause the

material to eject electrons. This, of course, could only take place if the photon energy

(which depends on the frequency) was high as some of its energy would be transferred to

the electrons. Many experiments were later conducted which led to the creation of solar

cells and were used in a satellite (VANGUARD) in 1958 and for terrestrial applications

such as navigation lights and communication stations in 1956.

Materials are classified as conductors (metals), insulators and semiconductors based on

their current capabilities. In a conductor, electrons in the outer shell (1 or 2) of an atom


are always free to roam about and hence conduct electricity. The reverse is the case in an

insulator as the electrons in the outer shell (6 or 7) are held tightly to the atom and do not

conduct electricity. For a semiconductor, which is in between the conductor and the

insulator, it has some electrons that become loose and roam about once there is a

temperature increase. The fraction of electrons f, having energy equal to or greater than E

is f = exp ( - E / kT ). This is as a result of the average energy, known as the energy gap,

Eg, acquired allowing its conductivity to rise. The conductivities of these materials are

given below (Table 1.1).

Table 1.1. Conductivities of materials [2]

Material Conductivity (Siemens/cm-1)

Conductor >106

Insulator < 10-10

Semiconductor 10-8 – 102

2
CHAPTER 2

SEMICONDUCTOR MATERIALS

Semiconductors exist in crystalline, polycrystalline or amorphous form. The table below

shows their atomic/molecular arrangement (Table 2.1).

Table 2.1. Arrangement of atoms/molecules in semiconductors

Crystalline Polycrystalline Amorphous

The atoms/molecules are The atoms are also The atoms are not an

symmetrically arranged symmetrically arranged but the ordered array on any

and the array periodic arrays are not well uniformed macroscopic scale

throughout the crystal with the atomic order and

regularity being totally

irregular

Since silicon is very common element, focus will be given to their study. As the silicon

solar cells operate in the same way as other solar cells; they will be used as an example.

Pure, intrinsic silicon (i.e., silicon having the desired characteristic) has very little

impurities (the impurities have a concentration of less than 1018 m-3) and electrical

resistivity ρ, of 2500 Ωm with four electrons at their outer shell, which have a covalent

bond with other neighbouring silicon atoms. As their crystals have different crystal

arrangements, their physical properties are not the same and as such affects the working

operation of the cells made from them. The properties of silicon can be looked in [2]. In a

3
silicon crystal, each atom is equally spaced with four other atoms and all have four

electrons in their outer shells that are covalently bonded as seen in Figure 2.1.

Figure 2.1. Silicon at absolute zero (Source: Savant C. J. et al, (1991) Electronic Design: Circuits and
System, 2nd ed., The Benjamin/Cummings Company Inc., California)

At absolute zero, the crystal behaves like an insulator. In terms of energy bands, the

valence band is occupied by the electrons while the conduction band is completely

empty. As a result, since they cannot move to the conduction band, there is no change in

electrical conductivity. But if a strong external energy is applied, the electrons that are

covalently bonded acquire this energy and become free to move leaving in its initial

position a void, as the electron binding energy (forbidden energy band) is low. This void

(called a hole) assumes a positive charge and can be occupied by another electron from a

neighbouring silicon atom (Figure 2.2). This in turn creates another hole in the presence

of an electric field and we have equal number of electrons and holes.

Figure 2.2. Conduction from broken covalent bond (Source: Savant C. J. et al, (1991) Electronic Design:
Circuits and System, 2nd ed., The Benjamin/Cummings Company Inc., California)

4
As can be seen in the band structure, the electrons move in the conduction band after

transferring from the valence band while the holes move about in the valence band from

the conduction band. This is shown below (Figure 2.3).

Figure 2.3. Band structure of intrinsic semiconductor (Source: Gibbs K. (1996) Advanced Physics, 2nd ed.,
Cambridge University Press, Cambridge)

The electrical conductivity can equally be affected by the addition of impurities

(commonly known as dopants) to the crystal of the host semiconductors. Such

semiconductors are termed extrinsic semiconductors. The dopants have energy levels

lying in the band gap of the parent material and help increase the number of electrons or

holes so that they become majority carriers. We have n-type doped crystals where the

dopants used are called donors. These donors are the Group V elements of the periodic

table having five valence electrons in their outer shell, that is, one electron more than that

of the silicon. They are antimony, phosphorus, arsenic and bismuth and are close to the

conduction band. When thermal lattice energy is applied, the excess electron is separated

thereby moving freely within the crystal. So the freely moving electrons are more than

the holes. For a p-type crystal, the dopants are called acceptors, which are close to the

valence band. They include the Group III elements such as boron, aluminium, gallium,

indium and thallium and are one electron less than silicon meaning three electrons at their

outer shell. It readily accepts an electron donated by the silicon without using energy as

5
such a hole is obtained. This hole is free to move about in the valence band of the crystal

(Figure 2.4).

Figure 2.4. Crystal structure of doped semiconductors (a) n-type (excess electron) silicon
(b) p-type (deficient electron) silicon (Source: Savant C. J. et al, (1991) Electronic Design: Circuits and
System, 2nd ed., The Benjamin/Cummings Company Inc., California)

The semiconductor junction is obtained when the p-type makes contact with the n-type

material in the crystal. The contact (i.e., junction) acts as a barrier to both electrons and

holes if the potential energy of the electrons (or holes) is greater in the p- (or n-) region

than in the n- (or p-) region. The presence of equal concentrations of impurities also

forms the junction (Figure 2.5).

Figure 2.5. Semiconductor junction [5]

There is diffusion across the junction whenever the carriers are not evenly distributed

with the resultant effect of a field, which causes current to flow in a direction opposite to

the diffusion current. There are several ways of obtaining this p-n junction and they

6
include: (a) ion implantation, (b) diffusion, (c) growing n-type impurity in p-type crystal

or vice versa, (d) alloy junction.

