Modell Ndjusafnui9fsh7hfsadundfsa6tgfad

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 47

Modelling and Control Design for Variable Speed Wind

Turbine Energy System

Abstract – Renewable energy from wind is the safest form of energy. Wind turbine-based
energy generators have the potential to generate high amount of electric power if there is a
proper wind velocity and control mechanisms. This can certainly reduce the dependency on
solar photovoltaic based energy systems, which needs huge space to install the solar
photovoltaic panels. However, the output power of wind turbine is affected by the uncertain
wind velocity. The output mechanical power has to be properly controlled. Hence, the wind
energy system efficacy depends on how well this uncertainty is addressed. The major
challenge is to design and control the wind turbine systems that has a suitable mediator
between the power generator and the load, which counters the damage to the load due to
variable voltages produced by the varying wind velocity. Keeping this in view, this paper
implements all-important converter design methods for wind energy application and
recommends the most suitable method for its controller design. The overall analysis is
presented via detailed quantitative results that are evaluated with the help of time-domain
performance index parameters.
CHAPTER 1
INTRODUCTION
In the world energy point of view, there are two types of energy sources i.e.,
nonconventional (renewable) energy sources and conventional energy sources. Renewable
energy sources are the type of resources which are derived from earth or its ecology, and are
plenty in quantity. There are different types of renewable resources in general such as, wind
energy, hydro-energy, solar energy, biomass, geothermal energy, etc., which are
inexhaustible in nature. Apart from the abundancy, these resources are pollution free in usage
when compared to the conventional fossil fuel-based energy systems. So, presentday energy
sector is depending more and more usage of renewables rather than fast depleting
conventional fossil fuels. Many researches has been done on the generation of the electricity
from these renewable energy resources [1]. Among all these renewable resources, wind
energy generation is cleanest and more efficient source of energy. The main advantage of this
is, it doesn’t pollute the air, water, or soil in any form and other major advantage is it doesn’t
contribute to global warming. However, the major challenge in the wind turbine-based energy
system is to design and control the wind turbine systems such that there should be a suitable
interface between the power generator and the load such that no (or minimum) damage to the
load due to variable voltage produced by the variable wind speed. Normally wind turbines
operates in two different ways one with variable speed and other with fixed speed. Wind
speed is uncontrollable and changes within a very short period of time, which makes the wind
power control operations very complex. If the wind power generated cannot be controlled, it
damages the load. This can be solved by controlling the speed of wind turbine generator
which can be used to control the power generated over a wide variety of speeds. So, to resist
the damage to the load, proper design of controller unit is required to control the wind
turbine’s speed against all the uncertain wind velocities. Many researches are on-going in this
area to regulate the wind turbine speed, such as, speed control for direct drive permanent
magnet wind turbine [2], non-variable speed control of wind turbine [3], sensor less control
scheme for wind turbine in 1.5 MW doubly-fed induction generator (DFIG) application with
maximum power point tracking (MPPT) facility [4], full range speed control mechanism
given for an electromagnetic doubly-salient wind turbine plant [5], disturbance observer
based speed control mechanism for fixed-pitch angle based wind turbine [6], and modern or
intelligent speed control novel methods, such as wind velocity forecasting concept based
predictive control method for wind turbines [7], adaptive MPPT based rotor speed control of
DFIG wind turbine [8], intelligent speed control integrated with MPPT facility for medium
power wind energy systems [9], protection and control mechanism for full wind speed range
of a wind turbine power system [10], imperialist competitive algorithm based induction
motor’s speed control provision that was supplied by wind turbine [11], and model predictive
control approach for speed control of a wind turbine energy system [12]. However, all these
controllers that are implemented in literature are normally based on PID (proportional cum
integral cum derivative) controllers [13]. The fundamental issue with the PID controller is,
this can suitably work well for linear error variations, however, for nonlinear deviations of
the system response may not be settled with a PID controller.

Variable-Speed and Constant- Speed Wind Turbines

A major distinction in the wind turbines includes variable and constant speed wind turbines
i.e. the rotor is allowed to run at variable speed or constrained to operate at constant speed.
The constant speed wind turbines allow the use of simple generators whose speed is fixed by
the frequency of the electrical system. Although the power electronics needed for variable
speed wind turbines are more expensive, this type of turbines can spend more time operating
at maximum aerodynamic efficiency than constant speed turbines. This can be seen clearly if
the performance coefficient, Cp of a wind turbine is plotted against the tip speed ratio,  . The
tip speed ratio,  , is defined as the ratio between the speed of the tips of the blades of the
wind turbine and the speed of the wind.

The coefficient of performance, Cp , is defined as the fraction of energy extracted by the


wind turbine of the total energy that would have flowed through the area swept by the rotor if
the turbine had not been there.

The coefficient of performance Cp has a theoretical optimum of 0.59.Only a portion of the


power in the wind can be converted to useful energy by wind turbine. The power available
for a wind turbine is equal to the change in kinetic energy of the air as it passes through the
rotor. This maximum theoretical Cp was formulated in 1919 by Betz and applies to all types
of wind turbines. A typical Cp vs  characteristics is depicted in the figure.1.The figure.1
represents the characteristics of coefficient of performance, Cp versus tip speed ratio,  .From
figure.1 we infer that maximum value of Cp i.e. 0.5 is achieved for a tip speed ratio of 10 and
pitch angle of 0. For a fixed-speed wind turbine, where ω is constant, this corresponds to a
particular wind speed. For all other wind speeds the efficiency of the turbine is reduced. The
aim of variable-speed wind turbines is to always run at optimal efficiency, keeping constant
the particular λ that corresponds to the maximum Cp , by adapting the blades velocity to the
wind speed changes. Hence, variable speed wind turbines are designed to operate at optimum
energy efficiency, regardless of the wind speed. On the other hand, due to the fixed-speed
operation for constant speed turbines, all fluctuations in the wind speed are transmitted as
fluctuations in the mechanical torque and then as fluctuations in the electrical power grid.
This together with the increased energy capture obtained by using a variable-speed wind
turbine provides enough benefit to make the power electronics cost effective .Therefore, the
wind industry trend is to design and construct variable-speed wind turbines.

The most common definition is that renewable energy is from an energy resource that
is replaced by a natural process at a rate that is equal to or faster than the rate at which that
resource is being consumed. Renewable energy is a subset of sustainable energy.

India has done a significant progress in the power generation in the country. The
installed generation capacity was 2300 megawatt (MW) at the time of Independence i.e.
about 60 year’s back. Which includes the generation through various sectors like Hydro,
Thermal and Nuclear. The power generation in the country is planned through funds provided
by the Central Sector, State Sector and Private Sector. The power shortages noticed is of the
order of 22%. In the opinion of the experts such short fall can be reduced through proper
management and thus almost 20% energy can be saved. It has been noticed that one watt
saved at the point of consumption is more than 2.5 watts generated. In terms of Investment it
costs around Rs.20 million to generate one MW of new generation plant, but if the same
Rs.20 million is spent on conservation of energy methods, it can provide up to 3 MW of
avoidable generation capacity.

There are about 80,000 villages yet to be electrified for which provision has been
made to electrify 62,000 villages from grid supply in the Tenth Plan. It is planned that
participation of decentralized power producers shall be ensured, particularly for
electrification of remote villages in which village level organizations shall play a crucial role
for the rural electrification programme.

Emphasis is given to the renewable energy programme towards gradual


commercialization. This programme is looked after by the Ministry of Non-Conventional
Sources of energy. Simultaneously private sector investments in renewable energy sources
are also increased to promote power generation. So far an excessive reliance was preferred on
the use of fossil fuel resources like coal, oil and natural gas to meet the power requirement of
the country which was not suitable in the long run due to limited availability of the fossil fuel
as well as the adverse impact on the environment and ecology. Since the availability of fossil
fuel is on the decline therefore, in this backdrop the norms for conventional or renewable
sources of energy (RSE) is given importance not only in India but has attracted the global
attention.

Types of Renewable Energy Sources


The main types under RES are as follows
i) Hydro Power

ii) Solar Power

iii) Wind Power

iv) Bio-mass Power

v) Energy from waste

vi) Ocean energy

vii) Alternative fuel for surface transportation

Wind Power
The wind power development in the country is largely of recent period which has
been found to be quite impressive. As per available data, it is 5320 MW by March 32, 2006,
through wind power. Earlier it was estimated that the potential for wind power in the country
was 20,000 MW which has been revised to 25000MW after collecting the data on the
potential available in the coastal and other areas of the country. At present India is fifth in the
world after Germany, USA, Denmark and Spain in terms of wind power. It has been observed
that the private sector is showing interest in setting of wind power projects.

The unit size of wind turbine generators which were earlier in the range of 55-200 kw
are now preferred in the range of 750-2000 kw. It has been observed that the productivity of
the larger machine is higher as compared to the smaller machine. In respect of cost
consideration, it has been noticed that the cost of such a project is about Rs.20 million to
Rs.50 million per MW which includes all local civil, electrical works and erection also. The
life of a wind power project is estimated to be about 20 years.

China has guaranteed all certified renewable energy producers in its service area that
the grid will purchase their power and the price will be spread out to all the users across the
grid. According to sources, such commitments can only spur further development in the
renewable energy sector.
CHAPTER 2
WIND ENERGY CONVERSION SYSTEM
2.1 Introduction

Wind energy is transformed into mechanical energy by means of a wind turbine that
has one or several blades. The turbine is coupled to the generator system by means of a
mechanical drive train. It usually includes a gearbox that matches the turbine low speed to the
higher speed of the generator. New wind turbine designs use multi pole, low speed
generators, usually synchronous with field winding or permanent magnet excitation, in order
to eliminate the gearbox. Some turbines include a blade pitch angle control for controlling the
amount of power to be transformed. Stall controlled turbines do not allow such control. Wind
speed is measured by means of an anemometer. A general scheme of Wind energy
conversion system is shown in Fig. 2.1.

