You are on page 1of 71

PROJECT REPORT

SOLAR PHOTO-VOLTAIC STREET LIGHT


OVERVIEW

• This system is designed for outdoor application


especially in un-electrified remote rural areas.

• This system is an ideal application for campus and


village street lighting.

• The system is provided with battery storage backup


sufficient to operate the light for significant hours
daily.

• The system is provided with automatic ON/OFF time


switch for dusk to down operation and overcharge /
deep discharge prevention cut-off with LED indicators.
RENEWABLE SOURCE OF
ENERGY

• The 89 petawatts of sunlight reaching the Earth's surface is


plentiful - almost 6,000 times more than the 15 terawatts of
average electrical power consumed by humans. Additionally,
solar electric generation has the highest power density among
renewable energies.

• Solar power is pollution-free during use. Production end-wastes


and emissions are manageable using existing pollution controls.
End-of-use recycling technologies are under development.

• PV installations can operate for many years with little


maintenance or intervention after their initial set-up, so after
the initial capital cost of building any solar power plant,
operating costs are extremely low compared to existing power
technologies.

• Solar electric generation is economically superior where grid


connection or fuel transport is difficult, costly or impossible.
Long-standing examples include satellites, island communities,
remote locations and ocean vessels.
SOLAR POWER

Solar power is the generation of electricity from sunlight. This can


be direct as with photovoltaics (PV), or indirect as with
concentrating solar power (CSP), where the sun's energy is focused
to boil water which is then used to provide power. Solar power has
the potential to provide over 1,000 times total world energy
consumption in 2008, though it provided only 0.02% of the total that
year. If it continues to double in use every two to three years, or
less, it would become the dominant energy source this century. The
largest solar power plants, like the 354 MW SEGS, are
concentrating solar thermal plants, but recently multi-megawatt
photovoltaic plants have been built. Completed in 2008, the 46 MW
Moura photovoltaic power station in Portugal and the 40 MW
Waldpolenz Solar Park in Germany are characteristic of the trend
toward larger photovoltaic power stations. Much larger ones are
proposed, such as the 100 MW Fort Peck Solar Farm, the 550 MW
Topaz Solar Farm, and the 600 MW Rancho Cielo Solar Farm.

Terrestrial solar power is a predictably intermittent energy source,


meaning that whilst solar power is not available at all times, we can
predict with a very good degree of accuracy when it will and will not
be available. Some technologies, such as solar thermal concentrators
have an element of thermal storage, such as molten salts. These
store spare solar energy in the form of heat which can be made
available overnight or during periods that solar power is not available
to produce electricity. Orbital solar power collection (as in solar
power satellites) avoids this intermittent issue, but requires
satellite launching and beaming of the collected power to receiving
antennas on Earth. The increased intensity of sunlight above the
atmosphere also increases generation efficiency.
APPLICATIONS
Solar power is the conversion of sunlight to electricity. Sunlight can
be converted directly into electricity using photovoltaics (PV), or
indirectly with concentrating solar power (CSP), which normally
focuses the sun's energy to boil water which is then used to provide
power, and technologies such as the Stirling engine dishes which use
a Stirling cycle engine to power a generator. Photovoltaics were
initially used to power small and medium-sized applications, from the
calculator powered by a single solar cell to off-grid homes powered
by a photovoltaic array.

Solar power plants can face high installation costs, although this has
been decreasing due to the learning curve. Developing countries have
started to build solar power plants, replacing other sources of
energy generation.

Solar power has great potential, but in 2008 supplied only 0.02% of
the world's total energy supply. However, use has been doubling
every two, or less, years, and at that rate solar power, which has the
potential to supply over 1,000 times the total consumption of energy,
would become the dominant energy source within a few decades.

Since solar radiation is intermittent, solar power generation is


combined either with storage or other energy sources to provide
continuous power, although for small distributed
producer/consumers, net metering makes this transparent to the
consumer. On a larger scale, in Germany, a combined power plant has
been demonstrated, using a mix of wind, biomass, hydro-, and solar
power generation, resulting in 100% renewable energy .
CONCENTRATING SOLAR
POWER

A legend claims that Archimedes used polished shields to


concentrate sunlight on the invading Roman fleet and repel them
from Syracuse. Auguste Mouchout used a parabolic trough to
produce steam for the first solar steam engine in 1866.

Concentrating Solar Power (CSP) systems use lenses or mirrors and


tracking systems to focus a large area of sunlight into a small beam.
The concentrated heat is then used as a heat source for a
conventional power plant. A wide range of concentrating technologies
exists; the most developed are the parabolic trough, the
concentrating linear fresnel reflector, the Stirling dish and the
solar power tower. Various techniques are used to track the Sun and
focus light. In all of these systems a working fluid is heated by the
concentrated sunlight, and is then used for power generation or
energy storage

A parabolic trough consists of a linear parabolic reflector that


concentrates light onto a receiver positioned along the reflector's
focal line. The receiver is a tube positioned right above the middle
of the parabolic mirror and is filled with a working fluid. The
reflector is made to follow the Sun during the daylight hours by
tracking along a single axis. Parabolic trough systems provide the
best land-use factor of any solar technology.The SEGS plants in
California and Acciona's Nevada Solar One near Boulder City,
Nevada are representatives of this technology. The Suntrof-Mulk
parabolic trough, developed by Melvin Prueitt, uses a technique
inspired by Archimedes' principle to rotate the mirrors.

Concentrating Linear Fresnel Reflectors are CSP-plants which use


many thin mirror strips instead of parabolic mirrors to concentrate
sunlight onto two tubes with working fluid. This has the advantage
that flat mirrors can be used which are much cheaper than parabolic
mirrors, and that more reflectors can be placed in the same amount
of space, allowing more of the available sunlight to be used.
Concentrating linear fresnel reflectors can be used in either large or
more compact plants.

A Stirling solar dish, or dish engine system, consists of a stand-alone


parabolic reflector that concentrates light onto a receiver
positioned at the reflector's focal point. The reflector tracks the
Sun along two axes. Paraboloidal coordinates ("parabolic") dish
systems give the highest efficiency among CSP technologies. The
500 m2 ANU "Big Dish" in Canberra, Australia is an example of this
technology.The Stirling solar dish combines a parabolic
concentrating dish with a Stirling heat engine which normally drives
an electric generator. The advantages of Stirling solar over
photovoltaic cells are higher efficiency of converting sunlight into
electricity and longer lifetime. A solar power tower uses an array of
tracking reflectors (heliostats) to concentrate light on a central
receiver atop a tower. Power towers are more cost effective, offer
higher efficiency and better energy storage capability among CSP
technologies. The Solar Two in Barstow, California and the Planta
Solar 10 in Sanlucar la Mayor, Spain are representatives of this
technology.

A solar bowl is a spherical dish mirror that is fixed in place. The


receiver follows the line focus created by the dish (as opposed to a
point focus with tracking parabolic mirrors).
PHOTOVOLTAICS

A solar cell, or photovoltaic cell (PV), is a device that converts light


into electric current using the photoelectric effect. This is based on
the discovery by Alexandre-Edmond Becquerel who noticed that
some materials release electrons when hit with rays of photons from
light, which produces an electrical current. The first solar cell was
constructed by Charles Fritts in the 1880s. Although the prototype
selenium cells converted less than 1% of incident light into
electricity, both Ernst Werner von Siemens and James Clerk
Maxwell recognized the importance of this discovery.[23] Following
the work of Russell Ohl in the 1940s, researchers Gerald Pearson,
Calvin Fuller and Daryl Chapin created the silicon solar cell in 1954.
These early solar cells cost 286 USD/watt and reached efficiencies
of 4.5–6%. As of late 2009, the highest efficieincy PV cells were
produced commercially by Boeing/SpectroLab at about 41%. Other,
similar, multi-layer cells are close. These are very expensive
however, and are used only for the most exacting applications. Thin
film PV cells have been developed which are made in bulk and are far
less expensive and much less fragile, but are at most around 20%
efficient. The most recent development (from Caltech, March 2010)
is the experimental demonstration of a new design which is 85%
efficient in plain sunlight and 95% efficient at certain wavelengths.
It has only been produced in experimental laboratory examples, but
may have some possibility for low cost bulk production in the future.
There are many competing technologies, including at least fourteen
types of photovoltaic cells, such as thin film, monocrystalline silicon,
polycrystalline silicon, and amorphous cells, as well as multiple types
of concentrating solar power. It is too early to know which
technology will become dominant.