7
CHAPTER 3

PHOTON ENERGY

The radiation from the sun consists of an almost continuous range of wavelengths from

less than 1 nm to many hundreds of metres. The solar spectrum is from X-rays through

visible light to low-energy microwaves but only wavelengths of between 0.25 μm in the

ultra-violet of the solar spectrum and 3.0 μm in the far infrared spectral region have about

98 % of the total emitted solar energy (Table 3.1, Figure 3.1) and so the sun is treated as a

heat radiator.

Table 3.1. Distribution of solar intensity by wavelength [3]

Wavelength interval (μm) Percentage of solar energy in interval

< 0.3 0

0.3-0.5 17

0.5-0.7 28

0.7-0.9 20

0.9-1.1 13

>1.1 22

The absorption lines seen in the solar spectrum is as a result of the reversing layer of the

sun absorbing some of the photosphere’s emitted light. The solar constant S, is defined as

the quantity of energy per unit time passing through unit area at right angles to the

direction of the solar beam measured outside the earth’s atmosphere, that is, it represents

the radiant energy flux integrated over wavelength reaching the top of the atmosphere at

the mean earth-sun distance. It therefore has to do with the solar intensity outside the

8
earth’s atmosphere, which has a value of 1370 W/m2 (according to satellite

measurements).

Figure 3.1. Sensitivity range of silicon solar cells, compared with the solar spectrum outside the atmosphere
and at ground level [4]

The Air Mass is the changes in the sun’s beam path length as a result of the changes in

the sun’s zenith angle during the day. Therefore, Air Mass Zero (AM-0) means that when

the sun is at zenith the solar radiation is at sea level, which means the conditions are the

same to solar intensity of 1370 W/m2. A better way to put it, AM-0 is the solar light

outside the atmosphere. Its curve is called the solar spectral irradiance curve. The

9
maximum light on the ground at sea level is called Air Mass One (AM-1) and its solar

intensity is nearly 1000 W/m2. In high mountains, the intensity increases to 1070 W/m2

(Figure 3.2).

Figure 3.2. Relationship of air mass 1, 2 and 3 and declination is shown (Source: Hunt D. V. (1982) Solar
Energy Dictionary, Industrial Press Inc., New York)

Solar light consists of minute energy packet of electromagnetic radiation called photons

and the energy of a photon depends on radiation frequency. When an electron interacts

with light, it gains energy and this photon energy is given by Ef = hν where h is Planck’s

constant, which has a value of 6.63 x 10 -34 Js and ν is the light frequency. Once this

electron has this energy, which must be greater than the energy gap of the electron, it will

escape from the atom. This implies that the number of photons alongside with the solar

radiation energy must be known beforehand.

As a large portion of solar irradiance can be found in the far infra-red region, solar cells

cannot exploit the energy there as they are insensitive to that part of the solar spectrum

(as the crystal melts or the electrons will not be excited). Instead, the different spectral

10
composition at sea level has the effect of concentrating more of the incoming energy in

the part of the spectrum where solar cells are sensitive (Figure 3.3).

Figure 3.3. Spectral response curve of a conventional silicon solar cell [4]

11
CHAPTER 4

ELECTRON-HOLE CONCENTRATION AND FERMI LEVEL

Electrons are usually excited to higher state when they interact with phonons and

photons. But the number of electrons that can be allowed at any excited energy level is

determined by Pauli Exclusion Principle. This energy level is called the Fermi level. If

energy is given to electrons that is above the Fermi level, the distribution of electrons in

the allowed levels can be described using the Fermi-Dirac distribution function f ( E ),

given as

f (E) = 1 / { 1 + exp [ (E – EF ) / kT ] } (4.1)

where E is the energy of an allowed state, EF is the Fermi energy, k is Boltzmann’s

constant and T is the absolute temperature. At T = 0 K, f ( E ) = 1 up to an energy EF. The

Fermi level EF, is any energy level having the probability that is exactly one-half filled

with electrons. In an intrinsic (pure) semiconductor, the Fermi level is halfway in

between the energy gap, that is, EF = Eg / 2 so there are equal concentrations of electrons

and holes (Figure 4.1(a)). But for the extrinsic semiconductors, which have been doped,

the Fermi levels are in different positions, whether it is an n- or p-type doped crystal as

seen in Figures 4.1 (b) and 4.1 (c) respectively. For the n-type doped crystal, the Fermi

level is closer to the donor level ED while for the p-type doped crystal, the Fermi level is

closer to the acceptor level EA.

12
Figure 4.1. Energy level diagram of semiconductor [2]

The exact locations of the Fermi level can be determined once the doping levels and the

absolute temperature are known. The energy difference for the n-type doped

semiconductor (EC – ED) is of the order of the thermal energy kT allowing electrons to

become excited into the conduction band. For the p-type doped semiconductor, the

electron energy in the acceptor EA, which is approximately equal to kT above the valence

band allows for electrons to be excited into the acceptor atom leaving behind a hole.

When there is thermal equilibrium in the n-type doped crystal, the electron density in the

conduction band n, is given as

n = NC exp [ ( EF – EC ) / kT ] (4.2)

where NC is called the effective density of states in the conduction band and is given as

NC = ( π / 2 )1/2 ( mdk )3/2 T3/2 / π2 h3 (4.3)

where md is the effective mass for the electrons in the conduction band and h being

Planck’s constant. Also, the density of holes in the valence band p, in a p-type material is

given as

p = NV exp [ ( EV – EF ) / kT ] (4.4)

where NV, called the effective density of states in the valence band is given as

13
NV = ( π / 2 )1/2 ( mvk )3/2 T3/2 / π2 h3 (4.5)

where mv is the effective mass for the holes in the valence band.

For an intrinsic semiconductor, since there are no impurities present, the number of holes

and electrons is equal. It is this number per unit volume that is called the intrinsic carrier

concentration ni.

n = p = ni (4.6)

Therefore, the product of the number of holes and electrons is constant when the

semiconductor is in thermal equilibrium, that is,

np = ni2 = NC NV exp [ - ( EC – EV ) / kT ]

= NC NV exp [ - ( Eg ) / kT ] (4.7)

From equations (4.6) and (4.7), the Fermi energy level can be obtained as

EF = ( EC + EV ) / 2 + kT / 2 ln ( NV / NC ) (4.8)

The implication of the above equation is that the Fermi level is exactly halfway in

between the conduction and valence band once they have equal numbers of electrons and

holes and moves closer to either of the bands whenever the density of states rises or falls.

