Вы находитесь на странице: 1из 86

Browse Journals & Magazines > Power Systems, IEEE Transacti ...

> Volume:11 Issue:4


Steady-state and dynamic models of unified power flow controller
(UPFC) for power system studies
Nabavi-Niaki, A. ; Dept. of Electr. & Comput. Eng., Toronto Univ., Ont., Canada ; Iravani, M.R.
This paper provides comprehensive development procedures and final forms of mathematical
models of a unified power flow controller (UPFC) for steady-state, transient stability and
eigenvalue studies. Based on the developed models, the impacts of control strategy, parameters
and location of UPFC on power system operating conditions are discussed. The accuracy of the
developed models is verified through comparing the study results with those obtained from
detailed time-domain simulation using the Electromagnetic Transients Program (EMTP)
Published in:
Power Systems, IEEE Transactions on (Volume:11 , Issue: 4 )
Date of Publication:
Nov 1996
Page(s):
1937 - 1943
ISSN :
0885-8950
INSPEC Accession Number:
5453383
Digital Object Identifier :
10.1109/59.544667
Date of Current Version :
06 August 2002
Issue Date :
Nov 1996
Sponsored by :
IEEE Power & Energy Society
Steady-state and dynamic models of unified power flow controller (UPFC) for
power system studies
Nabavi-Niaki, A. ; Iravani, M.R.
Power Systems, IEEE Transactions on
Volume: 11 , Issue: 4
Digital Object Identifier: 10.1109/59.544667
Publication Year: 1996 , Page(s): 1937 - 1943
Cited by: Papers (122)
IEEE Journals & Magazines


Distributed generation
From Wikipedia, the free encyclopedia
Distributed generation, also called on-site generation, dispersed generation, embedded
generation, decentralized generation, decentralized energy, distributed energy or district
energy,
[1]
generates electricity from many small energy sources. Most countries generate
electricity in large centralized facilities, such as fossil fuel (coal, gas powered), nuclear, large
solar power plants or hydropower plants. These plants have excellent economies of scale, but
usually transmit electricity long distances and can negatively affect the environment. Distributed
generation allows collection of energy from many sources and may give lower environmental
impacts and improved security of supply.


Local wind generator, Spain, 2010
Contents
1 Economies of scale
o 1.1 Grid parity
2 Types of distributed energy resources
o 2.1 Cogeneration
o 2.2 Solar panel
o 2.3 Wind turbine
o 2.4 Vehicle-to-grid
o 2.5 Waste-to-energy
3 Integration with the grid
4 Cost factors
5 Microgrid
6 Modes of power generation
7 Communication in DER systems
8 Legal requirements for distributed generation
9 See also
10 References
11 Further reading
12 External links
Economies of scale
Historically, central plants have been an integral part of the electric grid, in which large
generating facilities are specifically located either close to resources or otherwise located far
from populated load centers. These, in turn, supply the traditional transmission and distribution
(T&D) grid that distributes bulk power to load centers and from there to consumers. These were
developed when the costs of transporting fuel and integrating generating technologies into
populated areas far exceeded the cost of developing T&D facilities and tariffs. Central plants are
usually designed to take advantage of available economies of scale in a site-specific manner, and
are built as "one-off," custom projects.
These economies of scale began to fail in the late 1960s and, by the start of the 21st century,
Central Plants could arguably no longer deliver competitively cheap and reliable electricity to
more remote customers through the grid, because the plants had come to cost less than the grid
and had become so reliable that nearly all power failures originated in the grid.
[citation needed]
Thus,
the grid had become the main driver of remote customers power costs and power quality
problems, which became more acute as digital equipment required extremely reliable
electricity.
[2][3]
Efficiency gains no longer come from increasing generating capacity, but from
smaller units located closer to sites of demand.
[4][5]

For example, coal power plants are built away from cities to prevent their heavy air pollution
from affecting the populace. In addition, such plants are often built near collieries to minimize
the cost of transporting coal. Hydroelectric plants are by their nature limited to operating at sites
with sufficient water flow.
Low pollution is a crucial advantage of combined cycle plants that burn natural gas. The low
pollution permits the plants to be near enough to a city to provide district heating and cooling.
Distributed generation plants are mass-produced, small, and less site-specific. Their development
arose out of:
1. concerns over perceived externalized costs of central plant generation, particularly
environmental concerns,
2. the increasing age, deterioration, and capacity constraints upon T&D for bulk power,
3. the increasing relative economy of mass production of smaller appliances over heavy
manufacturing of larger units and on-site construction, and
4. Along with higher relative prices for energy, higher overall complexity and total costs for
regulatory oversight, tariff administration, and metering and billing.
Capital markets have come to realize that right-sized resources, for individual customers,
distribution substations, or microgrids, are able to offer important but little-known economic
advantages over Central Plants. Smaller units offered greater economies from mass-production
than big ones could gain through unit size. These increased valuedue to improvements in
financial risk, engineering flexibility, security, and environmental qualityof these resources
can often more than offset their apparent cost disadvantages.
[6]
DG, vis--vis Central Plants, must
be justified on a life-cycle basis.
[7]
Unfortunately, many of the direct, and virtually all of the
indirect, benefits of DG are not captured within traditional utility cash-flow accounting.
[2]

While the levelized generation cost of distributed generation is more expensive than
conventional sources on a kWh basis, this does not consider negative aspects of conventional
fuels. The additional premium for DG is rapidly declining as demand increases and technology
progresses,
[citation needed]
and sufficient and reliable demand may bring economies of scale,
innovation, competition, and more flexible financing, that could make DG clean energy part of a
more diversified future.
[citation needed]

Distributed generation reduces the amount of energy lost in transmitting electricity because the
electricity is generated very near where it is used, perhaps even in the same building. This also
reduces the size and number of power lines that must be constructed.
Typical distributed power sources in a Feed-in Tariff (FIT) scheme have low maintenance, low
pollution and high efficiencies. In the past, these traits required dedicated operating engineers
and large complex plants to reduce pollution. However, modern embedded systems can provide
these traits with automated operation and renewables, such as sunlight, wind and geothermal.
This reduces the size of power plant that can show a profit.
Grid parity
Grid parity(or 'socket parity') occurs when an alternative energy source can generate electricity at
a levelized cost (LCoE) that is less than or equal to the price of purchasing power from the
electricity grid. Reaching grid parity is considered to be the point at which an energy source
becomes a contender for widespread development without subsidies or government support. It is
widely believed that a wholesale shift in generation to these forms of energy will take place
when they reach grid parity.
Grid parity has been reached in some locations with on-shore wind power around 2000, and with
solar power it was achieved for the first time in Spain in 2013.
[8][9][10]

Types of distributed energy resources
Distributed energy resource (DER) systems are small-scale power generation technologies
(typically in the range of 1 kW to 10,000 kW) used to provide an alternative to or an
enhancement of the traditional electric power system. The usual problem with distributed
generators are their high initial capital costs.
Cogeneration
Distributed cogeneration sources use steam turbines, natural gas-fired fuel cells, microturbines or
reciprocating engines
[11]
to turn generators. The hot exhaust is then used for space or water
heating, or to drive an absorptive chiller
[12][13]
for cooling such as air-conditioning. In addition to
natural gas-based schemes, distributed energy projects can also include other renewable or low
carbon fuels including biofuels, biogas, landfill gas, sewage gas, coal bed methane, syngas and
associated petroleum gas.
[14]

Delta-ee consultants stated in 2013 that with 64% of global sales the fuel cell micro combined
heat and power passed the conventional systems in sales in 2012.
[15]
20.000 units where sold in
Japan in 2012 overall within the Ene Farm project. With a Lifetime of around 60,000 hours. For
PEM fuel cell units, which shut down at night, this equates to an estimated lifetime of between
ten and fifteen years.
[16]
For a price of $22,600 before installation.
[17]
For 2013 a state subsidy for
50,000 units is in place.
[16]

In addition, molten carbonate fuel cell and solid oxide fuel cells using natural gas, such as the
ones from FuelCell Energy and the Bloom energy server, or waste-to-energy processes such as
the Gate 5 Energy System are used as a distributed energy resource.
Solar panel
A primary issue with solar power is that it is intermittent. Popular sources of power for
distributed generation are solar heat collection panels and solar panels on the roofs of buildings
or free-standing. Solar heating panels are used mostly for heating water and when the water is
heated into steam it can effectively and economically used in steam turbines to produce
electricity.
The production cost for electricity produced from photovoltaic panels ranges from $0.99 to
2.00/W (2007) plus installation and supporting equipment unless the installation is Do it yourself
(DIY) bringing the cost to $0.525 to 0.750/W (2010).
[18]
This is comparable to coal power plant
costs of $0.582 to 0.906/W (1979),
[19][20]
adjusting for inflation. Nuclear power is higher at $2.20
to $6.00/W (2007).
[21]
Some "thin-film" solar cells have waste-disposal issues when they are
made with heavy metals such as Cadmium telluride (CdTe) and Copper indium gallium selenide
(CuInGaSe), and must be recycled, as opposed to silicon solar cells, which are mostly non-
metallic. Unlike coal and nuclear, there are no fuel costs, operating pollution, mining-safety or
operating-safety issues. Solar power has a low capacity factor, producing peak power at local
noon each day. Average capacity factor is typically 20%.
Wind turbine
Another source is small wind turbines. These have low maintenance, and low pollution, however
as with solar, wind energy is intermittent. Construction costs are higher ($0.80/W, 2007) per watt
than large power plants, except in very windy areas. Wind towers and generators have substantial
insurable liabilities caused by high winds, but good operating safety. In some areas of the US
there may also be Property Tax costs involved with wind turbines that are not offset by
incentives or accelerated depreciation.
[22]
Wind also tends to complement solar. Days without
sun there tend to be windy, and vice versa.
[citation needed]
Many distributed generation sites combine
wind power and solar power such as Slippery Rock University, which can be monitored online.
Vehicle-to-grid
Future generations of electric vehicles may have the ability to deliver power from the battery in a
vehicle-to-grid into the grid when needed.
[23]
An electric vehicle network could also be an
important distributed generation resource.
[citation needed]

Waste-to-energy
Municipal solid waste (MSW) and natural waste, such as sewage sludge, food waste and animal
manure will decompose and discharge methane-containing gas that can be collected as used as
fuel in gas turbines or micro turbines to produce electricity as a distributed energy resource.
Additionally, a California-based company, Gate 5 Energy Partners, Inc. has developed a process
that transforms natural waste materials, such as sewage sludge, into biofuel that can be
combusted to power a steam turbine that produces power. This power can be used in lieu of grid-
power at the waste source (such as a treatment plant, farm or dairy).
Integration with the grid
For reasons of reliability, distributed generation resources would be interconnected to the same
transmission grid as central stations. Various technical and economic issues occur in the
integration of these resources into a grid. Technical problems arise in the areas of power quality,
voltage stability, harmonics, reliability, protection, and control.
[24]
Behavior of protective devices
on the grid must be examined for all combinations of distributed and central station
generation.
[25]
A large scale deployment of distributed generation may affect grid-wide functions
such as frequency control and allocation of reserves.
[26]
As a result smart grid functions, virtual
power plants and grid energy storage such as power to gas stations are added to the grid.
Cost factors
Cogenerators are also more expensive per watt than central generators.
[citation needed]
They find
favor because most buildings already burn fuels, and the cogeneration can extract more value
from the fuel . Local production has no electricity transportation losses on long distance power
lines or energy losses from the Joule effect in transformers where in general 8-15% of the energy
is lost.
[27]

Some larger installations utilize combined cycle generation. Usually this consists of a gas turbine
whose exhaust boils water for a steam turbine in a Rankine cycle. The condenser of the steam
cycle provides the heat for space heating or an absorptive chiller. Combined cycle plants with
cogeneration have the highest known thermal efficiencies, often exceeding 85%.
In countries with high pressure gas distribution, small turbines can be used to bring the gas
pressure to domestic levels whilst extracting useful energy. If the UK were to implement this
countrywide an additional 2-4 GWe would become available. (Note that the energy is already
being generated elsewhere to provide the high initial gas pressure - this method simply
distributes the energy via a different route.)
Microgrid


Picture of a local microgrid, the Sendai Microgrid, located on the campus of Tohoku Fukushi University in
Sendai City in the Tohoku district in Japan
A microgrid is a localized grouping of electricity generation, energy storage, and loads that
normally operates connected to a traditional centralized grid (macrogrid). This single point of
common coupling with the macrogrid can be disconnected. The microgrid can then function
autonomously.
[28]
Generation and loads in a microgrid are usually interconnected at low voltage.
From the point of view of the grid operator, a connected microgrid can be controlled as if it were
one entity.
Microgrid generation resources can include fuel cells, wind, solar, or other energy sources. The
multiple dispersed generation sources and ability to isolate the microgrid from a larger network
would provide highly reliable electric power. Produced heat from generation sources such as
microturbines could be used for local process heating or space heating, allowing flexible trade
off between the needs for heat and electric power.
Micro-grids were proposed in the wake of the July 2012 India blackout:
[29]

Small micro-grids covering 3050 km radius
[29]

Small power stations of 510 MW to serve the micro-grids
Generate power locally to reduce dependence on long distance transmission lines and cut
transmission losses.
Modes of power generation
DER systems may include the following devices/technologies:
Combined heat power (CHP)
Fuel cells
Micro combined heat and power (MicroCHP)
Microturbines
Photovoltaic Systems
Reciprocating engines
Small Wind power systems
Stirling engines
Trigeneration
Communication in DER systems
IEC 61850-7-420 is under development as a part of IEC 61850 standards, which deals with the
complete object models as required for DER systems. It uses communication services mapped to
MMS as per IEC 61850-8-1 standard.
OPC is also used for the communication between different entities of DER system.
Legal requirements for distributed generation
In 2010 Colorado enacted a law requiring that by 2020 that 3% of the power generated in
Colorado utilize distributed generation of some sort.
[30][31]

See also

Energy portal

Sustainable development portal
Autonomous building
Demand response
Energy harvesting
Electric power transmission
Electricity generation
Electricity market
Electricity retailing
Energy demand management
Future energy development
Green power superhighway
Grid-tied electrical system
Hydrogen station
IEEE 1547 Standard for Interconnecting
Distributed Resources with Electric
Power Systems
Islanding
Microgeneration
Net metering
Relative cost of electricity generated by
different sources
Renewable energy development
Smart meter
Smart power grid
Solar Guerrilla
Stand-alone power system
Sustainable community energy system
Trigeneration
World Alliance for Decentralized Energy
References
1. Jump up ^ District Energy and CHP, www.clarke-energy.com, retrieved 16 September 2013
2. ^ Jump up to:
a

b
DOE; The Potential Benefits of Distributed Generation and Rate-Related Issues
that May Impede Their Expansion; 2007.
3. Jump up ^ Lovins; Small Is Profitable: The Hidden Economic Benefits of Making Electrical
Resources the Right Size; Rocky Mountain Institute, 2002.
4. Jump up ^ Takahashi, et al; Policy Options to Support Distributed Resources; U. of Del., Ctr. for
Energy & Env. Policy; 2005.
5. Jump up ^ Hirsch; 1989; cited in DOE, 2007.
6. Jump up ^ Lovins; Small Is Profitable: The Hidden Economic Benefits of Making Electrical
Resources the Right Size; Rocky Mountain Institute; 2002
7. Jump up ^ Michigan (Citation pending)
8. Jump up ^ http://www.forbes.com/sites/peterdetwiler/2012/12/26/solar-grid-parity-comes-to-
spain/
9. Jump up ^ http://oilprice.com/Latest-Energy-News/World-News/Spain-Achieves-Grid-Parity-for-
Solar-Power.html
10. Jump up ^ http://www.conergy.com/desktopdefault.aspx/tabid-136/281_read-831/
11. Jump up ^ Gas engine cogeneration, www.clarke-energy.com, retrieved 9.12.2013
12. Jump up ^ Cogeneration with absorptive chiller
13. Jump up ^ Trigeneration with gas engines, www.clarke-energy.com, retrieved 9.12.2013
14. Jump up ^ Gas engine applications, www.clarke-energy.com, retrieved 9th December 2013
15. Jump up ^ The fuel cell industry review 2013
16. ^ Jump up to:
a

b
Latest developments in the Ene-Farm scheme
17. Jump up ^ Launch of new 'Ene-Farm' home fuel cell product more affordable and easier to
install
18. Jump up ^ Solar panel production costs
19. Jump up ^ Osti 1979
20. Jump up ^ Coal power plants future
21. Jump up ^ How much?
22. Jump up ^ [1] Retrieved on 20 October 2010
23. Jump up ^ How electric vehicles are a part of distributed generation
24. Jump up ^ Tomoiag, B.; Chindri, M.; Sumper, A.; Sudria-Andreu, A.; Villafafila-Robles, R. Pareto
Optimal Reconfiguration of Power Distribution Systems Using a Genetic Algorithm Based on
NSGA-II. Energies 2013, 6, 1439-1455.
25. Jump up ^ P. Mazidi, G. N. Sreenivas; Reliability Assessment of A Distributed Generation
Connected Distribution System; International Journal of Power System Operation and Energy
Management(IJPSOEM), Nov. 2011
26. Jump up ^ Math H. Bollen, Fainan Hassan Integration of Distributed Generation in the Power
System, John Wiley & Sons, 2011 ISBN 1-118-02901-1, pages v-x
27. Jump up ^ How big are Power line losses?
28. Jump up ^ Stan Mark Kaplan, Fred Sissine, (ed.) Smart grid: modernizing electric power
transmission and distribution... The Capitol Net Inc, 2009, ISBN 1-58733-162-4, page 217
29. ^ Jump up to:
a

b
[2]
30. Jump up ^ "Going Solar Is Harder Than It Looks, a Valley Finds" article by Kirk Johnson in The
New York Times June 3, 2010
31. Jump up ^ "Colorado Increases Renewables Requirements" blog by Kate Galbraith on
NYTimes.Com March 22, 2010
Further reading
Brass, J. N.; Carley, S.; MacLean, L. M.; Baldwin, E. (2012). "Power for Development: A Review of
Distributed Generation Projects in the Developing World". Annual Review of Environment and
Resources 37: 107. doi:10.1146/annurev-environ-051112-111930. edit
Gies, Erica. Making the Consumer an Active Participant in the Grid, The New York Times,
November 29, 2010. Discusses distributed generation and the U.S. Federal Energy Regulatory
Commission.
Pahl, Greg (2012). Power from the people : how to organize, finance, and launch local energy
projects. Santa Rosa, Calif: Post Carbon Institute. ISBN 9781603584098.
External links
The UK District Energy Association - advocating the construction of locally distributed energy
networks
Decentralized Power as Part of Local and Regional Plans
IEEE P1547 Draft Standard for Interconnecting Distributed Resources with Electric Power
Systems
World Alliance for Decentralized Energy
The iDEaS project by University of Southampton on Decentralised Energy
Biofuels and gas pressure energy recovery
DER-Model
DERlab
Center for Energy and innovative Technologies
Decentralized Power System (DPS) in Pakistan
1. Introduction to Distributed Generation
1.1 What is Distributed Generation?
Distributed generation is an approach that employs small-scale technologies to produce
electricity close to the end users of power. DG technologies often consist of modular (and
sometimes renewable-energy) generators, and they offer a number of potential
benefits. In many cases, distributed generators can provide lower-cost electricity and
higher power reliability and security with fewer environmental consequences than can
traditional power generators.
In contrast to the use of a few large-scale generating stations located far from load
centers--the approach used in the traditional electric power paradigm--DG systems
employ numerous, but small plants and can provide power onsite with little reliance on
the distribution and transmission grid. DG technologies yield power in capacities that
range from a fraction of a kilowatt [kW] to about 100 megawatts [MW]. Utility-scale
generation units generate power in capacities that often reach beyond 1,000 MW.
Classic Electricity Paradigm--Central Power Station Model
The current model for electricity generation and distribution in the United States is
dominated by centralized power plants. The power at these plants is typically combustion
(coal, oil, and natural) or nuclear generated. Centralized power models, like this, require
distribution from the center to outlying consumers. Current substations can be anywhere
from 10s to 100s of miles away from the actual users of the power generated. This
requires transmission across the distance.
This system of centralized power plants has many disadvantages. In addition to the
transmission distance issues, these systems contribute to greenhouse gas emission, the
production of nuclear waste, inefficiencies and power loss over the lengthy transmission
lines, environmental distribution where the power lines are constructed, and security
related issues.
Many of these issues can be mediated through distributed energies. By locating, the
source near or at the end-user location the transmission line issues are rendered obsolete.
Distributed generation (DG) is often produced by small modular energy conversion units
like solar panels. As has been demonstrated by solar panel use in the United States, these
units can be stand-alone or integrated into the existing energy grid. Frequently,
consumers who have installed solar panels will contribute more to the grid than they take
out resulting in a win-win situation for both the power grid and the end-user.