The earliest significant application of solar cells was as a back-up


power source to the Vanguard I satellite in 1958, which allowed it to
continue transmitting for over a year after its chemical battery was
exhausted.The successful operation of solar cells on this mission was
duplicated in many other Soviet and American satellites, and by the
late 1960s, PV had become the established source of power for
them. After the successful application of solar panels on the
Vanguard satellite it still was not until the energy crisis, in the
1970s, that photovoltaic solar panels gained use outside of back up
power suppliers on spacecraft. Photovoltaics went on to play an
essential part in the success of early commercial satellites such as
Telstar, and they remain vital to the telecommunications
infrastructure today.

The high cost of solar cells limited terrestrial uses throughout the
1960s. This changed in the early 1970s when prices reached levels
that made PV generation competitive in remote areas without grid
access. Early terrestrial uses included powering telecommunication
stations, offshore oil rigs, navigational buoys and railroad crossings.
These off-grid applications accounted for over half of worldwide
installed capacity until 2004.

The 1973 oil crisis stimulated a rapid rise in the production of PV


during the 1970s and early 1980s. Economies of scale which resulted
from increasing production along with improvements in system
performance brought the price of PV down from 100 USD/watt in
1971 to 7 USD/watt in 1985. Steadily falling oil prices during the
early 1980s led to a reduction in funding for photovoltaic R&D and a
discontinuation of the tax credits associated with the Energy Tax
Act of 1978. These factors moderated growth to approximately 15%
per year from 1984 through 1996.

Since the mid-1990s, leadership in the PV sector has shifted from


the US to Japan and Europe. Between 1992 and 1994 Japan
increased R&D funding, established net metering guidelines, and
introduced a subsidy program to encourage the installation of
residential PV systems. As a result, PV installations in the country
climbed from 31.2 MW in 1994 to 318 MW in 1999, and worldwide
production growth increased to 30% in the late 1990s.
Germany became the leading PV market worldwide since revising its
feed-in tariffs as part of the Renewable Energy Sources Act.
Installed PV capacity in Germany has risen from 100 MW in 2000 to
approximately 4,150 MW at the end of 2007. After 2007, Spain
became the largest PV market after adopting a similar feed-in tariff
structure in 2004, installing almost half of the photovoltaics (45%)
in the world, in 2008, while France, Italy, South Korea and the U.S.
have seen rapid growth recently due to various incentive programs
and local market conditions. The power output of domestic
photovoltaic devices is usually described in kilowatt-peak (kWp)
units, as most are from 1 to 10 kW.

Concentrating photovoltaics (CVP) are another new method of


electricity generation from the sun. CPV systems employ sunlight
concentrated onto photovoltaic surfaces for the purpose of
electrical power production. Solar concentrators of all varieties may
be used, which are often mounted on a solar tracker in order to keep
the focal point upon the cell as the sun moves across the sky.
Tracking can increase flat panel photovoltaic output by 20% in
winter, and by 50% in summer.
SOLAR CELL

A solar cell is a device that converts the energy of sunlight directly


into electricity by the photovoltaic effect. Sometimes the term
solar cell is reserved for devices intended specifically to capture
energy from sunlight such as solar panels and solar cells, while the
term photovoltaic cell is used when the light source is unspecified.
Assemblies of cells are used to make solar panels, solar modules, or
photovoltaic arrays. Photovoltaics is the field of technology and
research related to the application of solar cells in producing
electricity for practical use. The energy generated this way is an
example of solar energy (also known as solar power).

HISTORY OF SOLAR CELLS


The term "photovoltaic" comes from the Greek φῶς (phōs) meaning
"light", and "voltaic", meaning electric, from the name of the Italian
physicist Volta, after whom a unit of electro-motive force, the volt,
is named. The term "photo-voltaic" has been in use in English since
1849.

The photovoltaic effect was first recognized in 1839 by French


physicist A. E. Becquerel. However, it was not until 1883 that the
first solar cell was built, by Charles Fritts, who coated the
semiconductor selenium with an extremely thin layer of gold to form
the junctions. The device was only around 1% efficient. Subsequently
Russian physicist Aleksandr Stoletov built the first solar cell based
on the outer photoelectric effect (discovered by Heinrich Hertz
earlier in 1887). Albert Einstein explained the photoelectric effect
in 1905 for which he received the Nobel Prize in Physics in 1921.
Russell Ohl patented the modern junction semiconductor solar cell in
1946, which was discovered while working on the series of advances
that would lead to the transistor.

APPLICATIONS &
IMPLEMENTATIONS
Solar cells are often electrically connected and encapsulated as a
module. Photovoltaic modules often have a sheet of glass on the
front (sun up) side, allowing light to pass while protecting the
semiconductor wafers from the elements (rain, hail, etc.). Solar cells
are also usually connected in series in modules, creating an additive
voltage. Connecting cells in parallel will yield a higher current.
Modules are then interconnected, in series or parallel, or both, to
create an array with the desired peak DC voltage and current.

The power output of a solar array is measured in watts or kilowatts.


In order to calculate the typical energy needs of the application, a
measurement in watt-hours, kilowatt-hours or kilowatt-hours per day
is often used. A common rule of thumb is that average power is equal
to 20% of peak power, so that each peak kilowatt of solar array
output power corresponds to energy production of 4.8 kWh per day
(24 hours x 1 kW x 20% = 4.8 kWh)

To make practical use of the solar-generated energy, the electricity


is most often fed into the electricity grid using inverters (grid-
connected photovoltaic systems); in stand-alone systems, batteries
are used to store the energy that is not needed immediately.
Solar cells can also be applied to other electronics devices to make
it self-power sustainable in the sun. There are solar cell phone
chargers, solar bike light and solar camping lanterns that people can
adopt for daily use.

THEORY
Simple explanation
1. Photons in sunlight hit the solar panel and are absorbed by
semiconducting materials, such as silicon.
2. Electrons (negatively charged) are knocked loose from their
atoms, allowing them to flow through the material to produce
electricity. Due to the special composition of solar cells, the
electrons are only allowed to move in a single direction.
3. An array of solar cells converts solar energy into a usable
amount of direct current (DC) electricity.

Photo generation of charge carriers


When a photon hits a piece of silicon, one of three things can
happen:

1. the photon can pass straight through the silicon — this


(generally) happens for lower energy photons,
2. the photon can reflect off the surface,
3. The photon can be absorbed by the silicon, if the photon
energy is higher than the silicon band gap value. This generates
an electron-hole pair and sometimes heat, depending on the
band structure.

When a photon is absorbed, its energy is given to an electron in the


crystal lattice. Usually this electron is in the valence band, and is
tightly bound in covalent bonds between neighboring atoms, and
hence unable to move far. The energy given to it by the photon
"excites" it into the conduction band, where it is free to move
around within the semiconductor. The covalent bond that the
electron was previously a part of now has one fewer electron — this
is known as a hole. The presence of a missing covalent bond allows
the bonded electrons of neighboring atoms to move into the "hole,"
leaving another hole behind, and in this way a hole can move through
the lattice. Thus, it can be said that photons absorbed in the
semiconductor create mobile electron-hole pairs.

A photon need only have greater energy than that of the band gap in
order to excite an electron from the valence band into the
conduction band. However, the solar frequency spectrum
approximates a black body spectrum at ~6000 K, and as such, much
of the solar radiation reaching the Earth is composed of photons
with energies greater than the band gap of silicon. These higher
energy photons will be absorbed by the solar cell, but the difference
in energy between these photons and the silicon band gap is
converted into heat (via lattice vibrations — called phonons) rather
than into usable electrical energy.

Charge carrier separation


There are two main modes for charge carrier separation in a solar
cell:

1. drift of carriers, driven by an electrostatic field established


across the device
2. diffusion of carriers from zones of high carrier concentration
to zones of low carrier concentration (following a gradient of
electrochemical potential).

In the widely used p-n junction solar cells, the dominant mode of
charge carrier separation is by drift. However, in non-p-n-junction
solar cells (typical of the third generation solar cell research such as
dye and polymer solar cells), a general electrostatic field has been
confirmed to be absent, and the dominant mode of separation is via
charge carrier diffusion.

P-N JUNCTION
The most commonly known solar cell is configured as a large-area p-n
junction made from silicon. As a simplification, one can imagine
bringing a layer of n-type silicon into direct contact with a layer of
p-type silicon. In practice, p-n junctions of silicon solar cells are not
made in this way, but rather by diffusing an n-type dopant into one
side of a p-type wafer (or vice versa).