Fig. 3.1 Block Diagram of Wind Energy Conversion System


The electrical generator transforms mechanical energy from the wind turbine into electrical
energy. The generator can be synchronous or asynchronous. In the first case, an excitation
system is included or permanent magnets are used. Variable speed systems require the
presence of a power electronic interface, which can adapt to different configurations. The
compensating unit may include power factor correction devices (active or passive) and filters.
Types of wind turbines
Wind turbines can be separated into two types based on the axis about which the
turbine rotates. Turbines that rotate around a horizontal axis are most common where as
vertical-axis turbines are less frequently used
Horizontal- and Vertical-Axis Wind Turbines
Wind turbines can be categorized based on the orientation of their spin axis into horizontal-
axis wind turbines (HAWT) and vertical-axis wind turbines (VAWT).In horizontal-axis wind
turbines, the orientation of the spin axis is parallel to the. The tower elevates the nacelle to
provide sufficient space for the rotor blade rotation and to reach better wind conditions. The
nacelle supports the rotor hub that holds the rotor blades and also houses the gearbox,
generator and, in some designs, power converters. The industry standard HAWT uses a three
blade rotor positioned in front of the nacelle, which is known as upwind configuration.
However, downwind configurations with the blades at the back can also be found in practical
applications. Turbines with one, two or more than three blades can also be seen in wind
farms.
In vertical-axis wind turbines, the orientation of the spin axis is perpendicular to the
ground. The turbine rotor uses curved vertically mounted airfoils. The generator and gearbox
are normally placed in the base of the turbine on the ground. The rotor blades of the VAWT
have a variety of designs with different shapes and number of blades. The design given in the
figure is one of the popular designs. The VAWT normally needs guide wires to keep the rotor
shaft in a fixed position and minimize possible mechanical vibrations. The vertical axis
machine has the shape of an egg beater, and is often called the Darrieus rotor after its
inventor. However, most modern turbines use horizontal axis design. In this paper the
dynamic model of a horizontal axis turbine is developed and simulated in the
MATLAB/SIMULINK 2009a.
Mathematical Model of wind turbine
Wind power is the use of air flow through wind turbines to provide the mechanical
power to turn electric generators. Wind power, as an alternative to burning fossil fuels, is
plentiful, renewable, widely distributed, clean, produces no greenhouse gas emissions during
operation, consumes no water, and uses little land. The net effects on the environment are far
less problematic than those of fossil fuel sources.
Wind farms consist of many individual wind turbines, which are connected to
the electric power transmission network. Onshore wind is an inexpensive source of electric
power, competitive with or in many places cheaper than coal or gas plants. Offshore wind is
steadier and stronger than on land and offshore farms have less visual impact, but
construction and maintenance costs are considerably higher. Small onshore wind farms can
feed some energy into the grid or provide electric power to isolated off-grid locations. Wind
power gives variable power, which is very consistent from year to year but has significant
variation over shorter time scales. It is therefore used in conjunction with other electric power
sources to give a reliable supply. As the proportion of wind power in a region increases, a
need to upgrade the grid and a lowered ability to supplant conventional production can occur.
[9][10]
Power-management techniques such as having excess capacity, geographically
distributed turbines, dispatchable sources, sufficient hydroelectric power, exporting and
importing power to neighboring areas, energy storage, or reducing demand when wind
production is low, can in many cases overcome these problems. Weather forecasting permits
the electric-power network to be readied for the predictable variations in production that
occur.
In 2018, global wind power capacity expanded 9.6% to 591 GW.[16]. In 2017, yearly
wind energy production grew 17% reaching 4.4% of worldwide electric power usage, [17] and
providing 11.6% of the electricity in the European Union. [18] Denmark is the country with the
highest penetration of wind power, with 43.4% of its consumed electricity from wind in
2017. At least 83 other countries around the world are using wind power to supply

Charles Brush's windmill of 1888, used for generating electric power.