Since the total macroscopic change neutrality is neither altered nor affected by the

addition of impurities, it fulfils the following condition:

p – n = N -A – N +D (4.9)

where N -A and N +D are the ionised acceptor and donor atoms and are given as

N -A = NA / { 1 + exp [ (EA – EF ) / kT ] } (4.10)

and

N +D = ND exp [ (ED – EF ) / kT ] / { 1 + exp [ (ED – EF ) / kT ] } (4.11)

14
where EA, ED are the electron energies of the acceptor and donor atoms while NA, ND are

the acceptor and donor impurity concentrations respectively.

The presence of impurities allows for the determination of the Fermi level position using

equations (4.2) and (4.4). Therefore,

ND = n = NC exp [ ( EF – EC ) / kT ]

Then,

ND / NC = exp [ ( EF – EC ) / kT ]

ln ND / NC = [ ( EF – EC ) / kT ]

EF – EC = kT ln ( ND / NC ) (4.12)

and

NA = p = NV exp [ ( EV – EF ) / kT ]

EV – EF = kT ln ( NA / NV ) (4.13)

The number of charge carriers determines the electrical conductivity σ, and is given as

the sum of both hole and electron conductivities;

σ = σn + σp = q (n μn + p μp ) (4.14)

where q is the electronic charge and μn and μp are the electron and hole mobilities

respectively and is given as

μ=v/E (4.15)

where v is the charge carrier velocity and E is the applied field. In a semiconductor, the

electron or hole is always in random motion colliding with other atoms of both the

material and the impurity and other charge carriers. The mobility depends on the impurity

concentration and temperature if the electric field is small. This is because the mobility

reduces if the semiconductor is heavily doped.

15
CHAPTER 5

A P-N JUNCTION

An n-type material is made to have contact with a p-type material forming the p-n

junction ( ) in a crystal. Prior to this formation, both materials are electrically

neutral. In the n-type material, the electrons diffuse to the p-type material through the

junction. This is as a result of a potential gradient. The diffusion current in, that results, is

then given as

in = q Dn dn / dx (5.1)

where x is the distance in the crystal and Dn is the diffusion constant, which is obtained

using the relation

Dn = μn kT / q (5.2)

Symmetrically, a diffusion current of holes ip, flows in the opposite direction;

ip = q Dp dp / dx (5.3)

Because of the diffusion of electrons to the p-type material, the n-type material is now

positively charged while the p-type material is negatively charged since there is a

deficiency of holes. As a result, we have an induced electric polarisation in an

unilluminated p-n junction. The diffusion effect of the charge carriers is shown (Figure

5.1).

16
Figure 5.1. Electron and hole concentration at a p-n interface in thermal equilibrium; currents across the
interface [4]

The depletion or space-charge region is where the potential gradient between the n- and

p-regions is concentrated and this is where the electric field is created. This drift-field E,

gives rise to a field current of electrons and holes and are written as

in = qμnnE; ip = qμppE (5.4)

As such, the electric field removes the minority carriers that come close to the region. If

the barrier (that is, the junction) is decreased, the current from the flow of the majority

carriers rises sharply since it is the barrier that is limiting the current. The implication of

this is that there is a zero net current when there is no applied external field across the

junction. Also, it is difficult to measure the potential barrier externally due to exactly

equal and opposite potential barriers that appear in the interface.

The concentration of the charge carriers at steady-state equilibrium is a function of the

distance inside depletion region and can be written as a Boltzmann distribution:

n(x) = nn exp ( + qV (x) / kT ); p(x) = pp exp ( - qV (x) / kT ) (5.5)

17
where nn, pp are the electron and hole equilibrium concentrations in the n- and p-layers, V

(x) the junction built-in electric potential and n, p the charge carrier concentrations.

The diffusion voltage VD, is defined as the total voltage across the junction usually of the

order of 1 V and is bigger than the photovoltage as seen later.

There is an energy difference qVD, which separates the charged carriers. Consequently,

there are bent energy bands at the junction resulting in an electric field and diffusion

voltage at the region. The charge carriers move in opposite directions to both sides of the

junction as a result of the energy qVD they have gained, that is, the electric field hastens

the movement of the carriers, which may drift through both sides of the junction. Hence,

the energy bands are shifted by an amount qVD (the conduction band EC is higher on the

p-side than on the n-side) as shown below (Figure 5.2).

Figure 5.2. Band diagram for crystal as in Figure 5.1 [4]

The diffusion voltage and the related energy qVD can also be described in terms of the

Fermi level. The energy shift as seen in Figure 5.2 is said to be as a result of the Fermi

levels in the n- and p-regions, which are at the same energy according to the Fermi level

definition (energy at which the occupancy of energy states with electrons is a half). The

18
electrons move from the valence band to the conduction band due to thermal energy and

so we have more carriers above (for the electrons) and below (for the holes) the Fermi

level whenever there is an increase in the temperature. The Fermi level is closer to the

conduction band for an n-type material (with donors) that has a high concentration of

electrons while the Fermi level is closer to the valence band for a p-type material since

the amount of electrons in it is low. This can be seen in Figure 5.2. Above a certain

temperature, the Fermi level lies between the n- and the p-layer since more electrons

become thermally ionised from the valence band to the conduction band. This is a

consequence of the intrinsic transitions where the crystal behaves as if it is not doped and

so the junction disappears since there is an equal amount of electrons in the conduction

band. The photovoltage then vanishes when the crystal is under light.

At thermal equilibrium, the rate at which electrons are occupied at any given energy must

be equal throughout the material. Therefore, thermal equilibrium means homogeneous

Fermi level. This thermal equilibrium can be disrupted once a voltage or light is applied.