Classic Electricity Paradigm

Distributed Generation (DG) Electricity Paradigm
(This figure has been taken from the European Commission Energy Research website:
http://ec.europa.eu/research/energy/nn/nn_rt/nn_rt_dg/article_1158_en.htm)
What are Some Examples of Distributed Generation Technologies?
Distributed generation takes place on two-levels: the local level and the end-point level.
Local level power generation plants often include renewable energy technologies that are
site specific, such as wind turbines, geothermal energy production, solar systems
(photovoltaic and combustion), and some hydro-thermal plants. These plants tend to be
smaller and less centralized than the traditional model plants. They also are frequently
more energy and cost efficient and more reliable. Since these local level DG producers
often take into account the local context, the usually produce less environmentally
damaging or disrupting energy than the larger central model plants.

Wind Turbines at Buffalo Mountain, TN

Photovoltaic (Solar) Panels help Power this Elementary School in Fairbanks, Alaska

A 300 kW Capstone Microturbine at a Demonstration Project at the Oak Ridge
National Laboratory in Oak Ridge, TN
Phosphorus fuel cells also provide an alternative route to a DG technology. These are not
as environmentally reliant as the previously mentioned technologies. These fuel cells are
able to provide electricity through a chemical process rather than a combustion process.
This process produces little particulate waste.
At the end-point level the individual energy consumer can apply many of these same
technologies with similar effects. One DG technology frequently employed by end-point
users is the modular internal combustion engine. For example, some departments here at
Virginia Tech use these power generators as a backup to the normal power grid. These
modular internal combustion engines can also be used to backup RVs and homes. As
many of these familiar examples show DG technologies can operate as isolated "islands"
of electric energy production or they can serve as small contributors to the power grid.

1.2 History of the Emergence of Distributed Generation
A Brief History of the Emergence of DG Technologies in the United States
Among a set of five laws proposed by Carter and passed by Congress (albeit in greatly
diluted form), the Public Utility Regulatory Policies Act (PURPA) of 1978 had the most
far-reachingand least intendedconsequences for power companies. It spurred
creation of radical technologies; it began the process of deregulation; and it challenged
the control held by power company managers. In the process, the law helped change the
momentum of the utility system.
A related provision of PURPA also spurred research on environmentally preferable
technologies that used water, wind, or solar power to produce electricity. More
successful than anyone originally anticipated, PURPA prompted work that cut the cost of
power produced by solar photovoltaic panels by about 70 percent between 1980 and
1995. More significantly, it contributed to work that lowered the cost of power produced
by wind turbines; by 2002, the average cost of wind-produced electricity dropped to
under 5 cents per kWh, a cost that compared favorably with electricity produced by
conventional utility plants burning natural gas or coal.
These smaller-scale generation technologies challenged the established paradigm of the
utility industry (and many other industries) that previously relied on large-scale
equipment to produce economies of scale. Now, it appeared, power plants producing
modest amounts of electricity proved economically viable. Moreover, since they were
smaller, they required less time to build, and they put less capital at risk during a period
of rapid price inflation. Finally, they matched the slower growth rate of consumption
more appropriately: with growth rates remaining under 3 percent per year, consumers
needed smaller increments of power to match their demand. Had utilities continued to
build their traditional behemoths, huge chunks of power would remain unused when the
plants were completed. Small scale, indeed, looked beautiful.
This unintended experiment with competition suggested to influential regulators and
legislators in the 1980s that more competition would benefit stakeholders in the electric
utility industry. (Not coincidentally, competition had already begun in the airline,
telecommunications, and natural gas industries.) Some academics and politicians
wondered if utility regulation still had merit, seeing that a traditional justification of
government oversightthe fact that power companies constituted natural monopolies
no longer appeared valid. After all, if nonutilities could produce power as cheaply as
could utilities, then the big power companies no longer deemed recognition as natural
monopolies. And if they were not natural monopolies, they no longer deserved special
status as noncompetitive entities that required regulation. Why not permit increased
competition to thrive outside the realm of the PURPA-inspired generation companies?
Already feeling their control of the system threatened by pressures resulting from
implementation of PURPA, utility managers had more to fear. After the Gulf War
focused attention again on the cost and security of energy supplies, Congress passed the
Energy Policy Act of 1992. The legislation sought to employ competitive forces to
increase domestic fuel production and to improve the efficiency of energy use. One
provision gave states the option of opening up their transmission network to use by
competitors. The network would serve as a common carrier so any electricity producer
could sell power to any customer. Essentially, the law gave states the right to begin
competition on the retail level. During the late 1990s, several states (with California
being among the first) passed legislation that established competitive retail frameworks
for power. By September 2001, 23 states (and the District of Columbia) had passed
similar legislation, while regulatory bodies in several other states had reduced their
oversight and had introduced market forces into the system.
The restructuring process, stimulated by PURPA and pursued by advocates of market
forces, meant that momentum in the utility system had been altered, largely because
utility managers lost control of their system. For almost a century, managers
commanded the huge-scale, incrementally improving conservative technologies that
produced and distributed electricity. But in recent decades, they began facing
competition from entrepreneurial companies that employed small-scale fossil fuel and
renewable energy technologies. In 1992, nonutility companies controlled 1.5 percent of
the nations total power capacity; that number rose to 34.7 percent in 2003. In addition,
power company managers lost the benefit of traditional regulation, which protected
companies from competition on the basis of their claimnow challengedto natural
monopoly status. At the same time, managers lost support from those stakeholders who
previously buttressed them: financiers, equipment manufacturers, and educational
institutions adapted to the changing environment and found new opportunities serving the
needs of the growing number of nonutility companies.
Assuming some of the political and economic power that managers once held, other
stakeholders began making new waves. Environmental advocates, for example, gained
impressive standing in the legislative process that led to creation of restructuring laws.
While supporting deregulation in general, they fought for (and in many cases won)
provisions in laws that guaranteed funding for renewable energy and energy-efficiency
initiatives. In Californias version of restructuring, for example, utilities earned the right
to receive payment from customers for building what turned out to be expensive power
plants during the era of regulation, but which would have little economic value in an era
of competition (called stranded assets). But environmental advocates won provision
for expenditures on energy efficiency work, renewable energy technologies, and for
development of research on new technologies that had not yet shown commercial
viability.
As stakeholders began renegotiating their positions in an altered utility system, the
California electricity crisis of 2000 and 2001 created a sense of chaos. Subsequent
blackouts in the Midwest during 2002 and in the Northeast and Canada during 2003
contributed further to that disharmony. In California, where one utility declared
bankruptcy as a result of the crisis, the state suspended competition altogether and
expanded its control of the wholesale market. Such events caused policy makers in other
states to reconsider their previous enthusiasm for restructuring and to slow down plans to
introduce market forces. At the same time, the Federal Energy Regulatory Commission
proposed new initiatives in response to the 2003 blackout that may give it greater control
over the increasingly fragile-looking transmission grid. That grid has witnessed serious
underinvestment since the 1990s as utility companies and nonutility entrepreneurs
remained concerned in an uncertain policy setting about how the grid will be employed
and which stakeholders will profit from its use.
The unsettled state of affairs in the power system has provided opportunities for
advocates of environmentally-friendly and distributed-generation technologies. Taking
advantage of the flux within the utility system, especially in states with strong traditions
of politically astute environmental advocates, activists won passage of laws for funding
of renewable energy and small-scale generation technologies. Customers paid into
public benefit funds (also known as system benefit funds) regardless of which
company (a traditional utility or nonutility company) provided them with electricity. And
in eighteen states (plus the District of Columbia), advocates convinced legislators to
enact laws creating renewable portfolio standards that required all competitive power
companies to produce (or to purchase) a certain amount of power coming from small-
scale, renewable resources.

<< Back Next >>
2007 Consortium on Energy Restructuring, Virginia Tech
1.3 Benefits of Distributed Generation
What are the Potential Benefits of DG Systems?
Consumer advocates who favor DG point out that distributed resources can improve the
efficiency of providing electric power. They often highlight that transmission of
electricity from a power plant to a typical user wastes roughly 4.2 to 8.9 percent of the
electricity as a consequence of aging transmission equipment, inconsistent enforcement
of reliability guidelines, and growing congestion. At the same time, customers often
suffer from poor power qualityvariations in voltage or electrical flowthat results
from a variety of factors, including poor switching operations in the network, voltage
dips, interruptions, transients, and network disturbances from loads. Overall, DG
proponents highlight the inefficiency of the existing large-scale electrical transmission
and distribution network. Moreover, because customers electricity bills include the cost
of this vast transmission grid, the use of on-site power equipment can conceivably
provide consumers with affordable power at a higher level of quality. In addition,
residents and businesses that generate power locally have the potential to sell surplus
power to the grid, which can yield significant income during times of peak demand.
Industrial managers and contractors have also begun to emphasize the advantages of
generating power on site. Cogeneration technologies permit businesses to reuse thermal
energy that would normally be wasted. They have therefore become prized in industries
that use large quantities of heat, such as the iron and steel, chemical processing, refining,
pulp and paper manufacturing, and food processing industries. Similar generation
hardware can also deploy recycled heat to provide hot water for use in aquaculture,
greenhouse heating, desalination of seawater, increased crop growth and frost protection,
and air preheating.
Beyond efficiency, DG technologies may provide benefits in the form of more reliable
power for industries that require uninterrupted service. The Electric Power Research
Institute reported that power outages and quality disturbances cost American businesses
$119 billion per year. In 2001, the International Energy Agency (2002) estimated that the
average cost of a one-hour power outage was $6,480,000 for brokerage operations and
$2,580,000 for credit card operations. The figures grow more impressively for the
semiconductor industry, where a two hour power outage can cost close to $48,000,000.
Given these numbers, it remains no mystery why several firms have already installed DG
facilities to ensure consistent power supplies.
Perhaps incongruously, DG facilities offer potential advantages for improving the
transmission of power. Because they produce power locally for users, they aid the entire
grid by reducing demand during peak times and by minimizing congestion of power on
the network, one of the causes of the 2003 blackout. And by building large numbers of
localized power generation facilities rather than a few large-scale power plants located
distantly from load centers, DG can contribute to deferring transmission upgrades and
expansionsat a time when investment in such facilities remains constrained. Perhaps
most important in the post-September 11 era, DG technologies may improve the security
of the grid. Decentralized power generation helps reduce the terrorist targets that nuclear
facilities and natural gas refineries offer, andin the event of an attackbetter insulate
the grid from failure if a large power plant goes down.
Environmentalists and academics suggest that DG technologies can provide ancillary
benefits to society. Large, centralized power plants emit significant amounts of carbon
monoxide, sulfur oxides, particulate matter, hydrocarbons, and nitrogen oxides. The
Environmental Protection Agency has long noted the correlation between high levels of
sulfur oxide emissions and the creation of acid rain. Because they concentrate the
amount of power they produce, large power plants also focus their pollution and waste
heat, frequently destroying aquatic habitats and marine biodiversity. On the other hand,
recent studies have confirmed that widespread use of DG technologies substantially
reduces emissions: A British analysis estimated that domestic combined heat and power
technologies reduced carbon dioxide emissions by 41% in 1999; a similar report on the
Danish power system observed that widespread use of DG technologies have cut
emissions by 30% from 1998 to 2001. Moreover, because DG technologies remain
independent of the grid, they can provide emergency power for a huge number of public
services, such as hospitals, schools, airports, fire and police stations, military bases,
prisons, water supply and sewage treatment plants, natural gas transmission and
distribution systems, and communications stations. Finally, DG can help the nation
increase its diversity of energy sources. Some of the DG technologies, such as wind
turbines, solar photovoltaic panels, and hydroelectric turbines, consume no fossil fuels,
while others, such as fuel cells, microturbines, and some internal combustion units burn
natural gas, much of which is produced in the United States. The increasing diversity
helps insulate the economy from price shocks, interruptions, and fuel shortages.
Table 1.1 Matrix of Distributed Generation Benefits and Services.

The benefits of distributed generation have been summarized well in a 2007 Department
of Energy report, which can be found at http://www.oe.energy.gov/epa_sec1817.htm.
(U.S. Department of Energy, The Potential Benefits of Distributed Generation and Rate-
Related Issues That May Impede Their Expansion [February 2007]). The matrix above
comes from that report (p. 1-11).
1.4 Usage of Distributed Generation
How Much are Renewable and DG Systems Used in the American Electric Utility
Sector?
Despite the immense environmental, technical, and financial promise of renewable
energy systems, such generators still constitute a very small percentage of electricity
generation capacity in the United States. Throughout the 1970s, some policy experts
expected renewable energy systems to be used for much more generation capacity than
they have. Dr. Arthur Rosenfeld, one of the five CEC commissioners serving from 2002
until the present, noted that President Carter had told him (during his presidency in the
late 1970s) that he expected renewable energy systems to reach 10 percent of national
electricity capacity by 1985. However, Carters expectation went unfulfilled: excluding
large hydroelectric generators, renewable energy technologies in 2003 comprised only
about 2 percent of the U.S. electricity generation mix.

U.S. Electricity Generation by Source, 2003


Electricity Generation by Fuel, 1970-2020 (billion kWh)
The relatively minor use of renewable energy systems has created a general attitude
among energy analysts, scholars, and laboratory directors that the technologies are not
viable sources of electricity supply. For example, Rodey Sobin, former Innovative
Technology Manager for the Virginia Department of Environmental Quality, argues that
in many ways, renewable energy systems were the technology of the future, and today
they still are. Ralph D. Badinelli, a professor of Business Information Technology at
Virginia Tech, explains that renewable energy technologies do not contribute
significantly to U.S. generation capacity because such sources have not yet proven
themselves Until they do, they will be considered scientific experiments as opposed to
new technologies. Similarly, Mark Levine, the Environmental Energy Technologies
Division Director at the Lawrence Berkeley National Laboratory, comments that despite
all of the hype surrounding renewable energy, such systems are still only excellent for
niche applications, but the niches arent large.
DG/CHP technologies have an only slightly better record. In 2004, the Energy
Information Administration characterized only 3.1 percent of electricity generation
capacity as commercial or industrial combined heat and power (33,217 MW out of 1.49
terrawatts [TW]). The EIA also estimated that in 2002 only 0.9 gigawatts (GW) of
distributed generation capacity existed in the United States. Similarly, the EIAs 2005
Annual Energy Outlook projected that CHP systems are not widely used in the electric
power sector, amounting to 0.053% of utility generation (197 billion kWh out of 3,700
billion kWh). Tom Casten, the Chair and Chief Executive Officer of Primary Energy, a
manufacturer of fuel processing cogeneration steam plants, notes that even though CHP
plants can reduce energy costs for industrial firms by over 40 percent, such plants remain
the exception instead of the rule.
Why Not Use More DG Technologies?
There are a mulitude of impediments to using DG technologies. A combination of social,
scientific, and technical impediments prevent a transition to a more friendly DG future.
Both proponents and opponents of DG technologies acknowledge that economic
considerations such as capital costs, utility preferences, and business practices tend to
deter people from investing in such technologies. DG systems are believed to have
higher, comparative capital costs (in dollars per kilowatt) than other generators, which
places many smaller, decentralized systems out of the price range of most residential
consumers. Moreover, entrepreneurs and business owners argue that the comparative
higher capital cost convinces them that investing in renewable or distributed systems is
too expensive and deviates from the core mission of their corporate goals. Even electric
utility managers generally shun renewable energy technologies, thinking that their power
output is more intermittent than their fossil-fueled and nuclear alternatives, thus making
them less viable providers of base-load and peaking power. Finally, since renewable and
distributed energy systems, by their very nature, are more diverse and context dependent,
transmission and distribution operators argue that they tend to be more difficult to permit,
monitor, interconnect, and maintain.
Furthermore, many analysts believe that the strong political support for DG technologies,
after the energy crisis of the 1970s, inflated expectations among the public that the use of
renewable energy resources would grow rapidly. Yet a number of unforeseen events
occurred: the Reagan administration shifted the energy policy of the country, fossil fuel
prices fell in the 1980s, and conventional technologies continued to improve. Advocates
of DG suggest that voters and politicians became disillusioned with renewable energy,
and relinquished whatever social capital they achieved after the energy crises to utility
managers and system operators. After the 1970s, when the country shifted completely
back into the fossil fuel paradigm, inconsistent political support for tax credits created
great uncertainty regarding DG technologies. This uncertainty deterred industry
investment in renewable and distributed energy systems. As a result, strong utility bias
became reflected in numerous state and federal regulations.
Moreover, deep and pervasive social attitudes prevent people from using more DG
technologies. The American electricity market shelters consumers from the true costs of
electricity generation. Thus, many Americans believe that they are entitled to cheap and
abundant sources of electricity, but they lack the necessary understanding of what needs
to occur so that they continue to have access to such a supply. As a result, renewable and
DG energy systems are often opposed not because they are a poor alternative to fossil
fuels, but because people simply dont comprehend why such technologies may be
needed. Far from becoming an unintentional side effect of the hub and spoke model of
electricity supply, some evidence suggests that public ignorance towards electricity is
an importantand plannedpart of the existing technological system. The lack of
public interest in the electricity sector allows utilities and system operators to more easily
maintain control over their system.
Thus, higher capital costs are believed to prevent investment in DG technologies;
lingering monopoly rules are blamed for continuing to favor utilities and fossil fueled
generators; advocates of DG complain that discriminatory practices have become
embedded in local regulations.