Wind power has been used as long as humans have put sails into the wind. For more
than two millennia wind-powered machines have ground grain and pumped water. Wind
power was widely available and not confined to the banks of fast-flowing streams, or later,
requiring sources of fuel. Wind-powered pumps drained the polders of the Netherlands, and
in arid regions such as the American mid-west or the Australian outback, wind
pumps provided water for livestock and steam engines.
The first windmill used for the production of electric power was built in Scotland in
July 1887 by Prof James Blyth of Anderson's College, Glasgow (the precursor of Strathclyde
University). Blyth's 10 metres (33 ft) high, cloth-sailed wind turbine was installed in the
garden of his holiday cottage at Marykirkin Kincardineshire and was used to
charge accumulators developed by the Frenchman Camille Alphonse Faure, to power the
lighting in the cottage, thus making it the first house in the world to have its electric power
supplied by wind power.[23] Blyth offered the surplus electric power to the people of Marykirk
for lighting the main street, however, they turned down the offer as they thought electric
power was "the work of the devil."[22] Although he later built a wind turbine to supply
emergency power to the local Lunatic Asylum, Infirmary and Dispensary of Montrose, the
invention never really caught on as the technology was not considered to be economically
viable.
Across the Atlantic, in Cleveland, Ohio, a larger and heavily engineered machine was
designed and constructed in the winter of 1887–1888 by Charles F. Brush,[24]. This was built
by his engineering company at his home and operated from 1886 until 1900.[25] The Brush
wind turbine had a rotor 17 metres (56 ft) in diameter and was mounted on an 18 metres
(59 ft) tower. Although large by today's standards, the machine was only rated at 12 kW. The
connected dynamo was used either to charge a bank of batteries or to operate up to
100 incandescent light bulbs, three arc lamps, and various motors in Brush's laboratory.
With the development of electric power, wind power found new applications in lighting
buildings remote from centrally-generated power. Throughout the 20th century parallel paths
developed small wind stations suitable for farms or residences, and larger utility-scale wind
generators that could be connected to electric power grids for remote use of power. Today
wind powered generators operate in every size range between tiny stations for battery
charging at isolated residences, up to near-gigawatt sized offshore wind farms that provide
electric power to national electrical networks.
Distribution of wind speed (red) and energy (blue) for all of 2002 at the Lee Ranch facility in
Colorado. The histogram shows measured data, while the curve is the Rayleigh model
distribution for the same average wind speed.
Wind energy is the kinetic energy of air in motion, also called wind. Total wind energy
flowing through an imaginary surface with area A during the time t is:
where ρ is the density of air; v is the wind speed; Avt is the volume of air passing
through A (which is considered perpendicular to the direction of the wind); Avtρ is therefore
the mass m passing through "A". Note that ½ ρv2 is the kinetic energy of the moving air per
unit volume.
Power is energy per unit time, so the wind power incident on A (e.g. equal to the rotor area of
a wind turbine) is:
Wind power in an open air stream is thus proportional to the third power of the wind speed;
the available power increases eightfold when the wind speed doubles. Wind turbines for grid
electric power therefore need to be especially efficient at greater wind speeds.
Wind is the movement of air across the surface of the Earth, affected by areas of high
pressure and of low pressure. The global wind kinetic energy averaged approximately 1.50
MJ/m2 over the period from 1979 to 2010, 1.31 MJ/m 2 in the Northern Hemisphere with 1.70
MJ/m2 in the Southern Hemisphere. The atmosphere acts as a thermal engine, absorbing heat
at higher temperatures, releasing heat at lower temperatures. The process is responsible for
production of wind kinetic energy at a rate of 2.46 W/m2 sustaining thus the circulation of the
atmosphere against frictional dissipation.[30]
Through wind resource assessment it is possible to provide estimates of wind power potential
globally, by country or region, or for a specific site. A global assessment of wind power
potential is available via the Global Wind Atlas provided by the Technical University of
Denmark in partnership with the World Bank.[27] [31][32] Unlike 'static' wind resource atlases
which average estimates of wind speed and power density across multiple years, tools such
as Renewables.ninja provide time-varying simulations of wind speed and power output from
different wind turbine models at an hourly resolution. [33]. More detailed, site specific
assessments of wind resource potential can be obtained from specialist commercial providers,
and many of the larger wind developers will maintain in-house modeling capabilities.
The total amount of economically extractable power available from the wind is considerably
more than present human power use from all sources.[34] Axel Kleidon of the Max Planck
Institute in Germany, carried out a "top down" calculation on how much wind energy there is,
starting with the incoming solar radiation that drives the winds by creating temperature
differences in the atmosphere. He concluded that somewhere between 18 TW and 68 TW
could be extracted.
Cristina Archer and Mark Z. Jacobson presented a "bottom-up" estimate, which unlike
Kleidon's are based on actual measurements of wind speeds, and found that there is 1700 TW
of wind power at an altitude of 100 metres over land and sea. Of this, "between 72 and 170
TW could be extracted in a practical and cost-competitive manner". [35] They later estimated
80 TW.[36] However research at Harvard University estimates 1 Watt/m2 on average and 2–10
MW/km2 capacity for large scale wind farms, suggesting that these estimates of total global
wind resources are too high by a factor of about 4.[37]
The strength of wind varies, and an average value for a given location does not alone indicate
the amount of energy a wind turbine could produce there.
To assess prospective wind power sites a probability distribution function is often fit to the
observed wind speed data. Different locations will have different wind speed distributions.
The Weibull model closely mirrors the actual distribution of hourly/ten-minute wind speeds
at many locations. The Weibull factor is often close to 2 and therefore a Rayleigh
distribution can be used as a less accurate, but simpler model.[39]
A wind farm is a group of wind turbines in the same location used for production of electric
power. A large wind farm may consist of several hundred individual wind turbines distributed
over an extended area, but the land between the turbines may be used for agricultural or other
purposes. For example, Gansu Wind Farm, the largest wind farm in the world, has several
thousand turbines. A wind farm may also be located offshore.
Almost all large wind turbines have the same design — a horizontal axis wind turbine having
an upwind rotor with three blades, attached to a nacelle on top of a tall tubular tower.
In a wind farm, individual turbines are interconnected with a medium voltage (often 34.5
kV), power collection system and communications network. In general, a distance of 7D (7 ×
Rotor Diameter of the Wind Turbine) is set between each turbine in a fully developed wind
farm.[51] At a substation, this medium-voltage electric current is increased in voltage with
a transformer for connection to the high voltage electric power transmission system.
Generator characteristics and stability
A permanent magnet synchronous generator is a generator where the excitation field is
provided by a permanent magnet instead of a coil. The term synchronous refers here to the
fact that the rotor and magnetic field rotate with the same speed, because the magnetic field is
generated through a shaft mounted permanent magnet mechanism and current is induced into
the stationary armature.
Synchronous generators are the majority source of commercial electrical energy. They are
commonly used to convert the mechanical power output of steam turbines, gas
turbines, reciprocating engines and hydro turbines into electrical power for the grid. Some
designs ofWind turbines also use this generator type.
In the majority of designs the rotating assembly in the center of the generator—the "rotor"—
contains the magnet, and the "stator" is the stationary armature that is electrically connected
to a load. As shown in the diagram, the perpendicular component of the stator field affects the
torque while the parallel component affects the voltage. The load supplied by the generator
determines the voltage. If the load is inductive, then the angle between the rotor and stator
fields will be greater than 90 degrees which corresponds to an increased generator voltage.
This is known as an overexcited generator. The opposite is true for a generator supplying a
capacitive load which is known as an underexcited generator. A set of three conductors make
up the armature winding in standard utility equipment, constituting three phases of a power
circuit—that correspond to the three wires we are accustomed to see on transmission lines.
The phases are wound such that they are 120 degrees apart spatially on the stator, providing
for a uniform force or torque on the generator rotor. The uniformity of the torque arises
because the magnetic fields resulting from the induced currents in the three conductors of the
armature winding combine spatially in such a way as to resemble the magnetic field of a
single, rotating magnet. This stator magnetic field or "stator field" appears as a steady
rotating field and spins at the same frequency as the rotor when the rotor contains a single
dipole magnetic field. The two fields move in "synchronicity" and maintain a fixed position
relative to each other as they spin.
Synchronous
They are known as synchronous generators because f, the frequency of the induced voltage in
the stator (armature conductors) conventionally measured in hertz, is directly proportional to
RPM, the rotation rate of the rotor usually given in revolutions per minute (or angular speed).
If the rotor windings are arranged in such a way as to produce the effect of more than two
magnetic poles, then each physical revolution of the rotor results in more magnetic poles
moving past the armature windings. Each passing of a north and south pole corresponds to a
complete "cycle" of a magnet field oscillation. Therefore, the constant of proportionality is ,
where P is the number of magnetic rotor poles (almost always an even number), and the
factor of 120 comes from 60 seconds per minute and two poles in a
In a permanent magnet generator, the magnetic field of the rotor is produced by
permanent magnets. Other types of generator use electromagnets to produce a magnetic field
in a rotor winding. The direct current in the rotor field winding is fed through a slip-ring
assembly or provided by a brushless exciter on the same shaft.
Permanent magnet generators (PMGs) or alternators (PMAs) do not require a DC supply for
the excitation circuit, nor do they have slip rings and contact brushes. A key disadvantage in
PMAs or PMGs is that the air gap flux is not controllable, so the voltage of the machine
cannot be easily regulated. A persistent magnetic field imposes safety issues during assembly,
field service or repair. High performance permanent magnets, themselves, have structural and
thermal issues. Torque current MMF vectorially combines with the persistent flux of
permanent magnets, which leads to higher air-gap flux density and eventually, core
saturation. In this permanent magnet alternators the speed is directly proportional to the
output voltage of the alternator.
Induction generators, which were often used for wind power projects in the 1980s and 1990s,
require reactive power for excitation, so substations used in wind-power collection systems
include substantial capacitor banks for power factor correction. Different types of wind
turbine generators behave differently during transmission grid disturbances, so extensive
modelling of the dynamic electromechanical characteristics of a new wind farm is required
by transmission system operators to ensure predictable stable behaviour during system faults.
In particular, induction generators cannot support the system voltage during faults, unlike
steam or hydro turbine-driven synchronous generators.
Induction generators aren't used in current turbines. Instead, most turbines use variable speed
generators combined with partial- or full-scale power converter between the turbine generator
and the collector system, which generally have more desirable properties for grid
interconnection and have Low voltage ride through-capabilities.[52] Modern concepts use
either doubly fed machines with partial-scale converters or squirrel-cage induction generators
or synchronous generators (both permanently and electrically excited) with full scale
converters.
Transmission systems operators will supply a wind farm developer with a grid code to
specify the requirements for interconnection to the transmission grid. This will include power
factor, constancy of frequency and dynamic behaviour of the wind farm turbines during a
system fault.
The wind power industry set new records in 2014 – more than 50 GW of new capacity was
installed. Another record breaking year occurred in 2015, with 22% annual market growth
resulting in the 60 GW mark being passed. [80] In 2015, close to half of all new wind power
was added outside of the traditional markets in Europe and North America. This was largely
from new construction in China and India. Global Wind Energy Council (GWEC) figures
show that 2015 recorded an increase of installed capacity of more than 63 GW, taking the
total installed wind energy capacity to 432.9 GW, up from 74 GW in 2006. In terms of
economic value, the wind energy sector has become one of the important players in the
energy markets, with the total investments reaching US$329bn (€296.6bn), an increase of 4%
over 2014.
Although the wind power industry was affected by the global financial crisis in 2009 and
2010, GWEC predicts that the installed capacity of wind power will be 792.1 GW by the end
of 2020 and 4,042 GW by end of 2050.[82] The increased commissioning of wind power is
being accompanied by record low prices for forthcoming renewable electric power. In some
cases, wind onshore is already the cheapest electric power generation option and costs are
continuing to decline. The contracted prices for wind onshore for the next few years are now
as low as 30 USD/MWh.
In the EU in 2015, 44% of all new generating capacity was wind power; while in the same
period net fossil fuel power capacity decreased.
Capacity factor
Since wind speed is not constant, a wind farm's annual energy production is never as much as
the sum of the generator nameplate ratings multiplied by the total hours in a year. The ratio of
actual productivity in a year to this theoretical maximum is called the capacity factor. Typical
capacity factors are 15–50%; values at the upper end of the range are achieved in favourable
sites and are due to wind turbine design improvements.
Online data is available for some locations, and the capacity factor can be calculated from the
yearly output. For example, the German nationwide average wind power capacity factor over
all of 2012 was just under 17.5% (45,867 GW·h/yr / (29.9 GW × 24 × 366) = 0.1746), and
the capacity factor for Scottish wind farms averaged 24% between 2008 and 2010.
Unlike fueled generating plants, the capacity factor is affected by several parameters,
including the variability of the wind at the site and the size of the generator relative to the
turbine's swept area. A small generator would be cheaper and achieve a higher capacity factor
but would produce less electric power (and thus less profit) in high winds. Conversely, a
large generator would cost more but generate little extra power and, depending on the type,
may stall out at low wind speed. Thus an optimum capacity factor of around 40–50% would
be aimed for.
A 2008 study released by the U.S. Department of Energy noted that the capacity factor of
new wind installations was increasing as the technology improves, and projected further
improvements for future capacity factors. In 2010, the department estimated the capacity
factor of new wind turbines in 2010 to be 45%. The annual average capacity factor for wind
generation in the US has varied between 29.8% and 34% during the period 2010–2015.
Wind energy penetration is the fraction of energy produced by wind compared with the total
generation. The wind power penetration in world electric power generation in 2015 was
3.5%.
There is no generally accepted maximum level of wind penetration. The limit for a
particular grid will depend on the existing generating plants, pricing mechanisms, capacity
for energy storage, demand management and other factors. An interconnected electric power
grid will already include reserve generating and transmission capacity to allow for equipment
failures. This reserve capacity can also serve to compensate for the varying power generation
produced by wind stations. Studies have indicated that 20% of the total annual electrical
energy consumption may be incorporated with minimal difficulty. [104] These studies have
been for locations with geographically dispersed wind farms, some degree of dispatch able
energy or hydropower with storage capacity, demand management, and interconnected to a
large grid area enabling the export of electric power when needed. Beyond the 20% level,
there are few technical limits, but the economic implications become more significant.
Electrical utilities continue to study the effects of large scale penetration of wind generation
on system stability and economics.
A wind energy penetration figure can be specified for different duration of time, but is often
quoted annually. To obtain 100% from wind annually requires substantial long term storage
or substantial interconnection to other systems which may already have substantial storage.
On a monthly, weekly, daily, or hourly basis—or less—wind might supply as much as or
more than 100% of current use, with the rest stored or exported. Seasonal industry might then
take advantage of high wind and low usage times such as at night when wind output can
exceed normal demand. Such industry might include production of silicon, aluminum, steel,
or of natural gas, and hydrogen, and using future long term storage to facilitate 100% energy
from variable renewable energy.[109][110] Homes can also be programmed to accept extra
electric power on demand, for example by remotely turning up water heater thermostats.
In Australia, the state of South Australia generates around half of the nation's wind power
capacity. By the end of 2011 wind power in South Australia, championed by Premier (and
Climate Change Minister) Mike Rann, reached 26% of the State's electric power generation,
edging out coal for the first time. At this stage South Australia, with only 7.2% of Australia's
population, had 54% of Australia's installed capacity.
Typically, conventional hydroelectricity complements wind power very well. When the wind
is blowing strongly, nearby hydroelectric stations can temporarily hold back their water.
When the wind drops they can, provided they have the generation capacity, rapidly increase
production to compensate. This gives a very even overall power supply and virtually no loss
of energy and uses no more water.
Alternatively, where a suitable head of water is not available, pumped-storage
hydroelectricity or other forms of grid energy storage such as compressed air energy
storage and thermal energy storage can store energy developed by high-wind periods and
release it when needed. The type of storage needed depends on the wind penetration level –
low penetration requires daily storage, and high penetration requires both short and long term
storage – as long as a month or more. Stored energy increases the economic value of wind
energy since it can be shifted to displace higher cost generation during peak demand periods.
The potential revenue from this arbitrage can offset the cost and losses of storage. For
example, in the UK, the 1.7 GW Dinorwig pumped-storage plant evens out electrical demand
peaks, and allows base-load suppliers to run their plants more efficiently. Although pumped-
storage power systems are only about 75% efficient, and have high installation costs, their
low running costs and ability to reduce the required electrical base-load can save both fuel
and total electrical generation costs.
In particular geographic regions, peak wind speeds may not coincide with peak demand for
electrical power. In the U.S. states of California and Texas, for example, hot days in summer
may have low wind speed and high electrical demand due to the use of air conditioning.
Some utilities subsidize the purchase of geothermal heat pumps by their customers, to reduce
electric power demand during the summer months by making air conditioning up to 70%
more efficient;[134]widespread adoption of this technology would better match electric power
demand to wind availability in areas with hot summers and low summer winds. A possible
future option may be to interconnect widely dispersed geographic areas with an HVDC
"super grid". In the U.S. it is estimated that to upgrade the transmission system to take in
planned or potential renewables would cost at least USD 60 bn, [135]while the society value of
added windpower would be more than that cost.[136]
Germany has an installed capacity of wind and solar that can exceed daily demand, and has
been exporting peak power to neighboring countries, with exports which amounted to some
14.7 billion kWh in 2012. A more practical solution is the installation of thirty days storage
capacity able to supply 80% of demand, which will become necessary when most of Europe's
energy is obtained from wind power and solar power. Just as the EU requires member
countries to maintain 90 days strategic reserves of oil it can be expected that countries will
provide electric power storage, instead of expecting to use their neighbors for net metering.
Capacity credit, fuel savings and energy payback[edit]
The capacity credit of wind is estimated by determining the capacity of conventional plants
displaced by wind power, whilst maintaining the same degree of system security. According
to the American Wind Energy Association, production of wind power in the United States in
2015 avoided consumption of 73 billion gallons of water and reduced CO
2 emissions by 132 million metric tons, while providing USD 7.3 bn in public health savings.
The energy needed to build a wind farm divided into the total output over its life, Energy
Return on Energy Invested, of wind power varies but averages about 20–25.[143][144] Thus, the
energy payback time is typically around a year.
Wind turbines reached grid parity (the point at which the cost of wind power matches
traditional sources) in some areas of Europe in the mid-2000s, and in the US around the same
time. Falling prices continue to drive the levelized cost down and it has been suggested that it
has reached general grid parity in Europe in 2010, and will reach the same point in the US
around 2016 due to an expected reduction in capital costs of about 12%.
CHAPTER 3