A charge dipole is created near the junction as a result of the field caused by the electrons

in the p-type material. So in thermal equilibrium there is no net current flow and therefore

no energy in the dipole, only a voltage.

Attention is turned to the energy bands of a silicon solar cell. The silicon cells are made

from a p-type raw material consisting of monocrystalline platelets having a resistivity

approximately 1 Ωcm doped with boron at a concentration of approximately 1016 cm-3.

The n-layer is obtained by diffusing phosphorus into the surface in such a way that a

concentration of about 10-19 cm acceptors extends to a depth of a fraction of a micrometer

from the surface (Figure 5.3).

19
Figure 5.3. Schematic cross-section of conventional silicon solar cell [4]

The energy gap for silicon is 1.12 eV at room temperature. Referring to Figure 5.4, it can

be seen that the distance of separation between the Fermi level and valence band is 0.2

eV in the p-layer and the Fermi level is nearly on the same level with the conduction

band in the n-layer.

Figure 5.4. Band diagram of a practical silicon solar cell, dark condition [4]

The diffusion voltage can be obtained since the energy gap for silicon is already known.

20
The value is nearly 0.9 eV (= 1.12 – 0.2 eV) at room temperature and is the photovoltage

upper limit.

Analysis of The Dark Currents Flowing In A P-N Junction When There

Is Voltage

Under ‘reverse bias’ condition with a positive d.c. voltage applied across the p-n junction

will cause no current to flow as a result of movement of the majority carriers from one

material to another. But there is a small current flow across the junction as a result of the

minority carriers (Figure 5.5).

Figure 5.5. p-n junction. (a) Reverse bias (b) Forward bias [Source: Basic Electronics and Linear Circuits,
Bhargava et al., (1984) TaTa McGraw-Hill Publishing Co. Ltd., New Delhi]

For a ‘forward bias’ condition with a negative voltage applied to the n-layer, the internal

energy barrier qVD drops. This results in reduced field currents as the electrons can easily

enter into the p-layer likewise vice versa for the holes. With the forward voltage equal to

VD and with much majority carriers on both sides of the junction, the barrier layer

vanishes and current rises sharply (Figures 5.5 & 5.6).

21
Figure 5.6. Band diagram of a practical silicon cell in the dark with a voltage of 0.5 V applied to the n-layer
(forward characteristic) [4]

The electron and hole currents can be described using Figures 5.1 & 5.3. At the point of

contact xi, the value of the electric current is the same as that of the hole current that is if

there is an equal amount of doping in both n- and p- materials of the crystal. Using the

Boltzmann distribution, at the point xn, the low concentration of holes at equilibrium is

given as

p(xn) = pn = pp exp ( - qVD / kT ) (5.6)

This concentration changes once there is an applied voltage VF. Thus,

p(xn)F = pp exp { - ( q [ VD - VF ] / kT ) } = pn exp ( qVF / kT ) (5.7)

The excess holes p(xn)F – p(xn) diffuses into the n- layer some distance and recombine

with electrons in the crystal defects or with metal ions of the n-material. This distance is

proportional to the diffusion constant Dp, measured cm2s-1, and the lifetime of the holes

τp, which is defined as the times needed for the number of holes to decay to 1/ exp of the

previous values (here, it is assumed high). The average distance that the holes travel

before it is recombines is called the diffusion length Lp. As a result, the excess holes

concentration decays exponentially and the decay as a function of the distance from xn

can be written as

22
p(x)F = pn { exp ( qV / kT ) }{ exp ( - [ x - xp ] / Lp ) } (5.8)

Likewise, at xp of the p-layer, the decay of excess electrons concentration can also be

written as

n(x)F = np { exp ( qV / kT ) }{ exp ( - [ xn - x ] / Ln ) } (5.9)

The illustration of the net concentrations is shown (Figure 5.7 – upper part).

Figure 5.7. Schematic of a symmetrical diode in the dark with voltage V forward bias [4]

The current flow throughout the crystal is also shown in Figure 5.7 – bottom part – and it
is explained as follows:

1. Since it is assumed that the n- and p-materials have an equal concentration of

dopants, at each side of the centre of the junction, the electron current is

symmetrical to the hole current on the opposite side of the junction.

23
2. There is a diffusion current flow into the layers at xp and when there is an increase

in the applied voltage, the diffusion current rises exponentially. This current is the

limiting process for the current flow through the crystal.

3. Inside the layers is a field current that reduces towards the junction.

4. Inside the junction, the field and diffusion currents have higher values than in the

layers. The effective voltage in the junction is VD – Vapplied. Therefore, the field

current in the junction flows in the opposite direction to that in the layers, where

the applied voltage is ε x Vapplied (ε << 1).

The current-voltage (I-V) characteristic equation can then be derived based on the

schematic description of the diode currents. The diffusion current is written as

ip = - q Dp dpF / dx (5.10)

dpF / dx is derived using equation (5.8) and this is given as

dp(x)F / dx = - ( 1 / Lp ) pn { exp ( qV / kT ) }{ exp ( - [ x - xp ] / Lp ) }

dp(x)F / dx = - ( 1 / Lp ) p(x)F (5.11)

Therefore, the diffusion current can be rewritten as

ip = ( q Dp / Lp ) p(x)F (5.12)

In the n-layer, the amount of holes fed into it can be written as

p(x)F = p(x) – pn (5.13)

where p(x) is the total concentration and pn the concentration of holes in the n-layer in

thermal equilibrium. At xn, we have

ip(xp) = ( q Dp / Lp ) ( p(x) – pn ) (5.14)

Using equation (5.7), ip(xp) can be rewritten as

ip(xp) = ( q Dp / Lp ) pn { exp ( qVF / kT ) - pn }

24
= ( q Dp / Lp ) { exp ( qVF / kT ) - 1 } pn (5.15)