Where are the technical and scientific reasons? If one is going to claim both (social and
technical), shouldn't we make sure we make those transparent to the average reader?
When I read this it seems like the social is influencing the techno-sci but not in other
direction. It seems like someone was holding on to a more "classic" STS "paradigm." Not
knowing the science behind this makes it difficult for me to fill in the technosci aspects
holding DG back...I think that it is generally useful to term it as a story of preventing
more DG from happening as this enables those reading it to write the DG future.

1.5 Questions

1. How does DG differ from conventional power production?
2. Give the potential benefits of DG, where do you think the systems would have the
greatest advantage?
3. What do you see as the biggest impediments to DG?
4. How do you think these impediments could be overcome?
2. Energy, Power, and Environment.**
Electric energy is produced through an energy conversion process. In the process energy is
converted from a form such as potential energy, kinetic energy, or chemical energy to electricity.
This module, Energy, Power & Environment, covers the basic concepts of energy, power, and
work. The common methods of converting energy to electric power are discussed, as well as
many of the environmental effects of energy conversion. This module requires a basic
understanding of physics. This module aims to identify different energy sources, methods of
storage and conversion, and to articulate the relationship between electric energy and the
environment. The study questions and solutions for this module are included in the text of the
sections.
2.1 Work, Energy and Power
Work
Work U is defined as the product of the force F applied to a mass to move the mass a linear
displacement l .
U = Fl
If the displacement is not in the direction of F , then the work performed is the product of the
displacement and the component of the force along the displacement. If a is the angle that F
makes with l , then
U = Fl cos(a)
Work is measured in Joules (J). One joule is the work done by a force of one Newton moving a
body one meter.
1 J = 1 N m
Energy
Energy is the capacity of a body to do work. Energy is expressed with the same unit as work
(J=N m). There are several types of energy. We will be focusing on electrical, mechnical, and
thermal energy. All three types have typical units of expression and each will be described.
Electric energy is typically expressed with a different unit, the watt-second (W s), where
1 W s = 1 J
More commonly, electric energy is expressed in terms of kilowatt-hours (kWh), where
1 kWh = 3.6 x 10
6
J .
You have probably seen this expressed on your electricity bill. Recently, I saw a 1000kWh bill!
Two of the most important forms of mechanical energy are kinetic energy and potential energy.
These are especially important for electric energy generation, as will be discussed later. A body
possesses kinetic energy through its motion. An object of mass M (kilograms), moving with a
velocity u (meters per second), has a kinetic energy

A body possesses potential energy through its position to an accelerating field. Gravitational
potential energy results from the body's position in a gravitational field. An object of mass M
(kilograms) at a height h (meters) has a potential energy due to the acceleration of gravity g
(meters per second per second) given by
PE = Mgh J
Thermal energy is measured in calories (cal). One calorie is the amount of heat required to raise
the temperature of one gram of water at 15C one degree Celsius. It has been found
experimentally that
1 cal = 4.186 J
A common unit of thermal energy is the British thermal unit (Btu), which is equivalent to:
1 Btu = 1.055 x 10
3
J = 0.252 x 10
3
cal
Power
Power is the time rate at which work is performed, or the time rate of change of energy.
Therefore, instantaneous power p may be expressed as

where U represents work and w represents energy. The unit of power is the watt (W). One watt is
equivalent to
1 W = 1 J/s
Power engineering commonly uses the kilowatt (10
3
) and the megawatt (10
6
) multiples of the
watt. The power ratings of electric motors are generally expressed in terms of horsepower (hp),
where
1 hp = 745.7 W
It is important to be able to convert between the common units for work and power and the
International System (SI) of units. The example problem should help both with conversion and
concept clarification.
Example Problem
The average person needs 2000 calories of energy per day to live. The typical candy bar contains
230 calories. How many candy bars does a person need to eat per day to live? What is the daily
power consumption of a person in megawatts?




**This chapter of the education module was written by Jaime DeLaRee Lopez, full citation.
Next >>

2.2 Sources of Energy
Electricity is produced by converting energy from one form to electricity. The process used may
be a direct conversion process, where the energy source is converted directly to electricity. An
example of this is solar photovoltaic cells, which converts the energy found in solar radiation
directly to electricity. An indirect conversion process consists of converting energy from one
form, to an intermediate form, to electricity. Coal-fired generating plants are an example of the
indirect process, as the chemical energy released as heat by burning the coal is changed to
rotating kinetic energy by the steam turbine, and then the rotating kinetic energy is converted to
electricity. The majority of the electricity today is produced through an indirect energy
conversion process. Both indirect and direct processes use the following major sources of energy
for the production of electricity: fossil fuels, nuclear energy, solar radiation, and hydroenergy.
Fossil fuels include coal, petroleum, and natural gas. Fossil fuels are a finite, non-renewable
resource. They remain the primary source for the production of electricity. The combustion of
these fuels releases their chemical energy, which produces heat to power steam turbines. The
steam turbines power rotating electric generators, which turn kinetic energy into electricity. No
energy conversion process converts all the energy present in one form completely into the new
form. Since the production of electricity from fossil fuels involves several energy conversion
steps fossil fuel power plants inefficiently produce power (somewhere around 40%). Table 2.1
summarizes the energy content of fossil fuels and the estimate of fossil fuel reserves for the
world. We must also consider the environmental effects of electricity generation from fossil
fuels. For example, flyash, physical matter left after coal combustion, is harmful to human
(animal and environmental) health. It demands environmentally friendly disposal. Combustion of
fossil fuels also produces carbon monoxide, carbon dioxide, sulfur dioxide, and nitrous oxides
gases. These "greenhouse gases" contribute to acid rain and global warming effects.
Table 2.1
Estimate of Fossil-Fuel Energy Content and World Fossil Fuel Reserve

Fuel
Approx. Energy
Content (Watts/yrs)
Estimated Reserves
Coal 937/ton 7.6 trillion metric tons
Petroleum 168/barrel 2 trillion barrels
Natural Gas 0.036/ft
3
10
16
ft
3

Nuclear energy, just like fossil fuels, is a finite, non-renewable, energy source that uses an
indirect conversion process to produce electricity. There are two basic forms of nuclear energy,
fission and fusion. Since the fusion reaction has never been performed, only the fission reaction
is used to produce electricity. The fission reaction involves the splitting of the nuclei of a heavy
element. The heat output from this reaction powers a steam turbine, which drives a rotating
electric generator, just as in a fossil fuel power plant. Nuclear fission produces a much greater
energy content than fossil fuels. Uranium contains an energy content of approximately 10
10

Btu/kg, which is about one million times the energy content of fossil fuels. There is an estimated
reserve of approximately 426 million kg in the United States. Nuclear fission also comes with
heavy environmental costs and risks. The fission process leaves the the nuclear fuel, the reactor
vessel where fission occurs, and the steam pipes highly radioactive. Also, plant failures can lead
to the release of radioactive steam into the atmosphere or worse.
Solar radiation includes energy used directly as intercepted solar radiation, or indirectly as wind
and hydropower. Solar radiation is a renewable energy source. The average incident power at the
earth's surface is 182 W/m
2
, which corresponds to a daily average energy of 4.4 kWh/m
2
. Direct
use of solar power includes active types involving photovoltaic cells, and passive types using
radiation to heat solar collectors. Photovoltaic cells directly convert sunlight into electricity. The
best photovoltaic cells have efficiencies in the 14% to 17% range. Environmentally attractive,
photovoltaics introduce an emissionless, wasteful by-product minimized, extractionless energy
conversion process. However, the most efficient solar cells use gallium arsenide, a toxic
material. It seems that solar cells are too new for an understanding of the disposal requirements
and costs involved for worn out cells. In addition to this, photovoltaic solar cells generate direct
current, hence requiring inverting equipment to obtain the desired alternating current for most
large-scale operations. Solar collectors are normally incorporated into a solar thermal system,
converting sunlight into heat for various forms of use, including space heating, water heating,
industrial process steam, and electricity production. At present several factors limit large-scale
utilization of solar energy, including the cost of solar cells and solar collector-heat exchanger
systems, and the requirement of an adequate energy storage system to smooth out the daily
variation. Yet, sunlight is available everywhere making the use of solar radiation for energy
production non-site specific.
Wind energy is also a form of indirect use of solar radiation. Solar radiation produces wind by
heating the air. During the day, the air over land is heated much faster than air over water bodies
because the land absorbs much less sunlight, and the evaporation is less. The heated air over land
expands, becomes lighter, and rises. The cooler, heavier air over large water bodies moves in to
replace the lighter and warmer air, creating a horizontal motion of air. During the night the land
cools faster than water, the cool air moves seaward to replace the warm air that rises from the
surface of the water. Wind energy to electricity is an indirect energy conversion process. It
requires turbine type wind generators to transform the kinetic energy of the wind into rotary-
shaft motion and, in turn, this is converted to electrical energy. Wind turbine generators are site
specific. They require a sustained wind speed of 20 km/hour. To meet this requirement the bulk
of wind generators are located on mountain passes or the coast. Wind turbines have a maximum
possible efficiency of 59.3%, with a more common efficiency of around 40%. The significant
environmental problems associated with wind turbines are noise, aesthetics, and interactions with
birds.
Hydropower is also an indirect means of using solar power to produce electricity, since
hydropower uses the stream-flow part of the hydrological cycle. In a hydropower plant, the
potential energy of a mass of water in a reservoir a distance above the stream bed is converted to
kinetic energy by flowing through a hydraulic turbine. The resulting kinetic energy of the turbine
drives an electric generator. Hydropower is available wherever a suitable site exists having
enough stream flow, potential drop and area. Industrialized nations contain about 30% of all
hydropower potential, and are responsible for about 80% of all electricity produced from
hydropower. Asia accounts for 30% of hydropower potential, and produces only 7% of
electricity such produced. Africa accounts for 20% of hydropower potential, yet produces only
2% of electricity such produced. Hydropower is very attractive because it is a non-polluting
renewable resource. It can be very disruptive environmentally. The dam and reservoir effect the
normal ecology of the stream and the surrounding habitat by altering water use, changing natural
water flow cycles of the stream, and taking up land area for the reservoir. For example,
hydroelectric operation in the northwest United States, within the Columbia River basin, is often
constrained in order to protect the yearly river migration of salmon. Additionally, new dams may
often require the relocation of people and buildings.
Tidal energy uses the tidal flow of oceans to run a hydropower plant to produce electricity.
Basically, a dam encloses a tidal pool. The tidal pool fills during periods of high tide, then
empties during periods of low tide. The water flow into and out of the pool drives a reversible
hydraulic turbine. Since the turbine is reversible, the flow of water into and out of the bay
produces electricity. Therefore, tidal power is available twice during each 12h 25min tidal
period. The ideal sites for tidal energy have a large difference in tides.
Many energy sources produce this diverse range of potential electrical power. Non-renewable
energy source include nuclear and fossil fuels. Renewable energy sources range from direct solar
to the indirect solar sources and geothermal energy (although it was not discussed). Several of
these are site specific, such as hydropower, wind power, and tidal power. Each source possesses
both benefits to general well-being and costs to the environment and future burdens.
The next section describes the efficiencies of different electrical systems and the costs of electric
production due these efficiencies.

<< Back Next >>
2007 Consortium on Energy Restructuring, Virginia Tech
2.3 Efficiency & Cost
Electrical energy results from converting one form of energy, generally mechanical energy, into
another form, electricity. Not all of the original energy input into an energy converter changes to
electricity. This leads to the idea of efficiency, or a rating of how well a converter changes
energy from one form to another form. Efficiency has different meanings in different types of
physical systems. In mechanical systems, use is made of thermal efficiency and mechanical
efficiency, which describe the efficiency of the two corresponding phases of a given process. In
the electrical systems that will be discussed, efficiency is defined as

This can also be expressed in terms of mechanical and electrical losses in either energy or power
terms as

The energy use or efficiency of an electric machine is becoming increasingly significant. It is one
of the more important design criteria today. Approximate maximum efficiencies of various types
of energy converters are shown in Table 2.2.
Table 2.2 Approximate maximum efficiencies of energy converters.
Energy Converter Efficency, percent
Synchronous electric generator greater than 95
Large electric motor 90
Small electric motor 65
Fuel cell 60
Steam turbine 45
Steam power plant 40
Diesel engine 38
Automobile engine 30
Solar cell 10
The efficiency of energy converters has a direct effect on the financial cost to produce electricity.
More efficient converters use less fuel or less space to produce the same amount of electricity.
This also decreases the associated financial costs of the system. Current standards necessiate
maximizing efficiency and minimizing the financial costs for electric systems.
In the next section, Electromechanical Energy Conversion, general conversion processes will be
discussed. This section is coupled with a section on Direct Energy Conversion, where energy
conversion occurs without using a mechnical turbine to electricity. Following these sections,
conversion types are discussed more specifically. These sections lead us into more of the
physical and engineering concepts behind electricity.


<< Back Next >>

2007 Consortium on Energy Restructuring, Virginia Tech
2.4 Conversion and Converters
The following sections outline the basic concepts behind energy conversion and specific
converters. There are two general types of energy conversion, electromechanical and direct
conversion. In the proceding, the energy source for conversion provides a logical way to group
the converter types discussed.
2.4a Conversion.
2.4a.1 Electromechanical Energy Conversion
The most common method for bulk power generation is by rotary generators located in electric
power stations. These generators are electromechanical energy converters, also known as electric
machines. In practice, a mechanical prime mover coupled to the generator rotates the electrical
conductors constituting the generator windings in a magnetic field, inducing a voltage in the
generator windings. These windings supply electrical load on the generator. Conversely, if a
current carrying conductor is placed in a magnetic field, the conductor experiences a force
according to Ampere's law. In general, electric machines are reversible, capable of operating
both as generators and motors. There are three major types of rotating electric machines: dc
commutator, induction, and synchronous machines.


Figure 2-1. Prime Mover Driving an Electric Generator

The source of mechanical energy for a rotating electric generator is known as the prime mover.
The prime mover is directly coupled to the generator. Energy sources for prime movers are
thermal, hydro, and wind. The prime movers normally are turbines, but some thermal units use
internal-combustion engines. A turbine is mechanical device that is forced to rotate by the
pressure of a gas (such as steam for thermal units or air for wind units) or fluid (such as water for
hydro units).
An electromechanical energy converter converts mechanical energy into electrical energy, and
vice versa. A generator converts energy from mechanical to electrical form, and modulates in
response to an electric signal. A motor converts energy from electrical to mechanical form, and
modulates in response to an electrical signal. Rotating machines, if lossless, operate on the
principle of electromechanical power equivalence as given by
P
mech
=Tw
m
= vi=P
electrical

where T is mechanical torque (N-m), w
m
is mechanical angular velocity (rad/s), v is
instantaneous electrical volts (volts), and i is instantaneous electrical current (amperes). Electric
generators are governed by Faraday's law of electromagnetic induction: an electromotive force
(emf) is induced in a conductor "cutting" magnetic lines of flux. Specifically if a conductor of
length l (m) moves with a velocity u in a uniform magnetic field B (tesla), such that l , u , and B
are mutually perpendicular, then the induced emf in the conductor is given by
E = Blu

Figure 2-2. Relative directions of E , u , and B
The following sections cover the basic types of electromechanical energy converters.
Sample Problems
2.1. An ideal energy converter develops 500 N-m of torque while running at 3600 rpm. If the
input voltage is 1200 volts, determine the input current for this generator.