DC/AC CONVERTERS

DC to AC converters produce an AC output waveform from a DC source.


Applications include adjustable speed drives (ASD), uninterruptable power supplies (UPS),
active filters, Flexible AC transmission systems (FACTS), voltage compensators, and
photovoltaic generators. Topologies for these converters can be separated into two distinct
categories: voltage source inverters and current source inverters. Voltage source inverters
(VSIs) are named so because the independently controlled output is a voltage waveform.
Similarly, current source inverters (CSIs) are distinct in that the controlled AC output is a
current waveform.

Being static power converters, the DC to AC power conversion is the result of power
switching devices, which are commonly fully controllable semiconductor power switches.
The output waveforms are therefore made up of discrete values, producing fast transitions
rather than smooth ones. The ability to produce near sinusoidal waveforms around the
fundamental frequency is dictated by the modulation technique controlling when, and for
how long, the power valves are on and off. Common modulation techniques include the
carrier-based technique, or pulse width modulation, space-vector technique, and the selective-
harmonic technique.[8]

Voltage source inverters have practical uses in both single-phase and three-phase
applications. Single-phase VSIs utilize half-bridge and full-bridge configurations, and are
widely used for power supplies, single-phase UPSs, and elaborate high-power topologies
when used in multicell configurations. Three-phase VSIs are used in applications that require
sinusoidal voltage waveforms, such as ASDs, UPSs, and some types of FACTS devices such
as the STATCOM. They are also used in applications where arbitrary voltages are required as
in the case of active filters and voltage compensators.[8]

Current source inverters are used to produce an AC output current from a DC current
supply. This type of inverter is practical for three-phase applications in which high-quality
voltage waveforms are required.

A relatively new class of inverters, called multilevel inverters, has gained widespread
interest. Normal operation of CSIs and VSIs can be classified as two-level inverters, due to
the fact that power switches connect to either the positive or to the negative DC bus. If more
than two voltage levels were available to the inverter output terminals, the AC output could
better approximate a sine wave. It is for this reason that multilevel inverters, although more
complex and costly, offer higher performance.[9]

Each inverter type differs in the DC links used, and in whether or not they require
freewheeling diodes. Either can be made to operate in square-wave or pulse-width
modulation (PWM) mode, depending on its intended usage. Square-wave mode offers
simplicity, while PWM can be implemented several different ways and produces higher
quality waveforms.[8]

Voltage Source Inverters (VSI) feed the output inverter section from an
approximately constant-voltage source.[8]

The desired quality of the current output waveform determines which modulation
technique needs to be selected for a given application. The output of a VSI is composed of
discrete values. In order to obtain a smooth current waveform, the loads need to be inductive
at the select harmonic frequencies. Without some sort of inductive filtering between the
source and load, a capacitive load will cause the load to receive a choppy current waveform,
with large and frequent current spikes.[8]

There are three main types of VSIs:

 Single-phase half-bridge inverter


 Single-phase full-bridge inverter
 Three-phase voltage source inverter

Single-phase half-bridge inverter


Figure 1: The AC input for an ASD.

FIGURE 2: Single-Phase Half-Bridge Voltage Source Inverter


The single-phase voltage source half-bridge inverters, are meant for lower voltage
applications and are commonly used in power supplies.[8] Figure 2 shows the circuit
schematic of this inverter.

Low-order current harmonics get injected back to the source voltage by the operation
of the inverter. This means that two large capacitors are needed for filtering purposes in this
design.[8] As Figure 2 illustrates, only one switch can be on at time in each leg of the inverter.
If both switches in a leg were on at the same time, the DC source will be shorted out.

Inverters can use several modulation techniques to control their switching schemes.
The carrier-based PWM technique compares the AC output waveform, v c, to a carrier voltage
signal, vΔ. When vc is greater than vΔ, S+ is on, and when vc is less than vΔ, S- is on. When the
AC output is at frequency fc with its amplitude at v c, and the triangular carrier signal is at
frequency fΔ with its amplitude at vΔ, the PWM becomes a special sinusoidal case of the
carrier based PWM.[8] This case is dubbed sinusoidal pulse-width modulation (SPWM).For
this, the modulation index, or amplitude-modulation ratio, is defined asma = vc / v∆ .

The normalized carrier frequency, or frequency-modulation ratio, is calculated using


the equation mf = f∆ / fc .

If the over-modulation region, ma, exceeds one, a higher fundamental AC output voltage will
be observed, but at the cost of saturation. For SPWM, the harmonics of the output waveform
are at well-defined frequencies and amplitudes. This simplifies the design of the filtering
components needed for the low-order current harmonic injection from the operation of the
inverter. The maximum output amplitude in this mode of operation is half of the source
voltage. If the maximum output amplitude, ma, exceeds 3.24, the output waveform of the
inverter becomes a square wave.[8]

As was true for PWM, both switches in a leg for square wave modulation cannot be
turned on at the same time, as this would cause a short across the voltage source. The
switching scheme requires that both S+ and S- be on for a half cycle of the AC output period.
[8]
The fundamental AC output amplitude is equal to vo1 = vaN.

Therefore, the AC output voltage is not controlled by the inverter, but rather by the
magnitude of the DC input voltage of the inverter.[8]

Using selective harmonic elimination (SHE) as a modulation technique allows the


switching of the inverter to selectively eliminate intrinsic harmonics. The fundamental
component of the AC output voltage can also be adjusted within a desirable range. Since the
AC output voltage obtained from this modulation technique has odd half and odd quarter
wave symmetry, even harmonics do not exist. [8] Any undesirable odd (N-1) intrinsic
harmonics from the output waveform can be eliminated.