Since xp is symmetrical to xn, in is then given as

in(xn) = ( q Dn / Ln ) { exp ( qVF / kT ) - 1 }np (5.16)

ip(xp) = ip(xn) when it crosses the junction since only a very small number of holes are

lost. So also, in(xn) = in(xp) as there is no reduction when it crosses the junction. The total

current for a biased diode can then be computed as either ip(xp) + in(xn) or ip(xn) + in(xp),

which gives

iF = q [ (Dp pn / Lp ) + ( Dn np / Ln ) ] { exp ( qVF / kT ) - 1 } (5.17)

(adding equations (5.15) and (5.16)). The first term is called the saturation (or dark)

current:

iS = q [ (Dp pn/ Lp ) + ( Dn np/ Ln ) ] (5.18)

This saturation current, therefore, depends on the diffusion of the minority carriers. The

values of np and pn can be obtained using the expressions

np = ni2 / Np = ni2 / p; pn = ni2 / Nn = ni2 / n (5.19)

Dp = ( kT / q ) μp; Dn = ( kT / q ) μn (5.20)

Equation (5.20) is the Einstein relations. Lp and Ln are also obtained using the

expressions

Lp = (Dp τp)½; Ln = (Dn τn)½ (5.21)

So the saturation current can be rewritten as

iS = q [ pn (Dp / τp )½ + np (Dn / τn )½ ] (5.22)

It is, thus, a function of equilibrium densities of carriers, diffusion rates and carrier

lifetimes. Going back to equation (5.17), it is then rewritten as

25
iF = iS { exp ( qVF / kT ) - 1 } (5.23)

iF, which is the total current can equally be called the junction current or better still the

diode current under forward bias. So in all, because of the junction solar cells have diode

characteristics in the dark and this is important for the photovoltaic effect.

Some considerations have to be taken into cognisance when dealing with the practical

aspect such as the assumption that recombination is small everywhere in the crystal

which may not necessarily be so. It has been observed that, particularly within the space

region, recombination centres might have some effect on the current. As a result, the

recombination current as calculated by Shockley is given as

iR = iRO [ exp ( qVF / AkT ) – 1 ] (5.24)

where A is a constant.

Attention is now directed towards what happens when light is focussed on the p-n

junction.

The photovoltaic effect is responsible for the photocurrent generated by the solar cells.

When light is made to fall upon the semiconductor junction, each photon with EF > Eg

(as a result of internal ionisation) can cause one minority and one majority carrier to be

produced as electron-hole pairs in either of the two regions. The concentration of the

electrons and holes depends on the light absorption rate and the carrier lifetime.

Equilibrium is attained in the flow of both the majority and minority carriers as a result of

the decrease in the potential barrier caused by some of the minority carriers that have

been swept across the junction. With this decrease, the potential barrier (in volts) can be

measured externally as seen in Figure 5.8 (a).

26
(a) without external load (b) with external load

Figure 5.8. Solar cell equivalent [5]

In Figure 5.8 (a), the current in the diode IJ, is equal to the added photon-generated

minority carrier current Iph, so that

Iph = IJ (5.25)

If a load resistor is connected in parallel with the diode junction as in Figure 5.8 (b), the

photon-generated current Iph, will be divided between the diode and the resistor. The

electric current I, in the load resistor will be the difference between the diode current IJ,

and the photon-generated current Iph

Iph = IJ + ( - I (RL) ) (5.26)

(the minus sign indicates flow in the opposite direction).

I (RL) = IJ - Iph

= IS { exp ( qVF / kT ) - 1 } - Iph (5.27)

This is for an idealised solar cell. With the recombination current given from equation

(5.24), the electric current becomes

I (RL) = IS { exp ( qVF / kT ) - 1 } + IRO [ exp ( qVF / AkT ) – 1 ] - Iph (5.28)

By varying the load resistance between zero and infinity, the current-voltage I-V,

characteristic of the solar cell may be gotten. For a p-n junction not illuminated, the I-V

characteristic is that as shown as curve (a) in Figure 5.9. Curve (b) is for an illuminated

p-n junction.

27
Figure 5.9. Typical current-voltage characteristics of a solar cell [1, 2, 4]

Curve (b) is identical to curve (a) except that it is shifted along the negative current axis.

The fourth (or power) quadrant in Figure 5.9 is where the photovoltaic energy conversion

takes place and so gives the ideal I-V output of a p-n junction solar cell. Three parameters

are used in describing the behaviour of a solar cell and they include: (a) the open-circuit

voltage VOC, that appears on the voltage axis; (b) the short-circuit current ISC, that

appears on the negative current axis, and (c) the fill factor F.F. The Voc (that is, when I =

0) is the voltage output from a solar cell when the load impedance is bigger than the

device impedance and therefore the maximum possible voltage. But when the load

impedance is smaller than the device impedance, the current output is the maximum

possible current, which is the short-circuit current that flows in the reverse direction to

the forward dark current. A linear displacement along the current axis shows a linear

dependence of light intensity on the short-circuit current while the open-circuit voltage

increases as a logarithmic function of light intensity. The fill factor is the ratio of the

smaller rectangle to the larger rectangle.

28
The effect of an internal series resistance Ri, in a steady-state equivalent cell is

considered below (see Figure 5.10).

Figure 5.10. Steady-state equivalent circuit of a solar cell [5]

Because of the conduction resistance in the thin diffused layer on top of the photovoltaic

cell, there is an internal series resistance, the voltage seen by the external load is given as

V = I RL. The voltage drop, that is, the internal diode forward bias voltage VF, is the same

as that seen in the series resistance. Therefore,

VF = - I ( RL + Ri ) (5.29)

Still using equation (5.27), the modified I-V characteristic is then given as

I (RL) = IJ - Iph

= IS { exp ( - { q I [ RL + Ri ] / kT } ) - 1 } - Iph

= IS { exp ( { q [ V - I Ri ] / kT ) } ) - 1 } - Iph (5.30)

The effect of this series resistance on the I-V characteristic for a typical silicon cell is

shown in Figure 5.11. The VOC is not affected but the ISC is, reducing the fill factor.

