Sei sulla pagina 1di 45

A

SEMINAR REPORT

ON

SOLAR PHOTOVOLTAICS

Submitted By:
Darshan Khunt
12BCH012

Guided By:
Prof. Leena Bora
Assistant Professor

CHEMICAL ENGINEERING DEPARTMENT


INSTITUTE OF TECHNOLOGY
NIRMA UNIVERSITY

CERTIFICATE
This is to certify that Mr. Darshan Khunt (12BCH012), student of Chemical
Engineering, V semester, of Nirma University, has satisfactorily completed the
seminar/project on SOLAR PHOTOVOLTAICS as a partial fulfillment towards the
degree of B. Tech. in Chemical Engineering.

Date:
Place:

Name of Guide
Designation of Guide

Prof. Leena Bora


Assistant Professor

CONTENT
Page No.
Acknowledgement

Abstract

II

List of Tables

III

List of Figures

III

Chapter Title

Page No.

No.
1.

2.

Introduction
1.1

Solar Energy Reaching the Earth

1.2

History of Photovoltaics

The Photovoltaic Effect

6-9

2.1

Effect of light on Silicon

2.2

The Potential Barrier

2.2.1 The Negative- Carrier Dopant

2.2.2 The Positive- Carrier Dopant

The Electric Current

2.3
3.

1-5

Solar Photovoltaics
3.1

11-35

Thin film solar Technology

11

3.1.1 Introduction

11

3.1.2 Thin film fabrication Methods

12

3.1.3 Amorphous Silicon Solar Cells

15

3.1.4 Cadmium Telluride Solar Photovoltaics

20

3.1.5 Copper Indium Gallium Selenide Solar Cell

22

3.2

Dye Sensitized Solar cells

24

3.3

Multi-junction Solar cell

28

3.4

Organic Solar cell

31

4.

Physical Aspects Of Solar Cell Efficiency

35

5.

Summary

36

References

37-38

ACKNOWLEDGEMENT

I would like to express my gratitude towards Head of Chemical Engineering Department,


Dr.Sanjay Patel for including seminar as part of curriculum in process of becoming engineer.
I would like thank the Library of our Institute and member of library for helping me out with
my topic. The resource of the library has provide me in depth knowledge of the seminar.

I have taken efforts in this Seminar. However, it would not have been possible without the kind
support and help of Prof. Leena Bora as my supporting guide. I am highly indebted to my guide
for their guidance and constant supervision as well as for providing necessary information
regarding the seminar & also for their support in completing the Seminar.

ABSTRACT

The word photovoltaic is a combination of two words - photo, meaning light, and voltaic,
meaning electricity. Therefore photovoltaic technology, the scientific term used to describe
solar energy, involves the generation of electricity from light. The secret to this process is the
use of a semi-conductor material which can be adapted to release electrons, the negatively
charged particles which form the basis of electricity. The most commonly used semi-conductor
material used in photovoltaic (PV) cells is silicon, an element most commonly found in sand.
The spectrum of solar energy is quite wide and its intensity varies according to the timing of
the day and geographic location. We review solar energy conversion into electricity with
particular emphasis on photovoltaic systems, solar cells and how to store electricity. Solar Cell
is differentiated on the basis of the material used form the solar cell construction like Cadmium
telluride PVs, Thin Film PVs, Plastic PVs etc. . Thin film Solar cell technology and recent
development of organic solar cell has been briefly described to have a close look on research
on solar PVs. Solar cells have many applications such as in remote area power systems, Earth
orbiting satellites, consumer systems, handheld calculators or wrist watches, remote radiotelephones and water pumping applications. Physical aspects of solar energy generation is
concisely discussed. Solar cells are regarded as one of the key technologies towards a
sustainable energy supply.
Keywords: Photovoltaic, silicon, solar cell, sustainable energy, Multijunction solar PVs,
Cadmium telluride, Copper Indium Gallium Selenide, Dye Sensitized, Organic Solar cell.

II

List of Tables
Table No.

Description

Page
No.

Table 1

Stabilized efficiencies of a-Si-based PV modules

18

List of Figures
Figure No.

Description

Page
No.

Figure 1

The earths atmosphere and clouds affect the way in which the

suns light reaches the surface of earth


b

Light from the sun at the outer fringes of the earth of the earths

atmosphere (AMO) Covers a broad range of wavelengths.


Figure 2-1

Representation of the silicon crystal lattice arrangement

Figure 2-2

Light of sufficient energy can generate electron-hole pairs in silicon,

both of which move for a time freely throughout


Figure 2-3

A potential barrier in a solar cell separates light-generated charge

carriers, creating a voltage.


Figure 2-4

When an impurity atom such as phosphorous with five valence

electrons is substituted into a silicon crystal, it has an extra,


unbounded electron
Figure 2-5

A three-valence-electron impurity in a silicon crystal is normally

bonded except one of the bonds is missing an electron, i.e., is a hole


Figure 2-6

Light incident on the cell creates electron-hole pairs, which are

10

separated by the potential barrier, creating a voltage that drives a


current through an external circuit.
Figure 3-1

Honeywell silicon-on-ceramic dip coating process

12

Figure 3-2

A schematic of an a-Si/a-SiGe tandem device structure fabricated on

17

a glass substrate.

III

Figure 3-3

A schematic of a triple-junction device structure fabricated on a

17

stainless steel substrate.


Figure 3-4

Film sequence of the CdTe thin-film solar cell.

21

Figure 3-5

Composition a conventional CuInGaSe2

22

Figure 3-6

Sphalerite or zinc blende structure of ZnSe (two unit cells);

23

Chalcopyrite structure of CuInSe2. The metal sites in the two

23

unit ceils of the sphalerite structure of ZnSe are alternately


occupied by Cu and In in the chalcopyrite structure.
Figure 3-7

Dye sensitized Solar cell

24

Figure 3-8

Certified I-V characteristic under simulated sunlight, of a dye-

25

sensitised photoelectrochemical cell, showing a conversion


efficiency > 10%.
Figure 3-9

Structure of a Triple-Junction PV Cell

29

Figure 3-10

Fundamental mechanism of the photon-to-electron conversion

32

process in bulk heterojunction solar cells.


Figure 3-11

Schematic diagrams of a conventional pn junction solar cell (left)


and an organic heterojunction solar cell (right). The diagram
highlights differences in carrier generation between the two types of
devices.

IV

33

Chapter 1. Introduction
Over the next few decades, it is possible that the demand for carbon-free electric power
generation will dramatically increase the use of intermittent renewable sources such as solar
photovoltaics (PV). In our previous analysis (Denholm and Margolis, 2007), we examined the
inherent limits of traditional electric power systems to accept very large amounts of PV energy.
A large fraction of PV electricity generation occurs when normal electricity demand is
relatively low, and the existence of large inflexible thermal steam plants results in unusable
PV, resulting in increased costs. At some point when PV is supplying in the range of 1020%
of a systems energy, the cost penalties and diminishing return of increasing PV generation
will likely limit the economic use of this generation technology. In this work, we examine
several options to increase the penetration of solar PV beyond 20% of a systems energy. We
begin by reviewing the cost impacts of PV at high penetration in conventional electric
power systems. We then discuss qualitatively, and analyse quantitatively three approaches that
could increase the usefulness of PV generation. The first is increasing the systems flexibility
by increasing the ramping capability and decreasing the minimum load on conventional
generators. The second is increasing the effective coincidence of PV supply and electricity
demand by the use of load shifting [1]. And the third is energy storage which provides the
ultimate solution by allowing excess PV generation to be stored and delivered at a later
time. This analysis includes results from simulations of each of these alternatives in an attempt
to quantify approaches to increase PV penetration in the electric power system. Photovoltaic
systems behave in an extraordinary and useful way: They react to light by transforming part of
it into electricity.