2.2. Calculate the power output of the ideal energy converter.

2.4a.2 Direct Energy Conversion
Direct energy conversion devices convert solar, thermal, chemical, and nuclear energies into
electricity without involving a rotating or reciprocating mechanical prime mover. Various direct
energy converters include:
Fuel cells and batteries
Photovoltaic, photoelectric, and electrostatic generators
Thermionic, thermoelectric, and ferroelectric generators
Magnetohydrodynamic generators
Piezoelectric generators
It is projected that, of all the direct energy converters, only solar power and fuel cells will
contribute to the production of any significant amount of electrical power in the near future.
During the decade 1981-1990, in the United States, considering wind, solar, geothermal,
cogeneration, solid waste, wood waste, and fuel cells as alternatives to conventional methods of
generating electrical power, no more than 12 percent has been generated from solar and fuel cells
in terms of installed capacity. Of these methods, only a fuel cell accomplishes energy conversion
directly.
Understanding how direct energy conversion devices operate requires knowledge of the laws of
thermodynamics. The first law of thermodynamics states that energy is conserved:
"The increase in the internal energy of a thermodynamic system is equal to the amount of heat
energy added to the system minus the work done by the sytem on the surroundings."
If heat Q is added to a system to increase its energy by an amount U (where U is the internal
energy of the system) resulting in mechanical output W (where W is the work done by the
system on the surroundings), then
Q = U + W

Figure 2-3. Relative directions of E , u , and B
Now, considering a steady-flow system, such as the one shown in Figure 2-3, where there is a
change in the pressure p , and volume V , of the system, then this equation becomes
Q = U + W + ( pV )
From Figure 2-3 , Q = Q
1
- Q
2
, U = U
1
- U
2
, and ( pV ) = p
1
V
1
- p
2
V
2
. The quantity ( pV
) reflects the difference in work required to move a unit mass into and out of the system. Both
U and ( pV ) are unique forms of stored energy in that they depend on the state variables p , V
, and T ( T is the temperature - recall pV=NRT, the ideal gas law, where N and R are constants)
of the system. Usually, the sum ( U+pV ) is considered as a single property H , known as the
enthalpy of the system.
H = U + pV
The second law of thermodynamics states that it is impossible to completely convert heat into
work (Lord Kelvin). According to the second law of thermodynamics, continuous conversion of
heat into work by a device receiving heat from a source at a temperature T
1
is possible only when
part of the heat is rejected into a sink at a lower temperature, T
2
(T
2
< T
1
). As a consequence, an
energy conversion process involving heat must have an efficiency of less than 100 percent, even
without loss. This ideal or maximum efficiency is called the Carnot efficiency,
c
, and is given
by

where T
1
and T
2
are absolute temperatures for a reversible cycle. The Carnot efficiency of a
modern large fuel-fired steam-turbine generating unit is about 61 percent. In an electric power
station the mechanical energy of the turbine is converted to electrical energy by an electric
generator, and the overall efficiency from fuel to electrical power can scarcely exceed 40
percent. Such efficiency limitations exist on all closed-cycle heat engines. On the other hand, a
fuel cell, which converts chemical energy into electrical energy, is not a heat engine and is not
subject to the Carnot efficiency limitation. Rather, the efficiency of a fuel cell is limited by the
thermodynamic quantities related to the fuel oxidation reaction. Fuel cells have efficiencies in
the range of 80 to 99 percent.
Heat flow is a function of temperature difference. If a quantity of heat is divided by its absolute
temperature, the quotient is called the entropy of the system. Entropy is a measure of degradation
of energy through usage. If heat Q enters a system at an absolute temperature T , then the
change in entropy of the system for a reversible process is given by
where S is entropy difference.
Note. The second law of thermodynamics can also be stated as: In an isolated system, a process
can occur only if it increases the total entropy of the system ( Rudolf Clausius).
In a reversible process, the state variables p , V , T , and U are such that if the process is reversed,
these state variables follow the same values in reverse order. Thus reversibility is measured by
entropy S . The second law of thermodynamics can be expressed as

where ">" is for a reversible process. For a reversible cycle the net change in entropy is zero.
Consider the thermodynamic limitation of energy conversion in an oxidation reaction, such as
that in a fuel cell. Gibbs free energy, G , is defined by
G = H - TS
where H is the enthalpy of the system at an absolute temperature T and entropy S. The change in
the Gibbs free energy, G , for an oxidation reaction is equal to the maximum amount of
electrical energy obtainable from the reaction. For an isothermal reaction,
G = H - T S
where T S is a measure of the heat absorbed by the system during a reversible change, and is
the unavailable energy.
2.4b Energy Conversion Types
2.4b.1 Electromechanical Energy Converters
Thermal
Thermal prime movers are the most common prime movers used to generate electricity. These
include steam turbines, gas turbines, gasoline engines, and diesel engines. In all cases, the prime
mover is a stationary device that rotates electrical conductors through a magnetic field, as
discussed in Section 2.4. To produce the steam necessary to drive steam turbines, coal, gas, or oil
is burned in boilers. Or for nuclear power plants heat resulting from nuclear fission is used to
produce the steam. Steam at maximum possible pressure and temperature is used to ensure
maximum efficiency of operation of the turbine. Turbine units with a rating of 500 MW and
above are common, as large turbine sizes result in lower per MW capital costs.
Gas turbines are essentially jet engines, driven by hot gases. These gases are produced in a
combustion chamber, in which a continuous combustion of injected fuel oil occurs in the
presence of compressed air simultaneously delivered to the chamber. In this regard, a gas turbine
is an internal-combustion engine, much like an automobile engine. The main advantage of a gas
turbine is its capability to quickly start and take up load. Such features are required to meet
sudden power system peak loads. Under normal continuous operating conditions, a gas turbine is
less economical than a steam turbine of the same rating.
Of the other internal-combustion engines, gasoline engines are seldom used to drive large (50
kW or more) electric generators, and diesel engines are usually used only for isolated units rather
than in central stations. The diesel engine differs from the gasoline engine mainly in the
mechanism of combustion. In a diesel engine heat is generated in the cylinder to ignite the fuel.
In a gasoline engine electric spark ignites the air-fuel mixture in the cylinder. Consequently, a
diesel engine has a higher compression and is bulkier, heavier, and more expensive than a
gasoline engine, but is relatively less expensive to operate.
Nuclear
Nuclear prime movers, as discussed above, are really only thermal prime movers. The heat
developed as a byproduct of the nuclear reaction is used to produce steam to drive a turbine. The
two basic forms of nuclear energy are fission and fusion. The fission reaction involves the
splitting of the nuclei of a heavy element, while the fusion reaction involves combing the nuclei
of light atoms. At the present time, only the fission reaction is used to produce electricity.
Uranium 235 (U 235 ) is the most suitable element for fission from an environmental standpoint.
However, fuel supplies are limited, since only 0.7% of all naturally occurring uranium is the U
235 isotope. Breeder reactors transform uranium 238 into fissionable plutonium 239 by
absorbing neutrons, or transform thorium 232 into fissionable uranium 233. In a breeder reactor,
1g of uranium produces 8.1 x 10
10
J of heat, which approximates the heat produced by 2.7 metric
tons of coal. Figure 2-4 shows a simple diagram of a fission reaction.

Figure 2-4. Fission Reaction
Fusion is the process that takes place on the sun, and it is a scientifically feasible and attractive
method to generate electricity. The main advantage is that the fusion reaction can use deuterium
as a fuel. Deuterium is an abundant (one atom of deuterium per 6,700 atoms of hydrogen)
isotope of hydrogen, and has a low cost to separate. Many engineering problems related to the
fusion process remain to solve. For fusion to occur, the nuclei must form plasma (atoms heated
to such a high temperature that they are stripped of their electrons). The plasma must be
contained or confined in a region of space such that the plasma density is high. Furthermore, the
plasma must be contained long enough for the fusion process to take place. Heating and plasma
confinement are major engineering problems for a fusion power generation. Therefore, no fusion
power plants exist, and implementation of a practical fusion power generating station does not
seem likely in the near future.
Wind
Wind turbines extract the kinetic energy present in the wind, and convert it to rotary shaft
motion. The shaft motion transmits power to generators by gearboxes, belts and pulleys, roller
chains, or by hydraulic transmissions. The power that can be extracted from the wind is given by

where r is the density of air (1.2 kg/m
3
), A is the wind turbine blade area in m
2
, and v is the
wind velocity in m/sec. The velocity of the wind, and the size of the turbine blade, determine the
amount of power available for a specific wind turbine. Since the velocity of wind increases with
height above the ground, the height of the wind turbine has an important effect on the power
extracted. The effectiveness of a wind turbine is measured by the power coefficient, C
p
. C
p
is
defined as

The maximum value of C
p
is 59.3%. Wind tunnel tests have shown that the ideal horizontal axis
two bladed wind turbine normally has a power coefficient of 42.0%. Two major factors
characterizing the design of a particular wind turbine are tip-speed ratio and blade pitch angle.
The tip-speed ratio, b, is the ratio between the blade tip speed and the wind speed. Modern wind
turbine generators have a b of 5.0 or more. A low b is associated with low efficiency, but an
ability to start under load. A higher b reflects higher efficiency, but inability to start under load.
The relationship between blade pitch angle and C
p
is not straightforward, as it is controlled by
lift and drag forces, choice of the aerofoil section, the air flow velocity over the blade, and the
blade width. Note that C
p
is not the same as the efficiency of the wind turbine, as rotational
losses are not considered in C
p
.

Figure 2-5. Possible positions of horizontal axis wind turbines
Wind turbines can be classified into two major groups: horizontal axis wind turbines and vertical
axis wind turbines. Horizontal axis wind turbines, the most common type, uses blades with a
horizontal axis of rotation. Under steady state conditions each blade section experiences a
constant angle of attack during one revolution. (The angle of attack is the angle between the
direction of wind flow and the direction of a line drawn parallel to the blade). Therefore it is
extremely important that the wind turbine yaw (or turn) maintain alignment with the direction of
the wind. Also, by turning the face of the blade away from or into the wind, the lift on the blade
(and thereby the velocity of the blade) can be increased or decreased. The horizontal axis wind
turbines may be either upwind or downwind types, as shown in Figure 2-5 . Small generators use
upwind horizontal axis wind turbines because the tail vane keeps the blades pointed into the
wind. The protection system in case of high winds is easily designed. Downwind turbines are
used for large horizontal axis wind turbines. Small downwind turbines have a natural tendency to
turn and align with wind flow, while large ones are steered by pilot wind vanes. The blades of
downwind turbines are designed to cone for protection against high winds. However, the tower
acts as a barrier to the windstream, forcing the blade to undergo stress transients each time it
passes the tower. Advantages of horizontal axis wind turbines include a high hub height, due to
the tower, and their ability to self-start. A disadvantage is that the generator and gearbox are
difficult to access, since they are mounted on the tower.
Vertical axis wind turbines, as shown in Figure 2-6, use blades with a vertical axis of rotation.
Therefore, the angle of attack experienced by each blade varies continuously through one
revolution. This means the vertical axis wind turbine has no need to yaw with changing wind
directions. Other advantages of the vertical axis wind turbine include a higher operating speed
than the horizontal axis wind turbine, and also that the generator and gearbox are located on the
ground. However, the vertical axis wind turbine is not self-starting, requiring an auxiliary means
of starting such as an electric motor.
Much experimenting has shown that large wind turbines (>2.5 MW) are impractical, due to the
cost of a tower large enough and inability to make reliable blades large enough. Whereas small
wind turbines may be practical for small standalone applications, they are not useful to replace
bulk power on the utility grid. For bulk power, many wind turbines are installed in groupings
known as "wind farms", usually having a total capacity of 500 MW or less, effectively turning a
collection of wind turbines into a large power station. Since the wind is a fluctuating energy
source, standalone wind turbines must have an energy supply during low wind periods; batteries
are often used to store electricity for this purpose. Popular home systems consist of a roof-
mounted wind turbine, a solar panel, and a battery storage system. Grid connected wind systems
cannot typically be used in "capacity planning" for serving peak loads because of this since there
is no guarantee that peak will not coincide with a low-wind period.

Figure 2-6. Vertical Wind Turbines
Hydro
Hydroelectric power plants and tidal power plants both use hydraulic turbines as prime movers,
which convert the potential energy of an elevated body of water to rotating kinetic energy. There
are two basic types of hydroelectric power plants: run-of-the-river plants and reservoir plants. As
its name implies, run-of-the-river plants are built so that the turbine blades are simply turned by
the water as it flows in the river.
Reservoir plants
But most hydro plants are reservoir type plants, which means there must be a dam to regulate the
water flow and add height to the source of the water. The powerhouse contains the hydraulic-
mechanical works consisting of turbines, the upstream waterways (penstock) carrying water from
the reservoir to the turbine, and the downstream discharge channels. It also contains the electric
generators. The height the water falls through the penstock is called the head. The power
available at the turbine is
P = gHQ kW
where g is the acceleration due to gravity, H is the head of the water in meters, and Q is the flow
rate of water through the turbine in cubic meters per second. Since a dam converts potential
energy into kinetic energy

giving

where v is the velocity of water in meters per second through the penstock. The velocity of the
water through the penstock can be converted to flow rate in cubic meters per second through

The power produced is proportional to the head and the flow rate. Dams are roughly classified
into low head (6 to 30 m), medium head (30 to 200 m), and high head (above 220 m). To
transmit water downstream, low-head plants utilize dams, and high-head plants use penstocks.
There are three types of hydraulic turbines available. Kaplan turbines are used for heads up to 60
m; Francis turbines are used for heads from 30 to 300 m, and Pelton wheels, for heads larger than
90 m. Maximum efficiencies of hydraulic turbines are between 85 and 95 percent. Francis
turbines of ratings exceeding 500 MW have been built and Pelton wheels of ratings of 40 MW
are in use. Hydraulic turbines can be started almost instantaneously from rest, and they have the
obvious advantage that no losses are incurred when at a standstill. Thus, working in parallel with
thermal power stations, hydroelectric plants can meet peak loads at minimum operating cost.
Pump-storage
The pump-storage unit represents a novel use of hydroelectric facilities where the plant may act
in either a generating mode or a pumping mode. In the generating mode, the water flows from
the upper reservoir to a lower reservoir, and the synchronous machine operates as a generator. In
the pumping mode, the synchronous machine operates as a motor, and water is pumped from the
lower reservoir to the upper reservoir. Such a system of moving water back and forth between
reservoirs seems strange, since there exist losses in such a system. The reason it is done is to take
advantage of the variation in the demand of the interconnected power system. When the load is
very heavy, the pump-storage facility is operated in the generating mode to help supply the
demand. But when the load is very light, the facility utilizes the least expensive plant in the
interconnection as a supply and is operated in the pumping mode. In this way, the cost of
operating the unit in the generating mode is equal to the cost of the least expensive unit in the
interconnection plus the cost of losses. This cost must be less than the cost of the most expensive
plant in the interconnection in order for pump-storage to be a good option.
Tidal power
Tidal power is essentially a specific form of hydro power, and therefore uses basically the same
equipment as a regular tidal station. The difference is in the available power to extract from the
tide. A reversible hydraulic turbine is used so the inflow and outflow of the tide can generate
electricity. The theoretical extractable energy for each tidal period is given by

where r is the density of water (1000 kg/m
3
), g is the acceleration due to gravity, A is the area of
the tidal pool, and R is the range of the tide. Average power for a tidal period is calculated by
dividing the extracted energy by the tidal period of 12h 25min, resulting in

Solar
As discussed previously, solar radiation may be connected directly to electricity, or it may be
converted to another form of energy to convert to electricity. This second method generally uses
solar radiation to heat a fluid. This heated fluid (which may be stored) is then run through a heat
exchanger to produce steam, which is converted to rotating kinetic energy in a steam turbine.
The turbine runs an electric generator. The fluid used may be water, or in some cases is a molten
salt such as nitrate salts or sodium. A central receiver system uses a field of sun-tracking mirrors
to concentrate sunlight onto a tower-mounted receiver. A parabolic dish system employs a point-
focusing reflector to concentrate sunlight at the focal point where a heat engine or a heat transfer
device is placed. The heat engine produces electricity directly, while the heat transfer device
produces a heated fluid.
2.4b.2 Direct Energy Converters
Fuel Cells
A fuel cell directly converts a hydrogen-rich fuel into electricity using a highly efficient
electrochemical process. The fuel cell uses the reverse electrolysis process to produce water and
electricity from hydrogen and oxygen. The hydrogen is normally produced from natural gas or
methanol. Chemical energy is stored outside the fuel cell, and the reaction products are
commonly rejected from the cells.

Figure 2-7. Fuel Cell
A fuel cell is shown schematically in Figure 2-7. The reverse electrolysis process is performed
by controlled continuous chemical reactions at the anode and cathode of the fuel cell. In a
hydrogen-oxygen fuel cell, hydrogen is the fuel and oxygen is the oxidant. The electrodes are
porous and are connected to the load. At the anode, the hydrogen fuel splits into positive
hydrogen ions and electrons as expressed in the following reaction:

After flowing through the load circuit, the electrons and the hydrogen ions combine with oxygen
at the cathode, as expressed in the following reaction:

The overall reaction in the fuel cell is given by

The oxidation product is water, which is removed from the cell. The energy change theoretically
available from the fuel cell is G , the Gibbs free energy change. The ideal electromotive force
(emf) E of a fuel cell is given by

where the Gibbs free energy is in kilocalories per gram mole, and n is the number of electrons
transferred per molecule of fuel oxidized. The net amount of energy liberated is the change in
enthalpy H , or the difference between the enthalpy of the reactants and the enthalpy of the
products.
For the standard fuel cell reaction, the value for the Gibbs free energy is G= -56.7 kcal at 25C
and the value for the enthalpy is H= -68.3kcal. The maximum possible emf for a fuel cell can
be calculated by