29
Figure 5.11. The effect of series resistance on the current voltage curve for a typical silicon solar cell [2]

The relation that describes the actual solar photovoltaic cells more accurately is given as

I = IS { exp ( { q [ V - I Ri ] / nkT ) } ) - 1 } - Iph (5.31)

where n is the junction perfection factor (n = 1 for perfect junction), accounts for space

charge generation and other effects. Its value usually is greater than unity and is nearly

constant with voltage. It also decreases with temperature and illumination intensity

(insolance).

From equation (5.27), VF is given as

VF = ( nkT / q ) ln {( Iph + I ) / IS + 1 } (5.32)

If I = 0, then we have an open-circuit voltage, given as

VOC = ( nkT / q ) ln {( Iph / IS ) + 1 } (5.33)

It is also known as the photovoltage. Its value is determined by the light density,

temperature and by semiconductor properties. Iph shows no significant change as a result

of the diffusion current of the photon-generated minority carriers not being well excited

sufficiently by the temperature. The diffusion constant D, and the lifetime τ, are rather

insensitive to temperature. With equation (5.32), the external voltage can easily be

obtained

30
- I [ RL + Ri ] = ( nkT / q ) ln {( Iph + I ) / IS + 1 }

- I = ( nkT / q ) [ RL + Ri ]-1 ln {( Iph + I ) / IS + 1 }

Since V = - IRL, then

V / RL = ( nkT / q ) [ RL + Ri ]-1 ln {( Iph + I ) / IS + 1 }

V = ( nkT / q ) [ ( RL / RL + Ri ) ] ln {( Iph + I ) / IS + 1 }

= ( nkT / q ) [ ( 1 + Ri / RL ) ]-1 ln {( Iph + I ) / IS + 1 } (5.34)

The short-circuit current, that is, when V = 0 is given as the relation, using equation

(5.31)

ISC = IS { exp ( - qIRi / nkT ) - 1 } - Iph (5.35)

It can be seen then that ISC depends on the spectral response of the photovoltaic cell and

light spectrum. This spectral response also depends on some parameters such as the

optical absorption coefficient α, junction depth between xp and xn, depletion region width

W, lifetimes and mobilities of both sides of the junction, surface recombination velocity

s, the presence and absence of electric field.

For high quality cells, Ri can be removed or made small by metallisation of the material.

It then follows that ISC ≈ Iph. While the VOC is a strong function of temperature, the ISC is

not so dependent on the operating temperature as will be seen below.

The ISC can be interrelated with VOC. From equation (5.33)

Iph = IS { exp ( qVoc / kT ) - 1 } (5.36)

Substituting equation (5.36) into equation (5.35), we have

ISC = IS { exp ( qVoc / kT ) - exp ( - qIRi / kT ) } (5.37)

If Ri is made small for a good quality cell, then

ISC = IS { exp ( qVoc / kT ) - 1 } (5.38)

31
The ISC can equally be written as

ISC = Ac T γ exp ( - Ego / kT ) exp ( qVoc / kT ) (5.39)

where Ac is the area, Eg0 is the band gap at zero temperature, T, the temperature and γ

determines the temperature dependence of other terms determining IS, which lies between

1 and 4. The temperature rate of change of VOC can be determined using equation (5.39).

From equation (5.39)

VOC = Eg0 / q + kT / q ln ( ISC / Ac T γ ) (5.40)

dVOC / dT = k / q ln ( ISC / Ac T γ ) - γ k / q

= k / q { ln ( ISC / Ac T γ ) - γ } (5.41)

From equation (5.40)

ln ( ISC / Ac T γ ) = ( - Eg0 + q VOC ) / q (5.42)

Therefore,

dVOC / dT = k / q [ { ( q VOC - Eg0 ) / kT } - γ ]

= {VOC - Vg0 - ( kT / q ) } / T (5.43)

where Vg0 = Eg0 / q. It can therefore be deduced from equation (5.43) that an increase in

temperature results in a linear decrease in VOC. This is as a result, say, in silicon solar

cell, of the electron concentration n(xp), which balances the photon-generated current in

open-circuit condition as it depends exponentially on qVOC / kT. By allowing VOC to be

proportional to kT, that is, temperature, it is the only way to keep n(xp) constant. The

figure below (Figure 5.13 (a)) shows the temperature effect on maximum current Im, for

various semiconductors while Figure 5.13 (b) shows that on maximum voltage Vm.

32
(a) (b)

Figure 5.13. The effect of temperature on (a) maximum current and (b) maximum voltage for a
semiconductor [2]

The power output from the cell and fed as input to RL is P = -IV and can be expressed in

terms of I and R or I and V.

P = - I kT / q [ ( 1 + Ri / RL ) ]-1 ln {( Iph + I ) / IS + 1 } (5.44a)

or

P = - V [ IS { exp ( { q [ V - I Ri ] / nkT ) } ) - 1 } - Iph ] (5.44b)

With equation (5.44b), a fraction of the VOC and the ISC product can be utilised. In any

practical circuit, ( Iph + I ) / IS ) >> 1. Therefore, equation (5.44a) can be written as

P ≈ - I kT / q [ ( 1 + Ri / RL ) ]-1 ln [ ( Iph + I ) / IS ) ] (5.45)

The maximum power can be obtained by differentiating equation (5.45) (the derivative is

with respect to I) and equating it to zero. Therefore,

dP / dI = - I kT / q[ ( 1 + Ri / RL ) ]-1 ln [ ( Iph + I ) / IS ] - I kT / q [ ( 1 + Ri / RL ) ]-1( Iph + I )

When dP / dI = 0, then

ln [ ( Iph + I )/ IS ] = - I / ( Iph + I ) (5.46)

33
This is a transcendental equation giving I in terms of Iph and IS. Equation (5.46) in

equation (5.45) gives the maximum available power from the solar cell as

Pmax = - Im kT / q [ ( 1 + Ri / RL ) ]-1 { - Im / ( Iph + Im ) }

= kT / q [ ( 1 + Ri / RL ) ]-1{ - Im / ( Iph + Im ) }(- Im )

Pm = Vm Im (5.47)

The efficiency of excited electrons collected determines the value of the maximum

current as it is less than ISC while the maximum voltage, which would be the voltage

across the load would probably be less than VOC / 2. This would give the power output as

nearly 32 % of the ideal value (48 %) as we shall see.