Moreover this conversion is novel and unique, since photovoltaics:


Have no moving parts (in the classical mechanical sense) to wear out
Contain no fluids or gases (except in hybrid systems) that can leak out, as do some solar
thermal systems
Consume no fuel to operate
Have a rapid response, achieving full output instantly
Can operate at moderate temperatures

Produce no pollution while producing electricity (although waste products from their
manufacture, and toxic gases in the event of catastrophic failure and disposal may be a
concern)
Require little maintenance if properly manufactured and installed
Can be made from silicon, the second most abundant element in the earth's crust are
modular permitting a wide range of solar-electric applications such as

Small scale for remote applications and residential use

Intermediate scale for business and neighbourhood supplementary power

Large scale for centralized energy farms of square kilometers size.

Have a relatively high conversion efficiency giving the highest overall conversion
efficiency from sunlight to electricity yet measured
Have wide power-handling capabilities, from microwatts to megawatts
Have a high power-to-weight ratio making them suitable for roof application
Are amenable to on-site installation, i.e., decentralized or dispersed power clearly,
photovoltaics have an appealing range of characteristics. However, there are ambivalent
views about solar, or photovoltaic, cells' ability to supply a significant amount of energy
relative to global needs.
Those pro, contend: Solar energy is abundant, inexhaustible, clean, and cheap.
Those can, claim: Solar energy is tenuous, undependable, and expensive beyond
practicality.
There is some truth to both of these views. The sun's energy, for all practical purposes, is
certainly inexhaustible. However, even though sunlight striking the earth is abundant, it comes
in rather a dilute form [1].

1.1 Solar Energy Reaching the Earth

Fig. 1 (a) the earths atmosphere and clouds affect the way in which the suns light reaches
the surface of earth
(b) Light from the sun at the outer fringes of the earth of the earths atmosphere
(AMO) covers a broad range of wavelengths.
Even though the sun ranks as a run-of-the-mill star, it releases a huge quantity of energy in
terms of human capacity or need. Power output per second is 3.86 x 1020 megawatts (MW),
several billion times the electric capacity of U.S. utilities. This energy fills the solar system,
bathing the earth's atmosphere with a near constant supply of 1.37 kilowatts per square meter
(kW/m2). Not all of the direct sunlight incident on earth's atmosphere arrives at the earth's
surface (Figure 1(a)). The atmosphere attenuates many parts of the spectrum (Figure 1(b)). For
example, X-rays are almost totally absorbed before reaching the ground. A good percentage of
ultraviolet radiation is also filtered out by the atmosphere. Some radiation is reflected back into
space. Some is randomly scattered by the atmosphere, which makes the sky look blue. It is
valuable to relate the amount of sunlight at the earth's surface to the quantity, or air mass (AM),
of atmosphere through which the light must pass. Radiation arriving at the surface of the earth
is measured against that reaching the fringes of the atmosphere, where there is no air, and the
air mass is zero (AMO). The light of the high-noon sun (and under further specified conditions)
passes through an air mass of one (AM1). The intensity of the sunlight reaching the ground
weakens for sun angles approaching the horizon since the rays have more atmosphere, or air
mass, to penetrate. The atmosphere is a powerful absorber and can cut the sun's energy reaching
3

the earth by 50% and more. The peak intensity of sunlight at the surface of the earth is about 1
kW/m2. However, not all areas of the earth get the same average amounts of sunshine
throughout the year. The most intensely bathed areas lie between 300 north and 300 south
latitude, since these areas have the least cloud cover. There also are, of course, seasonal
radiation variations caused by the tilt of the earth with respect to the sun. Thus, the winter sun
will daily provide less than 20% of the summer sun's energy at some locations because it is
lower in the sky and the days are shorter. All of these factors affecting the amount of local
radiation on earth have to be taken into account when designing photovoltaic systems. The sun
may be a constant source of energy, but at the earth's surface, the distribution of its energy and
the constancy of its radiation are hardly ideal. A good PV system cannot be designed without
providing for the variations associated with the energy spectrum and its local availability [1].

1.2 History of Photovoltaics


The physical phenomenon responsible for converting light to electricity-the photovoltaic
effect-was first observed in 1839 by a French physicist, Edmund Becquerel. Becquerel noted a
voltage appeared when one of two identical electrodes in a weak conducting solution was
illuminated. The PV effect was first studied in solids, such as selenium, in the 1870s. In the
1880s, selenium photovoltaic cells were built that exhibited 1%-2% efficiency in converting
light to electricity. Selenium converts light in the visible part of the sun's spectrum; for this
reason, it was quickly adopted by the then-emerging field of photography for photometric
(light-measuring) devices. Even today, light-sensitive cells on cameras for adjusting shutter
speed to match illumination are made of selenium. Selenium cells have never become practical
as energy converters because their cost is too high relative to the tiny amount of power they
produce (at 1% efficiency). Meanwhile, work on the physics of PV phenomena has expanded.
In the 1920s and 1930s, quantum mechanics laid the theoretical foundation for our present
understanding of PV. A major step forward in solar-cell technology came in the 1940s and
early 1950s when a method (called the Czochralski method) was developed for producing
highly pure crystalline silicon. In 1954, work at Bell Telephone Laboratories resulted in a
silicon photovoltaic cell with a 4% efficiency. Bell Labs soon bettered this to a 6% and then
11% efficiency, heralding an entirely new era of power-producing cells [2].

A few schemes were tried in the 1950s to use silicon PV cells commercially. Most were for
cells in regions geographically isolated from electric utility lines. But an unexpected boom in

PV technology came from a different quarter. Transistors and PV cells are made from similar
materials, and their workings ate determined by many of the same physical mechanisms. Solar
cells are usually divided into three main categories called generations. The First Generation
contains solar cells that are relatively expensive to produce, and have a low efficiency. The
Second Generation contains types of solar cells that have an even lower efficiency, but are
much cheaper to produce, such that the cost per watt is lower than in first generation cells. The
term Third Generation is used about cells that are very efficient. Most technologies in this
generation is not yet commercial, but there is a lot of research going on in this area [1]. The
goal is to make third generation solar cells cheap to produce. An enormous amount of research
and development has been expended in improving the third generation solar sell.

Today, photovoltaic systems are capable of transforming one kilowatt of solar energy falling
on one square meter into about a hundred watts' of electricity. One hundred watts can power
most household appliances: a television, a stereo, an electric typewriter, or a lamp. In fact,
standard solar cells covering the sun-facing roof space of a typical home can provide about
8500-kilowatthours of electricity annually, which is about the average household's yearly
electric consumption. By comparison, a modern, 200-ton electric-arc steel furnace, demanding
50,000 kilowatts of electricity, would require about a square kilometre of land for a PV power
supply [2][3].

Certain factors make capturing solar energy difficult. Besides the sun's low illuminating power
per square meter, sunlight is intermittent, affected by time of day, 8 Basic Photovoltaic
Principles and Methods climate, pollution, and season. Power sources based on photovoltaics
require either back-up from other sources or storage for times when the sun is obscured.

In addition, the cost of a photovoltaic system is far from negligible (electricity from PV systems
in 1980 cost about 20 times * that from conventional fossil fuel- powered systems).

Chapter 2: The Photovoltaic Effect


2.1 The Effect of Light on Silicon
The silicon atom has fourteen electrons arranged in such a way that the outer four can be given
to, accepted from, or shared with another atom. These four outer electrons are called valence
electrons. Large numbers of silicon atoms, through their valence electrons, can bond together
to form a solid. As a solid, each silicon atom usually shares each of its four valence electrons
with another silicon atom. Each basic silicon unit, forming a tetrahedral arrangement, thereby
contains five atoms. When light strikes a silicon crystal, it may be reflected, be absorbed, or
may go right through. Let's concentrate on the light that is absorbed. Usually when light of
relatively low energy is absorbed by a solid, it creates heat without altering the electrical
properties of the material. That is, low-energy light striking a silicon crystal causes atoms of
silicon to vibrate and twist in their bound positions, but do not break loose. Similarly, electrons
in bonds also gain more energy and are said to attain a higher energy level. Since these energy
levels are not stable, the electrons soon return to their original lower energy levels, giving off
as heat the energy they had gained [5].