If a piece of p-type silicon is placed in intimate contact with a piece


of n-type silicon, then a diffusion of electrons occurs from the
region of high electron concentration (the n-type side of the
junction) into the region of low electron concentration (p-type side
of the junction). When the electrons diffuse across the p-n junction,
they recombine with holes on the p-type side. The diffusion of
carriers does not happen indefinitely, however, because charges
build up on either side of the junction and create an electric field.
The electric field creates a diode that promotes charge flow, known
as drift current, that opposes and eventually balances out the
diffusion of electron and holes. This region where electrons and
holes have diffused across the junction is called the depletion region
because it no longer contains any mobile charge carriers. It is also
known as the space charge region.

Equivalent circuit of a solar cell


The equivalent circuit of a solar cell

The schematic symbol of a solar cell

To understand the electronic behavior of a solar cell, it is useful to


create a model which is electrically equivalent, and is based on
discrete electrical components whose behavior is well known. An
ideal solar cell may be modeled by a current source in parallel with a
diode; in practice no solar cell is ideal, so a shunt resistance and a
series resistance component are added to the model. The resulting
equivalent circuit of a solar cell is shown on the left. Also shown, on
the right, is the schematic representation of a solar cell for use in
circuit diagrams.

Characteristic equation
From the equivalent circuit it is evident that the current produced
by the solar cell is equal to that produced by the current source,
minus that which flows through the diode, minus that which flows
through the shunt resistor:

I = IL − ID − ISH

Where

 I = output current (amperes)


 IL = photo generated current (amperes)
 ID = diode current (amperes)
 ISH = shunt current (amperes).

The current through these elements is governed by the voltage


across them:

Vj = V + IRS

Where

 Vj = voltage across both diode and resistor RSH (volts)


 V = voltage across the output terminals (volts)
 I = output current (amperes)
 RS = series resistance (Ω).

By the Shockley diode equation, the current diverted through the


diode is:

Where

 I0 = reverse saturation current (amperes)


 n = diode ideality factor (1 for an ideal diode)
 q = elementary charge
 k = Boltzmann's constant
 T = absolute temperature
 At 25°C, volts.

By Ohm's law, the current diverted through the shunt resistor is:

Where

 RSH = shunt resistance (Ω).

Substituting these into the first equation produces the


characteristic equation of a solar cell, which relates solar cell
parameters to the output current and voltage:

An alternative derivation produces an equation similar in appearance,


but with V on the left-hand side. The two alternatives are identities;
that is, they yield precisely the same results.

In principle, given a particular operating voltage V the equation may


be solved to determine the operating current I at that voltage.
However, because the equation involves I on both sides in a
transcendental function the equation has no general analytical
solution. However, even without a solution it is physically instructive.
Furthermore, it is easily solved using numerical methods. (A general
analytical solution to the equation is possible using Lambert's W
function, but since Lambert's W generally itself must be solved
numerically this is a technicality.)

Since the parameters I0, n, RS, and RSH cannot be measured directly,
the most common application of the characteristic equation is
nonlinear regression to extract the values of these parameters on
the basis of their combined effect on solar cell behavior.

Effect of physical size


The values of I0, RS, and RSH are dependent upon the physical size of
the solar cell. In comparing otherwise identical cells, a cell with
twice the surface area of another will, in principle, have double the
I0 because it has twice the junction area across which current can
leak. It will also have half the RS and RSH because it has twice the
cross-sectional area through which current can flow. For this reason,
the characteristic equation is frequently written in terms of current
density, or current produced per unit cell area:

Where

 J = current density (amperes/cm2)


 JL = photo generated current density (amperes/cm2)
 J0 = reverse saturation current density (amperes/cm2)
 rS = specific series resistance (Ω-cm2)
 rSH = specific shunt resistance (Ω-cm2).

This formulation has several advantages. One is that since cell


characteristics are referenced to a common cross-sectional area
they may be compared for cells of different physical dimensions.
While this is of limited benefit in a manufacturing setting, where all
cells tend to be the same size, it is useful in research and in
comparing cells between manufacturers. Another advantage is that
the density equation naturally scales the parameter values to similar
orders of magnitude, which can make numerical extraction of them
simpler and more accurate even with naive solution methods.

There are practical limitations of this formulation. For instance,


certain parasitic effects grow in importance as cell sizes shrink and
can affect the extracted parameter values. Recombination and
contamination of the junction tend to be greatest at the perimeter
of the cell, so very small cells may exhibit higher values of J0 or
lower values of RSH than larger cells that are otherwise identical. In
such cases, comparisons between cells must be made cautiously and
with these effects in mind.

This approach should only be used for comparing solar cells with
comparable layout. For instance, a comparison between primarily
quadratical solar cells like typical crystalline silicon solar cells and
narrow but long solar cells like typical thin film solar cells can lead to
wrong assumptions caused by the different kinds of current paths
and therefore the influence of for instance a distributed series
resistance rS.

Cell temperature

Temperature affects the characteristic equation in two ways:


directly, via T in the exponential term, and indirectly via its effect
on I0 (strictly speaking, temperature affects all of the terms, but
these two far more significantly than the others). While increasing
T reduces the magnitude of the exponent in the characteristic
equation, the value of I0 increases exponentially with T. The net
effect is to reduce VOC (the open-circuit voltage) linearly with
increasing temperature. The magnitude of this reduction is inversely
proportional to VOC; that is, cells with higher values of VOC suffer
smaller reductions in voltage with increasing temperature. For most
crystalline silicon solar cells the reduction is about 0.50%/°C, though
the rate for the highest-efficiency crystalline silicon cells is around
0.35%/°C. By way of comparison, the rate for amorphous silicon solar
cells is 0.20-0.30%/°C, depending on how the cell is made.

The amount of photo generated current IL increases slightly with


increasing temperature because of an increase in the number of
thermally generated carriers in the cell. This effect is slight,
however: about 0.065%/°C for crystalline silicon cells and 0.09% for
amorphous silicon cells.

The overall effect of temperature on cell efficiency can be


computed using these factors in combination with the characteristic
equation. However, since the change in voltage is much stronger than
the change in current, the overall effect on efficiency tends to be
similar to that on voltage. Most crystalline silicon solar cells decline
in efficiency by 0.50%/°C and most amorphous cells decline by 0.15-
0.25%/°C. The figure above shows I-V curves that might typically be
seen for a crystalline silicon solar cell at various temperatures.

Reverse saturation current

If one assumes infinite shunt resistance, the characteristic equation


can be solved for VOC:
Thus, an increase in I0 produces a reduction in VOC proportional to
the inverse of the logarithm of the increase. This explains
mathematically the reason for the reduction in VOC that accompanies
increases in temperature described above. The effect of reverse
saturation current on the I-V curve of a crystalline silicon solar cell
are shown in the figure to the right. Physically, reverse saturation
current is a measure of the "leakage" of carriers across the p-n
junction in reverse bias. This leakage is a result of carrier
recombination in the neutral regions on either side of the junction.
SOLAR CELL EFFICIENCY
Energy conversion efficiency

Dust often accumulates on the glass of solar panels seen here as


black dots.

A solar cell's energy conversion efficiency (η, "eta"), is the


percentage of power converted (from absorbed light to electrical
energy) and collected, when a solar cell is connected to an electrical
circuit. This term is calculated using the ratio of the maximum power
point, Pm, divided by the input light irradiance (E, in W/m2) under
standard test conditions (STC) and the surface area of the solar cell
(Ac in m2).

STC specifies a temperature of 25 °C and an irradiance of 1000


W/m2 with an air mass 1.5 (AM1.5) spectrums. These correspond to
the irradiance and spectrum of sunlight incident on a clear day upon
a sun-facing 37°-tilted surface with the sun at an angle of 41.81°
above the horizon. This condition approximately represents solar
noon near the spring and autumn equinoxes in the continental United
States with surface of the cell aimed directly at the sun. Thus,
under these conditions a solar cell of 12% efficiency with a 100 cm2
(0.01 m2) surface area can be expected to produce approximately 1.2
watts of power.
The efficiency of a solar cell may be broken down into reflectance
efficiency, thermodynamic efficiency, charge carrier separation
efficiency and conductive efficiency. The overall efficiency is the
product of each of these individual efficiencies.