The fill factor (F.F.) can now be defined as the ratio of the maximum power to the

theoretical limit, that is, VOC and ISC product (Figure 5.9).

F.F. = Pm / VOC ISC (5.48)

which means the junction perfection factor and the series resistance are independent of

the fill factor. On practical silicon solar cells, fill factor is between 0.7 and 0.82. Typical

open-circuit voltages are around 550 to 600 mV usually decreasing by about 0.4 % for

each degree °C rise in temperature and the short-circuit currents are about 30 mAcm-2

when operated at room temperature above the earth’s atmosphere. The silicon’s cell

power output also decreases by about 0.35 to 0.45 % per Celsius of temperature increase

due to the open-circuit voltage reduction (Figure 5.15).

Figure 5.15. Temperature dependence of power [4]

34
The efficiency η, of a solar cell is defined as the percentage of power supplied (when

electrical energy is converted from optical energy) and absorbed, when a solar cell is

connected to an electrical circuit. It is therefore the ratio between the electric power

obtained from the maximum power point of the I-V curve and the incident light.

η = Pm / Pin = ImVm / EAC (5.49)

where E is the input light irradiance under ‘standard’ test (AM-1.5) conditions measured

in W/m2 and AC the solar cell surface area. E is calculated using the expression

E = ∫0 ∞ ( hc / λ ) f ( λ ) d λ (5.50)

where f ( λ ) is the number of photons per square meter per second per unit bandwidth

incident on the cell surface at wavelength λ, hc / λ is the photon energy. Substitution of

equations (5.48) and (5.50) into equation (5.49) gives efficiency as

η = ( F.F. VOC ISC ) / [ AC ∫0 ∞ ( hc / λ ) f ( λ ) d λ ] (5.51)

It can equally be expressed in terms of mean number of carriers collected and mean

number of photons in the spectrum, given as

η = Q (Im / Isc )( Vm q nph(Eg) / Nph Eav ) ( Aa / AC ) ( 1 – R ) (5.52)

where Q is the mean collection efficiency given as the ratio of number of carriers

collected to Nph (Eg), nph (Eg) is the number of photons having energy greater than Eg per

square cm per second, R is the average reflectivity, Eav the mean photon energy, Aa the

active area, AC the total device area. The table below (Table 5.1) shows the maximum

possible energy conversion in silicon at each wavelength.

35
Table 5.1 Energy conversion in silicon [3]

Wavelength interval (μm) Efficiency of conversion Proportion recovered (%)

< 0.3 - -

0.3 – 0.5 0.36 6

0.5 – 0.7 0.55 15

0.7 – 0.9 0.73 15

0.9 – 1.1 0.91 12

> 1.1 0 0

Total 48

The maximum conversion efficiency is 48 % and so high efficiency solar cells show

same characteristics and this can be gotten if the surface conditions, junction depth and

crystal perfection are near the maximum. However, in practical devices, such a figure

does not look attainable. The maximum gotten so far is 25 % and are used where their

premium cost is justified. For silicon cells at sea level, it is 20 % higher than that found

outside the atmosphere. In terrestrial applications, efficiencies (for commercial silicon

solar cells) lie between 10 and 15 %. To get optimum performance, cells are matched

based on whether they have high or low efficiencies under standard conditions. With this,

solar modules can be gotten and used in battery charging for circuit applications.

36
CHAPTER 6

ABSORPTION IN A SEMICONDUCTOR

The absorption of light in a semiconductor is a function of wavelength of light and is

possible only if the wavelength is shorter than the absorption edge of the material under

light. It is expressed as

f ( λ, x ) = f0 ( λ ) exp ( - α( λ ) x ) (6.1)

where f ( λ, x ) is the photon flux, measured in mW/cm2 or photons per square cm per

second, f0 ( λ ) is the photon flux at x = 0 and α ( λ ) is the absorption coefficient. For an

electron to move from the valence band to the conduction band, the photon provides the

minimum energy at the absorption edge. There are two kinds of intrinsic absorption

observed in semiconductors; direct absorption and indirect absorption. As a result we

have the direct band-gap and indirect band-gap semiconductor.

Direct band-gap semiconductor

Here, the crystal momentum is conserved in transition. In the diagram shown (Figure

6.1(a)), electron moves from the valence band to the conduction band when the photon

energy is the same as the band gap. This is when absorption starts.

(a) (b)

Figure 6.1. Optical absorption in a direct band semiconductor. (a) Direct optical transition from valence to
conduction band (b) Plot of absorption coefficient with photon energy [2]

37
The absorption coefficient α, near the minimum energy for absorption band gap is given

as

α = ( 3.38 x 107 / n ) ( me / mo )1/2 ( Eg / hυ ) ( hυ - Eg ) 1/2 (6.2)

where n is the index of refraction, mo is the free electron mass, me is the mass of electron.

( 3.38 x 107 / n ) ( Eg / hυ ) is for allowed transitions while with forbidden transitions, the

absorption coefficient is given as

α = f (E) ( hυ - Eg ) 1/2 (6.3)

where f (E) is a function of energy. The plot of equation (6.3) is given in Figure 6.1(b).

The momentum of the photon is negligible and they include GaAs, GaN, CdS, CdTe,

CdSe, CuInSe2, Sn (grey), InAs, InSb, InN, ZnS, ZnO, S, Te, etc.

Indirect band-gap semiconductor

Here, crystal momentum is not conserved as the maximum energy near the valence band

also plays a role in addition to the minimum energy near the conduction band as seen in

Figure 6.2 (a).