Fig. 2-1 Representation of the silicon


crystal lattice arrangement

Fig. 2-2. Lights of sufficient energy can


generate electron-hole pairs in silicon, both
of which move for a time freely throughout
the crystal.

Light of greater energy can alter the electrical properties of the crystal. If such light strikes a
bound electron, the electron is torn from its place in the crystal [4]. This leaves behind a silicon
bond missing an electron and frees an electron to move about in the crystal. A bond missing an
electron, rather picturesquely, is called a hole. An electron free to move throughout the crystal
is said to be in the crystal's conduction band (Figure 2-2), because free electrons are the means
by which electricity flows. Both the conduction-band electrons and the holes play important
parts in the electrical behaviour of PV cells. Electrons and holes freed from their positions in
the crystal in this manner are said to be light-generated electron-hole pairs.
A hole in a silicon crystal can, like a free electron, move about the crystal. The means by which
the hole moves is as follows: An electron from a bond near a hole can easily jump into the hole,
leaving behind an incomplete bond, i.e., a new hole. This happens fast and frequently-electrons
from nearby bonds trade positions with holes, sending holes randomly and erratically
throughout the solid [6]. The higher the temperature of the material, the more agitated the
electrons and holes and the more they move. The generation of electrons and holes by light is
the central process in the overall PV effect.

2.2 The Potential Barrier

Fig. 2-3 A potential barrier in a solar cell separates light-generated charge carriers, creating
a voltage [2] [5].
A PV cell contains a barrier that is set up by opposite electric charges facing one another on
either side of a dividing line. This potential barrier selectively separates light-generated
electrons and holes, sending more electrons to one side of the cell, and more holes to the other.
Thus separated, the electrons and holes are less likely to rejoin each other and lose their
7

electrical energy. This charge separation sets up a voltage difference between either end of the
cell (Figure 2-3), which can be used to drive an electric current in an external circuit.

2.2.1 The Negative-Carrier (Donor) Dopant

Fig. 2-4. When an impurity atom such as phosphorous with five valence electrons
is substituted into a silicon crystal, it has an extra, unbounded electrons.

Silicon crystal with numerous substituted phosphorus atoms would have many free, conduction
band electrons and a similar number of positive impurity ions locked into the crystal's structure
[3]. Overall, the whole crystal would remain neutral, since there are just as many positive ions
as free electrons; but the crystal's electrical properties would have been drastically altered.
Impurities introduced in this way are called dopants, and dopants that have one extra valence
electron (such as phosphorus introduced into a. silicon crystal) are called donors because they
donate an electron to the crystal. Such a donor-doped crystal is known as n-type because it has
free negative charges.

2.2.2 The Positive-Carrier (Acceptor) Dopant

Fig. 2-5 A three-valence-electron impurity in a silicon crystal is normally bonded except


one of the bonds is missing an electron, i.e., is a hole.
An appropriately altered material can be formed by substituting into the silicon crystal,
impurity atoms with one fewer valence electron than silicon. An impurity atom with three
valence electrons (such as boron) would sit in the position of the original silicon atom, but one
of its bonds with the silicon would be missing an electron, i.e., there would be a hole (Figure
2-5). As we saw before, holes can move about almost as freely as conduction band electrons.
In this way, a silicon crystal doped with many such boron atoms has many holes that act as if
they were free positive charges moving throughout the crystal lattice. A three-valence-electron
impurity in a silicon crystal is called an acceptor because its holes accept electrons from the
rest of the silicon crystal. An acceptor-doped silicon material is called p-type because of the
presence of free positive [5].

2.3 The Electric Current


If we connect the n-type side to the p-type side of the cell by means of an external electric
circuit, current flows through the circuit (which responds just as if powered by a battery}
because this reduces the light induced charge imbalance in the cell. Negative charges flow out
of the electrode on the n-type side, through a load (such as a light bulb}, and perform useful
work on that load (such as heating the light bulb's filament to incandescence}. The electrons
then flow into the p-type side, where they recombine wit holes near the electrode (Figure 2-8).
9

The light energy originally absorbed by the electrons is used up while the electrons power the
external circuit. Thus, an equilibrium is maintained: The incident light continually creates more
electron-hole pairs and, thereby, more charge imbalance; the charge imbalance is relieved by
the current, which gives up energy in performing work. The amount of light incident on the
cell creates a near proportional amount of current [1][6]. The amount of energy it takes to raise
an electron to the conduction band is the amount of energy the light originally imparts to the
electron and is, thus, the maximum that can be retrieved from the electron in the external circuit.
We have observed all the conditions necessary for current to flow: incident light to free the
charge carriers, a barrier to accelerate the carriers across the junction and keep them at opposite
ends of the cell, and a charge imbalance to drive a current (charged carriers) through a circuit.

Fig 2-6. Light incident on the cell creates electron-hole pairs, which are separated by the
potential barrier, creating a voltage that drives a current through an external circuit.

10

Chapter 3. Solar photovoltaics


3.1

Thin Film Solar Cell Technology

3.1.1 Introduction
Thin silicon solar cells are an important class of photovoltaics that are currently the subject
of intense research, development, and commercialization efforts. The potential cost
reductions realized by manufacturing solar cells in a thin device configuration are highly
compelling and have long been appreciated. However, most work on thin film approaches
to solar cells has centered on materials other than crystalline silicon because it was believed
that the optical properties of crystalline silicon would limit its usefulness as a thin-film
solar cell, and for various reasons, crystalline silicon did not readily lend itself to the
common thin film deposition technologies. Several early experimental efforts on making
thin film crystalline silicon solar cells seemed to confirm at least some of these perceived
difficulties. Still, there were proponents of thin crystalline silicon solar cells including
Redfield, Spitzer et al and Barnett, who advocated their potential advantages and articulated
design principles of light trapping and high open- circuit voltage needed to achieve high
efficiencies. Starting in the late 1980s and gaining momentum in the 1990s, thin-film
crystalline silicon solar cells emerged (or perhaps more aptlyre-emerged') as a promising
approach. This was partly due to several device design and materials processing
innovations proposed to overcome the difficulties and limitations in using crystalline
silicon as a thin film solar cell material, and also partly due to the stubborn lack of progress
in many competing photovoltaic technologies. Progress on thin crystalline silicon solar
cells has now reached a level where they are positioned to capture a significant share of the
photovoltaics market, and a thin silicon solar cell may well become the successor to the
conventional (thick) silicon wafer solar cells that are presently the mainstay of the
photovoltaics industry. This chapter reviews the issues and technical achievements that
motivate the current interest in thin silicon solar cells, and surveys developments and
technology options [7].
Thin silicon solar cells is actually an umbrella term describing a wide variety of silicon
photovoltaic device structures utilising various forms of silicon (monocrystalline,
multicrystalline, polycrystalline, microcrystalline, amorphous, and porous), and made with
an almost incredibly diverse selection of deposition or crystal growth processes and
11

fabrication techniques. Thin silicon solar cells are distinguished from traditional silicon
solar cells that are comprised of ~0.3 mm thick wafers or sheets of silicon. The common
defining feature of a thin silicon solar cell is a relatively thin (< O. 1 mm) 'active' layer or
film of silicon formed on, or attached to, a passive supporting substrate. Nevertheless, even
this very general description may not subsume all the different variations of 'thin silicon
solar cell' designs currently under investigation. One purpose of a review such as this is to
provide some perspective and objective criteria with which to assess the merits and
prospects of different approaches. However, such comparisons must be tempered with the
realization that the technology is still in a state of flux and relatively early development,
and there are many disparate solar cell applications, each with a different emphasis on cost
and performance [19].