Single-phase full-bridge inverter

FIGURE 3: Single-Phase Voltage Source Full-Bridge Inverter


FIGURE 4: Carrier and Modulating Signals for the Bipolar Pulsewidth Modulation
Technique
The full-bridge inverter is similar to the half bridge-inverter, but it has an additional
leg to connect the neutral point to the load. [8] Figure 3 shows the circuit schematic of the
single-phase voltage source full-bridge inverter.

To avoid shorting out the voltage source, S1+ and S1- cannot be on at the same time,
and S2+ and S2- also cannot be on at the same time. Any modulating technique used for the
full-bridge configuration should have either the top or the bottom switch of each leg on at any
given time. Due to the extra leg, the maximum amplitude of the output waveform is Vi, and is
twice as large as the maximum achievable output amplitude for the half-bridge configuration.
[8]

States 1 and 2 from Table 2 are used to generate the AC output voltage with bipolar
SPWM. The AC output voltage can take on only two values, either Vi or –Vi. To generate
these same states using a half-bridge configuration, a carrier based technique can be used. S+
being on for the half-bridge corresponds to S1+ and S2- being on for the full-bridge.
Similarly, S- being on for the half-bridge corresponds to S1- and S2+ being on for the full
bridge. The output voltage for this modulation technique is more or less sinusoidal, with a
fundamental component that has an amplitude in the linear region of ma less than or equal to
one[8] vo1 =vab1= vi • ma.

Unlike the bipolar PWM technique, the unipolar approach uses states 1, 2, 3 and 4
from Table 2 to generate its AC output voltage. Therefore, the AC output voltage can take on
the values Vi, 0 or –V [1]i. To generate these states, two sinusoidal modulating signals, Vc
and –Vc, are needed, as seen in Figure 4.

Vc is used to generate VaN, while –Vc is used to generate VbN. The following
relationship is called unipolar carrier-based SPWM vo1 =2 • vaN1= vi • ma.
The phase voltages VaN and VbN are identical, but 180 degrees out of phase with
each other. The output voltage is equal to the difference of the two phase voltages, and do not
contain any even harmonics. Therefore, if mf is taken, even the AC output voltage harmonics
will appear at normalized odd frequencies, fh. These frequencies are centered on double the
value of the normalized carrier frequency. This particular feature allows for smaller filtering
components when trying to obtain a higher quality output waveform.[8]

As was the case for the half-bridge SHE, the AC output voltage contains no even
harmonics due to its odd half and odd quarter wave symmetry.[8]

Three-phase voltage source inverter

FIGURE 5: Three-Phase Voltage Source Inverter Circuit Schematic

Single-phase VSIs are used primarily for low power range applications, while three-
phase VSIs cover both medium and high power range applications. [8] Figure 5 shows the
circuit schematic for a three-phase VSI.

Switches in any of the three legs of the inverter cannot be switched off simultaneously
due to this resulting in the voltages being dependent on the respective line current's polarity.
States 7 and 8 produce zero AC line voltages, which result in AC line currents freewheeling
through either the upper or the lower components. However, the line voltages for states 1
through 6 produce an AC line voltage consisting of the discrete values of Vi, 0 or –Vi.[8]

For three-phase SPWM, three modulating signals that are 120 degrees out of phase
with one another are used in order to produce out of phase load voltages. In order to preserve
the PWM features with a single carrier signal, the normalized carrier frequency, mf, needs to
be a multiple of three. This keeps the magnitude of the phase voltages identical, but out of
phase with each other by 120 degrees.[8] The maximum achievable phase voltage amplitude in
the linear region, ma less than or equal to one, is vphase = vi / 2. The maximum achievable line
voltage amplitude is Vab1 = vab • √3 / 2

The only way to control the load voltage is by changing the input DC voltage.

Current source inverters

FIGURE 7: Three-Phase Current Source Inverter

Figure 8: Synchronized-Pulse-Width-Modulation Waveforms for a Three-Phase Current


Source Inverter a) Carrier and Modulating Signals b) S1 State c) S3 State d) Output Current
Figure 9: Space-Vector Representation in Current Source Inverters

Current source inverters convert DC current into an AC current waveform. In


applications requiring sinusoidal AC waveforms, magnitude, frequency, and phase should all
be controlled. CSIs have high changes in current overtime, so capacitors are commonly
employed on the AC side, while inductors are commonly employed on the DC side.[8] Due to
the absence of freewheeling diodes, the power circuit is reduced in size and weight, and tends
to be more reliable than VSIs.[9] Although single-phase topologies are possible, three-phase
CSIs are more practical.

In its most generalized form, a three-phase CSI employs the same conduction sequence as a
six-pulse rectifier. At any time, only one common-cathode switch and one common-anode
switch are on.[9]

As a result, line currents take discrete values of –ii, 0 and ii. States are chosen such
that a desired waveform is outputted and only valid states are used. This selection is based on
modulating techniques, which include carrier-based PWM, selective harmonic elimination,
and space-vector techniques.[8]

Carrier-based techniques used for VSIs can also be implemented for CSIs, resulting in
CSI line currents that behave in the same way as VSI line voltages. The digital circuit utilized
for modulating signals contains a switching pulse generator, a shorting pulse generator, a
shorting pulse distributor, and a switching and shorting pulse combiner. A gating signal is
produced based on a carrier current and three modulating signals.[8]

A shorting pulse is added to this signal when no top switches and no bottom switches
are gated, causing the RMS currents to be equal in all legs. The same methods are utilized for
each phase, however, switching variables are 120 degrees out of phase relative to one
another, and the current pulses are shifted by a half-cycle with respect to output currents. If a
triangular carrier is used with sinusoidal modulating signals, the CSI is said to be utilizing
synchronized-pulse-width-modulation (SPWM). If full over-modulation is used in
conjunction with SPWM the inverter is said to be in square-wave operation.[8]

The second CSI modulation category, SHE is also similar to its VSI counterpart.
Utilizing the gating signals developed for a VSI and a set of synchronizing sinusoidal current
signals, results in symmetrically distributed shorting pulses and, therefore, symmetrical
gating patterns. This allows any arbitrary number of harmonics to be eliminated. [8] It also
allows control of the fundamental line current through the proper selection of primary
switching angles. Optimal switching patterns must have quarter-wave and half-wave
symmetry, as well as symmetry about 30 degrees and 150 degrees. Switching patterns are
never allowed between 60 degrees and 120 degrees. The current ripple can be further reduced
with the use of larger output capacitors, or by increasing the number of switching pulses.[9]

The third category, space-vector-based modulation, generates PWM load line currents
that equal load line currents, on average. Valid switching states and time selections are made
digitally based on space vector transformation. Modulating signals are represented as a
complex vector using a transformation equation. For balanced three-phase sinusoidal signals,
this vector becomes a fixed module, which rotates at a frequency, ω. These space vectors are
then used to approximate the modulating signal. If the signal is between arbitrary vectors, the
vectors are combined with the zero vectors I7, I8, or I9The following equations are used to
ensure that the generated currents and the current vectors are on average equivalent.

Multilevel inverters

FIGURE 10: Three-Level Neutral-Clamped Inverter


A relatively new class called multilevel inverters has gained widespread interest.
Normal operation of CSIs and VSIs can be classified as two-level inverters because the
power switches connect to either the positive or the negative DC bus. If more than two
voltage levels were available to the inverter output terminals, the AC output could better
approximate a sine wave.[8] For this reason multilevel inverters, although more complex and
costly, offer higher performance.[9] A three-level neutral-clamped inverter is shown in Figure
10.

Control methods for a three-level inverter only allow two switches of the four
switches in each leg to simultaneously change conduction states. This allows smooth
commutation and avoids shoot through by only selecting valid states. It may also be noted
that since the DC bus voltage is shared by at least two power valves, their voltage ratings can
be less than a two-level counterpart.

Carrier-based and space-vector modulation techniques are used for multilevel


topologies. The methods for these techniques follow those of classic inverters, but with added
complexity. Space-vector modulation offers a greater number of fixed voltage vectors to be
used in approximating the modulation signal, and therefore allows more effective space
vector PWM strategies to be accomplished at the cost of more elaborate algorithms. Due to
added complexity and number of semiconductor devices, multilevel inverters are currently
more suitable for high-power high-voltage applications. This technology reduces the
harmonics hence improves overall efficiency of the scheme.
CHAPTER 4
AC DC CONVERTERS
While this method may be suitable for low power applications it is unsuitable to applications
which need a “steady and smooth” DC supply voltage. One method to improve on this is to
use every half-cycle of the input voltage instead of every other half-cycle. The circuit which
allows us to do this is called a Full Wave Rectifier.

Like the half wave circuit, a full wave rectifier circuit produces an output voltage or current
which is purely DC or has some specified DC component. Full wave rectifiers have some
fundamental advantages over their half wave rectifier counterparts. The average (DC) output
voltage is higher than for half wave, the output of the full wave rectifier has much less ripple
than that of the half wave rectifier producing a smoother output waveform.

In a Full Wave Rectifier circuit two diodes are now used, one for each half of the cycle. A
multiple winding transformer is used whose secondary winding is split equally into two
halves with a common centre tapped connection, (C). This configuration results in each diode
conducting in turn when its anode terminal is positive with respect to the transformer centre
point C producing an output during both half-cycles, twice that for the half wave rectifier so
it is 100% efficient as shown below.
The Diode Bridge Rectifier

The full wave rectifier circuit consists of two power diodes connected to a single load
resistance (RL) with each diode taking it in turn to supply current to the load. When
point A of the transformer is positive with respect to point C, diode D1 conducts in the
forward direction as indicated by the arrows.