(a) (b)

Figure 6.2. Optical absorption in an indirect band semiconductor. (a) Indirect optical transition from
valence to conduction band (b) Plot of absorption with photon energy [2]

38
The process that happens in the direct band-gap semiconductor equally occurs here but in

a different manner. Here a phonon must be absorbed before transition can take place and

so

hν + Ep = Eg (6.4)

The minimum photon energy is therefore

hν = Eg - Ep (6.5)

where Ep is the absorbed phonon energy. The phonons have a high momentum but low

energy as compared with the photons. The absorption coefficient is given as

α = αa + αe (6.6)

where αa is due to phonon absorption while αe is the contribution due to phonon emission.

Both tend to zero if in the former hν < Eg - Ep and in the latter if hν < Eg + Ep. αa and αe

can be determined from these relations:

αa = A [ hν – ( Eg – Ep ) ]2 / [ exp ( Ep / kT ) – 1 ] (6.7a)

and

αe = A [ hν – ( Eg + Ep ) ]2 / [ 1– exp ( - Ep / kT ) ] (6.7b)

where A is a constant. For forbidden indirect transitions

α = f ( E, T ) [ hν – ( Eg – Ep ) ]2 (6.8)

where f ( E, T ) is a function of energy and temperature. With these, carriers gain more

energy than they have at thermal equilibrium. Compared to the direct band-gap, α is low.

Figure 6.2 (b) shows a plot of α1/2 and hν. The lower line intercepting at hν1 corresponds

to phonon absorption while the upper line intercepting at hν2 corresponds to phonon

emission. With this, the band gap and phonon energy can be written as

Eg = h ( ν1 + ν2 ) / 2; Ep = h ( ν1 - ν2 ) / 2 (6.9)

39
Examples of indirect band-gap semiconductor include Si, SiC, BaSi2, BeTe, AlAs, InP,

GaInAs, GaP, Ge, C (diamond), Se, etc. The optical absorption coefficient as a function

of photon energy is shown below (Figure 6.3).

Figure 6.3. Optical absorption coefficient as a function of photon energy for some semiconductors [2]

If the indirect band-gap material is used for solar cells, a very thick material is needed to

absorb most of the solar spectrum, as the absorption coefficient is inversely proportional

to the penetration depth of light. A thin direct gap material can equally be used though

they suffer from higher surface recombination losses.

In general, not every photon is able to produce an electron. The fraction of effective

photons is known as the quantum efficiency ηq, that is, it is an electron-hole pair that must

be produced from an incident photon. It is can be used in calculating the current, which is

given as

I = q ηq A Φbackground (6.10)

where Φbackground is the photon flux density, given in photons per second per square

centimetre, A is the junction area, and q is the electric charge respectively. The

expression A Φbackground is the light intensity H. So equation 6.10 can be rewritten as

40
I = q ηq H (6.11)

For a p-n junction, the charge carriers created within the depletion region are directly

exposed to the strong field and this allows the carriers not to follow the mobility relation

μ = v / E. The strong field is given as VD / W where VD is the diffusion voltage and W is

the barrier width. The reason is that, in a strong field, interaction takes place between

electrons and optical phonons of the semiconductor lattice. Because the field current in

the junction is constant, it does not vary with voltage. This is the reason photon-generated

current remains constant irrespective of the applied voltage value to the solar cell

provided it is lower than VD. The photon-generated current is restricted to one side of the

junction for a direct band-gap semiconductor.

41
CHAPTER 7

COMPARATIVE ANALYSIS OF SOLAR CELLS

The saturation current IS, can be written as a function of temperature. It is given as

IS = aT3 [ exp (– qEg / kT ) ] (7.1)

where a is obtained from semiconductor properties using a = q / kT and Eg is the

effective thermal energy gap. From equation (7.1), it is seen that maximum theoretical

efficiency of a solar cell is related with the semiconductor material energy gap. It is also a

function of spectral distribution of incident light and the temperature. If a p-n junction

has both the conduction band in the p-layer and the n-layer coinciding with the Fermi

level, then energy gap is equivalent to the maximum open-circuit voltage. This means an

increase in the band gap will see to a higher open-circuit voltage but will lead to a drop in

the short-circuit current and vice versa. This is because in materials with low band gaps,

their photon absorption edge is located at longer wavelengths and this shortcoming has to

be taken into account. The optimum band gap for solar cell of peak efficiency is between

1.4 and 1.6 eV. Figure 7.1 shows the maximum theoretical efficiency of semiconductors

as a function of their gap at room temperature.

Figure 7.1. Maximum theoretical conversion efficiency at room temperature for semiconductors, as a
function of band gaps [4]

42
A is used as a result of the recombination current in the junction. The only snag with the

diagram above is that certain problems are not accounted for which do not allow for

maximum efficiency. The next diagram below (Figure 7.2) shows the variation of the

maximum theoretical efficiency calculated using equation (7.1) for different

temperatures. But the efficiency losses are also ignored. For a uniformly high steady-state

temperature, high concentration devices can be used with them. The decrease in

temperature, which is as a result of the Fermi level in the middle of the band gap leads to

a sharp efficiency decrease.

Figure 7.2. Temperature dependence of maximum efficiency [4]

For silicon solar cell under monochromatic visible light, the theoretical maximum

efficiency is about 45 – 50%. For AM-0, it is about 19%. On ground about 23 %.

Analysing the total system including the material used, its thickness and fabrication

techniques can lead to the reduction in the costs of solar cells by employing thin-film or

light absorbing technology and so on.

43
REFERENCES

1. Bar-Lev A., (1997) Semiconductors and Electronic Devices, Prentice-Hall

International, London, 149-152.

2. Garg H. P., (1987) Heating, Agricultural and Photovoltaic Applications of Solar

Energy, Advances in Solar Energy Technology, Vol. 3, D. Reidel Publishing

Company, Holland, 279-285, 285-315, 325-329.

3. McMullan J. T. et al, (1979) Energy Resources and Supply, John Wiley and Sons,

New Jersey, 336-340.

4. Palz W., (1978) Solar Electricity: An Economic Approach To Solar Energy,

Butterworths and UNESCO, France, 179-189, 192, 217-223, 252-275.

5. Russell C. R., (1967) Elements Of Energy Conversion, 1 st ed., Pergamon Press, New

York, 245-254.

44

Вам также может понравиться