3.1.2 Thin Film Fabrication Methods


Melt Growth Techniques
Melt growth processes characteristically have both high growth rates and good material quality.
An example melt coating process is shown in Figure 3-1. In these examples, a substrate is
contacted with molten silicon, which wets the substrate and then solidifies as a silicon layer. In
many cases, the substrate is drawn through a bath of molten silicon. Generally, such processes
cannot produce layers much less than 100 microns in thickness [15].

Fig. 3-1. Honeywell silicon-on-ceramic dip coating process

12

Recrystallization of Silicon
Related to melt growth are various recrystallisation techniques. These are generally not
deposition processes per Se, but instead are used to melt already deposited silicon layers and
recrystallise them in order to achieve a more favorable grain structure. In this case, the grainstructure of the As-deposited silicon layer is not critical, and the deposition process can be
optimized for high-growth rates, large-areas and purity specifications. For instance, a plasmaenhanced CVD process can be used to plate a silicon layer of desired thickness on a suitable
substrate, such as a ceramic, which is compatible with a recrystallisation step. A zone-melting
recrystallisation (ZMR) process is effected by localised heating to create a melted zone that
moves or scans across the deposited silicon layer, melting the silicon at the leading edge and
resolidifying a silicon layer at the trailing edge. Such ZMR techniques can yield millimeter to
centimeter-sized grain structures. There are several ways to induce localised or zone melting
of layers including moving point- or line-focused infrared lamps, travelling resistively heated
strip heaters, and laser and electron beams[1][15].

High-Temperature Silicon Chemical Vapour Deposition


Chemical vapour deposition or CVD is defined as the formation of a solid film on a substrate
by reacting vapour-phase chemicals, or 'precursors', which contain the desired constituents. For
example, substrates can be coated with silicon layers by decomposition of gaseous silane (SiH4)
or trichlorosilane (SiHCl3). In fact, many precursors are possible for silicon CVD, and silane
or the chlorosilanes are probably the most commonly used, although for example, iodine and
bromine compounds are also sometimes considered as silicon precursors. In general, Silicon
CVD is a well-developed technology commonly used in integrated circuit fabrication, in which
case it is often used for epitaxial growth of silicon layers on monocrystalline silicon substrates.
For solar cell applications where CVD is used to deposit a 10-50 micron thick silicon layer for
subsequent, post-deposition ZMR, the CVD is optimised for high precursor utilisation (i.e., the
fraction of precursor converted to silicon), deposition efficiency (i.e., the fraction of deposited
silicon that ends up on the substrate rather than the walls of the reactor chamber or the substrate
susceptor), the deposition rate, the purity of the deposited silicon, areal uniformity, the potential
to recover unreacted precursors or reaction product, and various safety and environmental
issues[8][10].

13

For solar cell applications, three types of CVD are most used:

Atmospheric pressure chemical vapour deposition (APCVD)

Rapid thermal CVD(RTCVD)

Low-pressure CVD (LPCVD)

Low-Temperature Chemical Vapour Deposition


Some low-temperature chemical vapour techniques can be distinguished from the hightemperature (> 1000~ CVD processes discussed above. These methods are employed with
substrates such as glass which are not compatible with either a high-temperature deposition
step or a post-deposition recrystallization step. The relatively slow growth rates inherent in a
low temperature deposition process necessitates thin device structures on the order of several
microns thickness or less. The As-deposited silicon layers have average grain sizes of 1 micron
or less and are characterized as microcrystalline. Hydrogenated microcrystalline silicon solar
cells using a p-i-n structure which is similar to the amorphous silicon solar cell structure, can
achieve very respectable conversion efficiencies in excess of 10%. The benefits of such
microcrystalline silicon solar cells over amorphous silicon solar cells are a greater stability to
light-induced degradation processes. The deposition processes for microcrystalline silicon
solar cells are typically adaptations of those used for amorphous silicon solar cells.

Liquid-phase epitaxy (LPE).

Liquid-phase epitaxy is a metallic solution growth

technique that can be used to grow semiconductor layers on substrates. Silicon can be
precipitated from solutions of a number of molten metals in the temperature range 600-1200~
this method has been used to grow thin silicon solar cells on low-cost metallurgical grade (MG)
silicon substrates.

Other Operation of Fabrication

Physical vapor deposition (PVD)

Close-space sublimation (CSS)

Sputter deposition

Electrodeposition

Metal organic chemical vapor deposition (MOCVD)

Spray deposition

Screen-print deposition[10]

14

3.1.3 Amorphous Silicon Solar Cells


Introduction
Significant progress has been made over the last two decades in improving the performance of
amorphous silicon (a-Si) based solar cells and in ramping up the commercial production of aSi photovoltaic (PV) modules, which is currently more than 40 peak megawatts (MWp) per
year. The progress in a-Si solar cell technology can be attributed to concurrent advances in the
areas of new and improved materials, novel cell designs and in the development of large-area
deposition techniques suitable for mass production.

However, there are serious constraints imposed on the reduction in the thickness of the
junctions since this leads to a decrease in the absorption of sunlight (and a corresponding
decrease in short circuit currents) and also to an increase in shorts and shunts. The difficulty
with low light absorption in thin cells was greatly reduced with the development of efficient
optical enhancement obtained by introducing textured rather than smooth optical. Such optical
enhancement, which was first successfully applied to a-Si based solar cells, is now extensively
used in all types of thin film solar cells including thin film crystalline solar cells [16].

Physics of Operation
The operation of all solar cells is based on common physical principles. However, since
efficient a-Si based solar cells rely on material properties distinctly different from those of
crystalline silicon, the basic cell structures are somewhat different. In order to take advantage
of the excellent properties of the intrinsic (undoped) a-i:H and a-SiGe:H materials, p-i-n and
n-i-p heterojunction cell structures are used rather than the classic n/p junction structures in
crystalline silicon.

The p- and n-layers provide the built in potential of the junction in the device, however due to
the short lifetime in the highly defective doped materials the photogenerated carriers in the
doped layers are not collected and do not contribute to the cell photocurrents. The fabrication
of a p-i-n cell begins with the deposition of a p-type 'window layer' on the transparent
conductive oxide (TCO). An a-Si:H intrinsic layer (i-layer) is deposited to form the bulk
absorber region of the cell. The final step in forming the single-junction p-i-n cell is the n-layer
deposition [12].
15

Important considerations for the choice of TCO materials are their optical transmission,
conductivities and ability to form a good contact to the p-layers. The ideal TCO should have a
low sheet resistance, high optical transparency in the wavelength range 400 to 1000 nm, and
result in a small or ideally no potential barrier at the p/TCO interface. The band bending at the
interface depends on the front contact material, the p-layer bandgap, doping and densities of
states, as well as its thickness. In order to minimise optical absorption the p-type window layers
used in high efficiency cells are thin (~10 nm), which increases the likelihood that they are
fully depleted. In order to maximize the cell efficiency it is necessary to achieve a high Voc
using a very thin p-layer. If the p-layer is too thick the device performance will be adversely
affected by the loss in the photocurrent due to the higher absorption in the thick p-layer.

In a-Si solar cells, light that is absorbed in the i-layer will create electrons and holes, and the
collection of these photogenerated carriers is assisted by the internal electric field. Due to the
short carrier lifetimes associated with the localised gap states, the photogenerated carriers in aSi based cells must be collected primarily as a drift current, not as a diffusion current as is the
case in crystalline silicon solar cells. These gap states have important ramifications on the cell
performance since a large density of photogenerated carriers can become trapped in these
states. The native and light-induced defects in a-Si p-i-n devices adversely affect the carrier
collection in two ways- they act as recombination centers and also shield the electric field
produced by the doped layers which changes the electric field distribution in the i-layer.