Due to the difficulty in measuring these parameters directly, other


parameters are measured instead: thermodynamic efficiency,
quantum efficiency, VOC ratio, and fill factor. Reflectance losses are
a portion of the quantum efficiency under "external quantum
efficiency". Recombination losses make up a portion of the quantum
efficiency, VOC ratio, and fill factor. Resistive losses are
predominantly categorized under fill factor, but also make up minor
portions of the quantum efficiency, VOC ratio.

Thermodynamic efficiency limit


Solar cells operate as quantum energy conversion devices, and are
therefore subject to the "thermodynamic efficiency limit". Photons
with an energy below the band gap of the absorber material cannot
generate a hole-electron pair, and so their energy is not converted
to useful output and only generates heat if absorbed. For photons
with an energy above the band gap energy, only a fraction of the
energy above the band gap can be converted to useful output. When
a photon of greater energy is absorbed, the excess energy above
the band gap is converted to kinetic energy of the carrier
combination. The excess kinetic energy is converted to heat through
phonon interactions as the kinetic energy of the carriers slows to
equilibrium velocity.

Solar cells with multiple band gap absorber materials are able to
more efficiently convert the solar spectrum. By using multiple band
gaps, the solar spectrum may be broken down into smaller bins where
the thermodynamic efficiency limit is higher for each bin.
Quantum efficiency
As described above, when a photon is absorbed by a solar cell it can
produce a pair of free charge carriers, i.e. an electron-hole pair. One
of the carriers (the minority carrier) may then be able to reach the
p-n junction and contribute to the current produced by the solar
cell; such a carrier is said to be collected. Alternatively, the carrier
may give up its energy and once again become bound to an atom
within the solar cell without being collected; this process is then
called recombination since one electron and one hole recombine and
thereby annihilate the associated free charge. The carriers that
recombine do not contribute to the generation of electrical current.

Quantum efficiency refers to the percentage of photons that are


converted to electric current (i.e., collected carriers) when the cell
is operated under short circuit conditions. External quantum
efficiency (EQE) is the fraction of incident photons that are
converted to electrical current, while internal quantum efficiency
(IQE) is the fraction of absorbed photons that are converted to
electrical current. Mathematically, internal quantum efficiency is
related to external quantum efficiency by the reflectance (R) and
the transmittance (T) of the solar cell by IQE = EQE / (1 − R − T).
Please note that for a thick bulk Si solar cell T is approximately zero
and is therefore in practical cases often neglected.

Quantum efficiency should not be confused with energy conversion


efficiency, as it does not convey information about the fraction of
power that is converted by the solar cell. Furthermore, quantum
efficiency is most usefully expressed as a spectral measurement
(that is, as a function of photon wavelength or energy). Since some
wavelengths are absorbed more effectively than others in most
semiconductors, spectral measurements of quantum efficiency can
yield valuable information about the quality of the semiconductor
bulk and surfaces.
Maximum-power point
A solar cell may operate over a wide range of
voltages (V) and currents (I). By increasing the
resistive load on an irradiated cell continuously from
zero (a short circuit) to a very high value (an open
circuit) one can determine the maximum-power point,
the point that maximizes V×I; that is, the load for
which the cell can deliver maximum electrical power
at that level of irradiation. (The output power is zero
in both the short circuit and open circuit extremes).

A high quality, monocrystalline silicon solar cell, at 25


°C cell temperature, may produce 0.60 volts open-
circuit (VOC). The cell temperature in full sunlight,
even with 25 °C air temperature, will probably be
close to 45 °C, reducing the open-circuit voltage to
0.55 volts per cell. The voltage drops modestly, with
this type of cell, until the short-circuit current is
approached (Isc). Maximum power (with 45 °C cell
temperature) is typically produced with 75% to 80%
of the open-circuit voltage (0.43 volts in this case)
and 90% of the short-circuit current. This output
can be up to 70% of the VOC x ISC product. The short-
circuit current (Isc) from a cell is nearly proportional
to the illumination, while the open-circuit voltage
(VOC) may drop only 10% with a 80% drop in
illumination. Lower-quality cells have a more rapid
drop in voltage with increasing current and could
produce only 1/2 VOC at 1/2 ISC. The usable power
output could thus drop from 70% of the VOC x ISC
product to 50% or even as little as 25%. Vendors who
rate their solar cell "power" only as VOC x ISC, without
giving load curves, can be seriously distorting their
actual performance.

The maximum power point of a photovoltaic varies


with incident illumination. For systems large enough
to justify the extra expense, a maximum power point
tracker tracks the instantaneous power by
continually measuring the voltage and current (and
hence, power transfer), and uses this information to
dynamically adjust the load so the maximum power is
always transferred, regardless of the variation in
lighting.
SOLAR CELLS & ENERGY
PAYBACK

In the 1990s, when silicon cells were twice as thick, efficiencies


were much lower than today and lifetimes were shorter, it may well
have cost more energy to make a cell than it could generate in a
lifetime. In the meantime, the technology has progressed
significantly, and the energy payback time, defined as the recovery
time required for generating the energy spent for manufacturing of
the respective technical energy systems, of a modern photovoltaic
module is typically from 1 to 4 years depending on the module type
and location. Generally, thin-film technologies - despite having
comparatively low conversion efficiencies - achieve significantly
shorter energy payback times than conventional systems (often < 1
year). With a typical lifetime of 20 to 30 years, this means that
modern solar cells are net energy producers, i.e. they generate
significantly more energy over their lifetime than the energy
expended in producing them.

High-efficiency cells
High-efficiency solar cells are a class of solar cell that can
generate more electricity per incident solar power unit (watt/watt).
Much of the industry is focused on the most cost efficient
technologies in terms of cost per generated power. The two main
strategies to bring down the cost of photovoltaic electricity are
increasing the efficiency of the cells and decreasing their cost per
unit area. However, increasing the efficiency of a solar cell without
decreasing the total cost per kilowatt-hour is not more economical,
since sunlight is free. Thus, whether or not "efficiency" matters
depends on whether "cost" is defined as cost per unit of sunlight
falling on the cell, per unit area, per unit weight of the cell, or per
unit energy produced by the cell. In situations where much of the
cost of a solar system scales with its area (so that one is effectively
"paying" for sunlight), the challenge of increasing the photovoltaic
efficiency is thus of great interest, both from the academic and
economic points of view. Many groups have published papers claiming
possibility of high efficiencies after conducting optical
measurements under many hypothetical conditions. The efficiency
should be measured under real conditions and the basic parameters
that need to be evaluated are the short circuit current, open circuit
voltage

Thin-film solar cells


In 2002, the highest reported efficiency for thin film solar cells
based on CdTe is 18%, which was achieved by research at Sheffield
Hallam University, although this has not been confirmed by an
external test laboratory.

The US national renewable energy research facility NREL achieved


an efficiency of 19.9% for the solar cells based on copper indium
gallium selenide thin films, also known as CIGS (also see CIGS solar
cells).

NREL has since developed a robot that builds and analyzes the
efficiency of thin-film solar cells with the goal of increasing the
efficiency by testing the cells in different situations.

These CIGS films have been grown by physical vapour deposition in a


three-stage co-evaporation process. In this process In, Ga and Se
are evaporated in the first step; in the second step it is followed by
Cu and Se co-evaporation and in the last step terminated by In, Ga
and Se evaporation again.
Thin film solar has approximately 15% marketshare; the other 85%
is crystalline silicon. Most of the commercial production of thin film
solar is CdTe with an efficiency of 11%.

As of 28 April 2010, ZSW in Stuttgart have released a statement


claiming they have created CIGS-based cells with a new record
20.1% efficiency.

Crystalline Silicon
The highest efficiencies on silicon have been achieved on
monocrystalline cells. The highest commercial efficiency (22%) is
produced by SunPower, which uses expensive, high-quality silicon
wafers. The University of New South Wales has achieved 25%
efficiency on monocrystalline silicon in the lab, technology that has
been commercialized through its partnership with Suntech Power.
Crystalline silicon devices are approaching the theoretical limiting
efficiency of 29% and achieve an energy payback period of 1–2
years.

Light-absorbing materials
All solar cells require a light absorbing material contained within the
cell structure to absorb photons and generate electrons via the
photovoltaic effect. The materials used in solar cells tend to have
the property of preferentially absorbing the wavelengths of solar
light that reach the Earth surface. However, some solar cells are
optimized for light absorption beyond Earth's atmosphere as well.
Light absorbing materials can often be used in multiple physical
configurations to take advantage of different light absorption and
charge separation mechanisms.
Photovoltaic panels are normally made of either silicon or thin-film
cells:

Many currently available solar cells are configured as bulk materials


that are subsequently cut into wafers and treated in a "top-down"
method of synthesis (silicon being the most prevalent bulk material).