When point B is positive (in the negative half of the cycle) with respect to point C,
diode D2 conducts in the forward direction and the current flowing through resistor R is in the
same direction for both half-cycles. As the output voltage across the resistor R is the phasor
sum of the two waveforms combined, this type of full wave rectifier circuit is also known as a
“bi-phase” circuit.

Another type of circuit that produces the same output waveform as the full wave rectifier
circuit above, is that of the Full Wave Bridge Rectifier. This type of single phase rectifier
uses four individual rectifying diodes connected in a closed loop “bridge” configuration to
produce the desired output.

The main advantage of this bridge circuit is that it does not require a special centre tapped
transformer, thereby reducing its size and cost. The single secondary winding is connected to
one side of the diode bridge network and the load to the other side as shown below.

The four diodes labelled D1 to D4 are arranged in “series pairs” with only two diodes
conducting current during each half cycle. During the positive half cycle of the supply,
diodes D1 and D2 conduct in series while diodes D3 and D4 are reverse biased and the
current flows through the load as shown below.

As the current flowing through the load is unidirectional, so the voltage developed across the
load is also unidirectional the same as for the previous two diode full-wave rectifier, therefore
the average DC voltage across the load is 0.637Vmax.

However in reality, during each half cycle the current flows through two diodes instead of
just one so the amplitude of the output voltage is two voltage drops ( 2*0.7 = 1.4V ) less than
the input VMAX amplitude. The ripple frequency is now twice the supply frequency (e.g.
100Hz for a 50Hz supply or 120Hz for a 60Hz supply.)

Although we can use four individual power diodes to make a full wave bridge rectifier, pre-
made bridge rectifier components are available “off-the-shelf” in a range of different voltage
and current sizes that can be soldered directly into a PCB circuit board or be connected by
spade connectors.

The image to the right shows a typical single phase bridge rectifier with one corner cut off.
This cut-off corner indicates that the terminal nearest to the corner is the positive
or +ve output terminal or lead with the opposite (diagonal) lead being the negative or -
ve output lead. The other two connecting leads are for the input alternating voltage from a
transformer secondary winding.

We saw in the previous section that the single phase half-wave rectifier produces an output
wave every half cycle and that it was not practical to use this type of circuit to produce a
steady DC supply. The full-wave bridge rectifier however, gives us a greater mean DC value
(0.637 Vmax) with less superimposed ripple while the output waveform is twice that of the
frequency of the input supply frequency.

We can improve the average DC output of the rectifier while at the same time reducing the
AC variation of the rectified output by using smoothing capacitors to filter the output
waveform. Smoothing or reservoir capacitors connected in parallel with the load across the
output of the full wave bridge rectifier circuit increases the average DC output level even
higher as the capacitor acts like a storage device as shown below.
CHAPTER 5
BUCK CONVERTER
INTRODUCTION
A buck converter (step-down converter) is a DC-to-DC power converter which steps
down voltage (while stepping up current) from its input (supply) to its output (load). It is a
class of switched-mode power supply (SMPS) typically containing at least two
semiconductors (a diode and a transistor, although modern buck converters frequently replace
the diode with a second transistor used for synchronous rectification) and at least one energy
storage element, a capacitor, inductor, or the two in combination. To reduce voltage ripple,
filters made of capacitors (sometimes in combination with inductors) are normally added to
such a converter's output (load-side filter) and input (supply-side filter).

Fig. 1: Buck converter circuit diagram.


3 A buck converter
Switching converters (such as buck converters) provide much greater power efficiency as
DC-to-DC converters than linear regulators, which are simpler circuits that lower voltages by
dissipating power as heat, but do not step up output current.[2]
Buck converters can be highly efficient (often higher than 90%), making them useful for
tasks such as converting a computer's main (bulk) supply voltage (often 12 V) down to lower
voltages needed by USB, DRAM and the CPU (1.8 V or less).

Theory of operation

Fig. 2: The two circuit configurations of a buck converter: on-state, when the switch is
closed; and off-state, when the switch is open (arrows indicate current according to the
direction conventional current model).

Fig. 3: Naming conventions of the components, voltages and current of the buck converter.
Fig. 4: Evolution of the voltages and currents with time in an ideal buck converter operating
in continuous mode.
The basic operation of the buck converter has the current in an inductor controlled by
two switches (usually a transistor and a diode). In the idealised converter, all the components
are considered to be perfect. Specifically, the switch and the diode have zero voltage drop
when on and zero current flow when off, and the inductor has zero series resistance. Further,
it is assumed that the input and output voltages do not change over the course of a cycle (this
would imply the output capacitance as being infinite).
Concept
The conceptual model of the buck converter is best understood in terms of the relation
between current and voltage of the inductor. Beginning with the switch open (off-state), the
current in the circuit is zero. When the switch is first closed (on-state), the current will begin
to increase, and the inductor will produce an opposing voltage across its terminals in response
to the changing current.
This voltage drop counteracts the voltage of the source and therefore reduces the net
voltage across the load. Over time, the rate of change of current decreases, and the voltage
across the inductor also then decreases, increasing the voltage at the load. During this time,
the inductor stores energy in the form of a magnetic field. If the switch is opened while the
current is still changing, then there will always be a voltage drop across the inductor, so the
net voltage at the load will always be less than the input voltage source.
When the switch is opened again (off-state), the voltage source will be removed from
the circuit, and the current will decrease. The decreasing current will produce a voltage drop
across the inductor (opposite to the drop at on-state), and now the inductor becomes a Current
Source. The stored energy in the inductor's magnetic field supports the current flow through
the load. This current, flowing while the input voltage source is disconnected, when
concatenated with the current flowing during on-state, totals to current greater than the
average input current (being zero during off-state). The "increase" in average current makes
up for the reduction in voltage, and ideally preserves the power provided to the load. During
the off-state, the inductor is discharging its stored energy into the rest of the circuit. If the
switch is closed again before the inductor fully discharges (on-state), the voltage at the load
will always be greater than zero.
Continuous mode
A buck converter operates in continuous mode if the current through the inductor
never falls to zero during the commutation cycle. In this mode, the operating principle is
described by the plots in figure 4: The current through the inductor rises linearly (in
approximation, so long as the voltage drop is almost constant). As the diode is reverse-biased
by the voltage source V, no current flows through it;
When the switch is opened (bottom of figure 2), the diode is forward biased. The voltage
across the inductor is (neglecting diode drop). Current decreases.

Discontinuous mode

Fig. 5: Evolution of the voltages and currents with time in an ideal buck converter operating
in discontinuous mode.
In some cases, the amount of energy required by the load is too small. In this case, the
current through the inductor falls to zero during part of the period. The only difference in the
principle described above is that the inductor is completely discharged at the end of the
commutation cycle (see figure 5). This has, however, some effect on the previous equations.
The inductor current falling below zero results in the discharging of the output
capacitor during each cycle and therefore higher switching losses. A different control
technique known as Pulse-frequency modulation can be used to minimize these losses.
We still consider that the converter operates in steady state. Therefore, the energy in
the inductor is the same at the beginning and at the end of the cycle (in the case of
discontinuous mode, it is zero). This means that the average value of the inductor voltage
(VL) is zero; i.e., that the area of the yellow and orange rectangles in figure 5 are the same.
This yields:
Qualitatively, as the output capacitor or switching frequency increase, the magnitude of the
ripple decreases. Output voltage ripple is typically a design specification for the power supply
and is selected based on several factors. Capacitor selection is normally determined based on
cost, physical size and non-idealities of various capacitor types. Switching frequency
selection is typically determined based on efficiency requirements, which tends to decrease at
higher operating frequencies, as described below in Effects of non-ideality on the efficiency.
Higher switching frequency can also raise EMI concerns.
Output voltage ripple is one of the disadvantages of a switching power supply, and can also
be a measure of its quality.
Effects of non-ideality on the efficiency
A simplified analysis of the buck converter, as described above, does not account for non-
idealities of the circuit components nor does it account for the required control circuitry.
Power losses due to the control circuitry are usually insignificant when compared with the
losses in the power devices (switches, diodes, inductors, etc.) The non-idealities of the power
devices account for the bulk of the power losses in the converter.
Both static and dynamic power losses occur in any switching regulator. Static power losses
include (conduction) losses in the wires or PCB traces, as well as in the switches and
inductor, as in any electrical circuit. Dynamic power losses occur as a result of switching,
such as the charging and discharging of the switch gate, and are proportional to the switching
frequency.
For N-MOSFETs, the high-side switch must be driven to a higher voltage than Vi. To achieve
this, MOSFET gate drivers typically feed the MOSFET output voltage back into the gate
driver. The gate driver then adds its own supply voltage to the MOSFET output voltage when
driving the high-side MOSFETs to achieve a VGS equal to the gate driver supply voltage.
[7]
Because the low-side VGS is the gate driver supply voltage, this results in very
similar VGS values for high-side and low-side MOSFETs.
A complete design for a buck converter includes a tradeoff analysis of the various power
losses. Designers balance these losses according to the expected uses of the finished design.
A converter expected to have a low switching frequency does not require switches with low
gate transition losses; a converter operating at a high duty cycle requires a low-side switch
with low conduction losses.

Fig. 8: Simplified schematic of a synchronous converter, in which D is replaced by a second


switch, S2.
A synchronous buck converter is a modified version of the basic buck converter circuit
topology in which the diode, D, is replaced by a second switch, S2. This modification is a
tradeoff between increased cost and improved efficiency.
In a standard buck converter, the flyback diode turns on, on its own, shortly after the switch
turns off, as a result of the rising voltage across the diode.

Fig. 9: Schematic of a generic synchronous n-phase buck converter.