Electrode resistance, polarization of the electrolyte, and depletion of the electrolyte limit the
actual voltage produced. The ideal efficiency is the useful work per unit enthalpy change at a
constant temperature and pressure. Since the useful work is the Gibbs free energy, which means
that 17 percent of the energy is lost as heat.
Fuel cell power plants consist of more than just the fuel cell itself. A fuel processing section
converts natural gas or other hydrocarbon fuels into a hydrogen rich fuel. This is normally
accomplished through a steam catalytic reforming process. This fuel is then fed to the power
section, where it reacts with oxygen in fuel cells to produce direct current (dc) electricity, and
by-product heat as hot water or steam. Individual fuel cells are connected in series in a fuel cell
"stack" to produce usable amounts of electricity. The power conditioning section converts dc to
utility grade alternating current.
Fuel cells are classified by the type of electroltye used, and by their operating temperature. Low
temperature fuel cells operate in the range of 60C to 200C, and include phosphoric acid fuel
cells, alkaline fuel cells, and solid polymer fuel cells. High temperature fuel cells operate in the
range of 650C to 1000C, and include molten carbonate fuel cells and solid oxide fuel cells.
Fuel cells are currently used for special-purpose applications. For example, phosphoric acid fuel
cells were used on the Apollo spacecraft to provide electrical power, and alkaline fuel cells are
being tested to power cars and buses. Fuel cells have a lower environmental impact than normal
generating plants, since there are no emissions from the fuel cell itself, other than heat. Large
fuel cells are well suited for industrial cogeneration, because they produce steam as a by-product.
However, fuel cell plants are expensive to build, as they cost about $2000-$4000/kW. An 11
MW molten carbonate fuel cell power plant went into operation in California in 1996, which is
the first operating fuel cell power plant in the United States.
Batteries
Batteries are similar to fuel cells. They directly produce electricity as an output from a chemical
reaction. Unlike a fuel cell, where the reactants are continuously supplied, a battery stores the
reactants. The reactants are generated by electrical recharging of the battery.
The ability to recharge batteries with electricity makes them the most common form of energy
storage. Batteries are currently used in electric utility systems to provide backup power for
control and relay systems in substations. If the substation itself loses power, the battery system
ensures that the control systems in the substation continue to operate. Batteries are also used as
to provide reserve power for special cases such as standalone solar units and wind applications.
During times when the energy converter is operating, these units supply their load and recharge a
battery. When the units are not operating (such as nighttime for a solar unit), the battery supplies
the load. A major new use for batteries is as the power source for electric vehicles. The potential
growth of electric vehicles is leading to much research into battery technology.
The most common type of battery is the sealed lead-acid battery. Sealed lead-acid batteries are
very reliable. They also "recuperate," meaning the terminal voltage and energy density increase
slightly over a small time interval following a discharge characteristic. This means a lead-acid
battery will give a greater discharge capacity for a series of intermittent discharge cycles than for
a discharge at constant current of magnitude equal to the average of the intermittent currents.
This is a highly desired feature for electric vehicle applications. The lead-acid battery has an
average life of 400 cycles.
The nickel-iron battery is used where reliability and long-lived operations are needed. However,
it has a high operating temperature rise, high water consumption, and poor charge-acceptance in
comparison to lead-acid batteries.
The nickel-cadmium battery is widely used in vented or sealed form. It has a very high power
and energy density compared to the lead-acid battery, and can operate in a wide range of
temperatures. However, uneven cell voltages during charging can result in excessive gasing, cell
voltage reversal, and cell damage. The initial cost of this battery is very high, and requires
special manufacturing techniques for cadmium.
The nickel-zinc battery is a new type of battery still being developed. It is attractive due to its
potentially high energy density. This battery has an unusual charging problem due to the
instability of the zinc electrode, and current prototypes require careful hand re-charging.
The zinc-bromide battery has an energy density of 60 to 70 Wh/kg with a 70% energy efficiency.
Bromine spills pose a safety problem, so this battery is being developed for bulk energy storage
in electric utility networks. The battery requires a complete discharge every 5 to 10 cycles to
strip the zinc off the negative electrode.
The sodium-sulfur battery operates at around 300C, and individual cells have had lifetimes of
up to 6000 cycles. The sodium-sulfur battery uses a solid electrolyte made of thousands of tiny
hollow sodium-ion-conduction glass fibers giving it an energy efficiency in excess of 80%. This
electrolyte is made very inexpensively in large quantities, and it is highly resistant to impact
shock and vibration breakage. Energy densities of 125-150 Wh/kg are expected from this battery.
However, fires or explosions are possible if sodium leakage (due to battery damage) occurs.
Table 2.3 gives a summary of performance and expectations for the batteries discussed above.
Table 2.3
Performance of Various High-performance New Type Batteries
Energy Density Output Density Life (Cycle)
(Wh/kg) (W/kg)
Battery Current -> Future Current -> Future Current ->
Future
Sealed Type Lead
Acid
35->45 150->200 400->1000
Ni-Cd 50->60 160->200 500->1000
Ni-Fe 50->60 160->200 800->1000
Ni-Zn 70->85 160->220 200->500
Zn-Br 70->90 90->120 700->1000
Na-S 100->120 130->150 250->1000
Photovoltaic Energy Conversion
Photovoltaic systems, also known as solar cells, directly convert sunlight to electricity. In an
energy and environment conscious society the clean and seemingly inexhaustible source of
energy from photovoltaics provide an attractive option. For remote lighting and communications,
photovoltaics with battery backup provide the most cost-effective source of electricity.

Figure 2-8. Photovoltaic Effect
The photovoltaic effect is most prominent in various semiconductors. Most commercial solar
cells are made of crystalline and amorphous silicon materials. When sunlight strikes the solar
cell, part of the light spectrum imparts enough energy to create electron-hole pairs in the
semiconductor material. A potential barrier in the cell is set up by forming a junction between
dissimilarly-doped semiconductor layers. This separates the light-generated carriers (i.e.
electrons and holes) resulting in an induced voltage of about 1/2 volt. The available current is a
function of cell area and light intensity. The electricity is collected and transported by metallic
contacts placed on both surfaces of the cell. Photovoltaic cells are formed into modules by
connecting them in series and parallel in order to get more current and voltage. For even greater
power, modules can be interconnected in larger groups to form arrays. The dc electricity
generated by the solar cell array is usually passed through a power conditioner for voltage and
power regulation, and conversion to alternating current.
Commercially available photovoltaic modules can convert sunlight into electricity with
efficiencies ranging from 5% to 15%. The cost of photovoltaic cells has dropped from $1000 per
peak watt in the 1950's to under $5 per peak watt, so that they can produce electricity for as little
as 25 to 30 cents per kilowatt-hour.
Thermionic Energy Conversion
In a thermionic converter, heat energy is converted to electrical energy by thermionic emission,
whereby electrons are emitted from the surface of certain metals when the metals are sufficiently
heated. The three major components of a basic thermionic energy converter are the thermionic
emitter, the collector, and the working fluid, which may be an electron gas or a partially ionized
plasma. A thermionic energy converter is shown schematically in Figure 2-9. The input heat, Q
in
,
heats the emitter, and electrons are emitted. The cold collector receives some of these electrons
at an output heat, Q
out
. The difference (Q
in
-Q
out
) is the energy that drives the electrons through
the external circuit, and appears as electrical energy. The collector is cooled to remove the output
heat.

Figure 2-9: Thermionic Energy Converter
Vacuum converters, which have electron gas as their operating fluid, operate in the range
1200K to 1400K. They typically produce 1 W/cm
2
at an efficiency of 5 percent. Low-pressure
converters produce 10 W/cm
2
at an efficiency of 10 percent, operating at emitter temperatures up
to 2300K. High-pressure converters deliver 40 W/cm 2 at an efficiency of 20 percent, operating
at emitter temperatures up to 2200K.
Principal applications of thermionic energy converters are in regions not easily accessible, such
as outer space, undersea, and polar regions. The two important heat sources are the sun and
nuclear reactors.

Thermoelectric Energy Converters
Thermoelectric energy converters, or solar thermal systems, convert sunlight into heat for
various forms of end-use. One of three different conversion cycles may be used in a
thermoelectric system. The Seebeck effect results when the junctions of a loop made of two
different materials are at different temperatures such that an electromotive force (emf), and
consequently, a current is produced in the loop. If T
h
is the hot junction temperature, T
c
the cold
junction temperature, then V
12
, the open circuit voltage, is related to the temperature difference
by the Seebeck coefficienct a s represented by

When a current is passed through a loop of two different materials, one junction becomes hot and
the other junction becomes cold. This is the Peltier effect. The Peltier coefficient a P represents
the ratio of heat change at the junction to the current flow as shown by

When a current flows through a conductor in which a temperature gradient exists, heat is either
liberated or absorbed, depending on the direction of current flow. This is known as the Thomson
effect. The Thomson coefficient a T is the ratio of heat change per unit of current flow to the
local temperature, as shown by

These three effects are reversible and are interrelated.

Figure 2-10: Flat Plate Solar Collector
Solar collectors make use of one of these effects. There are two types of collectors - flat plate
and focusing. Flat plate collectors, illustrated in Fig. 2-10, do not use any device for
concentrating the sun's rays. These collectors still function when clouds cut off direct sunlight.
This advantage, along with their favorable cost, allows flat plate collectors to be used for low
temperature heating of up to 100C. Simpler flat plate collectors hold all the water that is to be
heated, while more advanced ones heat only a little water at a time. The heated water is kept in a
separate storage tank, to cut down heat losses from the fluid. Common uses of flat plate
collectors are for solar water heating, solar space heating, and solar cooling.
Focusing solar collectors, illustrated in Fig. 2-11, allow the use of much higher temperatures than
available from the best flat plate collectors. A focusing collector comprises a concentrator and a
receiver. The concentrator is the optical system that directs the sun's rays to the receiver. This
could be a parabolic dish with a point focus or a parabolic trough with a linear focus. Focusing
solar collectors are often used in systems that produce bulk electricity. They first concentrate the
incoming sunlight, convert it into heat, and convert the heat into electricity.

Figure 2-11. Focusing Solar Collector

Now, that you have been exposed to these key concepts (energy sources, efficiency, and
conversion/converters), you are ready for the next sections on energy storage and load. Energy
storage is an essential feature of electrical systems. Energy storage must be coordinated with the
load on the system. These concepts will be further explained in the next section.
<< Back Next >>
2007 Consortium on Energy Restructuring, Virginia Tech
2.5 Storage & Load
In the next two subsections delineate both different methods of energy storage and load growth
as well as the basic concepts behind energy storage and load growth.
2.5a Energy Storage
Electric energy production requires the conversion of energy into electricity. However,
conversion processes such as solar, wind, and hydro rely on a fluctuating fuel source. In these
cases, the power system must have some energy storage capability to overcome the fluctuations
in the energy supply. In other cases energy storage provides a means for harnesses excess energy
production, for example utilities produce more excess electricity at night. Energy storage
normally occurs through a conversion process from electrical energy to another form of potential
energy. Options for large scale energy storage include batteries, superconducting coils,
flywheels, and pumped storage.
As previously described, batteries provide a general solution to the conversion problem in remote
systems in conjunction with solar and wind sources, or as a backup in case of utility failure.
Current research includes using battery systems in the utility grid for bulk power storage.
Applications where batteries are being considered for utility power systems include load
levelling; voltage, VAR, and frequency control; and spinning reserve. Batteries provide quick
response time: response to load changes occurs in about 20 milliseconds.They are also quiet and
nonpolluting, making them ideal for installation in suburban areas, close to load centers.
A type of energy storage to consider is Superconducting Magnetic Energy Storage (SMES). This
type of energy storage involves converting off-peak power direct current and feeding it to a
doughnut shaped coil of superconducting wire. The coil is installed in a trench and kept at
superconductive temperature by a refrigeration system. With this process the unit can store and
discharge energy at an efficiency of greater than 90%, and charge in less than 20 milliseconds.
However, this system is very expensive, and some engineering problems related to
superconductors must be solved. In addition to these obstacles, SMES contains unknown health
effects due to the large magnetic field.
Flywheels are rotating wheels used to store kinetic energy, much like a spinning top. Electricity
is used to "wind" the wheel up through a system of gears. The flywheel then delivers rotational
energy to power an electric generator until friction dissipates it. The sum of the kinetic energy of
the individual mass elements that comprise the flywheel equals the energy stored. The kinetic
energy of a flywheel is given by
where I is the moment of inertia (the ability of an object to resist changes in its rotational
velocity), and w is the rotational velocity in rpm. The moment of inertia is defined as

where M is the mass, R is the radius, and k is the inertial constant. The inertial constant depends
on the shape of the object. Some common inertial constants are found below, in Table 2.4.
Table 2.4: Inertial constants
Shape I nertial Constant k
Wheel loaded at rim 1
Solid disk of uniform thickness 0.5
Solid sphere 0.4
Spherical shell 0.6667
Thin rectangular rod 0.5

To optimize the energy to mass ratio the flywheel needs to spin at the maximum possible speed.
Kinetic energy only increases linearly with mass but increases at the square of the rotational
speed, see the preceeding formula. However, centrifugal forces can rip apart rapidly rotating
objects. The centrifugal force for a rotating object is proportional to its density. Therefore, while
dense material can store more energy it is also subject to higher centrifugal force and fails at
lower rotational speeds than do low density material. This effectively means that tensile strength
is more important than density of material. For effective storage of energy, long rundown times
are required. Using frictionless bearings and a vacuum to minimize air resistance can result in
rundown times of 6 months. Flywheels provide about 80 percent efficiency. Figure 2-12 shows a
schematic of a flywheel energy storage system.

Figure 2-12: Flywheel Energy Storage System
Compressed-Air Energy Storage (CAES) plants use off-peak electricity to compress air into an
underground reservoir. When electricity is needed, the air is withdrawn, heated with gas or oil,
and run through expansion turbines to drive a generator. These plants burn about one-third of the
fuel of a conventional combustion turbine, and producing about one-third the pollutants.
Approximately three-fourths of the United States has the geologic potential for underground air
storage. Since this process uses an electromechanical converter to produce electricity, the
machinery is commercially available.
As described in the Conversion Section, pumped storage is a special use of hydroelectric energy.
Excess off-peak power is used to pump water to an elevated reservoir. When electricity is
needed, the potential energy of the water is released to flow through hydroelectric turbines,
exactly like a hydroelectric dam. Pumped storage plants require a large area with suitable
topography for the upper and lower reservoirs, limiting the number of desirable sites, and leading
to opposition from environmental groups. Pumped storage plants also have to be large (1000-
2000 MW capacity) to be economical, resulting in long lead times and high capital costs.
Some estimated costs for energy storage technologies are shown in Table 2.5.
Table 2.5
Estimated costs for Energy Storage Technologies
Technology

Power-
related cost
($/kW)
Energy-
related
cost
($/kW)
Hours
of
Storage
Total
Cost
($/kW)
CAES Small module (25-50 MW) 575 5 10 625
Large module (110-220
MW)
415 1 10 425

Pumped-hydro Conventional (500-1500
MW)
1000 10 10 1100
Underground (2000 MW) 1040 45 10 1490

Battery Lead-acid (target)(10 MW) 125 170 3 635
Advanced (target)(10MW) 125 100 3 425

SMES (Target)(1000 MW) 150 275 3 975

2.5a Load Growth
An electric load (or demand) is the power requirement of any device or equipment that converts
electric energy into light, heat, or mechanical energy. The total of all such loads connected to the
system constitutes the power system load. As such, the load varies daily, weekly, monthly, and
yearly with loads addition or subtraction from the power system. The minimum system load for a
given period is called the base load. The maximum system load for a given period is known as
the peak load or peak demand. The peak demand is usually quite short in duration. The operation
of generation plants must be closely coordinated with the load demands to ensure that enough
generation capacity is on-line. Since peak loads are generally only a few hours long, economical
fast-starting generators like pumped storage hydro are used. On weekdays, the base load
generally begins increasing at about 6:00 a.m., and hits peak load around 5:00 p.m. Maximum
yearly peak loads generally occur during the summer in the south, and during the winter for the
north.
As more people and businesses connect to an electric system, the amount of load on the system
increases. Load forecasting is performed to ensure that power system generating capacity will be
adequate to meet these future load demands. Power stations take years to build and require
advance planning, like load forecasting. One important part of load forecasting is the idea of the
load growth rate. This describes the estimated rate at which load on the power system increases
generally based on historical data. The growth rate of the system load L is mathematically
represented by

where a is the constant of proportionality, also known as the per-unit growth rate. The solution to
this equation is written as

where L
0
is the value of L at t = 0. At any two values of time, t
1
and t
2
, the ratio of the
corresponding L
1
and L
2
is

This equation may be used to determine the time t
k
such that L
2
= kL
1
and t
2
-t
1
= t
d
, given by

When talking about the growth rate of a quantity, the term "doubling time" is often used. This
term referes to the period necessary to double the initial value of load L , given a constant value
of a .

Doubling time is used to describe how long it will take, at a constant growth rate, to use twice
what is currently used. For example, assuming a present peak energy demand of 700 GW, and a
5 percent growth rate in the peak demand, the doubling time for energy demand is 14 years. This
means that peak energy demand in 14 years will equal 1400 GW, or twice the current peak
demand. Obviously, setting a steady growth rate for the use of any quantity is unrealistic, as the
growth rate depends on many factors. Load growth forecsting is still an important feature of
planning and constructing power plants and it involves more than simple linear growth
calculations.
The next section describes several of the environmental impacts of power production. These are
explored for both DG and non-DG technologies.



2.6 Environmental Impact of Electricity Generation and Transmission
All energy conversion methods used to produce electricity have some environmental impact. The
impact may have an active effect like the emission of airborne pollutants, or may have a passive
effect like aesthetics or habitat modification. Even methods considered environmentally friendly,
like wind, solar, and hydro, have some impact on the environment. Not only does the final
production of electricity have an environmental impact. The transmission of electricity with
concerns over electromagnetic fields, aesthetics, and land use, also impacts the environment.
The whole cycle of electricity generation must be considered when looking at the environmental
impact. This includes the production and transportation of fuel for the conversion process. This is
especially true of fossil fuel and nuclear power plants, which use large quantities of fuel taken
from the earth. Energy system environmental impact consists of fuel recovery and production,
fuel transportation, electicity transmission, and spent fuel emissions.
2.6.1 Environmental Effects of Raw Fuel Production
Coal Production
Coal is almost exclusively produced for electrical generation. Coal production is often viewed as
only a local environmental problem. Coal mining, particularly surface mining, has both long-
term and short-term effects on land, including dust, noise, and water drainage/runoff. Preparation
of coal produces both solid and liquid waste of which must be treated and disposed.
Transportation of coal produces dust while coal storage produces dust to control and also results
in water runoff problems.
Oil Production
The fuel oil burned in power plants is a byproduct of the petroleum industry, so electricity
production is partially responsible for environmental issues associated with oil and hydrocarbon
burning. This fuel burning produces many "greenhouse" gases. Other environmental impacts
associated with oil production include blowouts, spills, brine disposal, and the production of
hydrogen sulfide. Transportation of oil involves spill and leak hazards. Oil refining includes
environmental effects such as explosions, fires, air emissions, noise, odor, and water runoff.
Natural Gas Production
During the natural gas production, possible environmental effects include blowouts, leaks,
hydrocarbon emissions, and trace metal emissions. The treatment of natural gas involves air
emissions and the disposal of liquid residuals, while transportation and storage effects include
spills and explosions.
Uranium Production
Uranium mining involves concerns such as radioactive dust releases, mine water seepage,
protection of workers from radioactivity, and the disposal of a large quantity of mine waste
containing a low level of radioactivity. Uranium treatment must dispose of mill tailings
containing toxic metals and chemical wastes used in the treatment process, as well as
radiological waste. The treatment of raw uranium also must deal with radioactive dust releases.
The enrichment of uranium ore must account for liquid and gaseous effluent releases, and must
recycle fission products. Transportation involves the hazard of the accidental release of
radioactive particles.
2.6.2 Environmental Effects of Electricity Production
Fossil fuel power plants generally have the most widespread effect on the environment, as the
combustion process produces airborne pollutants that spread over a wide area. Nuclear power
plants have the most potentially dangerous effect. An operating accident at a nuclear station
could allow a large release of radioactive particles to occur. Solar, hydro, and wind power plants
generally have smaller effects on the environment.
Fossil Fuel Power Plants
Fossil fuel power plants produce environmental problems including land and water use, air
emissions, thermal releases, climatic and visual impacts from cooling towers, solid waste
disposal, ash disposal (for coal), and noise. Due to the need for large amounts of steam, plants
can have a great effect on water use. For example, a typical 500 MW coal fired power plant uses
25 x 10
9
l/GW-year of water, which must be taken from a water source, and then cooled to return
to the water source with as little environmental effect as possible. The biggest effect fossil fuel
plants have overall is the emission of air pollutants, particularly SO
X
, NO
X
, CO, CO
2
, and
hydrocarbons. Carbon monoxide, CO, carbon dioxide, CO
2
, and the hydrocarbons are the
"greenhouse gases," believed to be responsible for global warming. SO
X
and NO
X
produce acid
when released into the atmosphere, leading to the production of acid rain. Table 2.6 list
approximate amounts of airborne pollutants produced. Generally, air emissions are controlled by
the use of scrubbers and precipitators located at the plant.
Table 2.6
Power Plant Emissions (g/kWh)
Plant Type CO NO
X
SO
2
CO
2