Other materials are configured as thin-films (inorganic layers,


organic dyes, and organic polymers) that are deposited on supporting
substrates, while a third group are configured as nanocrystals and
used as quantum dots (electron-confined nanoparticles) embedded in
a supporting matrix in a "bottom-up" approach. Silicon remains the
only material that is well-researched in both bulk (also called wafer-
based) and thin-film configurations.
CRYSTALLINE SILICON

By far, the most prevalent bulk material for solar cells is crystalline
silicon (abbreviated as a group as c-Si), also known as "solar grade
silicon". Bulk silicon is separated into multiple categories according
to crystallinity and crystal size in the resulting ingot, ribbon, or
wafer.

1. monocrystalline silicon (c-Si): often made using the Czochralski


process. Single-crystal wafer cells tend to be expensive, and
because they are cut from cylindrical ingots, do not completely
cover a square solar cell module without a substantial waste of
refined silicon. Hence most c-Si panels have uncovered gaps at
the four corners of the cells.
2. Poly- or multicrystalline silicon (poly-Si or mc-Si): made from
cast square ingots — large blocks of molten silicon carefully
cooled and solidified. Poly-Si cells are less expensive to
produce than single crystal silicon cells, but are less efficient.
US DOE data shows that there were a higher number of
multicrystalline sales than monocrystalline silicon sales.
3. Ribbon silicon is a type of multicrystalline silicon: it is formed
by drawing flat thin films from molten silicon and results in a
multicrystalline structure. These cells have lower efficiencies
than poly-Si, but save on production costs due to a great
reduction in silicon waste, as this approach does not require
sawing from ingots.

Thin films
The various thin-film technologies currently being developed reduce
the amount (or mass) of light absorbing material required in creating
a solar cell. This can lead to reduced processing costs from that of
bulk materials (in the case of silicon thin films) but also tends to
reduce energy conversion efficiency (average 7 to 10% efficiency),
although many multi-layer thin films have efficiencies above those of
bulk silicon wafers.

They have become popular compared to wafer silicon due to lower


costs and advantages including flexibility, lighter weights, and ease
of integration.

Cadmium telluride solar cell


A cadmium telluride solar cell is a solar cell based on cadmium
telluride, an efficient light-absorbing material for thin-film cells.
Compared to other thin-film materials, CdTe is easier to deposit and
more suitable for large-scale production.

There has been much discussion of the toxicity of CdTe-based solar


cells. The perception of the toxicity of CdTe is based on the toxicity
of elemental cadmium, a heavy metal that is a cumulative poison.
While the toxicity of CdTe is presently under debate, it has been
shown that the release of cadmium to the atmosphere is impossible
during normal operation of the cells and is unlikely during fires in
residential roofs.[33] Furthermore, a square meter of CdTe contains
approximately the same amount of Cd as a single C cell Nickel-
cadmium battery, in a more stable and less soluble form.

Copper-Indium Selenide

Possible combinations of (I, III, VI) elements in the periodic table


that have photovoltaic effect

The materials based on CuInSe2 that are of interest for


photovoltaic applications include several elements from groups I, III
and VI in the periodic table. These semiconductors are especially
attractive for thin film solar cell application because of their high
optical absorption coefficients and versatile optical and electrical
characteristics which can in principle be manipulated and tuned for a
specific need in a given device.

CIS is an abbreviation for general chalcopyrite films of copper


indium selenide (CuInSe2), CIGS mentioned below is a variation of
CIS. CIS films (no Ga) achieved greater than 14% efficiency.
However, manufacturing costs of CIS solar cells at present are high
when compared with amorphous silicon solar cells but continuing work
is leading to more cost-effective production processes. The first
large-scale production of CIS modules was started in 2006 in
Germany by Würth Solar. Manufacturing techniques vary and include
the use of Ultrasonic Nozzles for material deposition. Electro-
Plating in other efficient technology to apply the CI(G)S layer.
When gallium is substituted for some of the indium in CIS, the
material is referred to as CIGS, or copper indium/gallium diselenide,
a solid mixture of the semiconductors CuInSe2 and CuGaSe2, often
abbreviated by the chemical formula CuInxGa(1-x)Se2. Unlike the
conventional silicon based solar cell, which can be modelled as a
simple p-n junction (see under semiconductor), these cells are best
described by a more complex heterojunction model. The best
efficiency of a thin-film solar cell as of March 2008 was 19.9% with
CIGS absorber layer. Higher efficiencies (around 30%) can be
obtained by using optics to concentrate the incident light or by using
multi-junction tandem solar cells. The use of gallium increases the
optical bandgap of the CIGS layer as compared to pure CIS, thus
increasing the open-circuit voltage, but decreasing the short circuit
current. In another point of view, gallium is added to replace indium
due to gallium's relative availability to indium. Approximately 70% of
indium currently produced is used by the flat-screen monitor
industry. However, the atomic ratio for Ga in the >19% efficient
CIGS solar cells is ~7%, which corresponds to a bandgap of ~1.15 eV.
CIGS solar cells with higher Ga amounts have lower efficiency. For
example, CGS solar cells (which have a bandgap of ~1.7 eV have a
record efficiency of 9.5% for pure CGS and 10.2% for surface-
modified CGS. Some investors in solar technology worry that
production of CIGS cells will be limited by the availability of indium.
Producing 2 GW of CIGS cells (roughly the amount of silicon cells
produced in 2006) would use about 10% of the indium produced in
2004. For comparison, silicon solar cells used up 33% of the world's
electronic grade silicon production in 2006.

Se allows for better uniformity across the layer and so the number
of recombination sites in the film are reduced which benefits the
quantum efficiency and thus the conversion efficiency .
Gallium arsenide multijunction
High-efficiency multijunction cells were originally developed for
special applications such as satellites and space exploration, but at
present, their use in terrestrial concentrators might be the lowest
cost alternative in terms of $/kWh and $/W. These multijunction
cells consist of multiple thin films produced using metalorganic
vapour phase epitaxy. A triple-junction cell, for example, may consist
of the semiconductors: GaAs, Ge, and GaInP2. Each type of
semiconductor will have a characteristic band gap energy which,
loosely speaking, causes it to absorb light most efficiently at a
certain color, or more precisely, to absorb electromagnetic radiation
over a portion of the spectrum. The semiconductors are carefully
chosen to absorb nearly the entire solar spectrum, thus generating
electricity from as much of the solar energy as possible.

GaAs based multijunction devices are the most efficient solar cells
to date, reaching a record high of 40.7% efficiency under "500-sun"
solar concentration and laboratory conditions.

This technology is currently being utilized in the Mars rover


missions.

Tandem solar cells based on monolithic, series connected, gallium


indium phosphide (GaInP), gallium arsenide GaAs, and germanium Ge
pn junctions, are seeing demand rapidly rise. In just the past 12
months (12/2006 - 12/2007), the cost of 4N gallium metal has risen
from about $350 per kg to $680 per kg. Additionally, germanium
metal prices have risen substantially to $1000–$1200 per kg this
year. Those materials include gallium (4N, 6N and 7N Ga), arsenic
(4N, 6N and 7N) and germanium, pyrolitic boron nitride (pBN)
crucibles for growing crystals, and boron oxide, these products are
critical to the entire substrate manufacturing industry.

Triple-junction GaAs solar cells were also being used as the power
source of the Dutch four-time World Solar Challenge winners Nuna
in 2005 and 2007, and also by the Dutch solar cars Solutra (2005)
and Twente One (2007).

The Dutch Radboud University Nijmegen set the record for thin film
solar cell efficiency using a single junction GaAs to 25.8% in August
2008 using only 4 µm thick GaAs layer which can be transferred
from a wafer base to glass or plastic film.