Fig. 10: Closeup picture of a multiphase CPU power supply for an AMD Socket 939
processor. The three phases of this supply can be recognized by the three black toroidal
inductors in the foreground. The smaller inductor below the heat sink is part of an input filter.
The multiphase buck converter is a circuit topology where basic buck converter circuits are
placed in parallel between the input and load. Each of the n "phases" is turned on at equally
spaced intervals over the switching period. This circuit is typically used with the synchronous
buck topology, described above.
This type of converter can respond to load changes as quickly as if it switched n times faster,
without the increase in switching losses that would cause. Thus, it can respond to rapidly
changing loads, such as modern microprocessors.
There is also a significant decrease in switching ripple. Not only is there the decrease due to
the increased effective frequency,[8] but any time that n times the duty cycle is an integer, the
switching ripple goes to 0; the rate at which the inductor current is increasing in the phases
which are switched on exactly matches the rate at which it is decreasing in the phases which
are switched off.
Another advantage is that the load current is split among the n phases of the multiphase
converter. This load splitting allows the heat losses on each of the switches to be spread
across a larger area.
This circuit topology is used in computer motherboards to convert the 12 VDC power
supply to a lower voltage (around 1 V), suitable for the CPU. Modern CPU power
requirements can exceed 200 W,[9] can change very rapidly, and have very tight ripple
requirements, less than 10 mV. Typical motherboard power supplies use 3 or 4 phases.
One major challenge inherent in the multiphase converter is ensuring the load current is
balanced evenly across the n phases. This current balancing can be performed in a number of
ways. Current can be measured "losslessly" by sensing the voltage across the inductor or the
lower switch (when it is turned on). This technique is considered lossless because it relies on
resistive losses inherent in the buck converter topology. Another technique is to insert a small
resistor in the circuit and measure the voltage across it. This approach is more accurate and
adjustable, but incurs several costs—space, efficiency and money.
Finally, the current can be measured at the input. Voltage can be measured losslessly, across
the upper switch, or using a power resistor, to approximate the current being drawn. This
approach is technically more challenging, since switching noise cannot be easily filtered out.
However, it is less expensive than emplacing a sense resistor for each phase.
CHAPTER 6
PROPOSED SYSTEM CONFIGURATION
Simulink is a software package for modeling, simulating, and analyzing dynamical systems.
It supports linear and nonlinear systems, modeled in continuous time, sampled time, or a
hybrid of the two. For modeling, Simulink provides a graphical user interface (GUI) for
building models as block diagrams, using click-and-drag mouse operations. Models are
hierarchical, so we can build models using both top-down and bottom-up approaches. We can
view the system at a high level, then double-click on blocks to go down through the levels to
see increasing levels of model detail. This approach provides insight into how a model is
organized and how its parts interact. After we define a model, we can simulate it, using a
choice of integration methods, either from the Simulink menus or by entering commands in
MATLAB's command window. Using scopes and other display blocks, we can see the
simulation results while the simulation is running. In addition, we can change parameters and
immediately see what happens, for "what if" exploration.
The simulation results can be put in the MATLAB workspace for post processing and
visualization. Simulink can be used to explore the behavior of a wide range of real-world
dynamic systems, including electrical circuits, shock absorbers, braking systems, and many
other electrical, mechanical, and thermodynamic systems.
Simulating a dynamic system is a two-step process with Simulink. First, we create a graphical
model of the system to be simulated, using Simulink's model editor. The model depicts the
time-dependent mathematical relationships among the system’s inputs, states, and outputs.
Then, we use Simulink to simulate the behavior of the system over a specified time span.
Simulink uses information that you entered into the model to perform the simulation.

BLOCK DIAGRAM

A Simulink block diagram is a pictorial model of a dynamic system. It consists of a set of


symbols, called blocks, interconnected by lines. Each block represents an elementary
dynamic system that produces an output either continuously (a continuous block) or at
specific points in time (a discrete block). The lines represent connections of block inputs
to block outputs. Every block in a block diagram is an instance of a specific type of block.
The type of the block determines the relationship between a block's outputs and its inputs,
states, and time. A block diagram can contain any number of instances of any type of
block needed to model a system. Blocks represent elementary dynamic systems that
Simulink knows how to simulate. A block comprises one or more of the following:

1) A set of inputs,
2) A set of states, and
3) A set of outputs.
A block's output is a function of time and the block's inputs and states (if any). The
specific function that relates a block's output to its inputs, states, and time depends on the
type of block of which the block is an instance. Continuous Versus discrete Blocks
Simulink's standard block set includes continuous blocks and discrete blocks. Continuous
blocks respond continuously to continuously changing input. Discrete blocks, by contrast,
respond to changes in input only at integral multiples of a fixed interval called the block's
sample time. Discrete blocks hold their output constant between successive sample time hits.
Each discrete block includes a sample time parameter that allows you to specify its sample
rate. The Simulink blocks can be either continuous or discrete, depending on whether they are
driven by continuous or discrete blocks. A block that can be either discrete or continuous is
said to have an implicit sample rate. The implicit sample time is continuous if any of the
block's inputs are continuous. The implicit sample time is equal to the shortest input sample
time if all the input sample times are integral multiples of the shortest time. Otherwise, the
input sample time is equal to the fundamental sample time of the inputs, where the
fundamental sample time of a set of sample times is defined as the greatest integer divisor of
the set of sample times.

Simulink can optionally color code a block diagram to indicate the sample times of
the blocks it contains, e.g., black (continuous), magenta (constant), yellow (hybrid), red
(fastest discrete), and so on. The block contains block name, icon, and block library that
contain the block, the purpose of the block
SIMULINK BLOCK LIBRARIES

Simulink organizes its blocks into block libraries according to their behavior.

1) The Sources library contains blocks that generate signals.


2) The Sinks library contains blocks that display or write block output.
3) The Discrete library contains blocks that describe discrete-time components.
4) The Continuous library contains blocks that describe linear functions.
5) The Math library contains blocks that describe general mathematics functions.
6) The Functions & Tables library contains blocks that describe general functions
and table look-up operations.
7) The Nonlinear library contains blocks that describe nonlinear functions.
8) The Signal & Systems library contains blocks that allow multiplexing and
demultiplexing, implement external input/output, pass data to other parts of the
model, and perform other functions.
9) The Subsystems library contains blocks for creating various types of subsystems.
10) The Block sets and Toolboxes library contains the Extras block library of specialized
blocks.
SUB SYSTEMS

Simulink allows to model a complex system as a set of interconnected subsystems


each of which is represented by a block diagram. We create a subsystem using Simulink's
Subsystem block and the Simulink model editor. We can embed subsystems with subsystems
to any depth to create hierarchical models. We can create conditionally executed subsystems
that are executed only when a transition occurs on a triggering or enabling input.

SOLVERS
Simulink simulates a dynamic system by computing its states at successive time step
solver a specified time span, using information provided by the model. The process of
computing the successive states of a system from its model is known as solving the model.
No single method of solving a model suffices for all systems. Accordingly, Simulink
provides a set of programs, known as solvers, that each embody a particular approach to
solving a model. The Simulation Parameters dialog box allows us to choose the solver most
suitable for our model.
Fixed-Step and Variable-Step Solvers
Fixed-step solvers solve the model at regular time intervals from the beginning to the
end of the simulation. The size of the interval is known as the step-size. We can specify the
step size or let the solver choose the step size. Generally decreasing the step size increases the
accuracy of the results while increasing the time required to simulate the system.
Variable-step solvers vary the step size during the simulation, reducing the step size to
increase accuracy when a model's states are changing rapidly and increasing the step size to
avoid taking unnecessary steps when the model's states are changing slowly. Computing the
step size adds to the computational overhead at each step but can reduce the total number of
steps, and hence simulation time, required to maintain a specified level of accuracy for
models with rapidly changing or piecewise continuous states.
Continuous and Discrete Solvers

Continuous solvers use numerical integration to compute a model's continuous states


at the current time step from the states at previous time steps and the state derivatives.
Continuous solvers rely on the model's blocks to compute the values of the model's discrete
states at each time step. Mathematicians have developed a wide variety of numerical
integration techniques for solving the ordinary differential equations (ODEs) that represent
the continuous states of dynamic systems. Simulink provides an extensive set of fixed-step
and variable-step continuous solvers, each implementing a specific ODE solution method.
Some continuous solvers subdivide the simulation time span into major and minor steps,
where a minor time step represents a subdivision of the major time step. The solver produces
a result at each major time step. It use results at the minor time steps to improve the accuracy
of the result at the major time step.