Construction
Generally, a p-i-n junction configuration is used with glass substrates so that the light is
incident on the glass and passes first through the p-layer side of the cell (this is sometimes
referred to a glass superstrate structure). An example of a glass superstrate cell structure is
shown in figure.
The first junction is formed by depositing a thin p-layer (~10 nm of a boron doped a-SiC:H
alloy) on the tin oxide, followed by H160 nm of an a-Si:H i-layer and then ~10 nm of
phosphorus-doped microcrystalline Si:H.

16

Fig. 3.2 A schematic of an a-Si/a-SiGe tandem device structure fabricated on a glass


substrate [15].
A tunnel or recombination junction is then formed by depositing ~ 10 nm of another a-SiC:H
p-layer. The second i-layer consists of ~ 100 nm of an a-SiGe:H alloy where the Ge content is
varied so that the band gap of the junction is graded. The second junction is completed by
depositing ~20 nm of a phosphorus-doped a-Si:H layer. The back contact is made by first
depositing about 100 nm of zinc oxide by low-pressure chemical vapour deposition (CVD) and
then sputter depositing about 300 nm of aluminum. The PV modules are then encapsulated
using EVA and another sheet of heat-strengthened glass.

Fig. 3.3 A schematic of a triple-junction device structure fabricated on a stainless steel


substrate.
17

This type of device is constructed using an n-i-p configuration where the first a-Si layer
deposited on the foil is an n-layer, and the triple-junction device has the configuration: stainless
steel foil/textured silver/zinc oxide/n-i3-p/n-i2-p/n-il-p/ITO/EVA/fluoropolymer where both i2
and i3 are a-SiGe:H alloys, il is an a-Si:H alloy, ITO is indium-tin-oxide and the fluoropolymer.

Performance
Present-day commercial a-Si based PV modules typically exhibit stabilised conversion
efficiencies in the range of 6-8% while those based on single-crystal or polycrystalline silicon
generally exhibit efficiencies in the range of 11-14%. The best stabilised efficiencies of a-Si
based PV modules reported by a number of companies are listed in Table 1.

Table 1. Stabilized efficiencies of a-Si-based PV modules


Company

Stabilized

Device configuration

aefficiency(%)/(aperture
area)
BP Solar

8.1%/(0.36 m2)

a-Si/a-SiGe tandem on glass

BP Solar

7.6%/(0.74 m2)

a-Si/a-SiGe tandem on glass

Fuji Electric

9.0%/(0.32 m2)

a-Si/a-SiGe tandem on plastic

Intersolar

~4.5-5.0%/(0.3 m2)

Single junction on glass

Lowa Thin Films

~4.5-5.5%/(0.45 m2)

Same gap tandem on plastic

Kaneka

8.1%/(0.41 m2)

Single junction on glass

Kaneka

~10%/(0.37 m2)

a-Si/c-Si tandem on glass

Phototronics

~6.0-6.5%/(0.55 m2)

Same gap tandem on glass

Sanyo

9.3%/(0.51 m2)

a-Si/a-SiGe tandem on glass

United Solar

10.1%/(0.09 m2)

Triple Junction on steel foil

United Solar

7.9%/(0.45 m2)

Triple Junction on steel foil

18

Future Trends
The key drivers behind the terrestrial PV market are module efficiency, selling price and
reliability. Crystalline silicon PV modules have exhibited good long-term reliability with many
arrays still in operation after more than two decades. Many first generation thin-film modules
exhibited reliability problems in outdoor testing, and the reliability of more recent thin-film
product has not yet been demonstrated since the product has only been in the field for a few
years.

Thus, the future of a-Si PV modules will depend critically on the ability to further increase the
stabilised efficiency, lower the manufacturing cost and improve the long-term reliability. The
major factor limiting the performance of a-Si PV modules is the light-induced degradation. At
this juncture, it is not clear that the light-induced degradation can be completely eliminated
since most investigators believe that the metastable defects are intrinsic to the a-Si alloys.
Nonetheless, there is a large R&D effort to develop a better understanding of the light-induced
degradation in a-Si alloys and to reduce the degradation. Considering the large potential payoff,
this effort is likely to continue for the near future [16].

The cost of thin-film PV modules and systems can be reduced by integrating thin-film PV into
building materials, but this approach requires both large-scale production of suitably sized
BIPV products and the integration of the products into the building industry infrastructure.

Another factor that affects the cost of a-Si PV modules is the relatively high capital cost of the
manufacturing equipment and facilities. This cost can be reduced by increasing the throughput
of the production equipment. At the present time, the production bottleneck and major
equipment cost are usually associated with the a-Si alloy deposition process where the
deposition rate is typically on the order of 0.1-0.2 nm/s.

In summary, while a-Si photovoltaics has become a $100 million dollar business, further
research, development and engineering will be required to increase the performance, lower the
manufacturing costs and improve the reliability in order to assure that a-Si PV will play a major
role in future world energy production [15].

19

3.1.4 Cadmium Telluride Solar Photovoltaics

Introduction
CdTe is very well suited for use as active material in thin-film solar cells due to four special
properties:

CdTe has an energy gap of 1.45 eV, and therefore is well adapted to the spectrum of
solar radiation.

The energy gap of CdTe is 'direct', leading to very strong light absorption.

CdTe has a strong tendency to grow as an essentially highly stoichiometric, but p-type
semiconductor film and can form a p-n heterojunction with CdS. (CdS has a rather wide
energy gap of 2.4 eV and grows n-type material under usual film deposition
techniques.)

Simple deposition techniques have been developed suited for low-cost production.

Current densities of up to 2 7 mA cm -2 and open-circuit voltages of 880 mV, leading to AM


1.5 efficiencies of 18%, can be expected for CdTe cells made under a mature technology [16].

Construction
In the preferred arrangement, first a transparent conducting film (typically In203 or SnO2 or a
combination of both) is deposited onto glass-plate used as transparent substrate. Then an nCdS film is deposited, followed by the active p-conducting CdTe film. A special treatment
improves the p-n junction between CdS and CdTe ('activation'). Finally a low-resistance
contact is deposited onto the CdTe, which can be opaque.

Working
Light enters the cell through the glass-substrate. Photons transverse the TCO and CdS layers.
These films are not active in the photovoltaic charge generation process although leading to
some- unwanted- absorption. The CdTe film is the active absorber of the solar cell. Electronhole pairs are generated close to the junction. The electrons are driven by the built-in field
through the interface into the n-CdS film. The holes remain in the CdTe and join the pool of

20

the holes promoting the p-conduction of this material and finally have to leave the cell via the
back contact. Electric power is drawn by metallic contacts attached to the TCO film and the
back contact [15].

Fig. 3.4 Film sequence of the CdTe thin-film solar cell.

Due to the strong light absorption in CdTe of about 105 cm -1 for light having a wavelength
below 800 nm, a film thickness of a few micrometers would be sufficient for complete light
absorption. For practical reasons a thickness of about 3-7 m is often preferred.

Intensive research has shown that this junction can be mastered so that the following basic
criteria for solar cells can be fulfilled under conditions of industrial production:

Effective generation of mobile minority charge carriers in the CdTe film.

Efficient separation of charge carriers by means of the internal electric field of the p-n
junction between n-CdS and p-CdTe.

Low loss-extraction of the photocurrent by means of ohmic contacts to the TCO and
back-contact films.

Simple fabrication technologies for low-cost, high-volume production.

Performance
Solar cells of efficiencies above 16% have been made in research laboratories and industrial
efforts have led to the recent start-up of industrial production units at three private companies
in the USA and Germany each aiming at large scale production of 100,000 m 2 per annum or
more. First large area modules have recently surpassed the 10% efficiency mark.