Coal 0.11 3.54 9.26 1090
Oil 0.19 2.02 5.08 781
Gas 0.20 2.32 0.004 490
Nuclear Power Plants
Nuclear power plants have one environmental issue no other form of electrical power plant does.
An accident at a nuclear power plant may release large amounts of radioactive particles, possibly
resulting in a direct loss of life, and rendering a large land area immediately around the plant
unlivable. The largest regular environmental impact is the disposal of the high level nuclear
waste contained in spent fuel rods, as this waste must be stored safely for thousands of years.
There is no site in the United States currently accepting high level nuclear waste, so utilities are
generally storing the waste in above ground casks at plant sites. A long term issue is the
decommissioning of nuclear power plants. Decommissioning is shutting down a nuclear plant
after its operational life is over. At this point the entire reactor vessel becomes a high level
radioactive waste that must be disposed. The current methods of decommissioning a plant are to
completely remove and dispose of all radioactive components, to entomb the reactor in concrete,
or simply to shut the plant down and restrict access until the radioactivity dies out.
Solar
The production of electricity from solar energy sources generally has a small effect on the
environment. There are no residuals produced in the energy conversion process. The only
exception is solar thermal processes, which have an operating fluid that must occasionally be
discharged. There are some environmental concerns, however. Bulk solar plants generally
require a large land area, and they produce a great deal of heat. An unknown quantity in solar
energy is the disposal of photovoltaic cells. The most promising solar cells use gallium arsenide,
a toxic substance.
Hydro
The use of hydropower to produce electricity can have both positive and negative effects on the
environment. At some sites, a dam may help with flood control, flow regulation, or the reservoir
may provide recreational opportunities. At other sites, the dam may have adverse effects on the
hydrological cycle, water quality of the stream, stream ecology, fish migration, and cause the
destruction of landscapes and ecosystems. Building new high-head dams requires the
displacement and compensation of populations. Low-head dams generally have a benign effect
on the environment. Dam failures can lead to catastrophic floods.
Wind
Wind generators biggest environmental effects come from visual pollution, noise, and TV
interference. This is particularly true of wind farms, where 50 or more wind turbines may be
mounted at the same site. Wind farms situated on a migratory path may pose major hazards to
birds. There is also a safety hazard in case of blade breakage.
2.6.3 Environmental Effects Of Electricity Transmission
Transmission of bulk electricity from the generating station to the load uses wires suspended on
large towers, known as transmission lines. Traditionally these lines have been viewed only as an
aesthetic nuisance that could cause communications interference and be a hazard to low flying
aircraft. Today, there are other issues considered about the effect of transmission lines on the
environment. Greater concern is placed on the effect of the lines on the natural habitat. The
major new issue is the effect of electromagnetic fields (EMFs) on human health.
More than 1000 studies have been performed since 1979 to assess the relationship between low
frequency magnetic fields and human health. Most of the research has focused on the
relationship between cancer rates and fields produced in the 50-60 Hertz range used for
electricity transmission. The studies generally focus on children with residential exposure to high
voltage transmission lines, or workers with high degrees of occupational exposures to EMFs.
Most of the studies have failed to establish a statistically significant relationship between cancer
and EMFs. One difficulty in the studies is estimating the amount of EMF exposure a person
receives.
Magnetic fields are produced when current flows through a conductor. The strength of the
magnetic field decreases quadratically with distance from the conductor. While the strength of a
magnetic field directly underneath a transmission line is in the range of 300 to 600 milliGauss
(mG), the field strength is only 10 to 100 mG 200 feet away. This is in comparison to the earth's
natural magnetic field of approximately 450 mG. Home appliances like hair dryers, microwaves,
and televisions produce magnetic fields, ranging in strength from 1 to 50 mG. Obviously, due to
the large number of different sources of EMFs, estimating a person's daily exposure becomes
extremely complex. The studies generally used a broad estimate of EMF exposure.
3. Introduction to DG & Intentional Islanding*

This chapter is an introduction to intentional islanding and its compatibility with distributed
generation energies from an electrical engineering perspective. First, the concept of intentional
islanding is explained. Then types of DG and power conversion systems are discussed with a
focus on the common elements between the various types. The final section of this chapter
discusses the design basics of power conversion systems and the challenges faced in designing a
standardizable component of the system.
3.1 DG & Intentional Islanding
As the demand for more reliable and secure power systems with greater power quality increases,
the concepts of distributed generation (DG) have become more popular [1-4]. This popularity of
DG concepts has developed simultaneoulsy with the decrease in manufacturing costs associated
with clean and alternative technologies, like fuel cells, biomass, micro-turbine, and solar cell
systems. Although the costs associated with these technologies have continued to decrease more
work is needed to make these technologies readily available. To make these distributed energy
resource (DER) technologies more economically viable and energy efficient, power-electronics
based conversion systems need to be developed for the proper conditioning of the energy to be
delivered to the current three-phase power system. These power conversion systems (PCS) allow
for increased reliability, security, and fewer downtimes by incorporating intentional islands into
the utility grid without having to add or replace the existing transmission system.
3.1.2 What is Intentional Islanding?
Intentional islanding is the purposeful sectionalization of the utility system during widespread
disturbances to create power "islands" [4]. These islands can be designed to maintain a
continuous supply of power during disturbances of the main distribution system. As in the
following figure when disturbances are present on a distributed utility system, the grid
sectionalizes itself. The distributed energy resources can then supply the load power demand of
the islands created until reconnection with the main utility system occurs.

Figure 1. Islanding Diagram.
In the following section, several types of distributed generation and power conversion systems
will be reviewed. The concluding section describes the basics of design for power conversion
systems. These sections are aimed at providing an introduction to the electrical engineering
aspects of distributed energy. If you would like a general introduction to electrical engineering
concepts, there are links available for this in the "Useful Links" section.

*This web module was developed with the Timothy Thacker's Master's research work. Full
citation.


3.2 Types and Commonalities of DG & PCS Systems
To be able to properly design a power conversion sytem (PCS) for distributed energy resource
(DER) interconnection, the type of distributed energy resource must be identified and taken into
account. There are two basic types of distributed energy resources, direct current (DC) and
alternating current (AC) voltage producing sources. For either type to be connected to the utility
system their raw outputs need to be processed.
The follwing figure, Figure 1, shows possible DC and AC distributed energy resource based
PCS. Though direct voltage conversion from a DC or AC source to an AC grid, ready voltage
can be obtained (top topologies in Figure (a) & (b)), these PCS topologies are not used due to
hardware protection issues [10-15].
In the remaining topologies of Figure 1, the first section of the PCS is dedicated to the processing
of raw DER energy into DC energy. This conversion cannot be easily standardized due to the
vast differences between DER technologies. For example: micro-turbines and wind-turbines can
both produce variable frequency AC voltages, but the range of frequencies for the micro-turbines
are generally higher than those of the wind-turbines. In addition to this, the voltage levels
between the two types also vary. Micro-turbines have a constant supply of fuel to generate
electricity as opposed to wind-turbines which are dependent on the weather conditions. Hence,
wind-turbines require energy storage to help manage power flow to the utility. Distributed
energy resource (DER) technologies often differ in produced range of frequencies, supply
source, and energy storage requirements. The processing portion is the site for accounting for
these differences. Due to this the processing portion of a PCS system cannot be standardized for
all DERs. However, the grid interfacing inverter has the potential for standardization. It will be
analyzed to explore this potential for standardization.

Figure 1. (a) DC DER based PCS; (b) AC DER based PCS
Table 1. summarizes various DER types, possible PCS topological configurations, and the
functions that the PCS need to perform to convert the raw DER energy into viable grid-ready
energy. From this it can be seen that the grid-interfacing inverters have a common feature for all
types of DER systems; all DER systems require grid inverters. These inverters produce three
phase AC grid voltage (60Hz) from the DER generated power. This commonality between all
DER PCS types makes the inverter a ideal site for standardizing the grid interconnection.

Table 1. Examples of specific DERs and the needed PCS functions for interconnections.
Additional figures depicting the relationship of area and local electric power systems (EPS) to
each other and the utility are presented in Figures 2(a) and 2(b). Figure 2(b) shows detailed PCS
configuration and interconnection with area/local EPS and utility. These diagrams also further
emphasize the potential for standardization at the grid interface.


Figure 2. (a) Top, Area EPSs of a Utility System showing DG interconnection. (b) Bottom,
Black diagram of DER, PCS, Area EPS, and the grid interconnection.
The grid interfacing inverter is the portion of the PCS with the most constraints and protection
demands placed upon it to be able to connect with a utility system. This makes the
standardization of inverter advantageous for the control design. A voltage source inverter (VSI)
was selected to be used in the PCS for this study. The following sections will detail the rationale
behind this selection.

3.3 Basics for the Design of a Distributed Generation Power Conversion System
Special issues and concerns must be addressed when dealing with medium and high power
systems. These issues arise in instances where a design is being done for the support of power
delivery to the utility and when incorporating the concepts of both DG and intentional islanding.
The following sections detail the reasoning behind the selection of using a VSI, common
standards and regulations used for DG interconnections, and descriptions of challenges faced
while implementing medium and high power systems.
3.3.1 Focus on VSI of the PCS
The inverter portion of the PCS was selected as a VSI as previous described. This type of
inverter was selected not only because of the readily available power electronics building block
(PEBB) based inverter system, but also because of the type of control systems to be
implemented. PEBBs have the ability to standardize the converter system. The PEBBs developed
and leveraged for this study were insulated gate bipolar transitors (IGBTs) with anti-parallel
diodes (APD) based switches. VSI typically consist of a transitor device (like the IGBT) with an
anti-parallel diode (APD) to allow for the bi-directional current flow. This makes the selection of
VSI a natural choice after the PEBBs were integrated into the design.
As mentioned above, power will be regulated through the PCS via the control. With the VSI's
inherent bi-directional current flow capabilities and the utility's fixed voltage, the PCS can now
simply control the power flow by means of current regulation [1, 5, 9, 17, 18]. The control
system must allow for both the grid-connection and islanding modes of operation. Direct
regulation, current for grid-connected mode and voltage for islanding mode, is preferable to the
alternatives. It allows for the system to achieve zero steady state error in output. The VSI's
bidirectional current flow fits the demands of DG technologies in an intentional islanding
framework rather nicely.
3.3.2 Standards and Common Practices for Grid Interconnections
The following standards are considered guidelines during the design process of the PCS and
control thereof:
ANSI/IEEE C84.1 - 1995, [19]
IEEE 519 - 1992, [20]
IEEE 929 - 2000, [21]
IEEE 1547 - 2003
UL 1741, [22]
These standards provide guidelines and specifications for the interconnection and control of
DERs to the utility grid. The following are brief summaries of each standard:
ANSI/IEEE C84.1 - 1995 standard deals with common line voltages at different distributions
levels (ie: residential power is single phase and an RMS voltage of 120 V, where as some
commercial sites have 3 phases with an RMS voltage of 240 V).
IEEE 519 - 1992 are recommended practices and requirements for the harmonic control of
electrical power systems. It sets maximum total harmonic distortion (THD) limits on voltages
and currents that a power system is allowed. Therefore, the PCS cannot inject harmonics into the
grid that cause the system to go above these limits set forth by the standard and the PCS should
filter these harmonics [23].
IEEE 929 - 2000 are recommended practices for the utility interface of photovoltaic (PV)
systems. Though written for PV inverters, the guidelines and specifications can be adapted to be
used for an inverter connecting a DER to the utility.
IEEE 1547 - 2003 is the standard for the interconnection of distributed resources to the utility
grid. This standard outlines requirements and specifications that the conversion systems of the
DER have to meet to be allowed to connect to the utility. This standard does not deal with the
concepts and issues of intentional islanding, and currently dictates that the DER shall disconnect
from the distribution system when islanding events occur. As noted above the standard does
leave open a section for consideration of intentional islanding in future revisions of the standard.
An analysis of 1547 raising questions to issues proposed by it can be found in [24].
UL 1741 is the Underwriters Laboratories' testing standards for equipment as they relate to IEEE
1547.
3.3.3 Challenges for Medium & High Power Inverters
Switching Frequency & Line Conditions -
The selected PCS is a pulsed-width modulation (PWM) based system. This makes it
advantageous to push the switching frequency as high as possible. Higher switching frequencies
translate to reduction in passive component (inductors and capacitors) sizes. However, with the
present semiconductor switching devices available for medium and high power systems (GTOs,
ETOs, IGBTs, etc...), device limitations require that the system switching frequencies be on the
order of kilohertz to the tens of kilohertz range. This, along with a utility line frequency of 60
Hz and low resonant frequency of the output line filter, brings about special considerations to the
design of control systems [10].
Due the fact that the resonant frequency of the VSI output filter tends to be a few decades below
the switching frequency (in order to allow filtering of the negative affects of the PWM
switching) and the frequency generally needs to be one to two decades above the line-frequency
to allow the system's natural dynamics to behave properly. This means that the control loop must
be designed within a narrow bandwidth. In addition to filter effects, the loading conditions of the
converter also affect the control design.
Digital Delay -
It has become common in converter systems for the control to be digitally implemented through
digital signal processors (DSPs); especially in 3 phase, medium to high power systems that
require complex calculations for coordinate transformations from stationary reference frames to
that of rotating reference frames for use in the control. Also, analog circuitry makes the
implementation of these transformations virtually impossible. Further, limiting of the states
(controlled variables), controller anti wind-up, and other protection protocols are easily
transported digitally to places where they are more complex to build in an analog control system.

Synchronization to the Grid -
Another fundamental aspect to consider in the design of DER to grid connected PCS is that of
synchronization to the utility system. The control needs measurements of the frequency and line-
angle of the utility to properly ensure that it can regulate the real and reactive power flow
through the PCS during the periods of time which the PCS is interconnected to the grid. These
measurements are obtained through the implementation of phase-lock loops (PLL), which will
use voltages of the PCS and grid to track the frequency and angles. Without these measurements
to synchronize the PCS to the utility, the power flow to/from the PCS will be incorrect, and
protection and safety issues arise [25].
The control system also needs these measurements for use in the coordinate transformation
calculations of the rotating reference frame .
Detection & Re-closure of/to the Grid -
One of the key features of operating a DER interconnected to the grid, and running it in both
islanding and grid-connected modes, is that the system has to have the capability to
autonomously detect when disturbances on the grid occur (over/under voltages and/or
frequencies, line faults, faults to ground, etc...). This ability to determine if it needs to
disconnect from the utility is to protect itself and the surrounding Area EPS from the grid
disturbances [1, 4].
The advantages of the PCS being able to autonomously detect potentially hazardous disturbances
on the grid not only aids in the creation and operation of fault protection schemes, but also helps
the control switch modes of operation from a "grid-connected" mode to an "islanding" mode.
Common practices and proposed new detection and re-closure schemes are not within the scope
of this module.
This module serves to present the rationale behind choosing the VSI as the common feature for
DG technologies. Both the common standards and regulations for DG technologies and the
challenges to medium and high power inverter design inform this choice.