Light-absorbing dyes (DSSC)


Typically a ruthenium metalorganic dye (Ru-centered) is used as a
monolayer of light-absorbing material. The dye-sensitized solar cell
depends on a mesoporous layer of nanoparticulate titanium dioxide
to greatly amplify the surface area (200–300 m2/g TiO2, as
compared to approximately 10 m2/g of flat single crystal). The photo
generated electrons from the light absorbing dye are passed on to
the n-type TiO2, and the holes are passed to an electrolyte on the
other side of the dye. The circuit is completed by a redox couple in
the electrolyte, which can be liquid or solid. This type of cell allows a
more flexible use of materials, and is typically manufactured by
screen printing and/or use of Ultrasonic Nozzles, with the potential
for lower processing costs than those used for bulk solar cells.
However, the dyes in these cells also suffer from degradation under
heat and UV light, and the cell casing is difficult to seal due to the
solvents used in assembly. In spite of the above, this is a popular
emerging technology with some commercial impact forecast within
this decade. The first commercial shipment of DSSC solar modules
occurred in July 2009 from G24i Innovations.
Organic/polymer solar cells
Organic solar cells and polymer solar cells are built from thin films
(typically 100 nm) of organic semiconductors such as polymers and
small-molecule compounds like polyphenylene vinylene, copper
phthalocyanine (a blue or green organic pigment) and carbon
fullerenes and fullerene derivatives such as PCBM. Energy conversion
efficiencies achieved to date using conductive polymers are low
compared to inorganic materials. However, it improved quickly in the
last few years and the highest NREL (National Renewable Energy
Laboratory) certified efficiency has reached 6.77%. In addition,
these cells could be beneficial for some applications where
mechanical flexibility and disposability are important.

These devices differ from inorganic semiconductor solar cells in


that they do not rely on the large built-in electric field of a PN
junction to separate the electrons and holes created when photons
are absorbed. The active region of an organic device consists of two
materials, one which acts as an electron donor and the other as an
acceptor. When a photon is converted into an electron hole pair,
typically in the donor material, the charges tend to remain bound in
the form of an exciton, and are separated when the exciton diffuses
to the donor-acceptor interface. The short exciton diffusion lengths
of most polymer systems tend to limit the efficiency of such
devices. Nanostructured interfaces, sometimes in the form of bulk
heterojunctions, can improve performance.
Silicon thin films
Silicon thin-film cells are mainly deposited by chemical vapor
deposition (typically plasma-enhanced (PE-CVD)) from silane gas and
hydrogen gas. Depending on the deposition parameters, this can
yield:

1. Amorphous silicon (a-Si or a-Si:H)


2. Protocrystalline silicon or
3. Nanocrystalline silicon (nc-Si or nc-Si:H), also called
microcrystalline silicon.

It has been found that protocrystalline silicon with a low volume


fraction of nanocrystalline silicon is optimal for high open circuit
voltage. These types of silicon present dangling and twisted bonds,
which results in deep defects (energy levels in the bandgap) as well
as deformation of the valence and conduction bands (band tails). The
solar cells made from these materials tend to have lower energy
conversion efficiency than bulk silicon, but are also less expensive to
produce. The quantum efficiency of thin film solar cells is also lower
due to reduced number of collected charge carriers per incident
photon.

Amorphous silicon has a higher bandgap (1.7 eV) than crystalline


silicon (c-Si) (1.1 eV), which means it absorbs the visible part of the
solar spectrum more strongly than the infrared portion of the
spectrum. As nc-Si has about the same bandgap as c-Si, the nc-Si
and a-Si can advantageously be combined in thin layers, creating a
layered cell called a tandem cell. The top cell in a-Si absorbs the
visible light and leaves the infrared part of the spectrum for the
bottom cell in nc-Si.
CONCENTRATING
PHOTOVOLTAICS (CPV)
Concentrating photovoltaic systems use a large area of lenses or
mirrors to focus sunlight on a small area of photovoltaic cells. High
concentration means a hundred or more times direct sunlight is
focused when compared with crystalline silicon panels. Most
commercial producers are developing systems that concentrate
between 400 and 1000 suns. All concentration systems need a one
axis or more often two axis tracking system for high precision, since
most systems only use direct sunlight and need to aim at the sun
with errors of less than 3 degrees. The primary attraction of CPV
systems is their reduced usage of semiconducting material which is
expensive and currently in short supply. Additionally, increasing the
concentration ratio improves the performance of high efficiency
photovoltaic cells. Despite the advantages of CPV technologies their
application has been limited by the costs of focusing, sun tracking
and cooling equipment. On October 25, 2006, the Australian federal
government and the Victorian state government together with
photovoltaic technology company Solar Systems announced a project
using this technology, Solar power station in Victoria, planned to
come online in 2008 and be completed by 2013. This plant, at 154
MW, would be ten times larger than the largest current photovoltaic
plant in the world.

Lifespan
Most commercially available solar cells are capable of producing
electricity for at least twenty years without a significant decrease
in efficiency. The typical warranty given by panel manufacturers is
for a period of 25 – 30 years, wherein the output shall not fall below
85% of the rated capacity.
Costs
Cost is established in cost-per-watt and in cost-per-watt in 24 hours
for infrared capable photovoltaic cells. Manufacturing costs are also
calculated including the energy required for manufacturing of the
cells and modules in a kWh basis. These figures are added to the end
price for solar investors and the energy payback is calculated from
the point of power plant initialization or connection to the grid.
another method of calculating the payback is to use the feed in
tariff mechanism in place for power plant remuneration. Solar-
specific feed in tariffs vary worldwide, and even state by state
within various countries. The energy payback time will vary
depending on the country of application and the level of the feed in
tariff.

Low-cost solar cell


Dye-sensitized solar cell, and luminescent solar concentrators are
considered low-cost solar cells.

This cell is extremely promising because it is made of low-cost


materials and does not need elaborate apparatus to manufacture, so
it can be made in a DIY way allowing more players to produce it than
any other type of solar cell. In bulk it should be significantly less
expensive than older solid-state cell designs. It can be engineered
into flexible sheets. Although its conversion efficiency is less than
the best thin film cells, its price/performance ratio should be high
enough to allow it to compete with fossil fuel electrical generation.
HARDWARE INVOLVED

1) Main Circuit
-Inverter cum charge controller circuit
-IC LM324N
-IC NE555P
2) PV Panel

3) Battery

4) CFL

The solar street light system comprise of

• 75 Wp Solar PV Module

• 12 V, 75 Ah VRLA Battery (c/10 Discharge rate)

• Charge Controller cum inverter

• 11 Watt CFL Lamp  

The SPV modules are reported to have a service life of 15-20 years.
VRLA Batteries provided with the solar street lighting system
require lower maintenance; have longer life and give better
performance as compared to pasted plate batteries used earlier.

The systems electronic provide for over-charge and over-discharge


cut-off essential for preventing battery and luminaries damages.
OP-AMP LM324N

These circuits consist of four independent, high gain,


internally frequency-compensated operational amplifiers.
They operate from a single power supply over a wide range
of voltages. Operation from split power supplies is also
possible and the low power supply current drain is
independent of the magnitude of the power supply voltage.

Features

■ Wide gain bandwidth: 1.3 MHz

■ Large voltage gain: 100 dB

■ Wide power supply range:

– Single supply: +3 V to +30 V

■ Dual supplies: ±1.5 V to ±15 V


BATTERY USED
Lead Acid Valve Regulated Battery (VRLA)

Advantages:

• The battery can be mounted in any position.

• There is no need to check the level of electrolyte or to


top up water lost due to electrolysis, reducing
inspection and maintenance.

We used 12 v battery with c/10 discharge rate


COMPACT FLUOROCENT
LIGHT (CFL)
Compared to incandescent lamps, compact fluorescent
lamps, when used properly have the following advantages:

 Last up to 10 times longer

 Use about one-fourth the energy

 Produce 90% less heat, while producing more light per


watt

We used 11 watt CFL (indo asian)


555 timer IC

NE555 from Signetics in dual-in-line package

Internal block diagram

The 555 Timer IC is an integrated circuit (chip) implementing a


variety of timer and multivibrator applications. The IC was
designed by Hans R. Camenzind in 1970 and brought to market in
1971 by Signetics (later acquired by Philips). The original name
was the SE555 (metal can)/NE555 (plastic DIP) and the part was
described as "The IC Time Machine".It has been claimed that the
555 gets its name from the three 5 kΩ resistors used in typical
early implementations, but Hans Camenzind has stated that the
number was arbitrary. The part is still in wide use, thanks to its
ease of use, low price and good stability. As of 2003, it is
estimated that 1 billion units are manufactured every year.

Depending on the manufacturer, the standard 555 package


includes over 20 transistors, 2 diodes and 15 resistors on a silicon
chip installed in an 8-pin mini dual-in-line package (DIP-8).
Variants available include the 556 (a 14-pin DIP combining two
555s on one chip), and the 558 (a 16-pin DIP combining four
slightly modified 555s with DIS & THR connected internally, and
TR falling edge sensitive instead of level sensitive).