21

3.1.5 Copper Indium Gallium Selenide Solar cell

CuInSe2 was synthesized for the first time by Hahn in 1953. In 1974, this material was proposed
as a photovoltaic material with a power conversion efficiency of 12% for a single-crystal solar
cell. In the years 1983-84, Boeing corp. reported efficiencies in excess of 10% from thin
polycrystalline films obtained from a three-source co-evaporation process. In 1987 Arco Solar
achieved a long-standing record efficiency for a thin-film cell of 14.1%. It took a further ten
years, before Arco Solar, at that time Siemens Solar Industries (now Shell Solar), entered the
stage of production. In 1998, the first commercial Cu(In,Ga)Se2 solar modules were available
.In parallel, a process which avoids the use of H2Se is being developed by Shell Solar in
Germany [16].

Fig. 3.5 Composition a conventional CuInGaSe2

22

Fig. 3.6 Unit cells of chalcogenide compounds. (a) Sphalerite or zinc blende structure of ZnSe
(two unit cells); (b) chalcopyrite structure of CuInSe2. The metal sites in the two unit ceils of
the sphalerite structure of ZnSe are alternately occupied by Cu and In in the chalcopyrite
structure.

Performance
With a power conversion efficiency of 18.8% on a 0.5 cm 2 laboratory cell and 16.6% for minimodules with an area of around 20 cm2 Cu(In,Ga)Se2 is today by far the most efficient thinfilm solar cell technology. As of September 2014, current conversion efficiency record for a
laboratory CIGS cell stands at 21.7%. The start of production at several places provides a new
challenge for research on this material. However, these recent achievements are based on a
long history of research and technological development.

23

3.2

Dye Sensitized Solar Cells

Fig. 3.7 Dye sensitized Solar cell


The constraints determining the dye selection have already been indicated, and the production
of a suitable formulation is a demanding exercise in synthetic chemistry. Firstly and evidently
there is the matter of an optical absorption spectrum closely matched to the application of the
photovoltaic device, whether it be to the solar spectrum for energy conversion or to artificial
light sources for indoor use. With a high optical absorption coefficient across the visible
spectrum, the excitation process should be rapid but subsequent relaxation slow, as already
explained. For attachment to the semiconductor surface the molecule should adsorb strongly,
by preference through a specific chemical bond, to the substrate, but avoid aggregation so that
monolayer coverage forms spontaneously on contact between the semiconductor and the dye
in solution [17].
The dye structures selected for intensive development arose originally from biomimetic
considerations, given the role of photosynthesis in the natural world. The prototype energyconverting dye provided by nature is of course chlorophyll, a molecule consisting of a central
magnesium atom surrounded by a nitrogen-containing porphyrin ring. Variations are due to
minor modifications of certain side groups. Chlorophyll is in turn similar in structure to
hemoglobin, the oxygen-carrying iron-based pigment found in blood. Given that the
development of the dye-sensitised cell was associated with an interest in artificial
photosynthesis, the adoption of porphyrin-like organometallic dyes as sensitizers was logical.
However, whereas plant photosynthetic processes rely on chlorophyll the synthetic chemist can
select from the whole range of complex forming metals to design an appropriate metal-ligand
charge transfer structure [3].
24

Fig. 3.8 Certified I-V characteristic under simulated sunlight, of a dye-sensitised


photoelectrochemical cell, showing a conversion efficiency > 10% [2].

By 1980 the idea of chemisorption of the dye, through an acid carboxylate group bonding to
the metal oxide surface had already emerged so that the sensitiser was immobilised and formed
a monomolecular film on the semiconductor substrate, thereby facilitating charge transfer by
electron injection. The carboxylated trisbipyridyl dye ('RuL3') therefore became the prototype
25

sensitiser for this type of electrochemical cell. The objective at that time was to
photoelectrolyse water using sensitised electrodes. Although other compounds have since been
assessed as sensitisers, such as zinc porphyrins and even Prussian blue analogues, the most
suitable dyes today are still modifications of the ruthenium-based pyridyl complexes. The
objective had also evolved to become a photovoltaic device, rather than photosynthesis [17].

Working
The dye-sensitised photoelectrochemical cell integrates all these considerations, the molecular
engineering of a suitable dye, its adsorption as a monomolecular film on a rough wide bandgap
stable semiconductor to obtain adequate optical absorption, an ohmic contact to deliver the
resulting current to an external circuit, and finally establishment of a suitable regenerative
system through a cathodic counterelectrode and redox electrolyte, in a single device [21]. The
established semiconductor choice for this application is titanium dioxide, TiO2, with a bandgap
of 3.1 eV. It has many advantages for sensitised photochemistry and photoelectrochemistry,
being a low cost, widely available, non-toxic and biocompatible material, and as such it is even
used in health care products as well as in domestic applications such as paint pigmentation.
Since light must enter the cell to photoexcite the dye- semiconductor composite, the ohmic
contact is usually made to a transparent conducting oxide (TCO) such as an indium-tin oxide
or zinc oxide supported on glass or a suitable polymer. Most research has used the iodine/iodide
(3I- ~ I3- + 2e-) redox system, with others, including transition metal complexes, now under
development.

Performance
Progress of the sensitised electrochemical photovoltaic device since 1991, with a conversion
efficiency at that time of 7.1% under solar illumination, has been incremental, by optimising
the synergy of structure, substrate roughness, dye photochemistry, counter-electrode kinetics
and electrolyte redox chemistry. That evolution has continued progressively since then, with
certified efficiency now over 10%.

26

Commercial or Future Prospectus


The status of the dye-sensitised device as the only verified alternative to solid state junction
devices has already been mentioned. However it must be recognised that the solid-state devices,
particularly the silicon p-n junction cells, benefit from over 40 years of industrial and
development experience, the technology transfer from the silicon-based electronics industry,
and even the widespread availability of high quality silicon at low cost resulting from the
expansion of that industry. The procedures for high-yield fabrication of silicon devices, both
crystalline and amorphous, are well understood, with costing precisely established, based on
decades of solid industrial experience. For the dye-sensitised cell, in contrast, fabrication
procedures require development and costing is based on estimates of the requirements of
chemical processes rather than those of the silicon metallurgy with its elevated temperatures
and vacuum technology as required for conventional cells. This may in fact turn to the
advantage of some variants of the dye-sensitised concept. Equally it is well known that the
substitution of an established technology by an upcoming alternative requires that the new
concept has definite advantages and no clear disadvantages. It is therefore noted with some
satisfaction that several companies in Europe, Japan and Australia have taken up the challenge
and are currently engaged under license in the venture of industrialization and
commercialization of dye-sensitised PV cells. More, the existence of a credible challenger is a
stimulus to the solid-state photovoltaic industry to improve its existing products and to remain
open to new concepts.

27

3.3

Multi-junction solar Cell

Introduction
Multi-junction solar cells are solar cells with multiple pn junctions made of different
semiconductor materials. Each material's p-n junction will produce electric current in response
to a different wavelength of light. The use of multiple semiconducting materials allows the
absorbance of a broader range of wavelengths, improving the cell's sunlight to electrical energy
conversion efficiency.

Physics of Energy Conversion


The most common materials used in multijunction solar cells are III-IV semiconductors. IIIIV compounds have become the basic materials for modern optoelectronic devices. These IIIV compounds such as gallium arsenide (GaAs), indium phosphide (InP) and gallium
antimonide (GaSb) have some excellent characteristics that allow for fabrication high
efficiency solar cells. Some of the characteristics of these materials are direct energy band gaps,
high optical absorption coefficients, and good values of minority carrier lifetimes and
mobilities. When designing a multijunction solar cell there are three basic design
considerations: band gap differentiation, lattice constant matching, and current matching[7].