Discrete solvers exist primarily to solve purely discrete models. They compute the
next simulation time-step for a model and nothing else. They do not compute continuous
states and they rely on the model's blocks to update the model's discrete states. We can use a
continuous solver, but not a discrete solver, to solve a model that contains both continuous
and discrete states. This is because a discrete solver does not handle continuous states. If you
select a discrete solver for a continuous model, Simulink disregards your selection and uses a
continuous solver instead when solving the model.
Simulink provides two discrete solvers, a fixed-step discrete solver and a variable-step
discrete solver. The fixed-step solver by default chooses a step size and hence simulation rate
fast enough to track state changes in the fastest block in our model. The variable-step solver
adjusts the simulation step size to keep pace with the actual rate of discrete state changes in
our model. This can avoid unnecessary steps and hence shorten simulation time for multirate
models.
MODEL EXECUTION PHASE

In the simulation model execution phase, Simulink successively computes the states
and outputs of the system at intervals from the simulation start time to the finish time, using
information provided by the model. The successive time points at which the states and
outputs are computed are called time steps. The length of time between steps is called the
step size. The step size depends on the type of solver used to compute the system's
continuous states, the system's fundamental sample time, and whether the system's
continuous states have discontinuities (Zero Crossing Detection). At the start of the
simulation, the model specifies the initial states and outputs of the system to be simulated. At
each step, Simulink computes new values for the system's inputs, states, and outputs and
updates the model to reflect the computed values. At the end of the simulation, the model
reflects the final values of the system's inputs, states, and outputs. At each time step:

1) Simulink Updates the outputs of the models' blocks in sorted order. Simulink
computes a block's outputs by invoking the block's output function. Simulink passes
the current time and the block's inputs and states to the output function as it may
require these arguments to compute the block's output. Simulink updates the output of
a discrete block only if the current step is an integral multiple of the block's sample
time.
2) Updates the states of the model's blocks in sorted order. Simulink computes a block's
discrete states by invoking its discrete state update function. Simulink computes a
block's continuous states by numerically integrating the time derivatives of the
continuous states. It computes the time derivatives of the states by invoking the
block's continuous derivatives function.
3) Optionally checks for discontinuities in the continuous states of blocks. Simulink uses
a technique called zero crossing detection to detect discontinuities in continuous
states.
4) Computes the time for the next time step.
Simulink repeats steps 1 through 4 until the simulation stop time is reached.

Block Sorting Rules


Simulink uses the following basic update rules to sort the blocks:

1) Each block must be updated before any of the direct-feed through blocks that it
drives. This rule ensures that the inputs to direct-feed through blocks will be valid
when they are updated.
2) Non direct-feed through blocks can be updated in any order as long as they are
updated before any direct-feed through blocks that they drive. This rule can be met by
putting all non direct-feed through blocks at the head of the update list in any order. It
thus allows Simulink to ignore non direct-feed through blocks during the sorting
process.
The result of applying these rules is an update list in which non direct-feed through
blocks appear at the head of the list in no particular order followed by direct-feed through
blocks in the order required to supply valid inputs to the blocks they drive. During the sorting
process, Simulink checks for and flags the occurrence of algebraic loops, that is, signal loops
in which an output of a direct-feed through block is connected directly or indirectly to one of
the block's inputs. Such loops seemingly create a deadlock condition since Simulink needs
the input of a direct-feed through block in order to compute its output. However, an algebraic
loop can represent a set of simultaneous algebraic equations (hence the name) where the
block's input and output are the unknowns. Further, these equations can have valid solutions
at each time step. Accordingly, Simulink assumes that loops involving direct-feed through
blocks do, in fact, represent a solvable set of algebraic equations and attempts to solve them
each time the block is updated during a simulation. The issue of global warming has been
raised due to the consumption of oil, coal and natural resources for the electricity producing
process. Apart from that, the amounts of these fuels on earth are now decreasing day by day.
Thus, the focus has been shifted to the green alternative energy which is not polluted and has
no impact on the environment. The power of wind is now being explored which the
researcher believes that it has all the qualifications to replace traditional fuel since it has less
effect to global warming.

Fig 1 Proposed circuit configuration


During the past decade, the amounts of wind capacity have been installed every three years.
Around 83% of wind capacities are located in these five countries, German, United states or
America, Denmark, India and Spain. Wind energy conversion system (WECS) consist of
wind turbine, pitch angle control, drive train, generator and power converter. There are
various kinds of generators used in WECS such as induction generator (IG), doubly fed
induction generator (DFIG) and permanent magnet synchronous generator (PMSG). The
PMSG based on WECS can connect to the turbine without using gearbox. The gearbox
causes in the cost of maintenance, and then it will decrease the weight of nacelle.

Fig 2 proposed system input wind speed vs time


MATLAB/SIMULINK is a graphical software package for modelling, simulating and
analysing dynamic systems. It supports linear and nonlinear systems both in continuous time
and sample time. There are various researches presented the PMSG based on WECS in
MATLAB/SIMULINK [8-9]. In order to understand the principle of producing electricity
from wind with WECS and analyze before it will be implemented in practical case. Thus, it is
needed to further study and investigate the PMSG based on WECS.

Fig 3 dc link voltage after wind conversion system vs time

Fig 4 inverter input voltage vs time


Fig 5 grid voltage vs time
CHAPTER 7
CONCLUSION
WECSs are widely employed nowadays in stand-alone systems for providing power
to isolated loads, as well as in distributed generation systems, microgrids, and smart grids. In
all of these applications, appropriate energy management processes must be performed in
order to maximize the energy production of the wind turbines and transfer the wind-generated
energy to the consumer with high efficiency and adequate power quality. This paper
presented the modelling of PMSG based on WECS. It consists of wind turbine, drive train,
PMSG, pitch angle control and ac/dc converter model. PMSG and converter are established
in proposed model.
REFERENCES
[1] Y. V. Pavan Kumar, Ravikumar Bhimasingu, “Renewable energy based microgrid system
sizing and energy management for green buildings,” Springer Journal of Modern Power
Systems and Clean Energy, Vol. 3, No. 1, pp. 1-13, 2015.
[2] Moein Lak, V. K. Ramachandaramurthy. “Speed control for direct drive permanent
magnet wind turbine,” in Proc. 2014 International 2020 International Conference on Artificial
Intelligence and Signal Processing (AISP) Conference on Renewable Energy Research and
Application (ICRERA), Milwaukee, WI, USA, pp. 317-321, 2014.
[3] Y. D. Song, B. Dhinakaran, “Nonlinear variable speed control of wind turbines,” in Proc.
1999 IEEE International Conference on Control Applications (Cat. No. 99CH36328), vol. 1,
pp. 814-819, 1999.
[4] Essam H. Abdou, Abdel-Raheem Youssef, Salah Kamel, Mohamed M. Aly, “Sensorless
Wind Speed Control of 1.5 MW DFIG wind turbines for MPPT,” in Proc. 2018 Twentieth
International Middle East Power Systems Conference (MEPCON), pp. 700-704, 2018.
[5] Ying Liu, Guofen Tang, Haidong Liu, Bo Zhou, Honghao Guo, Guangjie Zuo. “Speed
control method of double salient electro-magnetic wind power generator in full range of wind
speed,” in Proc. 2010 World Non-Grid-Connected Wind Power and Energy Conference,
Nanjing, China, pp. 1-5, 2010.
[6] Zurong Hu, Junqi Wang, Yundong Ma, Xing Yan. “Research on speed control system for
fixed-pitch wind turbine based on disturbance observer,” in Proc. 2009 World Non-Grid-
Connected Wind Power and Energy Conference, Nanjing, China, pp. 1-5, 2009.
[7] Mahinsasa Narayana, Ghanim Putrus, Milutin Jovanovic, Pak Sing Leung, “Predictive
control of wind turbines by considering wind speed forecasting techniques,” in Proc. 2009
44th International Universities Power Engineering Conference (UPEC), Glasgow, UK, pp. 1-
4, 2009.
[8] Dinh-Chung Phan, Shigeru Yamamoto, “Rotor speed control of doubly fed induction
generator wind turbines using adaptive maximum power point tracking,” Elsevier Energy
Journal, Vol. 111, pp. 377-388, Sep. 2016.
[9] Vlaho Petrovic, Carlo L. Bottasso, “Wind turbine envelope protection control over the full
wind speed range,” Elsevier Renewable Energy Journal, Vol. 111, pp.836-848, Oct. 2017.
[10] Ehab S. Ali, “Speed control of induction motor supplied by wind turbine via imperialist
competitive algorithm,” Elsevier Energy Journal, Vol. 89, pp. 593-600, Sep. 2015.
[11] Andrea Tilli, Christian Conficoni, “Speed control for medium power wind turbines: an
integrated approach oriented to mppt,” IFAC Proceedings Volumes, Vol. 44, No. 1, pp. 544-
550, Jan. 2011.
[12] Christian Leisten, Uwe Jassmann, Johannes Balshusemann, Dirk Abel, “Model
predictive speed control of a wind turbine system test bench,” IFAC-PapersOnLine, Vol. 51,
No. 32, pp.349-354, 2018.
[13] Zhenyu Zhang, Yongduan Song, Peng Li, WenLiang Wang, Ming Qin, “Adaptive and
robust variable-speed control of wind turbine based on virtual parameter approach,” in Proc.
32nd Chinese Control Conference, pp. 8886-8890, 2013.
[14] A. J. Mahdi, W. H. Tang, Q. H. Wu, “Derivation of a complete transfer function for a
wind turbine generator system by experiments,” in Proc. 2011 IEEE Power Engineering and
Automation Conference, Vol. 1, pp. 35-38, 2011.
[15] Y. Jaganmohan Reddy, Y. V. Pavan Kumar, K. Padma Raju, Anilkumar Ramsesh,
“Retrofitted hybrid power system design with renewable energy sources for buildings,” IEEE
Trans. Smart Grid, Vol. 3, No. 4, pp. 2174 – 2187, Dec. 2012.
[16] B. Vasu Murthy, Y. V. Pavan Kumar, U. V. Ratna Kumari, “Application of neural
networks in process control: Automatic/online tuning of PID controller gains for ±10%
disturbance rejection,” in Proc. 2012 IEEE International Conference on Advanced
Communication Control and Computing Technologies (ICACCCT), 2012. [17] B. V.
Murthy, Y. V. P. Kumar, U. V. R. Kumari, “Fuzzy logic intelligent controlling concepts in
industrial furnace temperature process control,” in Proc. 2012 IEEE International Conference
on Advanced Communication Control and Computing Technologies, 2012.

You might also like