Ultra-low power versions of the 555 are also available, such as


the 7555 and TLC555. The 7555 requires slightly different
wiring using fewer external components and less power.

The 555 has three operating modes:

 Monostable mode: in this mode, the 555 functions as a "one-


shot". Applications include timers, missing pulse detection,
bouncefree switches, touch switches, frequency divider,
capacitance measurement, pulse-width modulation
(PWM) etc

 Astable - free running mode: the 555 can operate as an


oscillator. Uses include LED and lamp flashers, pulse
generation, logic clocks, tone generation, security
alarms, pulse position modulation, etc.

 Bistable mode or Schmitt trigger: the 555 can operate


as a flip-flop, if the DIS pin is not connected and no
capacitor is used. Uses include bounce free latched
switches, etc.

Usage

Pinout diagram

The connection of the pins is as follows:

Pin Name Purpose

1 GND Ground, low level (0 V)

OUT rises, and interval starts, when this input


2 TRIG
falls below 1/3 VCC.

3 OUT This output is driven to +VCC or GND.

A timing interval may be interrupted by driving


4 RESET
this input to GND.

"Control" access to the internal voltage divider


5 CTRL
(by default, 2/3 VCC).

The interval ends when the voltage at THR is


6 THR
greater than at CTRL.

7 DIS Open collector output; may discharge a capacitor


between intervals.

Positive supply voltage is usually between 3 and 15


8 V+, VCC
V.

Monostable mode

Schematic of a 555 in monostable mode

The relationships of the trigger signal, the voltage on C and the


pulse width in monostable mode
In the monostable mode, the 555 timer acts as a “one-shot” pulse
generator. The pulse begins when the 555 timer receives a signal
at the trigger input that falls below a third of the voltage supply.
The width of the pulse is determined by the time constant of an
RC network, which consists of a capacitor (C) and a resistor (R).
The pulse ends when the charge on the C equals 2/3 of the supply
voltage. The pulse width can be lengthened or shortened to the
need of the specific application by adjusting the values of R and
C.

The pulse width of time t, which is the time it takes to charge C


to 2/3 of the supply voltage, is given by

where t is in seconds, R is in ohms and C is in farads. See RC


circuit for an explanation of this effect.

Bistable Mode

In bistable mode, the 555 timer acts as a basic flip-flop. The


trigger and reset inputs (pins 2 and 4 respectively on a 555) are
held high via pull-up resistors while the threshold input (pin 6) is
simply grounded. Thus configured, pulling the trigger momentarily
to ground acts as a 'set' and transitions the output pin (pin 3) to
Vcc (high state). Pulling the reset input to ground acts as a
'reset' and transitions the output pin to ground (low state). No
capacitors are required in a bistable configuration. Pin 8 (Vcc) is,
of course, tied to Vcc while pin 1 (Gnd) is grounded. Pins 5 and 7
(control and discharge) are left floating.

Astable mode
Standard 555 Astable Circuit

In astable mode, the '555 timer ' puts out a continuous stream of
rectangular pulses having a specified frequency. Resistor R 1 is
connected between VCC and the discharge pin (pin 7) and another
resistor (R2) is connected between the discharge pin (pin 7), and
the trigger (pin 2) and threshold (pin 6) pins that share a common
node. Hence the capacitor is charged through R1 and R2, and
discharged only through R2, since pin 7 has low impedance to
ground during output low intervals of the cycle, therefore
discharging the capacitor.

In the astable mode, the frequency of the pulse stream depends


on the values of R1, R2 and C:

The high time from each pulse is given by

and the low time from each pulse is given by


where R1 and R2 are the values of the resistors in ohms and C is
the value of the capacitor in farads.

Specifications
These specifications apply to the NE555. Other 555 timers can
have better specifications depending on the grade (military,
medical, etc).

Supply voltage (VCC) 4.5 to 15 V

Supply current (VCC = +5 V) 3 to 6 mA

Supply current (VCC = +15 V) 10 to 15 mA

Output current (maximum) 200 mA

Power dissipation 600 mW

Operating temperature 0 to 70 °C

Derivatives
Many pin-compatible variants, including CMOS versions, have been
built by various companies. Bigger packages also exist with two or
four timers on the same chip. The 555 is also known under the
following type numbers:

Manufacturer Model Remark

CMOS from
Custom Silicon
CSS555/CSS555C 1.2 V,
Solutions
IDD < 5 µA

Avago Technologies Av-555M

ECG Philips ECG955M

Exar XR-555

Fairchild
NE555/KA555
Semiconductor

Harris HA555

CMOS from
IK Semicon ILC555
2 V

Intersil SE555/NE555

Intersil ICM7555 CMOS

Lithic Systems LC555

CMOS from
Maxim ICM7555
2 V

Motorola MC1455/MC1555

National
LM1455/LM555/LM555C
Semiconductor

National LMC555 CMOS from


Semiconductor 1.5 V

NTE Sylvania NTE955M

Raytheon RM555/RC555

RCA CA555/CA555C

STMicroelectronics NE555N/ K3T647

Texas Instruments SN52555/SN72555

Dual timer 556

The dual version is called 556. It features two complete 555s in a


14 pin DIL package.

Quad timer 558

The quad version is called 558 and has 16 pins. To fit four 555s
into a 16 pin package the control voltage and reset lines are
shared by all four modules. Also for each module the discharge
and threshold are internally wired together and called timing.

OPERATIONAL AMPLIFIER
(LM324N)
Various op-amp ICs in eight-pin dual in-line packages ("DIPs")

An operational amplifier, which is often called an op-amp, is a


DC-coupled high-gain electronic voltage amplifier with a
differential input and, usually, a single-ended output. An op-amp
produces an output voltage that is typically millions of times
larger than the voltage difference between its input terminals.

Typically the op-amp's very large gain is controlled by negative


feedback, which largely determines the magnitude of its output
("closed-loop") voltage gain in amplifier applications, or the
transfer function required (in analog computers). Without
negative feedback, and perhaps with positive feedback for
regeneration, an op-amp essentially acts as a comparator. High
input impedance at the input terminals (ideally infinite) and low
output impedance at the output terminal(s) (ideally zero) are
important typical characteristics.

Op-amps are among the most widely used electronic devices


today, being used in a vast array of consumer, industrial, and
scientific devices. Many standard IC op-amps cost only a few
cents in moderate production volume; however some integrated or
hybrid operational amplifiers with special performance
specifications may cost over $100 US in small quantities. Op-amps
sometimes come in the form of macroscopic components, (see
photo) or as integrated circuit cells; patterns that can be
reprinted several times on one chip as part of a more complex
device.

The op-amp is one type of differential amplifier. Other types of


differential amplifier include the fully differential amplifier
(similar to the op-amp, but with two outputs), the instrumentation
amplifier (usually built from three op-amps), the isolation
amplifier (similar to the instrumentation amplifier, but with
tolerance to common-mode voltages that would destroy an
ordinary op-amp), and negative feedback amplifier (usually built
from one or more op-amps and a resistive feedback network).

Circuit notation
Circuit diagram symbol for an op-amp

The circuit symbol for an op-amp is shown to the right, where:

 : non-inverting input

 : inverting input

 : output

 : positive power supply

 : negative power supply

The power supply pins ( and ) can be labeled in different


ways. Despite different labeling, the function remains the same —
to provide additional power for amplification of the signal. Often
these pins are left out of the diagram for clarity, and the power
configuration is described or assumed from the circuit.

Operation
The amplifier's differential inputs consist of a input and a
input, and ideally the op-amp amplifies only the difference in
voltage between the two, which is called the differential input
voltage. The output voltage of the op-amp is given by the
equation,
where is the voltage at the non-inverting terminal, is the
voltage at the inverting terminal and AOL is the open-loop gain of
the amplifier. (The term "open-loop" refers to the absence of a
feedback loop from the output to the input.)

With no negative feedback, the op-amp acts as a comparator. The


inverting input is held at ground (0 V) by the resistor, so if the V in
applied to the non-inverting input is positive, the output will be
maximum positive, and if Vin is negative, the output will be
maximum negative. Since there is no feedback from the output to
either input, this is an open loop circuit. The circuit's gain is just
the GOL of the op-amp.