Design Structure of a Multijunction Solar Cell


The top of structure of a multijunction solar cell is very similar to the top of a single junction
solar cell. On top are metal contacts, usually made of Al, that touches the two sides of the
structure Figure 3.8. Under the top lies a p- or n-type doped semiconductor, in Figure 3.8 an
n-doped GaAs layer is shown. Also on the stop of the device is an anti-reflective coating. The
coat is generally a dual-layer dielectric stack. Common materials used to make the coating
are TiO2/Al2O3, Ta2O5/SiO2, or ZnS/MgF2. The reflective coating is designed to reduce the
large reflectance from around 30% to less than 1%. After the reflective coating is the top cell,
which is a p-n having the largest band gap of any other cell. The cells are connected via
tunnel junction. A tunnel junction is junction of two highly doped p- and n-type
semiconductors. The high doping creates an extremely thin depletion region that allows
tunneling to occur across the junction.. In order for the multijunction to preform properly the
band gap of the tunnel junction must be greater than the next cell. The tunnel junction helps
to separate the p-type first material from the n-type second material and it also connects the
28

two p-n junctions without have a large voltage drop. Following the tunnel junction is the ntype junction of the middle cell. The middle cells bandgap is less than the first cells but
larger than next cells if one exists. If another layer exists in the solar cell it is separated from
the previous cell with a tunnel junction. As seen in the Figure 3.8, between each cells p-n
junctions are two layers: a window layer and a back-surface field layer. Both layers create a
similar heterojunction, with the window layer being n-type and the back-surface field layer
being p-type. The purpose of the window later is to reduce surface combination while the
back-surface field layer decreases the scattering of carriers towards the tunnel junction [17].
The bandgap of the window must be greater than the back-surface field layer and both layers
lattice constant must be similar to the cells main material in order for the multijunction be
effective. Before the last layer of the multijunction solar cell are a buffer layer and a
nucleation layer. These layers are there to control the recombination of minority carriers and
to create a diffusion barrier for minority carriers. The final cell of the device is also the
substrate of the device. The material also has the smallest band gap and should theoretically
absorb most of the final photons available.

Fig. 3.9 Structure of a Triple-Junction PV Cell

29

Working
In order for a solar cell to work, the energy from incoming photons must have enough energy
to give electrons in the valence band sufficient energy to move to the conduction band.
Electrons are able to conduct to the conduction band without difficulty but are often conducted
so far beyond, or overshoot, the conduction band that a large portion of the energy is lost to
heat the in the crystal lattice as the electron moves to a lower energy state at the bottom of the
conduction band. If the energy of incoming photons is smaller than the band gap, electrons are
not given sufficient energy to move from the lower energy valence band to the higher energy
conduction band, and the incoming photons are not absorbed by the solar cell thus passing
through the material. This leaves a challenge to efficiently move electrons to the conduction
band and lose as little energy as possible to the lattice from overshooting. The solar spectrum
incorporates photons of different energy based on different wavelengths, so a single p-n
junction solar cell will only have limited range for conversion efficiency where electrons move
just beyond the conduction band and lattice heat loss is limited [19]. The solution to this
efficiency challenge is to use multiple p-n junctions stacked on top of each other with differing
band gap values such that together they are able to absorb a wider range of the solar spectrum.
This idea is referred to as Multijunction Solar Cells.

Performance
Multijunction or Tandum solar cells use a combination of different semiconductor materials
to form junctions to optimize the conversion of photons into electricity. Multijunction cells
were first studied in the 1960s and projected theoretical maximums in efficiency were between
38.2% and 51% depending on manufacturing technique. Advancements in semiconductor
technology have elevated the theoretical maximum efficiency of multijunction cells to 86.8%
from those first projected values in the 1960s. Progress in the development of multijunction
efficiency has shown promising results. From some of the initial demonstrations in the 1980s
of 16% to the current record of 43.7% set by Solar Junction in the middle of 2011 by using
Concentrated PVs with a dilute nitride cell architecture, it is expected that within the coming
years concentrated multijunction for commercial use will reach efficiencies of 50% [12].

30

3.4

Organic Solar Cell

Introduction
The last three years have seen an unprecedented growth of interest in solar cells made from
organic electronic materials. Organic materials are attractive for photovoltaics primarily
through the prospect of high throughput manufacture using processes such as reel-to-reel
deposition. Additional attractive features are the possibilities for ultra-thin, flexible devices
which may be integrated into appliances or building materials, and tuning of color through
chemical structure.
In 1959, Kallmann and Pope observed a photovoltaic effect in a single crystal of anthracene
when sandwiched between two identical electrodes and illuminated from one side (Kallmann
et al. 1959). While they could not completely explain the phenomenon, they postulated that
different exciton dissociation mechanisms must occur at the light and dark electrodes. Later,
they also observed a photovoltaic effect in a tetracenewater system (Geacintov et al. 1966).
Since this device was also completely symmetrical, except for illumination, they thought that
exciton dissociation via electron injection into the water, and hole transport by the organic
material away from the interface, could explain the observed behavior [2].

Mechanism of photon-to-electron conversion process


First, upon the absorption of light, an electron in the donor undergoes photoinduced excitation
from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular
orbital (LUMO) of the organic material, forming a Frenkel exciton (coulombically bound
electron (e-) and hole (h+)). The ratio of the generated Frenkel excitons to the total incident
photons, in terms of energy, is defined as the absorption efficiency. Excitons in organic
materials have a binding energy of 0.5 1 eV, due to their low dielectric permittivity (= 3 4).
In comparison, thermal energy (kT) at a room temperature (298 K) is approximately 0.025 eV,
substantially lower than the binding energy (0.5 1 eV) of an exciton in an organic material.
Thus, if the organic material is to serve as an electron donor in OPVs, a second material is
required as an electron acceptor to ensure a built-in internal field at the interface to break up
any excitons that diffuse there into free carriers. The most widely used acceptor materials are
fullerenes, which have electron affinities greater than those of polymers or small molecules.

31

Fig. 3.10 Fundamental mechanism of the photon-to-electron conversion process in bulk


heterojunction solar cells.

Second, the excitons must diffuse to the donoracceptor (DA) interfaces within the
diffusion length (LD) to prevent recombining to the ground state. Because the value of LD in
organic materials is typically 10 nm, the ideal donor or acceptor domain size is less than 20 nm.
This DA interface concept is analogousin terms of charge transportto a PN junction in
an inorganic semiconductor. The ratio of the number of excitons that reach the DA interface
to the total number of excitons generated through photoexcitation is defined as the exciton
diffusion efficiency.

Third, an exciton at a DA interface undergoes charge-transfer (CT) process at an


ultrafast pace (ca. 100 fs) to form a CT exciton, where the hole and electron remain in the donor
and acceptor phases, respectively, held together through coulombic attraction. The charge
separation efficiency is defined as the ratio of the number of excitons that have undergone the
CT process to the number of excitons that have reached the DA interface. Fourth, the CT
exciton dissociates, as a result of the builtin electric field, into free holes and electrons, which
are then transported through the donor and acceptor phases, respectively, to their respective
electrodes. The transport of free carriers to the respective electrodes occurs within a period of
32

time ranging from nano- to microseconds [2] [15]. The charge collection efficiency is defined
as the ratio of the number of carriers that have been collected at the electrodes to the number
of excitons that have undergone the CT process. Two recent developments in this area have
been it is possible for every absorbed photon resulting in a separated pair of charge carriers and
all photogenerated carriers being collected at the electrode for a push-pull polymer when an
TiOx optical spacer was incorporated in the device and (ii) when polymer and fullernes are
mixed at the molecular level, photogenerated excitons are next to the heterojunction and often
do not need to diffuse to the interfaces for CT process.