3.3 Basics for the Design of a Distributed Generation Power Conversion System
Special issues and concerns must be addressed when dealing with medium and high power
systems. These issues arise in instances where a design is being done for the support of power
delivery to the utility and when incorporating the concepts of both DG and intentional islanding.
The following sections detail the reasoning behind the selection of using a VSI, common
standards and regulations used for DG interconnections, and descriptions of challenges faced
while implementing medium and high power systems.
3.3.1 Focus on VSI of the PCS
The inverter portion of the PCS was selected as a VSI as previous described. This type of
inverter was selected not only because of the readily available power electronics building block
(PEBB) based inverter system, but also because of the type of control systems to be
implemented. PEBBs have the ability to standardize the converter system. The PEBBs developed
and leveraged for this study were insulated gate bipolar transitors (IGBTs) with anti-parallel
diodes (APD) based switches. VSI typically consist of a transitor device (like the IGBT) with an
anti-parallel diode (APD) to allow for the bi-directional current flow. This makes the selection of
VSI a natural choice after the PEBBs were integrated into the design.
As mentioned above, power will be regulated through the PCS via the control. With the VSI's
inherent bi-directional current flow capabilities and the utility's fixed voltage, the PCS can now
simply control the power flow by means of current regulation [1, 5, 9, 17, 18]. The control
system must allow for both the grid-connection and islanding modes of operation. Direct
regulation, current for grid-connected mode and voltage for islanding mode, is preferable to the
alternatives. It allows for the system to achieve zero steady state error in output. The VSI's
bidirectional current flow fits the demands of DG technologies in an intentional islanding
framework rather nicely.
3.3.2 Standards and Common Practices for Grid Interconnections
The following standards are considered guidelines during the design process of the PCS and
control thereof:
ANSI/IEEE C84.1 - 1995, [19]
IEEE 519 - 1992, [20]
IEEE 929 - 2000, [21]
IEEE 1547 - 2003
UL 1741, [22]
These standards provide guidelines and specifications for the interconnection and control of
DERs to the utility grid. The following are brief summaries of each standard:
ANSI/IEEE C84.1 - 1995 standard deals with common line voltages at different distributions
levels (ie: residential power is single phase and an RMS voltage of 120 V, where as some
commercial sites have 3 phases with an RMS voltage of 240 V).
IEEE 519 - 1992 are recommended practices and requirements for the harmonic control of
electrical power systems. It sets maximum total harmonic distortion (THD) limits on voltages
and currents that a power system is allowed. Therefore, the PCS cannot inject harmonics into the
grid that cause the system to go above these limits set forth by the standard and the PCS should
filter these harmonics [23].
IEEE 929 - 2000 are recommended practices for the utility interface of photovoltaic (PV)
systems. Though written for PV inverters, the guidelines and specifications can be adapted to be
used for an inverter connecting a DER to the utility.
IEEE 1547 - 2003 is the standard for the interconnection of distributed resources to the utility
grid. This standard outlines requirements and specifications that the conversion systems of the
DER have to meet to be allowed to connect to the utility. This standard does not deal with the
concepts and issues of intentional islanding, and currently dictates that the DER shall disconnect
from the distribution system when islanding events occur. As noted above the standard does
leave open a section for consideration of intentional islanding in future revisions of the standard.
An analysis of 1547 raising questions to issues proposed by it can be found in [24].
UL 1741 is the Underwriters Laboratories' testing standards for equipment as they relate to IEEE
1547.
3.3.3 Challenges for Medium & High Power Inverters
Switching Frequency & Line Conditions -
The selected PCS is a pulsed-width modulation (PWM) based system. This makes it
advantageous to push the switching frequency as high as possible. Higher switching frequencies
translate to reduction in passive component (inductors and capacitors) sizes. However, with the
present semiconductor switching devices available for medium and high power systems (GTOs,
ETOs, IGBTs, etc...), device limitations require that the system switching frequencies be on the
order of kilohertz to the tens of kilohertz range. This, along with a utility line frequency of 60
Hz and low resonant frequency of the output line filter, brings about special considerations to the
design of control systems [10].
Due the fact that the resonant frequency of the VSI output filter tends to be a few decades below
the switching frequency (in order to allow filtering of the negative affects of the PWM
switching) and the frequency generally needs to be one to two decades above the line-frequency
to allow the system's natural dynamics to behave properly. This means that the control loop must
be designed within a narrow bandwidth. In addition to filter effects, the loading conditions of the
converter also affect the control design.
Digital Delay -
It has become common in converter systems for the control to be digitally implemented through
digital signal processors (DSPs); especially in 3 phase, medium to high power systems that
require complex calculations for coordinate transformations from stationary reference frames to
that of rotating reference frames for use in the control. Also, analog circuitry makes the
implementation of these transformations virtually impossible. Further, limiting of the states
(controlled variables), controller anti wind-up, and other protection protocols are easily
transported digitally to places where they are more complex to build in an analog control system.

Synchronization to the Grid -
Another fundamental aspect to consider in the design of DER to grid connected PCS is that of
synchronization to the utility system. The control needs measurements of the frequency and line-
angle of the utility to properly ensure that it can regulate the real and reactive power flow
through the PCS during the periods of time which the PCS is interconnected to the grid. These
measurements are obtained through the implementation of phase-lock loops (PLL), which will
use voltages of the PCS and grid to track the frequency and angles. Without these measurements
to synchronize the PCS to the utility, the power flow to/from the PCS will be incorrect, and
protection and safety issues arise [25].
The control system also needs these measurements for use in the coordinate transformation
calculations of the rotating reference frame .
Detection & Re-closure of/to the Grid -
One of the key features of operating a DER interconnected to the grid, and running it in both
islanding and grid-connected modes, is that the system has to have the capability to
autonomously detect when disturbances on the grid occur (over/under voltages and/or
frequencies, line faults, faults to ground, etc...). This ability to determine if it needs to
disconnect from the utility is to protect itself and the surrounding Area EPS from the grid
disturbances [1, 4].
The advantages of the PCS being able to autonomously detect potentially hazardous disturbances
on the grid not only aids in the creation and operation of fault protection schemes, but also helps
the control switch modes of operation from a "grid-connected" mode to an "islanding" mode.
Common practices and proposed new detection and re-closure schemes are not within the scope
of this module.
This module serves to present the rationale behind choosing the VSI as the common feature for
DG technologies. Both the common standards and regulations for DG technologies and the
challenges to medium and high power inverter design inform this choice.


3.4 Questions.
1. How does intentional islanding offer a more secure alternative to traditional utility
configurations?
2. What makes the grid interface invertor ideal for standardization?
3. Name 2 challenges for medium- and high-power invertors.
(a). Propose a solution to one of the named challenges for a specific system.
4. List 2 standards used for grid interconnections.
3.5 Instructor's Guide
This instructor's guide provides answers to the question section.
3.4.1 Suggest that learners answer in a short paragraph. Both describing and refuting the security
advantages discussed in this module demonstrate interaction with the module.
3.4.2 From section 3.2, all distributed energy resource technologies reguire grid invertors. Grid
invertors perform the a common function for all of the DER technologies. They produce three
phase AC grid voltage from the DER generated power.
3.4.3 From section 3.3.3: switching frequency and line conditions, digital delay, synchronization
to he grid, anddetection and re-closure to the grid are some of the challenges medium and high
power invertors face. There are multiple other challenges.
(a) The answer to this questions demands an evaluation of the creative efforts put forward by the
learner.
3.4.4 ANSI/IEEE C84.1 - 1995, [19];IEEE 519 - 1992, [20]; IEEE 929 - 2000, [21];IEEE 1547 -
2003; UL 1741, [22]
4. Introduction to the Business of Energy
This section is intended to serve as an introduction to energy from a business perspective. The
main concepts to be discussed include: power management, unit commitment and optimal power
flow, energy markets, financial risk management, and demand forecasting. Some case studies
will also be presented to illustrate these concepts. The first of the concepts will be explored
below.
4.1 Power Management
4.1.1. The Infrastructure of a Power System.
A power system is a large, complex integration of large generators, a transmission grid,
distribution grids, distributed generators and loads, which are connected at points called busses .
In describing the management of a power system we must define several components of the
system:
Physical assets - generators, loads and grids
Asset owners - government, corporations and consumers
Asset operators - public and private utilities, consumers
Asset managers and energy markets - bi-lateral, wholesale, retail, ancillary-services markets
and grid management
Risk markets - financial risk portfolio managers, over-the-counter derivatives markets (OTC),
commodity/derivative exchanges
Asset managers - public and private utilities, contractors, independent service operators (ISO)
Figure 1.1 illustrates this system of interconnected components.

Figure 1.1: Power System Assets
4.1.1.1 Loads
A load is a demand for electric power. In power-grid management all loads are identified by the
bus at which they connect to the distribution grid. There are three categories of loads:
residential, commercial and industrial
Residential load = the demand for electric power from home-owners, apartment dwellers and
any other form of residence.
Commercial load = the demand for electric power from shopping malls, stores, government
agencies and light business activities.
I ndustrial load = the demand for electric power from manufacturing facilities and other forms of
heavy industry.
Most users of electricity do not purchase power directly from a generator. Instead they
purchase power from a retailer. These retailers are known as load serving entities (LSE) or load
aggregators . Retailers purchase power from wholesalers who, in turn, purchase power from
suppliers. The actual delivery of electric power is sub-contracted to power-line
owner/operators.
4.1.1.2 Islanded and interconnected DG
The source of electric power that is potentially nearest to a load is a small generator that is
connected directly to the load and serves that load only. This kind of generation is called
islanded distributed generation (DG). The qualifier "islanded" implies that this type of
generator cannot deliver power to loads other that the one to which it is connected. Examples of
islanded DG are small gas-turbine generators that provide power to commercial or industrial
sites such as universities, shopping malls and refineries or clusters of residential loads such as
apartment complexes. Other examples of islanded DG are solar panels atop houses and
buildings.
In the management of a power grid, islanded DG has the effect of reducing the average load
placed on the distribution grid as a DG unit can provide some or all of power required by the
facility to which it is connected. However, the reliability of a DG unit has an effect on the
variability of the load placed on the distribution grid as the shutdown of a DG unit suddenly
transfers the facility's demand for power from the DG unit to the distribution grid.
Some distributed generators may be connected to the distribution grid in addition to a particular
load. This kind of generation is called interconnected DG . The owner of an interconnected
distributed generator may be able to use the power of this generator for the load to which it is
connected as well as for sale to other loads within a distribution grid.
4.1.1.3 Generators, IOU, PU
Most generation comes from large power plants, which are typically fueled by coal, fissionable
radioactive elements or natural gas. Wind farms, hydroelectric generators, geothermal
generators and other new generation technologies round out the portfolio of large-scale
generation sources.
These generators are owned and operated either by public utilities ( PU ) or private, investor
owned utilities ( I OU ). Large generators are usually located long distances from the many
loads that they serve. Consequently, large generators make their power available to
transmission grids, which carry power from the generators over high-voltage power lines across
large geographical areas.
4.1.1.4 Transmission grids, basis points, ISO
A transmission grid is a network of high-voltage power lines, transformers and busses that
transfer electric power from large generators to distribution grids. A bus that connects a
distribution grid to the transmission grid is called a basis point . A transmission grid covers a
large geographical area and is connected to numerous large generation assets. The topology of
the transmission grid is complex and allows for power to flow in each branch of the network in
either direction as conditions require.
Transmission grids are owned and operated by regulated business entities called independent
service operators ( I SO ). ISO's are responsible for ensuring that the transmission grid is
maintained, all large generators have access to the grid and that all loads placed on the grid at
basis points are supplied.
Transmission grids are connected to one another to form a large power grid that can stretch
across national regions and even international borders. Transferring large amounts of power
from one transmission grid to another is a transaction known as wheeling.
4.1.1.5 Distribution grid
A conventional distribution grid is a network of low-voltage power lines, transformers and
busses that bring electric power to loads within a small geographical area such as a small city,
town or rural area within a county. The conventional topology of a distribution grid is
arborescent, which means that the power enters the grid at a basis point and is distributed to
loads through successive branching. Each bus in such a network has one line that brings power
into the bus and several lines that bring power away from the bus. The terminal points of a
distribution grid are the loads that the grid supplies. Figure 1.1.5 illustrates a distribution grid.

Figure 1.1.5 Distribution Grid Topology
Distribution grids are owned and operated by public utilities, cooperatives or investor-owned
utilities that are granted monopoly rights over the grid's power distribution. These service
organizations are responsible for the maintenance of the distribution grid.
The development of cost-effective DG has introduced two modifications to the conventional
power grid.
Islanded DG can reduce the load placed on the distribution grid at the discretion of the
operator of the DG asset. Although the average load placed on the distribution grid is reduced
by islanded DG, the volatility of this load may be increased due to the uncertainty in the DG
operation.
Interconnected DG can inject power into the distribution grid. This injection supplants some
of the power that the distribution grid would purchase from the transmission grid. However,
interconnected DG can cause power to flow in directions opposite to the conventional flow of
power from the trunk of the distribution grid outward to the ends of the terminal branches. This
reverse flow and the associated inverted bus voltages within the distribution grid raise safety and
reliability concerns. See Chapter 2 Environmental Impacts.

4.1.2 Hierarchical Planning in Power Management

The control of markets and electricity grids requires coordinated decision-making across five
decision domains. The scopes of these domains are identified below.
Capacity planning: installed generation capacity, grid configuration, market regulations
Power flow planning: bi-lateral contracts, wholesale bids & offers, unit availability
Power flow scheduling: unit commitment, ancillary services contracts, reserve requirements
Power dispatching: unit dispatch, demand management, regulation
Power system controlling : voltage control, frequency control, circuit protection
The determination of the optimal power flow in a grid over a sequence of time periods requires
the simultaneous optimization of all of the decisions listed above - a feat made impossible by the
large number of variables that these decisions encompass. Consequently, power grid and market
management is carried out through the application of a conventional hierarchical approach.
The conventional hierarchy of decision making conforms to the ordered list of decisions show
above. Every decision is characterized by three elements: decision variables, which specify the
alternatives available to the decision maker; performance measures, which specify the basis for
evaluating the feasibility and objectives of a decision; parameters, which specify the given,
uncontrollable factors that, together with the decision variables, determine the values of the
performance measures The basic idea behind hierarchical planning is that the solution to a
rough-cut representation of a decision in terms of aggregated decision variables can serve as a set
of guidelines and constraints for a refined decision in terms of detailed decision variables. For
example, the problems of determining the unit availability, unit commitment and unit dispatch
are all related through performance measures such as profit for asset owners, service level to
loads and total cost of power throughout a grid. Rather than attempt to find solutions to all three
decisions simultaneously so that a globally optimal solution is obtained, a hierarchical planning
approach would solve three separate decisions in stages. Specifically, the determination of unit
availability is based on approximate forecasts of total demand over the upcoming week. The
unit availability decision provides capacity constraints on the commitment and dispatching
decisions. The commitment decision, in turn, is based on a forecast of load variations over the
next 36 hours and consumes the bulk of the generation capacity, leaving a judicious amount of
capacity for support of the regulation dispatching decisions, which cure any imbalances between
loads and committed generation.
In order to implement a hierarchical planning scheme, each level of planning must approximate
the effects of the lower-level decision variables on the current stage's constraints and
performance measures. Furthermore, the solutions to higher-level decisions impose constraints
on the lower-level decisions.
Using the "hat" notation to indicate approximations, the hierarchical planning approach is
described as follows:
Suppose we have three sets of decision variables y
1
, y
2
, y
3
for the following decision model,

subject to:

By approximating the effects of variables y
2
, y
3
we construct the following aggregate planning
problem.

for j=1, ...,n
The first optimization in the hierarchy is,

subject to:

Resulting in a solution, y
1
*, which becomes a parameter in for all of the succeeding problems.
The second approximate decision model is,



subject to:


The remaining optimization problems are formulated in a similar manner.

<< Back Next >>
2007 Consortium on Energy Restructuring, Virginia Tech
4.1.3 Sequential Decision Processes and Adaptive Planning
The control of markets and electricity grids must be done on a continuous basis, which
necessitates ongoing decision-making regarding the supply availability, demand management,
unit commitment, dispatching, ancillary services and regulation. For practical reasons, this
decision process is carried out by updating plans at discrete points in time as opposed to
continuously adapting to changing conditions. The time interval between updates of a plan is
called the planning interval and the number of intervals into the future over which the plan is
specified is called the planning horizon . At the beginning of each planning interval, an asset
manager's updated plan is based on the current status of the asset and its environment and the
latest forecast of demand, supply and prices. Moreover, opportunity costs influence the plan, as
current decisions depend on future plans. The periodic updating of plans forms a sequential
decision process (SDP) , which is a general term that identifies any process through which a
sequence of decisions is made over time in a way that each decision is adapted to the legacy of
all previous decisions and to the outcomes of uncontrollable influences on the performance
measures that are pursued by the decisions.
A widely-used heuristic approach to solving SDP's is known as the rolling horizon planning
procedure. This procedure is executed through the combination of three planning techniques:
Rolling the plan: Plans are updated at regular intervals.
Planning over a horizon: Each plan extends over a number of future time periods.
Adapting the plan: At each update of the plan, the plan is adjusted within limits that are
determined by the system's constraints on the rates at which resource flows can change. The
planning horizon for each plan consists of a horizon over which the plan must be "frozen"
followed by a horizon over which adjustments are allowed. The boundary between the fixed
portion of a plan and the adjustable portion of a plan is called the planning "fence".
Of course the discrete representation of the time scale within a SDP for a process that changes
continuously introduces an approximation. However, the notion of developing a plan in finer
and finer detail as one proceeds through the levels of the hierarchical planning method described
in Section 1.2 applies to the time scale as well. Higher-level, more strategic decisions are given
longer planning horizon and longer planning intervals than lower-level tactical or operational
plans. As one moves down the hierarchy of decisions, the planning horizons and the planning
intervals are made shorter.
Another approximation that is inherent in a rolling horizon and adaptation procedure stems from
the use of a deterministic forecast for each plan update. The accuracy of this forecast increases
as the forecast horizon decreases. Consequently, the adaptation options with the shortest time
fences enjoy the most accurate forecasts and can be viewed as "fine tuning" actions with respect
to the "coarse tuning" of the plans produced by the longer-fence options.
Table 1.3a shows the basic scope and definition of the five levels of hierarchical planning, which
make up power system management.
Decision
Domain
Planning Horizon
(typical)
Planning Interval
(typical)
Configuring > 1 year > 1 month
Planning 1 day - 1 year 1 day
Scheduling 36 hours 1 hour
Dispatching 1 hour 5 minutes
Controlling 0.5 hour < 5 seconds
Table 1.3a: Planning horizons and periods
In the case of electricity scheduling and dispatch there are four options for specifying and
updating a plan. Each option is constrained to be exercised within the capacities that are set by
the capacity reservation decisions made at a higher level of the decision-making hierarchy (see
Section 1.2 ). Tables 1.3b - 1.3e define these options. The update intervals, planning horizons
and time fences given in Tables 1.3c and 1.3e are typical values in the operation of a large power
system.
Scheduling option Capacity constraint Demand
constraint
Day-ahead unit
commitment
Day ahead offers Day-ahead bids
Imbalance commitment Imbalance offers Imbalance bids
Regulation reserve
commitment
Regulation reserves
offers
Regulation
forecast
Spinning reserve
commitment
Spinning reserves
offers
Control error
forecast
Table 1.3b: Capacity and demand constraints on scheduling options

Scheduling option Update
interval
Planning
horizon
Time
fence
Day-ahead unit
commitment
24 hours 36 hours 12 hours
Imbalance commitment 24 hours 30 hour 6 hours
Regulation reserves 8, 16 hours 9, 17 hours 1 hour
Spinning reserves 8, 16 hours 9, 17 hours 1 hour