Adding negative feedback via the voltage divider Rf,Rg reduces


the gain. Equilibrium will be established when Vout is just
sufficient to reach around and "pull" the inverting input to the
same voltage as Vin. As a simple example, if Vin = 1 V and Rf = Rg,
Vout will be 2 V, the amount required to keep V– at 1 V. Because of
the feedback provided by Rf,Rg this is a closed loop circuit. Its
over-all gain Vout / Vin is called the closed-loop gain ACL. Because
the feedback is negative, in this case ACL is less than the AOL of
the op-amp.
The magnitude of AOL is typically very large—seldom less than a
million—and therefore even a quite small difference between
and (a few microvolts or less) will result in amplifier saturation,
where the output voltage goes to either the extreme maximum or
minimum end of its range, which is set approximately by the
power supply voltages. Additionally, the precise magnitude of AOL
is not well controlled by the manufacturing process, and so it is
impractical to use an operational amplifier as a stand-alone
differential amplifier. If linear operation is desired, negative
feedback must be used, usually achieved by applying a portion of
the output voltage to the inverting input. The feedback enables
the output of the amplifier to keep the inputs at or near the same
voltage so that saturation does not occur. Another benefit is that
if much negative feedback is used, the circuit's overall gain and
other parameters become determined more by the feedback
network than by the op-amp itself. If the feedback network is
made of components with relatively constant, predictable, values
such as resistors, capacitors and inductors, the unpredictability
and inconstancy of the op-amp's parameters (typical of
semiconductor devices) do not seriously affect the circuit's
performance.

If no negative feedback is used, the op-amp functions as a switch


or comparator.

Positive feedback may be used to introduce hysteresis or


oscillation.

Ideal and real op-amps


An equivalent circuit of an operational amplifier that models some resistive
non-ideal parameters.

An ideal op-amp is usually considered to have the following


properties, and they are considered to hold for all input voltages:

 Infinite open-loop gain (when doing theoretical analysis, a limit


may be taken as open loop gain AOL goes to infinity)

 Infinite voltage range available at the output (vout) (in practice


the voltages available from the output are limited by the
supply voltages and )

 Infinite bandwidth (i.e., the frequency magnitude response is


considered to be flat everywhere with zero phase shift).

 Infinite input impedance (so, in the diagram, , and zero


current flows from to )

 Zero input current (i.e., there is assumed to be no leakage or


bias current into the device)

 Zero input offset voltage (i.e., when the input terminals are
shorted so that , the output is a virtual ground or vout =
0).
 Infinite slew rate (i.e., the rate of change of the output
voltage is unbounded) and power bandwidth (full output voltage
and current available at all frequencies).

 Zero output impedance (i.e., Rout = 0, so that output voltage


does not vary with output current)

 Zero noise

 Infinite Common-mode rejection ratio (CMRR)

 Infinite Power supply rejection ratio for both power supply


rails.

In practice, none of these ideals can be realized, and various


shortcomings and compromises have to be accepted. Depending on
the parameters of interest, a real op-amp may be modeled to take
account of some of the non-infinite or non-zero parameters using
equivalent resistors and capacitors in the op-amp model. The
designer can then include the effects of these undesirable, but
real, effects into the overall performance of the final circuit.
Some parameters may turn out to have negligible effect on the
final design while others represent actual limitations of the final
performance that must be evaluated.

Recent trends
Recently supply voltages in analog circuits have decreased (as
they have in digital logic) and low-voltage opamps have been
introduced reflecting this. Supplies of ±5V and increasingly 5V
are common. To maximize the signal range modern op-amps
commonly have rail-to-rail inputs (the input signals can range from
the lowest supply voltage to the highest) and sometimes rail-to-
rail outputs.

Classification
Op-amps may be classified by their construction:

 discrete (built from individual transistors or tubes/valves)

 IC (fabricated in an Integrated circuit) - most common

 hybrid

IC op-amps may be classified in many ways, including:

 Military, Industrial, or Commercial grade (for example: the


LM301 is the commercial grade version of the LM101, the
LM201 is the industrial version). This may define operating
temperature ranges and other environmental or quality factors.

 Classification by package type may also affect environmental


hardiness, as well as manufacturing options; DIP, and other
through-hole packages are tending to be replaced by Surface-
mount devices.

 Classification by internal compensation: op-amps may suffer


from high frequency instability in some negative feedback
circuits unless a small compensation capacitor modifies the
phase- and frequency- responses; op-amps with capacitor built
in are termed "compensated", or perhaps compensated for
closed-loop gains down to (say) 5, others: uncompensated.

 Single, dual and quad versions of many commercial op-amp IC


are available, meaning 1, 2 or 4 operational amplifiers are
included in the same package.

 Rail-to-rail input (and/or output) op-amps can work with input


(and/or output) signals very close to the power supply rails.

 CMOS op-amps (such as the CA3140E) provide extremely high


input resistances, higher than JFET-input op-amps, which are
normally higher than bipolar-input op-amps.

 other varieties of op-amp include programmable op-amps


(simply meaning the quiescent current, gain, bandwidth and so
on can be adjusted slightly by an external resistor).

 manufacturers often tabulate their op-amps according to


purpose, such as low-noise pre-amplifiers, wide bandwidth
amplifiers, and so on.

Applications
Use in electronics system design

The use of op-amps as circuit blocks is much easier and clearer


than specifying all their individual circuit elements (transistors,
resistors, etc.), whether the amplifiers used are integrated or
discrete. In the first approximation op-amps can be used as if
they were ideal differential gain blocks; at a later stage limits can
be placed on the acceptable range of parameters for each op-
amp.

Circuit design follows the same lines for all electronic circuits. A
specification is drawn up governing what the circuit is required to
do, with allowable limits. For example, the gain may be required to
be 100 times, with a tolerance of 5% but drift of less than 1% in a
specified temperature range; the input impedance not less than
one megohm; etc.

A basic circuit is designed, often with the help of circuit modeling


(on a computer). Specific commercially available op-amps and
other components are then chosen that meet the design criteria
within the specified tolerances at acceptable cost. If not all
criteria can be met, the specification may need to be modified.

A prototype is then built and tested; changes to meet or


improve the specification, alter functionality, or reduce
the cost, may be made.

Basic single stage amplifiers


Non-inverting amplifier

An op-amp connected in the non-inverting amplifier configuration

In a non-inverting amplifier, the output voltage changes in the


same direction as the input voltage.

The gain equation for the op-amp is:

However, in this circuit V– is a function of Vout because of the


negative feedback through the R1R2 network. R1 and R2 form a
voltage divider, and as V– is a high-impedance input, it does not
load it appreciably. Consequently:

Where

Substituting this into the gain equation, we obtain:

Solving for Vout:

If AOL is very large, this simplifies to


.

Inverting amplifier

An op-amp connected in the inverting amplifier configuration

In an inverting amplifier, the output voltage changes in an


opposite direction to the input voltage.

As for the non-inverting amplifier, we start with the gain equation


of the op-amp:

This time, V– is a function of both Vout and Vin due to the voltage
divider formed by Rf and Rin. Again, the op-amp input does not
apply an appreciable load, so:

Substituting this into the gain equation and solving for Vout:

If AOL is very large, this simplifies to


.

A resistor is often inserted between the non-inverting input and


ground (so both inputs "see" similar resistances), reducing the
input offset voltage due to different voltage drops due to bias
current, and may reduce distortion in some op-amps.

A DC-blocking capacitor may be inserted in series with the input


resistor when a frequency response down to DC is not needed and
any DC voltage on the input is unwanted. That is, the capacitive
component of the input impedance inserts a DC zero and a low-
frequency pole that gives the circuit a bandpass or high-pass
characteristic.

Positive feedback configurations

Another typical configuration of op-amps is the positive


feedback, which takes a fraction of the output signal back to the
non-inverting input. An important application of it is the
comparator with hysteresis (i.e., the Schmitt trigger).

Other applications

 audio- and video-frequency pre-amplifiers and buffers

 voltage comparators

 differential amplifiers

 differentiators and integrators

 filters
 precision rectifiers

 precision peak detectors

 voltage and current regulators

 analog calculators

 analog-to-digital converters

 digital-to-analog converter

 voltage clamps

 oscillators and waveform generators

Most single, dual and quad op-amps available have a standardized


pin-out which permits one type to be substituted for another
without wiring changes. A specific op-amp may be chosen for its
open loop gain, bandwidth, noise performance, input impedance,
power consumption, or a compromise between any of these
factors.

You might also like