Conventional PVs V/s Organic PVs


In conventional cells, holes and electrons are generated together, in the same phase of the
material, and the photoinduced chemical potential gradient tends to drive them in the same
direction. This effect is greater on the minority carriers than on the majority carriers. In
addition, the built-in electric potential of inorganic devices drives the separation and flow of
holes and electrons. In contrast, in organic heterojunction devices, excitons dissociate at

Fig. 3.11 Schematic diagrams of a conventional pn junction solar cell (left) and an
organic heterojunction solar cell (right). The diagram highlights differences in carrier
generation between the two types of devices.
interfaces. So, the hole is generated in one phase (the donor phase) and the electron is generated
in the other phase (the acceptor phase). As a consequence of the free carriers being spatially
separated and existing in different phases, the photoinduced chemical potential drives them in
opposite directions. In heterojunction organic devices, built-in electrical fields may play a

33

smaller role in carrier movement, depending on device construction (solid state or dyesensitized)[18].

Performance
Based on current trends, efficiencies of 5-10% appear to be within reach, although stability
remains an obstacle.

Future trends
The great progress in OPVs in the last ten years can be attributed to advances in four fronts: (i)
a better understanding of the fundamental mechanism of photon-to-electron conversion; (ii)
advances in the molecular design of new materials with tailored energy levels and solubility;
(iii) new processing approaches to induce the optimal microstructure in the active layer and
new progress in the application of analytical tools; and (iv) new device architectures with
developments in optimizing the interfacial layers. To reach the holy grail of high-PCE OPVs
(cell: >15 %; module: >10 %) having large areas, fabricated through roll-to-roll solution
processes, and exhibiting device life times of greater than ten years will, however, require
concerted interdisciplinary efforts by both scientists and engineers to provide new approaches
toward solving critical technical issues. For instance, scientists must be able to prepare
multifunctional organic molecular structures that absorb into the visible and IR regions while
exhibiting good photo-stability, to synthesize new fullerene structures displaying
complementary absorptions and good solubility, and to perform simulations on material
properties with different chemical units; they also must have the ability to develop [18].

34

Chapter 4. Physical Aspects of Solar Cell Efficiency

Most of the energy that reaches a cell in the form of sunlight is lost before it can be converted
into electricity. Maximal sunlight-to-electricity conversion efficiencies for solar cells range up
to 30% (and even higher for some highly complex cell designs), but typical efficiencies are
10%-15% [1]. Most current work on cells is directed at enhancing efficiency while lowering
cost. Certain physical processes limit cell efficiency-some are inherent and cannot be changed;
many can be improved by proper design.
The major phenomena that limit cell efficiency are:
1. Reflection from the cell's surface
2. Light that is not energetic enough to separate electrons from their atomic bonds
3. Light that has extra energy beyond that needed to separate electrons from bonds
4. Light-generated electrons and holes (empty bonds) that randomly encounter each other
and recombine before they can contribute to cell performance
5. Light-generated electrons and holes that are brought together by surface and material
defects in the cell
6. Resistance to current flow
7. Self-shading resulting from top-surface electric contacts
8. Performance degradation at non optimal (high or low) operating temperatures [19]

35

SUMMARY

In this report Solar cell and other Solar PVs were given the importance have been given the
importance and their working conditions and application are included. The Principle of Solar
cell is elaborated to Study the basis behind the generation of electricity from the solar energy.
Sun playing an important role in supplying renewable energy to the earth. First generation solar
cell is the first of the kind solar cell, Second generation are the cell which were developed later
and Third generation Solar PVs were developed recently and their development in the history.
Past and Future scope of Solar cell in daily life. Physical Barriers for Solar cell efficiency is
also being considered. Semiconductor principle states that when the photons of sunlight pounds
on the semiconductor material creates a free electron and a hole which creates the potential
difference between semiconductors and thus electricity generation. We have deeply acquired
the knowledge of the Different types of solar photovoltaics and their principle behind their
operation. Thin film technology consisting CdTe, amorphous silicon and CIGS solar PVs
including Dye Sensitized, Multijunction Solar cell and Organic Solar cell are topics elaborated
to throw light on the solar PVs. Multijunction solar cell are used to increase efficiency of the
solar cell by a considerable increment. Use of organic PVs have reduced the cost of
manufacturing compared to conventional type of solar cell. Reason behind alleviation of the
solar efficiency is discussed to explain solar energy loss. Future prospects of solar
photovoltaics in the coming development age is precisely dependent upon the organic solar cell
and its type.

36

References:
1. Basic Photovoltaic Principles and Methods, Paul Hersch and Kenneth Zweibel,
2. Solar Cells: Materials, Manufacture and Operation by Elsevier publication,Tom
Markvart and Luis Castafier,215-433,
3. Rai G. D, Solar Energy Utilization by Nirali Prakashan, 4th edition.
4. Solar Photovoltaics: Fundamentals, Technologies and Applications, PHI Learning
Pvt. Ltd., Solanki Chetan Singh, Pg. 23-233
5. Ross, KG., Jr. 1976. "Interface Design Considerations for Terrestrial Solar Cell
Modules." Proceedings 12th IEEE Photovoltaic Specialists Conference1976. New
York, NY: Institute of Electrical and Electronic Engineers, Inc.; pp. 801-802.
6.

Paul Denholm, Robert M. Margolis, Evaluating the limits of solar photovoltaics


(PV) in electric power systems utilizing energy storage and other enabling
technologies, (2007), 4424-4433

7. Photovoltaic. Principles and. Me1hods. SERI/SP-290-1448. Solar Information


Module 6213,1982
8. Tim Jackson and Mark Oliver, The viability of solar photovoltaics, Centre for
Environmental Strategy, University of Surrey, Guildford, Surrey, Energy Policy 28
(2000), 983-988
9. Paul G. OBrien, Yang Yang, Along karn Chutinan, Pratish Mahtani, Keith Leong,
Daniel P.Puzzo, Leonardo D.Bonifacio, Chen-WeiLin, Geoffrey A.Ozin, Nazir
P.Kherani, Selectively transparent and conducting photonic crystal solar spectrum
splitters made of alternating sputtered indium-tin oxide and spin-coated silica
nanoparticle layers for enhanced photovoltaics, Solar Energy Materials & Solar
Cells,2012, 173183
10. Reindl, T., Krfihler, W., Pauli, M. and Mfiller,]., 1994. Electrical and Structural
Properties of the Si/C Interface in Poly-Si Thin Films on Graphite Substrates. Proc.
24th IEEE Photovoltaic Specialists Conf., Hawaii, pp. 1406-1409.
11. Brian E. McCandless and James R. Sites, Cadmium Telluride Solar Cells, Cadmium
Telluride cell, 2011, 617-667.
12. http://en.wikipedia.org/wiki/Solar_cell, 20 July
13. http://www.nature.com/nmat/journal/v7/n11/abs/nmat2312.html, 25 July
14. http://en.wikipedia.org/wiki/Photovoltaics, 29 July
37

15. http://www.nrel.gov/docs/legosti/fy98/24057.pdf ,28 Aug


16. http://ocw.tudelft.nl/fileadmin/ocw/courses/SolarCells/res00030/CH7_Thin_film_
Si_solar_cells.pdf, Sept 30
17. http://home.eng.iastate.edu/~sdmarch/documents/332%20Paper-Final%20DraftCherucheril,%20March,%20Verma.pdf, Oct 1
18. http://aerostudents.com/files/solarCells/CH8OrganicSolarCells.pdf, Oct 1
19. http://www.researchgate.net/profile/Dhandapani_Venkataraman/publication/7364
254_Organic_solar_cells_an_overview_focusing_on_active_layer_morphology/li
nks/0046352a3994ee4497000000, Oct 1

38

Potrebbero piacerti anche