Table 1.3c: Scheduling option parameters
Dispatch/Control
option
Capacity constraint Demand constraint
Day-ahead dispatch Day-ahead
commitments
Day-ahead
commitments
Real-time dispatch Imbalance commitments Demand forecast
Ancillary service
regulation
Regulation reserve
commitments
Regulation error
Voltage/frequency
control
Spinning reserve
commitments
Control error
feedback
Table 1.3.d: Capacity and demand constraints on dispatching options
Dispatch/control
option
Update
interval
Planning
horizon
Time fence
Day-ahead unit
commitment
8, 16 hours 9, 17 hours 1 hour
Real-time dispatch 1 hour 1.5 hours 30 minutes
Ancillary service
regulation
5 minutes 30 minutes 5 minutes
Voltage/frequency
control
4 seconds 30 seconds 4 seconds
Table 1.3.e: Dispatching option parameters
In the re-structuring energy markets of the United States, generation-unit commitment decisions
are made through a combination of self-scheduling decisions made by generation asset managers
and market clearing of bids and offers for electric power through markets that are managed by
ISO's. See Module 3 (under construction) for an explanation of the workings of these markets.
4.1.4 Potential Approaches to DG Power Management
The integration of DG capacity into the management of a regional power grid presents some new
opportunities and risks. Much work and experimentation needs to be done before the proper
role of DG in the power systems of the future can be determined. In this section we examine
some of the alternatives in the use of DG.
Four different categories of business entities appear likely to consider the implementation of DG
technology as an element of a strategic plan. These include:
Investor owned utilities (IOUs) and publicly owned utilities that may want to install DG
units for supplying peak demand in areas that are located behind congested transmission
lines.
Manufacturers of DG systems, which have already advanced the technologies for DG on
several fronts and, with the exception of large gas turbines, appear to be pacing their
capacity growth by market growth.
The new generator and consumer (NGC) that sees operation of DG units as a potential
substitute for some or all of its purchases of electricity from utilities. Included in this
category are industrial sites, apartment complexes, government agencies, military bases,
universities, hospitals, shopping malls, and the like. On the horizon, NGCs may also
become wholesale energy suppliers by interconnecting DG units.
The contractor industry that performs one or more of the functions of designing, building,
installing, and operating DG units. The viability of contractors depends on the rate of
adoption of DG by the above categories of business entities.
The most common business model for DG adoption and growth may consist of one that partners
new generators and consumers with DG contractors. Such partnerships relieve the NGCs of the
need to develop extensive expertise in power generation and management, which would distract
them from their core business operations. Instead, NGCs can engage the services of contractors
who have expertise in constructing and operating DG units.
A foray into power generation by an NGC is a risky venture, even with the assistance of a
qualified contractor. Several categories of uncertainty engender financial as well as non-
financial risks for the NGC.
Technological uncertainty: DG owners must deal with the chance that the DG
technology will not perform as reliably or as efficiently as its specifications. In
particular, interconnected DG systems may actually reduce the reliability of a distribution
grid due to the inability of grid operators to control unit dispatches under rapidly
changing conditions.
Fuel cost uncertainty: The price of natural gas, coal, and oil will affect the financial
performance of any DG unit that uses any of these fuels. Owners of renewable energy
technologies (such as wind turbines) will not need to worry about the cost of energy
resources, but the price they receive for surplus power will depend to a large extent on
the price of conventional fuels that provide competitive benchmarks prices.
Load uncertainty: The growth and volatility of electricity demand within the
transmission grid that serves the NGC must be considered. Ironically, efforts to reduce
the cost of electricity in the form of demand-side response could reduce the value of DG
units that are most beneficial in supplanting expensive peak-load power from utilities.
Electricity price uncertainty: The financial performance of a DG unit depends on the
cost of electricity from utilities that the unit supplants. Future prices of electric power in
the United States remain highly uncertain due to variability in fuel costs, regulation, and
technological change.
Regulatory and public policy uncertainty: The viability of DG projects depends, to a
certain extent, on the actions of government entities. NGCs need to consider the chance
that any tax incentives, subsidies, or easements associated with a DG implementation
may be offered or repealed by future legislatures and executive branches. Given the
spotty history of utility system restructuring and deregulation, it is difficult to predict the
effects of government policies on evolving electricity markets.
In spite of the risks associated with these uncertainties, there are opportunities for the profitable
and otherwise successful use of DG. One approach to the implementation of DG merits
consideration for its potential to produce profits for DG owners as well as benefits to grid
reliability, security and environmental impacts. This approach aggregates the operation of DG
units into a single management entity that can dispatch power to the grid. Through this
management DG owners would employ electricity and waste heat for their own needs when
power remains cheap, and would sell surplus power to the grid when spot prices for electricity
are high. However, in order to participate in the market for power, DG units must be dispatched
at times and locations where they are most needed. This kind of dispatch requires coordination
of all DG units in a distribution grid; hence, the need for a supply aggregator to manage multiple
DG units.
Through this managerial hierarchy, DG owners would contract with a supply aggregator to sell
electricity to the power grid in much the same way that consumers contract with load aggregators
(retailers) to purchase electricity from the grid. The supply aggregator trades off the cost of
generation from the DG units against the spot prices available from the wholesale market.
Furthermore, the supply aggregator has the expertise to protect DG owners from the financial
risks of price volatility through the trading of commodity derivatives. Given the special
expertise and close attention that such trading requires, DG owners would find the services of a
contractor beneficial in terms of financial performance as well as a relief from the need to
develop expertise outside of their core competencies. The supply aggregator of the power
systems of the future would design, install, maintain, and operate DG units for the owners as
well as trade the power produced by the units and manage risks associated with operation of the
units within an integrated distribution grid.
An experiment in New York, funded by the New York State Energy Research and Development
Authority (NYSERDA) and the U.S. Department of Energy, suggests that such dispatch can
occur efficiently. In this experiment, fifty backup generators (providing a capacity of 35 MW)
were linked to centralized control points for dispatch to the power grid. The centralized
dispatcher served as an aggregator that purchased bulk electricity from utilities and sold it in the
day-ahead market in different parts of New York. When real-time prices exceeded the cost of
generation from the DG units, the aggregator dispatched the DG units and earned revenues.
This aggregated DG system proved to be technically feasible and produced savings of $1.5
million in one year through the use of islanded DG units for load curtailment. The success of the
NYSERDA/DOE experiment should motivate serious interest by business and government
regulators in the potential for aggregated dispatching for DG.

<< Back
2007 Consortium on Energy Restructuring, Virginia Tech
5. A Critique & Analysis of Four Major Distributed Generation Technologies*
5.1 Introduction
Distributed generation (DG) is not a new concept. Large scale centralized power systems were
created based on an 'economy of scale' model and currently dominate the United Stated power
grid. Unfortunately, recent problems with the transmission system, the grid, and the dynamic fuel
market have created a need to stabilize, distribute, and equalize grid power flow. While some DG
technologies have not commercially or economically matured, the authors believe that DG
technologies will become a reality in the next few decades. In this module, the authors analyze
and rank four small-scale DG technologies, evaluating the technologies' current status and
potential status. This ranking provides the basis of comparison among the technologies as
practical solutions to the woes of the United States power grid. Distributed generation
technologies potentially hold the key to unlocking current grid power flow stabilization,
distribution, and equalization problems.
Scope of This Document
This study compares distributed generation in three technology groups: internal combustion
engines (ICE), gas turbines (two different types), and one type of fuel cell (solid oxide fuel
cells). In order to maximize applicability, we have focused our analysis on non-site specific DG
technologies. So for example, we have not compared photovoltaic and wind turbine technologies
to the previously mentioned technologies. Technologies such as photovoltaics and wind power
are dependent upon usable sunlight and average wind speed respectively which vary from
location to location. Further, these require a large amount of area per generated kW[1]. Another
location specific example is landfill DG technologes. A landfill gas production facility requires a
large landfill and a specific location. It must be noted that the authors of this paper feel these
location restricted technologies should be implemented as they become technologically and
economically viable. However, we consider the aforementioned to be very specific centralized
generation technologies that cannot be realized everywhere. While it is possible to put
photovoltaic arrays on top of every home, gas station and large corporation, current power
densities preclude profitability and practicality. Likewise, operating a two hundred foot tall wind
turbine in the heart of "the Windy City" (Chicago) is not a practical engineering solution to our
power demands. In this module, we have limited our scope to non-site specific practical
technologies. It is our hope that by limiting our scope to these technologies we have opened up
more possibilites in more places.
Distributed Generation Facility Sizing
There are three basic methods for determining the generating capacity of a DG system, base,
intermediate, and peak. A general overview of storage and load growth can be found at Chapter
2.5 Storage. For the base method, the capacity of the facility is defined as the average base power
required by the load. In this case, the DG system effectively generates maximum power all the
time. The end user will buy power from the grid system any time the base power of the DG
system is exceeded (peak power). For the intermediate method, the capacity is defined as the
mean power required. During peak power demand, the electricity is obtained from the grid, and
conversely, during low power demand, the electricity is sold back to the grid, thus maintaining
full utilization of the facility. The third method is peak capacity. The DG system operates at a
base level at all times, but has an auxiliary power supply that only provides power during peak
power requirements. This type of system can theoretically be grid independent because of the
auxiliary power capacity [1].
We propose a slightly different approach to ensure higher power quality (PQ) among demanding
customers. In our system, the DG facility is sized for the peak power demand of the high PQ
customers. Should the grid feed be lost, the DG station would shed non-critical loads freeing
capacity to cover the mission critical load. Under normal operation, the critical load's peak power
would still be sourced from the grid and the DG system would return to the base case generating
facility. This type of toggling facility would increase the security of power system through the
use of DG technologies while providing alternative forms of energy during non-critical times.
Penetration Depth
Penetration depth, described in a very simplistic fashion, is the amount of power the DG system
provides to the load in relation to the capacity of the feeder line (as known as the transmission
line). The penetration depth is determined by many factors, including but not limited to: distance
to substation, DG induced frequency fluctuations and damaging harmonics, recloser/relay
location, feeder line size and type of DG system. For example, the harmonics created by a line-
commutated based DG system will significantly limit penetration depth as compared to the same
system with a pulse width modulator (PWM) inverter. Here, the line-commutated inverter is
dependent on the feeder line for conversion from DC to AC power and ,therefore , it is
dependent on the feeder line harmonics. However, the PWM inverter produces nearly sinusoidal
output at minimum current harmonics.
The detailed calculation of penetration depth is outside of the scope of the paper, but is directly
related to the size of the DG system and the feeder line capacity. For this paper, we will assume a
typical 13.8kV feeder line with an 11MVA capacity. The typical feeder capacity should be large
enough to mitigate any complications caused by a 5MW DG system in normal grid connected
operations [1]. A 5MW DG system is sufficiently large to be credible as defined in the
distributed generation facility sizing section. A system of this size can carry a peak load and
perform as a toggling facility. This size also serves to minimize the distance to the load, reduce
transmission losses and improve reliability. Hence, the penetration depth will be maximized in
these terms.
The load density (power requirements) of the given area further dictates the proximity to the DG
system. It follows that in a large city, skyscraper and large building basements provide suitable
locations for enough DG systems to meet the city's energy needs. Likewise, in smaller cities and
communities demand far less energy and one or two DG systems located within the community
could provide ample power. Large-scale industrial and commercial complexes would be outfitted
with multiple 5MW, or larger units if the base load demand warrants extra capacity.
According to a model like this, each DG system will support a defined power 'borough' or
electrical community. Potentially, smaller communities' DG systems could be less than 5MW.
Further, the possibility of household power generation on the 1-2kW level and commercial
power generation on the level of 10-100kW could be realized. Presently, this is not a viable
solution as there are many unresolved system control, cost, and technology maturation issues.
This paper will discuss distributed generation technologies on the 5MW level and on a limited
basis include generation technologies down to the 250kW level.
Ranking system
We describe our ranking parameters below. The ranking system is broken down into several
different areas of interest. There are 4 technologies being considered and thus we have ranked
them from 1 to 4, where 1 is the highest and 4 is the lowest rank.
Technology Status This ranking is based on the status of the technology being described. This
includes a brief history with regard to when the technology was developed, global market
penetration and current competitive status with other DG technologies.
Overall Efficiency This ranking is based on electrical, fuel and overall efficiency. The efficiency
is very important for reasons of conserving natural resources. Higher efficiency technologies will
have lower fuel costs per kW because less fuel will need to be purchased to generate the same
amount of power.
Environmental and Noise Considerations This ranking is based on technology location
restrictions based on pollution and noise levels. Obvious reasons dictate the public's acceptance
(and subsequent ranking) level of DG systems when implemented near population centers in lieu
of centralized generation. This section also takes into account pollution impact on a global level.
Device Cost Per kW As suggested, it is the cost of the given technology per kW or monetary
efficiency. Lower cost solutions will enable faster implementation and increased acceptance to
investors.
Physical Size The physical size determines the amount of area a technology will occupy.
Obviously a higher power density is desired because the DG system will occupy less real estate
and will be able to be implemented in more locations. Smaller DG stations will also be less
unsightly to residents at some potential locations.
I ntended Application The intended application as defined in the scope is on the minimum level
of 250kW to maximum level of 5MW. This analysis compares how well the technology
performs in terms of power generation to the other technologies.
Potential For Improvement Realizing that ICEs are the oldest technology followed by CTs, MTs
and fuel cells, this section compares the technologies' potential for improvement. Basically, this
section determines the best of these technologies suited for DG when figured on a long time
frame.
*This module was originally written as a manuscript by Joel Gouker and Michael Schenck.
Next >>
2007 Consortium on Energy Restructuring, Virginia Tech
5.2 Internal Combustion Engines
Technology Overview
Internal combustion engines date back to the mid- nineteenth century. Nikolaus A. Otto invented
the four-stroke engine in 1876 and Rudolf Diesel patented the diesel engine in 1892. The diesel
engine has a theoretical efficiency limit of 75.6% while modern compression-ignition systems
have a fuel efficiency limit of up to 40%. These two technologies offer different solutions to
render energy from a combustion engine. A normal gasoline engine mixes gas and air using a
carburetor or fuel injection, compresses the air and then uses a spark to detonate the mixture.
This process is known as the Otto cycle. A diesel engine operates slightly differently using
compression technology. Instead of an air fuel mixture, a diesel engine takes in air, compresses
it, and then injects the fuel. The heat of the compressed air is enough to ignite the fuel.
For the power range we are concerned with, both types of engines are appropriate. However, the
generator set up (genset) and turnkey costs of diesel engines are about half the cost of gas
engines. Hence, we are only considering diesel engines in the comparison. Diesel engines can be
classified into three different categories that have been optimized for certain size power
generation installations:
Low speed (200-400 rpm) - 100MW
Medium speed (800-1200 rpm) - 600kW to 7000kW
High speed (1500-3600 rpm) - 200 to 800kW
The target installation size dictates the engine speed level and fuel type. The medium- and high-
speed diesel ICEs offer existing commercial solutions within our scope [12].
Technology Status
ICEs Ranked 1
st

ICEs as the fundamental prime mover are the oldest type of implemented distributed generation
system. Having existed in the commercial industry for nearly 100 years, they have evolved
significantly from the original ICEs. Diesel engines are in use in hundreds of thousands of trucks
worldwide and have been for many years. The compact and tailored designs make them the
preferred backup power generation for most major corporations worldwide. Diesel engines have
pervasively penetrated the world market. Due to this penetration and the relative age of diesel
engines, they have become familiar to many mechanics and engineers. Diesel engines also
possess a relatively simple design for maintenance purposes. Age, ease of maintenance, and
technical familiarity create a very highly reliable machine.
When compared to the theoretical efficiencies and the emissions efficiencies of other DG
technologies, diesel engines must be optimized to maintain their current market stronghold.
Recent developments in small diesel engines indicate a promising future. The small engine
market has traditional been ignored by the power industry. It is more difficult to emissions'
optimize small units than large units. At Cummins Diesel, their popular small truck engines
feature 'quiet diesel' with more precise full injections techniques that will undoubtedly carry over
to larger machines [3]. Other automotive manufactures have developed engines that can be
classified as 'ultra low emissions vehicle' horsepower range engines [4]. These engines present
the power industry with potential future diesel low-emission DG technologies.
We have conferred the first place rank to ICEs due to their proven history, their wide market
penetration, the ease of repair to these systems, and their current competitive status with other
DG technologies.
Environmental and Noise Considerations
ICEs Ranked 4
th

Despite the diesel engine's long history in the backup and uninterruptable power supply
generator set up (UPS genset) market, the diesel engine design did not have emissions reduction
as a priority. End users have accepted excessively high NO
x
and SO
x
emissions as the norm.
However, now our (the U.S.) government restricts most operating times to less than 150 hours
per year because of environmental concerns [10]. In addition to the environment considerations,
most sub 5MW diesel gensets lag behind automotive engine developments and create quite a bit
of noise. Diesel engines produce significant sources of environmental and noise pollution when
compared to the other DG technologies discussed.
Device Cost
ICEs Ranked 1
st

ICEs rank at the top of the list along with combustion turbines with a low $200-500 per kW
range. ICEs have been able to achieve their low cost because of their long history and mass
production. International price reduction agreements and contracts have been made with
everything from individual parts manufacturers down to mining and materials vendors.

Figure 1. 1.75MW diesel gensets for DG applications [5].
Physical Size
ICEs Ranked 1
st

Diesel engines have definitely benefited from technological advances of the automotive industry
pertaining to size reduction. In order for high power engines to be able to fit under the hood of a
16-wheeler, higher power density was demanded. As a result, the power density of a diesel
engine is .6 cubic feet per kW. Hence, diesel engine generators can be installed in more size
restrictive locations than the other DG technologies. Figure 1 illustrates 1.75MW trailer sized
Caterpillar gensets used in a DG application for peak shaving. In this form, the diesel engine
offers portability and convenience. However, where the footprint is more of a concern, the serial
nature of ICEs increases the length and width of the unit. Cursory examinations of product data
sheets illustrate the microturbine's small footprint area.
Intended Application
ICEs Ranked 2
nd

ICEs are very versatile in scaling. They can be built and are available from 1kW to more than
25MW for various duty cycles and load patterns [1]. The scope of this paper defines a 5MW
maximum where many commercial and proven ICEs are available. Caterpillar currently has off-
the-shelf solutions in this range as illustrated in Figure 2. Most ICEs are being used in backup
power and peak shaving applications. Diesel engines will fit the intended DG application
perfectly provided greenhouse gas emissions can be reduced to levels where they are operated
year round.
Potential for Improvement
ICEs Ranked 4
th

The long successful life of ICEs is not directly responsible for a low ranking in this category.
Though diesel engines have current fuel efficiencies of near 40%, no significant improvements
are expected in the near future. Noise reduction techniques are being researched and emissions
levels are being improved by the development of a new type of clean diesel fuel. Even though
reductions of 12% in NO
x
emissions and 24% in particulate matter emissions have been realized,
a definitive solution for large-scale units may still be years away [6]. For Diesel engines to be a
viable widespread DG technology, the greenhouse gas emissions must be significantly reduced.

<< Back Next >>
2007 Consortium on Energy Restructuring, Virginia Tech




http://www.dg.history.vt.edu/ch2/impact.html

Вам также может понравиться